You are on page 1of 1007

Technical Notes on Brick Construction

is a series of bulletins that contain design,


detailing and construction information based
on the latest technical developments in brick
masonry. Drawings, photographs, tables and
charts illustrate appropriate topics. They are
available individually or as a set. Registered
purchasers of a complete set will receive noti-
fication of new or revised editions via email.
Individual copies are available to view and
download free of charge from BIAs website
at www.gobrick.com.
To view, download or order Technical
Notes, go to the BIA website at www.
gobrick.com and select Technical Notes.
SUBJECT NUMBER
A
ACI 530/ASCE 5/TMS 402 Building Code 3
Admixtures in Mortar 1,8
Anchor Bolts 44
Arches 31-31C
Construction 31
Flashing 31
Semi-circular 31C
Structural Design 31A
ASTM International Standards
Anchors and Ties 7A, 28B, 44B
Brick 9A
Mortar 8
Pavers 9A, 14
Testing 39 Series
B
Barrier Walls 7
Beams 17B
Bearing Walls
(see Engineered Brick Masonry)
Bonds and Patterns in Brickwork 30
Paving Patterns 14, 29
Bond Breaks 18A, 21B
Bond - Mortar 8 Series
Reinforced Brick Masonry 17, 17A
Brick Sizes 9B, 10
C
Calculated Fire Resistance 16
Caps 36A
Cavity Walls 21-21C
Construction 21C
Detailing 21B
Glazed Brick 13
Materials 21A
Passive Solar Heating 43
Properties 21
Chimneys 19-19C
Classification of Brick 9A
Cleaning 20
Efflorescence 20
Coatings for Brick 6, 6A
Cold Weather Construction 1
Guide Specifications 11A
Color - Brick 9
Mortar 8
Columns and Pilasters 3B
Compressive Strength
Brick Masonry 3A, 39A, 42
Brick Units 9A, 39
Mortar 8 Series
Walls 39A, 42
Condensation 28B, 47
Control Joints (see Expansion Joints)
Copings 36
Corbels and Racking 36A
Corrosion
Metal Ties 7A, 44B
Shelf Angles 7A
Steel Lintels 7A, 31B
Coursing Tables for Brick 10
Cracking 18
Curtain and Panel Walls 17L
D
Dampproofing 46
Differential Movement 18, 18A
Bond Breaks 18A
Expansion Joints in Paving 14 Series
Expansion Joints in Walls 18A
Flexible Anchorage 18A
Material Properties 18
Structures without Shelf Angles 18A
Volume Changes and Effects 18
Dimensioning 10
Direct Gain, Passive Solar Heating 43
Drainage Walls 7
E
Efflorescence
Identification and Prevention 23
Causes and Prevention 23A
Removal 20
Empirical Design of Brick Masonry 42
Energy
Codes 4B
Embodied 48
Heat Transmission Coefficients 4
Engineered Brick Masonry
Allowable Design Stresses 3A, 39A
Bearing Wall 24
Building Code Requirements 3, 16
Construction 24F
Detailing 24G
Guide Specifications 11 Series
Material Properties 3A
Quality Control 39B
Section Properties 3B
Shear Wall Design 24C
Testing 39 Series
Wall Types and Properties 3B
Equivalent Thickness 16
Estimating Material Quantities 10
Expansion Joints 18A
Paving 14 Series
F
Fasteners for Brick Masonry 44A
Fences 29A
Field Panels 9B
Fireplaces, Residential 19-19E
Contemporary, Projected Corner,
Rumford, Multi-faced 19C
Details and Construction 19B
Finnish Style Masonry Heater 19E
Russian Style Masonry Heater 19D
Fire Resistance 16
Flashing, Types and Selection 7A
Arches 31
Details 7
Replacement 46
Flexible Paving Systems 14
Floor-Wall Connections, Bearing Wall 26
Freeze Thaw Durability 7A, 9A, 9B
Freezing, Protection from 1
G
Garden Walls 29A
Girders, Reinforced Brick Masonry 17M
Glazed Brick
Specifications 9A
Walls 13
Glossary 2
Green Building 48
Grout 17, 17A
Properties 3A
Testing 39
Guide Specifications 11 Series
H
Heat Transmission Coefficients 4
High-Lift Grouting 17A
Hollow Brick Masonry 41
Reinforced 17 Series
Hot Weather Construction 1
TECHNICAL NOTESon Brick Construction Index
June
2009
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
SUBJECT NUMBER
2009 Brick Industry Association, Reston, Virginia Page 1 of 4
Subject Index
SUBJECT NUMBER
www.gobrick.com | Brick Industry Association | TN INDEX | Subject Index | Page 2 of 4
I
Inspection 46
Reinforced Brick Masonry 17A
Initial Rate of Absorption 7A, 8B, 9A
9B, 39
L
Landscape Architecture 29-29B
Garden Walls 29A
Miscellaneous Applications 29B
Paving 14 Series
Pedestrian Applications 29
Lateral Forces, Shear Wall Design 24C
LEED 48
Lintels
Reinforced Brick 17B
Structural Steel 31B
Loadbearing Brick Homes 26
M
Maintenance 46
Cleaning 20
Manufacturing Brick 9
Material Properties 3A
Masonry Heaters 19D, 19E
Modular Brick Masonry 10
Moisture Control
Barrier Walls 7
Caps and Copings 7, 36A
Condensation 47
Corrosion 7A, 31B, 44B
Drainage Walls 7
Flashing 7A
Glazed Brick Walls 13
Maintenance 46
Mortar 8
Rain Screen Wall 27
Repointing 46
Water Repellent Coatings 6A
Weeps 7
Moisture Expansion 18
Mortars for Brickwork 8 Series
Cold Weather Construction 1
Efflorescence 8, 23 Series
Estimating Quantities 10
Guide Specifications 11E
Joints 7B, 21C
Materials 8
Mixing 8B
Paving Systems 14
Quality Assurance 8B
Reinforced Masonry 17A
Repointing 46
Selection 8B
Movement (see Differential Movement)
N
Noise Barrier Walls 45
Structural Design 45A
P
Painting Brick Masonry 6
Parapets 7, 18 Series, 36A
Passive Solar
Cooling 43C
Details 43G
Heating 43
Materials 43D
Patterns 30
Paving Systems 14 Series
Accessibility 14
Bases 14
Clay Pavers 9A, 14
Cleaning 20
Coatings 6A
Details 14 Series
Drainage 14 Series
Edge Restraint 14 Series
Expansion Joints 14 Series
Ice and Snow Removal 14
Installation 14 Series
Interlock 14A
Maintenance 14 Series
Patterns 14, 29
Permeable Pavements 14
Sand Setting Bed 14A
Traffic 14
Piers and Pilasters 3B
Portland Cement/Lime Mortar 8 Series
Prefabricated Brick Masonry
Introduction 40
Thin Brick 28C
Pressure-Equalized Rain Screen Wall 27
R
Rain Penetration (see Moisture Control)
Rain Screen Wall 27
Recycled Content 48
Reinforced Brick Masonry
Beams 17B
Curtain and Panel Walls 17L
Flexural Design 17B
Girders 17M
High-Lift Grouted 17A
History 17
Hollow Brick Masonry 26, 41
Inspection 17A
Lintel Design 17B
Materials 17A
Mortar and Grout 17A
Specifications 11 Series
Workmanship 17A
Repointing 46
Retrofit 28A
Rigid Paving Systems 14
Rumford Fireplaces 19C
R-Values 4, 4B
S
Salvaged Brick 15
Sealers (see Water Repellents)
Sealants 18A, 28
Section Properties 3B
Selection of Brick 9B
Serpentine Walls 29A
Shelf Angles
Typical Details 7, 28B
Corrosion Resistance 7A
Single-Wythe Bearing Walls 26
Sills 36
Sizes of Brick 9B, 10
Slip/Skid Resistance 14
Soffits 36
Solar Energy (see Passive Solar Systems)
Sound Barriers (see Noise Barrier Walls)
Sound Insulation 5A
Spalling 46
Specifications, General 11 Series
ACI 530.1/ASCE 6/TMS 602 3
Brick 9A
Cold and Hot Weather Construction 1
Mortars 8, 11E
Pavers 9A
Stains
Identification and Prevention 23
Removal 20
Steel Studs 28B
Steps and Ramps 29
Sustainability 48
Sustainable Development 29
T
Terminology 2
Terraces 29
Testing of Brick and Mortar 39 Series
Allowable Design Stresses 3A, 39A
Quality Control 39B
Thermal Expansion of Walls 18 Series
Thermal Storage Walls, Passive Solar
Heating 43
Thermal Transmission Coefficients 4
Thin Brick 28C
Ties and Reinforcement
Adjustable 44B
Corrosion Resistance 7A, 44B
Joint Reinforcement 44B
Specifications 11A
Tolerances 9A, 11C
Tooling 7B
Tuckpointing (see Repointing)
U
Used Brick 15
U-Values 4, 4B
V
Veneer Construction 28 Series
Existing Construction 28A
Hollow Brick 41
Steel Studs 28B
Thin Brick Veneer 28C
Wood Studs 28
W
Wall Ties 44B
Water Penetration (see Moisture Control)
Water Repellent Coatings 6A
Weeps 7, 46
Winter Construction
(See Cold Weather Construction)
Workmanship 7B, 21C
Reinforced Brick Masonry 17A
Specifications 11 Series
SUBJECT NUMBER SUBJECT NUMBER SUBJECT NUMBER
Numerical Index
www.gobrick.com | Brick Industry Association | TN INDEX | Numerical Index | Page 3 of 4
Technical Notes are rewritten to include new
technical information. The issue date of a current
Technical Note is shown between brackets [ ].
Current editions supersede earlier editions.
The designation Reissued indicates that the
edition of the Technical Note shown in brackets [ ]
has been thoroughly reviewed and found to be
technically accurate. Other editions dated on or
after the bracketed [ ] date are still valid; only
minor editorial changes have been made. The
reissued date appears in parentheses ( ).
Missing numbers have been discontinued.
1 [June 2006] Cold and Hot Weather
Construction
2 Rev [Jan./Feb. 1975] (Reissued March 1999)
Glossary of Terms Relating to Brick Masonry
3 Rev [July 2002] Overview of Building Code
Requirements for Masonry Structures
ACI 530-02/ASCE 5-02/TMS 402-02 and
Specifications for Masonry Structures ACI
530.1-02/ASCE 6-02/TMS 602-02
3A [Dec. 1992] Brick Masonry Material
Properties
3B [May 1993] Brick Masonry Section
Properties
4 Rev [Jan. 1982] (Reissued Sept. 1997) Heat
Transmission Coefficients of Brick Masonry
Walls
4B Rev [Feb. 2002] Energy Code Compliance
of Brick Masonry Walls
5A [June 1970] (Reissued Aug. 2000) Sound
Insulation Clay Masonry Walls
6 Rev [May 1972] (Reissued Dec. 1985)
Painting Brick Masonry
6A [Aug. 2008] Colorless Coatings for Brick
Masonry
7 [Dec. 2005] Water Penetration Resistance
Design and Detailing
7A [Dec. 2005] Water Penetration Resistance
Materials
7B [Dec. 2005] Water Penetration Resistance
Construction and Workmanship
8 [Jan. 2008] Mortars for Brickwork
8B [Oct. 2006] Mortars for Brickwork
Selection and Quality Assurance
9 [Dec. 2006] Manufacturing of Brick
9A [Oct. 2007] Specifications for and
Classification of Brick
9B Rev [Dec. 2003] Manufacturing,
Classification and Selection of Brick
Selection, Part III
10 [Feb. 2009] Dimensioning and Estimating
Brick Masonry
11 Rev [Dec. 1971] (Reissued Aug. 2001)
Guide Specifications for Brick Masonry,
Part I
11A Rev [June 1978] (Reissued Sept. 1988)
Guide Specifications for Brick Masonry,
Part II
11B Rev [Feb. 1972] (Reissued Sept. 1988)
Guide Specifications for Brick Masonry,
Part III
11C Rev [July 1972] (Reissued May 1998)
Guide Specifications for Brick Masonry,
Part IV
11D [Aug. 1972] (Reissued Sept. 1988) Guide
Specifications for Brick Masonry, Part IV
Continued
11E Rev [Sept. 1991] Guide Specifications for
Brick Masonry, Part V, Mortar and Grout
13 [Dec. 2005] Ceramic Glazed Brick Exterior
Walls
14 [Mar. 2007] Paving Systems Using Clay
Pavers
14A [Oct. 2007] Paving Systems Using Clay
Pavers on a Sand Setting Bed
15 Rev [May 1988] Salvaged Brick
16 [Mar. 2008] Fire Resistance of Brick
Masonry
17 Rev [Oct. 1996] Reinforced Brick Masonry,
Introduction
17A Rev [Aug. 1997] Reinforced Brick Masonry
Materials and Construction
17B Rev [Mar. 1999] Reinforced Brick Masonry
Beams
17L Rev [Feb./Mar. 1973] (Reissued Sept.
1988) Four-inch RBM Curtain and Panel
Walls
17M [July 1968] (Reissued Sept. 1988)
Reinforced Brick Masonry Girders
Examples
18 [Oct. 2006] Volume Changes Analysis and
Effects of Movement
18A [Nov. 2006] Accommodating Expansion of
Brickwork
19 Rev [Jan. 1993] Residential Fireplace
Design
19A Rev [May 1980] (Reissued Aug. 2000)
Residential Fireplaces, Details and
Construction
19B Rev [June 1980] (Reissued Apr. 1998)
Residential Chimneys Design and
Construction
19C Rev [Oct. 2001] Contemporary Brick
Masonry Fireplaces
19D [Jan. 1983] (Reissued June 1987) Brick
Masonry Fireplaces, Part I, Russian-Style
Heaters
19E [1983] (Reissued Feb. 1998) Brick
Masonry Fireplaces, Part II Fountain and
Contemporary Style Heaters
20 [June 2006] Cleaning Brickwork
21 Rev [Aug. 1998] Brick Masonry Cavity Walls
Introduction
21A Rev [Feb. 1999] Brick Masonry Cavity
Walls Selection of Materials
21B [Apr. 2002] Brick Masonry Cavity Walls
Detailing
21C Rev [Oct. 1989] Brick Masonry Cavity
Walls Construction
23 [June 2006] Stains Identification and
Prevention
23A [June 2006] Efflorescence Causes and
Prevention
24 Rev [June 2002] Brick Masonry Bearing
Walls Introduction
24C Rev [Sept./Oct. 1970] (Reissued May
1988) The Contemporary Bearing Wall
Introduction to Shear Wall Design
24F Rev [Nov./Dec. 1974] (Reissued Sept.
1988) The Contemporary Bearing Wall
Construction
24G [Dec. 1968] (Reissued Feb. 1987) The
Contemporary Bearing Wall Detailing
26 Rev [Sept. 1994] Single Wythe Bearing
Walls
27 Rev [Aug. 1994] Brick Masonry Rain Screen
Walls
28 Rev [Aug. 2002] Anchored Brick Veneer,
Wood Frame Construction
28A [Apr. 2008] Adding Brick Veneer to
Existing Construction
28B [Dec. 2005] Brick Veneer / Steel Stud
Walls
28C [Jan. 1986] (Reissued Jan. 2001) Thin
Brick Veneer Introduction
29 Rev [July 1994] Brick in Landscape
Architecture Pedestrian Applications
29A Rev [Nov. 1968] (Reissued Jan. 1999)
Brick in Landscape Architecture Garden
Walls
29B [Apr. 1967] (Reissued May 1988) Brick
in Landscape Architecture Miscellaneous
Applications
30 Rev [Mar. 1999] Bonds and Patterns in
Brickwork
31 Rev [Jan. 1995] Brick Masonry Arches
31A [Oct. 1967] (Reissued July 1986)
Structural Design of Brick Masonry Arches
31B Rev [Nov./Dec. 1981] (Reissued May
1987) Structural Steel Lintels
31C Rev [Feb. 1971] (Reissued Aug. 1986)
Structural Design of Semicircular Brick
Masonry Arches
36 Rev [July/Aug. 1981] (Reissued Jan. 1998)
Brick Masonry Details, Sills and Soffit
36A Rev [Sept./Oct. 1981] (Reissued Feb.
2001) Brick Masonry Details, Caps and
Copings, Corbels and Racking
39 Rev [Nov. 2001] Testing for Engineered
Brick Masonry Brick and Mortar
39A [July/Aug. 1975] (Reissued Dec. 1987)
Testing for Engineered Brick Masonry
Determination of Allowable Design Stresses
39B Rev [Mar. 1988] Testing for Engineered
Brick Masonry Quality Control
40 Rev [Aug. 2001] Prefabricated Brick
Masonry Introduction
41 [Jan. 2008] Hollow Brick Masonry
42 Rev [Nov. 1991] Empirical Design of Brick
Masonry
www.gobrick.com | Brick Industry Association | TN INDEX | Numerical Index | Page 4 of 4
43 Rev [May/June 1981] Passive Solar Heating
with Brick Masonry Part I Introduction
43C [Mar. 1980] (Reissued Feb. 2001) Passive
Solar Cooling with Brick Masonry, Part I
Introduction
43D [Sept./Oct. 1980] (Reissued Sept. 1988)
Brick Passive Solar Heating Systems, Part IV
Material Properties
43G [Mar./Apr. 1981] (Reissued Sept. 1986)
Brick Passive Solar Heating Systems, Part
VII Details and Construction
44 [Apr. 1986] Anchor Bolts for Brick Masonry
44A [May 1986] (Reissued Aug. 1997)
Fasteners for Brick Masonry
44B Rev [May 2003] Wall Ties for Brick
Masonry
45 [Feb. 1991] (Reissued July 2001) Brick
Masonry Noise Barrier Walls Introduction
45A [Apr. 1992] Brick Masonry Noise Barrier
Walls Structural Design
46 [Dec. 2005] Maintenance of Brick Masonry
47 [June 2006] Condensation Prevention and
Control
48 [June 2009] Sustainability and Brick
Changes to Technical Notes
(since December 2005)
New and Revised
1 June 2006
6A August 2008
7 December 2005
7A December 2005
7B December 2005
8 January 2008
8B October 2006
9 December 2006
9A October 2007
10 (replaces 10, 10A & 10B) February 2009
13 December 2005
14 March 2007
14A October 2007
16 (replaces 16 & 16B) March 2008
18 October 2006
18A November 2006
20 June 2006
23 June 2006
23A June 2006
28A April 2008
28B December 2005
41 January 2008
46 (replaces 7F) December 2005
47 (replaces 7C & 7D) June 2006
48 June 2009
Retired
7C (replaced by 47) June 2006
7D (replaced by 47) June 2006
7F (replaced by 46) December 2005
8A January 2008
10A (replaced by 10) February 2009
10B (replaced by 10) February 2009
16B (replaced by 16) March 2008
Cold and Hot Weather Construction
Abstract: This Technical Note defines cold and hot weather conditions related to brick masonry construction and describes
the unfavorable effects of these conditions on masonry materials and their performance. It provides information on weather
prediction necessary for construction planning and recommends practices to achieve optimum performance of masonry con-
structed during periods of extreme temperatures.
Key Words: absorption, ambient temperature, climatology, cold weather, evaporation, freezing, grout, hot weather, meteorology.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
1
June
2006
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Comply with cold and hot weather requirements of appli-
cable building codes
Follow requirements given in Table 1
Page 1 of 9
INTRODUCTION
Adequate planning and preparation can make brick construction possible in virtually all weather conditions. Cold
and hot weather can negatively affect masonry materials and the quality of constructed masonry. However, imple-
menting recommended changes to construction practices can usually ensure quality construction. Although nor-
mal, cold, and hot are relative terms, normal, used in this Technical Note, is any temperature between 40 F
and 100 F (4.4 C and 37.8 C). Cold is defined as temperature below 40 F (4.4 C); and hot, any temperature
above 100 F (37.8 C).
BUILDING CODE REQUIREMENTS
In many instances, building codes include mandatory measures intended to ensure the quality of masonry con-
structed during cold or hot weather. The International Building Code (IBC) [Ref. 1] includes a list of required cold
and hot weather construction provisions for masonry that are essentially identical to those found in Specification
for Masonry Structures (ACI 530.1/ASCE 6/TMS 602) [Ref. 11] and required by Building Code Requirements
for Masonry Structures (ACI 530/ASCE 5/TMS 402) [Ref. 6], both of which are referenced by the IBC. The
Specification for Masonry Structures provisions differ from those of the IBC in that they also require the submittal
and acceptance of a description of the hot and cold weather construction program prior to its use. The mandatory
cold and hot weather construction practices required by the IBC and Building Code Requirements for Masonry
Structures are summarized in Table 1.
Specific cold and hot weather provisions are not included within the International Residential Code (IRC) [Ref.
2]. However, the IRC states that mortar for use in masonry construction shall comply with ASTM C 270, which
requires mortar for other than masonry veneer to be prepared in accordance with the Masonry Industry Council's
"Hot and Cold Weather Masonry Construction Manual" [Ref. 8]. Hot and cold weather provisions apply to brick
veneer when the provisions of Building Code Requirements for Masonry Structures are used in lieu of the IRC
masonry provisions.
PLANNING FOR EXTREME WEATHER
To successfully build during periods of extreme weather conditions, designers and contractors utilize knowledge of
local meteorological conditions, as well as historic climatological information for a given area. During project plan-
ning, designers are concerned with climatological data such as the average and extreme daytime and nighttime
temperatures or average wind velocity for use in designing mechanical or structural systems. Contractors, how-
ever, are more concerned with meteorological conditions during construction, such as hourly temperatures and
mean daily temperature, as well as the predicted temperatures and wind velocities for the next few days. Mean
daily temperature is determined by adding together the maximum temperature for each day (24 hours, midnight
to midnight) and the minimum temperature for the same day and dividing by two. Ambient temperature as used in
this Technical Note is the outdoor temperature at the time considered.
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 2 of 9
Temperature
1
Preparation Requirements Construction Requirements Protection Requirements
(Prior to Work) (Work in Progress) (After Masonry Is Placed)
Shade materials and mixing
equipment from direct sunlight.
Comply with hot weather
requirements below.
Comply with hot weather requirements
below.
Maintain mortar and grout at a
temperature below 120 F (48.9 C).
Provide necessary conditions and
equipment to produce mortar having
a temperature below 120 F (48.9 C).
Flush mixer, mortar transport container,
and mortar boards with cool water
before they come into contact with
mortar ingredients or mortar.
Maintain sand piles in a damp,
loose condition.
Maintain mortar consistency by
retempering with cool water.
Use mortar within 2 hr of initial mixing.
N
o
r
m
a
l

W
e
a
t
h
e
r
100 F to 40 F
(37.8 C to 4.4 C)
Normal Procedures. Normal Procedures. Normal Procedures.
40 F to 32 F
(4.4 C to 0 C)
Do not lay masonry units having
either a temperature below 20F
(-6.7C) or containing frozen
moisture, visible ice, or snow on their
surface.
Remove visible ice and snow
from the top surface of existing
foundations and masonry to
receive new construction. Heat these
surfaces above freezing, using
methods that do not result in
damage.
Heat mixing water or sand to produce
mortar between 40 F (4.4 C) and 120
F (48.9 C).
Do not heat water or aggregates used in
mortar or grout above 140 F (60 C).
Heat grout materials when their
temperature is below 32 F (0 C).
Completely cover newly
constructed masonry with a
weather-resistive membrane for
24 hr after construction.
Comply with cold weather requirements
above.
Maintain mortar temperature above
freezing until used in masonry.
Heat grout materials so grout is at a
temperature between 70 F (21.1 C)
and 120 F (48.9 C) during mixing and
placed at a temperature above
70 F (21.1 C).
Comply with cold weather requirements
above.
Heat masonry surfaces under
construction to 40F (4.4C) and use
wind breaks or enclosures when the
wind velocity exceeds 15 mph
(24 km/h).
Heat masonry to a minimum of 40F
(4.4C) prior to grouting.
Comply with cold weather requirements
above.
32 F to 25 F
(0 C to -3.9 C)
Comply with cold weather
requirements above.
Comply with cold weather requirements
above.
Provide enclosure and heat to maintain
air temperatures above
32 F (0 C) within the enclosure.
20 F and Below
(-6.7 C and Below)
Comply with cold weather
requirements above.
Comply with cold weather
requirements above.
25 F to 20 F
(-3.9 C to -6.7 C)
1. Preparation and Construction requirements are based on ambient temperatures. Protection requirements, after masonry is placed, are based on mean daily temperatures .
Use cool mixing water for mortar and
grout. Ice must be melted or removed
before water is added to other mortar or
grout materials.
Above 115 F or
105 F with a wind
velocity over 8 mph
(46.1 C or 40.6 C
with a 12.9 km/hr wind)
H
o
t

W
e
a
t
h
e
r

Maintain newly constructed masonry
temperature above 32F (0C) for at
least 24 hr after being completed by
using heated enclosures, electric
heating blankets, infrared lamps, or
other acceptable methods. Extend time
period to 48 hr for grouted masonry,
unless the only cement in the grout is
Type III portland cement.
C
o
l
d

W
e
a
t
h
e
r

Completely cover newly
constructed masonry with weather-
resistive insulating blankets or equal
protection for 24 hr after completion of
work. Extend time period to 48 hr for
grouted masonry, unless the
only cement in the grout is Type III
portland cement.
Comply with hot weather requirements
below.
Fog spray newly constructed masonry
until damp, at least three times a day
until the masonry is
three days old.
Above 100 F or
90 F with 8 mph wind
(above 37.8 C or
32.2 C with a
12.9 km/hr
wind)
TABLE 1
Requirements for Masonry Construction in Hot and Cold Weather
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 3 of 9
Meteorological information can be obtained from the National Weather Service, a branch of the National
Oceanographic and Atmospheric Administration (NOAA). The National Weather Service has information centers
located at major airports in cities throughout the country. These centers provide current weather information and
regularly scheduled weather forecasts for the surrounding region.
Climatological information can be obtained from the National Climatic Data Center, also a branch of NOAA. The
National Climatic Data Center usually provides climatic information in the form of maps as shown in Figure 1.
These maps contain daily, monthly and annual data for a region and may be obtained free online or by contacting
the Center [Ref. 7].
NEGATING THE EFFECTS OF COLD WEATHER
Successful construction considers the effects of cold weather on masonry materials in the planning, scheduling
and set up of the masonry work and protection of the completed work. This section describes the properties of
masonry and masonry materials that are changed by low temperatures and code prescribed construction pro-
cedures that overcome these effects. In addition to anticipating specific weather conditions, these provisions,
presented in Table 1, assist the contractor in determining how to protect building materials, unfinished and newly
constructed masonry.
In regard to the quality of masonry constructed during cold weather, perhaps the most critical factor is ensuring
that mortar and grout maintain adequate heat for normal cement hydration. Without sufficient heat, cement hydra-
tion slows and may stop completely, arresting the development of the masonrys compressive and bond strengths.
Heating Materials
Masonry Units. Masonry units are the components of a masonry assembly least affected by below-normal tem-
peratures. The physical properties of masonry units are essentially unchanged by cold weather, however, the tem-
perature of brick and their absorption characteristics influence the rate of freezing of masonry. A cold masonry unit
will have a slightly smaller volume than one at normal temperatures.
Cold units draw heat from mortar and more rapidly reduce the temperature of mortar to points at which normal
cement hydration is retarded and freezing occurs. Preheating masonry units prior to laying helps to maintain heat
within the mortar and minimize the effect of cold temperatures on mortar hydration. When ambient temperatures
are below 20 F (-6.7 C), masonry units must be heated to a temperature of at least 40 F (4.4 C) prior to laying.
Masonry units having either a temperature below 20 F (-6.7 C) or containing frozen moisture, visible ice or snow
on their surface must not be laid. Frozen masonry units must be thawed and should by dried before use. Unit tem-
perature can be measured using a metallic surface contact thermometer or flat, instant-read thermometer.
It may be advantageous to heat brick even when ambient temperatures are above 20 F (-6.7 C). Preheated
brick will exhibit the same absorption characteristics as those laid at normal temperatures. Brick with higher Initial
Rates of Absorption (IRAs) more rapidly absorb water from mortar or grout, thereby reducing the risk of damage
from the expansive forces of freezing water in the mortar.
Mortar. Mortar mixed using cold materials has different properties from mortar mixed with materials at normal tem-
peratures. Low temperatures retard the hydration of the cement in mortar. Mortar mixed during cold weather often
Figure 1
Examples of Climatic Data Available
J anuary Mean Daily Maximum Temperature J anuary Mean Daily Minimum Temperature
LEGEND
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 4 of 9
has lower water content, increased air content, and reduced early strength compared with mortar mixed at normal
temperatures.
In freezing weather, ice may be present in mixing water and moisture in the sand may turn to ice. Ice in the mixing
water must be melted or removed before the water can be added to the mixer. Do not use sand containing frozen
particles or frost. At a minimum, any ice must be melted and additional heating may further improve mortar perfor-
mance.
Avoid freezing of mortar during construction in all cases. In cold weather, mix mortar in smaller amounts so it can
be used before it cools. In any case, use mortar within 2
1
/2 hours from the time of initial mixing. Mortar that freezes
may experience significant reductions in compressive strength. Further, bond strength, extent of bond and water
penetration resistance of masonry may be reduced. Mortar having a water content exceeding six percent of the
total volume may be damaged due to the increase in volume as freezing water is converted to ice.
Mortar mixed with heated materials can approximate the performance characteristics of mortar mixed at normal
temperatures. For these reasons, the codes include requirements for heating mortar materials.
When ambient temperatures fall below 40 F (4.4 C) sand or mixing water must be heated to produce mortar that
is between 40 F (4.4 C) and 120 F (48.9 C) at the time of mixing. Ideal temperatures for mortar are between
60 F (15.6 C) and 80 F (26.7 C). Mortar temperatures over 120

F (48.9 C) may lead to flash set, resulting in
lower compressive strength and reduced bond strength. Thus, do not heat sand or water above 140 F (60.0 C).
Water is the easiest and best material to heat because it does not lose heat readily. Heating prepackaged materi-
als such as portland, mortar and masonry cements and hydrated lime can be difficult. If the air temperature is
below 32 F (0 C), maintain the temperature of mortar above freezing until used.
Consider altering mortar constituents or proportions within permissible ranges to reduce the impacts of cold
weather. Increasing sand content provides a stiffer mortar that better supports the weight of subsequently laid
masonry. Using masonry or mortar cements, or reducing lime content allows mortars to lose water more rap-
idly, thus reducing the potential for freezing. High-early-strength (Type III) portland cement may also be used to
increase the rate of early strength gain. If a brick with a low IRA is used, the water content of the mortar should be
the minimum necessary for workability. Set accelerating admixtures, as discussed later in Other Cold Weather
Considerations may also be used, however heating and protection measures are still required.
Avoid freezing of mortar during construction in all cases, and protect mortar in newly completed masonry from
freezing. Specific requirements for protection of mortar are found in Table 1.
Grout. High water content is necessary in grout for ease of flow, but it greatly increases the amount of volumet-
ric expansion which can occur upon freezing. Thus grout, like mortar, must be mixed with heated materials if the
temperature of the materials is below 32 F (0 C), to prevent the damaging effects of freezing. If the ambient tem-
perature is below 32 F (0 C), grout aggregates and mixing water must be heated to produce a grout temperature
between 70 F (21.1 C) and 120 F (48.9 C) at the time of mixing. Do not heat grout aggregates and mixing
water above 140 F (60.0 C) and keep the grout temperature above 70 F (21.1 C) when it is placed. High-early-
strength (Type III) portland cement may be used to increase the rate of early strength gain of grout. Admixtures
may also be used, but heating and protection of the grouted masonry is still required. All grout must be placed
within 1
1
/2 hours of mixing.
Newly Constructed and Completed Masonry. Because the hydration of cement is a process that continues for
an extended period, it is necessary to ensure that masonry surfaces under construction do not extract excessive
heat from mortar and grout. Code provisions address this by requiring masonry under construction, or that is to
receive grout, to be heated to a minimum temperature of 40 F (4.4 C) when the ambient temperature reaches 25
F (-3.9 C) or below. If wind velocities exceed 15 mph, wind breaks or enclosures must be used during construc-
tion. In addition to these measures, if the ambient temperature falls to 20 F (-6.7 C) or below, the masonry under
construction must be enclosed and the air within the enclosure maintained at a temperature above 32 F (0 C).
Newly constructed masonry that is frozen may be moistened after thawing to reactivate the hydration process and
continue to develop strength.
If snow or ice is visible on existing foundations or masonry, the codes prohibit building new masonry on them.
There is danger of movement when the base thaws, and bond cannot be developed between the mortar bed and
frozen supporting surfaces. Ice and snow must be removed and the top surface must be heated to above freez-
ing, in a manner that does not damage the masonry.
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 5 of 9
Protecting Materials and Masonry
In addition to heating materials to adjust for cold weather, the
IBC and Building Code Requirements for Masonry Structures
require protection of masonry constructed in cold weather.
Protection is one of the most effective adjustments that can
be made to construction practices.
Material Storage. Careless material storage can increase
the cost of laying masonry if removal of ice and snow from
materials and thawing of masonry units is necessary before
construction begins. Measures to consider include covering
masonry materials with tarpaulins or polyethylene sheets to
keep them dry and free of ice and snow, locating sand so
that water does not drain into it and storing masonry units on
raised platforms to avoid contact with the ground.
Newly Constructed Masonry. As mentioned, the develop-
ment of strength and bond in masonry continues for some
time after the masonry is completed, and may be com-
promised if freezing occurs. Therefore, newly constructed
masonry must be protected so that it maintains enough heat
for cement hydration. When the mean daily temperature falls
to 40 F (4.4 C) or below, a series of protective measures
are required, beginning with covering newly constructed
masonry with a weather-resistive membrane for 24 hr after
completion. As temperatures decrease, more stringent pro-
tection is required. Specific provisions for the progressively
colder temperatures are presented in Table 1.
Materials used to cover brickwork should be weighted or oth-
erwise fixed in place and extend a minimum of 2 ft (0.6 m)
down each side of the wall, as shown in Photo 1, to prevent
contamination by water, ice or snow.
HEATING METHODS AND
EQUIPMENT
Individual Materials
There are many types of equipment are available as sources
of heat for cold weather construction. The type selected will
depend upon availability of equipment, fuel source, econom-
ics, size of project and severity of exposure. A few common
methods for heating individual materials are described below.
Materials may also be heated by placing them within heated
enclosures prior to use.
Both water and sand used in mortar and grout may be heat-
ed to provide proper temperatures for construction. Sand
may be heated by placing an electric heating pad on top of
the sand pile and covering with a weather-resistant tarpaulin,
as shown in Photo 2. The electric pad can safely heat the
sand overnight without exceeding a temperature of 100 F
(37.8 C). A more labor intensive method of heating the sand
is to place over a heated pipe or to pile the sand around a
horizontal metal culvert or smoke stack section in which a
slow fire is built, as shown in Photo 3.
Photo 1
Cover Protecting Newly Constructed
Masonry
Photo 2
Heating of a Sand Pile
with an Electric Blanket
Photo 3
Sand Pile Warmed by Heated Pipe
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 6 of 9
Other methods for heating sand involve the use of a steam lance or other steam heaters. Pay careful attention to
the fire or other heat source and the sand as it should be heated slowly to avoid scorching.
Alternatively, an electric rod can be used to heat mixing water and sand simultaneously. The electric heating rod
is placed in a drum of water in the center of a sand pile. The rod heats the water over several hours. The sand
surrounding the drum slowly absorbs heat from the drum and insulates the drum from further heat losses.
Mortar may also be placed on electrically heated mortar boards to help maintain proper temperature. Be careful
to avoid excessive drying of the mortar.
Newly Constructed Masonry (Enclosures)
Contractors have used several different methods to provide heat and protection for newly constructed masonry,
including complete and partial enclosures. Large tents, temporary wood structures covered with clear plastic,
and shelters built of prefabricated panels covered with clear plastic sheets are examples of complete enclosures.
Partial enclosures often consist of enclosed scaffolds which may be moved from floor to floor when necessary, as
shown in Photo 4. Commercial electric blankets may also be used to cover walls and provide heat during the cur-
ing period.
Forced air heaters (sometimes called torpedo heaters or salamanders) are widely used as a source of heat within
enclosures. When complete enclosure of the work area is provided, space heaters are recommended, as shown
in Photo 5. Cold weather provisions require circulation of warm air on both sides of the masonry wall within the
enclosure.
OTHER COLD WEATHER CONSIDERATIONS
Admixtures
Accelerators. Accelerators are admixtures used to speed the setting time of mortar and grout. By increasing the
rate of cement hydration, accelerators increase the rate of early strength gain. The most common accelerators
are inorganic salts such as calcium chloride, calcium nitrate, soluble carbonates and some organic compounds.
Evaluate any accelerator for deleterious effects on masonry strength and materials. Admixtures that contribute to
staining or efflorescence or cause corrosion of metal accessories are not desirable for use in masonry construc-
tion. Indiscriminate use of accelerators can adversely affect the performance of the completed masonry. Using
accelerators alone does not address all concerns related to cold weather construction and is not recommended.
Masonry constructed using accelerators in mortar or grout must still be protected from freezing as cement hydra-
tion essentially stops at temperatures below 40 F (4.4 C).
Calcium chloride, while highly effective as an accelerator and widely used in the past, is not recommended as it
causes corrosion of metals used in masonry such as ties, anchors and reinforcement. For this reason, admixtures
with more than 0.2 percent chloride ions are prohibited for use in mortar when masonry is constructed under
the provisions of Building Code Requirements for Masonry Structures. The incidence of efflorescence may also
increase if the accelerator contains excessive salts.
Photo 5
Space Heater in Enclosure
Photo 4
Scaffold Enclosures
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 7 of 9
Calcium nitrite and calcium nitrate are inorganic non-chloride compounds also used as accelerators. These com-
pounds require higher dosages by weight and are more costly than calcium chloride, but will not corrode metals or
contribute to efflorescence.
Antifreeze. Do not use antifreeze compounds. These admixtures are alcohols or combinations of salts. If used
in the quantities required to be effective, significant reductions in mortar compressive and bond strengths usually
result. Most commercial mortar antifreeze admixtures do not lower the freezing point of mortar or grout, but are
actually accelerators. However, some true antifreeze admixtures are available.
NON-MANDATORY COLD WEATHER RECOMMENDATIONS
In addition to the mandatory requirements for cold weather masonry construction found in Table 1, the following
items can be incorporated in the specifications of the project manual where applicable:
- Protect masonry units, cementitious materials and sand so that they are not contaminated by rain, snow
or ground water.
- Units with higher initial rates of absorption (up to 40 g/min/30 in.
2

(40 g/min/194 cm
2
)) may be used to
resist mortar freezing. However, units with suctions in excess of 30 g/min/30 in.
2
(30 g/min/194 cm
2
)
should be wetted, but not saturated, with heated water just prior to laying. Water used for wetting should
be above 70 F (21.1 C) when units are above 32 F (0 C). If units are 32 F (0 C) or below, water tem-
perature should be above 120 F (48.9 C).
If walls are properly covered when work is halted, ice or snow removal from walls should not be necessary.
However, in the event that the covering is displaced, the top course may be thawed with steam or a carefully
applied portable blowtorch. The heat should be sustained long enough to thoroughly dry the masonry. If portions
of the masonry are frozen or damaged, replace defective parts before progressing with new work.
NEGATING THE EFFECTS OF HOT WEATHER
This section describes the properties of masonry and masonry materials that are changed by high temperatures
and the code prescribed procedures that overcome these effects.
Periods of hot weather may also adversely affect the construction of masonry. The contractor must ensure that the
quality of masonry construction does not suffer from the effect of high temperatures. The IBC and Building Code
Requirements for Masonry Structures define hot weather as temperatures above 100 F (37.8 C); however, wind
speed, relative humidity and solar radiation also influence the absorption of masonry units, the rate of set, and the
drying rate of mortar.
High temperatures and high humidity are not as damaging to the performance of the masonry as are low tem-
peratures and low humidity. The increased rate of cement hydration and favorable curing conditions in hot, humid
weather will help develop masonry strength if sufficient water is present at the time of construction.
The primary concern during hot weather is rapid evaporation and absorption of water from the mortar. Without suf-
ficient water, cement hydration slows or stops and the bond strength and extent of bond between brick and mortar
is reduced. The integrity of the masonry may also be compromised if mortar that is too hot flash sets before it
completes hydration.
The adjustments to construction practices required by the IBC and Building Code Requirements for Masonry
Structures further improve the quality of masonry constructed in hot weather. These mandatory provisions, trig-
gered when the ambient air temperature reaches 100 F (37.8 C), or 90 F (32.3 C) with a wind velocity greater
than 8 mph (12.9 km/hr), are presented in Table 1 and are discussed below along with additional non-mandatory
recommendations for successful hot weather construction. Keeping materials cool during periods of hot weather
provides the best results.
Cooling Materials
Lowering the temperature of materials may be the easiest approach to achieving performance characteristics
associated with masonry constructed at normal temperatures.
Masonry Units. Masonry units are not significantly affected by hot weather. However, the interaction between the
masonry units and the mortar or grout is critical. Masonry units that are hot absorb more water from mortar and
increase the temperature of the masonry. Lower bond strength and extent of bond result if not enough water is
present in the mortar when the units are laid.
Keep masonry units cool by storing them in a shaded area. Shading of masonry units from direct sunlight is
required when ambient temperatures exceed 115 F (46.1 C) or 105 F (40.6 C) with a wind velocity over 8
mph (12.9 km/hr).
Brick with field IRAs over 30 g/min/30 in.
2
(30 g/min/194 cm
2
) may be required to be wetted prior to laying to
reduce their rate of absorption. Otherwise, they can draw too much water from the mortar too quickly. Brick may
be required to be surface dry at the time of laying and have an IRA less than 30 g/min/30 in.
2

(30 g/min/194 cm
2
).
Brick may be wetted immediately before laying, but the preferred method is to wet them 3 to 24 hours before use.
Mortar. Mortar mixed at high temperatures often has a higher water content, lower air content, and a shorter
board life than mortar mixed at normal temperatures. It also tends to lose plasticity rapidly due to evaporation of
water and the increased rate of cement hydration. Consider using mortar with a high lime content and high water
retention. Rapid stiffening of hot mortar, or flash set, occurs if mortar plasticity is lost before the cement hydrates
sufficiently. To avoid this, be sure mortar used during hot weather maintains a temperature less than 120 F (48.9 C).
Retempering of mortar with cool water should always be permitted, and is required for maintaining consistency
during hot weather. Use mortar within 2 hr of initial mixing. Hot weather provisions require that all mortar materials
be shaded from direct sunlight when the ambient temperature exceeds 115 F (46.1 C) or 105 F (40.6 C) with a
wind velocity over 8 mph (12.9 km/hr).
Sand. When ambient temperatures exceed 100 F (37.8 C) or 90 F (32.2 C) with winds exceeding 8 mph
(12.9 km/hr), keep sand in a damp, loose condition. This can be achieved by sprinkling sand piles with water, and
leaving them uncovered, which also reduces the temperature of the sand through evaporative cooling. Damp sand
takes longer to heat up.
Water. Cool water may be used to help control the temperature of mortar and grout. Cool mixing water for mortar
and grout are required by hot weather provisions when the ambient temperature exceeds 115 F (46.1 C) or 105 F
(40.6 C) with a wind velocity of 8 mph (12.9 km/hr). Ice is highly effective in reducing the temperature of the mix
water. Ice must be completely melted or removed before combining the water with any other ingredients.
Grout. Grout reacts to hot weather in a manner similar to mortar. Water evaporates more rapidly and thereby
reduces the water-cement ratio. Because grout requires a slump of at least 8 in. (203 mm) for use in masonry,
maintain a high water-cement ratio by initially mixing grout with adequate water to offset evaporation. Building
Code Requirements for Masonry Structures requires grout to be used within 1
1
/2 hours of mixing. As with mortar,
ice may be used to lower the mix water temperature.
Admixtures. The use of admixtures to increase plasticity is not recommended unless their full effect on the mor-
tar is known. Admixtures for grout that increase the flow rate or reduce the water content are not recommended.
Shrinkage compensating admixtures are recommended.
Equipment. A significant amount of heat can be absorbed by equipment that is exposed to sunlight for extended
periods during hot weather. Mixers, wheelbarrows and mortar pans can impart this heat to mortar, raising its tem-
perature. Mortar boards made of wood may also absorb more water from mortar. To prevent this from compro-
mising the quality of masonry, the IBC and Building Code Requirements for Masonry Structures require mixers,
containers and mortar boards to be flushed with cool water before they come in contact with mortar or mortar
materials. As with mortar materials, equipment is also required to be shaded from direct sunlight when the ambi-
ent temperature exceeds 115 F (46.1 C) or 105 F (40.6 C) with an 8 mph (12.9 km/hr) wind velocity.
Protecting Materials and Masonry
Wet curing or fog spraying may further improve masonry strength development during periods of high tempera-
tures and low relative humidity. Hot weather provisions require fog spraying of newly constructed masonry until
damp, at least three times a day for three days when the mean daily temperature exceeds 100 F (37.8 C) or
90 F (32.2 C) with a wind velocity over 8 mph (12.9 km/hr.)
Use wind breaks to prevent rapid drying of mortar during and after placement, and cover walls with a weather-
resistant membrane at the end of the work day to prevent rapid loss of moisture from the masonry assemblage.
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 8 of 9
SUMMARY
Construction and protection requirements in both cold and hot weather help ensure uninterrupted, quality masonry
construction. Performance characteristics associated with materials mixed and constructed during normal tem-
peratures can be achieved by following recommendations in this Technical Note. Table 1 summarizes practices
required by building codes for cold and hot weather construction.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. 2003 International Building Code, International Code Council, Inc., Falls Church, VA, 2003.
2. 2003 International Residential Code for One- and Two-Family Dwellings, International Code Council, Inc.,
Falls Church, VA, 2003.
3. ASTM C 270, Standard Specification for Mortar for Unit Masonry, Annual Book of ASTM Standards, Vol.
04.05, ASTM International, West Conshohocken, PA, 2006.
4. Brown, M.L., Speeding Mortar Setting in Cold Weather, The Magazine of Masonry Construction, Vol. 2,
No. 10, Addison, IL, October 1989.
5. Bigelow, O., Cold Weather Masonry Construction, Masonry Magazine, Schaumburg, IL, October 2005.
6. Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402), The Masonry Society,
Boulder, CO, 2005.
7. Color Climate Atlas Maps, National Climatic Data Center, retrieved March 20, 2006 from
http://gis.ncdc.noaa.gov/website/ims-climatls/index.html.
8. "Hot and Cold Weather Masonry Construction Manual," Masonry Industry Council, Schaumburg, IL, 1999.
9. Randall, J r., F.A., and W.C. Panarese, Concrete Masonry Handbook, Portland Cement Association,
Skokie, IL, 1991.
10. Schierhorn, C., Preventing Hot-Weather Construction Problems, The Magazine of Masonry
Construction, Vol. 7, Addison, IL, J une 1994.
11. Specification for Masonry Structures (ACI 530.1/ASCE 6/TMS 602), The Masonry Society, Boulder, CO,
2005.
12. Suprenant, B.A., Laying Masonry in Cold Weather, The Magazine of Masonry Construction, Vol. 1, No.
9, Addison, IL, December 1988.
13. Van der Klugt, L.J .A.R., Frost Damage to the Pointing and Laying Mortar of Clay Brick Masonry, TNO
Building Construction and Research, 9th International Brick/Block Masonry Conference, Rijswijk, The
Netherlands, October 1991.
www.gobrick.com | Brick Industry Association | TN 1 | Cold and Hot Weather Construction | Page 9 of 9

Technical Notes 2 - Glossary of Terms Relating to Brick Masonry
Jan/Feb 1975 (Reissued Mar. 1999)
ABSORPTION: The weight of water a brick unit absorbs, when immersed in either cold or boiling water for a stated
length of time. Expressed as a percentage of the weight of the dry unit. See ASTM Specification C 67.
ADMIXTURES: Materials added to mortar to impart special properties to the mortar.
ANCHOR: A piece or assemblage, usually metal, used to attach building parts (e.g., plates, joists, trusses, etc.) to
masonry or masonry materials.
ANSI: American National Standards Institute.
ARCH: A curved compressive structural member, spanning openings or recesses; also built flat.
Back Arch: A concealed arch carrying the backing of a wall where the exterior facing is carried by a lintel.
Jack Arch: One having horizontal or nearly horizontal upper and lower surfaces. Also called flat or straight arch.
Major Arch: Arch with spans greater than 6 ft and equivalent to uniform loads greater than 1000 lb. per ft. Typically
known as Tudor arch, semicircular arch, Gothic arch or parabolic arch. Has rise to span ratio greater than 0.15.
Minor Arch: Arch with maximum span of 6 ft and loads not exceeding 1000 lb. per ft. Typically known as jack arch,
segmental arch or multicentered arch. Has rise to span ratio less than or equal to 0.15.
Relieving Arch: One built over a lintel, flat arch, or smaller arch to divert loads, thus relieving the lower member
from excessive loading. Also known as discharging or safety arch.
Trimmer Arch: An arch, usually a low rise arch of brick, used for supporting a fireplace hearth.
ASHLAR MASONRY: Masonry composed of rectangular units of burned clay or shale, or stone, generally larger in
size than brick and properly bonded, having sawed, dressed or squared beds, and joints laid in mortar. Often the
unit size varies to provide a random pattern, random ashlar.
ASHRAE: American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc.
ASTM: American Society for Testing and Materials.
BACK FILLING: 1. Rough masonry built behind a facing or between two faces. 2. Filling over the extrados of an
arch. 3. Brickwork in spaces between structural timbers, sometimes called brick nogging.
BACKUP: That part of a masonry wall behind the exterior facing.
BAT: A piece of brick.
BATTER: Recessing or sloping masonry back in successive courses; the opposite of corbel.
BED JOINT: The horizontal layer of mortar on which a masonry unit is laid.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
1 of 10 9/13/2009 12:29 PM
BELT COURSE: A narrow horizontal course of masonry, sometimes slightly projected such as window sills which
are made continuous. Sometimes called string course or sill course.
BLOCKING: A method of bonding two adjoining or intersecting walls, not built at the same time, by means of offsets
whose vertical dimensions are not less than 8 in.
BOND: 1. Tying various parts of a masonry wall by lapping units one over another or by connecting with metal ties.
2. Patterns formed by exposed faces of units. 3. Adhesion between mortar or grout and masonry units or
reinforcement.
BOND BEAM: Course or courses of a masonry wall grouted and usually reinforced in the horizontal direction.
Serves as horizontal tie of wall, bearing course for structural members or as a flexural member itself.
BOND COURSE: The course consisting of units which overlap more than one wythe of masonry.
BONDER: A bonding unit. See Header.
BREAKING JOINTS: Any arrangement of masonry units which prevents continuous vertical joints from occurring in
adjacent courses.
BRICK: A solid masonry unit of clay or shale, formed into a rectangular prism while plastic and burned or fired in a
kiln.
Acid-Resistant Brick: Brick suitable for use in contact with chemicals, usually in conjunction with acid-resistant
mortars.
Adobe Brick: Large roughly-molded, sun-dried clay brick of varying size.
Angle Brick: Any brick shaped to an oblique angle to fit a salient corner.
Arch Brick: 1. Wedge-shaped brick for special use in an arch. 2. Extremely hard-burned brick from an arch of a
scove kiln.
Building Brick: Brick for building purposes not especially treated for texture or color. Formerly called common
brick. See ASTM Specification C 62.
Clinker Brick: A very hard-burned brick whose shape is distorted or bloated due to nearly complete vitrification.
Common Brick: See Building Brick.
Dry-Press Brick: Brick formed in molds under high pressures from relatively dry clay (5 to 7 percent moisture
content).
Economy Brick: Brick whose nominal dimensions are 4 by 4 by 8 in.
Engineered Brick: Brick whose nominal dimensions are 4 by 3.2 by 8 in.
Facing Brick: Brick made especially for facing purposes, often treated to produce surface texture. They are made
of selected clays, or treated, to produce desired color. See ASTM Specification C 216.
Fire Brick: Brick made of refractory ceramic material which will resist high temperatures.
Floor Brick: Smooth dense brick, highly resistant to abrasion, used as finished floor surfaces. See ASTM
Specification C 410.
Gauged Brick: 1. Brick which have been ground or otherwise produced to accurate dimensions. 2. A tapered arch
brick.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
2 of 10 9/13/2009 12:29 PM
Hollow Brick: A masonry unit of clay or shale whose net cross-sectional area in any plane parallel to the bearing
surface is not less than 60 percent of its gross cross-sectional area measured in the same plane. See ASTM
Specification C 652.
Jumbo Brick: A generic term indicating a brick larger in size than the standard. Some producers use this term to
describe oversize brick of specific dimensions manufactured by them.
Norman Brick: A brick whose nominal dimensions are 4 by 2 2/3 by 12 in.
Paving Brick: Vitrified brick especially suitable for use in pavements where resistance to abrasion is important. See
ASTM Specification C 7.
Roman Brick: Brick whose nominal dimensions are 4 by 2 by 12 in.
Salmon Brick: Generic term for under-burned brick which are more porous, slightly larger, and lighter colored than
hard-burned brick. Usually pinkish-orange color.
"SCR Brick" (Reg U.S. Pat Off., SCPI (BIA)): See SCR (Reg U.S. Pat. Off., SCPI (BIA)).
Sewer Brick: Low absorption, abrasive-resistant brick intended for use in drainage structures. See ASTM
Specification C 32.
Soft-Mud Brick: Brick produced by molding relatively wet clay (20 to 30 percent moisture). Often a hand process.
When insides of molds are sanded to prevent sticking of clay, the product is sand-struck brick. When molds are
wetted to prevent sticking, the product is water-struck brick.
Stiff-Mud Brick: Brick produced by extruding a stiff but plastic clay (12 to 15 percent moisture) through a die.
BRICK AND BRICK: A method of laying brick so that units touch each other with only enough mortar to fill surface
irregularities.
BRICK GRADE: Designation for durability of the unit expressed as SW for severe weathering, MW for moderate
weathering, or NW for negligible weathering. See ASTM Specifications C 216, C 62 and C 652.
BRICK TYPE: Designation for facing brick which controls tolerance, chippage and distortion. Expressed as FBS,
FBX and FBA for solid brick, and HBS, HBX, HBA and HBB for hollow brick. See ASTM Specifications C 216 and C
652.
BUTTERING: Placing mortar on a masonry unit with a trowel.
CAPACITY INSULATION: The ability of masonry to store heat as a result of its mass, density and specific heat.
C/B RATIO: The ratio of the weight of water absorbed by a masonry unit during immersion in cold water to weight
absorbed during immersion in boiling water. An indication of the probable resistance of brick to freezing and thawing.
Also called saturation coefficient. See ASTM Specification C 67.
CENTERING: Temporary formwork for the support of masonry arches or lintels during construction. Also called
center(s).
CERAMIC COLOR GLAZE: An opaque colored glaze of satin or gloss finish obtained by spraying the clay body with
a compound of metallic oxides, chemicals and clays. It is burned at high temperatures, fusing glaze to body making
them inseparable. See ASTM Specification C 126.
CHASE: A continuous recess built into a wall to receive pipes, ducts, etc.
CLAY: A natural, mineral aggregate consisting essentially of hydrous aluminum silicate; it is plastic when sufficiently
wetted, rigid when dried and vitrified when fired to a sufficiently high temperature.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
3 of 10 9/13/2009 12:29 PM
CLAY MORTAR-MIX: Finely ground clay used as a plasticizer for masonry mortars.
CLEAR CERAMIC GLAZE: Same as Ceramic Color Glaze except that it is translucent or slightly tinted, with a gloss
finish.
CLIP: A portion of a brick cut to length.
CLOSER: The last masonry unit laid in a course. It may be whole or a portion of a unit.
CLOSURE: Supplementary or short length units used at corners or jambs to maintain bond patterns.
COLLAR JOINT: The vertical, longitudinal joint between wythes of masonry.
COLUMN: A vertical member whose horizontal dimension measured at right angles to the thickness does not exceed
three times its thickness.
COPING: The material or masonry units forming a cap or finish on top of a wall, pier, pilaster, chimney, etc. It
protects masonry below from penetration of water from above.
CORBEL: A shelf or ledge formed by projecting successive courses of masonry out from the face of the wall.
COURSE: One of the continuous horizontal layers of units, bonded with mortar in masonry.
CULLS: Masonry units which do not meet the standards or specifications and have been rejected.
DAMP COURSE: A course or layer of impervious material which prevents capillary entrance of moisture from the
ground or a lower course. Often called damp check.
DAMPPROOFING: Prevention of moisture penetration by capillary action.
DOG'S TOOTH: Brick laid with their corners projecting from the wall face.
DRIP: A projecting piece of material, shaped to throw off water and prevent its running down the face of wall or
other surface.
EBM: See Engineered Brick Masonry.
ECCENTRICITY: The normal distance between the centroidal axis of a member and the parallel resultant load.
e1/e2: Ratio of virtual eccentricities occurring at the ends of a column or wall under design. The absolute value is
always less than or equal to 1.0.
EFFECTIVE HEIGHT: The height of a member to be assumed for calculating the slenderness ratio.
EFFECTIVE THICKNESS: The thickness of a member to be assumed for calculating the slenderness ratio.
EFFLORESCENCE: A powder or stain sometimes found on the surface of masonry, resulting from deposition of
water-soluble salts.
ENGINEERED BRICK MASONRY: Masonry in which design is based on a rational structural analysis.
FACE: 1. The exposed surface of a wall or masonry unit. 2. The surface of a unit designed to be exposed in the
finished masonry.
FACING: Any material, forming a part of a wall, used as a finished surface.
FIELD: The expanse of wall between openings, corners, etc., principally composed of stretchers.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
4 of 10 9/13/2009 12:29 PM
FILTER BLOCK: A hollow, vitrified clay masonry unit, sometimes salt-glazed, designed for trickling filter floors in
sewage disposal plants. See ASTM Specification C 159.
FIRE CLAY: A clay which is highly resistant to heat without deforming and used for making brick.
FIRE RESISTIVE MATERIAL: See Non-combustible Material.
FIREPROOFING: Any material or combination protecting structural members to increase their fire resistance.
FLASHING: 1. A thin impervious material placed in mortar joints and through air spaces in masonry to prevent water
penetration and/or provide water drainage. 2. Manufacturing method to produce specific color tones.
FROG: A depression in the bed surface of a brick. Sometimes called a panel.
FURRING: A method of finishing the interior face of a masonry wall to provide space for insulation, prevent moisture
transmittance, or to provide a level surface for finishing.
GROUNDS: Nailing strips placed in masonry walls as a means of attaching trim or furring.
GROUT: Mixture of cementitious material and aggregate to which sufficient water is added to produce pouring
consistency without segregation of the constituents.
High-Lift Grouting: The technique of grouting masonry in lifts up to 12 ft.
Low-Lift Grouting: The technique of grouting as the wall is constructed.
HACKING: 1. The procedure of stacking brick in a kiln or on a kiln car. 2. Laying brick with the bottom edge set in
from the plane surface of the wall.
HARD-BURNED: Nearly vitrified clay products which have been fired at high temperatures. They have relatively low
absorptions and high compressive strengths.
HEAD JOINT: The vertical mortar joint between ends of masonry units. Often called cross joint.
HEADER: A masonry unit which overlaps two or more adjacent wythes of masonry to tie them together. Often called
bonder.
Blind Header: A concealed brick header in the interior of a wall, not showing on the faces.
Clipped Header: A bat placed to look like a header for purposes of establishing a pattern. Also called a false
header.
Flare Header: A header of darker color than the field of the wall.
HEADING COURSE: A continuous bonding course of header brick. Also called header course.
INITIAL RATE OF ABSORPTlON: The weight of water absorbed expressed in grams per 30 sq. in. of contact
surface when a brick is partially immersed for one minute. Also called suction. See ASTM Specification C 67.
IRA: See Initial Rate of Absorption.
KILN: A furnace oven or heated enclosure used for burning or firing brick or other clay material.
Kiln Run: Brick from one kiln which have not been sorted or graded for size or color variation.
KING CLOSER: A brick cut diagonally to have one 2 in. end and one full width end.
LATERAL SUPPORT: Means whereby walls are braced either vertically or horizontally by columns, pilasters, cross
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
5 of 10 9/13/2009 12:29 PM
walls, beams, floors, roofs, etc.
LEAD: The section of a wall built up and racked back on successive courses. A line is attached to leads as a guide
for constructing a wall between them.
LIME, HYDRATED: Quicklime to which sufficient water has been added to convert the oxides to hydroxides.
LIME PUTTY: Hydrated lime in plastic form ready for addition to mortar.
LINTEL: A beam placed over an opening in a wall.
MASONRY: Brick, stone, concrete, etc., or masonry combinations thereof, bonded with mortar.
MASONRY CEMENT: A mill-mixed cementitious material to which sand and water must be added. See ASTM C 91.
MASONRY UNIT: Natural or manufactured building units of burned clay, concrete, stone, glass, gypsum, etc.
Hollow Masonry Unit: One whose net cross-sectional area in any plane parallel to the bearing surface is less than
75 percent of the gross.
Modular Masonry Unit: One whose nominal dimensions are based on the 4 in. module.
Solid Masonry Unit: One whose net cross-sectional area in every plane parallel to the bearing surface is 75
percent or more of the gross.
MORTAR: A plastic mixture of cementitious materials, fine aggregate and water. See ASTM Specifications C 270, C
476 or BIA M1-72.
Fat Mortar: Mortar containing a high percentage of cementitious components. It is a sticky mortar which adheres to
a trowel.
High-Bond Mortar: Mortar which develops higher bond strengths with masonry units than normally developed with
conventional mortar.
Lean Mortar: Mortar which is deficient in cementitious components, it is usually harsh and difficult to spread.
NOMINAL DIMENSION: A dimension greater than a specified masonry dimension by the thickness of a mortar joint,
but not more than 1/2 in.
NON-COMBUSTIBLE MATERIAL: Any material which will neither ignite nor actively support combustion in air at a
temperature of 1200 F when exposed to fire.
OVERHAND WORK: Laying brick from inside a wall by men standing on a floor or on a scaffold.
PARGETING: The process of applying a coat of cement mortar to masonry. Often spelled and/or pronounced
parging.
PARTITION: An interior wall, one story or less in height.
PICK AND DIP: A method of laying brick whereby the bricklayer simultaneously picks up a brick with one hand and,
with the other hand, enough mortar on a trowel to lay the brick. Sometimes called the Eastern or New England
method.
PIER: An isolated column of masonry.
PILASTER: A wall portion projecting from either or both wall faces and serving as a vertical column and/or beam.
PLUMB RULE: This is a combination plumb rule and level. It is used in a horizontal position as a level and in a
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
6 of 10 9/13/2009 12:29 PM
vertical position as a plumb rule. They are made in lengths of 42 and 48 in., and short lengths from 12 to 24 in.
POINTING: Troweling mortar into a joint after masonry units are laid.
PREFABRICATED BRICK MASONRY: Masonry construction fabricated in a location other than its final inservice
location in the structure. Also known as preassembled, panelized and sectionalized brick masonry.
PRISM: A small masonry assemblage made with masonry units and mortar. Primarily used to predict the strength of
full scale masonry members.
QUEEN CLOSER: A cut brick having a nominal 2 in. horizontal face dimension.
QUOIN: A projecting right angle masonry corner.
RACKING: A method entailing stepping back successive courses of masonry.
RAGGLE: A groove in a joint or special unit to receive roofing or flashing.
RBM: Reinforced brick masonry
REINFORCED MASONRY: Masonry units, reinforcing steel, grout and/or mortar combined to act together in
resisting forces.
RETURN: Any surface turned back from the face of a principal surface.
REVEAL: That portion of a jamb or recess which is visible from the face of a wall.
ROWLOCK: A brick laid on its face edge so that the normal bedding area is visible in the wall face. Frequently
spelled rolok.
SALT GLAZE: A gloss finish obtained by thermochemical reaction between silicates of clay and vapors of salt or
chemicals.
SATURATION COEFFICIENT: See C/B Ratio.
SCR (Reg U.S. Pat Off., SCPI (BIA)): Structural Clay Research (trademark Of the Structural Clay Products
Institute, BIA).
"SCR acoustile" (Reg U.S. Pat Off., SCPI (BIA) Pat. No 3,001,6O2): A side-construction two-celled facing tile,
having a perforated face backed with glass wool for acoustical purposes.
"SCR brick" (Reg U.S. Pat Off., SCPI (BIA)): Brick whose nominal dimensions are 6 by 2 2/3 by 12 in. (Reg U.S.
Pat Off., SCPI (BIA)):
"SCR building panel" (Reg U S. Pat Off., SCPI (BIA) Pat. No. 3,248,836): Prefabricated, structural ceramic
panels, approximately 2 1/2 in. thick.
"SCR insulated cavity wall" (Reg U.S. Pat Off., SCPI (BIA)): Any cavity wall containing insulation which meets rigid
criteria established by the Structural Clay Products Institute (BIA).
"SCR masonry process" (Reg. U.S. Pat Off., SCPI (BIA)): A construction aid providing greater efficiency, better
workmanship and increased production in masonry construction. It utilizes story poles, marked lines and adjustable
scaffolding.
SHALE: Clay which has been subjected to high pressures until it has hardened.
SHOVED JOINTS: Vertical joints filled by shoving a brick against the next brick when it is being laid in a bed of
mortar.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
7 of 10 9/13/2009 12:29 PM
SLENDERNESS RATIO: Ratio of the effective height of a member to its effective thickness.
SLUSHED JOINTS: Vertical joints filled, after units are laid, by "throwing" mortar in with the edge of a trowel.
(Generally, not recommended.)
SOAP: A masonry unit of normal face dimensions, having a nominal 2 in. thickness.
SOFFIT: The underside of a beam, lintel or arch.
SOFT-BURNED: Clay products which have been fired at low temperature ranges, producing relatively high
absorptions and low compressive strengths.
SOLAR SCREEN: A perforated wall used as a sunshade.
SOLDIER: A stretcher set on end with face showing on the wall surface.
SPALL: A small fragment removed from the face of a masonry unit by a blow or by action of the elements.
STACK: Any structure or part thereof which contains a flue or flues for the discharge of gases.
STORY POLE: A marked pole for measuring masonry coursing during construction.
STRETCHER: A masonry unit laid with its greatest dimension horizontal and its face parallel to the wall face.
STRINGING MORTAR: The procedure of spreading enough mortar on a bed to lay several masonry units.
STRUCK JOINT: Any mortar joint which has been finished with a trowel.
SUCTION: See Initial Rate of Absorption.
TEMPER: To moisten and mix clay, plaster or mortar to a proper consistency.
TIE: Any unit of material which connects masonry to masonry or other materials. See Wall Tie.
TOOLING: Compressing and shaping the face of a mortar joint with a special tool other than a trowel.
TOOTHING: Constructing the temporary end of a wall with the end stretcher of every alternate course projecting.
Projecting units are toothers.
TRADITIONAL MASONRY: Masonry in which design is based on empirical rules which control minimum thickness,
lateral support requirements and height without a structural analysis.
TUCK POINTING: The filling in with fresh mortar of cut-out or defective mortar joints in masonry.
VENEER: A single wythe of masonry for facing purposes, not structurally bonded.
VIRTUAL ECCENTRICITY: The eccentricity of a resultant axial load required to produce axial and bending stresses
equivalent to those produced by applied axial loads and moments. It is normally found by dividing the moment at a
section by the summation of axial loads occurring at that section.
VITRIFICATION: The condition resulting when kiln temperatures are sufficient to fuse grains and close pores of a
clay product, making the mass impervious.
WALL: A vertical member of a structure whose horizontal dimension measured at right angles to the thickness
exceeds three times its thickness.
Apron Wall: That part of a panel wall between window sill and wall support.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
8 of 10 9/13/2009 12:29 PM
Area Wall: 1. The masonry surrounding or partly surrounding an area. 2. The retaining wall around basement
windows below grade.
Bearing Wall: One which supports a vertical load in addition to its own weight.
Cavity Wall: A wall built of masonry units so arranged as to provide a continuous air space within the wall (with or
without insulating material), and in which the inner and outer wythes of the wall are tied together with metal ties.
Composite Wall: A multiple-wythe wall in which at least one of the wythes is dissimilar to the other wythe or wythes
with respect to type or grade of masonry unit or mortar
Curtain Wall: An exterior non-loadbearing wall not wholly supported at each story. Such walls may be anchored to
columns, spandrel beams, floors or bearing walls, but not necessarily built between structural elements.
Dwarf Wall: A wall or partition which does not extend to the ceiling.
Enclosure Wall: An exterior non-bearing wall in skeleton frame construction. It is anchored to columns, piers or
floors, but not necessarily built between columns or piers nor wholly supported at each story.
Exterior Wall: Any outside wall or vertical enclosure of a building other than a party wall.
Faced Wall: A composite wall in which the masonry facing and backings are so bonded as to exert a common
reaction under load.
Fire Division Wall: Any wall which subdivides a building so as to resist the spread of fire. It is not necessarily
continuous through all stories to and above the roof.
Fire Wall: Any wall which subdivides a building to resist the spread of fire and which extends continuously from the
foundation through the roof.
Foundation Wall: That portion of a loadbearing wall below the level of the adjacent grade, or below first floor
beams or joists.
Hollow Wall: A wall built of masonry units arranged to provide an air space within the wall. The separated facing
and backing are bonded together with masonry units.
Insulated Cavity Wall: See "SCR insulated cavity wall".
Loadbearing Wall: A wall which supports any vertical load in addition to its own weight.
Non-Loadbearing Wall: A wall which supports no vertical load other than its own weight.
Panel Wall: An exterior, non-loadbearing wall wholly supported at each story.
Parapet Wall: That part of any wall entirely above the roof line.
Party Wall: A wall used for joint service by adjoining buildings.
Perforated Wall: One which contains a considerable number of relatively small openings. Often called pierced wall
or screen wall.
Shear Wall: A wall which resists horizontal forces applied in the plane of the wall.
Single Wythe Wall: A wall containing only one masonry unit in wall thickness.
Solid Masonry Wall: A wall built of solid masonry units, laid contiguously, with joints between units completely filled
with mortar or grout.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
9 of 10 9/13/2009 12:29 PM
Spandrel Wall: That part of a curtain wall above the top of a window in one story and below the sill of the window in
the story above.
Veneered Wall: A wall having a facing of masonry units or other weather-resisting non-combustible materials
securely attached to the backing, but not so bonded as to intentionally exert common action under load.
WALL PLATE: A horizontal member anchored to a masonry wall to which other structural elements may be
attached. Also called head plate.
WALL TIE: A bonder or metal piece which connects wythes of masonry to each other or in other materials.
WALL TIE, CAVITY: A rigid, corrosion-resistant metal tie which bonds two wythes of a cavity wall. It is usually steel,
3/16 in. in diameter and formed in a "Z" shape or a rectangle.
WALL TIE, VENEER: A strip or piece of metal used to tie a facing veneer to the backing.
WATER RETENTIVITY: That property of a mortar which prevents the rapid loss of water to masonry units of high
suction. It prevents bleeding or water gain when mortar is in contact with relatively impervious units.
WATER TABLE: A projection of lower masonry on the outside of the wall slightly above the ground. Often a damp
course is placed at the level of the water table to prevent upward penetration of ground water
WATERPROOFING: Prevention of moisture flow through masonry due to water pressure.
WEEP HOLES: Openings placed in mortar joints of facing material at the level of flashing, to permit the escape of
moisture.
WITH INSPECTION: Masonry designed with the higher stresses allowed under EBM. Requires the establishing of
procedures on the job to control mortar mix, workmanship and protection of masonry materials.
WITHOUT INSPECTION: Masonry designed with the reduced stresses allowed under EBM.
WYTHE: 1. Each continuous vertical section of masonry one unit in thickness. 2. The thickness of masonry
separating flues in a chimney. Also called withe or tier.
ABSORPTION: The weight of water a brick unit absorbs, when immersed ... http://www.gobrick.com/BIA/technotes/t2.htm
10 of 10 9/13/2009 12:29 PM

Technical Notes 3 - Overview of Building Code Requirements for Masonry Structures (ACI
530-02/ASCE 5-02/TMS 402-02) and Specification for Masonry Structures (ACI
530.1-02/ASCE 6-02/TMS 602-02)
July 2002
Abstract: This Technical Notes provides a review of the national masonry design standard, ACI
530/ASCE 5/TMS 402, and its accompanying masonry specification, ACI 530.1/ASCE 6/TMS
602. New provisions and revisions of existing standards for masonry design are emphasized.
Subjects discussed pertaining to the design standard are: allowable stress and strength design of
unreinforced and reinforced masonry, prestressed masonry, empirical design, glass block
masonry, masonry veneer, quality assurance, and seismic provisions. Items addressed for the
masonry specification are: requirements checklist and submittals, masonry quality assurance and
inspection requirements, reinforcement and metal accessories, erection tolerances, construction
procedures and grouting requirements.
Key Words: adhered veneer, allowable stress design, anchored veneer, building code, design
standard, empirical design, inspection, prestressed masonry, specification, strength design.

INTRODUCTION
The American Concrete Institute (ACI), American Society of Civil Engineers (ASCE), and The
Masonry Society (TMS) promulgate a national consensus standard for the structural design of
masonry elements and a standard specification for masonry construction. These standards are
titled the Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402) and
the Specification for Masonry Structures (ACI 530.1/ASCE 6/TMS 602). They were developed to
consolidate and advance existing standards for the design and construction of masonry.
This Technical Notes, the first in a series, discusses various sections of the Building Code
Requirements for Masonry Structures and the Specification for Masonry Structures in brief
detail. Emphasis is placed on the new requirements in the 2002 edition of the standards.
Changes from prior masonry standards dealing with the design of brick masonry structures are
also presented. Other Technical Notes in this series provide material and section properties of
brick masonry members and more extensive discussion of the requirements of these standards.
For more information about the requirements of these standards and examples of their application,
the reader is referred to the Masonry Designer's Guide (MDG). The MDG is published by The
Masonry Society and contains an extensive number of design examples that illustrate the proper
application of the MSJC Code and Specification requirements.
In this Technical Notes, the Building Code Requirements for Masonry Structures and the
Specification for Masonry Structures are referred to as the Masonry Standards Joint Committee
(MSJC) Code and Specification, respectively. The pertinent section and article numbers from the
MSJC Code and Specification, are stated in parentheses following the discussion of particular
topics for quick reference.

HISTORY AND DEVELOPMENT
t45 http://www.gobrick.com/BIA/technotes/t3.htm
1 of 14 9/13/2009 12:30 PM
The development of this single masonry standard for the design and construction industry began in
1977. At that time, there were several design standards for masonry. These standards did not
have consistent requirements. It was difficult for engineers and architects to select the
appropriate design criteria for masonry elements. Concerned individuals representing masonry
materials and the design profession saw the need for a single, national consensus standard for the
design and construction of all types of masonry.
In 1977, ACI and ASCE agreed to jointly develop a consensus standard for masonry design with
the support of the masonry industry. The MSJC was formed with a balanced membership of
building officials, contractors, university professors, consultants, material producers and designers
who are members of ACI or ASCE. The Masonry Society joined as a sponsoring organization in
1991. Currently, the MSJC is comprised of over eighty regular (voting) and forty associate
members. The MSJC Code and Specification are available from each of the sponsoring
organizations or from the Brick Industry Association.
Changes to the MSJC Code and Specification are written, balloted and approved within the
MSJC. A review by the sponsoring organizations' technical activity committees follows. In order
to obtain a national consensus, the approved draft undergoes a public review. Approval by the
MSJC of the first edition of the MSJC Code and Specification occurred in June 1986. Public
review began in 1988 with the final approval of the 1988 MSJC Code and Specification in August
1989.
Commentaries for the MSJC Code and Specification were also developed. These documents
provide background information on the design and specification provisions. Considerations of the
MSJC members in determining requirements and references to research papers and articles are
included in the commentaries for further information.
The MSJC Code, Specification and Commentaries are revised on a three- or four-year cycle. The
first revision was issued in 1992. Most of the changes were editorial in nature or clarified intent or
omissions. In 1995 new chapters on glass unit masonry and anchored masonry veneers were
added, and the MSJC Specification was reformatted. Metric conversions were added throughout
the standards in accordance with the metrication policy of ASCE in addition to an index of key
words. The 1999 edition includes a number of significant changes. The MSJC Code and its
Commentary were reformatted. A chapter on prestressed masonry, a section on adhered veneer
and a quality assurance program were added. Other changes in the MSJC Code and
Specification include new design values for elastic moduli and masonry compressive strength and
the inclusion of mortar cement. In the 2002 edition there were significant changes to the seismic
design provisions, with prescriptive requirements for specific shear wall types. A chapter on
strength design was added. Other minor changes are documented in this Technical Notes.

Building Code Acceptance
The MSJC Code is to be adopted by a model building code and, subsequently, by a local
jurisdiction. State and local building code committees are encouraged to adopt the model building
codes which include the MSJC Code for the design of masonry. With adoption of the MSJC Code,
the Specification is automatically adopted because the MSJC Code requires that materials and
construction comply with the MSJC Specification. The local jurisdiction has the responsibility for
enforcement and compliance of masonry construction to the MSJC Specification once it is
adopted.
Two of the previous model building code organizations, the Standard Building Code Congress
International (SBCCI) and the Building Officials and Code Administrators (BOCA), chose to include
the MSJC Code in their documents. This adoption by reference began in 1988 and1989,
respectively. The International Council of Building Officials (ICBO) chose to maintain masonry
design criteria within the Uniform Building Code itself, rather than adopting the MSJC standards
by reference. However, many of the masonry design and construction requirements of the
t45 http://www.gobrick.com/BIA/technotes/t3.htm
2 of 14 9/13/2009 12:30 PM
Uniform Building Code have been changed over the last several years to be consistent with the
requirements of the MSJC Code and Specification.
The International Code Council (ICC) was formed by the three existing code organizations
(SBCCI), (BOCA) and (ICBO) with the charge to produce a single set of codes, referred to as the
I-codes. Two I-codes that are important to the brick industry are the International Building Code
(IBC) and International Residential Code (IRC). The National Fire Protection Association (NFPA)
is also developing another building code called NFPA 5000. The I-codes and NFPA 5000
reference the 2002 MSJC Code.

Benefits
The MSJC Code and Specification have had positive results; the design and construction
community has become more confident with their use. Designers have one national standard that
covers nearly all types of masonry construction. Architects are able to prepare and submit
complete, concise specifications more easily. Contractors have more consistent and better quality
specifications for projects. Owners obtain more uniform quality of masonry. Other benefits
presented by the MSJC Code and Specification are:
Nearly all forms of masonry are covered, including unreinforced, reinforced and prestressed
masonry, glass unit masonry, and adhered and anchored veneer masonry.
1.
Requirements for all masonry materials are covered, including clay and shale brick,
concrete block, stone, glass unit, mortar, grout and metal accessories.
2.
Differences in material properties are recognized and quantified. 3.
The same rational design procedures are utilized for clay and concrete masonry. 4.
Responsibilities and duties of the owner, designer, testing agency, and contractor are
clearly established.
5.
Quality assurance and inspection requirements are included. 6.
Design, materials and testing are the decision of the architect or engineer. 7.
Contract administration is easier. 8.
Since the introduction of the MSJC standards in 1988, there has been a shift in the masonry
design and construction communities. Designers and contractors use the MSJC Code and
Specification with more frequency. Indicative of this growth, the MSJC Code is now a required
reference for the Professional Engineer's Principles and Practice examination. The MSJC
Specification has placed greater demands on the masonry contractor with the use of masonry as a
structural material. Many requirements are performance related, which may require more site
inspection for verification of compliance. These demands are advantageous and vital to the
development of confidence that the masonry strengths assumed by the designer are met by the
constructed masonry.

THE MSJC CODE (ACI 530/ASCE 5/TMS 402)
The MSJC Code is the basis for masonry design by the architect or engineer. The provisions of
the MSJC Code will dictate the size and shape of masonry walls, beams, pilasters and columns.
Further, it influences the masonry materials the designer will require in the project specification. It
consists of seven chapters, which are listed below.
t45 http://www.gobrick.com/BIA/technotes/t3.htm
3 of 14 9/13/2009 12:30 PM
Chapter 1 - General Design Requirements for Masonry
Chapter 2 - Allowable Stress Design
Chapter 3 - Strength Design of Masonry (New Chapter)
Chapter 4 - Prestressed Masonry
Chapter 5 - Empirical Design of Masonry
Chapter 6 - Veneer
Chapter 7 - Glass Unit Masonry
Some relevant sections of the codes are discussed in this Technical Notes and are indicated in
parentheses for each of the chapters.

Chapter 1 - General Design Requirements for Masonry
Chapter 1 contains the scope of the minimum requirements for the design of any masonry
element. In this chapter, it states that the MSJC Code supplements the model building code
enforced in a jurisdiction. When the MSJC Code conflicts with the local building code, the local
building code governs. (1.1)
Project drawings and specifications must identify the individual responsible for their preparation.
Items required by the MSJC Code must be clearly marked such as: loads used in design, specified
compressive strength of masonry, reinforcement, anchors and ties with size and spacing, size and
location of all structural elements, provisions for differential movement, and size and location of
conduit, pipes and sleeves. Contract documents must include a quality assurance program. (1.2)
The MSJC Code permits alternative design methods from those stated in the MSJC Code. This is
to recognize new applications of masonry and different structural analysis techniques. (1.3)
Chapter 1 also includes the notation and definitions contained within the MSJC Code. Capital
letters are used for permitted stresses and lower case letters are used for calculated or applied
stresses. (1.5) For example, Fa is the notation for the allowable compressive stress due to axial
load, while fa denotes the calculated compressive stress due to axial load. The definitions are
specifically related to their meaning as used in the MSJC Code. Definitions in the MSJC Code are
coordinated with those in the MSJC Specification. Definitions of terms relating to strength design
of masonry and for prestressed masonry have been added. (1.6)
The following are brief summaries, highlights, of several sections within Chapter 1.
Section 1.7 - Loading. Service loads are used as the basis of design and are governed by the
building code that adopts the MSJC Code. If a building code is not enforced in the area under
consideration, then the MSJC Code requires that the load provisions of the 1993 edition of ASCE 7
Design Loads for Buildings and Other Structures apply to masonry structures. Allowable stresses
given in the MSJC Code are based on failure stresses with a factor of safety in the range of 2 to
5. The structural system must resist wind and earthquake loads and accommodate the resulting
deformations. (1.7.3) The effects of restraint of movement due to prestressing, vibrations, impact,
shrinkage, expansion, temperature changes, creep, unequal settlement of supports and differential
movement must also be considered in design. (1.7.4)
Section 1.8 - Material Properties. Material properties are included for both clay and concrete
masonry. The MSJC Code and Specification was the first national masonry standard to state
design coefficients for thermal expansion, moisture expansion, shrinkage and creep. For design
t45 http://www.gobrick.com/BIA/technotes/t3.htm
4 of 14 9/13/2009 12:30 PM
computations, the amount of shrinkage of brick masonry is taken as zero. The moduli of elasticity,
Em, of clay and concrete masonry is no longer based on the net area compressive strength of the
brick and the type of mortar used in construction. Em is now directly related to the specified
compressive strength of masonry, f'
m
. For clay masonry, E
m
is equal to 700 times f'
m
.
Alternately, Em may be determined by the chord modulus of elasticity taken between 0.05 and
0.33 of the maximum compressive strength of each prism determined by test in accordance with
Article 1.4 B.3 of the MSJC Specification. Refer to Technical Notes 18 Series for an extensive
discussion of differential movement of brick masonry elements. (1.8.2.2)
Section 1.9 - Section Properties. Section properties are used to determine stress computations.
Computations for stiffness, radius of gyration and flange design for intersecting walls are based on
the minimum net area of the section. This is normally the mortar-bedded area. When different
materials are combined in a single element, the transformed area must be used to account for
differences in elastic moduli of the dissimilar materials. Radius of gyration of the section, rather
than the minimum thickness, is used to determine the slenderness reduction for members in
compression. (1.9)
Section 1.10 - Deflection. Deflection limits are imposed for beams and lintels that support
unreinforced masonry. The deflection should not exceed the span length divided by 600 or 0.3 in.
(7.6 mm). Deflection of the masonry member should be calculated based upon uncracked section
properties. (1.10, 1.9.2)
Section 1.11 - Stack Bond Masonry. The MSJC Code requires that stack bond masonry be
reinforced with a prescriptive amount of horizontal reinforcement. This may be placed as joint
reinforcement or in bond beams spaced not more than 48 in. (1.2 m) on center vertically. (1.11)
Section 1.12 - Details of Reinforcement. The reinforcement detailing requirements given in this
chapter are similar to those for reinforced concrete under ACI 318, Building Code Requirements
for Reinforced Concrete. The maximum size of reinforcing bar permitted in masonry members,
designed by the allowable stress or empirical design methods, is a No. 11 (M #36) bar. Horizontal
joint reinforcement is permitted as structural reinforcement for the same design methods.
Placement limits for reinforcement include minimum grout spaces between the bars and masonry
units of 1/4 in. (6.4 mm) and 1/2 in. (12.7 mm) for fine and coarse grout, respectively. (1.12.2 -
1.12.3)
This section contains protection requirements for reinforcing steel. A minimum amount of masonry
cover is required, depending upon the exposure conditions. Corrosion protection is required for
joint reinforcement, wall ties, anchors and inserts in exterior walls. (1.12.4)
Minimum development lengths are stated for reinforcement. A 50 percent increase is
recommended for epoxy coated bars. (2.1.10.2) Standard hooks, minimum bend diameters, and
splice requirements are consistent with those for reinforced concrete members. (1.12.5, 1.12.6)
Chapter 3 contains variations in some of these requirements when strength design is used.
Section 1.13 - Seismic Design Requirements. These requirements apply to the design and
construction of all masonry, except glass unit masonry and masonry veneers, for all Seismic
Design Categories (SDC) as defined in ASCE 7-98. Early editions of the MSJC included seismic
design information as optional information in the Appendix and based the requirements on Seismic
Zones. Since 1995, the seismic requirements are mandatory parts of the Code. Seismic
provisions for masonry veneers are found in Chapter 6, Veneers.
Special seismic requirements in Section 1.13 are invoked by SDC. The requirements are additive
for each higher SDC. For example, buildings in category D must meet all the requirements for
buildings in categories A, B and C, plus the additional requirements stated in Section 1.13 for
buildings in category D.
Five types of shear walls that serve as the lateral force-resisting system are described. Each has
a required design method and prescriptive reinforcement requirements, see Table 1. Their use is
permitted by the seismic design category applicable to the structure under design.
t45 http://www.gobrick.com/BIA/technotes/t3.htm
5 of 14 9/13/2009 12:30 PM


TABLE 1
Requirement for Masonry Shear Walls Based on Shear Wall Designation
Shear Wall Designation Reinforcement Requirements Permitted SDC
Empirically Designed None SDC A
Ordinary Plain
(unreinforced)
None SDC A and B
Detailed Plain (unreinforced) Section 1.13.2.2.2.1 and
1.13.2.2.2.2
SDC A and B
Ordinary Reinforced Section 1.13.2.2.2.1 and
1.13.2.2.2.2
SDC A, B, and C
Intermediate Reinforced Section 1.13.2.2.4 SDC A. B, and C
Special Reinforced Section 1.13.2.2.5 SDC A, B, C, D, E and
F

In category A, the provisions of Chapters 2, 3, 4, or 5 of the MSJC Code apply. There is a
calculated story drift limit of 0.007 times the story height. Anchorage of masonry walls must meet
a minimum design force of 1000 times the effective peak velocity-related acceleration. (1.13.3)
For buildings in category B, the lateral force-resisting system must comply with the requirements
of Chapter 2, 3, or 4 of the MSJC Code. It cannot be designed in accordance with the empirical
requirements of Chapter 5. The lateral force-resisting system includes structural masonry
members such as columns, beams and shear walls. It does not include non-loadbearing elements,
such as partition walls. (1.13.4)
Masonry buildings in category C must meet more stringent requirements. Members that are not
part of the main lateral force-resisting system must be isolated so that they do not adversely affect
the response of the lateral force-resisting system. Connections are strengthened and minimum
amounts of reinforcement are required for shear walls and non-loadbearing masonry members in
order to provide more ductility to the structure. (1.13.5) Partition walls, screen walls and other
elements that are not designed to resist vertical or lateral loads other than their own weight must
be isolated from receiving these loads and designed to accommodate drift.
The special seismic provisions for categories D and E are still more restrictive. Minimum
reinforcement requirements are increased for all members. Type N mortar and masonry cement
mortars are not permitted for the lateral force-resisting system. (1.13.6, 1.13.7)
Section 1.14 - Quality Assurance. This section defines a quality assurance program with
different requirements based on the type of facility and method of design. Minimum tests,
submittals and inspection requirements are defined for three levels of quality assurance. (1.14.1)
The quality assurance program must include procedures for reporting, review and resolution of
noncompliances. (1.14.5) Qualifications for testing laboratories and for inspection agencies must
also be defined. (1.14.6)
The quality assurance program requires that each wythe of masonry and the grout, if present,
must meet or exceed the specified compressive strength of masonry, f'm. Compressive strength
of masonry must be verified in accordance with the provisions of the MSJC Specification. (1.14.2)
Section 1.15 - Construction. Construction of masonry must comply with the MSJC
Specification. Requirements for grouting are introduced in Section 1.15. The type of grout, either
t45 http://www.gobrick.com/BIA/technotes/t3.htm
6 of 14 9/13/2009 12:30 PM
fine or coarse, determines the minimum grout space dimensions and maximum grout pour height
permitted. New in the 2002 edition is the inclusion of a grout demonstration panel. The limits can
be exceeded if the panel indicates that the spaces are filled and adequately consolidated. Grout
must attain a minimum compressive strength of 2000 psi (13.8 MPa) at 28 days. (Table 1.15.1)
In addition, Section 1.15 contains provisions for pipes and conduits embedded in masonry
elements. The effect on structural performance of the opening caused by the embedded item
must be considered. Limitations on location, size, relative area and materials contained within
pipes and conduit are included. (1.15.2)
Chapter 2 - Allowable Stress Design
Allowable stress design (ASD) methodology has been used in masonry design for many years.
The ASD provisions of the MSJC Code are the most advanced to date for masonry members and
are reflective of the extensive amount of research and experience gained over the last century.
Chapter 2 of the MSJC Code states general provisions and establishes the scope of the rational
design requirements. The rational design provisions are based upon a few assumptions inherent in
the ASD approach, which are as follows:
Masonry materials are linearly elastic under service loads (materials rebound to original
position when unloaded, rather than deforming permanently).
1.
Stress is directly proportional to strain (applied load is directly proportional to
displacement).
2.
Masonry materials behave homogeneously (brick, mortar and grout behave as one element
rather than separately).
3.
Sections plane before bending remain plane after bending (flexural members do not warp). 4.
Service loads are used as the basis of allowable stress design. Allowable stresses given in the
MSJC Code are based on failure stresses with a factor of safety in the range of 2 to 5. Section
2.1.2 contains the loading combinations to be used for allowable stress design. For moment
strength design under Section 4.5.3.3.2, factored loads shall be combined as required by the
general building code. When the general building code does not provide load combinations,
structures or members shall use the most restrictive combinations of loads. (2.1.2)
The specified compressive strength of masonry, f'm, must be determined by the designer and
clearly stated in the contract documents. The specified compressive strength must be verified by
the contractor as required by the methods stipulated in the MSJC Specification. (2.1.3)
Anchor bolts consist of plate, headed and bent bar assemblies. Allowable loads for tension, shear
and combined tension and shear are given. Provisions for minimum embedment length are
provided to ensure proper transfer of load between the masonry and the anchor bolt. (2.1.4)
Refer to Technical Notes 44 for further discussion of the design of anchor bolts.
The MSJC Code requirements differentiate between multiwythe walls with respect to composite or
non-composite action. Composite action requires a rigid transfer of stress between wythes so
that the wythes act as a single element in resisting loads. The wythes must be bonded with a filled
collar joint and metal ties or with masonry headers. Prescriptive size and spacing limitations for
metal wall ties are taken from previous masonry standards. For multiwythe, composite walls,
criteria for allowable shear stresses at the interface between a wythe and a collar joint have been
introduced that were not included in previous masonry standards. These allowable shear stresses
are: a) 5 psi (34.5 kPa) for mortared collar joints, b) 10 psi (69.0 kPa) for grouted collar joints,
and c) the square root of the unit compressive strength of the header. (2.1.5.2.2)
When non-composite action occurs, each wythe is designed to individually resist the effects of
imposed loads. Loads are apportioned to wythes based upon their relative stiffnesses. As with
t45 http://www.gobrick.com/BIA/technotes/t3.htm
7 of 14 9/13/2009 12:30 PM
composite walls, prescriptive requirements for metal wall ties are based on past experience.
(2.1.5.3 ) Wall ties with drips are now prohibited.
Columns are isolated vertical members whose horizontal dimension at right angles to the thickness
does not exceed 3 times its thickness. Also, the member's height must be at least 3 times its
thickness. The minimum dimension of a column is 8 in. (203 mm) and the maximum ratio of
effective height to least nominal dimension (slenderness ratio) of a column is 25. Columns must
contain a minimum of four vertical reinforcing bars and a minimum amount of lateral ties. (2.1.6)
Pilasters are thickened elements of a wall which provide resistance to lateral loads or a
combination of axial and lateral loads. Design procedures consider the pilaster and wall to act
integrally, provided the two are properly bonded. Vertical reinforcement that is intended to resist
axial loads must be laterally tied in the same manner that is required for columns. (2.1.7)
Concentrated loads must be distributed over a prescribed length of wall. Requirements depend on
bond pattern, presence of bond beams and the width of the wall. The allowable bearing stress is
one-fourth of the specified compressive strength of masonry, but may be increased for smaller
bearing areas. (2.1.9)
Provisions for development of reinforcement are included. (2.1.10) Bars, hooks, welded wire
fabric, and splices are covered.
Section 2.2 - Unreinforced Masonry. Section 2.2 covers requirements for the design of masonry
structures in which tensile stresses in masonry are taken into consideration. This is known as
unreinforced (plain) masonry. Such members may, in fact, contain reinforcement for shrinkage or
other reasons, but this reinforcement is neglected in the structural design process.
The allowable axial compressive stress equation uses a different slenderness reduction factor from
that used in earlier masonry standards. The factor is a function of the radius of gyration of the
member's cross section, rather than its thickness. Additionally, the factor of safety changed from
5 in previous masonry standards to 4 in the MSJC Code. Unlike previous masonry design
standards, the MSJC Code does not place an arbitrary limit on the slenderness ratio of walls.
Rather, the slenderness reduction factor becomes very small for more slender walls. An equation
limiting the applied axial load to one-quarter of a modified Euler buckling load is included. The
classic Euler buckling load has been modified to reflect a member with negligible tensile strength.
The unity equation has been used to limit the combination of bending and axial load in masonry
design for many years. (2.2.3, 2.3.3)
Variables affecting flexural tension of masonry include the plane on which the stress acts, mortar
materials, unit cross-section, and presence of grout. The allowable flexural tension stresses for
grouted masonry normal to bed joints were modified in the 2002 edition. (2.2.3.2)
Allowable shear stresses are based upon a parabolic shear stress distribution rather than an
average shear stress distribution, as used in previous masonry standards. Consequently,
allowable shear stresses are approximately 1.5 times those in previous masonry standards. Four
allowable shear stresses for in-plane shear must be evaluated. No allowable shear stress values
are given for out-of-plane shear, but typically these same values for in-place shear are applied.
(2.2.5)
Section 2.3 - Reinforced Masonry. Section 2.3 contains requirements for the allowable stress
design of masonry elements neglecting the tensile strength of masonry. This is commonly termed
reinforced masonry. In this procedure, steel reinforcement is used to resist all tensile forces.
Reinforcement may also be required to resist shear forces. The MSJC Code does not prescribe a
minimum amount of reinforcement, except for masonry columns and for buildings in Seismic Design
Categories as given in Chapter 1. The size and placement of compressive, flexural and shear
reinforcement is determined by design requirements. (2.3.1) Allowable steel stresses are taken
from previous masonry standards. Reinforcement used to resist compressive stresses must be
laterally tied. (2.3.2.2)
When the applied shear stress exceeds the given allowable shear stress for reinforced masonry
t45 http://www.gobrick.com/BIA/technotes/t3.htm
8 of 14 9/13/2009 12:30 PM
without shear reinforcement, shear reinforcement is required. For reinforced masonry containing
shear reinforcement, allowable shear stresses are increased by a factor of 3.0 for flexural
members and 1.5 for shear walls. To use the increased allowable shear stresses, shear
reinforcement must be provided to resist 100 percent of the shear force. (2.3.5)
Chapter 3 - Strength Design of Masonry
This chapter is new in the 2002 edition of the MSJC Code. This chapter was developed from
research funded by the National Science Foundation and the masonry industry.
Strength design identifies the possible failure modes that the masonry element can exhibit. By
performing this type of analysis the engineer can preclude an undesirable failure. Strength design
provides for design of inelastic performance of masonry. The loads and stresses considered are
similar to those used in allowable stress design, but service level loads are replaced with strength
design loads and allowable stresses are replaced with nominal values based on research. The
required strength of the masonry must be greater than its nominal strength multiplied by a strength
reduction factor, . The strength reduction factors selected are similar to those used in concrete.
Strength design of masonry shall comply with the minimum requirements of this chapter. In
addition, the requirements of Chapter 1, Section 3.1, and either Section 3.2 or 3.3 also apply.
(3.1.1) The strength requirements are in accordance with the legally adopted building code.
When this information is not defined in the building code then the requirements of ASCE 7-98
govern. (3.1.2) Notations and definitions used in strength design are found in Sections 1.5 and
1.6, respectively.
The remainder of Chapter 3 covers design strength (3.1.3), strength reduction factors (3.1.4),
deformation requirements (3.1.5), headed and bent-bar anchor bolts (3.1.6), material properties
(3.1.7), reinforced masonry (3.2), and unreinforced (plain) masonry (3.3). Design equations are
similar to those for allowable stress design when possible. Perhaps the most significant difference
is in the development length. The strength design formula includes cover, bar size, and masonry
specified compressive strength as variables. This formula also applies to splices.
This chapter includes maximum reinforcement ratios chosen to prevent brittle failure of shear
walls. These are applied with specific limits on strain in the masonry and steel. There are also
dimensional limits for beams, piers, and columns.
It must be pointed out that Strength Design of Masonry may not be practical in many situations and
may in fact not provide the results a designer may seek.
Chapter 4 - Prestressed Masonry
Prestressed masonry is used to eliminate tensile stresses in masonry due to externally applied
loads. A controlled amount of precompression is applied to the masonry to offset the tensile
forces created under service loads. The use of prestressing is well documented in concrete
design and construction; however its use in masonry construction in the United States is limited.
The United Kingdom has a history of successful prestressed masonry construction for over two
decades.
The equipment for prestressed masonry is similar to that used in concrete construction. Some
proprietary systems have been developed specifically for use in prestressed masonry. Types of
structures that have utilized prestressed masonry in the United States include freestanding walls,
such as fences, bearing walls and masonry veneers designed to span between columns, rather
than span floor-to-floor.
Prestressing tendons placed in openings in the masonry may be grouted or ungrouted. The
tendons may be pre-tensioned or post-tensioned. Pre-tensioned tendons are stressed against
external abutments prior to placing the masonry. Post-tensioned tendons are stressed against the
masonry after it has been placed. Most construction applications to date have been
post-tensioned, ungrouted masonry because of the ease of construction and overall economy. As
t45 http://www.gobrick.com/BIA/technotes/t3.htm
9 of 14 9/13/2009 12:30 PM
a result, the MSJC Code focuses primarily on post-tensioned masonry.
Chapter 4 provides minimum requirements for the design of structures that are prestressed with
bonded or unbonded prestressing tendons. The general design requirements found in Chapter 1,
including seismic provisions, apply to prestressed masonry with a few modifications. (4.1)
Prestressed members are designed using elastic analysis and allowable stress design. A new
term, f'
mi
, is defined as the specified compressive strength of masonry at the time of transfer of
the prestress force. (4.2)
The remainder of Chapter 4 covers permissible stresses in the prestressing tendons, effective
prestress, axial compression and flexure, axial tension, shear, deflection, prestressing tendon
anchorages, couplers, end blocks, protection of prestressing tendons and accessories, and
development of bonded tendons.
Chapter 5 - Empirical Design of Masonry
Chapter 5 presents empirical requirements for masonry structures. These requirements are based
on past proven performance. Configuration of masonry structures for compliance with empirical
limits is a technique that predates rational design methods. The empirical provisions of previous
masonry standards have been modified and advanced in Chapter 5 to reflect contemporary
construction materials and methods. The requirements are essentially unchanged from the 1999
edition.
The empirical requirements in Chapter 5 may be applied to the following masonry elements:
The lateral force-resisting system for buildings in Seismic Design Categories (SDC) A, and
for other building elements in SDC A through C, as defined in ASCE 7-98. (5.1.2)
1.
Buildings subject to basic wind speed of 110 mph (145 km/hr) or less as defined by the
ASCE 7-98 standard. (5.1.2.2)
2.
Buildings not exceeding 35 ft (10.67 m) when the masonry walls are part of the main lateral
force-resisting system. (5.2)
3.
The empirical requirements may not be applied to structures resisting horizontal loads other than
those due to wind or seismic events, except that foundation walls may be as permitted in Section
5.6.3. The empirical requirements for foundation walls include limits on the height of backfill. There
are a number of restrictions on the backfill soil and the configuration of cross walls. (5.6.3.1) The
2002 Code also requires foundation piers to be a minimum of 8 in. (203 mm) in thickness. (5.6.4)
The empirical requirements of the MSJC Code are discussed in Technical Notes 42 Revised.
Chapter 6 - Veneers.
The requirements of Chapter 6 apply to masonry veneers. In the 2002 MSJC Code, provisions
address anchored masonry veneer and adhered masonry veneer. The requirements of this
chapter are especially important to the brick industry as the majority of brick produced in the
United States is used as veneer.
Section 6.2 - Anchored veneer. The majority of this chapter contains prescriptive requirements for
masonry veneer, but alternative design methods are permitted. (6.2.1) The prescriptive
requirements cannot be used in areas where the wind speed exceeds 110 mph (145 km/hr) as
given in ASCE 7-98. (6.2.2.1) Many of the requirements are based upon those found in Technical
Notes 28 Series on brick veneer walls and Technical Notes 44B on wall ties. (6.2.2.3-6.2.2.9)
Seismic requirements are included for buildings in SDC C, D, and E. (6.2.2.10)
Section 6.3 - Adhered veneer. Adhered veneer can be designed by the prescriptive requirements
contained in this section or by alternative design methods. (6.3.1) Prescriptive requirements found
in the 2002 MSJC Code are based on similar requirements that have been used in the Uniform
Building Code for over 30 years. These requirements limit unit size to no more than 2 5/8 in. (66.7
t45 http://www.gobrick.com/BIA/technotes/t3.htm
10 of 14 9/13/2009 12:30 PM
mm) in specified thickness, 36 in. (914 mm) in any face dimension and 5 ft
2
(0.46 m
2
) in total face
area. The weight of adhered veneer units is limited to15 lbs/ft
2
(718 Pa). (6.3.2)
Adhesion between the veneer units and the backing must have a shear strength of 50 psi (345
kPa) or greater based on gross unit surface area when tested in accordance with ASTM C 482.
Alternatively, adhered units may be applied using the procedure found in MSJC Specification
Article 3.3C. (6.3.2.4)
Chapter 7 - Glass Unit Masonry
Chapter 7 applies to glass unit masonry. The 2002 edition contains few changes from the 1999
version. The provisions are largely based upon those in the three previous model building codes.
Requirements are primarily prescriptive and empirical.
Maximum wall areas are imposed by a design wind pressure graph for standard units, 3 7/8 in.
(98.4 mm) thick. When 3 in. (76.2 mm) thick units are used, a maximum wind pressure of 20 psf
(958 Pa) is imposed and the maximum wall area is reduced. The size of interior wall panels is
limited to 250 ft2 (23.22 m2) and 150 ft2 (13.94 m2) for standard and thin units, respectively. (7.1,
7.2) Provisions regarding lateral support for panels limited to one unit wide or one unit high are
included. (7.3)
The MSJC Code also imposes requirements for expansion joints. (7.4)
Base surface treatment requires the surface on which glass unit masonry panels are placed to be
coated with an elastic waterproofing material. (7.5)
Glass unit masonry shall be built with Type S or N mortar. (7.6)
Glass unit masonry panels must contain a minimum amount of horizontal joint reinforcement. The
MSJC Code requires a minimum of two parallel W1.7 (MW11) wires spaced at 16 in. (406 mm)
o.c. vertically. Joint reinforcement is very important because the limitations on wall panel size are
based upon the failure of the reinforced section, rather than the first cracking strength of panels.
(7.7)

THE MSJC SPECIFICATION (ACI 530.1/ASCE 6/TMS 602)
The MSJC Specification is a reference standard that an architect or engineer may cite in the
contract documents for any project. The MSJC Specification contains requirements for the
contractor regarding materials, construction and quality assurance. The MSJC Code requires
compliance of construction of the masonry with the MSJC Specification, so it is an integral part of
the MSJC Code. The language is in imperative voice for ease of interpretation and enforcement.
The MSJC Specification should be referenced in the contract documents and may be modified as
required for the particular project.
The 2002 edition of the MSJC Specification consists of three components: a) Part 1 - General, b)
Part 2 - Products and c) Part 3 - Execution. The format was changed to the present one in 1995
to be more consistent with the Construction Specifications Institute's MASTERFORMAT.
Major changes in the 2002 edition relate to quality assurance and ease of use. Quality assurance
is established in conjunction with the MSJC Code and the MSJC Specification contains specific
instructions for the parties involved. The phrase When required was eliminated. Inclusion of this
phrase in earlier editions made it necessary for the user to extensively edit the MSJC Specification
for application to a particular project.
Requirements Checklists and Submittals
The requirements checklists help the designer to choose and specify the necessary products and
t45 http://www.gobrick.com/BIA/technotes/t3.htm
11 of 14 9/13/2009 12:30 PM
procedures found in the contract documents. Building codes set minimum requirements to protect
property and life safety. However, written contract documents may have more restrictive
requirements than provided in the building code. Adjustments for the particular project should be
made by the designer by reviewing the requirements checklists.
There are two checklists, mandatory and optional, that alert the designer to issues that must be
addressed. The mandatory list requires a choice on inspection, testing, material selection and
items not provided on the drawings or details of the project. The most significant change from the
1999 MSJC Specification in the mandatory checklist is exclusion of determining specified
compressive strength compliance. In addition, the 1999 MSJC Specification required that the level
of quality assurance be specified.
Part 1 - General
In Part 1 it is stated that the MSJC Specification covers requirements for materials and
construction of masonry elements. The provisions govern any project unless other requirements
are specifically stated in the contract documents. (1.1)
Definitions are provided and are coordinated with those found in the MSJC Code. (1.2) All
standards referenced in the MSJC Specification are listed. These standards include material
specifications, sampling procedures, test methods, detailing requirements, construction procedures
and classifications. The references are updated to the most current edition at the time of the
MSJC Code and Specification approval. (1.3)
The compressive strength of each wythe of masonry must equal or exceed that specified by the
engineer or architect. The compressive strength must be verified by the contractor by one of two
methods: unit strength or prism test. The unit strength method is a means to evaluate the strength
of masonry based upon the tested compressive strength of individual units and the mortar type
specified. The prism test method requires the sampling and testing of masonry prisms built with
the same types of materials that are used in the masonry construction. The MSJC Specification
specifies prism testing to be done in accordance with ASTM C 1314, Standard Test Method for
Compressive Strength of Masonry Prisms. (1.4B) Adhesion of adhered veneer units to their
backing is to be determined in accordance with ASTM C 482, Test Method for Bond Strength of
Ceramic Tile to Portland Cement. (1.4C)
Part 1 provides a list of items to be included in project submittals. Submittals should include
mortar and grout mix designs and test results, masonry unit samples and certificates, samples of
metal items such as reinforcement and wall ties. This also includes construction procedures for
cold- and hot-weather construction. (1.5)
Quality assurance is required by the MSJC Specification. The duties and services of the testing
agency, inspection agency and contractor are specified and are dependent upon the level of
quality assurance required. Article 1.6A outlines the responsibilities of the testing agencies.
Article 1.6B specifies the responsibilities of the inspection agency. Article 1.6C contains the
contractor's services and duties. The contractor must employ an independent testing laboratory to
perform required tests, to document submittals, certify product compliance, establish mortar and
grout mix designs, provide supporting data for changes requested by the contractor, or appeal
rejection of material found to be defective. The contractor must include in the submittals the
results of all testing performed to qualify the materials and to establish mix designs. Quality
assurances are actions taken by the owner or the owner's representative. They provide
assurance that actions of the contractor and supplier are in accordance with applicable standards
of good practice. Quality assurances are administrative policies and responsibilities related to
quality control measures that meet the owner's quality objectives. Quality control is the action
taken by the producer or contractor. This is simply systematic performance of construction,
testing and inspection to verify that proper materials and methods are used.
Quality assurance involves inspection and testing, preparation and erection of the masonry
structure. Inspection is assumed for every masonry project under the MSJC Code, a change from
previous masonry standards. The level of inspection and the amount of testing depend upon the
t45 http://www.gobrick.com/BIA/technotes/t3.htm
12 of 14 9/13/2009 12:30 PM
level of quality assurance specified. The level of quality assurance is determined according to
facility function, as defined by the general building code, and the method of design. The MSJC
Specification contains the same Quality Assurance tables that are found in the MSJC Code. (1.6)
Sample panels for masonry walls are required for Level 2 or 3 quality assurance. The construction
of a grout demonstration panel, used to depart from the requirements of Articles 3.5 C-E is also a
part of quality assurance. (1.6D)
Requirements for delivery, storage and handling of masonry materials are stated in order to avoid
contamination that might reduce the quality of the constructed masonry. (1.7) Project-specific
conditions such as support of construction loads by the masonry and shoring and weather
exposure during construction must be addressed. Cold- and hot-weather construction
requirements are included and are mandatory when they apply. The provisions for cold-weather
construction have been revised in the 2002 MSJC Specification. Provisions for both cold-and
hot-weather construction are separated into preparation, and construction protection. In most
cases the methods to achieve the requirements are left to the discretion of the contractor. (1.8)
Part 2 - Products
This section lists the available American Society for Testing and Materials (ASTM) standards for
masonry materials, including masonry units, mortar, grout, reinforcement and metal accessories.
Specific requirements are given if an appropriate ASTM standard does not exist. Referenced
ASTM standards for brick and tile are C 34, C 56, C 62, C 126, C 212, C 216, C 652, and C
1088. There are provisions for spacing of cross wires in joint reinforcement that are not included
in standard for this material. Minimum corrosion protection requirements for metal items are
stated including galvanized and epoxy coatings. Requirements for corrosion protection of bonded
and unbonded prestressing tendons are also included. Criteria are specified for prestressing
anchorages, couplers and end blocks. An accessories section provides requirements on
contraction joint material, expansion joint material, asphalt emulsions, masonry cleaners and joint
fillers. (2.1-2.5)
The MSJC Specification contains requirements for the mixing of mortar and grout. Time of mixing
and additives to mortar are limited. The grout must meet ASTM C 476 and be furnished and
placed with a slump between 8 in. (200 mm) and 11 in. (275 mm). (2.6)
Standard fabrication limits are stated for reinforcement and for prefabricated masonry panels.
These include bend and hook requirements for reinforcing bars. Prefabricated masonry panels
must conform to the provisions of ASTM C 901. (2.7)
Part 3 - Execution
The execution of the work includes initial inspection; preparation; masonry erection; reinforcement,
tie and anchor installation; grout placement; prestressing tendon installation and stressing
procedure; field quality control; and cleaning. Dimensional tolerances for foundations on which
masonry is placed are provided and should be measured prior to the start of masonry work. (3.1)
As part of the preparation requirements, clay or shale masonry units having initial absorption rates
in excess of one gram per minute per in
2
, as measured with ASTM C 67 must be pre-wetted, so
the initial rate of absorption will not exceed one gram per minute per in
2
when the units are used.
Cleanouts are required at the base of masonry to be grouted whenever pour heights exceed 5 ft
(1.5 m). (3.2)
Standard requirements for good workmanship are required by the MSJC Specification. These
include the requirement for completely filled mortar joints and grouted spaces. Proper support of
masonry and bracing during construction is required but is not prescribed. Dimensional tolerances
for the masonry are listed to ensure structural performance. The tolerances should not be used to
establish appearance criteria, unless specifically noted as such by the project specifications. (3.3)
Inspection of reinforcement and metal accessories is required to ensure that they have been
properly placed and are free of materials that hinder bond. Tolerances for locating and placing
t45 http://www.gobrick.com/BIA/technotes/t3.htm
13 of 14 9/13/2009 12:30 PM
reinforcing steel, wall ties, and veneer anchors are prescribed. Criteria for adjustable wall ties,
which are repeated from the MSJC Code, are included. Placement requirements for veneer
anchors have been added (3.4)
Prior to grout placement, debris must be removed from grout spaces. The grouting requirements
found in the MSJC Code are repeated in the MSJC Specification. Maximum grout pour heights are
determined by the type of grout used and the dimensions of the grout space. Consolidation of
grout is required to fill voids created by the loss of water from grout by absorption into the
masonry. Alternate grout placement requirements, established through the use of a grout
demonstration panel, are permitted. (3.5)
Prestressing tendon installation and stressing requirements include: tolerances; application and
measurement of the prestressing force; grouting bonded tendons; and burning and welding
operations. (3.6)
As part of field quality control, the specified compressive of masonry f'
m
is verified in accordance
with Article 1.6, Quality Assurance; grout is sampled and tested in accordance with Articles 1.4B
and 1.6. Provisions for cleaning exposed masonry surfaces complete the MSJC Specification.
(3.8)

SUMMARY
This Technical Notes provides an overview to the criteria contained in the MSJC Code and
Specification. The discussion centers on the design requirements to be followed by architects and
engineers and the masonry specifications to be implemented by the contractor during
construction. Changes to the Code and Specification in the 2002 editions are emphasized. The
MSJC Code and Specification provide the designer with coordination between the design and
construction phases of all masonry buildings.
The information and suggestions contained in this Technical Notes are based on the available data
and the experience of the engineering staff of the Brick Industry Association. The information
contained herein must be used in conjunction with good technical judgment and a basic
understanding of the properties of brick masonry. Final decisions on the use of the information
contained in this Technical Notes are not within the purview of the Brick Industry Association and
must rest with the project architect, engineer and owner.

t45 http://www.gobrick.com/BIA/technotes/t3.htm
14 of 14 9/13/2009 12:30 PM

Technical Notes 3A - Brick Masonry Material Properties
December 1992
Abstract: Brick masonry has a long history of reliable structural performance. Standards for the structural design of
masonry which are periodically updated such as the Building Code Requirements for Masonry Structures (ACI
530/ASCE 5/TMS 402) and the Specifications for Masonry Structures (ACI 530.1/ASCE 6/TMS 602) advance the
efficiency of masonry elements with rational design criteria. However, design of masonry structural members begins
with a thorough understanding of material properties. This Technical Notes is an aid for the design of brick and
structural clay tile masonry structural members. Clay and shale units, mortar, grout, steel reinforcement and
assemblage material properties are presented to simplify the design process.
Key Words: brick, grout, material properties, mortar, reinforcement, structural clay tile.
INTRODUCTION
The Masonry Standards Joint Committee (MSJC) has developed the Building Code Requirements for Masonry
Structures (ACI 530/ASCE 5/TMS 402) and the Specifications for Masonry Structures (ACI 530.1/ASCE 6/TMS
602). In this Technical Notes, these documents will be referred to as the MSJC Code and the MSJC Specifications,
respectively. Their contents are reviewed in Technical Notes 3. The MSJC Code and Specifications are periodically
revised by the MSJC and together provide design and construction requirements for masonry. The MSJC Code and
Specifications apply to structural masonry assemblages of clay, concrete or stone units.
This Technical Notes is a design aid for the MSJC Code and Specifications. It contains information on clay and shale
units, mortar, grout, steel reinforcement and assemblage material properties. These are used in the initial stages of
a structural design or analysis to determine applied stresses and allowable stresses. Material properties are
explained to aid the designer in selection of materials and to provide a better understanding of the structural
properties of the masonry assemblage based on the materials selected.
CONSTITUENT MATERIAL PROPERTIES
Because brick masonry is bonded into an integral mass by mortar and grout, it is considered to be a homogeneous
construction. It is the behavior of the combination of materials that determines the performance of the masonry as a
structural element. However, the performance of a structural masonry element is dependent upon the properties of
the constituent materials and the interaction of the materials as an assemblage. Therefore, it is important to first
consider the properties of the constituent materials: clay and shale units, mortar, grout and steel reinforcement. This
will be followed by a discussion of the behavior of their combination as an assemblage.
Clay and Shale Masonry Units
There are many variables in the manufacturing of clay and shale masonry units. Primary raw materials include
surface clays, fire clays, shales or combinations of these. Units are formed by extrusion, molding or dry-pressing
and are fired in a kiln at temperatures between 1800
o
F and 2100
o
(980
o
C and 1150
o
C). These variables in
manufacturing produce units with a wide range of colors, textures, sizes and physical properties. Clay and shale
masonry units are most frequently selected as a construction material for their aesthetics and long-term
performance. Consequently, material standards for clay and shale masonry units contain requirements to ensure that
units meet a level of durability and visual and dimensional consistency. Clay and shale masonry units used in
structural elements of building constructions are brick and structural clay tile. Material standards for brick and
structural clay tile include: ASTM C 216 (facing brick), ASTM C 62 (building brick), ASTM C 652 (hollow brick),
ASTM C 212 (structural clay facing tile) and ASTM C 34 (structural clay load-bearing tile).
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
1 of 10 9/13/2009 12:31 PM
While brick and structural clay tile are both visually appealing and durable, they are also well-suited for many
structural applications. This is primarily due to their variety of sizes and very high compressive strength. The material
properties of brick and structural clay tile which have the most significant effect upon structural performance of the
masonry are compressive strength and those properties affecting bond between the unit and mortar, such as rate of
water absorption and surface texture.
Unit Compressive Strength. The compressive strength of brick or structural clay tile is an important material
property for structural applications. In general, increasing the compressive strength of the unit will increase the
masonry assemblage compressive strength and elastic modulus. However, brick and structural clay tile are
frequently specified by material standard rather than by a particular minimum unit compressive strength. ASTM
material standards for brick and structural clay tile require minimum compressive strengths to ensure durability,
which may be as little as one-fifth the actual unit compressive strength. A recent Brick Institute of America survey of
United States brick manufacturers resulted in a data base of unit properties [6]. A subsequent survey of structural
clay tile manufacturers was conducted. The compressive strengths of brick and structural clay tile evaluated in these
surveys are presented in Table 1. As is apparent, all types of brick and structural clay tile typically exhibit
compressive strengths considerably greater than the ASTM minimum requirements. Compressive strength of brick
and structural clay tile is determined in accordance with ASTM C 67 Method of Sampling and Testing Brick and
Structural Clay Tile.
1
Extruded only.
2
Made from other materials or a combination of materials.
3
Based on gross area.
Unit Texture and Absorption. Unit texture and absorption are properties which affect the bond strength of the
masonry assemblage. In general, mortar bonds better to roughened surfaces, such as wire cut surfaces, than to
smooth surfaces, such as die skin surfaces. Cores or frogs provide a means of mechanical interlock. The bond
strength of sanded surfaces is dependent upon the amount of sand on the surface, the sand's adherence to the unit
and the absorption rate of the unit at the time of laying.
In practically all cases, mortar bonds best to a unit whose suction at the time of laying is less than 30 g/min/30 in
2
(1.55 kg/min/m
2
). Generally, molded units will exhibit a higher initial rate of absorption than extruded or dry-pressed
units. Unit absorption at the time of laying is an alterable property of brick and structural clay tile. In accordance with
the MSJC Specifications, units with initial rate of absorption in excess of 30 g/min/30 in.
2
(1.55 kg/min/m
2
) should be
wetted to reduce the rate of water absorption of the unit prior to laying. In addition, suction of very absorptive units
may be accommodated by using highly water-retentive mortars.
Mortar
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
2 of 10 9/13/2009 12:31 PM
The material properties of mortar which influence the structural performance of masonry are compressive strength,
bond strength and elasticity. Because the compressive strength of masonry mortar is less important than bond
strength, workability and water retentivity, the latter properties should be given principal consideration in mortar
selection. Mortar materials, properties and selection of masonry mortars are discussed in Technical Notes 8 Series.
Mortar should be selected based on the design requirements and with due consideration of the MSJC Code and
Specifications provisions affected by the mortar selected.
Laboratory testing indicates that masonry constructed with portland cement-lime mortar exhibit greater flexural bond
strength than masonry constructed with masonry cement mortar or air-entrained portland cement-lime mortar of the
same Type. This behavior is reflected in the MSJC Code allowable flexural tensile stresses for unreinforced
masonry, which are based on the mortar Type and mortar materials selected. In addition, masonry cement mortars
may not be used in Seismic Zones 3 and 4.
Other MSJC Code and Specifications provisions are the same for portland cement-lime mortars, masonry cement
mortars and air-entrained portland cement-lime mortars of the same Type. These include the modulus of elasticity of
the masonry, allowable compressive stresses for empirical design and the unit strength method of verifying that the
specified compressive strength of masonry is supplied. Following is a general description of the structural properties
of each Type of mortar permitted by the MSJC Code and Specifications.
Type N Mortar. Type N mortar is specifically recommended for chimneys, parapet walls and exterior walls subject
to severe exposure. It is a medium bond and compressive strength mortar suitable for general use in exposed
masonry above grade. Type N mortar may not be used in Seismic Zones 3 and 4.
Type S Mortar. Type S mortar is recommended for use in reinforced masonry and unreinforced masonry where
maximum flexural strength is required. It has a high compressive strength and has a high tensile bond strength with
most brick units.
Type M Mortar. Type M mortar is specifically recommended for masonry below grade and in contact with earth,
such as foundation walls, retaining walls, sewers and manholes. It has high compressive strength and better
durability in these environments than Type N or S mortars.
For compliance with the MSJC Specifications, mortars should conform to the requirements of ASTM C 270
Specification for Mortar for Unit Masonry. Field sampling of mortar for quality control should follow the procedures
given in ASTM C 780 Test Method for Preconstruction and Construction Evaluation of Mortars for Plain and
Reinforced Unit Masonry. Test procedures for masonry mortars are covered in Technical Notes 39 Series.
Grout
Grout is used in brick masonry to fill cells of hollow units or spaces between wythes of solid unit masonry. Grout
increases the compressive, shear and flexural strength of the masonry element and bonds steel reinforcement and
masonry together. For compliance with the MSJC Specifications, grout which is used in brick or structural clay tile
masonry should conform to the requirements of ASTM C 476 Specification for Grout for Masonry. Grout proportions
of portland cement or blended cement, hydrated lime or lime putty, and coarse or fine aggregate are given in Table
2.
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
3 of 10 9/13/2009 12:31 PM
1
Aggregate measured by volume in a damp, loose condition.
The amount of mixing water and its migration from the grout to the brick or structural clay tile will determine the
compressive strength of the grout and the amount of grout shrinkage. Tests indicate that the total amount of water
absorbed from grout by hollow clay units appears to be more dependent on the initial water content of the grout than
the absorption properties of the unit [3]. Grouts with high initial water content exhibit more shrinkage than grouts with
low initial water contents. Consequently, use of a non-shrink grout admixture is recommended to minimize the
number of flaws and shrinkage cracks in the grout while still producing a grout slump of 8 to 11 in. (200 to 280 mm),
unless otherwise specified.
The MSJC Specifications require grout compressive strength to be at least equal to the specified compressive
strength of masonry, f'm, but not less than 2,000 psi (13.8 MPa) as determined by ASTM C 1019 Method of
Sampling and Testing Grout. Test procedures for grout are explained in more detail in Technical Notes 39 Series. In
general, the compressive strength of ASTM C 476 grout by proportions will be greater than 2,000 psi (13.8 MPa).
Prediction of the compressive strength of grout which is proportioned in accordance with ASTM C 476 is difficult
because of the many possible combinations of materials, types of materials and construction conditions. However,
ASTM C 476 grout proportions produce a rich mix which is recommended to complement the high compressive
strength of brick and structural clay tile.
Steel Reinforcement
Steel reinforcement for masonry construction consists of bars and wires. Reinforcing bars are used in masonry
elements such as walls, columns, pilasters and beams. Wires are used in masonry bed joints to reinforce individual
masonry wythes or to tie multiple wythes together. Bars and wires have approximately the same modulus of
elasticity, which is stated in the MSJC Code as 29,000 ksi (200,000 MPa). In general, wires tend to achieve greater
ultimate strength and behave in a more brittle manner than reinforcing bars. Common bar and wire sizes and their
material properties are given in Table 3. As stated in the MSJC Specifications, steel reinforcement for masonry
structural members should comply with one of the material standards given in Table 4.
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
4 of 10 9/13/2009 12:31 PM
1
From reference [5].
ASSEMBLAGE MATERIAL PROPERTIES
The properties of the constituent materials discussed previously combine to produce the brick or structural clay tile
masonry assemblage properties. Following is a discussion of the material properties of the masonry assemblage.
Compressive Strength
Perhaps the single most important material property in the structural design of masonry is the compressive strength
of the masonry assemblage. The specified compressive strength of the masonry assemblage, f'm, is used to
determine the allowable axial and flexural compressive stresses, shear stresses and anchor bolt loads given in the
MSJC Code.
The compressive strength of the masonry assemblage can be evaluated by the properties of each constituent
material, termed in the MSJC Specifications the "Unit Strength Method," or by testing the properties of the entire
masonry assemblage, termed the "Prism Testing Method." These methods are not to be used to establish design
values; rather, they are used by the contractor to verify that the masonry achieves the specified compressive
strength, f'm.
Unit Strength Method. A benefit of verifying compliance of the compressive strength of masonry by unit, mortar
and grout properties is the elimination of prism testing. Each of the materials in the masonry assemblage must
conform to ASTM material standards mentioned in previous sections of this Technical Notes. For compliance with
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
5 of 10 9/13/2009 12:31 PM
these material standards, the compressive strength of the unit and the proportions or properties of the mortar and
grout must be evaluated. Not surprisingly, there have been attempts by numerous researchers to accurately
correlate the assemblage compressive strength with unit, mortar and grout compressive strengths. Testing an
assemblage of three materials produces a large scatter of compressive strengths covering all possible combinations
of materials. Therefore, estimates of the masonry assemblage compressive strength based on unit, mortar and
grout properties are necessarily conservative. The correlations provided in the MSJC Specifications, shown in Table
5, between unit compressive strength, mortar type and the masonry assemblage compressive strength represent a
lower-bound to experimental data. In addition, the MSJC Specifications unit strength method does not directly
address variable grout strength, multi-wythe construction or the influence of joint reinforcement on the compressive
strength of the masonry assemblage. Consequently, compliance with the specified compressive strength of masonry
by prism testing will always produce a more accurate and optimum use of brick or structural clay tile masonry's
compressive strength than the unit strength method.
The conservative nature of Table 5 should not be overlooked by the designer. A comparison of the predicted
assemblage compressive strength by the unit strength method in the MSJC Specifications and a data base of actual
brick masonry prism test results [1] reveals this conservatism. The average compressive strength of prisms of solid
brick units was found to be about 1.7 times the masonry compressive strength predicted by Table 5. The average
compressive strength of prisms of hollow units ungrouted and grouted was found to be 1.9 and 1.4 times the
compressive strengths predicted by Table 5, respectively.
1
Linear Interpolation is permitted.
Prism Test Method. Prism testing of brick or structural clay tile masonry provides a number of advantages over
constituent material testing alone. The primary benefit of prism testing is a more accurate estimation of the
compressive strength of the masonry assemblage. Another benefit of prism testing is that it provides a method of
measuring the quality of workmanship throughout the course of a project. Low prism strengths may indicate mortar
mixing error or poor quality grout.
The MSJC Specifications permit testing of masonry prisms to show conformance with the specified compressive
strength of masonry, f'm. In addition, the material components must meet the appropriate standards of quality.
Masonry prisms are tested in accordance with ASTM E 447 Test Methods for Compressive Strength of Masonry
Prisms, Method B as modified by the MSJC Specifications. At least three prisms are required by the MSJC
Specifications for each combination of materials. The average of the three tests must exceed f'm. Further
explanation of prism testing procedures is provided in Technical Notes 39B.
Shear Strength
The shear strength of a masonry assemblage may be separated into four parts: 1) the shear strength of the unit,
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
6 of 10 9/13/2009 12:31 PM
mortar and grout assemblage, 2) the effect of the shear span-to-depth ratio, M/Vd, 3) the enhancement of shear
strength due to compressive stress, and 4) the contribution of shear reinforcement in the masonry assemblage. All
four phenomenon are represented in the allowable shear stresses provided in the MSJC Code. However, only the
first and fourth items are controlled by material properties. Items two and three vary with member size and applied
loads.
The shear strength of the masonry assemblage is directly related to the properties of the unit, mortar and grout.
Shear failure of a unit-mortar assemblage is by splitting of units, step-cracking in mortar joints, or a combination of
the two. Unit splitting strength is increased by increasing the compressive strength of the unit. In general, unit
splitting is not a common shear failure mode of brick or structural clay tile masonry. Unit splitting occurs in masonry
assemblages of weak units and strong mortar and may also occur in shear walls which are heavily axially loaded.
Cracking in mortar joints is the more common shear failure mode for brick and structural clay tile masonry
assemblages. Mortar joint failure occurs by sliding along bed joints and separation of head joints. Mortar joint shear
failure is affected by bond strength and the frictional characteristics between the mortar and the unit. In general, a
unit-mortar combination which provides greater bond strength will also provide greater shear strength. Grouting the
masonry assemblage will also increase shear strength by providing a shear key between courses. The shear
strength of a masonry assemblage may be evaluated in accordance with ASTM E 519 Test Method for Diagonal
Tension (Shear) in Masonry Assemblages. The contribution of unit, mortar and grout to the allowable shear stresses
stated in the MSJC Code are based on ASTM E 519 tests of masonry assemblages.
Steel reinforcement may be added to the masonry assemblage to increase shear strength. Shear reinforcement
should be provided parallel to the direction of applied shear force. The MSJC Code also requires a minimum amount
of reinforcement perpendicular to the shear reinforcement of one-third the area of shear reinforcement. When shear
reinforcement is provided in accordance with the MSJC Code, allowable shear stresses given in the MSJC Code for
reinforced masonry are increased three times for flexural members and one and one-half times for shear walls.
Flexural Tensile Strength
Reinforced brick and structural clay tile masonry is considered cracked under service loads and the flexural tensile
strength of the masonry is neglected in design. However, cracking of an unreinforced brick or structural clay tile
masonry member constitutes failure and must be avoided. Thus, flexural tensile strength is an important design
consideration for unreinforced masonry. Flexural tensile strength is the bond strength of masonry in flexure. It is a
function of the type of unit, type of mortar, mortar materials, percentage of grouting of hollow units and the direction
of loading. Workmanship is also very important for flexural tensile strength, as unfilled mortar joints or dislodged units
have no mortar-to-unit bond strength.
Allowable flexural tensile stresses stipulated in the MSJC Code for unreinforced masonry are given in Table 6. The
allowable flexural tensile stresses for portland cement-lime mortars are based on full-size wall tests in accordance
with ASTM E 72 Method of Conducting Strength Tests of Panels for Building Construction. Values for masonry
cement and air-entrained portland cement-lime mortars are based on reductions obtained with comparative testing.
Flexural tensile strength may be evaluated by testing small-scale prisms in accordance with ASTM E 518 Test
Method for Flexural Bond Strength of Masonry or ASTM C 1072 Test Method for Measurement of Masonry Flexural
Bond Strength, but these results may not directly correlate to the allowable flexural tensile stresses in the MSJC
Code.
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
7 of 10 9/13/2009 12:31 PM
1
For partially grouted masonry allowable stresses shall be determined on the basis of linear interpolation between hollow units which are fully grouted or
ungrouted and hollow units based on amount of grouting.
Elastic Modulus
The elastic modulus of the masonry assemblage, in combination with the moment of inertia of the section,
determines the stiffness of a brick or structural clay tile masonry structural element. Elastic modulus is the ratio of
applied load (stress) to corresponding deformation (strain). The elastic modulus is roughly proportional to the
compressive strength of the masonry assemblage. Testing of brick masonry prisms indicates that the elastic
modulus of brick masonry falls between 700 and 1200 times the masonry prism compressive strength [4]. If the Unit
Strength Method is used to show compliance with the specified compressive strength of masonry, f'm, an accurate
estimation of the actual compressive strength of the masonry assemblage may not be known. Consequently, the
elastic modulus of the masonry assemblage is determined by the mortar type and the unit compressive strength.
See Table 7. The data in Table 1 can be used to estimate the modulus of elasticity of the masonry assemblage for
the type of unit selected.
The elastic modulus of grout is computed as 500 times the compressive strength of the grout in accordance with the
MSJC Code. In general, the elastic modulus of grout and the elastic moduli of brick or structural clay tile and mortar
masonry assemblages are comparable and are often considered equal for design calculations. However, the MSJC
Code recommends that the method of transformation of areas based on relative elastic moduli be used for
computation of stresses in grouted masonry elements.
1
MSJC Code Table 5.5.1.2.
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
8 of 10 9/13/2009 12:31 PM

Dimensional Stability
Dimensional stability is also an important property of the masonry assemblage. Expansion and contraction of the
brick or structural clay tile masonry may exert restraining stresses on the masonry and surrounding elements.
Material properties which affect dimensional stability of clay and shale unit masonry are moisture expansion, creep
and thermal movements. Effects of these phenomenon may be evaluated by the coefficients provided in the MSJC
Code, which are listed in Table 8. The coefficients in Table 8 represent average quantities for moisture expansion
and thermal movements and an upper-bound value for creep. Moisture expansion and thermal expansion and
contraction are independent and may be added directly. The magnitude of creep of clay or shale unit masonry will
depend upon the amount of load applied to the masonry element.
1
Conversion based on equivalent deformation at 100
o
F (38
o
C).
SUMMARY
This Technical Notes contains information about the material properties of brick and structural clay tile masonry.
This information may be used in conjunction with the MSJC Code and Specifications to design and analyze structural
masonry elements. Typical material properties of clay and shale masonry units, mortar, grout, reinforcing steel and
combinations of these are presented.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Institute of America. The information contained herein must be used
in conjunction with good technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the purview of the Brick
Institute of America and must rest with the project architect, engineer and owner.
REFERENCES
1. Atkinson, R.H., "Evaluation of Strength and Modulus Tables for Grouted and Ungrouted Hollow Unit
Masonry," Atkinson-Noland and Associates, Inc., Boulder, CO, November 1990, 47 pp.
2. Building Code Requirements for Masonry Structures and Commentary (ACI 530/ASCE 5/TMS 402-92)
and Specifications for Masonry Structures and Commentary (ACI 530.1/ASCE 6/TMS 602-92), American
Concrete Institute, Detroit, MI, 1992.
3. Kingsley, G.R., et al., "The Influence of Water Content and Unit Absorption Properties on Grout
Compressive Strength and Bond Strength in Hollow Clay Unit Masonry," Proceedings 3rd North American
Masonry Conference, The Masonry Society, Boulder, CO, June 1985, pp. 7:1-12.
4. Plummer, H.C., Brick and Tile Engineering, Brick Institute of America, Reston, VA, 1977, 466 pp.
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
9 of 10 9/13/2009 12:31 PM
5. "Steel Reinforcement Properties and Availability," Report of ACI Committee 439, Journal of the American
Concrete Institute, Vol. 74, Detroit, MI, 1977, p. 481.
6. Subasic, C.A., Borchelt, J.G., "Clay and Shale Brick Material Properties - A Statistical Report," submitted
for inclusion, Proceedings 6th North American Masonry Conference, The Masonry Society, Boulder, CO,
June 1993, 12 pp.
t3a http://www.gobrick.com/BIA/technotes/t3a.htm
10 of 10 9/13/2009 12:31 PM

Technical Notes 3B - Brick Masonry Section Properties
May 1993
Abstract: This Technical Notes is a design aid for the Building Code Requirements for Masonry Structures (ACI
530/ASCE 5/TMS 402-92) and Specifications for Masonry Structures (ACI 530.1/ASCE 6/TMS 602-92). Section
properties of brick masonry units, steel reinforcement and brick masonry assemblages are given to simplify the
design process. Section properties are used to calculate stresses and to determine the allowable stresses given in
the ACI 530/ASCE 5/TMS 402-92 Code.
Key Words: brick, dimensions, section properties, steel reinforcement.
INTRODUCTION
An assemblage's geometry determines its ability to resist loads. Section properties are properties of a masonry
assemblage which are based solely on its geometry. Section properties are used in design and analysis of brick
masonry structural elements. Section properties are used to determine allowable stresses which may be applied to
brick masonry elements, as well as to calculate an element's stress under applied loads. Because brick is a small
building unit, it may be used to construct assemblages of nearly any configuration. While this is a benefit of
construction with brick masonry, it can make design tedious because each masonry assemblage will have unique
section properties. To simplify the design process, this Technical Notes presents the section properties of brick
units, steel reinforcement and typical brick masonry assemblages. The section properties are based on specified
dimensions of the units and assemblages.
This Technical Notes is a design aid for the Building Code Requirements for Masonry Structures (ACI 530/ASCE
5/TMS 402-92) and the Specifications for Masonry Structures (ACI 530.1/ASCE 6/TMS 602-92). These documents,
which are promulgated by the Masonry Standards Joint Committee (MSJC), will be referred to as the MSJC Code
and the MSJC Specifications, respectively. References are made to the MSJC Code and Specifications to indicate
where each section property applies. Other Technical Notes in this series provide an overview of the MSJC Code
and Specifications and material properties of brick masonry.
NOTATION
Following are notations used in the text, figure and tables in this Technical Notes. Where applicable, notations are
the same as used in the MSJC Code and Specifications.
An Net cross-sectional area of masonry, in.
2
(mm

2
)
As Area of steel, in.
2
(mm
2
)
b Width of section, in. (mm)
bflange Width of flange, in. (mm)
bweb Width of web, in. (mm)
d Distance from extreme compression fiber ot the centroid of tension reinforcement, in. (mm)
Em Elastic modulus of masonry, psi (MPa)
Es Elastic modulus of steel, psi (MPa)
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
1 of 10 9/13/2009 12:31 PM
I Moment of inertia, in.
4
(m
4
)
j Ratio of distance between centroid of flexural compressive forces and centroid of tensile forces to
depth
k Ratio of distance between compression face and neutral axis to distance between compression
face and centroid of tensile forces
n Elastic moduli ratio, E
s
/E
m
Q First moment about the neutral axis of a section of that portion of the cross section lying between
the neutral axis and extreme fiber, in.
3
(m
3
)
r Radius of gyration, in. (mm)
S Section modulus, in.
3
(m
3
)
SECTION PROPERTIES OF CONSTITUENT MATERIALS
The constituent materials of units, mortar, grout and reinforcement combine to form brick masonry assemblages.
The section properties of each constituent material may be required in the design process. The section properties of
clay and shale masonry units are the basis for the section properties of the total brick masonry assemblage. The
section properties of steel reinforcement are used to determine the size and spacing of reinforcement within a brick
masonry assemblage.
Clay and Shale Masonry Units
Clay and shale masonry units are manufactured in a number of sizes and shapes. Clay and shale masonry units are
classified as either solid units or hollow units. Solid units may contain up to 25 percent void area as a percentage of
the gross cross-sectional area of the unit. Hollow units are classified as H40V for units with a total void area greater
than 25 percent and less than 40 percent of the gross cross-sectional area, or H60V for units with a total void area
greater than 40 percent and less than 60 percent of the gross cross-sectional area. The number and size of voids
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
2 of 10 9/13/2009 12:31 PM
vary with unit size and manufacturing equipment.
The range of sizes of clay and shale masonry units is given in Table 1. The names given for unit sizes in Table 1
were established by consensus of United States brick manufacturers and are standard terminology for the brick
industry. Further information on masonry unit sizes and coursing of brickwork can be found in the Technical Notes 10
series on estimating brickwork.


One criteria for unit selection may be accommodation of reinforcement within the unit itself. Placement of steel
reinforcement within the cores or cells of hollow units or solid cored units is permitted by the MSJC Code and
Specifications. A core is a void area less than or equal to 1 1/2 in.
2
(970 mm
2
). A cell is a void area which is larger
than 1 1/2 in.
2
(970 mm
2
). When placing reinforcement within a unit, adequate space for grouting must be provided.
Specifically, MSJC Code Section 8.3.5 requires that the minimum distance between the steel reinforcement and the
surrounding masonry unit be 1/4 in. (6 mm) when fine grout is used and 1/2 in. (13 mm) when coarse grout is used.
In certain instances, the cross-sectional area of masonry units may need to be determined. For example, the
compressive strength of a masonry prism is determined based on the unit's gross cross-sectional area when the
prism is constructed of solid units or fully grouted hollow units, and on the unit's net cross-sectional area when the
prism is constructed of hollow units. For solid units which contain cores, the gross cross-sectional area is used as
the net cross-sectional area. Unit cross-sectional area may be determined in accordance with ASTM C 67 Methods
of Sampling and Testing Brick and Structural Clay Tile.
The shell and web thickness of hollow units may need to be determined because hollow unit brick masonry walls are
typically face-shell bedded, while columns, pilasters and the first course of walls must be fully bedded. Minimum
thickness requirements for shells and webs of hollow units are established by ASTM C 652 Specification for Hollow
Brick (Hollow Masonry Units Made From Clay or Shale). These limits are given in Table 2. Many manufacturers
exceed the minimum thickness requirements given in Table 2, so it is advisable to request actual unit dimensions for
design purposes.
TABLE 2
Hollow Unit Section Properties
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
3 of 10 9/13/2009 12:31 PM
1
Cores greater than 1 in.2 (650 mm2) in cored shells shall be not less than 1/2 in. (13 mm) from any edge. Cores not greater than 1 in.2
(650 mm2) in shells cored not more than 35% shall be not less than 3/8 in. (10 mm) from any edge.
2
The thickness of webs shall not be less than 1/2 in. (13 mm) between cells, 3/8 in. (10 mm) between cells and cores or 1/4 in. (6 mm)
between cores.
Steel Reinforcement
Steel reinforcement for brick masonry assemblages consists of bars and wires. Reinforcing bars are placed in
grouted cavities, pockets, cores, cells or bond beams of brick masonry walls, columns, pilasters and beams. Steel
wire reinforcement is placed in brick masonry mortar joints to reinforce individual assemblages or to tie structural
elements together, such as the wythes of a multi-wythe wall. Common bar and wire section properties are given in
Table 3. The sizes of reinforcement listed in Table 3 are those permitted by the MSJC Code. The cross-sectional
area of reinforcement is used in MSJC Code Eq. 7-10 to determine the spacing of shear reinforcement. The
diameter of reinforcement is used to establish placement limits and minimum reinforcement development length
requirements given in Chapter 8 of the MSJC Code.
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
4 of 10 9/13/2009 12:31 PM
SECTION PROPERTIES OF BRICK MASONRY ASSEMBLAGES
The section properties of the assemblage of the constituent materials, along with the strength of the materials, will
determine the magnitude of loads the assemblage can resist. Consider the section properties required in the design
of brick masonry assemblages following the MSJC Code. The width of the brick masonry assemblage, b, is used in
MSJC Code Eqs. 6-7 and 7-3 and the effective depth of reinforcement, d, is used in MSJC Code Eqs. 7-3, 7-5, 7-8
and 7-10. Moment of inertia, I, is used in MSJC Code Eqs. 6-6 and 6-7. Radius of gyration, r, is used in MSJC Code
Eqs. 6-3, 6-4, 6-6, 7-1 and 7-2 to determine allowable compressive stresses and axial load. The first moment of
area, Q, is used in MSJC Code Eq. 6-7 to determine the shear stress in an unreinforced masonry element. The
dimensionless quantities k and j are used to determine a cracked, reinforced masonry element's compressive stress
and the allowable shear stress given in MSJC Code Eq. 7-3. The quantities k and j are functions of the area of
reinforcement, As, and the moduli ratio, n. The moduli ratio, n, is the ratio of the modulus of elasticity of steel, Es, to
the modulus of elasticity of masonry, Em.
Following is a discussion of the section properties of typical brick masonry assemblages. Tables 4 through 7 provide
section properties of these assemblages based on the dimensions indicated, which are based on the least specified
brick unit dimensions given in Tables 1 and 2. The MSJC Code requires that the computation of stresses be based
on the minimum net cross-sectional area of the element under consideration, An. For ungrouted, hollow brick units
laid with face-shell bedding, the minimum net cross-sectional area is the mortar bedded area. The computation of
stiffness of a brick masonry element may be based on the average net cross-sectional area of the element. The
average cross-sectional area is permitted for stiffness computations, because the distribution of material within an
element may be non-uniform. Examples of structural elements which have a non-uniform distribution of materials
include partially grouted or ungrouted hollow unit masonry walls.
Walls
Brick masonry walls may be constructed of a single wythe (one unit in thickness) or multiple wythes and can be
reinforced or unreinforced. Brick masonry walls may be loaded perpendicular to the plane of the wall or in the plane
of the wall. Out-of-plane loads may be caused by wind or earth pressures or by earthquake induced ground motions.
In-plane loads may be the dead weight of the structure, live loads or the result of the transfer of out-of-plane loads
through wall connections.
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
5 of 10 9/13/2009 12:31 PM
Section properties used in the MSJC Code's design equations for unreinforced masonry walls are I, r and Q. Section
properties used in the MSJC Code's design equations for reinforced masonry walls are j, b and d. Additional section
properties used to compute applied stresses are An, S and k. Effective areas for partially grouted, hollow unit
masonry walls are illustrated in Figure 1. Shading indicates net uncracked area, net cracked area and shear area for
a cracked cross section. For all illustrations in this Technical Notes, cross-hatching indicates mortar bedded areas.
In Figure 1(b), the effective width, b, is taken as the least of s, 6t and 72 in. (1.8 m). In Figure 1(c), the effective
width, b, is taken as the width of the grout space plus the thicknesses of the adjacent web and end web.
Effective Areas for Partially Grouted, Hollow Unit Masonry Walls
FIG. 1
Section properties for typical ungrouted and grouted brick masonry walls are given in Tables 4 and 5, respectively.
The quantities k and j are not provided in this Technical Notes because they are dependent upon the quantity of
reinforcement provided, the elastic moduli of the masonry and the steel and the loading conditions. The elastic
moduli of the masonry and the steel will determine the moduli ratio, n. The moduli ratio is used to determine the state
of stress in the steel and the masonry under loads. The loading conditions may be a combination of out-of-plane and
in-plane loads. Walls which are subject to flexural and axial loads must be designed considering the interaction of
axial load and bending moment, which may be accomplished by the use of a moment-load interaction diagram. The
method of development of a moment-load interaction diagram is beyond the scope of this Technical Notes.
TABLE 4
Ungrouted Wall Section Properties1
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
6 of 10 9/13/2009 12:31 PM
1
Per foot (305 mm) of wall.
2
Section properties are based on minimum solid face shell thickness (see Table 2) and face shell bedding.


TABLE 5
Grouted Wall Section Properties1
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
7 of 10 9/13/2009 12:31 PM
1
Per foot (305 mm) of wall. Section properties are based on minimum solid face shell thickness (see Table2) and face shell
bedding of hollow unit masonry.
Columns
Columns, as defined by the MSJC Code, are isolated elements whose horizontal dimension measured at a right
angle from the thickness dimension does not exceed three times the thickness dimension and whose height is at
least three times its thickness. Brick masonry columns are used to support large axial loads. Axial loads are typically
due to the permanent weight of the structure and the transient floor or roof load which is tributary to the column.
According to the MSJC Code, columns must be reinforced with a minimum of four reinforcing bars, and the area of
reinforcement, As, must be at least 0.0025 but not more than 0.04 times the column's net cross-sectional area, An.
The minimum nominal dimension of a column is 8 in. (200 mm) and the ratio of height to least lateral dimension must
not exceed 25. These requirements will influence the brick masonry column cross section selected. Typical brick
masonry column configurations and section properties are given in Table 6. Section properties are based on
uncracked cross sections. Typically, a brick masonry column will be in compression and will not crack under loads.
However, columns which are loaded by a eccentric axial load or a large lateral load may crack in flexure. A
moment-load interaction diagram should be used to design and analyze such columns, considering the section
properties of the cracked cross section. The method of development of a moment-load interaction diagram is
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
8 of 10 9/13/2009 12:31 PM
beyond the scope of this Technical Notes.
TABLE 6
Column Section Properties
Pilasters
A pilaster is simply an increase in the effective thickness of a wall at a specific location. To work together, the wall
and the thickened section must be integrally constructed. MSJC Code Section 5.10 permits three methods of
bonding a pilaster to create integral construction: 1) interlocking fifty percent of the masonry units, 2) toothing at 8 in.
(200 mm) maximum offset and attachment with metal ties and 3) providing reinforced bond beams at a maximum
spacing of 4 ft (1.2 m) on centers vertically. The length of the wall or flange that is considered to act integrally with
the pilaster from each edge of the pilaster or web is the lesser of six times the thickness of the wall or the actual
length of the wall. Typical brick masonry pilaster configurations and uncracked section properties are given in Table
7. As noted previously, cracked section properties such as k and j must be determined based on the amount of
reinforcement, the moduli ratio and the loading conditions. Pilasters which are loaded both out-of-plane and in-plane
must be designed considering the interaction of axial load and bending moment, which may be accomplished by the
use of a moment-load interaction diagram. The method of development of a moment-load interaction diagram is
beyond the scope of this Technical Notes.
TABLE 7
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
9 of 10 9/13/2009 12:31 PM
Pilaster Section Properties
1
Section properties are based on minimum solid face shell thickness (see Table 2) face shell bedding of the flange and full bedding of the
web.
Beams
Reinforced brick masonry beams may be used to span over wall openings such as windows and doors. Brick
masonry beams provide a number of advantages over precast concrete or steel lintels. For example, brick masonry
beams are a more efficient use of materials and produce a visually appealing brick masonry soffit. Some typical
brick masonry beam configurations and their section properties are given in Technical Notes 17H and 17J.
SUMMARY
Section properties of brick masonry materials and assemblages are required whenever a rational design of brick
masonry structural elements is developed following the criteria of the MSJC Code and Specifications. This Technical
Notes provides a summary of section properties of brick masonry. Section properties of clay and shale masonry
units, steel reinforcement and typical brick masonry assemblages are given.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Institute of America. The information contained herein must be used
in conjunction with good technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the purview of the Brick
Institute of America and must rest with the project architect, engineer and owner.
REFERENCES
1. Building Code Requirements for Masonry Structures and Commentary (ACI 530/ASCE 5/TMS 402-92)
and Specifications for Masonry Structures and Commentary (ACI 530.1/ASCE 6/TMS 602-92), American
Concrete Institute, Detroit, MI, 1992.
t3b http://www.gobrick.com/BIA/technotes/t3b.htm
10 of 10 9/13/2009 12:31 PM

Technical Notes 4 - Heat Transmission Coefficients of Brick Masonry Walls
Rev [Jan. 1982] (Reissued Sept. 1997)
Abstract: A procedure to analyze the heat flow through the opaque walls of a building envelope is provided. The
design coefficients of heat transmission are provided for commonly used construction materials. Methods of
calculating heat transmission coefficients and examples of heat loss calculations under steady-state conditions are
provided for opaque wall assemblies.
Key Words: brick, conductance, conductivity, energy, heat loss, rate of heat flow, resistance, resistivity,
steady-state conditions, series and parallel path, thermal transmission.
INTRODUCTION
Because of the finite supply of fossil fuels and the high cost of energy, the need to design energy-efficient buildings
that are also economical becomes important. Various industry groups are continually updating and refining energy
conservation standards and guidelines for use in the design of new buildings. These standards and guidelines may
be used to assist the building designers. The designer is confronted with the fact that no two buildings are exactly
identical, nor are the methods or modes of operation similar. Thus, the energy performance of each building, as a
whole, must be evaluated relative to the real performance of its materials, systems and equipment.
This Technical Notes provides information and methods of calculating transmission coefficients and heat transfer
values of brick masonry walls under static conditions. These may be used in energy conservation studies and
comparisons for predicting thermal performance of building components. However, ASHRAE (American Society of
Heating, Refrigerating and Air-Conditioning Engineers) cautions the designer that heat flow through a building
envelope is actually not static, and although steady-state calculations provide an estimate of energy consumption,
they do not take into account dynamic conditions such as the thermal storage capacity of materials, direct solar
radiation, wind and other variables. The term "steady-state" means that all ambient conditions are assumed to be
constant, which in the real world is virtually never the case.
BUILDING THERMAL DESIGN
The ASHRAE Handbook of Fundamentals states the following concerning heat transfer calculations:
"Current methods for estimating the heat transferred through floors, walls and roofs of buildings are largely based on
a steady-state or steady-periodic heat flow concept (Equivalent Temperature Difference Concept). The engineering
application of these concepts is not complicated and has served well for many years in the process of design and
selection of heating and cooling equipment for buildings. However, competitive practices of the building industry
sometimes require more than the selection or design of a single heating or cooling system. Consultants are
requested to present a detailed comparison of alternative heating and cooling systems for a given building, including
initial costs as well as short- and long-term operating and maintenance costs. The degree of sophistication required
for costs may make it necessary to calculate the heating and cooling load for estimating energy requirements in
hourly increments for a year's time for given buildings at known geographic locations. Because of the number of
calculations involved, computer processing becomes necessary. The hour-by-hour heating and cooling load
calculations, when based upon a steady heat flow or steady-periodic heat flow concept, do not account for the heat
storage effects of the building structure, especially with regard to net heat gain to the air-conditioned spaces."
The Handbook of Fundamentals also suggests that the designer consider the following factors when performing
heating load calculations: 1) building construction-heavy, medium or light; 2) presence of insulation; 3) infiltration and
ventilation loads; 4) glass area-normal or greater than normal; 5) occupancy nature and schedule; 6) presence of
auxiliary heating devices; and 7) expected cost of energy.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
1 of 18 9/13/2009 12:32 PM
Actual heat flow through a wall under normal weather conditions will involve daily cycles of solar radiation and air
temperature, changing wind speeds and directions, and radiation to the night sky. In studies ("Effective U-Values",
New Mexico Energy Institute, 1978) of dynamic heat transmission through a building envelope, it was found that
consideration of solar heat gain and material thermal storage effects provided results significantly different from
steady-state heat flow calculations. These studies also showed that the optimum economic insulation level varies
with wall orientations, and that changing the color of East, West and South walls was more cost-effective in some
instances than insulating. For a detailed description of the thermal storage effects of brick masonry walls, see
Technical Notes 43 and 43D.
The actual rate of heat flow through typical masonry building walls may be up to 20% less than the calculated rate
based on published U-values. This is indicated by past research (Structural Clay Products Research Foundation,
Studies of Heat Transfer.), which points out that the rate of heat transfer can be 20% to 60% greater than the
calculated rate for wood frame walls and metal panel walls, respectively.
Masonry walls have a more favorable rate of heat transfer because of their greater heat storage capacity, which is
sometimes referred to as thermal mass, or capacity insulation. The heat flows calculated by steady-state methods
are 29% to 60% greater than those measured under dynamic conditions for masonry walls. (Dynamic Thermal
Performance of an Experimental Masonry Building, Building Science Series 45, National Bureau of Standards.) This
means that massive masonry walls may be up to 60% better at retarding heat flow than steady-state U-values
indicate. A method to modify the steady-state calculations, in order to account for the effect of mass, is provided in
Technical Notes 4B.
The overall coefficient of heat transmission (U-value) of various walls discussed in this Technical Notes is used in
steady-state heat transfer and steady-periodic heat gain calculations.
Computer programs, such as those used by the National Bureau of Standards, (National Bureau of Standards Loads
Determination (NBSLD) Computer Program, T. Kasuda, "NBSLD-National Bureau of Standards Heating and Cooling
Load Determination Program", Journal, Automated Procedures for Engineering Consultants (APEC), Winter
1973-1974.) give values much closer to the actual performance of walls than is possible under the steady-state
concept of heat transfer. Government agencies and industry groups are continuing to examine simplified methods to
calculate dynamic heat flow without the use of computers.
TERMINOLOGY
Commonly used terms relative to heat transmission are defined below in accordance with ASHRAE Standard 12-75,
Refrigeration Terms and Definitions. All of these terms describe the same phenomenon, however, some are
described as determined by material dimensions and boundaries.
U = Overall Coefficient of Heat Transmission.
The rate of heat flow through a unit area of building envelope material or assembly, including its boundary
films, per unit of temperature difference between the inside and outside air. The term is commonly called the
"U-value". The Overall Coefficient of Heat Transmission is expressed in Btu/(hr
0
F ft
2
). Note that in computing
U-values, the component heat transmissions are not additive, but the overall U-value is actually less (i.e.,
better) than any of its component layers. Normally, the U-value is calculated by determining the resistance (R,
defined below) of each component, and then taking the reciprocal of the total resistance.
k = Thermal Conductivity.
The rate of heat flow through a homogeneous material, 1-in. thick, per unit of temperature difference between
its two surfaces. A material is considered homogeneous when the value of its thermal conductivity does not
depend on its dimensions (within the range normally used in construction). Thermal Conductivity is expressed
in (Btu in)/(hr
0
F ft
2
)
C = Thermal Conductance.
The rate of heat flow through a unit area of material per unit of temperature difference between its two
surfaces for the thickness of construction given, not per in. of thickness. Note that the conductance of an air
space is dependent on height, depth, position, character and temperature of the boundary surfaces.
Therefore, the air space must be fully described if the values are to be meaningful. For a description of other
than vertical air spaces, see the 1981 ASHRAE Handbook of Fundamentals, Chapter 23. Thermal
Conductance is expressed in Btu/(hr
0
F ft
2
)
h = Film or Surface Conductance.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
2 of 18 9/13/2009 12:32 PM
The rate of heat exchange between a unit or surface area and the air it is in contact with. Subscripts i and o
are used to denote inside and outside conductances, respectively. Film or surface conductance is expressed
in Btu/(hr
0
F ft
2
).
R = Thermal Resistance.
The reciprocal of a heat transfer coefficient, as expressed by U, C, or h. R is in (hr
0
F ft
2
)/Btu. For example,
a wall with a U-value of 0.25 would have a resistance value of R = I/U = 1/0.25=4.0. The value of R is also
used to represent Thermal Resistivity, the reciprocal of the thermal conductivity. Thermal Resistivity is
expressed in (hr
0
F ft
2
)/(Btu in)
Btu = British Thermal Unit.
It is the approximate heat required to raise 1 lb. of water 1 deg Fahrenheit, from 59
0
F to 60
0
F.
The difference of thermally homogeneous materials and thermally heterogeneous materials is shown in Figure 1.
There is a directly proportional relationship between the R and C of the thermally homogeneous material, at twice
the thickness the R is twice as great and the C is halved. For the thermally heterogeneous material, there is no
directly proportional relationship to the R or C and the material thickness. Fig. 1 also shows the horizontal path of
heat flow through a 1 ft
2
surface area of the wall component.
Thermal Transmittance Through Materialsa
FIG. 1
a
It is important to note that not all materials are isotropic with respect to heat transmission. In such thermally heterogeneous materials, the specific
thermal property under consideration could vary with temperature and material orientation. For this reason, care must be taken that the direction of
heat flow through a material is suitable for the material's intended use. Materials in which heat flow is identical in all directions are considered
thermally homogeneous.
CALCULATION OF OVERALL COEFFICIENTS
General
Conductance and resistance coefficients of various wall elements are listed in Table 1. These coefficients were
taken from the 1981 ASHRAE Handbook of Fundamentals, Chapter 23, which states:
"The most exact method of determining heat transmission coefficients for a given combination of building materials
assembled as a building section is to test a representative section in a guarded hot box. However, it is not
practicable to test all the combinations of interest. Experience has indicated that U-values for many constructions,
t4 http://www.gobrick.com/BIA/technotes/t4.htm
3 of 18 9/13/2009 12:32 PM
when calculated by the methods given in this chapter using accurate values for component materials, and with
corrections with framing member heat loss, are in good agreement with the values determined by guarded hot box
measurements, when there are no free air cavities within the construction. "Remember, the values shown for
materials in calculating overall heat transmission are representative of laboratory specimens tested under idealized
conditions. In actual practice, if insulation is improperly installed (for example), shrinkage, settling, insulation
compression, and similar factors may have a significant effect on the overall U-value numbers. Materials that are
field fabricated and consequently especially sensitive to the skills of the mechanic, are especially prone to variations
resulting in performance less than the idealized number."
Calculation Methods
Conductances and resistances of homogeneous material of any thickness can be obtained from the following
formula:
C
x
=k/x, and R
x
=x/k
where:
x=thickness of material in inches.

This calculation for a homogeneous material is shown in Fig. 1. The calculation only considers the brick component
of the wall assembly. Whenever an opaque wall is to be analyzed, the wall assembly should include both the outside
and inside air surfaces. The inclusion of these air surfaces makes all opaque wall assemblies layered construction.
In computing the heat transmission coefficients of layered construction, the paths of heat flow should first be
determined. If these are in series, the resistances are additive, but if the paths of heat flow are in parallel, then the
thermal transmittances are averaged. The word "series" implies that in cross-section, each layer of building material
is one continuous material. However, that is not always the case. For instance, in a longitudinal wall section, one
layer could be composed of more than one material, such as wood studs and insulation, hence having parallel paths
of heat flow within that layer. In this case, a weighted average of the thermal transmittances should be taken.
For layered construction, with paths of heat flow in series, the total thermal resistance of the wall is obtained by:
R
1
=R
1
+R
2
+...

and the overall coefficient of heat transmission is:
U=1/R
1
A solid 8-in. face brick wall would be a layered construction assembly in regard to thermal analysis:
R
(hr *
0
F * ft
2
)
--------------
BTU
Outside Air Surface 0.17
8-in. Face Brick 0.88
Inside Air Surface 0.68
Total: R
1
=1.73
U = 1/R
1
= 0.578 Btu/(hr *
0
F * ft
2
Average transmittances for parallel paths of heat flow may be obtained from the formula:
u
avg
[A
A
(U
A
) + A
B
(U
B
) + ...] / A
t
t4 http://www.gobrick.com/BIA/technotes/t4.htm
4 of 18 9/13/2009 12:32 PM
or
Uavg = [1/ (RA/AA) + 1/(RB/AB)...]/AT
where:
A
A
, A
B
, etc. = area of heat flow path, in Ft
2
,
U
A
,U
B
, etc.= transmission coefficients of the respective paths,
R
A
, R
B
, etc.=thermal resistance of the respective paths.
A
t
= total area beign considered (A
A
+A
B
+...), in Ft
2
Such an analysis is important for wall construction with parallel paths of heat flow when one path has a high heat
transfer and the other a low heat transfer, or the paths involve large percentages of the total wall with small
variations in the transfer coefficients for the paths.
Thermal bridges built into a wall may increase heat transfer substantially above the calculated amount if the bridge is
ignored. Thermal bridges occur in several types of walls. Three examples of these are shown. Different methods are
used in calculating the Uavg for metallic and non-metallic bridges. Examples of both are shown.
The brick veneer-frame wall shown in Fig. 2 has thermal bridges which occur at the wood studs. The parallel path
method allows the average U-value of the wall to be calculated by first calculating the U-values in series of the two
paths involved. Using the heat transmission coefficients for the various materials found in Table 1, the calculation is
shown in Fig. 2. The path at the wood stud is Path A and the path at the insulation is Path B.
Brick Veneer/Wood Stud
t4 http://www.gobrick.com/BIA/technotes/t4.htm
5 of 18 9/13/2009 12:32 PM
FIG. 2a
Brick Veneer/Wood Stud
FIG. 2b
This calculation reveals that, if the thermal bridge formed by the stud is considered, the Uavg exceeds the U of the
wall having the insulation (Path B) by approximately 6 per cent. It is common practice to calculate the U-values for
the insulation path by the series method and then multiply this value by 1.08 to obtain the Uavg for the wood frame
walls.
This method of correcting for wood framing in the walls is still used in many energy calculation guidelines
procedures, although it is no longer provided in the ASHRAE Handbook of Fundamentals. It should be noted that the
correction factor should be higher because this value properly predicts the Uavg for the studs, but does not
appropriately adjust the U-value for jambs, heads, sills, and top and toe plates. Also, if 2 in. x 6 in. wood studs are
used, the correction factor may no longer be appropriate.
Most masonry walls have parallel paths of heat flow which result from bonding the separate wythes together. This
may be by masonry bonders or metal ties. However, for conventional constructions, the effect of the bonders is not
significant, because of the relatively small area of the metal ties per sq ft of wall, and the slight differences in
conductivity or conductance of masonry units.
However, if masonry bonded cavity walls with insulation in the cavity of walls with a large amount of headers are
being considered, the parallel path method of calculation should be used. This is illustrated by the calculated
U-values of the brick cavity wall, shown in Fig. 3.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
6 of 18 9/13/2009 12:32 PM
Brick Masonry Cavity Wall(Masonry Bonded)
FIG. 3a
t4 http://www.gobrick.com/BIA/technotes/t4.htm
7 of 18 9/13/2009 12:32 PM
Brick Masonry Cavity Wall(Masonry Bonded)
FIG. 3b
If the thermal bridge at the bonder were ignored, the U-value would be the same as UB, which is 0.088. This is
approximately an 18 per cent differential between the series and parallel path calculated transmission coefficients.
The metal-tied cavity wall shown in Fig. 4 requires the parallel path method of calculation. However, a slightly
modified parallel path method should be used because the ASHRAE Handbook of Fundamentals requires that
calculations for metallic thermal bridges be done by the Zone Method. Under this method a slightly larger area is
assumed to be affected by the metallic bridge than just the area of the metal. The wall is divided into two zones,
Zone A, containing the metal; and Zone B, the remaining portion of the wall.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
8 of 18 9/13/2009 12:32 PM
Brick Masonry Insulated Cavity Wall
FIG. 4a
t4 http://www.gobrick.com/BIA/technotes/t4.htm
9 of 18 9/13/2009 12:32 PM
Brick Masonry Insulated Cavity Wall
FIG. 4b
The Handbook of Fundamentals also prescribes a method for determining the size and shape of Zone A. The
surface shape of Zone A in the case of a metal beam would be a strip of width, W, centered on the beam. In the
wall shown in Fig. 4, the shape of Zone A, due to the circular tie, would be a circle of diameter W. W is calculated
from the following formula:
where:
W = width or diameter of the zone, in in.,
m = width or diameter of the metal heat path, in in.,
d = distance from the panel surface to the metal, in in. The value of d should not be taken
as less than 0.5 in.
Calculations for W should be run for both surfaces and the larger of the two values used.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
10 of 18 9/13/2009 12:32 PM
For the insulated cavity wall with one metal tie provided for each 4 1/2 sq ft of wall surface, the calculations in Fig. 4
show that there is about 3.2 per cent increase in the heat loss through the wall when the ties are considered as
compared to the heat loss through the wall without consideration of the ties.
For a cavity wall which does not contain any insulation, the effect of the metal ties is much less. By subtracting out
the effects of the insulation and the metal ties through the insulation shown in Fig. 4, the effect of the wall tie through
a 1-in. air space may be determined:
This calculation procedure shows that the effect of a metal tie across a 1-in. air space is negligible. Fig. 5 shows the
calculations for an uninsulated cavity wall and again the effect is negligible. These calculations demonstrate that the
effect of a metal tie would be negligible in the 1-in. air space in brick veneer construction and also in uninsulated
cavity walls. There will be minor variations, depending on the type, size and spacing of metal ties, but the effect may
usually be ignored. However, as demonstrated in the calculations in Fig. 4, if the metal tie passes through insulation,
the effect of the metal tie on the thermal performance of the wall may become more significant. It should be noted
that as the R-value of the material the metal tie penetrates in creases, the per cent of heat loss due to the metal tie
also increases.


t4 http://www.gobrick.com/BIA/technotes/t4.htm
11 of 18 9/13/2009 12:32 PM
Brick Masonry Cavity Wall
FIG. 5a

t4 http://www.gobrick.com/BIA/technotes/t4.htm
12 of 18 9/13/2009 12:32 PM
Brick Masonry Cavity Wall
FIG. 5b
Another factor which affects the thermal performance of walls containing metal is the location of the metal in the
wall. The farther the metal is located from the face of the wall, the larger the area of the zone affected by the metal
tie. This may be demonstrated with brick veneer/steel stud systems. Consider the brick veneer/steel stud system
shown in Fig. 6. The steel stud backup system consists of 6-in., 20 gage steel studs at 24 in. o.c., with 6-in. batt
insulation between the steel studs. The width of Zone A is determined from the exterior flange of the steel stud to
the exterior face of the brick veneer, as shown in Fig. 6. The zone, including the metal, is quite wide for this type of
construction. In accordance with steady-state analysis, assuming that the 1-in. air space is a material of the system,
the width of the zone becomes 10.5359 in. The 1 5/8-in. wide flange of the metal studs, being relatively thin as
compared to the wall section, is not considered in the analysis because it will not significantly affect the average
thermal performance of the system.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
13 of 18 9/13/2009 12:32 PM
Brick Veneer/Steel Stud
FIG. 6a

Brick Veneer/Steel Stud
t4 http://www.gobrick.com/BIA/technotes/t4.htm
14 of 18 9/13/2009 12:32 PM
FIG. 6b
Without consideration of sills, jambs, heads, and toe and top channels, the performance of the brick veneer/ steel
stud system analyzed is almost 50 per cent less than the value calculated through the insulation. This performance is
calculated using the procedures in the 1981 ASHRAE Handbook and Product Directory. However, actual tests of the
heat transmission and more precise calculation procedures will probably demonstrate that the calculated heat loss is
considerably higher than the actual heat loss.
The intent of this example is simply to show that the thermal performance of brick veneer/metal stud systems is not
the same as brick veneer over wood frame. The designer should be aware of this discrepancy and the accuracy, or
inaccuracy of the approximation of thermal performance by simplified calculation procedures. The thermal
performance of the brick veneer/metal stud system would require a correction factor for the framing which greatly
exceeds the 8 per cent or the 1.08 U adjustment factor allowable for wood frame given in the previous brick veneer
example. Even for the wood frame, because of the presence of fire stops, heads, jambs, sills and top and toe
plates, it is recommended that the 1.08 factor for wood frame be increased to about 1.20, and that an even larger
factor be used for metal studs.
HEAT LOSS AND HEAT GAIN
Building envelope heat losses and heat gains are calculated using the overall heat transmission coefficients and other
known data.
Even though heat losses and heat gains are calculated using U-values in the steady-state and steady-periodic
formulae in lieu of the more accurate methods available, other factors greatly affect the performance of the building
envelope in conserving energy. It should be remembered that the values obtained from the steady-state and steady-
periodic calculations are merely an estimate of the thermal performance of the envelope.
The designer should be aware that several factors, other than U-values, determine the actual performance of the
envelope in conserving energy. Some of these factors are: 1) building orientation and aspect ratio (The aspect ratio
is the proportion of length to width. As the ratio approaches 1, the surface area to volume ratio decreases, and
generally there will be less loss of thermal energy from interior spaces through the building envelope); 2) exterior
surface color of envelope materials; 3) color of inside walls and ceilings; 4) mass and specific heat of envelope
materials; 5) wind velocities; 6) infiltration through the envelope; and 7) orientation, area and external shading of
glazing.
These factors are not considered in the steady-state calculations. However, if their effects on heat transmission are
kept foremost in the designer's mind, he can utilize the energy-conserving characteristics of each of these factors.
The resulting structure will be more thermally efficient than is shown by the steady-state calculations. Note that some
of these factors are accounted for by the CLTD values in heat gain calculations.
The steady-state method of calculation for heat loss is straightforward and simple to perform. The outdoor design
temperatures required can be found in the 1981 ASHRAE Handbook of Fundamentals. The inside design
temperature should be 72
o
F, or as prescribed by governing codes. The formula for calculating heat loss is as
follows:
where:
H = heat loss transmitted through the walls or other elements of the building envelope, in
Btu/hr,
A = area of the walls or other elements, in ft
2
,
U = overall coefficient of heat transmission of the walls or other elements, in Btu/(hr
o
F
ft
2
),
ti = indoor design temperature, in
o
F,
to = outdoor design temperature, in
o
F.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
15 of 18 9/13/2009 12:32 PM
t4 http://www.gobrick.com/BIA/technotes/t4.htm
16 of 18 9/13/2009 12:32 PM
a
From ASHRAE Handbook of Fundamentals, except as noted.
b
Face brick and common brick do not always have these specific densities. When the density is different from that shown, there will be a change in
the thermal conductivity.
c
Calculated data based upon hollow brick (25% to 40% cored) of one manufacturer. Based upon coring and density given. R figures based upon
coring and density of supplier using parallel path method. Vermiculite fill in cores.
d
From NCMA TEK 38
e
Values for metal siding applied over flat surfaces vary widely depending upon the amount of ventilation of air space beneath the siding, whether the
air space is reflective or non-reflective, and on the thickness, type. and application of insulating backing-board used. Values given are averages
intended for use as design guide values and were obtained from several guarded hot-box tests (ASTM C 236) on hollow-backed types and on types
made using backer-board of wood-fiber, foamed plastic, and glass fiber. Departures of +/- 50%. or more, from the values given may occur.
t4 http://www.gobrick.com/BIA/technotes/t4.htm
17 of 18 9/13/2009 12:32 PM
f
Thicknesses can vary. R values must be stamped on batt.
g
Based upon values as commercially produced. For calculations use specific manufacturer's specified values.
h
Time-aged values for board stock with gas-barrier quality (0.001 in thickness or greater) aluminum foil facers on two major surfaces.
CONCLUSION
Present-day technology for heat transmission (steady-state and steady-periodic) does not permit the designer to
take full advantage of the thermal mass of the element. While these design methods are relatively easy to
understand and calculate, they are not a true measure of the performance of massive elements. These methods do
give the designer an approximate solution which is on the conservative side in relation to the actual performance of
massive walls.
The designer should take into account the higher performance of massive construction which in many cases, may
provide savings in operational costs, efficiency of operation and energy. To provide a more accurate prediction of
these savings, a detailed computer study of the thermal performance of the structure is usually warranted.
Other Technical Notes in this series discuss heat gain through opaque walls, thermal transmission corrections for
dynamic conditions, balance point temperatures and energy conservation including worksheets, examples and data
tables.
METRIC CONVERSION
Because of the possible confusion inherent in showing dual unit systems in calculations, the metric (Sl) units are not
given in the data, equations or examples. Table 2 provides metric (Sl) conversion for the more commonly used heat
transmission units. This table is provided so that the user may use the data and procedures with Sl units.


REFERENCES
1. 1997 ASHRAE Handbook and Product Directory, Fundamentals Volume.
2. 1981 ASHRAE Handbook and Product Directory, Fundamentals Volume.


t4 http://www.gobrick.com/BIA/technotes/t4.htm
18 of 18 9/13/2009 12:32 PM

Technical Notes 4B Revised- Energy Code Compliance of Brick Masonry Walls
February 2002

Abstract: All buildings designed today must comply with energy code requirements. Building energy performance
requirements may be embodied in a model building code or in a separate energy standard. These documents typically
contain requirements for the building envelope, including walls, windows, doors, roofs and floors. Brick masonry, as a
high mass building material, has the inherent energy saving feature of thermal storage capacity (thermal mass). This
Technical Notes describes how to quantify thermal mass and calculate the heat capacity of several brick masonry walls.
The procedure for addressing thermal mass in residential and commercial construction when determining building
envelope compliance with widely used energy standards and codes is also described.
Key Words: brick, building codes, building envelope, energy, heat capacity, standards, thermal mass.

INTRODUCTION
All buildings designed today must comply with energy code requirements. Energy performance requirements may be
found in such documents as the 2000 International Residential Code [8], the 2000 International Energy Conservation
Code [7], and the ASHRAE/IES Standard 90.1-1999: Energy Efficient Design of New Buildings Except New Low-Rise
Residential Buildings [4]. These standards and codes specify energy efficient design through overall building
performance criteria or by a component prescriptive approach. The element in overall building performance discussed in
this Technical Notes is the building envelope. Brick masonry walls provide a uniquely energy efficient envelope due to
their high thermal mass. Thermal mass is the characteristic of heat capacity and surface area capable of affecting
building thermal loads by storing heat and releasing it at a later time. Materials with high thermal mass react more slowly
to temperature fluctuations and thereby reduce peak energy loads. Economic, energy efficient designs may be achieved
by recognizing this inherent aspect of brick masonry and incorporating it in the building envelope design.
The benefits of thermal mass have been known for a long time. Research in 1975-76 during the energy crisis, sponsored
by the Masonry Industry Committee, led to the development of a simplified method for quantifying thermal mass benefits
[10]. This method, called the M Factor, was developed for use by designers to compare wall systems with respect to
energy performance during the heating cycle. The M Factor was not intended for sizing of mechanical equipment, but
rather as a comparative analysis tool. By knowing the weight of a wall and the annual heating degree days (HDD), a
designer could determine the correction factor (M Factor) to convert the calculated U- or R-value of a wall to an
equivalent U- or R-value accounting for thermal mass and its effect of slowing heat transmittance. The U- and R-values
are measures of steady-state heat transmittance. The corrected U- or R-value was then used to comply with prescribed
energy requirements. At that time codes and standards did not incorporate thermal storage concepts when prescribing
limits on heat transfer.
Today, energy codes and standards specify energy requirements as a function of wall type. Adjustment factors are
included for masonry and other high mass walls as well as for walls built with steel studs that create thermal bridges. In
the case of masonry walls, a higher maximum permissible value for the coefficient of thermal transmittance (U-value) for
the building envelope is given depending upon where the insulation is located relative to the wall mass. Some codes
additionally specify a maximum overall thermal transfer value for walls (OTTVw) of mechanically cooled spaces. The
OTTVw is also a measure of heat transmittance and is a function of the wall temperature difference (TDEQ) which is also
related to wall weight.
This Technical Notes instructs the user on the methods for determining compliance of various brick masonry walls with
the building envelope requirements of several energy standards and codes. Those included are the 1999 ASHRAE/IES
Standard 90.1, the 2000 International Residential Code [8], and the 2000 International Energy Conservation Code [7].
The methods by which these energy standards and codes criteria reflect thermal mass properties of brick masonry are
explained.
The user of this Technical Notes is assumed to have a working knowledge of heat transmittance and familiarity with the
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
1 of 13 9/13/2009 12:33 PM
energy codes and standards listed. Procedures for calculating the actual U-values for walls can be found in Technical
Notes 4 Revised, Section 8.4 of ASHRAE/IES Standard 90.1, and in the ASHRAE Handbook of Fundamentals [5].
Because of the possible confusion inherent in showing dual unit systems in calculations, metric (SI) units are not given in
the data, equations, or examples in this Technical Notes. Table 1 provides metric (SI) conversion factors for the more
commonly used energy units.

NOTATION
Ad door area, ft2
Af fenestration area, ft2
Ag glazing area, ft2
Ao gross wall area above grade, ft2
Aw opaque wall area, ft2
c specific heat, Btu/(lb -F)
dt temperature difference between exterior and interior design conditions, F
HC heat capacity, Btu/(ft2-F)
HDD annual heating degree days
HDD65 annual Fahrenheit heating degree days, 65 F base
OTTVw overall thermal transfer value - walls, Btu/(hr-ft2)
SC shading coefficient of the fenestration, dimensionless
SF solar factor value, Btu/(hr-ft2)
TDEQ temperature difference value, F
Ud thermal transmittance of the door area, Btu/(hr-ft2-F)
Uf thermal transmittance of the fenestration area, Btu/(hr-ft2-F)
Ug thermal transmittance of the glazing area, Btu/(hr-ft2-F)
Uo average thermal transmittance of the gross wall area, Btu/(hr-ft2-F)
Uow overall thermal transmittance of the wall assembly, Btu/(hr-ft2-F)
Uw thermal transmittance of the opaque wall area, Btu/(hr-ft2-F)
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
2 of 13 9/13/2009 12:33 PM
w weight, lb/ft2

HEAT CAPACITY
In most energy codes, the thermal characteristics of high mass walls are quantified by measuring the heat capacity of the
wall. Heat capacity represents the amount of thermal energy which may be stored by a material. For walls constructed
of multiple materials, total heat capacity is calculated as the sum of the heat capacities of the individual components. In
most energy codes and standards in the United States, heat capacity (HC) of a wall is calculated as the product of
weight per unit area and specific heat (HC = w x c). Since the specific heats of most building materials are roughly
equal, the heat capacity of a wall is directly proportional to its weight. Those materials which are relatively lightweight,
such as insulation, do not have a significant effect on heat capacity and are often ignored when determining heat
capacity. Use of the adjustment factors for mass walls in the 2000 International Residential Code [8] and the 2000
International Energy Conservation Code [7] is limited to walls having a heat capacity greater than or equal to 6 Btu/ft
2
.
Sample calculations of heat capacity for several brick walls are provided in Figure 1. Brick walls with a nominal thickness
or 4 in. or greater have heat capacities greater than or equal to 6 Btu/ft
2
.

ENERGY CODE COMPLIANCE
Each energy code and standard is slightly different in scope and criteria for compliance. The ASHRAE/IES Standard
90.1 is only applicable to non-residential buildings. Both residential and non-residential criteria may be found in the model
building codes or the International Energy Conservation Code [7]. Each code and standard is discussed individually
below.

ASHRAE/IES Standard 90.1
The ASHRAE/IES Standard 90.1 covers the energy performance design of new buildings except residential buildings of
three stories or less. Compliance with this standard may follow one of two paths: the Building Energy Cost Budget
Method or the System/Component Method.
The Building Energy Cost Budget Method (BECBM) is to be used with innovative design concepts which cannot be
accommodated by the System/Component Method or when a design fails the System/Component approach. The
BECBM requires a detailed energy analysis to determine the estimated design energy cost. The BECBM permits any
design whose design energy cost does not exceed the specified energy cost budget and meets the other requirements of
the method. A complete description of this method can be found in Section 13 of the ASHRAE/IES Standard 90.1.
The System/Component Method can be divided into two compliance paths for the building envelope: Prescriptive Criteria
found in Section 8.5 and System Performance Criteria found in Section 8.6. These methods give minimum requirements
to satisfy both heating and cooling cycle conditions. As the use of the ASHRAE/IES Standard 90.1 may be somewhat
confusing, the National Codes and Standards Council of the Concrete and Masonry Industries has published a handbook
which discusses the benefits of thermal mass and the design provisions of ASHRAE/IES Standard 90.1 [2]. In addition to
the examples given in this Technical Notes, the reader is also urged to refer to this handbook.
FIG. 1
Heat Capacities of Several Brick Walls
_________________________________________________________________________________

(a) 4 IN. BRICK AND WOOD STUD WALL

t4b http://www.gobrick.com/BIA/technotes/t4b.htm
3 of 13 9/13/2009 12:33 PM
4 IN. BRICK w = (130 lb/ft3) x [(0.75)(3.63 in.) / 12in./ft] = 29.5 lb/ft2
(>75% SOLID) c = 0.20 Btu/(lb-F)
HC = 29.5 x 0.20 = 5.9 Btu/(ft2-F)
4 IN. STUD w = 45 lb/ft3 x [(3.5 in. x 1.5 in.) / (144 in.2/ft2)] x (12 in./ft / 16 in.)
= 1.23 lb/ft2
c = 0.30 Btu/(lb-F)
HC = 1.23 x 0.30 = 0.4 Btu/(ft2-F)
(2) 1/2 IN. w = 50 lb/ft3 x [(2)(0.5 in.) / 12 in./ft] = 4.2 lb/ft2
GYPSUM BOARD c = 0.26 Btu/(lb-F)
HC = 4.2 x 0.26 = 1.1 Btu/(ft2-F)
INSULATION NEGLIGIBLE
TOTAL HC = 5.9 + 0.4 + 1.1 = 7.4 Btu/(ft2F)
_________________________________________________________________________________

(b) 4 IN. BRICK AND 8 IN. LIGHTWEIGHT CMU WALL

4 IN. BRICK HC = 5.9 Btu/(ft3F) (from Fig. 1a)
8 IN. LIGHTWEIGHT CMU w = 90 lb/ft3 x [(0.52)(7.63 in.) / 12in./ft] = 29.7 lb/ft2
(52% SOLID) c = 0.21 Btu/(lb-F)
HC = 29.7 x 0.21 = 6.2 Btu/(ft2-F)
INSULATION NEGLIGIBLE
TOTAL HC = 5.9 + 6.2 = 12.1 Btu/(ft2-F)
_________________________________________________________________________________

t4b http://www.gobrick.com/BIA/technotes/t4b.htm
4 of 13 9/13/2009 12:33 PM
(c) 4 IN. BRICK AND 6 IN. LIGHTWEIGHT CMU WALL

4 IN. BRICK HC = 5.9 Btu/(ft2-F) (from Fig. 1a)
6 IN. LIGHTWEIGHT CMU w = 90 lb/ft3 x [(0.55)(5.63 in.) / 12 in./ft] = 23.2 lb/ft2
(55% SOLID) c = 0.21 Btu/(lb-F)
HC = 23.2 x 0.21 = 4.9 Btu/(ft2-F)
INSULATION NEGLIGIBLE
TOTAL HC = 5.9 + 4.9 = 10.8 Btu/(ft2-F)
_________________________________________________________________________________

(d) 6 IN. HOLLOW BRICK WALL

6 IN. BRICK w = 130 Ib/ft3 x [(0.60)(5.63 in.) / 12 in./ft] = 36.6 Ib/ft2
(60% SOLID) c = 0.20 Btu/ (Ib-F)
TOTAL HC = 36.6 x 0.20 = 7.3 Btu/(ft2-F)
IF GROUTED, HC WOULD BE EVEN GREATER
_________________________________________________________________________________

Prescriptive Criteria. Section 8.5 of the ASHRAE/IES Standard 90.1 provides precalculated Alternate Component
Package (ACP) tables based on the System Performance Criteria in Section 8.6 for a set of climate ranges. These ACP
tables list the maximum permissible percentage of fenestration in a wall area, maximum thermal transmittance U-values,
and minimum thermal resistance R-values as a function of the building's internal energy load, type and characteristics of
fenestration and wall construction. The many climatic variables which influence the building envelope are grouped
together in each ACP table for a range of climates. Thus, the criteria found in the ACP tables address a worst case
condition and may be more stringent than the System Performance Criteria in Section 8.6.
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
5 of 13 9/13/2009 12:33 PM
The maximum permissible overall thermal transmittance value of an opaque wall (Uow) using the prescriptive envelope
criteria and the appropriate ACP table. The following example illustrates the benefits of using a thermal mass wall by
comparing the maximum permissible Uow-value of a lightweight and a high mass wall. The UOW-value is a function of the
wall weight (represented by HC); the building's internal cooling loads due to heat generated by lights, equipment, and
people (ILD); the placement of the insulation either internal to or integral with the wall mass (INT INS) or outside the wall
mass (EXT INS); and the percentage of total wall area consisting of doors, windows and other glazing (PCT FEN).
Refer to Fig. 2 of this Technical Notes and Section 8.6 of the ASHRAE/IES Standard 90.1 for the tables and terms used
in this example.

EXAMPLE 1: ASHRAE/IES Standard 90.1 Prescriptive Criteria
Office Building
Determine the maximum permissible overall wall thermal transmittance value (Uow) of a 12,000 ft2 office building located
near Albuquerque, NM. The building is constructed of 4 in. nominal brick veneer with 8 in. nominal concrete masonry
loadbearing walls with insulation as shown in Fig. 1b. The building's fenestration is 30 percent of the total wall area. To
determine the maximum permissible Uow-value, use the following steps.
Step 1: To use the Prescriptive Envelope Criteria, first determine the appropriate ACP table from the locations listed in
Table 8A-0 in Attachment 8A of the Standard. Find Albuquerque, NM in Table 2 of this Technical Notes. From Table 2,
determine that the appropriate ACP table is Table 8A-23 of the Standard. The ACP table for Albuquerque, NM (8A-23)
is shown in Fig. 2 of this Technical Notes.
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
6 of 13 9/13/2009 12:33 PM
1Table 8A-0, reprinted by permission from ASHRAE/IES Standard 90.1-1989 published by ASHRAE
FIG. 2

Step 2: Calculate the heat capacity (HC) of the wall in question. The HC of the brick and concrete masonry wall shown
in Fig. 1b has already been calculated to be 12.1 Btu/(ft2-F) .
Step 3: Calculate internal load density (ILD) of the building. Section 8.5.5.2 of the Standard defines ILD as the sum of
Lighting Power Density (LPD), Equipment Power Density (EPD) and Occupant Load Adjustment (OLA). Values for LPD
are found in Table 6-5 of the Standard. (Note that Unit Lighting Power Allowance (ULPA) equals LPD.) For this office
building example, LPD equals 1.81 W/ft2. The EPD can be selected from Table 8-4 of the Standard. For an office, EPD
equals 0.75 W/ft2. OLA is a measure of the heat generated by living objects and is discussed in Section 8.5.5.2 of the
Standard. For this example, assume OLA equals 0.0 W/ft2. Therefore,

t4b http://www.gobrick.com/BIA/technotes/t4b.htm
7 of 13 9/13/2009 12:33 PM
ILD = LPD + EPD + 0LA = 1.81 + 0.75 + 0.0 = 2.56 W/ft2.

Using the ACP table for Albuquerque, NM shown in Fig. 2, enter the appropriate row based on the ILD of the building.
Since ILD for this example is 2.56 W/ft2, enter the row for ILD 1.51 - 3.00.
Step 4: The ACP tables contain criteria for both fenestration and opaque portions of the building envelope. This example
addresses only the opaque wall requirements. Therefore, to determine the maximum permissible U OW-value for the wall
assembly in this example, move to the far right to the box under the heading OPAQUE WALL. Since HC of the wall in
question is greater than 5 Btu/(ft2-F), go to the subheading MASS WALL and find the box corresponding to the ILD row
found in Step 3.
Step 5: The maximum UOW-value is also a function of the location of insulation in the wall assembly. The insulation in this
example is placed between or integral with the wall mass. Therefore, select the column for interior or integral insulation,
INT INS. See Section 8.5.5.3 of the Standard for a complete discussion of insulation location.
Step 6: Find the appropriate rows under MASS WALL corresponding to the HC of the wall in question. In this example,
HC equals 12.1 Btu/(ft2-F), so use the rows HC greater than or equal to 10. Follow these rows to where they intersect
the INT INS column. There are two possible values of Uow based on the percentage of fenestration (PCT FEN) in the
envelope.
Step 7: Follow the rows for HC greater than or equal to 10 to PCT FEN equal to 11 and PCT FEN equal to 57. Recall
that the building's fenestration (PCT FEN) equals 30 percent of the wall area in this example. The UOW-value
corresponding to 11 percent equals 0.15 Btu/(hr-ft2-F), and Uow-value corresponding to 57 percent equals 0.14 Btu/(hr-
ft2-F). Linearly interpolate for PCT FEN equal to 30 or use the lower of the two values. Using the lower value as the
maximum permissible value, Uow must be less than or equal to 0.14 Btu/(hr-ft2-F).
Step 8: To comply with the Standard, the calculated UOW-value of the wall in question may not exceed the maximum
permissible value as determined from the ACP table. Using the steps found in Technical Notes 4 Revised or the
ASHRAE Hand/book of Fundamentals, the thermal transmittance of the wall in Fig. 1b is calculated to be 0.10 Btu/(hr-
ft2-F). Since the calculated UOW-value is less than the maximum permissible value of 0.14 Btu/(hr-ft2-F), the wall
construction complies with the Building Envelope Requirements of ASHRAE/IES Standard 90.1.
Compare the maximum permissible Uow-value of the thermal mass wall in this example with the maximum permissible
Uow-value for a lightweight wall with HC less than 5 Btu/(ft2-F). The box under the heading OPAQUE WALL shows that
the maximum Uow-value is only 0.10 Btu/(hr-ft2-F) for a lightweight wall. In terms of R-values, this thermal mass wall
must have a minimum R-value of 7.1 (hr-ft2-F)/Btu, whereas a lightweight wall must have an R-value of at least 10.0
(hr-ft2-F)/Btu.
System Performance Criteria. A system approach for compliance with envelope requirements is provided in Section
8.6 of the ASHRAE/IES Standard 90.1. This method is more flexible than the Prescriptive Criteria when considering
thermal mass for several reasons. The external wall criteria are based on annual energy calculations for a specific
location, rather than for a group of climates. Calculations allow for variations in internal loads and wall heat capacity by
separating the building into zones. Furthermore, wall assemblies with HC greater than or equal to 7 Btu/(ft2-F) do not
have limits on the permissible Uow-value as they do in the ACP tables. Compliance with the System Performance Criteria
is achieved if the calculated energy loads do not exceed the criteria specified in Section 8.6. The System Performance
Criteria approach requires numerous mathematical calculations by hand or a computer program. Information on an
acceptable computer program, ENVSTD, is part of Appendix D of the ASHRAE/IES Standard 90.1. The program models
the building envelope's performance and fully accounts for the effects of thermal mass. For this reason, ENVSTD is
recommended for use with the System Performance Criteria when determining energy compliance of brick masonry
walls, particularly when passive solar technologies are employed in the design.
International Residential Code
The 2000 International Residential Code (IRC) covers all aspects of residential design and construction. Section N1102
of Chapter 11 - Energy Efficiency contains prescriptive requirements for energy compliance of the building envelope. This
Code is only applicable for climates with Heating Degree Days (HDD) of less than 13,000. Further restrictions on the use
of this Code limit the glazing area of Type A-1 Residential buildings to 15 percent or less and 25 percent or less for Type
A-2 Residential buildings. Thermal performance criteria in the form of minimum required insulation R-values and maximum
permissible U-factors are specified for each element in the building envelope. This Technical Notes covers the
requirements for exterior walls only.
Section N1102.1.1 contains tables with minimum R-values for walls, ceilings, floors, etc. based on the climates Heating
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
8 of 13 9/13/2009 12:33 PM
Degree Days (HDD). Two tables are included specifically for mass walls. The first, Table N1102.1.1.1(1), specifies the
minimum R-values for mass walls. The requirements vary depending upon the location of insulation and HDD. Walls with
insulation placed on the exterior of the entire masonry mass are considered to have exterior insulation. An example of this
type of construction is EIFS with a masonry backup. Walls that have insulation sandwiched between two roughly equal
layers of masonry or mixed with the mass materials are considered to have integral insulation. Examples of this type of
construction include masonry cavity walls and concrete masonry walls with insulated cores. Log walls are also considered
to have integral insulation. Walls with interior insulation have the entire mass material on the exterior side of the insulation,
such as in the case of brick veneer walls.
The required R-values found in Table N1102.1.1.1(1) for mass walls with exterior or integral insulation are the same. Mass
walls that do not meet the definitions for exterior or integral insulation are grouped into the Other mass walls category.
The R-value requirements are lowest for walls having exterior or integral insulation. However, even Other mass walls
reflect a considerable reduction in R-value as compared with non-mass walls. This savings is shown in Example 2(a).
The second Table in this section provides a listing of the R-values of common mass wall assemblies. To comply with the
minimum R-value requirements of Table N1102.1.1.1(1), find the R-value of the mass assembly from Table N1102.1.1.1(2)
and add to it the R-value of any insulation or other layers in the wall assembly. See Example 2(b).
EXAMPLE 2: 2000 International Residential Code - Mass Wall Requirements
Determine the required amount of insulation for the walls of a single family home in Raleigh, North Carolina framed with
wood construction and a brick veneer. The glazing area is 12 percent. What is the required insulation R-value with vinyl
siding instead of brick veneer? What is the required insulation R-value if the house is built with loadbearing brick masonry
cavity walls?
Ex. 2a: Determine the required R-values for brick veneer wall, a vinyl sided wall, and a loadbearing brick masonry cavity
wall with integral insulation.
Step 1: Determine the Climate Zone and annual Fahrenheit Heating Degree Days (HDD) for the location given.
Climate Zones are listed in Table N1101.2 of the IRC. The Climate Zone for Raleigh, North Carolina which is
located in Wake County is Zone 7.
Step 2: Brick veneer construction is considered to have interior insulation. On Table N1102.1.1.1(1), Mass Wall
Prescriptive Building Envelope Requirements, find Zone 7 and the column for Other mass walls. The required
mass wall assembly (insulation and masonry) R-value is 10.8 (hrft
2
F)/Btu.
Step 3: For the vinyl-sided wall, use Table N1102.1. Zone 7 corresponds to HDD 3,000 - 3,499. From the walls
column determine that R-13 insulation is required for the non-mass wall system.
Step 4: For the loadbearing brick masonry cavity wall, use Table N1102.1.1.1(1). Under the column for integral
insulation, find that the mass wall assembly (insulation and masonry) R-value must be R-8.9 (hrft
2
F)/Btu.
Ex. 2b: Determine the required insulation values for the brick veneer wall and the loadbearing brick masonry cavity wall.
Step 1: For the brick veneer wall, From Table N1102.1.1.1(2) determine that brick veneer alone has an R-value of
2.0(hrft
2
F)/Btu. The required insulation R-value is calculated as 10.8 - 2.0 = 8.8(hrft
2
F)/Btu
Step 2: For the brick masonry cavity wall, assume ungrouted cells are not insulated and the only insulation is
located in the cavity. From Table N1102.1.1.1(2) determine that the mass assembly R-value is equal to 3.7
(hrft
2
F)/Btu. Calculate the required insulation R-value as 8.9 - 3.7 = 5.2 (hrft
2
F)/Btu.

2000 International Energy Conservation Code
The 2000 International Energy Conservation Code (IECC) prepared by the International Code Council, Inc. is applicable
to residential dwellings as well as commercial, institutional and other buildings. This code sets limits on the permissible
thermal transmission (U-value) of the building envelope. Residential buildings may comply with this code by adhering to
one of three chapters : Chapter 4 - Residential Building Design by Systems Analysis and Design of Buildings Utilizing
Renewable Energy Sources; Chapter 5 - Residential Building Design by Component Performance Approach; or Chapter
6 - Simplified Prescriptive Requirements for Residential Buildings, Type A-1 and A-2.
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
9 of 13 9/13/2009 12:33 PM
RESIDENTIAL REQUIREMENTS - Chapters 4, 5, and 6
Chapter 4 - Systems Analysis and Renewable Energy Source Analysis. Chapter 4, as its title implies, is
separated into two sections: Systems Analysis and Renewable Energy Source Analysis. Both sections require an
analysis of the annual energy usage of the proposed system. Requirements and procedures for analysis are specified.
Compliance is achieved if annual energy consumption is not greater than that of a similar residential building designed
according to IECC Component Performance Approach found in Chapter 5.
Chapter 5 - Component Performance Approach. The component performance approach presented in Chapter 5 of
the IECC has requirements for residential building envelope (Section 502) as well as building mechanical systems, water
heating, and electrical power and lighting. Residential requirements of Section 502 are divided into two types of
residential construction, A-1 and A-2. Type A-1 are buildings with glazing areas that do not exceed 15 percent of the
gross area of exterior walls. Detached one- and two-family dwellings are commonly Type A-1 buildings. Type A-2 have
glazing areas that do not exceed 25 percent of the gross area of the exterior walls. Section 502.2, Heating and Cooling
Criteria, specifies the maximum thermal transmission U-value for each building component (walls, roof, slab on grade,
etc.). In residential construction, the maximum permissible UW-value for walls is a function of the heat capacity of the wall
in question. The example that follows illustrates how the maximum permissible Uow-value may be increased if the HC of
the wall in question is greater than or equal to 6 Btu/(ft2-F). All 4 in. brick veneer walls have a HC of at least 6
Btu/(ft2-F). This example utilizes the Compliance by Performance on an Individual Component Basis found in Section
502.2.1. Other provisions in this section contain criteria for compliance using Acceptable Practices (Section 502.2.3) and
Prescriptive Criteria (502.2.4).
EXAMPLE 3: International Energy Conservation Code Component Performance Criteria
Single Family Home
Determine the maximum permissible thermal transmittance of the opaque wall area (UW-value) of a 2,000 ft2 two-story
single family home located in a suburb of Washington, D.C. The house is brick veneer over wood frame constructed as
shown in Fig. 1a. The home's fenestration is 20% of the total wall area: 15% glazing, 5% doors. Thermal transmittance
values for the fenestration are: Ug = 0.48 and Ud = 0.48. The following steps are suggested to determine the maximum
permissible UW-value.
Step 1: Determine the annual Fahrenheit heating degree days (HDD, 65 F base) for the location given. For Washington,
D.C., HDD equals 4224. HDD for many U.S. cities can be found in the 2000 International Energy Conservation Code [7].
Other resources include Table B7.1 of Building Control Systems [1] or in the 1981 ASHRAE Handbook of Fundamentals
[5]. A single family home with a glazing area of 20% is classified by IECC Section 101.3.1 as building Type A-2. Using
this information, determine the maximum UO-value for the gross wall area from Fig. 3 of this Technical Notes to be 0.215
Btu/(hr-ft2-F).

U
O
WallsType A-1 and A-2 Residential BuildingsHeating'
FIG. 3
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
10 of 13 9/13/2009 12:33 PM

Step 2: Calculate Uw using Eq. 1 and knowing the U-values of the glazing and door areas and the gross wall area (Uo).
Equation 1 in this Technical Notes is Eq. 5-1 found in Section 502 of the IECC, solved for Uw.
Eq. 1

U
w
= 0.118 Btu/(hr-ft
2
-F) or R 8.42 (hr-ft
2
-F)/Btu. This Uw-value is the maximum permissible value for wall
constructions having a heat capacity less than 6 Btu/(ft2-F).
Step 3: Determine the HC of the wall in question. The HC of the brick veneer and wood stud wall has already been
calculated to be 7.4 Btu/(ft2-F), see Fig. 1a. The maximum permissible Uw-value for a wall having a HC of 6 Btu/(ft2-F)
or greater may be increased to account for the effects of thermal mass using Tables 3a-3c in this Technical Notes. The
values in these tables, taken from the IECC, are a function of climate (represented by HDD65); wall construction (HC
greater than or equal to 6 Btu/(ft2-F)); and the placement of insulation outside the thermal wall mass (Table 3a), on the
interior of the wall mass (Table 3b) or integral with the wall mass (Table 3c).
In this example, the insulation in the wall shown in Fig. 1a is placed interior of the wall mass. Therefore, Table 3b should
be used. Enter the row in Table 3b for HDD equal to 4001-5500 and the column for Uw equal to 0.118 Btu/(hr-ft2-F).
Uw equal to 0.118 is between the columns in the table labeled 0.10 and 0.12. Linearly interpolate the table to determine
the maximum permissible thermal transmittance, Uw, to be 0.137 Btu/(hr-ft2-F). This U-value corresponds to an R-value
greater than or equal to 7.30 (hr-ft2-F)/Btu.

TABLE 3a
Required Uw for Wall With a Heat Capacity
Equal to or Exceeding 6 Btu/(ft2 oF)
With Insulation Placed on the Exterior of the Wall Mass
TABLE 3b
Required Uw for Wall With a Heat Capacity
Equal to or Exceeding 6 Btu/(ft2 oF)
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
11 of 13 9/13/2009 12:33 PM
With Insulation Placed on the Interior of the Wall Mass

TABLE 3c
Required Uw for Wall With a Heat Capacity
Equal to or Exceeding 6 Btu/(ft2 oF)
With Integral Insulation (Insulation and Mass Mixed, Such as a Log Wall)
Step 4: To determine if the wall in question complies with Section 502 of the IECC, compare the maximum permissible
thermal transmittance, Uw-value, determined in Step 3 with the calculated Uw-value. The calculated Uw-value, determined
using the procedures contained in Technical Notes 4 Revised or the ASHRAE Handbook of Fundamentals, is 0.071
Btu/(hr-ft2-F). Since the calculated Uw-value is less than the maximum Uw-value (0.157 Btu/(hr-ft2-F)), this wall
construction meets the requirements of the IECC Section 502.
For comparison, in this example the maximum permissible Uw-value for a lightweight wall is 0.118 Btu/(hr-ft2-F), but for a
thermal mass wall, the maximum Uw-value is 0.137 Btu/(hr-ft2-F). The allowable Uw-value for the thermal mass wall is 16
percent greater.
Chapter 6 - Simplified Prescriptive Requirements for Residential Buildings, Type A-1 and A-2
Chapter 6 contains a simplified prescriptive approach that does not reflect different percentages of glazing or trade-offs
between building envelope components. It does, however, allow for decreased R-value requirements for mass walls
similar to those found in Chapter 5.

COMMERCIAL CONSTRUCTION - Chapters 7 and 8

Chapter 7 - Building Design for All Commercial Buildings. Chapter 7 of the IECC simply references the requirements
of ASHRAE/IES Energy Code for Commercial and High-Rise Residential Buildings. All commercial buildings must meet
t4b http://www.gobrick.com/BIA/technotes/t4b.htm
12 of 13 9/13/2009 12:33 PM
the requirements of the ASHRAE/IES Energy Code or the requirements of Chapter 8 of the IECC.
Chapter 8 - Design by Acceptable Practice for Commercial Buildings. Chapter 8 of the IECC is applicable to
buildings that have a window and glazed door area not greater than 50 percent of the gross wall area. Buildings with
glazing areas over 50 percent must comply with the ASHRAE/IEC Energy Code. Chapter 8 contains requirements for
individual building components (walls, roof, floors). If any of these requirements are not met, the ASHRAE/IEC Energy
Code can be used for that portion of the building envelope. Differences for mass walls are reflected in the required
values for all but the warmest climates.

SUMMARY
This Technical Notes continues the discussion of the energy efficiency of thermal mass brick masonry walls. Direction is
provided on how to treat thermal mass when considering the envelope requirements of several energy codes or
standards. Methods for complying with these requirements are described in detail. Sample calculations quantifying
thermal mass as heat capacity (HC) are given.
The information and suggestions contained in this Technical Notes are based on the available data and the experience of
the engineering staff of the Brick Industry Association. The information contained herein must be used in conjunction with
good technical judgment and a basic understanding of the properties of brick masonry. Final decisions on the use of the
information contained in this Technical Notes are not within the purview of the Brick Industry Association and must rest
with the project architect, engineer and owner.

REFERENCES
1. Bradshaw, V., Building Control Systems, John Wiley & Sons, New York, NY, 1985.
2. Thermal Mass Handbook, Concrete and Masonry Design Provisions Using the ASHRAE/IES Standard 90.1.
National Codes and Standards Council of the Concrete and Masonry Industries, Herndon, VA, 1993.
3. Energy Conservation in New Building Design (ASHRAE Standard 90A), American Society of Heating, Refrigerating
and Air-Conditioning Engineers, Inc., Atlanta, GA, 1980.
4. Energy Efficient Design of New Buildings Except New Low-Rise Residential Buildings (ASHRAE/IES Standard
90.1-1989 and Addendum-1992), American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. and
Illuminating Engineering Society of North America, Atlanta, GA.
5. Handbook of Fundamentals, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc., Atlanta,
GA, 1981 edition and 1997 edition.
6. Heat Transmission Coefficients of Brick Masonry Walls, Technical Notes on Brick Construction 4 Revised,, Brick
Industry Association, Reston, VA, January 1982.
7. International Energy Conservation Code (IECC), International Code Council (ICC), Falls Church, VA, 2000.
8. International Residential Code (IRC), International Code Council (ICC), Falls Church, VA, 2000.
9. "Report on the Effect of Wall Mass on the Storage of Thermal Energy," Hankins and Anderson, Inc., Richmond, VA
and Boston, MA, 1976.


t4b http://www.gobrick.com/BIA/technotes/t4b.htm
13 of 13 9/13/2009 12:33 PM

Technical Notes 5A - Sound Insulation - Clay Masonry Walls
(Reissued August 2000)
INTRODUCTION
The sound insulation or sound transmission loss of a wall is that property which enables it to resist the passage of
noise or sound from one side to the other. This should not be confused with sound absorption which is that property
of a material which permits sound waves to be absorbed, thus reducing the noise level within a given space and
eliminating echoes or reverberations. Only sound insulation will be discussed in this Technical Notes.
MEASUREMENT OF SOUND
The sound insulation of a building assembly is expressed as a reduction factor in decibels (dB). The decibel is
approximately the smallest change in energy the human ear can detect, and the decibel scale is used for measuring
ratios of sound intensities. The reference sound intensity used to measure absolute noise levels is that
corresponding to the faintest sound a human ear can hear (0 dB). However, a difference of 3 or less dB is not
especially significant, because the human ear cannot detect a change in sounds of less than 3 dB.
Figure 1 shows the intensity level of common sounds on the decibel scale. These data are reproduced from "How
Loud is Loud? Noise, Acoustics and Health", by Lee E. Farr, M.D., published in the February 1970 issue of
Architectural & Engineering News.
SOUND TRANSMISSION LOSS
It is desirable to have a single number rating as a means for describing the performance of building elements when
exposed to an "average" noise. In the past it was customary to use the numerical average of the transmission loss
values at nine frequencies. This rating, termed the nine-frequency average transmission loss, is often quite
inaccurate in comparing an assembly of materials having widely differing TL-frequency characteristics. One single
number rating method which has been recently proposed is the sound transmission class (STC). This rating is
based on the requirements that the value of transmission loss at any of the eleven measuring frequencies does not
fall below a specified TL-frequency contour. The shape of this contour is drawn to represent the more common
types of noise, and generally covers the requirements for speech privacy.
5a http://www.gobrick.com/BIA/technotes/t5a.htm
1 of 9 9/13/2009 12:33 PM
FIG. 1
The following are conclusions in a report entitled, "Measurements of Sound Transmission Loss in Masonry", by
William Siekman of Riverbank Acoustical Laboratories, June 1969.
"In conclusion, changes in results of transmission loss measurements have been studied. They indicate that
deficiencies in earlier test methods and environments have apparently been corrected. Although data reported today
are lower than ever before, they agree very well with data taken in field situations, and consequently provide
5a http://www.gobrick.com/BIA/technotes/t5a.htm
2 of 9 9/13/2009 12:33 PM
assurance that laboratory tests can be relied upon to achieve the desired noise reductions. The performance of
walls near the coincidence frequency cannot be predicted yet on a theoretical basis, nor can the performance of
walls having a compound structure, but test specimen sizes are now large enough to be representative of typical
walls and to provide data over the present frequency range of interest.
"Since the principal deviation due to specimen size is apt to occur at the lower frequencies, users of transmission
loss data are urged to avoid dependence upon single figure ratings, even such a relatively good one as is
recommended by the Proposed Classification for Determination of Sound Transmission Class, ASTM RM 14-2
(1966). The decision to use a particular construction should always be based upon the total curve and the
requirements at individual frequencies."
DESCRIPTION OF SPECIMENS
The specimens discussed in this issue of Technical Notes were constructed at the Riverbank Acoustical
Laboratories in a testing frame having inside dimensions of 14 ft 4 in. wide by 9 ft 4 in. high. The joints were of
typical thickness and were staggered. Mortar was mixed in a ratio by volume of 1 part cement, 2 parts lime and 9
parts sand. All specimens were constructed by a professional mason. The curing time was 28 days or more. The
transmission area, S. used in the computations was generally 126 sq ft.
Following are the descriptions of tests, performed at the Riverbank Acoustical Laboratories starting with the lowest
Sound Transmission Class (STC):
STC 39. 4-in. Structural Clay Tile Wall
Tile dimensions: 3-9/16 by 4-7/8 by 11-3/4 in
Wall thickness: 3-9/16 in.
Average weight: 22.3 psf
Test: TL 67-59
STC 41. 4 in. Structural Clay Tile Wall, with 5/8-in. plaster one face
Tile dimensions: 3-9/16 by 4-7/8 by 11-3/4 in.
Wall thickness: 4-3/16 in.
Average weight: 25.3 psf
Test: TL 67-82
STC 45. 8-in. Structural Clay Tile Wall
Tile dimensions: 7-5/8 by 4-7/8 by 11-3/4 in.
Wall thickness: 7-5/8 in.
Average weight: 40.6 psf
Test: TL 67-69
STC 45. 4-in. Face Brick Wall
Brick dimensions: 2-1/4 by 3-3/4 by 8-1/4 in.
Wall thickness: 3-3/4 in.
5a http://www.gobrick.com/BIA/technotes/t5a.htm
3 of 9 9/13/2009 12:33 PM
Average weight: 38.7 psf
Test: TL 67-70
STC 49. 6-in. "SCR brick" (Reg. U.S. Pat. Off., SCPI) Wall, with 3/8-in. gypsum board
over 1-in. styrofoam insulation one face
Brick dimensions: 2-1/4 by 5-1/2 by 11-1/2.
Wall thickness: 6-7/8 in.
Average weight: 57.7 psf
Test: TL 70-39
NOTE: The styrofoam was placed with adhesive, spot applied 12 in. o.c. both vertically and horizontally, to the brick
wall on one side. A single layer of 3/8 in. gypsum board was applied vertically over the foam with adhesive, spot
applied 12 in. o.c. vertically and horizontally in the field and 6 in. o.c. at the joints. The external joints were finished
with a typical drywall joint system.
STC 50. 8-in. Face Brick and Structural Clay Tile Composite Wall
Brick dimensions: 2-1/4 by 3-3/4 by 8-1/4 in.
Tile dimensions: 4 in. nominal thickness
Wall thickness: 8 in.
Average weight: 63.8 psf
Test: TL 67-65
STC 50. 10-in. Face Brick Cavity Wall, with 2-in. air space
Brick dimensions: 2-1/4 by 3-3/4 by 8-1/4 in.
Wall thickness: 10 in.
Average weight: 81.0 psf
Test: TL 68-31
NOTE: The 2 wythes of masonry were tied together with metal wall ties.
STC 50. 4-in. Brick Wall, with 1/2-in. sanded plaster, two-coat one face
Brick dimensions: 2-1/4 by 3-5/8 by 7-5/8 in.
Wall thickness: 4-1/8 in.
Average weight: 42.4 psf
Test: TL 69-283
STC 51. 6-in. "SCR brick" (Reg. U.S. Pat. Off., SCPI) Wall
Brick dimensions: 2-1/4 by 5-1/2 by 11-1/2 in.
5a http://www.gobrick.com/BIA/technotes/t5a.htm
4 of 9 9/13/2009 12:33 PM
Wall thickness: 5-1/2 in.
Average weight: 55.8 psf
Test: TL 69-286
STC 52. 8-in. Solid Face Brick Wall
Brick dimensions: 2-1/4 by 3-3/4 by 8-1/4 in.
Wall thickness: 8 in.
Average weight: 83.3 psf
Test: TL 67-68
STC 53. 8-in. Solid Brick Wall, with 1/2-in. gypsum board on furring strips one face
Brick dimensions: 2-1/4 by 3-5/8 by 7-5/8 in.
Wall thickness: 9-1/4 in.
Average weight: 86.7 psf
Test: TL 69-287
NOTE: The 3/4-in. collar joint was filled with mortar. Metal Z ties were used between wythes spaced at 24 in. o.c.
both vertically and horizontally. The 1 by 3 wood vertical furring strips were spaced at 16 in. o.c. and nailed at the
mortar joints approximately 12 in. o.c. The gypsum board was applied vertically and attached with nails spaced 12
in. o.c. in the field and 8 in. o.c. along the edges. The joints and nail heads were finished with standard drywall
system.
STC 53. 6-in. "SCR brick" (Reg. U.S. Pat. Off., SCPI) Wall, with 1/2-in. plaster one face
Brick dimensions: 2-1/4 by 5-1/2 by 11-1/2 in.
Wall thickness: 6 in.
Average weight: 60.8 psf
Test: TL 70-70
STC 55. 12-in. Face Brick and Structural Clay Tile Composite Wall
Brick dimensions: 2-1/4 by 3-3/4 by 8-1/4 in.
Tile dimensions: 7-5/8 by 4-7/8 by 11-3/4 in.
Wall thickness: 12 in.
Average weight: 84.1 psf
Test: TL 67-62
STC 59. 12-in. Solid Brick Wall
Brick dimensions:
5a http://www.gobrick.com/BIA/technotes/t5a.htm
5 of 9 9/13/2009 12:33 PM
Face: 2-1/4 by 3-3/4 by 8-1/4 in.
Building: 2-1/4 by 3-5/8 by 8 in.
Wall thickness: 12 in.
Average weight: 116.7 psf
Test: TL 67-32
NOTE: The outside wythes were of face brick. The interior wythe was of common brick.
STC 59. 10-in. Reinforced Brick Masonry Wall (RBM)
Brick dimensions: 2-1/4 by 3-5/8 by 7-5/8 in
Wall thickness: 9-1/2 in.
Average weight: 94.2 psf
Test: TL 70-6
NOTE: The 2-1/4-in. grouted cavity contained No. 6 bars at 48 in. o.c. vertically and No. 5 bars at 30 in. o.c.
horizontally.
SOUND TRANSMISSION CLASS
Sound transmission class contours (see Fig. 2) may be constructed in accordance with ASTM RM 14-2 on
conventional semi-logarithmic paper as follows: a horizontal line segment from 1250 to 4000 Hz (cycles per second);
a middle line segment decreasing 5 dB in the interval 1250 to 400 Hz; and a low frequency segment decreasing 15
dB in the interval 400 to 125 Hz.
5a http://www.gobrick.com/BIA/technotes/t5a.htm
6 of 9 9/13/2009 12:33 PM
The sound transmission loss of the tested specimen is shown by the curved line in the above graph. The broken line
is the limiting sound transmission class contour.
The theoretical transmission loss of that limp mass having the same weight per square foot as the specimen can be
located by drawing a straight line between the two slash marks on the edges of the grid. This was derived from the
equation: TL = 20 log W + 20 log F - 33, where W is weight in pounds per square foot, and F is frequency in Hertz
(cycles per second).
FIG. 2
The STC contour is shifted vertically relative to the test curve until some of the measured TL values for the test
specimen fall below those of the STC contour and the following conditions are fulfilled: The sum of the deficiencies
(that is; the deficiencies of test points below the contour) shall not be greater than 32 dB, and the maximum
deficiency of any single test point shall not exceed 8 dB. The sound transmission class for the specimen is the TL
(transmission loss) value corresponding to the intersection of the sound transmission class contour and the 500-Hz
ordinate.
Table 1 shows the decibel losses for 18 frequencies of test specimens listed above. Deficiencies or deviations from
the contour (see graph) are tabulated to correspond with the proper frequencies.
These measurements were made using a one-third octave bank of pink noise, swept in 13 min from 100 to 5000 Hz.
Runs were made before and after a system interchange, during which the ratio of sound pressure levels in the two
rooms was directly recorded graphically. The final results were obtained by averaging the runs, with a resultant
precision within a 90 per cent confidence limit of 1 dB.
The sound transmission class is computed in accordance with the Tentative Recommended Practice for Laboratory
Measurement of Airborne Sound Transmission Loss of Building Partitions, ASTM E 90-66T, and ASTM RM 14-2.
The STC number is intended to be used as a preliminary estimate of the acoustical properties of the specimen. Final
decisions for design use should be based upon the entire TL curve for the values at all the test frequencies.
5a http://www.gobrick.com/BIA/technotes/t5a.htm
7 of 9 9/13/2009 12:33 PM
1
Transmission Loss, decibels
2
Deficiencies, decibels
MASKING EFFECT
The sound insulation required in a structure to give satisfactory results depends not only upon the noise level outside
of the building or in adjoining rooms, but also upon the noise level within the room under consideration.
If it is to be assumed that there is no noise within the room to be insulated against sound transmission and the noise
level in the adjoining room is 60 dB, it will require a partition having a reduction factor of 60 dB to render the noise in
the adjoining room inaudible. However, if the noise level in the room under consideration is 30 dB, a partition having a
sound reduction factor of approximately 40 dB (see Fig. 3) will make the sound in the adjoining room inaudible.
Experiments have shown that for one sound to mask another, there must be at least 10 dB difference between the
two sounds. This effect of the sound within the room under consideration is known as the "masking effect". Figure 3
illustrates this "masking effect" principle.
5a http://www.gobrick.com/BIA/technotes/t5a.htm
8 of 9 9/13/2009 12:33 PM
Effect of Masking Noise on "Listening" Side of Wall
FIG. 3
CONCLUSION
This issue of Technical Notes has discussed recent test data for, and sound insulation performance of, brick and tile
walls and partitions. Future issues of Technical Notes will contain some suggestions and recommendations for the
control of sound transmission through brick and tile walls and partitions.
5a http://www.gobrick.com/BIA/technotes/t5a.htm
9 of 9 9/13/2009 12:33 PM

Technical Notes 6 - Painting Brick Masonry
Rev [May 1972] (Reissued December 1985)
INTRODUCTION
Although some masonry walls require protective coatings to impart color and help in resisting rain penetration, clay
masonry requires no painting or surface treatment. Brick are generally selected because, among other
characteristics, they have integral and durable color and, when properly constructed, are resistant to rain
penetration.
Clay masonry walls may be painted to increase light reflection or for decorative purposes. Most paint authorities
agree that, once painted, exterior masonry will require repainting every three to five years.
This issue of Technical Notes discusses general applications of paint to interior and exterior brick walls, and a brief
discussion on specific paints suitable for brick masonry.
GENERAL
It is often erroneously assumed that brick masonry walls that are to be painted can be built with less durable
materials and, in some instances, with less than extreme care in workmanship than would normally be used for
unpainted brick walls. This is not the case. When a brick wall is to be painted, the selection of materials, both brick
units and mortar, and the workmanship used in constructing the wall should all be of the highest quality; at least as
good in quality as when the walls are to be left exposed. Every care should be taken to see that joints are properly
filled with mortar to avoid the entrance of moisture into the wall, since it may become trapped behind the paint and
cause problems. Every care should be taken to see that there are no efflorescing materials in the wall, either in the
mortar, brick units or in the backup, since efflorescence beneath the paint film can also cause problems. See
Technical Notes 23 Series.
Brick. Brick units to be used for walls that are to be painted should conform to the applicable requirements of the
ASTM Specifications for Building Brick or Facing Brick, C 62 or C 216, respectively. The grade of units (which
designates their durability) should not be lower than would be used if the wall were not to be painted. Grade SW is
recommended. It may be acceptable to use brick units which are durable but differ in color in a wall to be painted.
However, care should be taken that the units have similar absorption and suction characteristics so that the paint
applied will adhere to all of the surfaces and have a uniform acceptable appearance.
Mortar. Mortar for brick masonry walls to be painted should conform to the Specifications for Mortar for Unit
Masonry, ASTM C 270, Proportion Specifications. It is suggested that the mortar consist of portland cement and
lime, and that the mortar type be selected on the basis of the structural requirements of the wall. See Technical
Notes 8.
Paint. Paint for application to brick masonry walls should be durable, easy to apply and have good adhesive
characteristics. It should be porous if applied on exterior masonry, thereby permitting the wall to breathe and
preventing the trapping of free moisture behind the paint film.
CONSIDERATIONS FOR PAINTING CLAY MASONRY
In selecting a paint system for a brick masonry wall, the primary concern should be the characteristics of the surface
and the exposure conditions of the wall. A primer coat may be of particular importance, especially where unusual or
severe conditions exist.
Alkalinity. The chemical property of masonry which may have a significant effect on paint durability and performance
t6 http://www.gobrick.com/BIA/technotes/t6.htm
1 of 7 9/13/2009 12:34 PM
is the alkalinity of the wall. Brick are normally neutral, but are set in mortars which are chemically basic. Paint
products, which are based on drying oils, may be attacked by free alkali and the oils can become saponified. To
prevent this occurrence, an alkaline-resistant primer is recommended.
Efflorescence. The deposit of water-soluble salts on the surface of masonry, efflorescence, is another factor that
can hamper the performance of painted masonry. Efflorescence, which is present on the surface, should be
removed and, once removed, the surface should be observed for reoccurrence prior to being painted. Methods of
preventing and removing efflorescence are discussed in Technical Notes 23 Series, "Efflorescence-Causes,
Prevention and Control".
Water and Moisture. Water or moisture in a masonry system will generally hamper the satisfactory performance of
the painted surface. Moisture may enter masonry walls in any of several ways; through the pores of the material,
through incompletely bonded or only partially filled mortar joints, copings, sills and projections, through incomplete
caulked joints and improperly installed flashing or where flashing is omitted. In general, brick wall surfaces should be
dry for painting. Acceptable moisture conditions for masonry walls to receive paint are listed in Table 1. The use of
an electrical moisture meter may be used to measure the moisture content of a wall
1
Some manufacturers offer special porous, highly pigmented emulsion paints which may give somewhat better results in very adverse conditions
where delay is not acceptable.
SURFACE PREPARATION
General. Proper surface preparation is as important as paint selection. Because each coat is the foundation for all
future coats, success or failure depends largely upon surface preparation. Thoroughly examine all surfaces to
determine the required preparation. Previously painted surfaces often require the greatest effort. Before painting,
remove all loose matter. Take special care when cleaning surfaces for emulsion paints and primers. They are
nonpenetrating and require cleaner surfaces than solvent-based paints. Some paints can or should be applied to
damp surfaces. Others must not. Be sure to follow directions accompanying proprietary brands.
New Masonry. As a general rule, new clay masonry is seldom painted. It is difficult to justify the extra expenditure
for initial and future painting. However, if for any reason painting new masonry is desired, there are a few
precautions necessary for reasonable success.
Do not wash new clay masonry walls with acid cleaning solutions. Acid reactions can result in paint failures. Use
alkali-resistant paints. If low-alkali portland cement is not used in the mortar, it may be necessary to neutralize the
wall to reduce the possibility of alkali-caused failures. Zinc chloride or zinc sulfate solution, 2 to 3 1/2 lb per gal of
t6 http://www.gobrick.com/BIA/technotes/t6.htm
2 of 7 9/13/2009 12:34 PM
water, is often used for this purpose.
Existing Masonry. Examine older unpainted masonry for evidence of efflorescence, mildew, mold and moss. While
these conditions are not common, they all indicate the presence of moisture. Examine all possible entry points for
water. Where necessary, repair flashing and caulking; tuckpoint defective mortar joints.
Remove all efflorescence by scrubbing with clear water and a stiff brush. A wall which has effloresced for a long
time may present difficulties. The presence of moisture, the deposition of salts and the probable presence of alkalies
are all factors which may contribute to the deterioration of paints.
If moss has accumulated on damp, shaded masonry, apply an ordinary weed killer. Wet the wall with clear water
before applying weed killers to prevent them from being drawn into the wall. Chemical weed killers may contain
solubles which can contribute to efflorescence or react unfavorably with paint, and should be removed after being
used by scrubbing the wall with a stiff brush while rinsing with clear water.
Mildew seldom occurs on unpainted masonry. However, where present, treat it the same as on painted surfaces,
discussed in the following paragraphs. Be sure to wet the wall before applying any cleaning solution. Clean small
areas and rinse thoroughly. For further discussion on cleaning brick see Technical Notes 20 Revised, "Cleaning Clay
Products Masonry".
Painted Surfaces. Previously painted surfaces normally require extensive preparation prior to repainting (refer to
Table 2 for typical paint failures). Under humid conditions, mildew may have developed. Mildew may feed on a paint
film or on particles trapped by the painted surface. If present, remove it completely before applying paint.
Otherwise, growth will continue, damaging new paint. Mildew has been successfully removed by steam cleaning and
sand blasting. The following is also effective:
3 oz trisodium phosphate (Soilax, Spic and Span, etc.), plus
1 oz detergent (Tide, All, etc.), plus
1 qt 5 per cent sodium hyperchlorite (Chlorox, Purex, etc.), plus
3 qt warm water, or enough to make 1 gal of solution.
Use this solution to remove mildew and dirt. Scrub with a medium soft brush until the surface is clean; then rinse
thoroughly with fresh water. For small areas, use an ordinary household cleanser. Scrub with a medium soft brush
and then rinse thoroughly. Use masonry paints containing a mildewcide to help prevent molds from recurring.
Remove all peeled, cracked, flaked or blistered paint by scraping, wire brushing or sand blasting. In some instances,
old paint may be burned off, but this should be done only by skilled operators. Like efflorescence, paint blistering is
caused by water within the masonry. Search for the water's source and take the necessary corrective measures to
keep water out of the wall.
If alligatoring exists, remove the entire finish. There is no other means of correction.
If slight chalking has occurred, brush the surface thoroughly. However, if chalking is deep, remove by scrubbing with
a stiff fiber brush and a solution of trisodium phosphate and water. Rinse the surface thoroughly afterwards. Use a
penetrating primer to improve adhesion of the final coat.
Excessive paint buildup results from too many coats or excessively thick coats. Where it occurs, remove all paint and
treat as a new surface.
Completely remove cement-based paints before repainting with other types. An exception to this rule is the use of
cement-based paints as primers which will be covered by another paint within a relatively short time. If the wall will
be repainted with another cement-based paint, wire brushing and scrubbing will suffice, providing treatments for
mildew, efflorescence, etc. are not required.
TABLE 2
Types of Paint Failure
t6 http://www.gobrick.com/BIA/technotes/t6.htm
3 of 7 9/13/2009 12:34 PM

MASONRY PAINTS
Because all paints have distinct properties and because surfaces vary considerably, even the most experienced
painting contractors carefully examine a surface before making recommendations. However, the following will
generally indicate the proper use of masonry paints.
CEMENT-BASED PAINTS
For many years, cement-based paints have been satisfactory coatings for masonry surfaces. They achieved
popularity because they have relatively good adherence and tendency to make a wall less permeable to free water.
Cement-based paints are permeable, permitting the wall to breathe. Their main components are portland cement,
lime and pigments. Additives, binders and sands may be added.
Although cement-based paints are more difficult to apply than other types, good surface protection results when
properly applied. While they are not complete waterproofers, cement-based paints help to seal and fill porous areas,
excluding large amounts of free water. White and light colors tend to be the most satisfactory. It is difficult to obtain
a uniform coating with darker shades. Lighter colors tend to become translucent when wet, and dark colors become
darker. Color returns to normal as the wall surface dries. Cement-based paints can provide a good base for other
paints applied within a relatively short time.
The following procedure for applying paint on a properly prepared surface generally applies:
1. Cure new masonry walls for approximately one month before applying cement-based paints.
2. Dampen wall surfaces thoroughly by spraying with water.
3. Cement-based paints are packaged in powdered form. Because their cementitious components begin to hydrate
upon contact with water, mix immediately prior to application for optimum results.
4. Apply heavy coats with a stiff brush, allowing at least 24 hr to elapse between coats.
5. During this time, keep the wall damp by periodically spraying it with water.
t6 http://www.gobrick.com/BIA/technotes/t6.htm
4 of 7 9/13/2009 12:34 PM
6. Apply additional coats in the same manner.
7. Keep the final coat damp for several days to properly cure.
WATER-THINNED EMULSION PAINTS
General Characteristics. Water-thinned emulsion paints, commonly referred to as latex paints, are relatively easy
to apply. Water-thinned emulsions may be brush, roller or spray-applied. However, brush application is preferable,
especially on coarse-textured masonry. Emulsion paints dry quickly, have practically no odor and present no fire
hazard. They may be applied to damp surfaces, permitting painting shortly after a rain or on walls damp with
condensation.
As a group, these paints are alkali-resistant. Hence, neutralizing washes and curing periods are not usually
necessary before painting. Water emulsion paints possess high water vapor permeability and are known to have
performed well on brick substrates that have been properly prepared.
Emulsion paints will not adhere well to moderately chalky surfaces. If possible, repainting should be done before the
previous coat chalks excessively. However, specifically formulated latex paints are available containing emulsified
oils or emulsified alkyds which facilitate wetting of chalky surfaces. This property enables the paint to bond the chalk
together and to the substrate.
The principal water-thinned emulsion paint types are: butadiene-styrene, vinyl, acrylic, alkyd and multicolored
lacquers.
Butadiene-Styrene Paints. These relatively low-cost, rubber-based latex paints develop water resistance more
slowly than vinyl or acrylic emulsions. They are most satisfactory in light tints as chalking rate may be excessive in
deep colors.
Vinyl Paints. Polyvinyl acetate emulsion paints dry faster, have improved color retention and a more uniform, lower
sheen than rubber-based latex paints.
Acrylic Emulsion Paints. Acrylic emulsions have excellent color retention, permit recoating in 30 min or less, and
have good alkali resistance. Acrylics have high resistance to water spotting and may be scrubbed easily.
Alkyd Emulsion Paints. Alkyd emulsions are related to solvent-thinned alkyd types, but have all the general
characteristics of latex paints. They do have more penetration than most water-thinned emulsions, achieving better
adhesion on chalky surfaces. Compared to other emulsion paints, these are rather slow to dry, have more odor, are
not as resistant to alkalies, and have poorer color retention. Under normal exposure conditions, alkyd emulsions can
serve as a finished coat over a suitable primer.
Multicolored Lacquers. A specialized paint group, multicolored lacquers are applied only by spray gun. The finished
film appears as a base color with separate dots or particles of contrasting colors. These paints will cover many
surface defects and irregularities. However, they must be applied over a base coat of another type; for example,
polyvinyl acetate or acrylic emulsion paints.
FILL COATS
Fill coats are base coats for exterior masonry. They are similar in composition, application and uses to
cement-based paints. However, fill coats contain an emulsion paint in place of some water, giving improved adhesion
and a tougher film than unmodified cement paints. Fill coats have greater water retention, giving the cement a better
chance to cure. This is particularly valuable in arid areas where it is difficult to keep the painted surface moist during
the curing period.
SOLVENT-THINNED PAINTS
The five major solvent-thinned paints are oil-based, alkyd (synthetic resin), synthetic rubber, chlorinated rubber and
epoxy. Oil-based and alkyd paints are not recommended for exterior masonry. Solvent-thinned paints should be
applied only to completely dry, clean surfaces. They produce relatively nonporous films and should be used only on
interior masonry walls not susceptible to moisture penetration. The exception to this is special purpose paint, such as
t6 http://www.gobrick.com/BIA/technotes/t6.htm
5 of 7 9/13/2009 12:34 PM
synthetic rubber, chlorinated rubber and epoxy paints.
Oil-Based Paints. Oil-based paints have been used for many years. They are relatively non-porous and
recommended for interior use only. Although several coats may be required for uniform color and good appearance,
they bind well to porous masonry. As with most solvent-based paints, they have good penetration on relatively
chalky surfaces, but are highly susceptible to alkalies. New masonry must be thoroughly neutralized to avoid
saponification. Available in a wide color range, oil-based paints are moderately easy to apply. Several days' drying is
generally required between coats.
Alkyd Paints. Alkyd paints are similar to oil-based paints in most general characteristics. They may have slightly
less penetration, resulting in somewhat better color uniformity at the cost of adhering power. Alkyd paints are more
difficult to brush, dry faster and give a harder film than oil-based paints. These, too, are nonpermeable and are
recommended for interior use only.
Synthetic Rubber and Chlorinated Rubber Paints. These paints have excellent penetration and good adhesion to
previously painted, moderately chalky surfaces as well as new surfaces. They are reported to be more resistant to
efflorescence and are generally good in alkali resistance. They may be applied directly to alkaline masonry surfaces,
but are more difficult to brush on than oil paints. Darker colored synthetic rubber paints lack color uniformity. Both
types have high resistance to corrosive fumes and chemicals. For this reason, they are often specified for industrial
applications. Both types require very strong volatile solvents, a fire hazard which may prove undesirable.
Epoxy Paints. Epoxy paints are of synthetic resins generally composed of two parts, a resin base and a liquid
activator. They must be used within a relatively short time after mixing. Epoxies can be applied over alkaline
surfaces, have very good adhering power, and good corrosion and fume resistance. However, some types chalk
excessively if used outdoors. Epoxies are relatively expensive and somewhat difficult to apply.
"HIGH-BUILD" PAINT COATINGS
High-build paint coatings are generally used on interiors to give the effect of glazed brick. Some coatings are based
on two-component urethane polyesters and epoxies. Others are of an emulsion-based coat with acrylic lacquer.
These paint systems usually include fillers to smooth out surface irregularities.
OTHER COATINGS
Heavily applied coatings of the so-called "breathing type" are available with either a water or solvent base. They are
generally composed of asbestos fiber and sand, and applied thickly to hide minor surface imperfections. The
presence of moisture on the surface of a masonry wall generally will not harm the latex type. Lower application
temperatures of 35 F to 50 F on the other hand are less damaging to the solvent type.
For both types, adhesion is mostly mechanical because of low binder and high pigment content. Some coatings
require special primers to insure adhesion. Although these coatings are reported to have given good performance on
masonry, they tend to show stains where water runoff occurs.
These coatings are capable of allowing passage of water vapor, but cannot transmit large quantities of water that
may enter through construction defects. Failure may occur as a result of freezing of water accumulation behind the
film.
PAINTING NEAR UNPAINTED MASONRY
Often windows and trim of masonry buildings are painted with self-cleaning paints to keep surfaces fresh and clean.
Unfortunately, self-cleaning is generally achieved through chalking. The theory is that rain will wash away chalked
paint, constantly exposing a fresh paint surface. The theory works well, but too often no provision is made to keep
chalk-contaminated rain water away from masonry surfaces. The result is usually more unsightly than dirty paint on
trim or windows. Avoid this staining by choosing nonchalking paints for windows and trim and by providing a
means of draining water away from wall surfaces.
REFERENCES

t6 http://www.gobrick.com/BIA/technotes/t6.htm
6 of 7 9/13/2009 12:34 PM
1. Manual on the Selection and Use of Paints, Technical Report #6, National Research Council of Canada,
Division of Building Research, 1950, Ottawa, Canada.
2. Paints for Exterior Masonry Walls, BMS110, National Bureau of Standards, 1947, Washington, D.C.
3. Field Applied Paints and Coatings, Publication 653, Building Research Institute, 1959, Washington, D.C.
4. Paints and Coatings, Publication 706, Building Research Institute, 1960, Washington, D.C.
5. Painting Walls; 1, Building Research Station Digest (2nd Series), No. 55, Building Research Station, 1965,
Garston, Herts., England.
6. Coatings for Masonry Surfaces, by H. E. Ashton, Canadian Building Digest, CBD 131, November 1970,
Ottawa, Canada.
7. Coatings for Masonry and Cementitious Materials, by Walter Bayer, Construction Specifier, November
1970, Washington D.C.
t6 http://www.gobrick.com/BIA/technotes/t6.htm
7 of 7 9/13/2009 12:34 PM
2008 Brick Industry Association, Reston, Virginia Page 1 of 14
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
6A
August
2008
Colorless Coatings for Brick Masonry
Abstract: This Technical Note discusses common reasons for applying colorless coatings to above-grade brick masonry and
the appropriateness of such actions. The types of products often used and the advantages and disadvantages of each are
presented. Issues to consider prior to application of a clear coating to brick masonry are provided.
Key Words: clear, colorless coatings, film former, graffiti-resistant, penetrant, silane, siloxane, water penetration, water
repellent.
General
Application of a water repellent coating is not necessary
to achieve water resistance in brickwork subjected to
normal exposures where proper material selection,
detailing, construction and maintenance have been
executed
Application is not recommended on newly constructed
brick veneer or cavity walls or on new or existing
pavements using clay pavers
Correct conditions contributing to water penetration
before applying a coating to brickwork
Consider providing vents at top of drainage spaces when
a water repellent coating is applied
General Selection Criteria
Consult the brick manufacturer prior to the selection of a
coating
Select only coatings intended for use on clay brickwork
Consider the effects of all coating properties on
brickwork, not just the desired property
Select coatings that have demonstrated consistent
performance on similar installations, materials and
exposures for a minimum period of five years

Except for anti-graffiti applications, use only breathable


coatings with a water vapor permeability of 0.98 or
greater as measured by ASTM E96
Consider the use of a siloxane or siloxane/silane pre-
blended coating
Use comparative testing of treated and untreated walls
using ASTM E514 or ASTM C1601 to determine coating
effectiveness
Do not apply film-forming coatings to brickwork located in
freeze-thaw environments
Specific Selection Criteria
For exterior brickwork, consider a condensation analysis
to determine whether coating affects the dew point
location within the wall
For paving, consider the effects of coating on pavement
slip resistance and the abrasion resistance of the coating
Application
Use a contractor with a minimum of five years experience
installing selected coating on similar installations
Apply the coating according to the coating manufacturer
directions

INTRODUCTION
Colorless coatings are available in many types and are designed for a variety of uses. When needed, colorless
coatings for brick masonry should be selected based on their intended use, documented performance and
chemical and physical properties [Refs. 4, 6, 10, 13]. Clear coatings formulated for use on other masonry materials
may not be appropriate for brick masonry and may in fact be detrimental to brick. Clay brick masonry has physical
and chemical properties that are different from stone, concrete or concrete masonry. Brick masonry has a different
pore structure and is generally less absorptive, less permeable and less alkaline than concrete masonry. The
recommendations included herein are applicable only to clay brick masonry.
The type of exposure the brickwork is subject to also plays an important role in coating selection. Coatings suitable
for interior brick masonry may not be suitable for exterior exposures. Similarly, coatings applied to floors or
pavements are subject to conditions different from those in brick walls.
Specific recommendations regarding the reasons for, selection of and use of colorless coatings are found
throughout this Technical Note. Opaque coatings, such as damp-proofing or waterproofing coatings, are not
addressed. For further information about opaque coatings, refer to Technical Note 6, which covers painting of brick
masonry.
SUMMARY OF RECOMMENDATIONS:
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 2 of 14
REASONS FOR USE
Clear coatings may be applied to brick masonry in an effort to facilitate cleaning, to resist graffiti, to provide gloss
or to reduce water absorption or penetration. Often, a single product is used to achieve several of these objectives.
Selection of a coating should be based on the desired appearance, resistance to water penetration, application of
brickwork, material substrate, economics, life span or other criteria set by the designer or user. The disadvantages
of using colorless coatings should also be considered during selection.
Water Penetration Resistance
It is desirable to minimize the penetration and absorption of water in brickwork to avoid problems encountered in
walls. Problems caused by excessive water penetration include freezing and thawing deterioration; corrosion of
metal ties, metal studs and other items; rotting of wood members; mold growth; and damage to interior finishes.
The most effective means of minimizing water penetration include exercising care during material selection,
designing and detailing brick masonry properly, constructing high-quality brickwork, and performing proper
maintenance. Detailed discussions of these issues are provided in the Technical Note 7 Series. Drainage-type
walls, such as brick veneer and cavity walls, are designed to accommodate water penetration of the exterior
brickwork without damage to the interior components of the wall system through its drainage system.
Nonetheless, water-repellent coatings are sometimes suggested to reduce the amount of water that penetrates
brickwork. Research indicates varied effectiveness of clear water repellents in reducing water leakage through a
brick masonry wythe. [Refs. 3, 7, 11] Water-repellent coatings are most effective at reducing the amount of water
absorbed by brick masonry. But water usually penetrates brick masonry at separations and cracks between brick
and mortar or at junctures with other materials. Thus, a change in the absorption properties of brick masonry
provided by a water-repellent coating may not significantly reduce water penetration through brickwork. Water-
repellent coatings cannot stop water penetration caused by design or construction deficiencies such as ineffective
sills, caps or copings, or incompletely filled mortar joints. Penetrating water-repellent coatings seldom stop water
penetration through cracks more than 0.02 in. (0.5 mm) wide, and their effectiveness under conditions of wind-
driven rain is limited. As a result, the use of water-repellent coatings to eliminate water penetration in a wall with
existing defects can be futile.
Water repellents can be useful for barrier walls, chimneys, parapets and other brickwork that is particularly
vulnerable to water absorption and penetration, especially in climates that receive large amounts of rain. When
a water-repellent coating is considered for use on these elements, the benefits must be weighed against the
possible disadvantages. Past successful performance of the proposed coating, for a number of years in the
same exposure conditions and on the same type of brick and mortar, should be required. In climates that
experience freezing and thawing cycles, the effect of a coating on the durability of the brickwork is of particular
concern.
The age of construction and limitations of different types of water repellents are described in the sections that
follow. Methods for evaluating the effectiveness of water repellents are discussed under Performance Criteria.
New Construction Use. Water repellents sometimes are specified for newly constructed brick masonry to protect
against water penetration due to imperfections in construction. As discussed previously, water repellents have
limited effectiveness and cannot compensate for poor construction or design. Furthermore, most brick masonry
wall systems do not require a water repellent to effectively manage water and prevent water intrusion into the
interior of a building. For these reasons, the use of water repellents on newly constructed drainage walls is not
recommended.
Remedial Use. Water-repellent coatings most often are applied in an attempt to reduce or eliminate water
penetration in existing brickwork experiencing water penetration problems. As noted previously, water repellents
cannot prevent water from penetrating cracks wider than 0.02 in. (0.5 mm). Therefore, the source of water
penetration should be determined and necessary repairs completed prior to the application of a water-repellent
coating. Exterior walls should be inspected to determine the condition of caps and copings, flashing, weeps,
sealant joints, mortar joints, brick units and general execution of details. Technical Note 46 provides an inspection
checklist for areas of concern. Repair and replacement of missing, broken, failed or disintegrating items identified
during the inspection and essential to the water resistance of the brickwork should be completed prior to
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 3 of 14
application of a water repellent. The application of a water repellent is rarely effective and is not recommended in
lieu of the following common repairs:
1. Removal of failed sealant, and cleaning, priming and replacement with an appropriate grade of elastomeric
sealant at all windows, copings, sills, expansion joints and between brick masonry and other materials.
2. Repointing of incompletely filled, cracked or disintegrated mortar joints.
3. Removal and replacement of brick with spalled faces or cracks extending through the face shell.
4. Surface grouting of separations between the brick units and the mortar.
These remedial measures are described in Technical Note 46.
Other repairs, which are generally more difficult and costly to complete, include the following:
1. Clearing of mortar blockage from weeps and the air space or cavity.
2. Removal and replacement of damaged, omitted or improperly installed flashing.
The latter repairs are considered by some to be unnecessary or uneconomical if a water repellent is applied.
However, these repair techniques provide long-term solutions to water penetration problems. Not completing them
may allow water within a wall to become trapped, resulting in failure of the coating or deterioration of brickwork.
After remedial repairs have been completed and inspected, it is advisable to wait a period of several months to
determine whether a water repellent is necessary. Moisture penetration problems often will be corrected by these
initial repairs, and further consideration of coatings can be dismissed.
If water penetration remains a problem, or long-term solutions are judged to be too costly despite their benefits, the
application of a water repellent can be considered. If water absorption appears to be the problem, a water repellent
can be particularly effective. However, water repellents are not a permanent solution and will require reapplication.
See the discussion under Durability of Coating for further information on the life span of coatings.
Stain Resistance and Efflorescence Prevention
By reducing the amount of water absorbed by brickwork, colorless coatings may help reduce staining and
efflorescence. As a result, colorless coatings are sometimes used on brickwork that is subject to severe exposures
or on units that have a relatively high absorption. Brick manufacturers sometimes apply colorless coatings to units
during manufacture to reduce staining or initial rate of absorption. ASTM standards for face brick require that
the brick manufacturer report the presence of such coatings. Selection of a coating for any of these uses should
be based on demonstrated successful performance on similar brick with comparable exposures. Staining and
efflorescence may not be completely eliminated by application of a coating. If staining or efflorescence occur on
masonry treated with a colorless coating, the stains and salts may be difficult or impossible to remove. Further, for
film-forming coatings and water repellents with a vapor permeability less than 0.98, efflorescing salts may become
trapped under the coating, causing damage to the brick.
Appearance Change
Another common reason for using a colorless coating
is to achieve a darker, wet or glossy appearance.
In some cases, a colorless coating may result in
an undesired sheen or gloss. Such gloss may be
an indication of an improperly applied coating or of
poor coating selection (see Photo 1). Satisfactory
appearance of a treated surface is best judged by
examining a sample panel or test area of masonry
before and after treatment.
Graffiti Resistance
Resistance to graffiti and ease of cleaning can be
important attributes for public structures such as
schools, government buildings, libraries and noise
barrier walls, where brick masonry is chosen for its
Photo 1
Undesired Gloss Due to a Colorless Coating
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 4 of 14
appearance and low maintenance. Colorless coatings are sometimes applied to brick masonry to keep graffiti or
dirt on the surface of the brickwork for easier removal. Glazed brick often are used in similar installations to provide
the same benefits. Note that some coatings used for graffiti resistance are sacrificial, meaning that the coating
itself is removed when the graffiti is removed.
TYPES OF COLORLESS COATINGS
Colorless coatings for brick masonry can be classified into two general categories: film formers and penetrants.
The two types have significantly different physical properties and performance. As the name implies, film formers
produce a continuous film on the surface of the masonry. Penetrants enter up to in. (10 mm) into the brick
masonry and do not form a surface film.
Colorless coatings may be either waterborne or solvent-borne. Carrier type influences permissible application
conditions. Originally, better penetration and performance were attained using solvent-borne solutions. However,
manufacturers are increasingly producing waterborne solutions that have lower volatile organic compound (VOC)
content. Coatings with higher solids content also may have lower VOC content. VOC content is regulated by
the Environmental Protection Agency nationwide because of its connection with poor air quality. In addition,
many green building guidelines have limits on VOC content in coatings. Product data and test results should
be examined carefully to compare performance. Temperature range, substrate moisture content, environmental
regulations and effects on adjacent materials and vegetation must be considered.
Colorless coatings are discussed in the following sections according to generic chemical type. Most colorless
coating manufacturers will provide information on the generic chemical composition of their products. In addition,
handbooks are available that classify many proprietary coatings according to their generic chemical composition.
Film Formers
Typically, film-forming products adhere to the brick masonry and form a film on the surface. Surface preparation
can be important in achieving satisfactory adhesion of a film-forming coating. Film-forming products should be
applied only to dry surfaces. Film materials, continuity and product concentration determine the performance
characteristics.
Film-forming products are effective at preventing water from penetrating into brick masonry. Film formers can
bridge the small, hairline cracks that are commonly the source of water penetration. If the crack is active, such as
one created by wind load or thermal fluctuations, a film-forming product may also crack. This obviously reduces its
effectiveness. However, a film-forming product's ability to exclude water from the exterior also inhibits evaporation
of water within the masonry through the exterior face and can result in clouding (see Photo 2) and spalling
(see Photo 3) if the source of moisture is not addressed. The reduced water vapor transmission rate, or lack of
breathability, is of special concern in exterior brick masonry subject to freezing and thawing cycles. Thus, film-
forming products are not recommended for brick masonry in such environments.
Photo 2
Clouding of Brick Masonry Wall
with a Film-Forming Coating
Photo 3
Spalling of Brick Masonry Wall
with a Film-Forming Coating
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 5 of 14
A film on a masonry wall may facilitate cleaning by keeping surface contaminants from penetrating into the
masonry. This characteristic leads to such products' use as graffiti-resistant coatings. When an appearance
change is desired, film formers typically are used. Film-forming products, by their nature, tend to produce a sheen
or gloss when applied. When used in high concentrations, they may darken the appearance of a wall (the wet
look).
Acrylics, stearates, mineral gum waxes and urethanes are among the products that form a film when applied to
brick masonry. The large molecular size of these products prevents them from penetrating into the masonry.
Acrylics. Acrylics can be effective as water repellents. They often are used when a high gloss is desired. Acrylics
are available in two forms, waterborne and solvent-borne. Acrylic emulsions are waterborne. Acrylic solutions
are solvent-borne. Because of increasing regulation of solvent-borne products, acrylic emulsions are more widely
used. Coating manufacturers typically recommend that acrylics be applied to substrates that are thoroughly dry.
If applied to a damp substrate, the acrylic film can separate from the masonry, giving it a cloudy, or whitened,
appearance. Some acrylics can create a slippery surface, which is a concern in pavements. However, some
acrylics increase slip resistance. When stabilized against degradation in ultraviolet (UV) light, acrylics can last
approximately five to seven years.
Stearates. Stearates promoted for use on masonry are generally aluminum or calcium stearates. They are
sometimes known as metallic soaps. Stearates form a water-repellent surface by reacting with free salts in
mineral building materials and plugging the pores. Some formulations are used as integral water repellents in
concrete masonry and mortar. Their effectiveness as applied water repellents varies, and typically film-forming
stearates must be reapplied every year. Stearates also have the potential to turn cloudy if moisture gets behind
the coating.
Mineral Gum Waxes. Paraffin wax and polyethylene wax are commonly referred to as mineral gum waxes.
These products are typically solvent-borne and can be good water repellents, able to bridge hairline cracks. As
with other coating types, mineral gum waxes can be used to protect units from staining. However, they have
been known to darken the substrate and, in cases where moisture gets behind the coating, turn the surface a
milky white. If the sources of moisture are not addressed, clouding and eventual spalling of the masonry may
occur.
Urethanes. Urethanes, chemically referred to as polyurethanes, are isocyanate resins. They are classified
as either aromatic or aliphatic, depending on the resulting chemical. They are considered one-part urethanes
if cured by moisture in the substrate or air and two-part if they require a chemical catalyst to cure. While
urethanes can be excellent water repellents and provide good gloss, they can break down under UV light and
have very low vapor permeability. Chemical additives often are used in urethanes to prevent UV degradation
and yellowing and to improve gloss retention. Urethanes with such additives usually last from one to three
years.
Penetrants
Penetrating type coatings are characterized by their penetration into the substrate, typically to depths up to in.
(10 mm). They repel water by changing the capillary force, or contact angle with water, of the pores in the face of
the masonry from positive (suction) to negative (repellency). Penetrating coatings are typically more resistant to
UV degradation because of their chemical composition and because they penetrate below the masonry surface.
Because they coat the pores rather than bridge them, penetrants tend to have better water vapor transmission
characteristics. The solids content of these materials commonly ranges from 5 to 40 percent by weight. Higher
solid content typically indicates better water penetration resistance. Penetrants can be categorized into six
groups siloxanes, silanes, silicates, methyl siliconates silicone resins and RTV silicone rubber and blends of
these.
Siloxanes. Siloxanes have a larger molecular structure than silanes and provide good penetration and water
repellency. Siloxanes bond chemically with silica- or alumina-containing materials, such as brick and mortar, to
make the material water-repellent. This results in a long life, up to 10 years or more, and makes the coating more
difficult to remove. Some siloxanes can also be applied to slightly damp surfaces. Siloxanes are less volatile than
silanes and react with chemically neutral substrates without a chemical catalyst. Siloxanes are typically used
in solutions having 5 to 12 percent solids by weight. Siloxanes have been known to work well on certain brick
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 6 of 14
masonry installations. However, siloxanes are highly reactive with silica and will bond with glass that is not properly
protected.
Silanes. Silanes used as clear water repellents have a smaller molecular structure than siloxanes, which allows
good penetration on dense substrates. They are used in relatively high concentrations (typically 20 percent or
greater solids content). Like siloxanes, silanes bond chemically with silica- or alumina-containing materials and
can bond with unprotected glass. Silanes can be applied to slightly damp substrates. An alkaline substrate, such
as concrete or concrete masonry, acts as a catalyst to speed the reaction to form a water-repellent surface.
Chemical catalysts also are used with silanes to improve the chemical reaction on less alkaline substrates such
as brick.
Silicates. Ethyl silicates are commonly used in restoration of deteriorated masonry as consolidants for natural
stone and occasionally brick masonry. Consolidants are designed to react with and stabilize the substrate to
which they are applied. Their use on brick is uncommon. None are effective water repellents, and they are not
recommended for this use on brick masonry.
Methyl Siliconates. Methyl siliconates are alkaline solutions that react with silica-containing materials in the
presence of carbon dioxide to form a water-repellent surface. Siliconates are sometimes injected into brick
masonry to form a horizontal barrier to rising damp.
Silicone Resins. Silicone resins come in many weights and forms. The 5 percent silicone resin is the most
common penetrating formula. Silicones do not chemically bond with the substrate and as a result have a short life.
Many silicones require reapplication on a yearly basis, although some last longer.
RTV Silicone Rubber. Room temperature vulcanizing (RTV) silicone rubber is a penetrating water repellent
that contains petroleum distillates. It does not require the presence of alkali to react with the substrate. Once
cured, RTV silicone rubber retains its elasticity, helping it to bridge hairline cracks. Asphalt, plastic rubber and
glass surfaces must be protected from contact with it. RTV silicone rubber is commonly used in anti-graffiti
coatings.
Blends. Colorless coatings also are made from blends of the materials listed above. Blends are created to
produce products with the benefits of the constituent materials. As such, they reflect the properties of the
constituent materials, but the properties will be modified somewhat. Thus, it is important to review product data
and test results for products, especially blended ones. For quality assurance that a blend is formulated in the
correct proportions, select a product that is pre-blended by the manufacturer.
PERFORMANCE CRITERIA
Any coating applied to brick masonry will change the physical properties of the masonry. The most critical
properties of colorless coatings to be evaluated are water vapor transmission, water penetration and repellency,
durability, compatibility with the substrate, gloss, slip resistance, graffiti resistance, VOC content and environmental
considerations. A variety of industry standard tests for evaluating these properties exist; however, it can be difficult
to compare products because the reported performance characteristics of each product may be based on a
different set of tests.
Another difficulty exists in correlating test results with in-service performance of coatings applied to brickwork. For
example, one method of evaluating water repellency of a coating is by comparing the cold water absorption of an
untreated brick to that of a treated brick, using the method described in ASTM C67, Test Methods of Sampling
and Testing Brick and Structural Clay Tile. Although such a test may indicate the ability of a coating to reduce
the amount of water absorbed through the faces of individual brick, it neglects the effect of mortar joints on the
water penetration resistance of brickwork. The presence of partially filled mortar joints, hairline cracks and minute
separations that occur in brickwork will often reduce, and sometimes completely negate, the tested effectiveness
of a coating.
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 7 of 14
Until standard tests better correlate with performance of brickwork in service, good judgment and experience are
necessary in establishing performance criteria. The properties discussed in the following sections can be useful
in comparing colorless coating alternatives. Table 1 presents a relative comparison of several colorless coating
properties.
TABLE 1
Typical Properties of Colorless Coatings for Brick Masonry
1
Water Vapor
Transmission
Water
Repellency
Life Span,
Years
Available with
Glossy Finish
Graffiti
Resistance
Film Formers
Acrylics Poor Very good 5 to 7 Yes Yes
Stearates Poor Varies 1 No No
Mineral gum waxes Poor Good Varies No No
Urethanes Poor Very good 1 to 3 Yes Yes
Penetrants
Siloxanes Very good Very good 10+ No No
Silanes Very good Very good 10+ No No
Silicates Poor Poor Varies No No
Methyl siliconates Good Fair Varies No No
Silicone resins Fair Varies 1 Yes No
RTV silicone rubber Good Good 5 to 10 No Yes
Blends Varies Varies Varies No No
1. Refs. 6, 14
Water Vapor Transmission
Rate and Permeability
The most important property to consider when
selecting a coating for application on exterior brick
masonry is the water vapor transmission rate. The
water vapor transmission, or breathability, determines
the rate and amount of water that can evaporate
through the face of the brickwork. Coatings that have
low water vapor transmission rates inhibit evaporation
and can trap water within the brickwork, leading to
clouding of the coating, as shown in Photo 2 and
Photo 4.
Low water vapor transmission may also result in the
premature deterioration of brickwork. Water that is
unable to pass through a coating increases risks of
masonry deterioration due to freeze-thaw cycles and
deposition of water-soluble salts behind the coating.
As these salts crystallize, they grow significantly in
size and can create enough expansive pressure to
cause spalling of brick.
For these reasons, the effect of a coating on the water vapor transmission rate of brickwork should be carefully
considered, particularly for walls exposed to freezing and when moisture problems such as rising damp and
condensation are known to exist. Coatings with a water vapor permeability of 0.98 or higher allow natural
evaporation to occur, thus reducing the potential for problems. However, even a highly breathable coating may
lower the vapor transmission of a wall by preventing moisture migration to the exterior surface where evaporation
occurs. A condensation analysis, as described in Technical Note 47, should be performed before applying a
coating to determine the effect of the coating on the location of condensation within the wall system.
Photo 4
Clouding of a Colorless Coating
on a Brick Pavement
M
i
k
e

D
i
c
k
e
y
,

F
r
i
e
z
e

&

A
s
s
o
c
i
a
t
e
s
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 8 of 14
At present, there is no definitive test establishing the effect of colorless coatings on the water vapor transmission
rate and durability of brick masonry. However, the water vapor transmission rate of a coating can be measured
using ASTM E96, Test Methods for Water Vapor Transmission of Materials. To accurately replicate field conditions
(air rather than water on one side of the brick), the desiccant method is preferred. Using this test, the effect of a
coating can be evaluated through a comparative measurement between an untreated and a treated brick. For
comparative testing, a maximum 10 percent reduction in the rate of vapor transmission is the recommended limit.
Another method to evaluate the potential of a colorless coating to entrap damaging salts and cause spalling is
proposed by Binda [Ref. 2]. Individual brick are treated with the colorless coating on their exposed faces. The
sides of the units are sealed with rubber to prevent evaporation except through the treated face. The units are
subjected to cycles of immersion in a salt solution for four hours and air drying for 44 hours. The cross-sectional
size is measured after each cycle. Deterioration is typically by delaminations of the treated brick face, hence a
reduction in brick cross section. A correlation of the number of cycles to deterioration in this test to the durability
of a masonry assemblage has not yet been established. However, this method is one means of assessing salt
crystallization damage potential when evaluating colorless coatings.
Water Repellency
Water repellency is an important criterion when a coating is intended to reduce water penetration resistance.
However, water repellency of most coatings is based on reducing the amount of water absorbed by a substrate.
Water repellency is often evaluated by comparing the absorptions of treated and untreated brick using the
ASTM C67 test for cold water absorption. As discussed previously, this approach has significant limitations.
Because most water penetrates brickwork through voids or cracks in mortar joints and minute separations between
brick and mortar, tests of water repellents on individual brick cannot accurately indicate the performance of a
water repellent on brickwork. The effectiveness of water-repellent coatings in reducing water penetration through
brickwork is more accurately evaluated by using representative brickwork panels.
ASTM E514, Test Method for Water Penetration and Leakage Through Masonry, is the preferred laboratory
test for evaluating the ability of a coating to reduce the water penetration of brickwork. The test can be used
to compare the water penetration resistance of brickwork treated with water-repellent coatings to untreated
brickwork. Testing should be performed on a minimum of three identical wall specimens of the intended materials
and construction. The amount of water penetration should be measured on each specimen in accordance with
ASTM E514 before and after coating with the clear water repellent. The percentage reduction in water penetration
is a measure of the water repellent effectiveness. A 90 percent reduction in maximum leakage rate; and a
75 percent reduction in percent area of dampness on the back face of the wall and total water collected after 24
hours of testing [Ref. 3] as compared to the untreated wall panel is recommended. ASTM E514 has its limitations.
Performance of coatings in laboratory tests may differ from results on actual brickwork due to the variables
inherent in construction. Thus, a tested percent reduction rate for a laboratory test does not automatically translate
into the same percent reduction in water leakage through the exterior brickwork of a constructed building.
ASTM C1602, Test Method for Field Determination of Water Penetration of Masonry Wall Surfaces, provides a
means to evaluate the effectiveness of a coating in the field. The test can be used on existing masonry walls or
field mock-ups. A sheet of water is to be developed and maintained on the wall surface during testing. If the sheet
of water does not consistently form, the results of this test may be inaccurate. After a preconditioning period, a
specified water flow rate and air pressure are maintained. The amount of water applied to the face of the wall
during the test is measured and the water loss calculated. Again, a coating should provide at least a 75 percent
reduction in loss of water.
Durability of Coating
The durability of a coating is an important selection criterion. Greater depth of penetration or film thickness and
greater resistance to degradation in UV light and harmful environments imply longer life for coatings applied to
exterior brickwork. Durability of coatings applied to brick pavements may also depend on resistance to abrasion.
A coatings durability also determines how often it must be reapplied, which may have permeability and ongoing
maintenance implications.
Most coatings must be reapplied every five to 15 years, and some last considerably shorter periods of time. Many
coatings are warranted by the manufacturer to last 10 years or more. It is common for film-forming products to
require reapplication more often than penetrants, particularly if they are applied to brick floors subject to significant
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 9 of 14
amounts of traffic. Evaluation of a coatings resistance to abrasion is difficult, because there are no direct test
methods for measurement on brick. Reapplication of a coating (especially if carried out prematurely) may decrease
the vapor permeability of the brickwork. This may be a concern for exterior brick masonry walls, particularly in
climates subject to freezing and thawing.
One way to evaluate the durability of a coating is with laboratory tests that simulate outdoor exposure.
ASTM G154, Practice for Operating Fluorescent Light Apparatus for UV Exposure of Nonmetallic Materials, is one
often specified. The difficulty with using laboratory tests to measure the life span of a coating is trying to correlate
laboratory test results to field performance. Coating characteristics, such as gloss or water repellency, can be
measured before and after exposure and the results compared, but such tests have not been correlated to the
actual life expectancy of the coating.
Periodic evaluations of field performance can also be used to determine whether a coating continues to be
effective. Results of field tests conducted on a specified area of a newly treated wall can be compared to tests
performed in the same location after some period of service. Such evaluation will indicate if the coating has met its
warranted life and help to determine when reapplication may be necessary.
Compatibility. Compatibility of a coating with the brickwork and its existing surface treatments should be
determined prior to application. Only coatings specifically formulated for use on brickwork should be selected.
Incompatibility of a coating with the brickwork or an existing coating may adversely affect durability, appearance
or otherwise prevent the coating from performing as intended. Penetrating coatings are typically incompatible with
existing film-forming coatings. In some cases, reapplication of a coating may cause clouding and may be difficult or
impossible to remove.
It may be necessary to remove any existing coating, following the coating manufacturers recommendations before
reapplication or application of a different coating. This procedure may involve hazardous chemicals often regulated
or restricted from use by local, state or federal environmental regulations. Thus, an existing coating may have to
remain in place until it wears off, even if deterioration of the masonry calls for its removal.
Environmental Considerations
Possible environmental hazards are also of concern when considering a colorless coating. Often the chemicals
used in colorless coatings are highly reactive and can etch glass, damage paint, kill vegetation and emit harmful
vapors. This requires attention to worker safety and proper protection of adjacent surfaces.
Appearance
Some coatings, particularly film-formers, may impart a gloss, sheen or darkening to brickwork. Acceptable
appearance is a subjective matter and should be determined by the designer or owner prior to application. Gloss
is best evaluated by treating half of a test area representing the entire range of brick colors and textures and
comparing the treated half to the untreated half. An accepted test area should be retained as a means of judging
acceptability of other treated areas. When necessary, a number of ASTM test methods can be used to evaluate
differences and to establish tolerances [Ref. 1, Volume 6.01].
Slip Resistance
A coating can adversely affect the slip resistance of a brick floor or pavement. The slip resistance of coated floors
or pavements should be evaluated for safety reasons, especially in public access areas and in areas where water
may contact the floor or pavement. The slip resistance of coatings often is measured in the laboratory using ASTM
D2047, Test Method for Static Coefficient of Friction of Polish-Coated Floor Surfaces as Measured by the James
Machine [Ref. 1]. A value of 0.5, measured by the James machine, is the recognized minimum value for slip-
resistant walking surfaces in courts of law in the United States. Slip resistance can be measured in the field using
portable devices such as the NBS-Brungraber machine (also known as the Mark I Slip Tester). The United States
Access Board recommends coefficient of friction values of 0.6 for a level surface and 0.8 for ramps, as measured
using the NBS-Brungraber machine [Ref. 14].
Graffiti Resistance
Effective graffiti resistance depends on the ability of a coating to prevent penetration of unwanted markings into
brickwork and facilitate their removal. Often, water repellency, appearance, durability and other properties are also
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 10 of 14
important selection criteria for anti-graffiti coatings. A method for determining the effectiveness of an anti-graffiti
coating is described in ASTM D7089, Practice for Determination of the Effectiveness of Anti-Graffiti Coating for
Use on Concrete, Masonry and Natural Stone Surfaces by Pressure Washing [Ref. 1]. Satisfactory performance
is indicated by successful removal of intentionally applied graffiti. Always consult the coating manufacturer prior to
testing, as reactions between the cleaner and the coating may be hazardous.
Anti-graffiti coatings generally employ either a barrier or sacrificial strategy to resist graffiti. Barrier or permanent
coatings must be resistant to cleaning chemicals so that they remain on the surface of brickwork after graffiti is
removed. Conversely, sacrificial coatings should be easy to remove. Removal of graffiti should always follow
coating manufacturers recommendations, because many anti-graffiti coatings are intended to be used with a
particular removal method or cleaning product.
As anti-graffiti coatings provide a barrier to paint and other staining, they also provide a barrier to water
evaporation through the outer face of the brick, similar to that of glazed brick. Therefore, most of the drying of the
brickwork occurs by evaporation through the back face of the brick, into the air space. It is important that when
an anti-graffiti coating is used, the cavity behind the brick be vented at top and bottom to help remove the excess
moisture in the air space created by this evaporation.
CONSIDERATIONS PRIOR TO COATING
Selection of a colorless coating for use on brick masonry should be based on the desired performance, the
information discussed in this Technical Note and literature from the coating manufacturer. Additional items
to be considered prior to application of a colorless coating follow. Whenever possible, consult with the brick
manufacturer for specific recommendations regarding coating of a particular brick. Properties of each brick are
unique and can affect coating performance.
1. It is suggested that the designer or user require test reports for relevant performance criteria and a written
warranty from the coating manufacturer for the performance of the coating over a designated period of time.
2. The coating should be that of a well-known manufacturer who has been in business for at least five years.
It is suggested that a brand name be used that has a good track record over a period of at least five years.
References of projects with similar installations, materials and exposure should be investigated.
3. The coating should be applied at the application rate and under the climatic conditions recommended for
clay brick masonry substrates by the coating manufacturer. Typically, temperatures above 40 F (4 C) and
below 100 F (38 C) are required. Application on windy days should be avoided when possible.
4. Repair and replacement of brick and mortar joints and other necessary repairs should be completed prior to
applying a colorless coating.
5. A minimum of one month should pass after close-in of the building before a water repellent is applied to
newly constructed brickwork. This period allows the evaporation of moisture from the building materials to
occur naturally, unimpeded by a coating on the brickwork, and permits the walls to cure sufficiently. In fact,
many colorless coating manufacturers recommend application only to a relatively dry substrate. A delay of
one year is preferred so that efflorescence due to water absorbed during construction, often known as new
building bloom, is not entrapped by the coating. For a more complete discussion of efflorescence, refer to
the Technical Note 23 Series.
6. There should have been no efflorescence or, at the maximum only a minor occurrence of efflorescence, on
the brick masonry to be treated. Walls with a history of efflorescence should be coated only after the source
of moisture has been addressed.
7. The wall must be clean at the time of application [Ref. 9]. Heavy accumulations of dirt will interfere with
proper penetration or adhesion of the coating and result in poor performance and shorter life. See ASTM
D5703, Practice for Preparatory Surface Cleaning for Clay Brick Masonry [Ref. 1], for a discussion of
cleaning techniques that may be required. In addition, freshly repointed mortar and repaired sealant joints
should cure for a minimum of 72 hours before a coating is applied [Ref. 11].
8. The brickwork should have a moisture content consistent with that recommended by the coating
manufacturer. Moisture content of the brick masonry should be checked at several locations by the method
recommended by the coating manufacturer.
9. Apply samples of the selected coating to test areas of at least 10 ft (1 m) on the building at a location
representative of the area to be treated or on a sample panel. Allow these test areas to cure as
recommended by the coating manufacturer. Inspect and test them to determine satisfactory performance
with respect to the performance criteria established.
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 11 of 14
10. The application contractor should know the work to be performed and should protect adjacent and
surrounding surfaces from over-spray as necessary. Qualifications of the contractor should be verified.
These steps must be taken in conjunction with the recommendations contained within the applicable sections of
this Technical Note. They cannot guarantee successful performance but will greatly increase the likelihood that the
colorless coating will perform as intended. The coating manufacturer often will have additional recommendations
regarding coating selection, substrate preparation, curing, application methods and coverage rates. The coverage
rate is especially critical, because over-application of the coating can reduce its breathability. Failure to consider
these items can result in poor performance of the coating and can cause severe harm to the masonry or
surrounding elements.
RECOMMENDATIONS FOR USE
Selection of a specific product should be based on recommended performance criteria described herein and any
other criteria set by the designer to address the particular conditions involved. In addition, the brick manufacturer
should be consulted for recommendations on the use of colorless coatings prior to coating selection.
There are a variety of reasons that colorless coatings may be considered for application to brickwork. However, it
is important to recognize that coatings change the physical properties of the brickwork to which they are applied.
Therefore, the potential advantages of colorless coatings should be carefully weighed against their disadvantages.
Exterior Walls
Penetrating coatings are preferred for exterior brick masonry walls because they permit water vapor transmission.
Only coatings with a water vapor permeability of 0.98 or greater as measured by ASTM E96 should be used.
If a water repellent is to be used, siloxanes are recommended. Siloxanes provide the advantage of good water
repellency and long-term performance and have been shown to be effective on many brick masonry walls. Silanes
containing chemical catalysts also have been used successfully.
Because of the effect of a film on the breathability of masonry, use of film-forming coatings is discouraged,
particularly in freezing climates. Some film-forming coatings have been known to perform successfully; however,
there can be significant risks. If use of a film-forming coating, such as an anti-graffiti coating, is necessary, select
only products known to successfully perform in a similar climate, wall type and exposure on brick masonry with
similar physical properties.
When a drainage wall is treated with a colorless coating, the use of vents at the top and bottom of wall cavities can
promote evaporation of moisture from the brickwork.
Chimneys and Parapets. These
elements can be subject to
premature deterioration because
of severe exposure. They are often
exposed to wind-driven rain and
water rundown on the exterior walls
from the crown or coping. Because
of the large amount of moisture
that can contact the surface of
a chimney or parapet wall, a
clear water-repellent coating can
sometimes be effective in reducing
water-related problems. Conditions
in which a clear water repellent
may be recommended on chimneys
and parapet walls include climates
with a driving rain index above
3 (see Figure 1) and on sloped
or horizontal projections of such
elements where water and snow
can accumulate.
Driving Rain Index
1 2 3 4 0 5 >5
Figure 1
Driving Rain Index Map [Ref. 9]
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 12 of 14
Interior Walls
Colorless coatings are generally applied to interior walls to facilitate cleaning or to provide a gloss. Water
repellency and breathability of interior walls is generally not a concern. Film-forming products, specifically water-
borne acrylics (acrylic emulsions) and urethanes, typically will give the best results when gloss and ease of
cleaning are desired. However, some penetrating coatings may also provide this effect. Acrylics in particular are
known to provide a high gloss. Both acrylics and urethanes are durable in installations with no UV exposure.
In the case of exterior brick masonry walls that have their interior faces exposed, water vapor transmission may
be a concern. Film-forming products should be used cautiously, only after the effect of the film on the water vapor
transmission of the wall system has been evaluated.
Pavements and Floors
Coatings may be desired on brick pavements to resist staining or to decrease moss and mildew growth.
The exposure and construction of brick pavements are significantly different from those of vertical brickwork.
Lack of a drainage cavity or air space to aid in drying increases the severity of exposure. There are several
disadvantages associated with the use of a colorless coating on pavement surfaces. Colorless coatings can
decrease the slip resistance of the pavement or floor, especially when wet. Also, pavements and interior floors are
subject to abrasion due to foot traffic, which shortens the life expectancy of most coatings compared to vertical
installations. Exterior brick pavements are subjected to more severe weathering exposures than exterior vertical
walls. Pavements often have prolonged contact with moisture due to their horizontal orientation and are seldom
protected by overhangs.
Any joint sand stabilizers needed to protect sand in joints from erosion are typically applied before coatings. For
more information about these products, refer to Technical Note 14A.
Exterior Pavements. By the nature of their construction, pavements allow evaporation of moisture from the
masonry through only one face, the wearing surface. As a result, the potential for problems associated with
reduced water vapor transmission are significant. These disadvantages usually outweigh any potential benefit. For
this reason, colorless coatings are not recommended for use on exterior brick pavements subject to freezing and
thawing. In exterior environments not subject to freezing, the water vapor transmission rate of the coating must be
high. Clouding of the coating is a particularly common problem (see Photo 4).
Interior Floors. Colorless coatings are often applied to interior brick floors to provide a glossy finish and to
facilitate cleaning. Mortarless brick pavements also can be coated to help retain the jointing sand in the joints.
Urethanes, acrylics, waxes and some penetrating coatings that meet the performance criteria discussed herein,
and those set by the designer, can be used on interior brick masonry floors not subject to freezing. The primary
disadvantage of most colorless coatings used on floors is their tendency to reduce the skid resistance of the
floor. New epoxy-based coatings show promise in this area. Film-forming coatings may separate from the brick
paving and turn cloudy if moisture from the brickwork or supporting members migrates through the brick floor.
Consequently, film-forming coatings should be applied only when the brick floor and supporting members are dry.
Past successful performance is the best measure of a satisfactory coating.
SUMMARY
This Technical Note has discussed both the reasons for and the suitability of colorless coatings for brick masonry.
For most exterior brick masonry, use of colorless coatings is discouraged. Furthermore, clear water repellents
are not necessary on properly designed and constructed brick masonry. However, under certain conditions, clear
water repellents and other colorless coatings may be beneficial.
The information and suggestions contained in this Technical Note are based on the available
data and the experience of the engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 13 of 14
REFERENCES
1. Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2007:
Volume 4.05
C67 Standard Test Methods for Sampling and Testing Brick and Structural Clay Tile
C1601 Standard Test Method for Field Determination of Water Penetration of Masonry Wall
Surfaces
E514 Standard Test Method for Water Penetration and Leakage Through Masonry
Volume 4.06
E96/E96M Standard Test Methods for Water Vapor Transmission of Materials
Volume 6.01
D523 Standard Test Method for Specular Gloss
D3134 Standard Practice for Establishing Color and Gloss Tolerances
D4449 Standard Test Method for Visual Evaluation of Gloss Differences Between Surfaces of
Similar Appearance
Volume 6.02
D5703 Standard Practice for Preparatory Surface Cleaning of Clay Brick Masonry
D7089 Standard Practice for Determination of the Effectiveness of Anti-Graffiti Coating for Use
on Concrete, Masonry and Natural Stone Surfaces by Pressure Washing"
Volume 15.04
D2047 Standard Test Method for Static Coefficient of Friction of Polish-Coated Floor Surfaces
as Measured by the James Machine
G154 Standard Practice for Operating Fluorescent Light Apparatus for UV Exposure of
Nonmetallic Materials
2. Binda, L., Experimental Study on the Durability of Preservation Treatments of Masonry Surfaces: Use of
Outdoor Physical Models, Proceedings of the Workshop The Degradation of Brick and Stone Masonries
Due to Moisture and Salt Content and the Durability of Surface Treatments, Politecnico di Milano, Milan,
Italy, January 1991, pp. 1-8.
3. Brown, R.H., Initial Effects of Clear Coatings on Water Permeance of Masonry, Masonry: Materials,
Properties, and Performance, ASTM STP 778, J.G. Borchelt, ed., ASTM International, West
Conshohocken PA, 1982, pp. 221-236.
4. Clark, E.J., Campbell, P.G., and Frohnsdorff, G., Waterproofing Materials for Masonry, NBS Technical
Note 883, National Bureau of Standards, Gaithersburg, MD, October 1975.
5. Clear Water Repellents for Above Grade Masonry and Horizontal Concrete, Sealant, Waterproofing &
Restoration Institute, Kansas City, MO, 1994.
6. Clear Water Repellent Treatments for Concrete Masonry, Concrete Masonry Association of California and
Nevada and the Masonry Institute of America, Los Angeles, CA, 1993, pp. 38-40.
7. Coney, W.B., and Stockbridge, J.G., The Effectiveness of Waterproofing Coatings, Surface Grouting, and
Tuckpointing on a Specific Project, Masonry: Materials, Design, Construction, and Maintenance, ASTM
STP 992, H.A. Harris, ed., ASTM International, West Conshohocken, PA, 1988, pp. 220-224.
8. Grimm, C.T., A Driving Rain Index for Masonry Walls, Masonry Materials, Properties, and Performance,
ASTM STP 778, American Society for Testing and Materials, West Conshohocken, PA, 1982, pp. 171-
177.
9. Mack, R.C., and Grimmer, A., Assessing Cleaning and Water-Repellent Treatments for Historic Masonry
Buildings, Preservation Briefs, No. 1, U.S. National Park Service, Washington, DC, November 2000.
10. McGettigan, E., Selecting Clear Water Repellents, The Construction Specifier, Vol. 47, No. 6,
Construction Specifications Institute, Alexandria, VA, June 1994, pp. 121-132.
www.gobrick.com | Brick Industry Association | TN 6A | Colorless Coatings for Brick Masonry | Page 14 of 14
11. Roller, Sandra, A Comparison of ASTM E 514 and MAT Tube Water Penetration Testing Methods
Including an Evaluation of Saver Systems Water Repellents, Department of Civil and Architectural
Engineering, University of Wyoming, Laramie, WY, October 1994.
12. Roth, M., Comparison of Silicone Resins, Siliconates, Silanes and Siloxanes as Water Repellent
Treatments for Masonry, Technical Bulletin 983-1, ProSoCo, Inc., Kansas City, KS, 1985.
13. Suprenant, B.A., Water Repellents: Selection and Usage, Magazine of Masonry Construction, Aberdeen
Group, Addison, IL, December 1993, pp. 527-532.
14. Technical Bulletin: Ground and Floor Surfaces, United States Access Board, Washington, DC, August
2003.
Water Penetration Resistance -
Design and Detailing
Abstract: Proper design, detailing and construction of brick masonry walls are necessary to minimize water penetration into
or through a wall system. Many aspects of design, construction and maintenance can influence a wall's resistance to water
penetration. The selection of the proper type of wall is of utmost importance in the design process as is the need for com-
plete and accurate detailing. In addition to discussing various wall types, this Technical Note deals with proper design of brick
masonry walls and illustrates suggested details which have been found to be resistant to water penetration.
Key Words: barrier, design, detailing, drainage, flashing, installation, rain, wall types, weeps.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
7
December
2005
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Wall System Selection:
Drainage walls provide maximum protection against water
penetration
Barrier walls are designed to provide a solid barrier to
water penetration and provide good water penetration
resistance
Single wythe masonry walls require careful detailing and
construction practices to provide adequate water penetra-
tion resistance
Through Wall Flashing Locations:
Install at wall bases, window sills, heads of openings, shelf
angles, tops of walls and roofs, parapets, above projec-
tions, such as bay windows, and at other discontinuities in
the cavity
Through-Wall Flashing Installation:
Lap continuous flashing pieces at least 6 in. (152 mm)
and seal laps
Turn up the ends of discontinuous flashing to form end
dams
Extend flashing beyond the exterior wall face
Terminate UV sensitive flashings with a drip edge
Weeps:
Open head joint weeps spaced at no more than 24 in.
(610 mm) o.c. recommended
Most building codes permit weeps no less than
3
/16 in. (4.8
mm) diameter and spaced no more than 33 in. (838 mm)
o.c.
Wick and tube weep spacing recommended at no more
than 16 in. (406 mm) o.c.
Page 1 of 9
INTRODUCTION
This Technical Note is the first in a series addressing water resistance of brick masonry. Design considerations
and details are provided to illustrate the principles involved in addressing water penetration issues. The other
Technical Notes in this series provide detailed guidance in the areas of material selection (7A) and construction
(7B). Technical Notes 7C and 7D provide information on condensation.
When masonry walls encounter problems, water-related issues are often one of the primary factors. If a wall is
saturated with water, freezing and thawing may cause cracking, crazing, spalling and disintegration over time.
Water can cause masonry to experience dimensional changes, metals to corrode, insulation to lose its effective-
ness, interior finishes to deteriorate and efflorescence to appear on exterior surfaces. Water penetration may also
provide the moisture necessary for the development of mold growth on susceptible wall elements.
Water resistance of a masonry wall depends on four key factors: design, including detailing; materials; construc-
tion; and maintenance. Attention to all four is necessary to produce a satisfactorily performing wall. Failure to prop-
erly address any one factor can result in water penetration problems.
Water is abundant in many forms. Rain and snow contact building materials, wetting them. Water vapor is present
in the air from many sources. As a result, since water cannot be completely eliminated, water penetration must
be controlled. When water passes through brick masonry walls, it typically does so through minute separations
between the brick units and the mortar joints. Under normal exposures, it is virtually impossible for significant
amounts of water to pass directly through the brick units or through the mortar. Highly absorbent brick will absorb
some water, but certainly do not contribute to an outright flow of water through a wall.
Before brick veneer became popular, masonry walls usually functioned as both the structural system and as the
exterior skin of the building. As a result, these masonry walls were quite massive, ranging in thickness from 12 in.
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 2 of 9
(305 mm) up to 6 ft (1.83 m) of solid brick. These masonry walls, both because of their thickness and their being
in constant compression due to the structural loads, worked quite well in keeping water out of the interior of the
building. Many older masonry walls were built with cornices and other ornamentation which helped to protect the
faces of the buildings from excessive water rundown and subsequent water penetration to the interior.
Walls used today are much less massive, and the masonry may be only 3 in. (76 mm) or less in thickness. In
many cases, they have minimal overhang at the top, allowing sheeting of the rain water from the roof or para-
pet down to the ground. As a result of these newer wall systems, rain water is allowed to be in contact with the
masonry in larger quantities and for longer periods of time, thus leading to more opportunity for water penetration
problems.
The successful performance of a masonry wall depends on limiting the amount of water penetration and control-
ling any water that enters the wall system. If water penetration can be minimized, for all practical purposes, the
wall system will perform well.
DESIGN
The first factor in evaluating water penetration resistance of masonry is that of design. Proper design of masonry
does not mean just proper structural design. Design includes fire resistance, heat transmission, structural integrity,
material compatibility, sound reduction, aesthetics and water resistance. Other Technical Notes provide guidance
on each of these different design factors.
Design for water resistance requires evaluation of several items, including: (1) sources of moisture; (2) selection
of wall type; and (3) flashing and weeps. Each of these items will be addressed separately.
Sources of Moisture
Moisture is present almost everywhere in various
forms, i.e., rain, snow, condensation, ground water,
construction water, etc. Some of these lend them-
selves to control; some do not. This section deals
with wind-driven rain. Interstitial condensation and
its control are discussed in Technical Notes 7C and
7D.
Wind-Driven Rain. The exposure to which a
masonry wall will be subjected is very important to
the proper design of the wall. No single standard
design can be expected to perform equally well
under all exposures.
Exposures vary greatly throughout the United
States, from severe on the Atlantic Seaboard and
Gulf Coast, where rains of several hours' duration
may be accompanied by high velocity winds; to
moderate in the Midwest and Mississippi Valley,
where wind velocities are usually lower; to slight in
the arid areas of the West. Refer to Figure 1.
Selection of Wall Type
The selection of the proper wall type to use in any
given situation is very important. Under normal
conditions, it is nearly impossible to keep a heavy
wind-driven rain from penetrating a single wythe of
brickwork, regardless of the quality of the materials
or the degree of workmanship used.
The best approach to designing a water resistant
wall is to design the wall assuming some water
penetrates the surface. Therefore, the objective is
Figure 1
Wind Speed and Precipitation
Source: National Climatic Data Center
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 3 of 9
to control the moisture once it begins to penetrate the wall. Two basic wall systems are used for this purpose: the
drainage wall and the barrier wall.
Drainage Wall Systems. Drainage wall systems include cavity walls (metal-tied and masonry-bonded hollow walls
in historical applications), and anchored veneer walls as shown in Figures 2 through 5. The basic concept behind
the drainage wall assumes a heavy, wind-driven rain will penetrate the exterior wythe of brickwork. When it does,
the wall is designed to allow the water to flow inward to
the air space or cavity between the wythes. The water
then flows down the back face of the outer brick wythe,
where it is collected on the flashing and redirected out of
the wall system through the weeps. Properly designed,
detailed and constructed drainage wall systems are excel-
lent with respect to water penetration resistance. Specific
detailed information on all aspects of cavity wall systems
can be found in the Technical Notes 21 Series. The
Technical Notes 28 Series addresses anchored veneer
wall systems.
Barrier Wall Systems. Barrier wall systems, such as
the one shown in Figure 6, include multi-wythe walls with
mortar- or grout-filled collar joints (including composite
brick and concrete block walls), reinforced brick masonry
walls and adhered veneer walls. The basic concept is that
when a wind-driven rain penetrates the exterior wythe of
Figure 4
Insulated Brick/CMU Wall
Figure 5
Masonry Bonded Hollow Wall
Figure 2
Brick Veneer/Wood Stud Wall
Figure 3
Brick Veneer/Steel Stud Wall
Figure 6
Reinforced Barrier Wall
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 4 of 9
masonry it migrates inward toward a filled collar joint that acts as a barrier to prevent further inward movement.
The water then migrates back out of the wall system. The key item is that the collar joint must be completely filled
with grout or mortar to provide a monolithic barrier to moisture. Grouting is the most effective method of ensuring
that collar joints are completely filled. However, grouting spaces less than
3
/4 in. (19.1 mm) is not recommended.
In these instances, the face of the inner masonry wythe should be parged and the back of brick in the exterior
wythe buttered in order to fill the collar joint. Placing mortar in the collar joint with a trowel after the individual
wythes are laid, commonly referred to as "slushing", does not result in completely filled joints, and is not recom-
mended. Flashing is also integrated into barrier walls to aid in controlling water that penetrates the exterior wythe.
Properly designed, detailed and constructed barrier wall systems work well with respect to water penetration resis-
tance.
Single-Wythe Walls. Single-wythe masonry walls can be considered a variation of a barrier wall system. Single-
wythe brick masonry construction can be designed with either solid or hollow units. In single-wythe walls, the
masonry wythe usually exceeds the thickness of a nominal 4 in. (102 mm) exterior brick wythe. In addition to the
added thickness, grouted cells help to prevent water from penetrating to the interior of the wall system. The single-
wythe wall design is not inherently as resistant to water penetration as are drainage wall systems or multi-wythe
barrier wall systems and may not be appropriate for some severe exposures. With careful detailing and good
construction practices however, they can perform well. For example vertically reinforced and grouted brickwork
often provides good water penetration resistance. With single-wythe masonry, it is especially important to use a
mortar joint profile that sheds, rather than collects water. Concave and "V" joints are preferred over raked joints,
for example. See Technical Note 7B for further information. Penetrating water repellents can increase the water
resistance of single-wythe walls. See Technical Note 6A for further information.
DETAILING
Through-Wall Flashing
Through-wall flashing is a membrane, installed in a masonry wall system, that collects water that has penetrated
the exterior wythe and facilitates its drainage back to the exterior. Such flashing is essential in a drainage wall sys-
tem, and is required as a second line of defense in a barrier wall system. Proper design requires flashing at wall
bases, window sills, heads of openings, shelf angles, projections, recesses, bay windows, chimneys, tops of walls
and at roofs. Flashing should extend vertically up the backing a minimum of 8 in. (203 mm). The water-resistant
barrier on the backing should lap the top of the flashing. Examples of water-resistant membranes include No. 15
asphalt felt, building paper, certain high-density polyethylene or polypropylene plastics (housewraps) and certain
water-resistant sheathings. Various types of flashing materials which may be used in the design of brick masonry
and composite walls are covered in Technical Note 7A.
In regard to flashing, the designer must also address the following considerations:
Extension Through Wall. When possible, flashing should extend beyond the face of the wall to form a drip as
shown in Figure 7. When using a flashing that deteriorates with UV exposure, a metal or stainless steel drip edge
can accomplish this. It is imperative that flashing be extended at least to the face of the brickwork.
Continuity. Flashing is not usually installed in one
long, continuous sheet. As a result, pieces must be
fitted together on the job. Flashing pieces should be
lapped at least 6 in. (152 mm) and the laps sealed
with mastic or an adhesive compatible with the flash-
ing material. Self-adhesive flashing should be consid-
ered as an alternate.
Flashing Around Corners. To achieve flashing con-
tinuity around corners, preformed corner pieces are
available or the pieces of flashing may be cut, lapped
and sealed to conform to the shape of the structure.
End Dams. Where the flashing is not continuous, such
as over and under openings in the wall and on each
side of vertical expansion joints, the ends of the flash-
ing should be extended beyond the jamb lines on both
Figure 7
Shelf Angle Flashing
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 5 of 9
sides and turned up into the head joint at least 1 in. (25.4
mm) at each end to form a dam. Preformed end dams may
also be used. Refer to Figure 8.
Flashing at Vertical Supports. In some cases, connec-
tions that support shelf angles make it necessary to cut,
puncture or otherwise interrupt the flashing. When this
occurs, it is important to make sure that all openings in the
flashing are tightly sealed, and that the flashing is attached
to these supports with mastic.
Weeps
In order to properly drain any water collected on the flash-
ing, weeps are required immediately above the flashing at
all locations. An open head joint, formed by leaving mortar
out of a joint, is the recommended type of weep. Open
head joint weeps should be at least 2 in. (51 mm) high.
Weep openings are permitted by most building codes to
have a minimum diameter of
3
/16 in. (4.8 mm). The practice
of placing weeps in one or more courses of brick above the
flashing can cause a backup of water and is not recom-
mended. Non-corrosive metal, mesh or plastic screens can
be installed in open head joint weeps if desired. Refer to
Figure 9.
Spacing of open head joint weeps is recommended at no
more than 24 in. (610 mm) on center. Spacing of wick and
tube weeps is recommended at no more than 16 in. (406
mm) on center. Weep spacing is permitted by most building
codes at up to 33 in. (838 mm) on center. Wicks should be
at least 16 in. (406 mm) long and extend through the brick,
into the air space and along the back of the brick.
Drainage
The air space must be kept clear of mortar and mortar drop-
pings to allow proper drainage. Drainage materials may be
specified that prevent mortar from entering the air space or
catch mortar droppings at the wall base. These materials
are usually made of a plastic mesh or fabric porous enough
to allow passage of water, but catch or deter mortar from
collecting at the base of the air space. The effects of mortar
collection devices should be considered carefully as they
may require modifications to typical details such as extend-
ing flashing more than 8 in. (203 mm) vertically up the back-
ing. While it is not mandatory to include drainage materials,
they may help to keep the air space open for drainage.
However, the use of drainage materials should not preclude
good workmanship and an effort to keep the air space
clean.
Critical Locations
Wall Base. Moisture that enters a wall gradually travels downward. Continuous flashing must be placed above
grade at the base of walls to divert this water to the exterior. In addition, base flashing prevents water from ris-
ing up into the wall system due to capillary action and helps prevent efflorescence. The elevation of flashing and
weeps should be above planting beds, ground covering, sidewalks, etc. that are placed immediately adjacent to
the wall. Once the designer has determined the level for placing flashing in the wall in accordance with the grad-
Figure 9
Flashing and Weeps
Figure 8
End Dam Detail
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 6 of 9
ing plans, care should be taken that field modifications do not result in any section of flashing being below grade.
Refer to Figure 10.
The top of the foundation stem wall should be above the elevation of the base flashing to prevent water from
being directed toward the building interior. The cavity below base wall flashing should be solidly filled with mortar
or grout.
Window Sills. Window sills should be sloped to drain; 15 degrees is recommended. Through-wall flashing must
be placed under all sills as shown in Figures 11 through 13 and turned up at the ends to form dams. Soffits and
deep reveals may require special flashing considerations. The Technical Notes 36 Series contains further details
and information.
Steel Lintels. Through-wall flashing should be installed over all openings including door and window heads as
shown in Figure 14. An exception may be those completely protected by overhangs. The flashing should be
placed on a thin bed of mortar directly on top of the lintels and turned up at the ends to form dams. Figure 15
shows several examples of lintels. Weeps are recommended above all lintels which require flashing.
Figure 10
Wall Base Flashing
(a)
(b)
(c) (d)
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 7 of 9
Shelf Angles. In concrete or steel frame buildings with the brick wythe supported on shelf angles, the entire face
of spandrel beam may be flashed or the flashing may be inserted in a continuous reglet installed in the spandrel
beam or integrated with moisture-proofing on the spandrel beam. Refer to Figure 8.
Projections, Recesses and Caps. Projections, recesses and caps tend to collect rain water and snow. They
should be sloped away from the wall to drain and be flashed where possible as shown in Figure 16. Other details
and information can be found in the Technical Notes 36 Series.
Tops of Walls and Parapets. The tops of all walls and parapets should have an adequate cap or coping, and
there should be flashing beneath the coping. Drainage-type parapet walls as shown in Figures 17 and 18 are rec-
Figure 12
Window Sill in Cavity Wall
Figure 11
Window Sill in Brick Veneer/Frame Wall
Figure 14
Window Head in Brick Veneer/Frame Wall
Figure 13
Precast Concrete or Stone Sill
Figure 15
Structural Steel Lintels
Double Angle
Solid Wall
Double Angle
Hollow Wall
Steel Shape
Suspended Plate
Steel Shape
Attached Plate
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 8 of 9
Counter Flashing
Precast or
Stone Coping
Anchorage Varies
Sealant (Typ.)
Flashing
Air Space, Min.
2 in. (51 mm)
Recommended
Overhang, Min.
1 in. (38 mm)
Recommended
1
2
/
J oint Reinforcement
with Eye & Pintle
Through-Wall
Flashing
Figure 16
Projections and Caps
Figure 17
Precast Concrete or Stone Coping on Cavity Wall
Figure 18
Metal Coping on Cavity Wall Parapet
Figure 19
Non-Parapet Wall
www.gobrick.com | Brick Industry Association | TN 7 | Water Penetration Resistance - Design and Detailing | Page 9 of 9
ommended as the best parapet system for resistance to water penetration. The Technical Notes 36 Series provide
more details and information on these subjects.
Metal copings, as shown in Figure 19, are preferable to brick, cast stone, concrete or stone copings. Metal cop-
ings should extend down the face of the wall at least 8 in. (203 mm) with the bottom edge sealed against the
masonry to prevent wind-blown rain from entering the wall. Copings of cast stone, concrete or stone must have
joints between each element closed with sealants.
Roof Flashing. Because roof flashing is placed at very vulnerable points, it must be designed and installed with
great care. Roof flashing design may depend upon the type of roofing used. Where the roof flashing is metal, the
counter-flashing should also be metal, extending into the wall and overlapping the roof flashing a minimum of 3 in.
(76 mm). Refer to Figures 17 and 18.
SUMMARY
Masonry walls constructed of brickwork have performed well for centuries and are a testament to the performance
and durability of brick. Design and detailing that maximizes the water penetration resistance of brickwork is need-
ed to achieve this level of service.
Selection of the wall type should be based on the project's location, environmental conditions and building use.
Water penetration resistance of brickwork is enhanced by including appropriate details that reduce water penetra-
tion at key points in the brickwork.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. ASTM E 2266, " Standard Guide for Design and Construction of Low-Rise Frame Building Wall
Systems to Resist Water Intrusion", Annual Book of Standards, Vol. 04.12, ASTM International, West
Conshohocken, PA, 2005.
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Water Penetration Resistance - Materials
Abstract: This Technical Note discusses considerations for the selection of materials used in brickwork and their impact on
its resistance to water penetration. Minimum recommended property requirements and performance characteristics of typical
materials are described.
Key Words: anchors, brick, coatings, corrosion resistance, flashing, grout, lintels, mortar, sealants, shelf angles, ties, water-
resistant barrier, weeps
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction 7A
December
2005
Brick and Mortar:
Select brick from the appropriate ASTM standard, desig-
nated for exterior exposures
Choose mortar materials and types that are compatible
with the brick selected
Use mortar type with lowest compressive strength meeting
project requirements
Ties and Anchors:
Use galvanizing, stainless steel or epoxy coatings to pro-
vide corrosion resistance
Water-Resistant Barriers:
Install when brick veneer is anchored to wood or steel
studs
Protect from or avoid prolonged ultraviolet (UV) exposure
Use No. 15 asphalt felt conforming to ASTM D 226 or
building paper, polymeric films (building wraps) or water-
resistant sheathings deemed equivalent or conforming to
AC 38
Tape or seal all joints of insulation or sheathings with fac-
ings intended to act as a water-resistant barrier
Flashing:
Select flashing that is waterproof, durable, UV resistant
and compatible with adjacent materials
Flashing materials should conform to applicable ASTM
specifications
Do not use aluminum, sheet lead, polyethylene sheeting
or asphalt-saturated felt, building paper or house wraps
Use a metal drip edge to extend flashings that degrade
when exposed to UV light
Weeps:
Open head joint weeps recommended
Sealant Joints:
Use backer rods in joints wide enough to accommodate
them.
Use sealants meeting the requirements of ASTM C 920 for
joints subject to large movements
Page 1 of 10
INTRODUCTION
This Technical Note is the second in a series addressing water resistance of brick masonry and provides guidance
regarding material selection of brick masonry components. Other Technical Notes in the series address brickwork
design and details (7), construction techniques and workmanship (7B) and condensation (7C and 7D).
The use of quality construction materials in brickwork is of prime importance in attaining a satisfactory degree of
water resistance. Requiring that materials meet the minimum criteria of appropriate material specifications helps to
ensure that they are of an acceptable quality.
The most recognized and widely used building material specifications for the determination of quality construction
materials are those developed by ASTM International (ASTM). The requirements of ASTM specifications alone
cannot predict performance levels of products because they are also affected by design, detailing and workman-
ship. However, the requirements are based on laboratory tests and field experience and, in the case of brick, are
the result of experience gained over a time span exceeding 100 years.
BRICK UNITS
Selection of quality brick is very important. Units are normally chosen based on color, texture, size and cost.
However, characteristics that can affect water penetration resistance should also be considered. These include
durability and those properties that influence brick/mortar compatibility.
Under normal exposures, it is virtually impossible for significant amounts of water to pass directly through brick
units. Brick may absorb some water, but this does not contribute to an outright flow of water through the brickwork.
Durability
Because exterior masonry will be exposed to moisture and the elements, durability is a primary concern. Durability
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 2 of 10
of the brickwork is affected not only by the durability of individual materials, but also the compatibility of materials,
how the assembly is designed, how materials are installed and the conditions to which the masonry is exposed.
The ASTM specifications for brick are written to provide guidance in choosing a suitable quality of brick based on
specific exposure conditions. The requirements for compressive strength, absorption and saturation coefficient are
established to indicate the resistance of the brick to damage by freezing and thawing when saturated. Cracking,
crazing, spalling and disintegration can occur if an improper choice of brick is made.
The ASTM requirements are not intended to serve as an indicator of the degree of water resistance of the mason-
ry. The degree of water resistance is related to the durability of the masonry insofar as the more water that enters
the system, the greater the probability that the masonry will be in a saturated condition during freeze/thaw cycles.
Brick Standards. Each kind of brick currently in use has its own designated ASTM standard, with specific require-
ments for durability stipulated by physical properties of the brick. The most commonly used brick standards and
the classification for the most severe exposures are:
ASTM C 216, Grade SW - Facing Brick (Solid Masonry Units Made From Clay or Shale)
ASTM C 652, Grade SW - Hollow Brick (Hollow Masonry Units Made From Clay or Shale)
ASTM C 62, Grade SW - Building Brick (Solid Masonry Units Made From Clay or Shale)
ASTM C 1405, Class Exterior - Glazed Brick (Single Fired, Brick Units)
ASTM C 126, (does not include physical requirements for the brick body, use Grade SW within ASTM C 216 or C
652) - Ceramic Glazed Structural Clay Facing Tile, Facing Brick, and Solid Masonry Units
MORTAR AND GROUT
Choosing the proper type of mortar or grout to use in a particular application is very important. To minimize water
penetration the primary concern is to choose a mortar and/or grout that will result in the most complete bond with
the masonry units chosen. The Technical Notes 8 Series provides detailed information on mortar. Technical Note
3A provides further information on grout.
Mortar
The most commonly used standard for specifying mortars for unit masonry is ASTM C 270. Four types of mortar
(M, S, N and O) are covered in the standard, although building codes typically require the use of Types M, S or
N. ASTM C 270 addresses mortars made with portland cement-lime combinations and those made with mortar
cements and masonry cements. Detailed information on ASTM C 270, mortar types and properties can be found
in Technical Note 8.
No single type of mortar is best for all purposes. The basic rule for the selection of a mortar for a particular project
is: Always select the mortar type with the lowest compressive strength that meets the performance requirements
of the project.
This general rule must be tempered with good judgment. For example, it would be uneconomical and unwise to
continuously change mortar types for various parts of a structure. However, the general intent of the rule should
be followed, using good judgment and economic sense. For most brick veneer applications, Type N mortar is
appropriate.
Grout
In some barrier masonry walls, grout is used to form a collar joint that bonds the outer and inner masonry wythes
together. Collar joints are the primary means of providing water penetration resistance in contemporary barrier
wall construction. When properly constructed, collar joints provide a solid cementitious layer deterring water entry
into the inner masonry wythe.
Grout for brickwork should conform to ASTM C 476. Two types of grout, fine and coarse, are addressed in this
standard. Coarse grout differs from fine grout in that, in addition to sand, it contains coarse aggregates such as
pea gravel. Grout may be specified by proportions or by strength requirements. Specification by proportions is
recommended for grout used in brickwork. Volumes of materials used in grout specified by proportions should be
consistently measured throughout the project.
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 3 of 10
Specification for Masonry Structures [Ref. 9] contains requirements for the maximum height of grout pour, the min-
imum width of grout space and the minimum dimensions of cells receiving grout for each grout type. Fine grout
requires a minimum grout space width of in. (19.1 mm) and any cells receiving grout to be a minimum dimen-
sion of 1 x 2 in. (38 x 51 mm). Coarse grout requires a minimum grout space width of 1 in. (38 mm) and any
cells receiving grout to be a minimum dimension of 1 x 3 in. (38 x 76 mm).
BRICK/MORTAR COMPATIBILITY
When water passes through brick masonry walls, it does so through separations that form between the brick
and the mortar at the time of laying or through cracks that form after the mortar has cured. The dominant prop-
erty affecting the amount of water entering brickwork from a materials selection standpoint is the extent of bond
between the brick and the mortar. Extent of bond is a measure of the area of contact at the interface between
brick and mortar surfaces.
Not to be confused with extent of bond, bond strength is a measure of the adhesion between brick and mortar.
Bond strength is one factor that determines if cracks form after the mortar cures. Brick and mortar combinations
that have high bond strengths may not have an extent of bond that would provide high resistance to water pen-
etration. Consequently, extent of bond is more important to water penetration resistance of brick masonry than
bond strength.
Extent of bond is influenced by both brick and mortar properties and is best achieved when both are considered.
Initial rate of absorption is the key property of the brick related to brick/mortar compatibility. Mortar properties
include water retention, air content and workability.
The initial rate of absorption (IRA) of a brick is a measure of the amount of water taken into a 30 in.
2
(194 cm
2
)
brick surface area within one minute. A bricks IRA can be measured in the laboratory under controlled drying
conditions or in the field. The field IRA of a brick will vary depending on its moisture condition at the time it is mea-
sured.
Tests over the years have shown that the most complete bond is achieved when the initial rate of absorption (IRA)
of a brick, at the time of laying, is below 30 g/min30 in.
2
(30 g/min194 cm
2
). As a result, Specification for Masonry
Structures requires brick with initial rates of absorption in excess of this value to be wetted prior to laying. Water
penetration tests of masonry built with low and high IRA brick [Ref. 4 and 5] indicate that water penetration gen-
erally increases as brick IRA increases and as mortar water retention decreases. Thus, low IRA brick should be
combined with mortars that exhibit low water retention and high IRA brick should be combined with mortars with
high water retention, See Technical Note 8B for mortar recommendations with brick of various IRAs.
Mortar air content will also affect extent of bond. Higher air content mortars such as masonry cement mortars and
those made with air-entrained cements or lime are more likely to increase water penetration.
Several studies have shown that workmanship is critical with respect to water penetration. Thus, mortars with
better workability should be used. There are no recognized tests to determine mortar workability, but it typically
increases with air content and lower compressive strength mortars.
TIES AND ANCHORS
Ties and anchors in a masonry wall system connect two or more wythes together or attach the brick veneer to
a structural backing. Ties and anchors do not directly influence water penetration, except when related to crack-
ing of the brickwork and resulting water entry. All ties and anchors must be corrosion-resistant. Applicable ASTM
standards for corrosion-resistance of masonry ties and anchors are discussed later in this Technical Note. More
detailed information on ties and anchors can be found in Technical Note 44B.
Truss-type joint reinforcement that engages the brick wythe with fixed diagonal cross wires is only permitted in
multiwythe walls with a filled collar joint. In other walls, it can restrict differential in-plane movement between
masonry wythes, which can lead to cracking and subsequent water penetration.
Additional Considerations
Drips. A drip is a bend or crimp in a tie or anchor that helps any moisture traveling across the tie to drip off before
reaching the interior masonry wythe or backing. Ties and anchors with drips are not permitted [Ref. 6] because
the drips reduce the compressive and tensile capacity of the ties when transferring the lateral loads between the
wythes.
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 4 of 10
Corrosion Resistance. Corrosion resistance is usually provided by zinc coatings or by using stainless steel. The
level of corrosion protection required for wall ties and anchors varies with their intended exposure conditions, as
follows [Ref. 6]:
- when exposed to earth or weather or to a mean relative humidity exceeding 75%, ties and anchors are
required to be stainless steel, hot-dip galvanized or epoxy-coated
- in other exposures, ties and anchors must be mill galvanized, hot-dip galvanized or stainless steel.
In addition, the designer should consider the potential for corrosion due to contact between dissimilar metals.
Items protected by zinc coatings may be hot-dip or mill galvanized. With mill galvanizing, the steel is galvanized
before the joint reinforcement or wall tie is fabricated. Therefore, ends cut during or after the manufacturing pro-
cess are not coated. With hot-dip galvanizing, the finished item is galvanized, providing more complete coverage.
Stainless steel items should be AISI Type 304 or Type 316 and conform to the appropriate specification listed
below. Building Code Requirements for Masonry Structures, also known as the MSJC Code [Ref. 6] also allows
epoxy coatings to be used as corrosion protection.
To ensure adequate resistance to corrosion, coatings or materials should conform to the following:
Zinc Coatings - ASTM A 123 or A 153 Class B (for sheet metal ties and sheet metal anchors) or 1.50 oz/ft
2

(458 g/m
2
) (for joint reinforcement, wire ties and wire anchors)
ASTM A 641, 0.1 oz/ft
2
(0.031 kg/m
2
) (minimum for joint reinforcement)
ASTM A 653, Coating designation G60 (for sheet metal ties and sheet metal anchors)
Stainless Steel - ASTM A 240 (for sheet metal anchors and sheet metal ties)
ASTM A 480 (for sheet metal anchors and sheet metal ties and for plate and bent-bar anchors)
ASTM A 580 (for joint reinforcement, wire anchors and wire ties)
ASTM A 666 (for plate and bent-bar anchors)
Epoxy Coatings - ASTM A 884 Class A, Type 1- less than or equal to 7 mils (175 m) (for joint reinforcement)
ASTM A 899, Class C 20 mils (508 m) (for wire ties and wire anchors)
MASONRY HEADERS
A header is a masonry unit laid perpendicular to the wythe that may be used to connect two wythes of masonry.
Although the MSJC Code allows wythes of masonry designed for composite action to be bonded by masonry
headers, they are not commonly used in contemporary construction. These units provide a direct path for water
penetration from the outside of the wall to the interior along the head and bed joints. As a result, they are not rec-
ommended.
WATER-RESISTANT BARRIERS
Water-resistant barriers are membranes placed behind claddings as a secondary measure to prevent the passage
of liquid water to underlying materials such as sheathing and other wall elements susceptible to moisture damage.
This function is distinct from those provided by vapor retarders, intended to prevent water vapor diffusion, and air
barriers, intended to prevent air flow through the wall system. However, some materials can serve all three func-
tions. A water-resistant barrier should keep out any water which finds its way across the air space via anchors,
mortar bridging or splashing.
A water-resistant barrier is required in exterior walls when brick veneer is anchored to wood or steel framing and
can be provided by No. 15 asphalt felt or other approved materials as described below. While a membrane is pre-
ferred, sheathing or rigid insulation boards with an inherent resistance to moisture penetration may serve as the
water-resistant barrier when all edges and joints are completely taped or sealed.
Sheet Membranes
Typically, mechanically attached membranes should not be left exposed to UV light for an extended period of time,
as they deteriorate and become less water-resistant.
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 5 of 10
Asphalt Saturated Felt. One layer of No. 15 asphalt felt is prescribed by most codes as the material for water-
resistant barriers. The felt should conform to Type I of ASTM D 226, Specification for Asphalt-Saturated Organic
Felt Used in Roofing and Waterproofing. The durability of asphalt-saturated felt is adequate; however it may be
torn during or after installation. Asphalt-saturated felt typically has a high water vapor permeability.
Building Paper. Asphalt saturated kraft paper (generally referred to as building paper) has a long history as an
approved and common substitution for No. 15 asphalt felt. Building paper for use as a water-resistant barrier
should conform to the requirements of Federal Specification UU-B-790a, Type I, Grade D. Characteristics of build-
ing paper are similar to those of asphalt saturated felt. Building paper typically has less asphalt and lower perme-
ance than felts and can offer better resistance to bending damage.
Polymeric Films. Some plastic films (building-wraps) have been approved for use as water-resistant barriers.
These films may have qualities similar to those of other water-resistant barriers, but ascertaining the effectiveness
of a particular plastic as a water-resistant barrier can be difficult as a standard specification is yet to be developed.
Some plastic membranes act as vapor retarders and can potentially trap water vapor inside the stud wall where
it can condense if the temperature in the wall drops below the dew point. Thus, all plastic membranes should not
be considered suitable and caution should be exercised when specifying them as water-resistant barriers. AC38,
Acceptance Criteria for Water-Resistive Barriers [Ref. 1], developed by the International Code Council Evaluation
Service, Inc., is typically used to establish the suitability of a polymeric film as a water-resistant barrier. Perforated
films are not recommended because they do not consistently resist water penetration in commonly used perfor-
mance tests. PVC is not recommended because of its tendency to become brittle with age.
Polymeric films are highly resistant to tearing and often function concurrently as air barriers; however, they do not
tend to seal themselves when penetrated by fasteners as felts sometimes do. Some manufacturers suggest fas-
teners with large heads or plastic caps be used rather than standard fasteners to enhance water penetration resis-
tance at fastener locations. Polymeric films can often be installed with fewer lap joints than felt and building paper,
as they are supplied in larger rolls up to 10 feet (3.1 m) wide.
Liquid Applied Films
Liquid applied films often have the capability of serving as vapor and air barriers and sometimes thermal insula-
tion, in addition to providing water resistance. These coatings are varied in type and may be spray, roller or trowel
applied; however they generally have the benefit of providing a seamless, monolithic membrane that adheres to
most substrates. Although these materials can be applied rapidly, they require skilled applicators to ensure quality
and performance.
These membranes have a unique set of service requirements as a result of being bonded to a substrate. The
effects of wet substrates, expansion and contraction at substrate joints, volume changes of building materials, and
stresses caused by lateral loads must be considered so that the membrane performs successfully during its life.
Quality installations are more difficult to achieve on substrates with rough surfaces and may require increased
thicknesses.
Board Products
Sheathings and other board products that are inherently water-resistant or have water-resistant facings are per-
mitted to serve as water-resistant barriers when the edges and joints of boards are completely taped or sealed.
To perform successfully, the materials providing this seal must maintain their integrity and performance when
subjected to moisture and other environmental conditions for the entire service life of the wall. Board products that
act as water-resistant barriers should be vapor permeable except when they are also intended to serve as a vapor
retarder.
SHELF ANGLES AND LINTELS
Although similar, shelf angles and lintels differ in the way each is incorporated into brickwork. A shelf angle sup-
ports brick veneer and is anchored to the structure. A lintel, on the other hand, is a structural beam placed over an
opening to carry superimposed loads. As such, it is supported by the masonry on each side of the opening and is
not attached to the structure.
Lintels may be loose steel angles, stone, precast concrete or reinforced masonry. The proper specification of
material for lintels is important for both structural and serviceability requirements.
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 6 of 10
Nongalvanized and non-stainless steel angles and lintels should be primed and painted as a minimum to inhibit
corrosion. For severe climates and exposures, such as coastal areas, consideration should be given to the use of
galvanized or stainless steel shelf angles and lintels. Even where galvanized or stainless steel shelf angles and
lintels are used, continuous flashing should be installed to protect the angle. To ensure adequate resistance to
corrosion, shelf angles should be protected by a zinc coating conforming to ASTM A 123, or be made of stainless
steel conforming to ASTM A 167, Type 304. Additional discussion and details of shelf angles and lintels may be
found in Technical Notes 21, 21A, 28B, 31 and 31B.
FLASHING
Selection of a proper flashing material is of utmost importance because the flashing is a critical element to the
drainage of water that may penetrate the wall system. Flashing materials should be waterproof, durable and resist
puncture and cracking during and after construction. Because flashing may be installed in advance of the exterior
brick wythe, it should be able to endure some exposure to ultraviolet (UV) light without significant deterioration.
The flashing should also resist damage from contact with metal, mortar or water and be compatible with adjacent
adhesives and sealants. Minimum recommended flashing thicknesses are included below for each type of flash-
ing. In general, thicker flashings are more durable, but may be more difficult to form.
Flashing materials generally fall into three categories: sheet metals, composite materials (combination flashings)
and plastic or rubber compounds. The selection is largely determined by cost and suitability. It is suggested that
only superior quality materials be selected, since replacement in the event of failure may be expensive. Materials
such as polyethylene sheeting, asphalt-impregnated building felt, building paper and house wraps should not be
used as flashing materials. These materials are easily damaged during installation and in many cases, turn brittle
and decay over time.
Sheet Metals
Stainless Steel. Stainless steel is an excellent flashing material that has excellent chemical resistance and does
not stain masonry. Stainless steel flashing should conform to ASTM A 167, Type 304. The minimum thickness
should be at least 0.01 in. (0.25 mm).
Because it is difficult to form, preformed shapes are commonly used, although these are difficult to bend on-site if
field adjustments are required. Mastic can be used to seal joints between individual flashing pieces, as stainless
steel can be difficult to solder.
Copper. Copper is another excellent flashing material that is durable, easy to form and solder, and is available
in preformed shapes. Exposed copper may stain adjacent masonry, but it is not damaged by the caustic alkalies
present in masonry mortars. It can be safely embedded in fresh mortar and will not deteriorate in continuously
saturated, hardened mortar, unless excessive chlorides are present. When using copper flashing, prohibit the use
of mortar admixtures containing even small amounts of chloride ions.
Copper flashing should conform to ASTM B 370, Standard Specification for Copper Sheet and Strip for Building
Construction, or B 882, Specification for Pre-Patinated Copper for Architectural Applications. The Copper
Development Association recommends minimum weights of 12 oz./ft
2
and 16 oz./ft
2
for High Yield and standard
cold rolled copper, respectively, used as through-wall flashing. If copper flashing is used adjacent to other metals,
proper care should be taken to account for separation of the materials. Laminated copper flashing and combina-
tions of copper sheet and other materials are discussed below in the Composites section.
Galvanized Steel and Zinc Alloys. Galvanized coatings are subject to corrosion in fresh mortar, thus the use
of galvanized steel as through wall flashing is not recommended. Although corrosion forms a very compact film
around zinc, its extent cannot be accurately predicted. Bending steel items cracks the galvanized coating, thereby
reducing its durability. Some zinc-alloy flashings are available, but, like many alloys, these may have properties
considerably different from those of the pure metal.
Aluminum. Aluminum should not be used as a flashing material in brick masonry. The caustic alkalies in fresh,
unhardened mortar will attack aluminum. Although dry, seasoned mortar will not affect aluminum, corrosion can
continue if the adjacent mortar becomes wet.
Sheet Leads. Thin lead sheet is not recommended as a flashing material in brick masonry. Lead, like aluminum,
is susceptible to corrosion in fresh mortar. Furthermore, where lead is partially embedded in mortar with moisture
present, galvanic action can occur resulting in the gradual disintegration of the lead.
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 7 of 10
Plastic and Rubber Flashing
Plastic and rubber flashings are resilient, corrosion resistant materials that are easy to form and join. However,
because the chemical compositions of these products vary widely, the durability of these materials is variable.
Thus, it is necessary to rely on performance records of the material, the reputation of the manufacturer, and where
possible, test data to ensure satisfactory performance. Some of the critical areas are: (1) resistance to degrada-
tion in UV light; (2) compatibility with alkaline masonry mortars; (3) compatibility with joint sealants and (4) resis-
tance to tear and puncture during construction. A minimum thickness of 30 mil (0.76 mm) is recommended for
plastic and rubber flashings.
Polyvinyl chloride (PVC). PVC degrades under exposure to UV light and should be cut flush with the face of the
wall or used with a metal drip edge to extend beyond the wall face.
Ethylene Propylene Diene Monomer (EPDM). EPDM is a synthetic rubber that is used as a single ply roofing
membrane as well as flashing. It has better low temperature performance the PVC, and better weathering resis-
tance than butyl rubber. It is commonly available in a thickness of 40 mils (1.0 mm) or greater, reducing concerns
of fragility during construction. Dimensional stability may be a concern.
Self-Adhesive Rubberized Asphalt. Self-adhesive rubberized asphalt flashing adheres to other building materi-
als and itself, thus speeding flashing installation and making it easier to seal flashing laps and terminations. These
flashings are also self-healing, making them less susceptible to small punctures. Substrates should be dry and
clean for proper adhesion. In addition, when self-adhesive flashings are used, care should be taken to ensure
compatibility between the flashing adhesive and sealants used in the wall. Primers may be necessary to ensure
adequate adhesion of self-adhering flashings to some substrates.
Composites
The most common type of composite or combination flashing is a thin layer of metal sandwiched between one or
two layers of another material, such as bitumen, kraft paper or various fabrics. The metal layer is usually copper,
lead or aluminum. Composite flashings utilize the better properties of each of their component materials. In the
case of lead and aluminum composite flashings, the paper and fabric laminates reduce the potential for corrosion
resulting from the metal foil contacting the mortar or adjacent dissimilar metals. These flashings also allow the use
of thinner metal sheet, making them less expensive and easier to form, but also more prone to tearing and punc-
tures. The laminate must either be durable and stable under UV exposure or these flashings should be used with
stainless steel drip edges. It is beyond the scope of this Technical Note to describe the various types of composite
flashing and their properties. The manufacturer's literature should be consulted for the various types of composite
flashing available.
DRAINAGE MATERIALS AND MORTAR DIVERTERS
When a high probability of mortar falling into the air space exists, such as for tall brick veneer without shelf angles,
drainage materials and mortar diverters may be useful to help prevent mortar from bridging the air space or block-
ing weeps. It is beyond the scope of this Technical Note to characterize the widely varying types of materials used
for these purposes. Manufacturers literature should be used to compare and determine the suitability of drainage
materials and mortar diverters. The use of drainage materials should not preclude good workmanship and an
effort to keep the air space clean of excess mortar droppings.
WEEPS
Although open head joint weeps are the recommended type of weep, some weeps are made using plastic or
metal tubes, or using rope wicks. These alternate weeps should be spaced more closely as they do not drain
water as quickly. Weep openings are permitted by most building codes to have a minimum diameter of
3
/16 in (4.8
mm). Rope wicks should be at least 16 in. (406 mm) long and made from cotton sash cord or other materials that
wick. Items used to form weeps should not easily deteriorate or stain the brickwork. Open head joint weeps may
have non-corrosive plastic, mesh or metal screens installed if desired. Vent-type weeps can serve a dual function
of allowing water to drain, but can also allow air to enter the cavity resulting in more drying action. There is no
single method that produces the best weep for all situations.
SEALANTS
Sealants are an important element in preventing water penetration around openings in masonry walls. Too fre-
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 8 of 10
quently, sealants are relied on as a means of correcting or hiding poor workmanship rather than as an integral
part of construction.
A discussion of the characteristics of joint sealants is beyond the scope of this Technical Note, but a few com-
ments are in order. Sealants should be selected for their durability, extensibility, compressibility and their compat-
ibility with other materials. Other important considerations in sealant selection may include curing time, UV resis-
tance, color stability, resistance to staining and the ability to handle a broad range of joint sizes. A sealant should
be able to maintain these qualities under the temperature extremes of the climate in which the building is located.
Trial applications of sealants under consideration are always helpful in determining suitability for a particular appli-
cation. Additional discussion of sealants may be found in Technical Notes 18 and 18A.
Oil-based caulks and acetoxic silicone sealants that attack cement in mortar should not be applied to masonry.
Solvent-based acrylic sealant or a butyl caulk should only be used where little or no movement is expected, such
as joints around windows and other openings. For joints subject to large movements, such as expansion joints, an
elastomeric joint sealant conforming to the requirements of ASTM C 920 should be used. This includes silicones,
urethanes and polysulfides. Application of a sealant primer may be required to preclude staining of some sealants
on certain brick.
Backer rods are recommended behind sealants in joints large enough to accommodate them. Backer rods should
be plastic foam or sponge rubber. Backer rods should be capable of resisting permanent deformation before and
during sealant application, non-absorbent to liquid water and gas, and should not emit gas which may cause bub-
bling of the sealant. A bond breaking tape may be used when there is not sufficient space for a backer rod. For
further information on sealants, refer to ASTM C 1193, Guide for Use of Joint Sealants.
COATINGS
The use of external coatings, such as paint or clear coatings, on brick masonry should be considered only after
a detailed evaluation of the possible consequences. Although coatings are not required on properly designed,
specified and constructed brick masonry, they may be used successfully to correct certain deficiencies or alter the
walls appearance.
Coatings intended to reduce water penetration (water repellents) are most effective when their intended use corre-
sponds with the nature of the water penetration problem. Use of coatings for reasons outside their intended appli-
cation rarely reduces water penetration and often leads to more serious problems. Considerations in the choice of
coating include: compatibility with brick masonry, water and air permeability, ability to span cracks, applicability to
exterior exposure, potential lifespan and aesthetic considerations. Technical Notes 6 and 6A should be consulted
when considering a coating for brick masonry.
SUMMARY
This, the second in a series of Technical Notes on water resistance of brick masonry, has provided information on
properly selecting quality materials for masonry work. This Technical Note cannot cover all available materials or
all conditions. Lack of specific reference to a material should not preclude its use providing that it results in water-
resistant brick masonry.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. Acceptance Criteria for Water Resistive Barriers, AC38, ICC Evaluation Service, Inc., Whittier CA, 2004.
2. Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2005:
Volume 1.03 - A 167, Standard Specification for Stainless and Heat-Resisting Chromium-Nickel Steel
Plate, Sheet, and Strip
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 9 of 10
A 240/A 240M, Standard Specification for Chromium and Chromium-Nickel Stainless Steel
Plate, Sheet, and Strip for Pressure Vessels and for General Applications
A 480/A 480M, Standard Specification for General Requirements for Flat-Rolled Stainless
and Heat-Resisting Steel Plate, Sheet, and Strip
A 580/A 580M, Standard Specification for Stainless Steel Wire
A 666, Standard Specification for Annealed or Cold-Worked Austenitic Stainless Steel
Sheet, Strip, Plate, and Flat Bar
A 884/A 884M, Standard Specification for Epoxy-Coated Steel Wire and Welded Wire
Reinforcement
A 899, Standard Specification for Steel Wire, Epoxy-Coated
Volume 1.06 - A 123/A 123M, Standard Specification for Zinc (Hot-Dipped Galvanized) Coatings on Iron
and Steel Products
A 153/A 153M, Standard Specification for Zinc Coating (Hot-Dip) on Iron and Steel
Hardware
A 641/A 641M, Standard Specification for Zinc-Coated (Galvanized) Carbon Steel Wire
A 653/A 653M, Standard Specification for Steel Sheet, Zinc-Coated (Galvanized) or Zinc-
Iron Alloy-Coated (Galvannealed) by the Hot-Dip Process
Volume 2.01 - B 370, Standard Specification for Copper Sheet and Strip for Building Construction
B 882, Specification for Pre-Patinated Copper for Architectural Applications
Volume 4.04 - D 226, Standard Specification for Asphalt-Saturated Organic Felt Used in Roofing and
Waterproofing
Volume 4.05 - C 62, Standard Specification for Building Brick (Solid Masonry Units Made From Clay or
Shale
C126, Standard Specification for Ceramic Glazed Structural Clay Facing Tile, Facing Brick,
and Solid Masonry Units
C 216, Standard Specification for Facing Brick (Solid Masonry Units Made from Clay or
Shale)
C 270, Standard Specification for Mortar for Unit Masonry
C 476, Standard Specification for Grout for Masonry
C 652, Standard Specification for Hollow Brick (Hollow Masonry Units Made From Clay or
Shale)
C 1405, Standard Specification for Glazed Brick (Single Fired, Brick Units)
Volume 4.07 - C 920, Standard Specification for Elastomeric Joint Sealants
C 1193, Standard Guide for Use of Joint Sealants
2. Beall, C., "Selecting a Joint Sealant", Masonry Construction, Hanley Wood, LLC, December 1996.
3. Bomberg, M. and Onysko D., "Characterization of Exterior Sheathing Membranes," Symposium on
Membranes in Enclosure Wall Systems, Building Environment and Thermal Envelope Council, June 10-
11, 2004.
4. Borchelt, J.G. and Tann, J.A., Bond Strength and Water Penetration of Low IRA Brick and Mortar,
Proceedings of the Seventh North American Masonry Conference, South Bend, IN, The Masonry Society,
June 1996.
www.gobrick.com | Brick Industry Association | TN 7A | Water Penetration Resistance - Materials | Page 10 of 10
5. Borchelt, J.G., Melander, J.M. and Nelson, R.L., Bond Strength and Water Penetration of High IRA Brick
and Mortar Proceedings of the Eight North American Masonry Conference, Austin, TX, The Masonry
Society, June 1999.
6. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
7. Lies, K.M., "Weather Resistant Barrier Performance and Selection", Symposium on Membranes in
Enclosure Wall Systems, Building Environment and Thermal Envelope Council, June 10-11, 2004.
8. Pickett, M., "Fluid Applied Wall Membrane Systems", Symposium on Membranes in Enclosure Wall
Systems, Building Environment and Thermal Envelope Council, June 10-11, 2004.
9. Specification for Masonry Structures (ACI 530.1-05/ASCE 6-05/TMS 602-05), The Masonry Society,
Boulder, CO, 2005.
10. Yorkdale, A.H. , Initial Rate of Absorption and Mortar Bond, Masonry: Materials, Properties and
Performance, STP 778, J.G. Borchelt, Ed., ASTM, September, 1982.
Water Penetration Resistance -
Construction and Workmanship
Abstract: This Technical Note covers essential construction practices needed to assure water-resistant brick masonry.
Procedures for preparing materials to be used in brick construction are recommended, including proper storage, handling and
preparation of brick, mortar, grout and flashing. Good workmanship practices are described, including the complete filling of all
mortar joints, tooling of mortar joints for exterior exposure and covering unfinished brick masonry walls to protect them from
moisture.
Key Words: air space, brick, construction, flashing, initial rate of absorption, joints, mortar, tooling, weeps, workmanship.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction 7B
December
2005
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
General
Store materials on the job site to avoid wetting and con-
tamination
For drainage walls, keep the air space free of excessive
mortar droppings
Do not disturb newly laid masonry
Cover tops of unfinished walls until adjacent construction
protects them from water entry
Brick
Pre-wet brick with a field measured initial rate of absorp-
tion (IRA) exceeding 30 g/min30 in.
2
(30 g/min194 cm
2
)
Mortar
When mixing mortar, use accurate batching measure-
ments and maximum amount of water that produces a
workable mortar
For brick with an IRA exceeding 30 g/min30 in.
2
(30 g/
min194 cm
2
), increase water or maximize water retention
by increasing lime proportions within limits of ASTM C 270
For brick with an IRA lower than 5 g/min30 in.
2

(5 g/min194 cm
2
), reduce water or minimize water reten-
tion by decreasing lime proportions within limits of ASTM
C 270
Joints
In exterior wythes, completely fill all mortar joints intended
to have mortar
Minimize furrowing of bed joints and prohibit slushing of
head joints
Fill collar joints completely with grout or mortar, preferably
grout; do not slush collar joints
Tool mortar joints when thumbprint hard with a concave,
V or grapevine jointer
Flashing and Weeps
Do not stop flashing behind face of brickwork
Where required, turn up flashing ends into head joint a
minimum of 1 in. (25.4 mm) to form end dams
Lap continuous flashing pieces at least 6 in. (152 mm) and
seal laps
Where installed flashing is pierced, make watertight with
sealant or mastic compatible with flashing
Install weeps immediately above flashing
Page 1 of 7
INTRODUCTION
The best design, detailing and materials will not compensate for poor construction practices and workmanship.
Proper construction practices, including preparation of materials and workmanship, are essential to achieve a
water-resistant brick masonry wall.
This Technical Note discusses construction techniques and workmanship and is the third in a series of Technical
Notes addressing water penetration resistance of brick masonry. Other Technical Notes in the series address
brickwork design and details (7), materials (7A) and condensation (7C and 7D). Maintenance of brick masonry is
addressed in Technical Note 46. All of these items are essential to obtain water-resistant brick masonry walls.
PREPARATION OF MATERIALS
Preparation of masonry materials before bricklaying begins is very important. Specific procedures must be followed
to ensure satisfactory performance and avoid future problems. Preparation includes material storage, mixing mor-
tar and grout and, in some cases, wetting the brick.
Storage of Materials
All materials at the jobsite should be stored to avoid contamination. Masonry units, mortar materials, ties and rein-
forcement should be stored off the ground, preferably in a dry location. In addition, all materials should be covered
with tarpaulins or other weather-resistant materials to protect them from the elements.
www.gobrick.com | Brick Industry Association | TN 7B | Water Penetration Resistance - Construction and Workmanship | Page 2 of 7
Wetting Brick
Brick with an initial rate of absorption (IRA) greater than 30 g/min30 in.
2
(30 g/min194 cm
2
) at the time of laying
tend to draw too much moisture from the mortar before initial set. As a result, construction practices should be
altered when using brick with high IRA to achieve strong, water-resistant masonry. The IRA of brick in the field will
typically be less than that reported in laboratory tests. Laboratory test results may be used to determine if measur-
ing IRA in the field is necessary. ASTM C 67, Test Methods for Sampling and Testing Brick and Structural Clay
Tile, includes a standard procedure for measuring IRA in the field.
A crude method of indicating whether brick need to be wetted prior to placement consists of drawing, with a wax
pencil, a circle 1 in. (25.4 mm) in diameter on the brick surface that will be in contact with the mortar. A quarter
can be used as a guide for the circle. With a medicine dropper, place 20 drops of water inside this circle and note
the time required for the water to be absorbed. If the time exceeds 1
1
/2 minutes, the brick should not need wetting;
if less than 1
1
/2 minutes, adjustments to typical construction practice are recommended.
Specification for Masonry Structures [Ref. 4] requires that brick
with an IRA exceeding 30 g/min30 in.
2
(30 g/min194 cm
2
) be
wetted prior to laying to produce an IRA less than 30 g/min30
in.
2
(30 g/min194 cm
2
) when the units are placed. However, exe-
cution of this method may be impractical on large-scale construc-
tion projects and the contractor may consider other alternatives,
as discussed in the following section, Mixing of Mortar and Grout.
If brick are to be wetted, the method of wetting is very important.
Sprinkling or dipping the brick in a bucket of water just before lay-
ing would produce the surface wet condition which may not be
sufficient, as shown in Figure 1b. The units should have a satu-
rated interior, but be surface dry at the time of laying, as shown in
Figure 1d.
Satisfactory procedures for wetting the brick consist of letting
water run on the cubes or pallets of brick, or placing them in a
large tank of water. This should be done the day before the units
are laid, or not later than several hours before the units will be
used so that the surfaces have an opportunity to dry before the
brick are laid. Wetting low-absorption brick or excessive wetting
of brick may result in saturation, as shown in Figure 1c. This can
cause bleeding of the mortar joints and floating of the brick.
Mixing of Mortar and Grout
Typically, a high water content in the mortar is necessary to obtain complete and strong bond between mortar and
brick. In general the mortar should be mixed with the maximum amount of water that produces a workable mor-
tar. Factors such as the jobsite environment and the IRA of the brick should be considered when determining the
proper amount of water to include in the mortar.
Mortar to be used with brick that have an IRA greater than 30 g/min30 in.
2
(30 g/min194 cm
2
) should be mixed to
maximize water retention by increasing mixing water or lime content within the limits of ASTM C 270. This is par-
ticularly important when pre-wetting the brick to reduce their IRA is impossible or impractical. Admixtures designed
to increase the water retention of the mortar may also be used to improve the compatibility of mortar with high IRA
brick. Only admixtures with test data showing no deleterious effects should be used.
Mortar for use with brick that have an IRA less than 5 g/min30 in
2
(5 g/min194 cm
2
) should be mixed with
reduced amounts of water or lime to minimize water retention. Lime proportions should remain within the limits of
ASTM C 270.
When brick with widely different absorption rates are used together in brickwork, it is important to maintain the cor-
rect water content in the mortar used with the different brick.
All cementitious materials and aggregates must be mixed for at least 3 minutes and not more than 5 minutes in
a mechanical batch mixer. If, after initial mixing, the mortar stiffens due to the loss of water by evaporation, addi-
c) Saturated
a) Dry b) Surface Wet
d) Surface Dry c) Saturated
a) Dry b) Surface Wet
d) Surface Dry
Figure 1
Moisture Content of Brick
www.gobrick.com | Brick Industry Association | TN 7B | Water Penetration Resistance - Construction and Workmanship | Page 3 of 7
tional water should be added and the mortar remixed (retempered). All mortar should be used within 2
1
/2 hr (2 hr
in hot weather conditions, see Technical Note 1) of initial mixing and grout should be used within 1
1
/2 hour of intro-
ducing water into the mix. No mortar or grout should be used after it has begun to set.
One of the most common problems with mortar is oversanding. Oversanded mortar is harsh, unworkable and
results in poor extent of bond and reduced bond strength, thus increasing the potential for water penetration prob-
lems. The cause of oversanding is frequently the use of the shovel method of measuring the sand. The amount of
sand that a shovel will hold varies depending on the moisture content of the sand, the person doing the shoveling
and the different size of shovels used on the jobsite. To alleviate this problem, proper batching methods must be
used. Measurement of sand by shovel should not be permitted without periodically gauging the shovel count using
a bucket or box of known volume. Technical Note 8B provides detailed guidelines for various methods of more
accurately batching mortar.
Blending of Brick
While not related to water penetration resistance, blending of brick at the jobsite is an important preparation task
related to workmanship and the acceptable appearance of brickwork. Because brick is made from natural materi-
als that vary in physical properties, variations in color may occur between production runs and occasionally within
the same run. Modern manufacturing processes use automatic equipment which may not permit inspection of
each brick, also resulting in minor color and texture variations. For these reasons, straps of brick from different
cubes should be placed together around the wall. The mason should then select brick from adjacent straps when
laying a given section of brickwork. By blending the brick throughout the wall in this manner, the effect of potential
color variations on the finished brickwork is minimized.
WORKMANSHIP
The importance of good workmanship to attain quality brickwork
cannot be overemphasized. While design and the quality of mate-
rials contribute to the water penetration resistance of brickwork,
workmanship is a highly important factor in the construction of
water-resistant masonry.
Placing Flashing and Weeps
Flashing must be installed properly and integrated with adjacent
materials to form an impervious barrier to moisture movement. The
flashing should be wide enough to start outside the exterior face
of the brick wythe, extend across the cavity, and turn up vertically
against the backing or interior wythe at least 8 in. (203 mm). The
top (vertical) edge should be placed in a mortar joint of the back-
ing wythe, in a reglet in concrete backing, or attached to sheathing
with a termination bar, as shown in Figure 2. Sections of flashing
are to be overlapped at least 6 in. (152 mm) and the lap sealed
with a compatible adhesive. Water-resistant sheet membranes
should overlap the flashing in a shingle fashion by at least 6 in.
(152 mm).
Flashing that is placed so that the outside edge projects from the
face of the wall may be cut flush with the face of the brickwork. In
no circumstances should the flashing be stopped behind the face
of the brickwork. Continuity at corners and returns is achieved by
cutting and folding straight sections or using preformed corner
pieces. Discontinuous flashing should terminate with an end dam in
a head joint, rising at least 1 in. (25.4 mm) as shown in Figure 3.
Flashing must be placed without punctures or tears. Openings cre-
ated for reinforcement or anchors must be closed with a compatible
sealant. Protection may be needed around bolts fastening shelf
angles to the structure.
Termination Bar
Water-Resistant
Barrier on
Exterior Sheathing
Weep
Flashing
Filled Cavity
Beneath Flashing
Figure 2
Wall Base Flashing Detail
Flashing
End Dam
Figure 3
End Dam Detail
www.gobrick.com | Brick Industry Association | TN 7B | Water Penetration Resistance - Construction and Workmanship | Page 4 of 7
Weeps are required, and should be formed in mortar joints imme-
diately above the flashing. Open head joints, formed by leaving
mortar out of a joint, are the recommended type of weep. Open
head joint weeps should be at least 2 in. (51 mm) high. Weep
openings are permitted by most building codes to have a mini-
mum diameter of
3
/16 in. (4.8 mm). The practice of specifying the
installation of weeps one or more courses of brick above the flash-
ing can cause a backup of water and is not recommended. Non-
corrosive metal, mesh or plastic screens can be installed in open
head joint weeps if desired.
Spacing of open head joint weeps at no more than 24 in. (610
mm) on center is recommended. Spacing of wick and tube weeps
is recommended at no more than 16 in. (406 mm) on center.
Weep spacing is permitted by most building codes up to 33 in.
(838 mm) on center. If other than an open head joint weep is
used, be sure the weep is clear of all mortar to allow the wall to
drain (see Technical Note 21C). Rope wicks should be flush with,
or extend
1
/2 in. (12.7 mm) beyond the face of the wall to promote
evaporation. The rope should continue into the bottom of the air
space, placed along the back of the brick and be at least 16 in.
(406 mm) long.
Filling Mortar Joints
To reduce water penetration, there is no substitute for proper
filling of all mortar joints that are designed to receive mortar.
Improperly filled mortar joints can result in leaky walls, reduce
the strength of masonry, and may contribute to disintegration and
cracking due to water penetration and subsequent freezing and
thawing.
A uniform bed of mortar should be spread over only a few brick,
and furrowed lightly, if at all. Filled joints result when plenty of mor-
tar is placed on the end of the brick to be laid and it is shoved into
place so that mortar is squeezed out of the top of the head joint,
as shown in Photo 1. After placement, mortar squeezed out of bed
joint should be cut off prior to tooling, as shown in Photo 2. When
placing closures, plenty of mortar is needed on the ends of brick in
place and on the ends of the brick to be laid. The closure should
be shoved into place without disturbing brick on either side, as
shown in Photo 3.
Bed Joints. A bed joint is the horizontal layer of mortar on which
a brick is laid. The length of time between placing the bed joint
mortar and laying the succeeding brick influences the resulting
bond. If too long a time elapses, poor extent of bond will result.
Brick should be laid within 1 minute or so after the mortar is placed.
For solid brick, bed joints should be constructed without deep
furrowing of the mortar, as full bed joints (covering the entire bed-
ding surface) are an inherent requirement for water-resistant brick
masonry construction. For hollow brick, bed joints may be laid with
face shell bedding (mortar placed only on the front and back face
shells). Both face shells must be completely covered with mortar.
Head Joints. A head joint, sometimes called a cross joint, is the vertical mortar joint between two brick. For both
solid and hollow brick it is important that head joints be completely filled. The best head joints are formed by com-
pletely buttering the ends of the brick with mortar and shoving the brick into place against previously laid brick.
Photo 1
Shoving Brick into Place
Photo 2
Cutting Excess Mortar
Bad Bad Good
Photo 3
Placing the Closure
Figure 4
Head Joints
www.gobrick.com | Brick Industry Association | TN 7B | Water Penetration Resistance - Construction and Workmanship | Page 5 of 7
Slushing (throwing mortar into the joint with the edge of a trowel)
does not adequately fill joints or compact the mortar, resulting in
joints that are less resistant to water penetration. The results of head
joint forming are shown in Figure 4.
Tooling of Mortar Joints
Proper tooling, or striking, of mortar joints helps seal the wall sur-
face against moisture penetration. Mortar joints should be tooled
when they are thumbprint hard, (pressing the thumb into the mortar
leaves an indentation, but no mortar is transferred to the thumb)
with a jointer slightly larger than the joint. It is important that joints
are tooled at the appropriate time as this affects both their effective-
ness and appearance. J oints that are tooled too early often smear
and result in rough joints. If tooling is delayed too long the surface of
the joint cannot be properly compressed and sealed to the adjacent
brick. Each portion of the completed brickwork should be allowed to
set for the same amount of time before tooling in order to ensure a
uniform mortar shade. Early tooling often results in joints of a lighter
color. Later tooling results in darker shades.
Concave, V and grapevine joints best resist water penetration in
exterior brickwork. These joints produce a more dense and weather-
tight surface, as the mortar is pressed against the brick, as shown in
Photos 4 and 5. For interior masonry work, other joints such as the
weathered, beaded, struck, flush, raked or extruded joints shown in
Figure 5 can also be used.
Collar Joints
The vertical, longitudinal joint between wythes of masonry is called a
collar joint. The manner in which these joints are filled is very impor-
tant. Grouting is the most effective method of ensuring that collar
joints are completely filled. However, grouting spaces less than
3
/4 in.
(19.1 mm) is not permitted. Mortar protrusions (fins) that extend more
than
1
/2 in. (12.7 mm) into a cell or cavity that will be grouted must be
removed prior to grouting. For mortar-filled collar joints, the outer face
of the inner masonry wythe should be parged and the back of brick in
the exterior wythe buttered in order to fill the collar joint.
Slushing of collar joints is not effective since it does not completely
fill all voids in the joint, as shown in Photo 6. Frequently, the mortar is
Figure 5
Typical Mortar Joints
Photo 6
Poorly Filled Collar Joint
Photo 4
Concave Mortar Joints
Photo 5
"V" Mortar Joints
www.gobrick.com | Brick Industry Association | TN 7B | Water Penetration Resistance - Construction and Workmanship | Page 6 of 7
caught and held before it reaches the bottom of the joint, leaving openings between the face brick and the back-
ing. Even when this space is filled, there is no way to compact the mortar. The mortar does not bond with the brick
over its entire surface and channels are left between the mortar and the brick. Some of these channels may allow
water to reach the back of the wall. A properly constructed collar joint is completely filled with grout or mortar.
Parging
Parging is the process of applying a coat of portland cement mortar to masonry. Parging the outer face of the
inner wythe of a multiwythe wall with Type M or S mortar as damp proofing may help resist rain penetration and
can also reduce air leakage. Membranes or liquid-applied materials usually provide superior performance to parg-
ing, which will crack if the wythe cracks. However, parging can provide a smooth base for these materials. If parg-
ing alone is to resist water penetration, proper curing is necessary to reduce shrinkage cracks. Parging the back
side of the exterior wythe is not recommended for drainage-type walls, as this may result in more debris in the air
space or break the brick/mortar bond.
The face of the wall to be parged must not have any mortar protrusions. Protruding mortar can cause bond breaks
in the parge coat, resulting in a leaky wall. When applied in multiple layers, each should be a minimum thickness
of in. (6.4 mm). The first coat should be allowed to partially set, roughened, and allowed to cure for 24 hours.
It is then moistened for application of the second coat. The parged surface should be troweled smooth so that it
sheds water easily. When completed in adjacent areas, the edges of the parging should be feathered and new
parging should overlap existing parging by a minimum of 6 in. (152 mm). Lap joints should be spaced no closer
than 6 feet (1.83 m).
Keeping Air Spaces Clean
In a drainage wall system, such as a cavity wall or an anchored veneer wall, it is essential that the air space be
kept clean. If it is not, mortar droppings may clog the weeps, protrusions may span the air space and water pen-
etration to the interior may occur.
To the greatest extent possible, mortar droppings should be prevented from falling into the air space or cavity. An
aid to prevent this is to bevel the bed joint away from the air space or cavity, as shown in Figure 6. When brick
are laid on a beveled bed joint, a minimum of mortar is squeezed out of the joint, as shown in Photo 7. The mor-
tar squeezed from the joints on the air space or cavity side may be troweled onto the units. This same procedure
may be used for laying the exterior wythes of grouted and reinforced brick cavity walls.
Another method allows access to the base of the cavity for cleaning. When the brickwork is initially constructed,
every third brick or so in the course above the flashing of the exterior wythe is omitted. Once the brickwork is com-
plete, mortar droppings at the base of the cavity can be easily removed and weeps provided when the omitted
brick are placed in the wall with mortar.
Alternately, a wooden or metal strip, slightly smaller than the cavity width, can be placed in the air space. This
strip rests on the wall ties as the wall is built. Wire or rope is attached to the strip so the strip can be lifted out as
the mason builds the wall. Care should be taken when raising or removing the strip to not disturb the brickwork.
a
b
Beveled
Bed J oints
Beveled
Bed J oints
Beveled
Bed J oints
Figure 6
Beveled Bed Joints
Photo 7
Beveled and Conventional Mortar Joints
(a) Beveled J oint; (b) Conventional J oint
www.gobrick.com | Brick Industry Association | TN 7B | Water Penetration Resistance - Construction and Workmanship | Page 7 of 7
Drainage materials and mortar dropping control devices may also be used to keep the air space adjacent to the
weeps free from mortar. Use of these devices does not guarantee that bridging of the air space will not occur, thus
the amount of mortar droppings should be limited as much as possible.
Disturbance of Newly Laid Masonry
Newly laid brick should never be pushed, shoved, tapped or otherwise disturbed once they are laid in their final
position and the mortar has begun to set. Any disturbance at this point will break the bond and may lead to a leak.
If adjustments are necessary, the incorrectly placed brick should be removed and re-laid in fresh mortar.
Protection of Unfinished Brickwork
Covering of masonry walls at the end of each work day, and especially in times of inclement weather, is essential
for satisfactory performance. Covering unfinished walls with tarpaulins or other water-resistant materials, securely
tied or weighted in position, should be rigorously enforced. Mortar boards, scaffold planks and light plastic sheets
weighted with brick should not be accepted as suitable cover. Metal clamps, similar to bicycle clips, are commer-
cially available in a variety of sizes to meet various wall thicknesses. These are used in conjunction with plastic
sheets or water-repellent tarpaulins and offer excellent protection for extended periods of time.
Tops of walls should also be covered after the masons work is finished if a permanent coping is not attached
immediately after the brickwork is completed. Protection of openings in brickwork such as those for windows,
movement joints, etc. should also be considered as they may allow moisture ingress from rain and snow and can
lead to moisture-related problems such as efflorescence, and in some cases could affect the final mortar color.
SUMMARY
Quality construction practices and good workmanship are essential to achieve brickwork that is resistant to water
penetration. This Technical Note does not cover all construction practices, but describes material storage and
preparation procedures, construction practices and installation techniques that are indicative of high quality and,
when combined with proper design, detailing and materials, result in brickwork that is resistant to water penetra-
tion.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. The BDA Guide to Successful Brickwork, Second Edition, The Brick Development Association, Arnold (a
member of the Hodder Headline Group), London, England, 2000.
2. Drysdale, R.G., Hamid, A.A., and Baker, L.R., Masonry Structures: Behavior and Design, Second Edition,
The Masonry Society, Boulder, CO, 1999.
3. Koski, J .A., Waterproof the Backup Wythe, Masonry Construction, August 1992.
4. Specification for Masonry Structures, ACI 530.1-05/ASCE 6-05/TMS 602-05, The Masonry Society,
Boulder, CO, 2005.
2008 Brick Industry Association, Reston, Virginia Page 1 of 11
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
8
January
2008
Mortars for Brickwork
Abstract: This Technical Note addresses mortars for brickwork. The major ingredients of mortar are identified. Means of
specifying mortar are covered. Mortar properties are described, as well as their effect on brickwork. Information is provided for
selection of the appropriate materials for mortar and properties of mortars.
Key Words: hardened mortar properties, mortar, plastic mortar properties, specifications, Types of mortar.
General
Use mortar complying with ASTM C270
For typical project requirements, use proportion
specifications of ASTM C270
Select mortar Type using recommendations of Technical
Note 8B
Use Type N mortar for normal use, including most veneer
applications
Avoid combining two air-entraining agents in mortar
Mortar Materials
Cementitious:
Use cement complying with ASTM C150 (portland
cement), ASTM C595 (blended hydraulic cement), or
ASTM C1157 (hydraulic cement) in combination with
either hydrated lime complying with ASTM C207, Type S,
or lime putty complying with ASTM C1489
Use mortar cement complying with ASTM C1329
Use masonry cement complying with ASTM C91
Aggregate:
Use natural or manufactured sand complying with
ASTM C144
Water:
Use potable water free of deleterious materials
Mortar Admixtures
Use admixtures complying with ASTM C1384
When using a bond enhancer admixture, do not use an
air-entraining agent
When using a set retarding admixture, do not retemper
mortar
Do not use water-repellent admixtures
Pigments
Use pigments complying with ASTM C979
Use as little pigment as possible
For metallic oxide pigments, limit quantity to 10 percent of
cement content by weight
For carbon black pigment, limit quantity to 2 percent of
cement content by weight
Avoid using pigments containing Prussian blue, cadmium
lithopone and zinc and lead chromates
Premix cement and coloring agents in large, controlled
quantities
Do not retemper colored mortar
INTRODUCTION
Mortar is the bonding agent that integrates brick into a masonry assembly. Mortar must be strong, durable
and capable of keeping the masonry intact, and it must help to create a water-resistant barrier. Also, mortar
accommodates dimensional variations and physical properties of the brick when laid. These requirements are
influenced by the composition, proportions and properties of mortar ingredients.
Because concrete and mortar contain the same principal ingredients, it is often erroneously assumed that good
concrete practice is also good mortar practice. In reality, mortar differs from concrete in working consistencies,
methods of placement and structural performance. Mortar is used to bind masonry units into a single element,
developing a complete, strong and durable bond. Concrete, however, is usually a structural element in itself.
Mortar is usually placed between absorbent masonry units and loses water upon contact with the units. Concrete
is usually placed in nonabsorbent metal or wooden forms, which absorb little if any water. The importance of the
water/cement ratio for concrete is significant, whereas for mortar it is less important. Mortar has a high water/
cement ratio when mixed, but this ratio changes to a lower value when the mortar comes in contact with the
absorbent units.
The most frequently used means of specifying mortar is ASTM C270, Standard Specification for Mortar for Unit
Masonry [Ref. 1]. This standard contains information on specifying and using mortar. This Technical Note uses
ASTM C270 as a basis and addresses the materials, properties and means of specifying mortars. The other
Technical Note in this series addresses the selection and quality control of mortars.
SUMMARY OF RECOMMENDATIONS:
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 2 of 11
MATERIALS
Historically, mortar has been made from a variety of materials. Burned gypsum and sand were used to make
mortar in ancient Egypt, while lime and sand were used extensively in this country before the 1900s. Currently,
the basic dry ingredients for mortar include some type of cement, hydrated lime and sand. Each of these materials
makes a definite contribution to mortar performance.
Portland and Other Hydraulic Cements
Portland cement, a hydraulic cement, is the principal cementitious ingredient for cement-lime mortar. It contributes
to durability, high strength and early setting of the mortar. Portland cement used in masonry mortar should conform
to ASTM C150, Standard Specification for Portland Cement [Ref. 1]. Of the eight portland cement Types covered
by ASTM C150, only three are recommended for use in masonry mortars:
Type I - For general use when the special properties of Types II and III are not required.
Type II - For use when moderate sulfate resistance or moderate heat of hydration is desired.
Type III - For use when high early strength is desired.
ASTM C270 permits the use of other hydraulic cements in mortar. Some of these materials may slow the strength
gain or may affect the color of mortar. The material standards for these cements are ASTM C595, Standard
Specification for Blended Hydraulic Cements [Ref. 1], such as portland blast-furnace slag cement, portland-
pozzolan cement and slag cement; and ASTM C1157, Standard Performance Specification for Hydraulic Cement
[Ref. 1]. The use of blended hydraulic cements is not recommended unless the mortar containing such cements
meets the property specifications of ASTM C270.
Because high air entrainment can significantly reduce the bond between the mortar and brick or reinforcement, the
use of air-entrained portland, blended hydraulic or hydraulic cements is not recommended. Most building codes
have lower allowable flexural tensile stress values for mortar made with air-entrained cementitious materials.
Masonry Cements
Masonry cements are proprietary cementitious materials for mortar. They are widely used because of their
convenience and good workability. ASTM C91, Standard Specification for Masonry Cement [Ref. 1], defines
masonry cement as a hydraulic cement, primarily used in masonry and plastering construction, consisting of
a mixture of portland or blended hydraulic cement and plasticizing materials (such as limestone, hydrated or
hydraulic lime) together with other materials introduced to enhance one or more properties such as setting time,
workability, water retention, and durability. ASTM C91 provides specific criteria for physical requirements and
performance properties of masonry cements. The constituents of masonry cement may vary depending on the
manufacturer, local construction practices and climatic conditions.
Masonry cements are classified into three Types by ASTM C91: Types M, S and N. The current edition of ASTM
C91 requires a minimum air content of 8 percent (by volume) and limits the maximum air content to 21 percent
for Type N masonry cement and 19 percent for Types S and M masonry cements. Mortar prepared in the field will
typically have an air content that is 2 to 3 percent lower than mortar tested under laboratory conditions.
In the model building codes, allowable flexural tensile stress values for masonry built with masonry cement mortar
are lower than those for masonry built with non-air-entrained portland cement-lime mortar. Therefore, the use of
masonry cement should be based on the requirements of the specific application.
Mortar Cements
Mortar cements are hydraulic cements, consisting of a mixture of portland or blended hydraulic cement, plasticizing
materials such as limestone or hydrated or hydraulic lime, and other materials intended to enhance one or more
of the properties of mortar. In this respect, mortar cement is similar to masonry cement. However, ASTM C1329,
Standard Specification for Mortar Cement [Ref. 1], includes requirements for maximum air content and minimum
flexural bond strength that are not found in the masonry cement specification. Because of the strict controls on
air content and the minimum strength requirement, mortar cement and portland cement-lime mortars are treated
similarly in the Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05) [Ref. 5].
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 3 of 11
Three Types of mortar cements are specified in ASTM C1329: Types M, S and N. Physical requirements vary
depending upon mortar cement Type. Air content for all three Types must be a minimum of 8 percent. The
maximum air content is 14 percent for Types M and S and 16 percent for Type N. Flexural bond strength, as
measured by the test method in ASTM C1072, Standard Test Method for Measurement of Masonry Flexural
Bond Strength [Ref. 1], is also specified. The minimum flexural bond strength for these mortar cements is 115 psi
(0.8 MPa) for Type M, 100 psi (0.7 MPa) for Type S and 70 psi (0.5 MPa) for Type N.
Hydrated Lime and Lime Putty
Hydrated lime is a derivative of limestone that has been through two chemical reactions to produce calcium
hydroxide. Lime contributes to extent of bond, workability, water retention and elasticity.
Hydrated lime in ASTM C207, Standard Specification for Hydrated Lime for Masonry Purposes [Ref. 1], is
available in four Types. Only Type S hydrated lime should be used in mortar. Type N hydrated lime contains no
limits on the quantity of unhydrated oxides. Types NA and SA lime contain air-entraining additives that reduce the
extent of bond between the mortar and masonry units or reinforcement, and are therefore not recommended for
mortar.
ASTM C1489, Standard Specification for Lime Putty for Structural Purposes [Ref. 1], is prepared from hydrated
lime and is often used in restoration projects.
Because lime hardens only upon contact with carbon dioxide in the air, hardening occurs over a long period of
time. However, if small hairline cracks develop, water and carbon dioxide that penetrate the joint will react with
calcium hydroxide from the mortar and form calcium carbonate. The newly developed calcium carbonate will seal
the cracks, limiting further water penetration. This process is known as autogenous healing.
Aggregates
Aggregates (sand) act as a filler material in mortar, providing for an economical mix and controlling shrinkage.
Either natural sand or manufactured sand may be used. Gradation limits are given in ASTM C144, Standard
Specification for Aggregates for Masonry Mortar [Ref. 1].
Gradation can be easily and inexpensively altered by adding fine or coarse sands. Sometimes the most feasible
method requires proportioning the mortar mix to suit the available sand, rather than requiring sand to meet a
particular gradation. However, if the sand does not meet the grading requirement of ASTM C144, it can only be
used provided the mortar meets the property specifications of ASTM C270.
Water
Water that is clean, potable and free of deleterious acids, alkalis or organic materials is suitable for masonry
mortars.
Admixtures
Admixtures are sometimes used in mortar to obtain a specific mortar color, increase workability, decrease setting
time, increase setting time, increase flexural bond strength or act as a water repellent [Ref. 2]. Admixtures to
achieve a desired color of the mortar are the most widely used. Although some admixtures are harmless, some
are detrimental to mortar and the resulting brickwork. Because the properties of both plastic and hardened mortars
are highly dependent on mortar ingredients, the use of admixtures should not be considered unless their effect on
the mortar is known. Admixtures also should be examined for their effect on the masonry, masonry units and items
embedded in the brickwork. For example, admixtures containing chlorides promote corrosion of embedded metal
anchors and therefore should not be used. ASTM C1384, Standard Specification for Admixtures for Masonry
Mortars [Ref. 1], provides methods to evaluate the effect of admixtures on mortar properties. The admixtures
represented in ASTM C1384 are as follows:
Bond Enhancers. Bond enhancers improve flexural bond strength, surface density and freeze-thaw resistance.
They are typically used to increase bond strength to smooth, dense surface units and applications such as
copings and pavers. Bond enhancers should not be used with air-entraining agents.
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 4 of 11
Set Accelerators. Set accelerators shorten the time required for cement hydration to occur and typically reduce
the setting time by 30 to 40 percent. They are typically used to reduce the time required for cold weather protective
measures. Set accelerators typically increase short-term compressive strengths and may affect color.
Set Retarders. Set retarders increase the board life of fresh mortar by increasing the time required for cement
hydration to occur. They are typically used in conjunction with hot weather protective measures or to aid in
reducing the rapid suction associated with high initial rate of absorption (IRA) brick. Mortar with set retarders
should not be retempered, and severely retarded mortar may require moist curing to maintain hydration. Set
retarders typically reduce short-term compressive strength and may affect color.
Water Repellents. Water repellent admixtures are typically used in conjunction with concrete masonry units where
the admixture is added to both the mortar and to the concrete masonry units. When water-repellent admixtures are
used in the mortar alone, they may inhibit bond and are not recommended for use with brick.
Workability Enhancers. Workability enhancers add viscosity to mortar mixes, allowing easier placement of mortar
on masonry units. The benefits of workability enhancers are subjective, and their use is more to suit the liking of
the mason. They should be reviewed to ensure that there are no deleterious effects on the mortar.
Colored Mortar
Colored mortars may be obtained through the use of colored aggregates or suitable pigments. The use of colored
aggregates is preferable when the desired mortar color can be obtained. White sand, ground granite, marble
or stone usually have permanent color and do not weaken the mortar. For white joints, use white sand, ground
limestone or ground marble with white portland cement and lime.
Most pigments that conform to ASTM C979, Standard Specification for Pigments for Integrally Colored Concrete
[Ref. 1], are suitable for mortar. Mortar pigments must be sufficiently fine to disperse throughout the mix, capable
of imparting the desired color when used in permissible quantities, and must not react with other ingredients to
the detriment of the mortar. These requirements are generally met by metallic oxide pigments. Carbon black and
ultramarine blue also have been used successfully as mortar colors. Avoid using organic colors and, in particular,
those colors containing Prussian blue, cadmium lithopone and zinc and lead chromates. Paint pigments may not
be suitable for mortars.
Use as little pigment as is needed to produce the desired results; an excess may seriously impair strength and
durability. The maximum permissible quantity of most metallic oxide pigments is 10 percent of the cement content
by weight. Although carbon black is a very effective coloring agent, it will greatly reduce mortar strength when used
in greater proportions. Therefore, limit carbon black to 2 percent of the cement content by weight.
For best results, use cement and coloring agents premixed in large, controlled quantities. Premixing large
quantities will ensure more uniform color than can be obtained by mixing smaller batches in the field. A consistent
mixing sequence is essential for color consistency when mixing smaller batches in the field. Further, use the same
source of mortar materials throughout the project.
Color uniformity varies with the amount of mixing water, the moisture content of the brick when laid and whether
the mortar is retempered. The time and degree of tooling and cleaning techniques also will influence final mortar
color. Color permanence depends upon the quality of pigments and the weathering and efflorescing qualities of the
mortar.
SPECIFYING MORTAR
Masonry mortars are classified by ASTM C270 into four Types: M, S, N and O. Each mortar Type consists of
aggregate, water and one or more of the four cementitious materials (portland or hydraulic cement, mortar cement,
masonry cement and lime) listed in the previous section.
There are two methods of specifying mortar by Type in ASTM C270: proportion specifications and property
specifications. A cement-lime mortar, a mortar cement mortar, or a masonry cement mortar is permitted. The type
of cementitious material desired should be specified.
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 5 of 11
Proportion Specifications
The proportion specifications require that mortar materials be mixed according to given volumetric proportions. If
mortar is specified by this method, no laboratory testing is required, either before or during construction. Table 1
lists proportion requirements of the various mortar Types. Note that masonry cement and mortar cement may be
used alone to produce Type M, S, N or O mortars. Additionally, Type N mortar cement or masonry cement may be
combined with portland cement to produce a Type M or Type S mortar.
Mortar Type
Proportions by Volume (Cementitious Materials)
Aggregate Ratio
(Measured in
Damp, Loose
Conditions)
Portland
or Blended
Cement
Mortar Cement Masonry Cement
Hydrated Lime
or Lime Putty
M S N M S N
Cement
Lime
M 1

Not less than


2 and not
more than 3
times the sum
of the separate
volumes of
cementitious
materials
S 1

over to
N 1

over to 1
O 1

over 1 to 2
Mortar
Cement
M 1

1

M

1

S

1

S

1

N

1

O

1

Masonry
Cement
M 1

1

M

1

S

1

S

1

N

1

O

1

TABLE 1
Proportion Specification Requirements
Note: Two air-entraining materials shall not be combined in mortar
The volumetric proportions given in Table 1 can be converted to weight proportions using assumed weights per
cubic foot (cubic meter) for the materials as follows:
Portland cement 94 lb (1506 kg)
Masonry, mortar and blended cements Varies, use weight printed on bag
Hydrated lime 40 lb (641 kg)
Lime putty 80 lb (1281 kg)
Sand, damp and loose 80 lb (1281 kg) of dry sand
Property Specifications
The property specifications require a mortar mix of the materials to be used for construction to meet the specified
properties under laboratory testing conditions. If mortar is specified by the property specifications, compressive
strength, water retention and air content tests must be performed prior to construction on mortar mixed in the
laboratory with a controlled amount of water. The material quantities determined from the laboratory testing are
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 6 of 11
then used in the field with the amount of water determined by the mason. Table 2 lists property requirements of
the various mortar Types. Properties of field-mixed mortar cannot be compared to the requirements of the property
specifications because of the different amounts of water used in the mortars, the use of different mixers and the
different curing conditions. Field sampling of mortar, where specified, is typically performed for tracking project
consistency from beginning to end. It is not to be used for compliance with property specifications. Additional
information about this type of quality assurance testing can be found in Technical Note 8B.
Proportion vs. Property Specifications
The specifier should indicate in the project specifications whether the proportion or the property specifications are
to be used. If the specifier does not indicate which should be used, then the proportion specifications govern by
default. The specifier also should confirm that the mortar Types selected and the materials indicated in the project
specifications are consistent with the structural design requirements of the masonry.
Mortar prepared by the proportion specifications is not to be compared to mortar of the same Type prepared by the
property specifications. A mortar that is mixed according to the proportion specification will have a higher laboratory
compressive strength than that of the corresponding mortar Type under the property specification [Ref. 7].
PHYSICAL PROPERTIES OF MORTAR
Mortars have two distinct, important sets of properties: those in the plastic state and those in the hardened
state. The plastic properties help to determine the mortars compatibility with brick and its construction suitability.
Properties of plastic mortar include workability, water retention, initial flow and flow after suction. Properties of
hardened mortars help determine the performance of the finished brickwork. Hardened properties include flexural
bond strength, durability, extensibility and compressive strength. Properties of plastic mortar are more important to
the mason, while the properties of hardened mortar are more important to the designer and owner.
Workability
Workability is the most important physical property of plastic mortar. A mortar is workable if its consistency allows
it to be spread with little effort and if it will readily adhere to vertical masonry surfaces. This results in good extent
of bond between the mortar and the brick, which provides resistance to water penetration. Although experienced
masons are good judges of the workability of a mortar and have developed various methods to determine
suitability, there is no standard laboratory or field test for measuring this property.
TABLE 2
Property Specification Requirements
1
Mortar Type
Average Compressive
Strength at 28 Days,
min. psi (MPa)
Water
Retention,
min. %
Air Content,
max. %
Aggregate Ratio (Measured
in Damp, Loose Conditions)
Cement
Lime
M 2500 (17.2) 75 12
Not less than 2 and
not more than 3 times
the sum of the separate
volumes of cementitious
materials
S 1800 (12.4) 75 12
N 750 (5.2) 75 14
2
O 350 (2.4) 75 14
2
Mortar
Cement
M 2500 (17.2) 75 12
S 1800 (12.4) 75 12
N 750 (5.2) 75 14
2
O 350 (2.4) 75 14
2
Masonry
Cement
M 2500 (17.2) 75 18
S 1800 (12.4) 75 18
N 750 (5.2) 75 20
3
O 350 (2.4) 75 20
3
1. Laboratory prepared mortar only.
2. When structural reinforcement is incorporated in cement-lime or mortar-cement mortar, the maximum air content shall be 12 percent.
3. When structural reinforcement is incorporated in masonry-cement mortar, the maximum air content shall be 18 percent.
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 7 of 11
Water retention, flow and resistance to segregation affect workability. In turn, these are affected by properties of
the mortar ingredients. Because of this complex relationship, quantitative estimates of workability are difficult to
obtain. Until a test is developed, the requirements for water retention and aggregate gradation must be relied upon
to provide a quantitative measure of workability.
Water Content
Water content is possibly the most misunderstood aspect of masonry mortar, probably due to the similarity
between mortar and concrete materials. Many designers mistakenly base mortar specifications on the assumption
that mortar requirements are similar to concrete requirements, especially with regard to the water/cement ratio.
Many specifications incorrectly require mortar to be mixed with the minimum amount of water consistent with
workability. Often, retempering of the mortar is prohibited. These provisions result in mortars that have higher
compressive strengths but lower bond strengths. Mixing mortar with the maximum amount of water consistent
with workability will provide maximum bond strength within the capacity of the mortar. As a result, water content
normally should be determined by the mason or bricklayer to produce the best workability. Retempering is
permitted, but only to replace water lost by evaporation. This is usually controlled by the requirement that all mortar
be used within 2 hours after initial mixing, or as determined for hot weather construction.
Water Retention
Water retention is the ability of a mortar to hold water when placed in contact with absorbent masonry units. The
laboratory value of water retention is the ratio of flow after suction to the initial flow, expressed in a percentage.
Flow after suction, as described in ASTM C91, is determined by subjecting the mortar to a vacuum and
remeasuring the flow of the mortar. A mortar that has low water retention will lose moisture more rapidly. This is
used in conjunction with the IRA of the brick to select mortar materials and Type.
In general, the following will increase water retention:
1. Addition of sand fines within allowable gradation limits.
2. Use of highly plastic lime (Type S lime).
3. Increased air content.
4. Use of hydraulic cement containing very fine pozzolans.
Initial Flow
Initial flow is essentially a measure of the mortars water content. It can be measured by either of two methods:
ASTM C109, Standard Test Method for Compressive Strength of Hydraulic Cement Mortars [Ref. 1], or ASTM
C780, Standard Test Method for Preconstruction and Construction Evaluation of Mortars for Plain and Reinforced
Unit Masonry [Ref. 1].
In ASTM C109, a truncated cone of mortar is formed on a flow table, which is then mechanically raised 1 in.
(25.4 mm) and dropped 25 times in 15 seconds. During this test, the mortar will flow, increasing the diameter of the
mortar specimen. The initial flow is the ratio of the increase in diameter from the initial 4 in. (102 mm) cone base
diameter, expressed in a percentage. Flow rates are laboratory tests.
In ASTM C780, a 3 in. (89 mm) high hollow cylinder is filled with mortar, and a cone-shaped plunger, whose
point is placed at the top of the cylinder, is dropped into the mortar. The depth of the cone penetration into the
mortar is measured in millimeters. The greater the penetration of the cone into the mortar, the greater its flow or
water content. Cone penetration can be measured in the laboratory or in the field.
Laboratory mortars are mixed to have an initial flow of only 105 to 115 percent. Construction mortars normally have
initial flows in the range of 130 to 150 percent (sometimes higher in hot weather) to produce workability satisfactory to
the mason. Requirements for laboratory-prepared mortar should not be applied to field-prepared mortar. Test results
of laboratory-prepared mortar should not be compared to test results of field-prepared mortar without considering the
initial flow of each. The lower initial flow requirements for laboratory mortars were set to allow for more consistent test
results on most available laboratory equipment, and to compensate for water absorbed by the units.
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 8 of 11
Extensibility and Plastic Flow
Extensibility is another term for maximum tensile strain at failure. It reflects the maximum elongation possible
under tensile forces. High-lime mortars exhibit greater plastic flow than low-lime mortars. Plastic flow, or creep,
acting with extensibility will impart some flexibility to the masonry, permitting slight movement. Where greater
resiliency for movement is desirable, the lime content may be increased while still satisfying other requirements.
Flexural Bond Strength
Flexural bond strength is perhaps the most important physical property of hardened mortar. For veneer
applications, the bond strength of mortar to brick units provides the ability to transfer lateral loads to veneer
anchors. For loadbearing applications, the bond influences the overall strength of the wall for resisting lateral and
flexural loads. Variables that affect the bond strength include texture of the brick, suction of the brick, air content
of the mortar, water retention of the mortar, pressure applied to the joint during forming, mortar proportions and
methods of curing.
Brick Texture. The texture of a brick affects the mechanical bond between the brick and mortar [Ref. 8]. Mortar
bond is greater to roughened surfaces, such as wire-cut surfaces, than to smooth surfaces, such as die-skin
surfaces. Sanded and coated surfaces can reduce the bond strength depending upon the amount and type of
material on the surface and its adherence to the surface.
Brick IRA (Suction). The laboratory-measured initial rate of absorption (IRA) of brick indicates the bricks suction
and whether it should be considered for wetting before use. It is the IRA at the time of laying that influences bond
strength. In practically all cases, mortar bonds best to brick with an IRA less than 30 g/min/30 in.
2
(30 g/min/
194 cm
2
) when laid. If the bricks IRA exceeds this value, then the brick should be wetted three to 24 hours before
laying. Wetted brick should be surface dry when they are laid in mortar.
Several researchers have shown that IRA appears to have little influence on bond strength when the appropriate
mortar is used [Refs. 3, 4 and 9].
Air Content. Available information indicates a definite relationship between air content and bond strength of
mortar. Provided that other parameters are held constant, as air content is increased, compressive strength and
bond strength are reduced, while workability and resistance to freeze-thaw deterioration are increased [Ref. 10].
Water Content. Mortar with a high water content, or flow, at the time of use is beneficial because it can satisfy the
suction of the brick and can allow greater control of the mortar for the bricklayer. For all mortars, and with minor
exceptions for all brick suction rates, bond strength increases as flow increases. However, excessive water can
reduce both workability and bond strength.
The time lapse between spreading mortar and placing brick will affect mortar flow, particularly when mortar is
spread on brick with high suction rates, or when construction takes place during hot, dry weather. In such cases,
mortar will have less flow by the time brick are placed than when it was first spread. Conceivably, bond to brick
placed on this mortar could be materially reduced. For highest bond strength, reduce the time interval between
spreading the mortar and laying brick on top of it to a minimum.
Because not all mortar is used immediately after mixing, some of its water may evaporate while it is on the mortar
board. The addition of water to mortar (retempering) to replace water lost by evaporation should be encouraged,
when necessary. Although compressive strength may be slightly reduced and mortar color lightened if mortar
is retempered, bond strength may be lowered if it is not. ASTM C270 requires that all mortar be used within
2 hours after mixing since the mortar will begin to set. This time may be affected by hot or cold weather, as
discussed in Technical Note 1.
Materials and Proportions. There is no precise combination of materials that will always produce optimum bond.
Mortars made with cement-lime and mortar cement cementitious materials typically have higher flexural bond
strengths than do masonry cement mortars [Refs. 3, 4, 6]. Building codes prescribe the same bond strength values
to Type S and M mortars [Ref. 5].
Test Methods. Because many variables affect bond, it may be desirable to achieve reproducible results from
a small-scale laboratory test. The bond wrench test, ASTM C1072, Standard Test Method for Measurement of
Masonry Flexural Bond Strength [Ref. 1], appears to fulfill this need. It evaluates the flexural bond strength of each
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 9 of 11
joint in a masonry prism. The apparatus shown in
Figure 1 consists of a stack-bonded prism clamped in
a stationary frame. A cantilevered arm is clamped to
the top brick over the joint to be tested. The free end
of the cantilevered arm is loaded until failure, which
occurs when the clamped brick is wrenched off. The
bond wrench test has replaced previous tests of full-
sized wall specimens and prisms in which only one
joint was tested.
In general, to increase the flexural bond strength:
1. Bond mortar to a wire-cut or roughened
surface rather than a die-skin surface.
2. Wet brick with an IRA greater than 30
grams/min/30 in.
2
(30 g/min/194 cm
2
) when
laid.
3. Use Type S portland cement-lime mortar,
Type S mortar cement or Type S masonry
cement mortar with air content in the low to mid-range of ASTM C91 limits.
4. Mix mortar to the maximum flow compatible with workmanship. Use maximum mixing water
and permit retempering.
Compressive Strength
As with concrete, the compressive strength of mortar primarily depends upon the cement content and the water/
cement ratio. However, because compressive strength of masonry mortar is less important than bond strength,
workability and water retention, the latter properties should be given principal consideration in mortar selection.
The water/cement ratio of mortar as mixed in the field is reduced due to absorption of water by the adjacent brick.
Proportions. Compressive strength increases with an increase in cement content of mortar and decreases with
an increase in water content, lime content or over-sanding. Occasionally air entrainment is introduced to obtain
higher flows with lower water content. The reasoning here is that lower water/cement ratios will provide higher
compressive strengths. However, this generally proves futile since compressive strength decreases with an
increase in air content.
Test Methods. Compressive strength is measured by testing 2 in. (51 mm) mortar cubes or 2 in. (51 mm) or 3 in.
(76 mm) diameter cylinders. Procedures for molding and testing cubes appear in ASTM C109, and procedures for
molding and testing both cubes and cylinders appear in ASTM C780.
Durability
The durability of mortar in unsaturated masonry is not a serious problem. The durability of mortar is shown in the
number of masonry structures that have been in service for many years.
In general, mortar contains sufficient entrapped and entrained air to resist freeze-thaw damage. Though increasing
air content may theoretically increase the durability of masonry mortar, a decrease in bond strength, compressive
strength and other desirable properties will result. For this reason, the use of air-entraining admixtures to increase
air content is not recommended.
Volume Change
Volume changes in mortars can result from four causes: chemical reactions in hardening, temperature changes,
wetting and drying, and unsound ingredients that chemically expand. Differential volume change between brick
and mortar in a given wythe has no significant effect on performance. However, total volume change can be
significant.
Volume change caused by cement hydration (hardening) is often termed shrinkage and depends upon curing
conditions, mix proportions and water content. Mortars hardened in contact with brick exhibit considerably less
shrinkage than those hardened in nonabsorbent molds. An increase in water content will cause an increase in
Figure 1
Bond Wrench Test Apparatus
Load
Bearing Plate
Clamping
Bolts
Test Specimen
Adjustable
Base Support
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 10 of 11
shrinkage during hardening of mortar if the excess water is not removed. Change in temperature will lead to
expansion or contraction of mortar. Thermal expansion and contraction of masonry and means to accommodate
the expected movement are discussed in the Technical Note 18 Series.
Mortar swells as its moisture content increases and shrinks as it decreases. Moisture content changes with
normal cycles of wetting and drying. The magnitude of volume change due to this effect is smaller than that from
shrinkage. Unsound ingredients or impurities such as unhydrated lime oxides or gypsum can cause significant
volume change and are thus limited by ASTM C207.
Efflorescence
Efflorescence is a crystalline deposit of water-soluble salts on the surface of masonry. Mortar may be a major
contributor to efflorescence since it is a primary source of calcium hydroxide. This chemical can produce
efflorescence on its own and can react with carbon dioxide in the air or solutions from the brick to form insoluble
compounds. Mortar can contain other soluble constituents, including alkalis, sulfates and magnesium hydroxide.
Currently there is no standard test method to determine the efflorescence potential of mortar or of a brick/mortar
combination. Researchers have concluded that mortars will effloresce under any standard test.
RECOMMENDED MORTAR USES
Selection of a particular mortar Type and materials is usually a function of the needs of the finished masonry
element. Type N mortar is recommended for normal use and in most veneer applications. In applications where
high lateral strength is required, mortar with high flexural bond strength should be chosen. For loadbearing
walls and reinforced brick masonry, high compressive strength may be the governing factor. In some projects,
considerations of durability, color and flexibility may be of utmost concern. Factors that improve one property of
mortar often do so at the expense of others. For this reason, when selecting a mortar, evaluate properties of each
Type and materials and choose the combination that will best meet the particular end-use requirements. No single
mortar Type is best for all purposes. Refer to Technical Note 8B for more information on selection of mortar Type.
GREEN BUILDING/SUSTAINABILITY
Sustainability or Green Building is a movement to use resources efficiently, create healthier environments and
enhance the quality of buildings while minimizing social and environmental impacts on future generations. For
further information about the sustainability of brick masonry, refer to Technical Note 48.
While materials used to make mortar are readily abundant and produce a durable material, sustainability can be
improved further by using recycled products such as blast furnace slag cement and cements with fly ash in the
mortar to partially replace portland cement. Blast furnace slag is a by-product from the production of iron. The
waste from the production is processed to produce slag cement. When slag cement is used in mortar, it typically
makes the cement hydration process more efficient, increases long-term compressive strength, produces a tighter
pore structure and increases workability of mortar during placement. Fly ash comes from coal-fired plants used in
generating electrical power. It can replace a portion of the cement in mortar materials. Fly ash increases strength
and durability by increasing density.
SUMMARY
Mortar requirements differ from concrete requirements, principally because the primary function of mortar is to
bond masonry units into an integral element. Properties of both plastic and hardened mortars are important. Plastic
properties determine construction suitability; hardened properties determine performance of finished elements.
When selecting a mortar, evaluate all properties, and then select the mortar providing the best results overall for
the particular requirements.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
www.gobrick.com | Brick Industry Association | TN 8 | Mortars for Brickwork | Page 11 of 11
REFERENCES
1. Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2007:
Volume 4.01
C91 Standard Specification for Masonry Cement
C109 Standard Test Method for Compressive Strength of Hydraulic Cement Mortars (Using 2-in. or
[50-mm] Cube Specimens)
C150 Standard Specification for Portland Cement
C207 Standard Specification for Hydrated Lime for Masonry Purposes
C595 Standard Specification for Blended Hydraulic Cements
C1157 Standard Performance Specification for Hydraulic Cement
C1489 Standard Specification for Lime Putty for Structural Purposes
C1329 Standard Specification for Mortar Cement
Volume 4.02
C979 Standard Specification for Pigments for Integrally Colored Concrete
Volume 4.05
C144 Standard Specification for Aggregate for Masonry Mortar
C270 Standard Specification for Mortar for Unit Masonry
C780 Standard Test Method for Preconstruction and Construction Evaluation of Mortars for Plain and
Reinforced Unit Masonry
C1072 Standard Test Method for Measurement of Masonry Flexural Bond Strength
C1384 Standard Specification for Admixtures for Masonry Mortars
2. Beall, Christine, A Guide to Mortar Admixtures, Magazine of Masonry Construction, October 1989, pp.
436-438.
3. Borchelt, J .G., Melander, J .M., and Nelson, R.L., Bond Strength and Water Penetration of High IRA
Brick and Mortar, Proceedings of the Eighth North American Masonry Conference, The Masonry Society,
Boulder, CO, J une 1999, pp. 304-315.
4. Borchelt, J .G., and Tann, J .A., Bond Strength and Water Penetration of Low IRA Brick and Mortar,
Proceedings of the Seventh North American Masonry Conference, The Masonry Society, Boulder, CO,
J une 1996, pp. 206-216.
5. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
6. Matthys, J .H., Brick Masonry Flexural Bond Strength Using Conventional Masonry Mortar, Proceedings
of the Fifth Canadian Masonry Symposium, University of Vancouver, Vancouver, BC, 1992, pp. 745-756.
7. Melander, J .M., and Conway, J .T., Compressive Strengths and Bond Strengths of Portland Cement-Lime
Mortars, Masonry, Design and Construction, Problems and Repair, ASTM STP 1180, American Society
for Testing and Materials, Philadelphia, PA, 1993, pp. 105-120.
8. Ribar, J .W., and Dubovoy, V.S., Investigation of Masonry Bond and Surface Profile of Brick, Masonry:
Materials, Design, Construction and Maintenance, ASTM STP 992, American Society for Testing and
Materials, Philadelphia, PA, 1988, pp. 33-37.
9. Wood, S.L., Flexural Bond Strength of Clay Brick Masonry, The Masonry Society Journal, Vol. 13, No. 2,
The Masonry Society, Boulder, CO, February 1995, pp. 45-55.
10. Wright, B.T., Wilkin, R.D., and J ohn, G.W., Variables Affecting the Strength of Masonry Mortars, Masonry,
Design and Construction, Problems and Repair, ASTM STP 1180, American Society for Testing and
Materials, Philadelphia, PA, 1993, pp. 197-210.
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Mortars for Brickwork -
Selection and Quality Assurance
Abstract: This Technical Note discusses the selection and specification of mortar Type.
Key Words: bond strength, extent of bond, lime, masonry cement, mortar, mortar cement, portland cement, quality
assurance, sand, testing, workability.
TECHNICAL NOTES on Brick Construction 8B
October
2006
Select a mortar Type with the lowest compressive strength
meeting project requirements
Select mortar appropriate for application, project conditions
and workability
Type N mortar is recommended for normal use, including
most veneer applications
Create a quality assurance program, where appropriate, to
obtain consistent mortar
Follow recommended procedure and sequence for mixing
mortar
Measure mortar materials by volume
2006 Brick Industry Association, Reston, Virginia Page 1 of 6
INTRODUCTION
Selection of an appropriate mortar helps to ensure durable brickwork that meets performance expectations.
Mortar Type and mortar material selection should consider multiple aspects of a project, including design, brick
or masonry materials, exposure and required level of workmanship. Improper mortar selection may lead to lower
performance of the finished project.
This Technical Note provides guidance for selecting the appropriate mortar Type. It also describes a quality
assurance program to ensure the desired results. Technical Note 8 addresses specific properties of mortar, mortar
materials and their selection as well as the specification of mortar.
SELECTION OF MORTAR
Mortar bonds individual brick together to function as a single element. In its hardened state, mortar must be
durable and must help resist moisture penetration. Mortar also must have certain properties in its plastic state so
that it is both economical and easy to place.
One property of mortar that is often overemphasized is compressive strength. Stronger is not necessarily better
when specifying mortar. In fact, the opposite is often true. Mortar selection should be based on properties such as
durability and workability in addition to compressive strength.
Mortar for each project should be selected to balance the construction requirements with the performance of
the completed masonry. High lateral loads from wind or seismic activity may require a mortar that develops high
flexural tensile strength. Allowable flexural tensile and compressive stresses for unreinforced structural masonry
are given in the building code. Building code requirements may limit the use of some mortar Types under certain
conditions. For example, Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402) [Ref. 8]
does not permit the use of Type N or masonry cement mortars in any part of the lateral force-resisting system for
structures located in Seismic Design Categories D, E or F.
Other considerations may include durability (below grade or in retaining walls), color uniformity, flexibility,
workability or other desired properties. The combination of the mortar and brick properties may dictate the
selection of a certain mortar.
SUMMARY OF RECOMMENDATIONS:
These are the fundamental guidelines of mortar selection:
No single mortar is best for all purposes.
Select a mortar Type with the lowest compressive strength meeting the project requirements.
Of course, these guidelines must be used with good judgment. For example, it could be uneconomical and unwise
to use different mortars for various portions of the same structure.
Mortar Type Characteristics
Mortars are classified by ASTM C 270, Standard Specification for Mortar for Unit Masonry [Ref. 2], into four Types:
M, S, N and O. These four Types of mortar can be made with portland cement, masonry cement, mortar cement
or blended cements some of which are combined with hydrated lime.
Each mortar Type has some basic characteristics:
Type N mortar - General all-purpose mortar with good bonding capabilities and workability
Type S mortar - General all-purpose mortar with higher flexural bond strength
Type M mortar - High compressive-strength mortar, but not very workable
Type O mortar - Low-strength mortar, used mostly for interior applications and restoration
Although the descriptions above provide basic mortar characteristics, each mortar Type can be used in a variety
of applications. No single mortar is best for all purposes.
Simplistic Mortar Selection
The easiest method to select mortar is to remember the following mnemonic:
Type N for normal brickwork applications
Type S for stronger brickwork applications
Normal applications include most veneer. Stronger applications are needed in high seismic and high wind areas
and in reinforced brickwork.
Mortar Selection Based on Use
More explicit guidance on mortar selection based on the location and use of the building segment is given in Table
1. More durable mortar Types are recommended for more severe exposures.
TABLE 1
Mortar Recommendations Based on Use
Brick Properties Influencing Mortar Selection
In general, the bond between brick and mortar is the most important property to consider when selecting mortar
Type. Bond actually has two components: extent of bond and bond strength. Extent of bond refers to the amount
of intimate contact between the mortar and brick, which is enhanced by good mortar workability. Good extent
of bond provides durability and resistance to water penetration. Bond strength refers to the force required to
separate the mortar from the brick. Good bond strength provides resistance to cracking.
www.gobrick.com | Brick Industry Association | TN 8B | Mortars for Brickwork - Selection and Quality Assurance | Page 2 of 6
Recommended Alternate
Exterior, above grade Reinforced or Loadbearing walls S N
Veneer or Non-loadbearing walls N S
Parapets, Chimneys N S
Exterior, at or below grade Foundation walls, Retaining walls M S
Sewers, Manholes
Interior Loadbearing walls N S
Partitions N O or S
Mortar Type
Building Segment Location
Brick properties, particularly the initial rate of absorption (IRA), also can affect bond. Brick with a high IRA should
be used with mortar that has a greater ability to retain mixing water. Conversely, brick with a low IRA should be
used with mortar that does not retain water as easily. Bed joint surface texture also may influence bond strength
and extent of bond, but to a lesser degree than IRA.
Table 2 can be used to select a mortar based on IRA. These recommendations are based on Bond Strength and
Water Penetration of Low IRA Brick and Mortar [Ref. 6] and Bond Strength and Water Penetration of High IRA
Brick and Mortar [Ref. 7]. The mortar recommendations in Table 2 are applicable for construction in temperatures
from 40 to 100 F (4 to 37.8 C). Under colder or hotter temperatures, other brick and mortar combinations may
be preferable. Refer to Technical Note 1 for hot and cold weather construction recommendations. In addition, there
may be other brick/mortar combinations that perform as well. Bond strength of particular combinations can be
tested using ASTM C 1357, Standard Test Methods for Evaluating Masonry Bond Strength [Ref. 4].
TABLE 2
Mortar Recommendations Based on Brick Unit IRA
1
Mortars for Special Applications
Certain applications may require special considerations for mortar selection. Several of these follow:
Repointing Mortars. Repointing mortars are used in maintenance and restoration projects. Compatibility between
existing brick and mortar is the most important consideration in selecting a repointing mortar. Hence, it may be
necessary to use a weaker mortar for older masonry than would be used for new construction. In general, the
compressive strength of a repointing mortar should not exceed that of the existing mortar. If necessary, the existing
mortar can be tested to determine proportions of ingredients for the repointing mortar. Type O mortar often is used
for repointing older brickwork. Type N mortar may be suitable for repointing newer brickwork.
Repointing mortars should be pre-hydrated. In this process the mortar materials are mixed dry, and then just
enough water is added to produce a damp mix which will retain its shape when formed into a ball. After one to
one and half hours, additional water should be added to bring the mortar to the proper consistency for placement.
Refer to Technical Note 46 for more information about repointing.
Paving. Paving applications are more likely to be in a saturated condition than walls. Because of this, the mortar
typically must be more durable to resist the harsher exposure. Type M mortar is recommended with Type S as
the alternate. A mortar with a latex modifier conforming to ANSI A118.4, Specification for Latex-Portland Cement
Mortar [Ref. 1], may provide a more durable assembly. Flexible brick paving, which uses sand rather than mortar
to fill joints between pavers, is less susceptible to damage from exposure and should be considered as an
alternative to mortared paving. Refer to Technical Note 14A for more information about paving materials.
www.gobrick.com | Brick Industry Association | TN 8B | Mortars for Brickwork - Selection and Quality Assurance | Page 3 of 6
Up to 10 g/min/30 in.
(Up to 0.0005 g/min/mm)
Type S
(Type N)
Type S
(Type N)
Type S
(Type N)
10 to 30 g/min/30 in.
(0.0005 to 0.0016 g/min/mm)
Type N or S Type N or S Type N or S
Above 30 g/min/30 in.
(Above 0.0016 g/min/mm)
Dry when laid
Type N
(Type S)
__
2
__
2
Above 30 g/min/30 in.
(Above 0.0016 g/min/mm)
Wetted prior to laying
Type N
(Type S)
Type N
(Type S)
Type S
(Type N)
2
Not recommended unless verified with testing
Masonry Cement
Mortar
Initial Rate of Absorption
Range of Brick
Portland or Blended Cement:
Lime Mortar
Mortar Cement
Mortar
1
Alternate Types listed in parentheses
Stain-Resistant Mortar. Where resistance to staining is desired, aluminum tristearate, calcium stearate or
ammonium stearate may be added to the mortar. Where maximum stain resistance is desired, use mortar
consisting of one part portland cement, one-eighth part lime and two parts graded fine (80 mesh) sand,
proportioned by volume. To this, add aluminum tristearate, calcium stearate or ammonium stearate equal to
2 percent of the portland cement by weight.
Chemical-Resistant Mortar. Chemical-resistant masonry often is used in food processing plants, refineries
or breweries. Chemical-resistant mortars may include silicate mortars, sulfur mortars, various resin mortars
or cementitious mortars. For further information on chemical-resistant mortar, refer to Corrosion & Chemical
Resistant Masonry Materials Handbook [Ref. 9].
MIXING REQUIREMENTS
Although most mortar is mixed on-site, preblended mortar also is available. Preblended mortar is supplied in
consistent proportions without the need for on-site batching and measurement controls. While each mortar Type has
specified ranges of material quantities, accurate and consistent material quantities are desired throughout the job.
Material measuring and batching should be by volume or by weight to ensure that the specified mortar proportions
are accurately controlled and maintained. For material weights and recommended proportions, refer to ASTM
C 270 or Technical Note 8. When using a mechanical mixer, the ingredients should be added in such a manner
that the mix remains damp. Typically, about half the mix water is added to the mixer, followed by about half of the
sand, then any and all lime. The cement and the remainder of the sand are then added, followed by the remainder
of the water. These materials should be mixed for three to five minutes. If admixtures are to be used, they should
consistently be added at the same stage in the
mixing process. The same quantities of materials
should be added in the same order from batch to
batch to help ensure uniform results throughout
the job.
Every effort should be made to keep the
materials agitated by the paddles. This may
require changing the sequence in which water
is added. If ingredients are added too fast or if
not enough water is added to the mixer before
the dry ingredients, the mixer may not be able
to combine them, and the dry materials will stick
around the bowl.
Cement and lime should be placed in the mixer
in whole (preferable) or half bags. The mixer
should be sized accordingly, also depending
upon the project requirements and the size of the
masonry crew.
Photo 1 shows an example of batching and measurement controls that are both economical and accurate. Sand
can be measured with a 1-cubic foot (0.028 m
3
) box or a 5-gallon bucket equal to 2/3-cubic-feet (0.019 m
3
).
Alternatively, the number of shovels of sand required to fill the box or bucket can be calibrated. Shovel count
calibration should be done every morning and afternoon or whenever the shovel size or individual shovelling sand
is changed.
QUALITY ASSURANCE
A quality assurance program provides policies, procedures and requirements intended to ensure compliance with
the contract documents. Quality assurance requirements may be set by the owner, designer or governing building
code. Quality control is a part of the quality assurance program that may involve testing, inspection, or both. Some
quality assurance programs require the contractor to submit documentation showing conformance to the contract
documents. Building Code Requirements for Masonry Structures [Ref. 8] assumes that all masonry is constructed
under a quality assurance program.
Photo 1
Obtaining Accurate Sand Quantities
www.gobrick.com | Brick Industry Association | TN 8B | Mortars for Brickwork - Selection and Quality Assurance | Page 4 of 6
For mortar specified by ASTM C 270, the key to quality assurance is adherence to the material proportions
added to the mixer. ASTM C 270 prescribes the volumes of the materials in each mortar Type when the
proportion specification is used. When the property specification is used, laboratory testing establishes the
material proportions that will be used in the field. Observation during measuring and mixing is thus an essential
component of the quality assurance program. Testing may be included as a second component. ASTM C 1586,
Standard Guide for Quality Assurance for Mortars [Ref. 5], explains how to use ASTM C 270 and ASTM C 780 for
evaluating laboratory-prepared and field-prepared mortars.
Inspection
Inspection is often a part of the quality assurance programs required by the contract documents or building code.
Mortar inspection typically entails verifying that the specified materials are used and that they are in the proper
proportions. Inspection also may include verifying proper mix time, retempering, mortar placement and tooling.
Testing
Field testing of mortar is not necessary on most projects. When the ASTM C 270 property specification is used,
however, laboratory testing is necessary to establish mortar mix proportions, which are then used to prepare mortar
in the field. If inspection during mixing is not possible, some physical testing of the mortar may be appropriate.
ASTM C 780, Standard Test Method for Preconstruction and Construction Evaluation of Mortars for Plain and
Reinforced Unit Masonry [Ref. 3], provides methods for sampling and testing mortar in the laboratory and in the
field. It defines procedures for measuring properties of plastic mortar such as consistency, the aggregate ratio,
air content and water content. Finally, it defines procedures for measuring properties of hardened mortar, such as
compressive strength. These test results are used to verify mortar consistency from batch to batch.
For test results to be useful there must be a basis of comparison. Preconstruction testing with the materials to be
used during the actual construction provides the benchmark for field testing results. Proper interpretation of mortar
test results requires a thorough knowledge of mortar specifications and test methods.
For example, compressive strength test results from field-sampled mortar cannot be compared with the minimum
requirements of the ASTM C 270 property specification. The different sampling and mixing requirements of ASTM
C 780 will yield different results from those determined according to ASTM C 270. ASTM C 270 is for laboratory-
prepared and tested mortars, while ASTM C 780 is mainly for field sampling and testing. Compressive strength
results obtained according to ASTM C 780 can be expected to be lower and more variable than ASTM C 270
laboratory test results; the two are not comparable.
ASTM C 780 can be used to determine whether the proper proportions are being used in the field. Freshly
sampled mortar is placed in a jar with isopropyl or methyl alcohol to prevent hydration. The sand used in the
mortar also is sampled to determine its gradation. After weighing the materials, the fine material is filtered out of
the mortar using a sieve. The remaining material is assumed to be sand, from which the sand to cement ratio can
be determined. This can be compared with the specified proportions.
Interpreting Test Results
If ASTM C 780 field test methods are used, the results must be properly interpreted and compared with
preconstruction test results. Observations should include mortar sampling, test specimen preparation, specimen
handling during transportation, storage at the test facility and test procedures. If there is a substantial difference
between preconstruction and field results, the following should be investigated:
Change of mortar materials or proportions
Change in brick properties (different brick or wet brick) resulting in a change to the amount of water added
to the mortar
Change in time between mortar mixing and sampling
Proper construction of specimens
Unusual curing conditions
Damage to specimens during transit or storage
Proper adherence to test procedures
Accuracy of calculations
www.gobrick.com | Brick Industry Association | TN 8B | Mortars for Brickwork - Selection and Quality Assurance | Page 5 of 6
This information can be used to help identify the possible cause(s) of inconsistent test results. If questions about
mortar quality remain, additional masonry testing may be required. In some cases, prism tests of masonry
specimens from the project can be conducted to determine the structural capacity of the masonry.
SUMMARY
Mortar, although it comprises a relatively small portion of brickwork, has a significant impact on overall
performance. A range of mortars is available to suit the needs of all brick projects. Taking into consideration the
brick unit properties as well as the project requirements when specifying mortar Type will contribute to a properly
performing brick structure, as will implementing a good quality assurance plan.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
ANSI A118.4, Specification for Latex-Portland Cement Mortar, American National Standards Institute,
Washington, DC, 2006.
ASTM C 270, Standard Specification for Mortar for Unit Masonry, Annual Book of Standards, Vol. 04.05,
ASTM International, West Conshohocken, PA, 2006.
ASTM C 780, Standard Test Method for Preconstruction and Construction Evaluation of Mortars for Plain and
Reinforced Unit Masonry, Annual Book of Standards, Vol. 04.05, ASTM International, West Conshohocken,
PA, 2006.
ASTM C 1357, Standard Test Methods for Evaluating Masonry Bond Strength, Annual Book of Standards,
Vol. 04.05, ASTM International, West Conshohocken, PA, 2006.
ASTM C 1586, Standard Guide for Quality Assurance for Mortars, Annual Book of Standards, Vol. 04.05,
ASTM International, West Conshohocken, PA, 2006.
Borchelt, J .G. and Tann, J .A., Bond Strength and Water Penetration of Low IRA Brick and Mortar,
Proceedings of the Seventh North American Masonry Conference, The Masonry Society, Boulder, CO, 1996.
Borchelt, J .G., Melander, J .M., and Nelson, R.L., Bond Strength and Water Penetration of High IRA Brick
and Mortar, Proceedings of the Eighth North American Masonry Conference, The Masonry Society, Boulder,
CO, 1999.
Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402) and Specification for
Masonry Structures (ACI 530.1/ASCE 6/TMS 602), The Masonry Society, Boulder, CO, 2005.
Sheppard, Walter Lee J r., Editor, Corrosion and Chemical Resistant Masonry Materials Handbook, Noyes
Publications, Park Ridge, NJ , 1986.
1.
2.
3.
4.
5.
6.
7.
8.
9.
www.gobrick.com | Brick Industry Association | TN 8B | Mortars for Brickwork - Selection and Quality Assurance | Page 6 of 6
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Manufacturing of Brick
Abstract: This Technical Note presents fundamental procedures for the manufacture of clay brick. The types of clay used,
the three principal processes for forming brick and the various phases of manufacturing, from mining through storage, are
discussed. Information is provided regarding brick durability, color, texture (including coatings and glazes), size variation,
compressive strength and absorption.
Key Words: absorption, clays, color, cooling, compressive strength, de-hacking, drying, durability, firing, forming, hacking,
manufacturing, mining, preparation, shales, size variation, texture.
SUMMARY:
TECHNICAL NOTES on Brick Construction 9
December
2006
Brick is made of clay or shale formed, dried and fired into
a durable ceramic product.
There are three ways to form the shape and size of a
brick: extruded (stiff mud), molded (soft mud) and dry-
pressed. The majority of brick are made by the extrusion
method.
Brick achieves its color through the minerals in the fired
clay or through coatings that are applied before or after
the firing process. This provides a durable color that never
fades or diminishes.
Brick shrink during the manufacturing process as
vitrification occurs. Brick will vary in size due to the
manufacturing process. These variations are addressed by
ASTM standards.
The method used to form a brick has a major impact on
its texture. Sand-finished surfaces are typical with molded
brick. A variety of textures can be achieved with extruded
brick.
Brick manufacturers address sustainability by locating
manufacturing facilities near clay sources to reduce
transportation, by recycling of process waste, by
reclaiming land where mining has occurred, and by taking
measures to reduce plant emissions. Most brick are used
within 500 miles of a brick manufacturing facility.
2006 Brick Industry Association, Reston, Virginia Page 1 of 7
INTRODUCTION
The fundamentals of brick manufacturing have not changed over time. However, technological advancements
have made contemporary brick plants substantially more efficient and have improved the overall quality of the
products. A more complete knowledge of raw materials and their properties, better control of firing, improved kiln
designs and more advanced mechanization have all contributed to advancing the brick industry.
Other Technical Notes in this series address the classification and selection of brick considering the use, exposure
and required durability of the finished brickwork.
RAW MATERIALS
Clay is one of the most abundant natural mineral materials on earth. For brick manufacturing, clay must possess
some specific properties and characteristics. Such clays must have plasticity, which permits them to be shaped or
molded when mixed with water; they must have sufficient wet and air-dried strength to maintain their shape after
forming. Also, when subjected to appropriate temperatures, the clay particles must fuse together.
Types of Clay
Clays occur in three principal forms, all of which have similar chemical compositions but different physical
characteristics.
Surface Clays. Surface clays may be the upthrusts of older deposits or of more recent sedimentary formations. As
the name implies, they are found near the surface of the earth.
Shales. Shales are clays that have been subjected to high pressures until they have nearly hardened into slate.
Fire Clays. Fire clays are usually mined at deeper levels than other clays and have refractory qualities.
Surface and fire clays have a different physical structure from shales but are similar in chemical composition. All
www.gobrick.com | Brick Industry Association | TN 9 | Manufacturing of Brick | Page 2 of 7
three types of clay are composed of silica and alumina with varying amounts of metallic oxides. Metallic oxides
act as fluxes promoting fusion of the particles at lower temperatures. Metallic oxides (particularly those of iron,
magnesium and calcium) influence the color of the fired brick.
The manufacturer minimizes variations in chemical composition and physical properties by mixing clays from
different sources and different locations in the pit. Chemical composition varies within the pit, and the differences
are compensated for by varying manufacturing processes. As a result, brick from the same manufacturer will have
slightly different properties in subsequent production runs. Further, brick from different manufacturers that have the
same appearance may differ in other properties.
MANUFACTURING
Although the basic principles of manufacture are fairly uniform, individual manufacturing plants tailor their
production to fit their particular raw materials and operation. Essentially, brick are produced by mixing ground
clay with water, forming the clay into the desired shape, and drying and firing. In ancient times, all molding was
performed by hand. However, since the invention of brick-making machines during the latter part of the 19
th

century, the majority of brick produced in the United States have been machine made.
Phases of Manufacturing
The manufacturing process has six general phases: 1) mining and storage of raw materials, 2) preparing raw
materials, 3) forming the brick, 4) drying, 5) firing and cooling and 6) de-hacking and storing finished products (see
Figure 1).
Mining and Storage. Surface clays, shales and some
fire clays are mined in open pits with power equipment.
Then the clay or shale mixtures are transported to
plant storage areas (see Photo 1).
Continuous brick production regardless of weather
conditions is ensured by storing sufficient quantities
of raw materials required for many days of plant
operation. Normally, several storage areas (one for
each source) are used to facilitate blending of the
clays. Blending produces more uniform raw materials,
helps control color and allows raw material control for
manufacturing a certain brick body.
Figure 1
Diagrammatic Representation of Manufacturing Process
Photo 1
Clay or Shale Being Crushed
and Transported to Storage Area
www.gobrick.com | Brick Industry Association | TN 9 | Manufacturing of Brick | Page 3 of 7
Preparation. To break up large clay lumps and stones, the material is processed through size-reduction machines
before mixing the raw material. Usually the material is processed through inclined vibrating screens to control
particle size.
Forming. Tempering, the first step in the forming process, produces a homogeneous, plastic clay mass. Usually,
this is achieved by adding water to the clay in a pug mill (see Photo 2), a mixing chamber with one or more
revolving shafts with blade extensions. After pugging, the plastic clay mass is ready for forming. There are three
principal processes for forming brick: stiff-mud, soft-mud and dry-press.
Stiff-Mud Process - In the stiff-mud or extrusion process (see Photo 3), water in the range of 10 to 15
percent is mixed into the clay to produce plasticity. After pugging, the tempered clay goes through a de-
airing chamber that maintains a vacuum of 15 to 29 in. (375 to 725 mm) of mercury. De-airing removes
air holes and bubbles, giving the clay increased workability and plasticity, resulting in greater strength.
Next, the clay is extruded through a die to produce a column of clay. As the clay column leaves the die,
textures or surface coatings may be applied (see PROPERTIES, Textures, Coatings and Glazes). An
automatic cutter then slices through the clay column to create the individual brick. Cutter spacings and die
sizes must be carefully calculated to compensate for normal shrinkage that occurs during drying and firing
(see PROPERTIES, Size Variation). About 90 percent of brick in the United States are produced by the
extrusion process.
Soft-Mud Process - The soft-mud or molded process is particularly suitable for clays containing too
much water to be extruded by the stiff-mud process. Clays are mixed to contain 20 to 30 percent water
and then formed into brick in molds. To prevent clay from sticking, the molds are lubricated with either
sand or water to produce sand-struck or water-struck brick. Brick may be produced in this manner by
machine or by hand.
Dry-Press Process - This process is particularly suited to clays of very low plasticity. Clay is mixed with
a minimal amount of water (up to 10 percent), then pressed into steel molds under pressures from 500 to
1500 psi (3.4 to 10.3 MPa) by hydraulic or compressed air rams.
Drying. Wet brick from molding or cutting machines contain 7 to 30 percent moisture, depending upon the forming
method. Before the firing process begins, most of this water is evaporated in dryer chambers at temperatures
ranging from about 100 F to 400 F (38 C to 204 C). The extent of drying time, which varies with different clays,
usually is between 24 to 48 hours. Although heat may be generated specifically for dryer chambers, it usually is
supplied from the exhaust heat of kilns to maximize thermal efficiency. In all cases, heat and humidity must be
carefully regulated to avoid cracking in the brick.
Hacking. Hacking is the process of loading a kiln car or kiln with brick. The number of brick on the kiln car is
determined by kiln size. The brick are typically placed by robots or mechanical means. The setting pattern has
Photo 2
Clay is Thoroughly Mixed with Water
in Pug Mill Before Extrusion
Photo 3
After Mining, Clay is Extruded Through a Die and
Trimmed to Specified Dimension Before Firing
www.gobrick.com | Brick Industry Association | TN 9 | Manufacturing of Brick | Page 4 of 7
some influence on appearance. Brick placed face-to-
face will have a more uniform color than brick that are
cross-set or placed face-to-back.
Firing. Brick are fired between 10 and 40 hours,
depending upon kiln type and other variables. There
are several types of kilns used by manufacturers.
The most common type is a tunnel kiln, followed by
periodic kilns. Fuel may be natural gas, coal, sawdust,
methane gas from landfills or a combination of these
fuels.
In a tunnel kiln (see Photo 4), brick are loaded onto
kiln cars, which pass through various temperature
zones as they travel through the tunnel. The heat
conditions in each zone are carefully controlled, and
the kiln is continuously operated. A periodic kiln is one
that is loaded, fired, allowed to cool and unloaded,
after which the same steps are repeated. Dried brick
are set in periodic kilns according to a prescribed
pattern that permits circulation of hot kiln gases.
Firing may be divided into five general stages: 1)
final drying (evaporating free water); 2) dehydration;
3) oxidation; 4) vitrification; and 5) flashing or
reduction firing. All except flashing are associated
with rising temperatures in the kiln. Although the
actual temperatures will differ with clay or shale, final
drying takes place at temperatures up to about 400 F
(204 C), dehydration from about 300 F to 1800 F
(149 C to 982 C), oxidation from 1000 F to 1800 F
(538 C to 982 C) and vitrification from 1600 F to
2400 F (871 C to 1316 C).
Clay, unlike metal, softens slowly and melts or vitrifies
gradually when subjected to rising temperatures.
Vitrification allows clay to become a hard, solid mass
with relatively low absorption. Melting takes place in
three stages: 1) incipient fusion, when the clay particles become sufficiently soft to stick together in a mass when
cooled; 2) vitrification, when extensive fluxing occurs and the mass becomes tight, solid and nonabsorbent; and
3) viscous fusion, when the clay mass breaks down and becomes molten, leading to a deformed shape. The key
to the firing process is to control the temperature in the kiln so that incipient fusion and partial vitrification occur but
viscous fusion is avoided.
The rate of temperature change must be carefully controlled and is dependent on the raw materials, as well as
the size and coring of the brick being produced. Kilns are normally equipped with temperature sensors to control
firing temperatures in the various stages. Near the end, the brick may be flashed to produce color variations (see
PROPERTIES, Color).
Cooling. After the temperature has peaked and is maintained for a prescribed time, the cooling process begins.
Cooling time rarely exceeds 10 hours for tunnel kilns and from 5 to 24 hours in periodic kilns. Cooling is an
important stage in brick manufacturing because the rate of cooling has a direct effect on color.
De-hacking. De-hacking is the process of unloading a kiln or kiln car after the brick have cooled, a job often
performed by robots (see Photo 5). Brick are sorted, graded and packaged. Then they are placed in a storage
yard or loaded onto rail cars or trucks for delivery. The majority of brick today are packaged in self-contained,
strapped cubes, which can be broken down into individual strapped packages for ease of handling on the jobsite.
The packages and cubes are configured to provide openings for handling by forklifts.
Photo 4
Brick Enter Tunnel Kiln for Firing
Photo 5
Robotic Arm Unloading Brick After Firing
www.gobrick.com | Brick Industry Association | TN 9 | Manufacturing of Brick | Page 5 of 7
PROPERTIES
All properties of brick are affected by raw material composition and the manufacturing process. Most
manufacturers blend different clays to achieve the desired properties of the raw materials and of the fired brick.
This improves the overall quality of the finished product. The quality control during the manufacturing process
permits the manufacturer to limit variations due to processing and to produce a more uniform product.
The most important properties of brick are 1) durability, 2) color, 3) texture, 4) size variation, 5) compressive
strength and 6) absorption.
Durability
The durability of brick depends upon achieving incipient fusion and partial vitrification during firing. Because
compressive strength and absorption values are also related to the firing temperatures, these properties, together
with saturation coefficient, are currently taken as predictors of durability in brick specifications. However, because
of differences in raw materials and manufacturing methods, a single set of values of compressive strength and
absorption will not reliably indicate the degree of firing.
Color
The color of fired clay depends upon its chemical composition, the firing temperatures and the method of firing
control. Of all the oxides commonly found in clays, iron probably has the greatest effect on color. Regardless of its
natural color, clay containing iron in practically any form will exhibit a shade of red when exposed to an oxidizing
fire because of the formation of ferrous oxide. When fired in a reducing atmosphere, the same clay will assume a
dark (or black) hue. Creating a reducing atmosphere in the kiln is known as flashing or reduction firing.
Given the same raw material and manufacturing method, darker colors are associated with higher firing
temperatures, lower absorption values and higher compressive strength values. However, for products made from
different raw materials, there is no direct relationship between strength and color or absorption and color.
Texture, Coatings and Glazes
Many brick have smooth or sand-finished textures
produced by the dies or molds used in forming. A
smooth texture, commonly referred to as a die skin,
results from pressure exerted by the steel die as the
clay passes through it in the extrusion process. Most
extruded brick have the die skin removed and the
surface further treated to produce other textures using
devices that cut, scratch, roll, brush or otherwise
roughen the surface as the clay column leaves the
die (see Photo 6). Brick may be tumbled before or
after firing to achieve an antique appearance.
Many manufacturing plants apply engobes (slurries)
of finely ground clay or colorants to the column.
Engobes are clay slips that are fired onto the ceramic
body and develop hardness, but are not impervious
to moisture or water vapor. Sands, with or without
coloring agents, can be rolled into an engobe or
applied directly to the brick faces to create interesting
and distinctive patterns in the finished product.
Although not produced by all manufacturers, glazed brick are made through a carefully controlled ceramic glazing
procedure. There are two basic variations of glazing; single-fired and double-fired. Single-fired glazes are sprayed
on brick before or after drying and then kiln-fired at the normal firing temperatures of the brick. Double-fired glazes
are used to obtain colors that cannot be produced at higher temperatures. Such a glaze is applied after the brick
body has been fired and cooled, then refired at temperatures less than 1800 F (982 C). Glazes are available in a
wide variety of colors and reflectances. Unlike engobes, glazes are impervious to water and water vapor.
Photo 6
Some Brick Textures are Applied by Passing
Under a Roller After Extrusion
www.gobrick.com | Brick Industry Association | TN 9 | Manufacturing of Brick | Page 6 of 7
Size Variation
Because clays shrink during both drying and firing, allowances are made in the forming process to achieve the
desired size of the finished brick. Both drying shrinkage and firing shrinkage vary for different clays, usually falling
within the following ranges:
Drying shrinkage: 2 to 4 percent
Firing shrinkage: 2.5 to 4 percent
Firing shrinkage increases with higher temperatures, which produce darker shades. When a wide range of colors
is desired, some variation between the sizes of the dark and light units is inevitable. To obtain products of uniform
size, manufacturers control factors contributing to shrinkage. Because of normal variations in raw materials and
temperature variations within kilns, absolute uniformity is impossible. Consequently, specifications for brick allow
size variations.
Compressive Strength and Absorption
Both compressive strength and absorption are affected by properties of the clay, method of manufacture and
degree of firing. For a given clay and method of manufacture, higher compressive strength values and lower
absorption values are associated with higher firing temperatures. Although absorption and compressive strength
can be controlled by manufacturing and firing methods, these properties depend largely upon the properties of the
raw materials.
ENVIRONMENTAL ISSUES
Brick manufacturing is one of the most efficient uses of materials to produce a product. Brick plants are typically
located close to raw material sources. Processed clay and shale removed in the forming process before firing are
returned to the production stream. Brick not meeting standards after firing are culled from the process and ground
to be used as grog in manufacturing brick or crushed to be used as landscaping material. There is virtually no
waste of raw materials in manufacturing brick.
Brick manufacturing uses readily available raw
materials, including some waste products. The
primary ingredient, clay, has been termed an
abundant resource by many authorities including the
American Institute of Architects [Ref. 1], confirming
that depletion of clay is not a concern. Nonhazardous
waste products from other industries are sometimes
used. Examples include using bottom- and fly-ash
from coal-fired generators, using other ceramic
materials as grog, using lubricants derived from
processing organic materials in the forming of brick,
and using sawdust as a burnout material.
The brick industrys goal is to reduce resources used
in the manufacturing process. Although water is used
in brick manufacturing, it is not chemically altered but
is evaporated into the atmosphere. By using storage
tanks to recirculate and reuse water, potable water
demand can be cut dramatically. Brick manufacturers
are continuously looking for ways to minimize use of
water. Photo 7 shows one plant using a storage tank to
hold recirculated water for reuse in brick production.
While natural gas is the most frequently used energy source for brick manufacturing, many manufacturers are
using waste products, such as methane gas from landfills and sawdust, for brick firing.
The brick industry recognizes the need for compliance with state and federal regulations for clean air and the
environment. Air emissions are minimized with controls such as scrubbers installed on kiln exhausts. Dust in plants
Photo 7
Left Storage Tank Captures Used Water from
Manufacturing Process; Water is Cleaned, Cooled
and Moved to Holding Tank for Reuse
www.gobrick.com | Brick Industry Association | TN 9 | Manufacturing of Brick | Page 7 of 7
is controlled through the use of filtering systems, vacuums, additives and water mists. Mined areas are reclaimed
by replacing overburden and topsoil so the resulting property can be used for a wide variety of functions, including
farmland, residential and commercial sites, and even wetlands.
Current manufacturing processes for brick are similar in scope to those used for the past 3500 years. Over this
period of time, it has been demonstrated that brick are safe and durable products for society. The long service life
of brickwork is a key component of sustainable structures and pavements.
The Brick Industry Association has adopted the following environmental policy statement:
The brick industry recognizes that the stewardship of our planet lies in the hands of our generation. Our
goal is to continually seek out innovative, environmentally friendly opportunities in the manufacturing
process and for the end use of clay brick products. As demonstrated over time, we are committed
to manufacturing products that provide exceptional energy efficiency, durability, recyclability, and low
maintenance with minimal impact on the environment from which they originate. We will ensure that our
facilities meet or exceed state and federal environmental regulations, and we will continue to partner with
building professionals to help them in using our products to create environmentally responsible living and
working spaces for todays and future generations.
SUMMARY
This Technical Note on manufacturing brick is the first in a series covering the manufacturing, classification and
selection of brick. It provides a synopsis of the manufacturing process and discusses the various properties that
are a function of this process. More detailed descriptions of the ceramic properties of brick are not within the
purview of the Brick Industry Association. This type of information is more readily available through the National
Brick Research Center, ceramic engineers and educators.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment and a
basic understanding of the properties of brick masonry. Final decisions on the use of the information
contained in this Technical Note are not within the purview of the Brick Industry Association and
must rest with the project architect, engineer and owner.
REFERENCES
1. American Institute of Architects, Environmental Resource Guide, The American Institute of Architects,
Canada, 1998.
2. Campbell, J . W. P. and Pryce, W., Brick, A World History, Thames and Hudson, New York, NY, 2003.
2007 Brick Industry Association, Reston, Virginia Page 1 of 13
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
9A
October
2007
Specifications for and Classification
of Brick
Abstract: This Technical Note describes the predominant-consensus standard specifications for brick and the various
classifications used in each. Specific requirements including physical properties, appearance features and coring are
described. Additional requirements for each brick specification also are covered.
Key Words: appearance, ASTM standards, brick, chippage, classification, CSA standard, dimensions, distortion, durability,
exposure, grade, physical properties, specification, tolerances, type, use.
Identify the appropriate brick specification for the intended
use
Specify each classification in the specification or verify
that the default classification is valid
Specify each required action of the purchaser and
specifier
Evaluate and specify any optional requirement
Use requirements in consensus-based specifications;
deviate from them only with consideration of effect on
performance and cost
INTRODUCTION
Brick selection is made according to the specific application in which the brick will be used. Standards for brick
cover specific uses of brick and classify the brick by performance characteristics. The performance criteria include
strength, durability and aesthetic requirements. Selection of the proper specification and classification within that
specification, along with proper design and construction, should result in expected performance.
ASTM International (ASTM) publishes the most widely accepted standards on brick. These standards are voluntary
consensus standards that are reviewed and updated periodically to contain the most recent information. All
have been through a thorough review process by a balanced committee of interested ASTM members classified
as producers, users and general interest. All of the model building codes in the United States reference ASTM
standards for brick.
Standards used in Canadian building codes are prepared by the Canadian Standards Association (CSA). The
process used to prepare and revise CSA standards is similar to ASTMs. The sole CSA standard for brick, A82
Fired Masonry Brick Made from Clay or Shale, is similar in content to the ASTM standards for face brick and
hollow brick. It also includes test methods.
This Technical Note identifies the standards for brick and the specific requirements for its various classifications.
Other Technical Notes in this series address the fundamentals of brick manufacturing and the proper selection of
brick.
SUMMARY OF RECOMMENDATIONS:
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 2 of 13
BRICK SPECIFICATIONS
Depending on its use, brick is covered by one of several specifications. See Table 1. Because firebox brick,
chemical resistant brick, sewer and manhole brick, and industrial floor brick are special uses, they will not be
addressed in this Technical Note.
Title of Specification ASTM Designation
1
CSA Designation
2
Building Brick C 62
Facing Brick C 216 A82
Hollow Brick C 652 A82
Thin Veneer Brick Units Made from Clay or Shale C 1088
Pedestrian and Light Traffic Paving Brick C 902
Heavy Vehicular Paving Brick C 1272
Ceramic Glazed Structural Clay Facing Tile, Facing
Brick, and Solid Masonry Units
C 126
Glazed Brick, Single Fired C 1405
Firebox Brick, Residential Fireplaces C 1261
Chemical-Resistant Masonry Units C 279
Sewer and Manhole Brick C 32
Industrial Floor Brick C 410
TABLE 1
Specifications for Brick
1. ASTM International, 100 Bar Harbor Drive, West Conshohocken, PA 19428.
2. Canadian Standards Association, 5060 Spectrum Way, Suite 100, Mississauga, Ontario, L4W 5N6 Canada.
Beginning with the 2007a edition of ASTM C 216, an appendix has been added. The appendix is designed to
explain the specification, noting subtleties and relationships that might not otherwise be clear. In many instances
the use of brick is similar to the title of its ASTM specification.
Facing Brick
Facing brick are intended for use in both structural and nonstructural masonry, including veneer, where
appearance is a requirement.
Hollow Brick
Hollow brick are used as either building or facing brick but have a greater void area. Most hollow brick are used
as facing brick in anchored veneer. Hollow brick with very large cores are used in reinforced brickwork and contain
steel reinforcement and grout.
Building Brick
Building brick are intended for use in both structural and nonstructural brickwork where appearance is not a
requirement. Building brick are typically used as a backing material.
Thin Brick
Thin veneer brick have normal face dimensions but a reduced thickness. They are used in adhered veneer
applications.
Paving Brick
Paving brick are intended for use as the wearing surface on clay paving systems. As such they are subject to
pedestrian and light or heavy vehicular traffic.
Glazed Brick
Glazed brick have a ceramic glaze finish fused to the brick body. The glaze can be applied before or after the firing
of the brick body. These brick may be used as structural or facing components in masonry.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 3 of 13
CLASSIFICATIONS
There are several classifications used in each standard. Classifications include grade, class, type, application and
use. The criteria for these classifications may include exposure or use conditions; appearance items; physical
properties needed for performance; tolerances on dimensions and distortion; chippage; and void area.
Brick qualify for a particular classification based on their properties after manufacturing. While most brick can be
manufactured to attain all the attributes desired by a user, certain attributes may be dictated by the production
method, durability classification or appearance classification designated by the user. For example, a molded brick
cannot be made to meet the classification for the tightest dimensional tolerances since the production method uses
a higher percentage of water that may result in greater shrinkage. Brick manufactured by the extrusion process
can be made to meet the classification for tight or loose dimensional tolerances.
When specifying brick each classification should be designated. Some ASTM brick specifications default to a
certain classification if it is not designated. The default classification may not be suitable for the intended use.
Table 2 contains a listing of the classifications in ASTM and CSA brick specifications.
Classification
Durability Appearance Void Area Use
ASTM Specification
C 62
Building Brick
Grade None None None
C 216
Facing Brick
Grade Type None None
C 652
Hollow Brick
Grade Type Class None
C 1088
Thin Veneer Brick
Grade Type None None
C 902
Pedestrian and Light Traffic Paving Brick
Class and Type Application None Type
C 1272
Heavy Vehicular Paving Brick
Type Application None Type
C 126
Ceramic Glazed Facing Brick
None Grade and Type None None
C 1405
Single Fired Glazed Brick
Class Grade and Type Division None
CSA Specification
A82
Fired Masonry Brick
Made from Clay or Shale
Grade Type None
1
None
TABLE 2
Classifications in Specifications for Brick
1. No classification given, but solid, cored and hollow brick are defined. See Void Area.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 4 of 13
Durability
Classification
More Severe
Exposure
Less Severe
Exposure
ASTM Specification
C 62 Building Brick Grade SW MW NW
C 216 Facing Brick Grade SW MW
C 652 Hollow Brick Grade SW MW
C 1088 Thin Veneer Brick Grade Exterior Interior
C 902 Pedestrian and Light Traffic Paving Brick
Class SX MX NX
Type I II III
C 1272 Heavy Vehicular Paving Brick Type F R
C 126 Ceramic Glazed Facing Brick None
C 1405 Single Fired Glazed Brick Class Exterior Interior
CSA Specification
A82 Fired Masonry Brick Made from Clay or Shale Grade Exterior (EG) Interior (IG)
TABLE 3
Durability Classifications
For durability classifications the letters S, M and N in C 62, C 216, C 652 and C 902 indicate the following
exposure conditions:
S indicates severe weathering.
M indicates moderate weathering.
N indicates negligible or no weathering.
Physical Property Requirements. The physical property requirements in most specifications are compressive
strength, water absorption and saturation coefficient. These properties must be determined in accordance with
ASTM C 67, Standard Methods of Sampling and Testing Brick and Structural Clay Tile [Ref. 1] or CSA A82 [Ref.
3]. The minimum compressive strength, maximum water absorption and maximum saturation coefficient are
used in combination to predict the durability of the bricks in use. The saturation coefficient, also referred to as the
C/B ratio, is the ratio of 24-hour cold water absorption to the five-hour boiling absorption. The physical property
requirements for each standard are listed in Table 4.
Some brick are durable but cannot be classified under the physical requirements shown in Table 4. Using
alternates and alternatives in the specifications allows brick that are known to perform well to meet the durability
requirement. A brick qualifying for a classification by an alternate or alternative does not signify that it is of a lower
quality.
The Absorption Alternate is found in ASTM C 62, C 216, C 652, C 1088, C 902 and C 1405. The Freezing
and Thawing Alternative is found in ASTM C 62, C 216, C 652, C 1088, C 902, C 1272 and C 1405. The Low
Weathering Index Alternative is found in ASTM C 62, C 216 and C 1088. CSA A82 includes a freeze-thaw test
as an alternative if the brick does not meet the physical property requirements. Other unit specifications include
alternates as well. These are discussed in the Additional Requirements section.
Absorption Alternate- The saturation coefficient requirement does not apply, provided the cold water absorption of
any single brick of a random sample of five brick does not exceed 8 percent.
Durability and Exposure
Since the environmental and service conditions that brick are subjected to vary, each brick specification classi-
fies brick for its specific durability. The classification is based on the severity of weather and the exposure of the
brick. The classification assigned to the brick is typically based on physical properties of the brick. See Technical
Note 9B for selection of the appropriate level of durability. The durability classifications for each specification are
listed in Table 3.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 5 of 13
Minimum
Compressive
Strength, Gross
Area
1
psi (MPa)
Maximum Cold
Water Absorption,
%
Maximum
Five-Hour Boiling
Absorption, %
Maximum Saturation
Coefficient
Minimum Breaking
Load, lb/in. (kN/mm)
Average
of 5 brick
Individual
Average
of 5 brick
Individual
Average
of 5 brick
Individual
Average
of 5 brick
Individual
Average
of 5 brick
Individual
ASTM Specification and Classification
C 62
Grade
SW
3000
(20.7)
2500
(17.2)
17.0 20.0 0.78 0.80
MW
2500
(17.2)
2200
(15.2)
22.0 25.0 0.88 0.90
NW
1500
(10.3)
1250
(8.6)
No limit No limit No limit No limit
C 216
Grade
SW
3000
(20.7)
2500
(17.2)
17.0 20.0 0.78 0.80
MW
2500
(17.2)
2200
(15.2)
22.0 25.0 0.88 0.90
C 652
Grade
SW
3000
(20.7)
2500
(17.2)
17.0 20.0 0.78 0.80
MW
2500
(17.2)
2200
(15.2)
22.0 25.0 0.88 0.90
C 1088
Grade
Ext. 17.0 20.0 0.78 0.80
Int. 22.0 25.0 0.88 0.90
C 902
Class
SX
8000
[4000]
2
(55.2)
[(27.6)]
2
7000
[3500]
2
(48.3)
[(24.1)]
2
8.0
[16.0]
2
11.0
[18.0]
2
0.78 0.80
MX
3000
(20.7)
2500
(17.2)
14.0 17.0 No limit No limit
NX
3000
(20.7)
2500
(17.2)
No limit No limit No limit No limit
C 1272
Type
F
10,000
(69.0)
8800
(60.7)
6.0 7.0
475
(83)
333
(58)
R
8000
(55.2)
7000
(48.3)
6.0 7.0
C 126
Coring
Vert.
3000
(20.7)
2500
(17.2)

Horiz.
2000
(13.8)
1500
(10.3)

C 1405
Class
Ext.
6000
(41.4)
5600
(34.8)
7.0 0.78 0.80
Int.
3000
(20.7)
2500
(17.2)

CSA Specification and Classification
A82
Ext.
3000
(20.7)
2500
(17.2)
8.0
3
17.0 0.78
3

Int.
2500
(17.2)
2200
(15.2)
22.0 25.0 0.88 0.90
TABLE 4
Physical Properties in Brick Specifications
1. Brick in bearing position or loaded in the same direction as in service.
2. Numbers in brackets are for molded brick and apply provided the requirements for saturation coefficient are met.
3. Either of these requirements must be met, not both.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 6 of 13
Freezing and Thawing Alternative- The requirements for five-hour boiling water absorption and saturation
coefficient do not apply, provided a sample of five brick, meeting the strength requirements, passes the freezing
and thawing test as described in the Rating section of the Freezing and Thawing test procedures of ASTM C 67
with a weight loss not greater than 0.5 percent in dry weight of any individual brick (for Grade SW). Unlike ASTM
C 67, CSA A 82 stipulates that brick must be kept in a frozen state during any interruption of the freeze-thaw test.
Low Weathering Index Alternative- If the brick are intended for use where the weathering index is less than 50
and have a minimum average compressive strength of 2500 psi (17.2 MPa), the requirements given for five-hour
boiling water absorption and for saturation coefficient shall not apply.
Consult the appropriate ASTM specification for specific alternates.
Appearance
Classification related to the appearance may include limits tolerances on dimensions, distortion, out-of-square
and chippage. The appearance classification is established on the size and precision attained in manufacturing.
The classifications for appearance of brick for each specification are listed in Table 5, and requirements for size
variation, distortion and chippage are listed in Table 6, Table 7 and Table 8, respectively. There are no color-related
tolerances in the ASTM standards for brick. Those are dictated by the sample panel or project specification.
Appearance
Classifications
More Stringent
Requirements
Less Stringent
Requirements
ASTM Specification
C 62 Building Brick None
C 216 Facing Brick Type FBX FBS FBA
C 652 Hollow Brick Type HBX HBS HBA HBB
C 1088 Thin Veneer Brick Type TBX TBS TBA
C 902 Pedestrian and Light Traffic
Paving Brick
Application PX PS PA
C 1272 Heavy Vehicular Paving Brick Application PX PS PA
C 126 Ceramic Glazed Facing Brick
Grade SS S
Type II I
C 1405 Single Fired Glazed Brick
Grade SS S
Type II I
CSA Specification
A82 Fired Masonry Brick Made from
Clay or Shale
Type X S A
TABLE 5
Appearance Classifications
For appearance classifications the letters X, S and A have the following meanings:
X indicates extreme or extra control in the criteria.
S indicates standard production.
A indicates architectural or aesthetic criteria that must be specified and in many specifications must be less
stringent than the S designation.
Dimensional Tolerances. Variations in raw materials and the manufacturing process will result in brick that vary
in size. Permitted size variation is based on the brick classification and the relative dimensional range measured.
These permitted variations in size are listed in Table 6A, Table 6B and Table 6C. The variation is plus or minus
from the specified dimension. Size variation becomes important when vertical alignment of brick (stack bond)
is used, when bands of brick from different production runs are combined, or when a short horizontal extent of
brickwork is constructed, such as between closely spaced window openings.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 7 of 13
Specified Dimension or
Average Brick Size
in Job Lot Sample, in. (mm)
Maximum Permissible Variation, in. (mm), plus or minus from:
Column A (for Specified
Dimension)
Column B (for Average Brick Size in Job
Lot Sample)
2
Type FBX Type FBS Type FBX
Type FBS
Smooth
3
Type FBS
Rough
4
3 (76) and under 1/16 (1.6) 3/32 (2.4) 1/16 (1.6) 1/16 (1.6) 3/32 (2.4)
Over 3 to 4 (76 to 102), inclusive 3/32 (2.4) 1/8 (3.2) 1/16 (1.6) 3/32 (2.4) 1/8 (3.2)
Over 4 to 6 (102 to 152), inclusive 1/8 (3.2) 3/16 (4.8) 3/32 (2.4) 3/32 (2.4) 3/16 (4.8)
Over 6 to 8 (152 to 203), inclusive 5/32 (4.0) 1/4 (6.4) 3/32 (2.4) 1/8 (3.2) 1/4 (6.4)
Over 8 to 12 (203 to 305), inclusive 7/32 (5.6) 5/16 (7.9) 1/8 (3.2) 3/16 (4.8) 5/16 (7.9)
Over 12 to 16 (305 to 406), inclusive 9/32 (7.1) 3/8 (9.5) 3/16 (4.8) 1/4 (6.4) 3/8 (9.5)
TABLE 6A
Dimensional Tolerances for ASTM C 216 and CSA A82
1
1. Dimensional tolerances for Type FBA and A in C 216 and A82,
respectively, shall be as specified by the purchaser, but not more
restrictive than Type FBS and S (Rough), respectively.
2. Lot size shall be determined by agreement between purchaser and
seller. If not specified, lot size shall be understood to include all
brick of one size and color in the job order.
3. Type FBS Smooth brick have relatively fine texture and smooth
edges, including wire cut surfaces. These definitions relate to
dimensional tolerances only.
4. Type FBS Rough bricks are molded brick or extruded brick with tex-
tured, rounded or tumbled edges or faces. These definitions apply
to dimensional tolerances only.
TABLE 6B
Dimensional Tolerances
ASTM
Specification
and
Classification
Maximum Permissible Variation, in. (mm), plus or minus
3 (76) and
under
Over 3 to 4
(102) inclusive
Over 4 to 6
(152) inclusive
Over 6 to 8
(204) inclusive
Over 8 to 12
(306) inclusive
Over 12 to 16
(408) inclusive
C 62 3/32 (2.4) 1/8 (3.2) 3/16 (4.8) 1/4 (6.4) 5/16 (8.0) 3/8 (9.5)
C 652
HBX 1/16 (1.6) 3/32 (2.4) 1/8 (3.2) 5/32 (4.0) 7/32 (5.6) 9/32 (7.1)
HBS
and
HBB
3/32 (2.4) 1/8 (3.2) 3/16 (4.8) 1/4 (6.4) 5/16 (7.9) 3/8 (9.5)
HBA As specified by the purchaser, but not more restrictive than HBS and HBB
C 1088
TBX 1/16 (1.6) 3/32 (2.4) 1/8 (3.2) 5/32 (4.0) 7/32 (5.6) 9/32 (7.2)
TBS 3/32 (2.4) 1/8 (3.3) 3/16 (4.8) 1/4 (6.4) 5/16 (8.0) 3/8 (9.5)
TBA As specified by the purchaser
C 126 See ASTM C 126
C 902
and
C 1272
PX 1/16 (1.6) 3/32 (2.4) 1/8 (3.2) 7/32 (5.6)
PS 1/8 (3.2) 3/16 (4.8) 1/4 (6.4) 5/16 (8.0)
PA No limit No limit No limit No limit
Specified Dimension
or Average Brick Size
in Job Lot Sample, in. (mm)
Maximum Permissible Variation in Dimensions, in. (mm) plus or minus from:
Column A (for Specified Dimension)
Column B (for Average Brick Size
in Job Lot Sample)
1
Grade S Grade SS Grade S Grade SS
3 (76) and under 1/16 (1.6) 1/16 (1.6) 1/16 (1.6) 1/16 (1.6)
Over 3 to 4 (76-102), inclusive 3/32 (2.4) 1/16 (1.6) 1/16 (1.6) 1/16 (1.6)
Over 4 to 6 (102-152), inclusive 1/8 (3.2) 1/16 (1.6) 3/32 (2.4) 1/16 (1.6)
Over 6 to 8 (152-203), inclusive 5/32 (4.0) 1/16 (1.6) 3/32 (2.4) 1/16 (1.6)
Over 8 to 12 (203-305), inclusive 7/32 (5.6) 1/16 (1.6) 1/8 (3.2) 1/16 (1.6)
Over 12 to 16 (305-406), inclusive 9/32 (7.1) 1/16 (1.6) 3/16 (4.8) 1/16 (1.6)
TABLE 6C
Dimensional Tolerances for ASTM C 1405
1. Lot size shall be determined by agreement between purchaser and seller. If not specified, lot size shall be understood to include all brick of
one size and color in the job order.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 8 of 13
Distortion. Permitted distortion, or warpage, of brick is
listed in Table 7. The amount of distortion is based on
the brick specification and face dimension. Distortion
may be convex or concave and may be in the plane of
the wall or perpendicular to it, as illustrated in Figure 1.
Other terms for distortion are bowed or banana
brick. A brick that is over the distortion limitations is
difficult to lay and is easily noticeable in the brickwork.
Maximum Permissible Distortion, in. (mm)
8 (204) and under
Over 8 to 12 (306),
inclusive
Over 12 to 16 (408),
inclusive
ASTM Specification and Classification
C 62 No limit No limit No limit
C 216
FBX 1/16 (1.6) 3/32 (2.4) 1/8 (3.2)
FBS 3/32 (2.4) 1/8 (3.2) 5/32 (4.0)
FBA As specified by the purchaser
C 652
HBX 1/16 (1.6) 3/32 (2.4) 1/8 (3.2)
HBS 3/32 (2.4) 1/8 (3.2) 5/32 (4.0)
HBA As specified by the purchaser
C 1088
TBX 1/16 (1.6) 3/32 (2.4) 1/8 (3.2)
TBS 3/32 (2.4) 1/8 (3.2) 5/32 (4.0)
TBA As specified by the purchaser
C 902 and
C1272
PX 1/16 (1.6) 3/32 (2.4) 1/8 (3.2)
PS 3/32 (2.4) 1/8 (3.3) 5/32 (4.0)
PA No limit.
C 126 Special requirements see ASTM C 126
C 1405
SS 1/16 (1.6) 3/32 (2.4) 3/32 (2.4)
S 1/16 (1.6) 3/32 (2.4) 1/8 (3.2)
CSA Specification and Classification
A82
X (1.5) (2.5) (3.0)
S (2.5) (3.0) (4.0)
A As specified by purchaser, but not more restrictive than Type S (Rough)
TABLE 7
Distortion Tolerances
Concave Surface Convex Surface
Average of
4 Corners
Concave Edge Convex Edge
Maximum
Maximum Maximum
Figure 1
Distortion Measurements
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 9 of 13
Chippage. Brick may be damaged or chipped during packaging, shipping or on the job site. Limitations to the
size and number of chips on individual brick are listed in Table 8. The amount of chippage is based upon the brick
specification and classification.
A delivery of brick may contain up to 5 percent broken brick or brick chipped beyond the limits in Table 8. The
chippage requirements in Table 8 are based on the remaining 95 percent of the shipment. The chips are measured
from an edge or a corner, and the total length of these chips may not be greater than 10 percent of the perimeter
of the face of the brick. Chips are more noticeable on brick that have a surface color different from the body of the
brick. Chips on through-body color brick are less noticeable.
Specification and Type or Application
Percent
Allowed
Chippage in From
Percent
Allowed
Chippage in From
ASTM
C 216
ASTM
C 652
ASTM
C 1088
ASTM
C 902
ASTM
C 1272
CSA
A82
Edge,
in. (mm)
Corner,
in. (mm)
Edge,
in. (mm)
Corner,
in. (mm)
FBX HBX TBX X
95 to
100%
0 to 1/8
(0 to 3.2)
0 to 1/4
(0 to 6.4)
5% or
less
1/8 to 1/4
(3.2 to 6.4)
1/4 to 3/8
(6.4 to 9.5)
FBS
2
HBS
2
TBS
2
S
2
90 to
100%
0 to 1/4
(0 to 6.4)
0 to 3/8
(0 to 9.5)
10% or
less
1/4 to 5/16
(6.4 to 7.9)
3/8 to 1/2
(9.5 to
12.7)
FBS
3
HBS
3
TBS
3
S
3
85 to
100%
0 to 5/16
(0 to 7.9)
0 to 1/2
(0 to 12.7)
15% or
less
5/16 to 7/16
(7.9 to
11.1)
1/2 to 3/4
(12.7 to
9.1)
FBA
HBA
TBA PA PA
4
A As specified by the purchaser
5
HBB
PS
PS
100%
5/16
(7.9)
1/2
(12.7)

PX
PX 100%
1/4
(6.4)
3/8
(9.5)

1. There are no chippage requirements for C 62, C 126 or C 1405.
2. Extruded brick with unbroken natural die finish face and dry-pressed brick.
3. Extruded brick with finished face sanded, combed, scratched, scarified, or broken by mechanical means such as wire cutting or wire brush-
ing, and molded brick.
4. No limit.
5. Not more restrictive than FBS (Textured) in C 216 or HBS (altered).
TABLE 8
Maximum Permissible Range of Chippage
1
ADDITIONAL REQUIREMENTS
Void Area
In ASTM standards brick are generally classified as solid or hollow. A solid brick is defined as a unit whose net
cross-sectional area in every plane parallel to the bearing surface is 75 percent or more of its gross cross-sectional
area measured in the same plane. Thus, a solid brick has a maximum coring or void area of 25 percent. A hollow
brick is defined as a unit whose net cross-sectional area in every plane parallel to the bearing surface is less than
75 percent of its gross cross-sectional area measured in the same plane. A hollow brick has a minimum coring or
void area greater than 25 percent, and a maximum of 60 percent. Brick are cored or frogged at the option of the
manufacturer.
Cores. Holes in brick less than or equal to 1 square inches (9.68 cm
2
) in cross-sectional area, referred to as
cores, are used to aid in the manufacturing process and shipping of brick. The cores permit better utilization of raw
materials, create more uniform drying and firing of the brick, reduce the amount of fuel necessary to fire the brick
and reduce shipping costs by reducing weight. Additional advantages, such as aiding in mechanical bond in a wall,
easier laying of the brick, etc., also may result from brick manufactured with cores. Cores are found only in brick
manufactured by the extrusion or dry-press process. Limits to the amount of coring allowed in brick, the distance
from a core to a face, and web thickness where applicable are listed in Table 9.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 10 of 13
Cells. Cells are similar to cores except that a cell is larger in minimum dimension and has a cross-sectional
area greater than 1 square inches (9.68 cm2). Some requirements for cells are shown in Table 9. Additional
requirements for cells can be found in ASTM C 652, C 126 and C 1405 and CSA A82.
Frogs. Frogs are depressions in brick, usually located on one bed surface, and are included for the same reasons
as cores and cells. Frogs are found in brick manufactured by the molded process. Panel frogs are limited to a
specified depth and a specified distance from a face. Requirements for panel frogs are listed in Table 9. Deep
frogs are depressions deeper than 3/8 in. (10 mm), and must conform to the requirements for coring, hollow
spaces and void area of the applicable standard.
The Canadian Standards Association takes a different approach. CSA A82 defines a solid brick as one without
cores, cells or frogs deeper than 3/8 in. (10 mm); cored brick as those of which the net cross-sectional area in any
plane parallel to the bed face shall be at least 75 percent of the gross cross-sectional area measured in the same
plane; and hollow brick as brick whose net cross-sectional area in a plane parallel to the bed face is not less than
40 percent and not more than 75 percent of its gross cross-sectional area measured in the same plane. Further,
there is a required minimum dimension of 1/2 in. (6 mm) between cores; 1 in. (13 mm) between cells; and 3/4 in.
(19 mm) to an edge from a core, cell or frog.
ASTM
Specification
Void
Area, %
Cores Frogs Cells
a A b c E e f g h
in.
(mm),
min.
in.
(cm),
max.
in.
(mm),
min.
in.
(mm),
min.
in.
(cm),
max.
in.
(mm),
min.
in.
(mm),
min.
in.
(mm),
min.
in.
(mm),
min.
C 62 < 25
3/4
(19.1)

3/4
(19.1)
3/8
(9.5)
No Requirements for Cells
C 216 < 25
3/4
(19.1)

3/4
(19.1)
3/8
(9.5)
No Requirements for Cells
C 652
2
H40V
> 25,
40
5/8
(16)
1
(9.68)
5/8
(16)
3/8
(9.5)
< 1
(9.68)
3/4
(19.1)
3/4
(19.1)
1/2
(13)

H60V
3
> 40,
60
5/8
(16)
1
(9.68)
5/8
(16)
3/8
(9.5)
> 1
(9.68)
3/4
(19.1)
3/4
(19.1)
1/2
(13)

C 1088 No Requirements for Cores, Frogs or Cells


C 902 No Requirements for Cores, Frogs or Cells
C 1272 Cores and Cells Not Permitted
C 126
4

No Requirements
for Cores or Frogs
> 1
(9.68)
3/4
(19.1)
3/4
(19.1)
1/2
(13)
5
1/2
(13)
C 1405
2

Solid 25
3/4
(19.1)

3/4
(19.1)
3/8
(9.5)
No Requirements for Cells
H40V
> 25,
40
5/8
(16)
1
(9.68)
5/8
(16)
3/8
(9.5)
> 1
(9.68)
3/4
(19.1)
3/4
(19.1)
1/2
(13)

H60V
3
> 40,
60

5/8
(16)
1
(9.68)
5/8
(16)
3/8
(9.5)
1
(9.68)
3/4
(19.1)
3/4
(19.1)
1/2
(13)
g
e
f
h
E
A
a
c
b
1. Deep frogs shall meet coring requirements of the applicable specification (see ASTM C 62, C216, C 652 and C 1405).
2. Cored-shell and double-shell hollow brick shall meet additional coring requirements of applicable specification in ASTM C 652 and C 1405.
3. Based on 3 in. (76 mm) and 4 in. (102 mm) nominal width (for larger dimensions see C 652 and C 1405).
4. Cells shall meet additional requirements of ASTM C 126.
5. Web thickness in cored brick shall meet additional requirements of ASTM C 126.
TABLE 9
Requirements for Void Areas
1
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 11 of 13
Efflorescence
Efflorescence is a crystalline deposit of water-soluble salts that can form on the surface of some brickwork. The
principal objection is an unsightly appearance, though it typically is not harmful to brick. The test for efflorescence
is described in ASTM C 67 and CSA A82. Brick tested under C 67 are given a rating of effloresced or not
effloresced. The specifier must invoke this part of the standard for the requirement of not effloresced to apply.
CSA A82 also includes a rating of slightly effloresced, and it is this rating that must be met if efflorescence testing
is invoked. Requirements on efflorescence are not included in C 62 and C 126.
Strength
Brickwork may be used as a structural material, so there may be instances when it is important to specify a
minimum compressive strength of the brick. This possibility is noted in ASTM C 62, C 216, C 652 and C 1405.
Most brick have compressive strengths considerably higher than the minimum compressive strengths required for
durability and abrasion resistance.
Initial Rate of Absorption
The initial rate of absorption (IRA) is a measure of how quickly the brick will remove water from mortar spread on
it. IRA is not a qualifying property or condition of brick in the ASTM or CSA specifications. IRA values may be of
interest when selecting mortar and in use of the brick on the jobsite. If the purchaser wishes to learn the IRA of the
brick, the IRA test must be requested. Initial rate of absorption information is included in ASTM C 62, C 216, C 652
and C 1405.
Sampling and Testing
All brick under ASTM specifications are sampled and tested in accordance with ASTM C 67. The purchaser
designates the place of selection of the brick for testing when the order is placed. Brick for efflorescence testing
must be sampled at the point of manufacturer. This is because the brick may be contaminated by efflorescing
materials after leaving the brick plant. Brick are sampled and tested for compliance to their specification prior
to use. ASTM C 126 and C 1405 include additional tests for properties of the glaze. These are described in the
following section on Glazed Brick.
CSA A82 includes sampling and test methods as part of the standard.
Facing Brick, ASTM C 216 and CSA A82
An additional tolerance is found in the ASTM standard for solid facing brick specification and in CSA A82. The
amount that the exposed face of a brick can be out-of-square is limited. This is more critical as brick height
increases. The maximum permitted dimension for out-of-square of the exposed face of the brick in C 216 is 1/8 in.
(3.2 mm) for Type FBS brick and 3/32 in. (2.4 mm) for Type FBX brick. Tolerances on out-of-square for Type FBA
brick shall be specified by the purchaser.
CSA A82 contains similar requirements: Type S of 3.0 mm and Type X of 2.5 mm. Tolerances on out-of-square for
Type A brick shall be specified be specified but shall not be more restrictive than for Type S (Rough) brick.
Paving Brick, ASTM C 902 and C 1272
Not only must paving brick conform to the physical properties required in Table 4, but they also must have
additional alternatives for durability and must meet requirements for abrasion resistance.
Alternative Performance Requirements. If information on the performance of brick in a pavement subject to
similar exposure and traffic conditions is documented, then the physical property requirements in Table 4 may be
waived. This is identified as the Performance Alternative.
An optional test for the freeze and thaw test is ASTM C 88 Test Method for Soundness of Aggregates by Use of
Sodium Sulfate. The sulfate soundness test, like the freeze and thaw test, is not required unless the paving brick
do not meet the saturation coefficient and absorption requirements.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 12 of 13
Abrasion Resistance. Since paving brick are used in a horizontal application and are exposed to traffic, they
must meet a specified abrasion limit. Pedestrian and light traffic paving brick (C 902) are assigned a Type by the
traffic or abrasion expected. Type I pavers are exposed to extensive abrasion, such as driveways or public entries.
Type II pavers are exposed to high levels of pedestrian traffic, such as in stores, restaurant floors or exterior
walkways. Type III pavers are exposed to light pedestrian traffic, such as floors or patios in homes.
Heavy vehicular paving brick (C 1272) are assigned a Type depending on their intended installation. Type R
pavers are intended to be set in a mortar or asphalt setting bed supported by an adequate base. Type R pavers
must be at least 2 in. (57.2 mm) thick. Type F pavers are intended to be set in a sand setting bed, with sand
joints, and supported by an adequate base. Type F pavers must be at least 2 in. (66.7 mm) thick. The abrasion
requirements are the same for Type F and Type R pavers.
The abrasion resistance index can be determined in either of two ways: 1) by dividing the absorption by the
compressive strength and multiplying by 100, or 2) by determining the volume abrasion loss in accordance with
ASTM C 418 Test Method for Abrasion Resistance of Concrete by Sandblasting. The abrasion requirements are
listed in Table 10.
ASTM Specification Traffic Type Abrasion Index, Max.
Volume Abrasion Loss,
Max. (cm
3
/cm
2
)
C 902 Pedestrian and
Light Traffic Paving Brick
Type I 0.11 1.7
Type II 0.25 2.7
Type III 0.50 4.0
C 1272 Heavy Vehicular Paving Brick Types F and R 0.11 1.7
TABLE 10
Abrasion Resistance Requirements for Pavers
Glazed Brick, ASTM C 126 and C 1405
ASTM C 126 and C 1405 are specifications for glazed brick and contain requirements for properties of the glaze.
These properties include imperviousness, opacity, resistance to fading, resistance to crazing, flame spread, fuel
contribution and smoke density, toxic fumes, hardness, and abrasion resistance.
SUMMARY
This Technical Note identifies brick specifications used in the United States and Canada. Classification
designations for each brick specification and the criteria used to qualify for them are explained. Potential
performance issues can be minimized by designating the proper brick specification and applicable classifications
based on the environmental and service conditions of the project.
The information and suggestions contained in this Technical Note are based on the available
data and the experience of engineering staff and members of the Brick Industry Association.
The information contained herein should be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of the
information discussed in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
www.gobrick.com | Brick Industry Association | TN 9A | Specifications for and Classification of Brick | Page 13 of 13
REFERENCES
1. Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA 2006:
Volume 04.02 Concrete and Aggregate
ASTM C 88 Test Method for Soundness of Aggregates by Use of Sodium Sulfate
ASTM C 418 Test Method for Abrasion Resistance of Concrete by Sandblasting
Volume 4.05 Chemical Resistant Nonmetallic Materials; Vitrified Clay Pipe; Concrete Pipe; Fiber-Reinforced
Cement Products; Mortars and Grouts; Masonry; Precast Concrete
C 32, Standard Specification for Sewer and Manhole Brick (Made From Clay or Shale)
C 62, Standard Specification for Building Brick (Solid Masonry Units Made From Clay or Shale)
C 67, Standard Test Methods for Sampling and Testing Brick and Structural Clay Tile
C 126, Standard Specification for Ceramic Glazed Structural Clay Facing Tile, Facing Brick, and Solid
Masonry Units
C 216, Standard Specification for Facing Brick (Solid Masonry Units Made from Clay or Shale)
C 279, Standard Specification for Chemical-Resistant Masonry Units
C 410, Standard Specification for Industrial Floor Brick
C 652, Standard Specification for Hollow Brick (Hollow Masonry Units Made from Clay or Shale)
C 902, Standard Specification for Pedestrian and Light Traffic Paving Brick
C 1088, Standard Specification for Thin Veneer Brick Units Made from Clay or Shale
C 1261, Standard Specification for Firebox Brick for Residential Fireplaces
C 1272, Standard Specification for Heavy Vehicular Paving Brick
C 1405, Standard Specification for Glazed Brick (Single Fired, Brick Units)
2. Borchelt, J. G., Danforth, L.. Jr., and Hunsicker, R., Specifying Brick: Getting what you want for appearance and
function, The Construction Specifier, Construction Specifications Institute, Alexandria, VA, January 2006, pp.
20-28.
3. CSA A82, Fired Masonry Brick Made from Clay or Shale, Canadian Standards Association, Mississauga,
Ontario, Canada, 2006.

Technical Notes 9B - Manufacturing, Classification, and Selection of Brick, Selection, Part 3
Revised December 2003
Abstract: This Technical Notes addresses the selection of brick. Evaluation of the properties and applications of brick
determines the durability, appearance, and impression of a project. Information is provided regarding aesthetics, cost
and availability.
Key Words: abrasion, absorption, aesthetics, availability, brick, color, compressive strength, cost, durability, size,
texture.
INTRODUCTION
The selection of brick is important in that it determines a project's durability and appearance, and results in a lasting
impression. It is necessary to identify which qualities and properties of brick are appropriate to consider in selecting a
brick. Brick with a wide variety of strength, color, texture, size, shape and cost are available. The owner or designer
must decide which characteristics of brick are most critical. This selection process can dictate the success of any
project.
This Technical Notes addresses the properties and characteristics which must be considered in the selection of the
appropriate brick for a project. Other Technical Notes in this series provide the fundamentals of brick manufacturing and
classification of brick.
GENERAL
Brick selection is based on a number of factors. Not only are aesthetics and durability important, but strength,
absorption, availability and cost are important to the owner, designers and contractors. The selection process can be
difficult since each group is trying to satisfy different requirements. Typically, the final selection is based on a
compromise from all parties involved.
Aesthetics
The use of brick as a building material dates back centuries. Because of brick's enduring qualities and limitless
appearances, designers can satisfy their creative styles with brick. Brick is readily available in many sizes, colors,
textures and shapes. These can be adapted to achieve virtually any desired style or expression.
A variety of common brick sizes are shown in Figure 1. Brick's small module can be related to the scale of the wall.
These sizes can be combined in such a way as to create different appearances and patterns. Not only does brick size
influence scale and appearance, but the size of brick influences wall cost because larger units require fewer brick,
normally resulting in less labor. When specifying the size of units, dimensions should be listed in the following order:
thickness (width) by height by length.
9b http://www.gobrick.com/BIA/technotes/t9b.htm
1 of 8 9/13/2009 12:40 PM
Brick Sizes (Nominal Dimensions)
FIG. 1
Brick manufacturers also offer a wide variety of colors to choose from. Units whose colors range from reds and
burgundies to whites and buffs are manufactured today. Many manufacturers produce over 100 colors. Many of these
color variations are created during the firing process. Temperature variations and the order in which the units are
stacked in the kiln determine shades of light and dark. Ceramic glazes, slurries or sand coatings can be applied to the
surface to achieve colors not possible with some clays. The possibilities of using units of contrasting colors in bands or
other patterns are endless. Sample panels, or mockups, can aid in selecting the desired color by showing the finished
appearance.
Another aesthetic feature to consider when selecting brick is the texture. Textures on brick can be smooth, wirecut
(velour), stippled, tumbled, brushed, rolled, and more. The texture interacts with light and creates differing and
interesting shadows.
Unique design features can easily be achieved by using special brick shapes. Brick can be molded and formed into any
shape, from simple sloped sill shapes to fancy watertable brick. For most manufacturers, molded shapes are easier to
produce than extruded shapes, because the molded, or soft-mud process is more adaptable to making brick shapes than
the extruded process. Making very large shapes can be difficult in either process because of problems with proper
drying and firing.
Physical Properties
There are many physical properties which may influence the selection of brick. Some of these include durability,
absorption, compressive strength and abrasion resistance. This Technical Notes will provide a basic understanding of
these properties to aid in selection of the proper brick. Physical properties required for proper performance are given in
the appropriate American Society for Testing and Materials (ASTM) specification for brick.
Durability. Currently, there are two accepted methods for demonstrating durability under ASTM standards: 1)
durability as predicted by compressive strength, absorption, and saturation coefficient, or 2) durability as determined by
compressive strength and passing 50 cycles of the freeze and thaw test. Criteria in each ASTM specification determine
grade or class designations. Because of the varying climates and applications of brick, specific physical properties are
required. Brick are classified into these grades or classes according to their resistance to freezing when wet. Table 1
gives the recommended grade of facing, building and hollow brick, based on weathering index and exposure. Fig. 2
indicates the approximate weathering indices of areas across the U.S. Technical Notes 9A describes this in more detail.
Most manufacturers make brick to meet the designation for the most severe weathering exposure, SW or SX, so they
9b http://www.gobrick.com/BIA/technotes/t9b.htm
2 of 8 9/13/2009 12:40 PM
may ship brick to all parts of the country. Some manufacturers produce brick complying only with the designation for
moderate weathering, MW or MX. Grade NW or NX brick are typically confined to interior applications, or where they
are protected from water absorption and freezing. Brick manufacturers can furnish certification that their product will
meet a certain grade or class.

Weathering Indices in the United States
FIG. 2


Absorption. Absorption can be broken into two distinct categories - absorption and initial rate of absorption (IRA). Both
are important in selecting the appropriate brick. Absorption of a brick is expressed as a percentage, and defined as the
ratio of the weight of water that is taken up into its body divided by the dry weight of the unit. Water absorption is
measured in two ways: 1) submerging the test specimen in room temperature water for a period of 24 hours, and 2)
9b http://www.gobrick.com/BIA/technotes/t9b.htm
3 of 8 9/13/2009 12:40 PM
submerging the test specimen in boiling water for five hours. These are known as the 24 hour cold water absorption, and
the 5 hour boiling water absorption, respectively. These two are used to calculate the saturation coefficient by dividing
the 24 hour cold water absorption by the 5 hour boiling. The saturation coefficient is used to help predict durability.
The initial rate of absorption (IRA) or suction is the rate of how much water a brick draws (sucks) in during the first
minute after contact of the bed surface with water. The suction has a direct bearing on the bond between brick and
mortar. It has been shown by test results that when a brick has high suction (over 30 grams/min/30 in [30
grams/min./194 cm
2
]), a strong, watertight joint may not be achieved. Therefore, high suction brick should be wetted
prior (3 hrs to 24 hrs) to laying to reduce the suction and allow the brick's surface to dry. Very low suction brick should
be covered and kept dry on the jobsite. Brick manufacturers can furnish values of IRA and saturation coefficient of the
selected units. The material specifier or supplier should inform the mason contractor about the suction of the brick prior
to construction.
Compressive Strength. The strength of a unit is used to determine durability and also compressive strength of the
resulting brick masonry. Typically, most materials are judged on the basis of strength. However, it is important not to
sacrifice properties of durability and bond for higher compressive strengths. Most brick currently produced have
strengths ranging from 3,000 psi (20.7MPa) to over 20,000 psi (138 MPa), averaging around 10,000 psi (68.9 MPa).
Achieving sufficient compressive strength with brick is seldom a problem.
Abrasion Resistance. This property is important when brick is used as paving. The resistance to abrasion is affected
by the degree of firing and by the nature of the raw material. Abrasion resistance is predicted in two manners. It is
evaluated in terms of cold water absorption and compressive strength. These two properties produce an abrasion index
which is used to determine the type of traffic which is suitable for a particular brick. Alternately, volume loss is
determined by sand blasting the paver surface.
Application
A building must perform the functions for which it is designed. The materials selected for a project must also perform as
intended. The designer must consider all factors which a wall or material must withstand. Some of the more important
factors include moisture penetration, temperature variations and structural loads. No one standard assembly is suitable
for all localities, occupancies, or designs; therefore, the designer must evaluate each factor and its relative effect on the
selection of a material or assembly.
Moisture Penetration. The use of quality materials and workmanship is essential in obtaining a satisfactory degree of
water resistance. When water passes through brick masonry walls, it invariably does so through separations or cracks
between the brick units and the mortar. It is virtually impossible for significant amounts of water to pass directly through
a brick unit. Therefore, brick units which develop a complete bond with mortar offer the best moisture resistance. Brick
and mortar properties should be compared to provide compatible materials which result in more watertight walls.
Currently, there are no requirements for the degree of water resistance of a wall.
Temperature Variations. Brick must withstand daily temperature cycles and seasonal extremes (-30F to 120F [-34C
to 49C]) depending on location, throughout its life. Thermal expansion and contraction of brick is not critical to the
selection of brick, but it is important to designers and this movement should be provided for in design and construction.
Brick also withstands temperature extremes in fires. Since brick is a fired material, it will not burn and acts as an
excellent barrier to fire because it is non-combustible.
Structural Loads. Ability to withstand either gravity or lateral loads relies heavily on brick strength, mortar strength and
dimensions of the wall assembly. Compressive strength requirements found in the ASTM specifications for brick are
based on durability performance. Structural analysis may require a higher compressive strength in order to resist the
applied loads. Compressive strength of masonry may be a governing criterion in loadbearing or reinforced brick masonry
projects.
Cost
Material selection is often based on cost, usually initial cost only. Although initial cost is important, lifecycle cost is a
better tool for making critical decisions. When deciding between different materials, all costs involved including labor and
maintenance costs, future value and life expectancy should be considered.
The selling price of brick is governed by many factors, including manufacturing methods and appearance of the unit.
2
9b http://www.gobrick.com/BIA/technotes/t9b.htm
4 of 8 9/13/2009 12:40 PM
When considering different brick, one must take into account shipping costs. Since most prices quoted are plant prices,
distance between the manufacturing plant and the jobsite is a major determinant of these shipping costs. Brick
manufacturers and distributors can supply brick prices and shipping prices. Brick price is only one part of the in-place
costs. Labor and overhead costs are approximately twice the brick and mortar costs. Many of the Masonry Institutes
throughout the country provide cost comparisons between different materials.
Availability
The availability of brick fluctuates with the time of the year and current construction trends and demands. On the
average, brick production time runs about 5 days, from pugging of the clay to the finished, fired product. This can
change depending on many factors such as variations in raw materials, forming process, and kiln types. Many brick
manufacturers have stockpiles of brick, but usually only a small quantity of each brick type. This may satisfy smaller
jobs, but for large projects requiring large quantities of brick, a special production run must be made for the job. Most
manufacturers have a set schedule as to when they produce a certain brick shade. It is at this time that the size of the
run will be increased to accommodate the large order. It is wise to determine the brick's availability from the
manufacturer.
It is best to purchase all brick from the same production run because there are typically slight color variations between
runs. All manufacturers have quality controls to keep this at a minimum.

SUMMARY
This Technical Notes has described which characteristics of brick are important in selecting a particular unit. There is a
wide selection of brick from which to choose. Selecting the appropriate material is important to the project's longevity
and appearance.
The remainder of this Technical Notes is "Recommended Practices Relating to the Responsibilities and Relationships in
Brick Construction". This document, developed jointly by the Brick Industry Association and the Mason Contractors
Association of America, explains some potential problems that may occur during and after the selection process, and
how to avoid them.
The information and suggestions contained in this Technical Notes are based on the available data and the experience of
the technical staff of the Brick Industry Association. The information and recommendations contained herein must be
used along with good technical judgment and a basic understanding of the properties of brick masonry. Final decisions
on the use of the information discussed in this Technical Notes are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
More detailed information on subjects discussed here can be found in the following publications:
1.Brick Industry Association Technical Notes 7 Series - Water Resistance of Brick Masonry
2.ASTM Standard Specifications for Brick.

Most brick construction projects are completed with the result that the building owner, architect, general contractor,
mason contractor, brick distributor, and brick manufacturer are completely satisfied with the final product. On rare
occasions, however, a mistake is made or a misunderstanding occurs that spoils what would otherwise be a rewarding
and profitable experience for all concerned.
Recognizing this fact, representatives from the manufacturing, sales and distribution, and installation segments have
9b http://www.gobrick.com/BIA/technotes/t9b.htm
5 of 8 9/13/2009 12:40 PM
developed these Recommended Practices which identify the areas in which misunderstandings are most likely to
occur and suggest procedures to be followed that will minimize the effects when mistakes do occur.
The Recommended Practices Relating to the Responsibilities and Relationships in Brick Construction was developed
through the cooperative efforts of the Brick Industry Association and the Mason Contractors Association of America.
Draft copies of the complete document were distributed to other construction industry associations for review and,
where appropriate, their advice was included in these final Recommended Practices.
INTRODUCTION
The purpose of these recommended practices is to prevent misunderstandings that might result from improper sampling
procedures, ordering, or examination of the field work.
As in all business relationships, there are responsibilities among all parties involved - manufacturers, distributors, general
contractors, mason contractors, construction managers, architects, engineers, owners and/or their respective
representatives or agents - in producing an acceptable masonry project. It is to the mutual advantage of all concerned
that problems, when encountered, be identified and addressed in a timely manner.
Contract Allowances
The practice of using only dollar value allowances for brick in construction documents is not recommended because this
method does not provide sufficient information to make an informed bid. Items such as special shapes often are too
complicated to use an allowance. However, if an allowance is used, the following variables should be included: unit
specification (ASTM standard), grade, type and size (width by height by length). The construction documents should
clearly state whether taxes, delivery, handling, and/or installation are included in the allowance. In the initial
establishment of an allowance, the parties should take into consideration the extra cost of special shapes and any other
special units required by the project.
Ordering
All brick orders should be submitted in writing by the purchaser to the distributor or manufacturer, whichever is
appropriate. The order should include and clearly identify the following:

A.
B.
C.
D.
E.
Job name and type;
Location;
Owner;
Architect;
General contractor and/or mason contractor;

F. Material quantities, including types and quantities of special or non-standard items, should be accurately
determined so that the order may be shipped in its entirety. Brick should be described by specified dimensions
(width by height by length) rather than by generic or trade name;
G. Unit prices, including conditions such as escalation of prices, freight rates and terms;
H. Delivery schedules, including anticipated start date and quantity of each shipment;
I.
Other information pertinent to the order, such as a copy of that portion of the specifications which applies to the
brickwork.
If special shapes are required, detailed large-scale drawings should be supplied by the purchaser through appropriate
channels at the earliest possible time.
Most orders are processed through a chain of purchasing which begins with the signing of the owner-general contractor
agreement and ends with the receipt of an order by the manufacturer. Other parties may be involved in this process as
intermediaries or secondary parties, including, among others, the owner's representative, the general contractor, the
mason contractor and the distributor. Each party in the chain should endeavor to promptly process the order and give
approvals as necessary so as to cause minimal delays in the schedule of the project. Upon receipt of the order, the
manufacturer typically acknowledges the order and should promptly advise the parties through the chain of purchasing
about any unacceptable or impractical terms. The acknowledgement should then be thoroughly scrutinized by the
responsible parties.
It should also be understood by all parties that by the placement of a written order the purchaser incurs the specific
payment responsibility for all special and/or non-standard items.
9b http://www.gobrick.com/BIA/technotes/t9b.htm
6 of 8 9/13/2009 12:40 PM
Certificates and Testing
Contract documents may require a letter of certification from the manufacturer to verify that the quality and
characteristics of the brick meet ASTM standards. Test reports from an independent testing laboratory, supplied by the
manufacturer, should be considered current if they are 24 months old or less. The cost of such tests is borne by the
seller. If testing of the production run that is intended for shipment to the project is required, the cost of testing is
typically borne as follows: if the results of the tests show that the brick do not conform to the requirements of the
product specification, the cost is typically borne by the seller; if the results of the tests show that the brick conform to the
requirements of the products specification, the cost is typically borne by the purchaser. The cost of any additional testing
is typically borne by the purchaser.
All testing shall be done in accordance with ASTM procedures and specifications.
Selection and Sampling
Brick is subject to variations in color between production runs and occasionally within the same run. Modern
manufacturing processes encompass the use of automatic equipment, which may also result in minor differences in color
and texture of the brick.
The selection of the size, color, texture and type of brick is the responsibility of the owner and/or owner's representative.
Usually, small samples are used for the preliminary selection and may not exactly represent the complete range of colors
and textures encountered in production runs.
Sometimes, a small sample is sufficient for determining the final selection. However, when large quantities of brick are to
be erected, the prudent owner, general contractor, mason contractor, distributor and/or manufacturer should direct or
request that the final selection be made from a field panel (also known as a field sample or mock-up). A field panel is
typically constructed as a freestanding element that will later be torn down when the project is complete. Usually, a
quantity of brick equal to 100 modular-size brick (approximately 15 square feet) will be used for the construction of the
freestanding field panel.
If an owner or the owner's representative requires the field panel, the distributor or manufacturer may not have control
over the actual erection that is frequently performed by a mason contractor. The party or parties who have control over
the work of the mason contractor (either by direct contract or by other powers) should take appropriate action during the
erection of the field panel to assure that no additions or deletions are made to the brick supplied by the distributor and
manufacturer, unless written approval has been received from the manufacturer for such a change.
Field panels should be constructed from the production run that is intended for shipment to the project. In the event that
the field panel has to be constructed for inspection and final selection before the production run for that project, the
owner and the manufacturer should agree in writing upon such a use. The manufacturer may reserve the right to
resample from the actual run before shipment commences. The owner or owner's representative should inspect and
approve the new panel.
When the field panel has been formally approved, it is the manufacturer's responsibility to provide brick as represented in
the field panel. A strap or control sample may be retained at the plant.
Typically, the general contractor and mason contractor are responsible for preserving and maintaining the integrity of the
field panel which is considered the project standard for bond, mortar, workmanship and appearance and as the standard
for comparison until the masonry has been completed and accepted by the owner or the owner's representative. If the
owner or owner's representative elects not to have a field panel erected, the first 100 square feet of actual construction
shall serve as the field panel.
Inspection and Examination
The general contractor or mason contractor normally receives the brick when they are delivered to the job site. The
general contractor or mason contractor should properly protect the brick from the weather and damage. It is critical that
the contractor inspect the brick before they are placed in the wall. If there are any discrepancies, the manufacturer or
distributor should be notified immediately.
The owner or the owner's representative is responsible for acceptance of the work and, therefore, should inspect, as
necessary, while the work progresses. This is especially critical at the start of the project to ensure that the color,
texture and workmanship are representative of the field panel and are acceptable.
9b http://www.gobrick.com/BIA/technotes/t9b.htm
7 of 8 9/13/2009 12:40 PM
The selling party, whether the manufacturer or distributor should visit the job site, as necessary, and, in addition, should
be available for meetings and consultation in the event the owner or the owner's representative discovers a problem.
In the event the work does not meet with the approval of the owner or the owner's representative, the owner should
immediately notify the general contractor, and appropriate action should be taken to correct the problem. If necessary,
this may require that the work be stopped and that all interested parties meet to resolve the problem.
References
Brick Industry Association, 11490 Commerce Park Dr., Reston, VA 20191, (703) 620-0010, www.gobrick.com.
Mason Contractors Association of America, 33 S. Roselle Rd., Schaumburg, IL 60193, (800) 536-2225,
www.masonryshowcase.com.
9b http://www.gobrick.com/BIA/technotes/t9b.htm
8 of 8 9/13/2009 12:40 PM
2009 Brick Industry Association, Reston, Virginia Page 1 of 11
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
10
February
2009
Dimensioning and Estimating Brick
Masonry
Abstract: This Technical Note presents information for determining the basic layout of brick masonry walls, including both
structural and veneer applications. Modular and non-modular brick masonry is discussed, including overall dimensioning of
masonry walls using various brick unit sizes. Finally, guidelines are presented to aid the designer in estimating the amount of
materials needed for brick masonry.
Key Words: actual dimension, construction, estimating, modular masonry, nominal dimension, size, specified dimension.
Brick and Mortar Joint Sizes:
Specify brick using standardized nomenclature and
specified size (width by height by length)
For modular brick, specify mortar joint thicknesses such
that when added to the specified brick size, the intended
modular dimensions result
When possible, select brick size to minimize cutting
Bond Pattern:
Select one-half running bond for applications when brick
width is one-half of brick length; select one-third running
bond when brick width is one-third of brick length
Dimensioning:
When using modular brick sizes, use multiples of
brick dimension plus mortar joint to determine nominal
dimensions
For horizontal dimensions of elements longer than four
brick lengths, use nominal dimensions as intended
constructed dimensions
When nominal dimensions are used on plans but are
not intended to be used for construction, note plans
accordingly
Estimating:
Use wall area method and tables to determine number
of brick and quantity of mortar per wall area
Modify brick estimates for bond pattern, breakage and
waste
Modify mortar estimates for bond pattern, collar joints
and waste
Include partial brick in estimates to maintain bond at
corners
Determine approximate mortar material quantities based
on brick size and bond pattern
INTRODUCTION
Brick are made in a number of sizes and laid in a variety of patterns. Most patterns of brickwork will adhere
to a common module that facilitates the dimensioning of the brickwork and any masonry openings. Generally,
designers can minimize the number of cuts of whole brick by dimensioning to a module. Knowing the size of the
brick and bond pattern will allow an estimate of the number of brick and amount of mortar needed for the project.
This Technical Note presents information to help the designer to choose a brick size, lay out modular dimensions
using the chosen size, and develop a materials estimate for brick and mortar.
Metric Measurements
Throughout this Technical Note, dimensions are based on the inch-pound system with conversions given for
the metric system. The measurements and dimensions correspond to brick manufactured primarily in the United
States to a typical module of 4 in. (102 mm). Brick manufactured for projects requiring metric dimensions typically
conform to a module of 100 mm (3.94 in.). Although the principles presented here are the same for either system,
care should be used when using the conversions given here for metric modular units and construction.
BRICK SIZES
Brick is a building material with a human scale. Brick sizes have varied over the centuries but have always been
similar to present-day sizes. Some sizes were developed to meet specific design, production or construction
needs. For example, larger brick were developed to increase bricklaying economy, and thinner brick help conserve
resources.
SUMMARY OF RECOMMENDATIONS:
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 2 of 11
Brick Orientation
A brick has three dimensions: width (sometimes
referred to as thickness), height and length. Although
brick can be laid in six different orientations (see
Figure 1), these dimensions as referenced apply
to a brick laid as a stretcher. Height and length are
sometimes called face dimensions, because these are
the dimensions exposed when the brick is laid as a
stretcher.
Brick Dimensions
There are three different sets of dimensions used
with brick: nominal, specified and actual. Each must
be used with care and accuracy to avoid confusion
during design and construction.
Nominal dimensions. Nominal dimensions apply
to modular brick and are the result of the specified
dimension of the brick plus the intended thickness of
its intended mortar joint. Generally, these dimensions
will fall into round numbers to produce modules
of 4 in. or 8 in. for imperial units or 100 mm for SI units. They are also a quick way to refer to a given brick size
without having to include fractions.
Specified dimensions. Specified dimensions are the anticipated manufactured dimensions of the brick, without
consideration for mortar, which are to be used in project specifications and purchase orders. They are also used
by the structural engineer in rational design of brick masonry. In non-modular construction, only the specified
dimensions are used; thus the absence of corresponding nominal dimensions in Table 2.
Actual dimensions. Actual dimensions are the measurements of the brick as manufactured. Generally the actual
dimensions will be within a tolerance of the specified dimensions. The allowable tolerances are dependent upon
the type and size of the brick and are given within the applicable ASTM standard specifications, such as those in
ASTM C216, Standard Specification for Facing Brick and C652, Standard Specification for Hollow Brick [Ref. 1].
Bond Pattern
For most brick sizes, one-half running bond is the basic pattern when laying a wall or pavement; i.e.,
approximately half of the bricks length overlaps the brick below. This pattern is the most frequently used pattern
in homes, schools and offices. However, some sizes lend themselves best to other bond patterns. As an example,
a utility-sized brick has a nominal length three times its nominal thickness. At corners, where the thickness of the
wythe is exposed as the brickwork turns the corner, laying a one-half running bond with utility-sized brick would
require cutting at least one brick in every course to maintain bond around the corner. So for utility-sized brick, one-
third running bond is much easier to install. These two patterns, as well as some of the more historic patterns that
use headers to tie together multiple wythes of masonry, are presented in greater detail in Technical Note 30.
Modular and Non-Modular Brick
Modular brick are sized such that the specified dimension plus the intended mortar joint thickness equal a
modular dimension. Generally, modular dimensions are whole numbers without fractions that result in modules of
4 in. or 8 in. for imperial units or 100 mm for SI units. A modular brick has a set of nominal, specified and actual
dimensions as referenced above. A non-modular brick has a set of specified and actual dimensions but does not
have nominal dimensions.
Brick are available in many sizes and are referred to by many different names, depending on region. In addition,
the name of a brick and its size, whether modular or non-modular, can vary depending on the manufacturer.
Modular brick and their nominal and specified dimensions are shown in Table 1 and Figure 2. Non-modular brick
and their specified sizes are shown in Table 2 and Figure 3.
Stretcher
Sailor
Header
Rowlock
Stretcher Soldier
Rowlock
Exposed faces shaded.
Figure 1
Brick Positions in a Wall
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 3 of 11
2 / " - 3"
3
4
2 / " - 2 / "
5
8
3
4
9 / " - 9 / "
5
8
3
4
8"
2 / " - 2 / "
3
4
13
16
2 / "
1
4
3 / "- 3 / "
1
2
5
8
8 "
Standard
Engineer
Standard
3 / "- 3 / "
1
2
5
8
Closure
Standard
8"
3 / "- 3 / "
1
2
5
8
3 / "- 3 / "
1
2
5
8
2 / " - 3"
3
4
2 / "
3
4
7 / " - 8"
5
8
King Queen
3"
2 / " - 2 / "
5
8
3
4
8 / "
5
8
4"
16"
4"
4"
12"
4"
Utility
16"
8"
Meridian
16"
8"
16"
4"
16"
4"
8"
Double Through-
Wall Meridian
8"
8-in. Through-
Wall Meridian
6"
6-in. Through-
Wall Meridian
4"
Double
Meridian
3 / "
1
5
12"
2 / "
2
3
12"
2
12"
4"
Norman
4"
Engineer
Norman
Roman
4"
8"
4"
8"
4"
3 / "
1
5
4"
Closure
Modular
Modular
Engineer
Modular
2 / "
2
3
8"
4"
12"
8"
4"
6"
12"
4"
12"
6"
3 / "
1
5
8"
8"
4"
8"
4"
6"
16"
4"
2 / "
2
3
Figure 2
Modular Brick Sizes (Nominal Dimensions)
Figure 3
Non-modular Brick Sizes (Specified Dimensions)
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 4 of 11
TABLE 1
Modular Brick Sizes
Brick Designation
1
Nominal Dimensions, in. (mm) Joint
Thickness,
3

in. (mm)
Specified Dimensions,
4
in. (mm)
Vertical
Coursing
W H L W H L
Modular 4 (102) 2 (68) 8 (203)
(9.5)
(12.7)
3 (92)
3 (89)
2 (57)
7 (194)
7 (191)
3C = 8 in.
(203 mm)
Engineer Modular 4 (102) 3
1
5 (81) 8 (203)
(9.5)
(12.7)
3 (92)
3 (89)
2
13
16 (71)
2 (70)
7 (194)
7 (191)
5C = 16 in.
(406 mm)
Closure Modular 4 (102) 4 (102) 8 (203)
(9.5)
(12.7)
3 (92)
3 (89)
3 (92)
3 (89)
7 (194)
7 (191)
1C = 4 in.
(102 mm)

2
4 (102) 6 (152) 8 (203)
(9.5)
(12.7)
3 (92)
3 (89)
5 (143)
5 (140)
7 (194)
7 (191)
2C = 12 in.
(305 mm)

2
4 (102) 8 (203) 8 (203)
(9.5)
(12.7)
3 (92)
3 (89)
7 (194)
7 (191)
7 (194)
7 (191)
1C = 8 in.
(203 mm)
Roman 4 (102) 2 (51) 12 (305)
(9.5)
(12.7)
3 (92)
3 (89)
1 (41)
1 (38)
11 (295)
11 (292)
2C = 4 in.
(102 mm)
Norman 4 (102) 2 (68) 12 (305)
(9.5)
(12.7)
3 (92)
3 (89)
2 (57)
11 (295)
11 (292)
3C = 8 in.
(203 mm)
Engineer Norman 4 (102) 3
1
5 (81) 12 (305)
(9.5)
(12.7)
3 (92)
3 (89)
2
13
16 (71)
2 (70)
11 (295)
11 (292)
5C = 16 in.
(406 mm)
Utility 4 (102) 4 (102) 12 (305)
(9.5)
(12.7)
3 (92)
3 (89)
3 (92)
3 (89)
11 (295)
11 (292)
1C = 4 in.
(102 mm)

2
6 (152) 3
1
5 (81) 12 (305)
(9.5)
(12.7)
5 (143)
5 (140)
2
13
16 (71)
2 (70)
11 (295)
11 (292)
5C = 16 in.
(406 mm)

2
6 (152) 4 (102) 12 (305)
(9.5)
(12.7)
5 (143)
5 (140)
3 (92)
3 (89)
11 (295)
11 (292)
1C = 4 in.
(102 mm)

2
8 (203) 4 (102) 12 (305)
(9.5)
(12.7)
7 (194)
7 (191)
3 (92)
3 (89)
11 (295)
11 (292)
1C = 4 in.
(102 mm)

2
4 (102) 2 (68) 16 (406)
(9.5)
(12.7)
3 (92)
3 (89)
2 (57)
15 (397)
15 (394)
3C = 8 in.
(203 mm)
Meridian 4 (102) 4 (102) 16 (406)
(9.5)
(12.7)
3 (92)
3 (89)
3 (92)
3 (89)
15 (397)
15 (394)
1C = 4 in.
(102 mm)
Double Meridian 4 (102) 8 (203) 16 (406)
(9.5)
(12.7)
3 (92)
3 (89)
7 (194)
7 (191)
15 (397)
15 (394)
1C = 8 in.
(203 mm)
6-in. Through-Wall
Meridian
6 (152) 4 (102) 16 (406)
(9.5)
(12.7)
5 (143)
5 (140)
3 (92)
3 (89)
15 (397)
15 (394)
1C = 4 in.
(102 mm)
8-in. Through-Wall
Meridian
8 (203) 4 (102) 16 (406)
(9.5)
(12.7)
7 (194)
7 (191)
3 (92)
3 (89)
15 (397)
15 (394)
1C = 4 in.
(102 mm)
Double Through-
Wall Meridian
8 (203) 8 (203) 16 (406)
(9.5)
(12.7)
7 (194)
7 (191)
7 (194)
7 (191)
15 (397)
15 (394)
1C = 8 in.
(203 mm)
1. Some manufacturers may use a brick designation different from that shown.
2. No brick designation is provided due to inadequate consensus among manufacturers.
3. Common joint sizes used with length and width dimensions. Actual bed joint thicknesses vary based on vertical coursing and actual brick
height.
4. Specied dimensions may vary within this range from manufacturer to manufacturer.
Although a size not listed in Table 1 or Table 2 might be desired for a specific project, special sizes are typically
avoided where possible in order to not increase costs unnecessarily. The use of specified dimensions when
ordering and specifying brick is strongly recommended, since a brick name can vary from manufacturer to
manufacturer, and a non-modular brick will not have nominal dimensions. To avoid confusion, specify brick using
the stretcher position with width first, followed by height, then length. In other words, a modular brick would be
specified as 3 in. 2 in. 7 in. (92 mm 57 mm 194 mm).
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 5 of 11
MODULAR MASONRY
There are relationships between the width, height and
length of brick that were developed as brick masonry
construction began. The most common of these
dimensional relationships are:
two brick widths plus one mortar joint
equal one brick length, and
three brick heights plus two mortar joints
equal one brick length.
Use of these relationships allows corners and
openings in brick walls to be constructed with little
waste and limited cutting of brick. These relationships
allow rowlocks and headers to tie adjacent wythes
together and courses of brick in different orientations
to align vertically (see Photo 1). This has given
rise to the rich variety of detailing that is part of the
architectural vernacular of brickwork.
Because of greater ease in design and construction,
the vast majority of contemporary brickwork uses
modular-sized brick and modular dimensioning.
The most common modular dimension system for
brickwork utilizes a 4 in. (102 mm) grid. The 4 in.
(102 mm) grid is used in modular coordination
between brick and concrete masonry units [Ref. 1]
and fits the modular dimensions of other construction
materials.
TABLE 2
Non-Modular Brick Sizes
Brick Designation
1
Joint
Thickness,
3

in. (mm)
Specified Dimensions,
4
in. (mm)
Vertical
Coursing W H L
Queen
(9.5)
(12.7)
2 (70)
3 (76)
2 (70)
7 (194)
8 (203)
5C = 16 in.
(406 mm)
King
(9.5)
(12.7)
2 (70)
3 (76)
2 (67)
2 (70)
9 (244)
9 (248)
5C = 16 in.
(406 mm)

2 (9.5)
(12.7)
3 (76)
2 (67)
2 (70)
8 (219)
5C = 16 in.
(406 mm)
Standard
(9.5)
(12.7)
3 (92)
3 (89)
2 (57) 8 (203)
3C = 8 in.
(203 mm)
Engineer Standard
(9.5)
(12.7)
3 (92)
3 (89)
2
13
16 (71)
2 (70)
8 (203)
5C = 16 in.
(406 mm)
Closure Standard
(9.5)
(12.7)
3 (92)
3 (89)
3 (92)
3 (89)
8 (203)
1C = 4 in.
(102 mm)
1. Some manufacturers may use a brick designation different from that shown.
2. No brick designation is provided due to inadequate consensus among manufacturers.
3. Common joint sizes used with length and width dimensions. Actual bed joint thicknesses vary based on vertical coursing
and actual brick height.
4. Specied dimensions may vary within this range from manufacturer to manufacturer.
Photo 1
Dimensional Relationships
A
r
c
h
i
t
e
c
t
:

D
i
M
e
l
l
a

S
h
a
f
f
e
r
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 6 of 11
Modular dimensions are sometimes called nominal dimensions, because they represent round numbers without
accounting for the fractions of an inch represented by mortar joint thicknesses. For masonry elements, the
relationship between modular dimensions and the actual dimensions constructed in the field can depend upon
the overall length of the masonry element. For longer masonry wall lengths made of modular-sized brick and
about four or more brick lengths long, the actual constructed length of the element often will be the modular
dimension. This is possible because during construction, the mason typically will adjust the horizontal layout
of the brick to allow slightly larger or smaller head joints so that the brickwork meets the required dimension.
For shorter masonry wall lengths made of modular-sized brick and less than about four brick lengths long, the
designer may want to consider the specified dimension of the brick and joint thickness when dimensioning the
wall. This is because the amount of adjustment necessary to the thickness of head joints between brick will be
larger. Additionally, the mason will adjust the number of courses and the bed joint thicknesses in order to meet
fixed vertical dimensions. When the completed elevation is viewed, any slight deviation in mortar joint width or the
number of courses generally is not obvious in the brickwork.
Overall Dimensioning
The choice of whether nominal or specified dimensions are to be used on drawings is often determined by
the type of information that the drawing provides. For drawings that cover large areas, such as elevations and
floor plans, use of nominal dimensions is recommended. The overall intent and appearance of the project can
be presented without the precision of specified dimensions. When nominal dimensions are used on plans, the
drawings must be clearly noted to advise the mason of the intended actual size of the completed masonry
elements.
For drawings that provide specific information to other trades, those that coordinate the installation of materials,
and for shop drawings, the use of specified dimensions is recommended. An easy manner to remember this is to
use nominal dimensions for drawings in which the scale is smaller than in. per foot. Use specified dimensions
for drawings shown in in. per foot and larger, Of course Computer Aided Drafting (CAD) and Building
Information Modeling (BIM) programs often have the specified dimensions of the brick and mortar joint as input
options. Thus, at the designers discretion, specified dimensions that utilize fractions can be used throughout the
drawings to indicate the desired constructed dimensions of the brickwork. However, doing so involves fractions
and may complicate the dimensioning process.
Non-Modular Horizontal Dimensioning. Non-modular brick by definition do not conform to a 4 in. (203 mm)
module. However, all non-modular brick of a certain size create a module equal to the sum of one brick length
and one mortar joint width. This module can be used to establish modular dimensioning for the brickwork in a
fashion similar to that used for modular brick. Non-modular brick that are approximately three times as long as
they are wide are usually laid in one-third running bond. When laid in one-half running bond, brick near wall ends
and openings must usually be cut to maintain the bond.
Vertical Dimensioning. The vertical coursing of both modular and non-modular sized brick is similar. A certain
number of courses will correspond to 4, 8, 12 or 16 in. (102, 203, 305 or 406 mm) in height. This dimension
establishes the vertical modular grid used on the brickwork. For example, for a non-modular engineer standard
brick, a vertical grid of 16 in. (406 mm) is used since five courses of brick equal 16 in. (406 mm). For a wall
constructed of modular brick, a vertical grid is established by three courses (three brick and three mortar joints)
equaling 8 in. (203 mm). Table 3 gives the vertical dimensions for numbers of courses (stretcher or header
positions) and corresponding mortar joints using various sized brick, rounded to the nearest
1
16 in.
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 7 of 11
TABLE 3
Vertical Coursing
No. of
Courses
Vertical Coursing of Unit
2C = 4 in.
(102 mm)
3C = 8 in.
(203 mm)
5C = 16 in.
(406 mm)
1C = 4 in.
(102 mm)
ft in. m ft in. m ft in. m ft in. m
1 0 2 0.051 0 2
11

16
0.068 0 3
3

16
0.081 0 4 0.102
2 0 4 0.102 0 5
5
16 0.135 0 6 0.163 0 8 0.203
3 0 6 0.152 0 8 0.203 0 9 0.244 1 0 0.305
4 0 8 0.203 0 10
11
16 0.271 1
13
16 0.325 1 4 0.406
5 0 10 0.254 1 1
5
16 0.339 1 4 0.406 1 8 0.508
6 1 0 0.305 1 4 0.406 1 7
3
16 0.488 2 0 0.61
7 1 2 0.356 1 6
11
16 0.474 1 10 0.569 2 4 0.711
8 1 4 0.406 1 9
5
16 0.542 2 1 0.65 2 8 0.813
9 1 6 0.457 2 0 0.61 2 4
13
16 0.732 3 0 0.914
10 1 8 0.508 2 2
11
16 0.677 2 8 0.813 3 4 1.02
11 1 10 0.559 2 5
5
16 0.745 2 11
3
16 0.894 3 8 1.12
12 2 0 0.61 2 8 0.813 3 2 0.975 4 0 1.22
13 2 2 0.66 2 10
11
16 0.881 3 5 1.06 4 4 1.32
14 2 4 0.711 3 1
5
16 0.948 3 8
13
16 1.14 4 8 1.42
15 2 6 0.762 3 4 1.02 4 0 1.22 5 0 1.52
16 2 8 0.813 3 6
11
16 1.08 4 3
3
16 1.3 5 4 1.63
17 2 10 0.864 3 9
5
16 1.15 4 6 1.38 5 8 1.78
18 3 0 0.914 4 0 1.22 4 9 1.46 6 0 1.83
19 3 2 0.965 4 2
11
16 1.29 5
13
16 1.54 6 4 1.93
20 3 4 1.02 4 5
5
16 1.36 5 4 1.63 6 8 2.03
21 3 6 1.07 4 8 1.42 5 7
3
16 1.71 7 0 2.13
22 3 8 1.12 4 10
11
16 1.49 5 10 1.79 7 4 2.24
23 3 10 1.17 5 1
5
16 1.56 6 1 1.87 7 8 2.34
24 4 0 1.22 5 4 1.63 6 4
13
16 1.95 8 0 2.44
25 4 2 1.27 5 6
11
16 1.69 6 8 2.03 8 4 2.54
26 4 4 1.32 5 9
5
16 1.76 6 11
3
16 2.11 8 8 2.64
27 4 6 1.37 6 0 1.83 7 2 2.2 9 0 2.74
28 4 8 1.42 6 2
11
16 1.9 7 5 2.28 9 4 2.85
29 4 10 1.47 6 5
5
16 1.96 7 8
13
16 2.36 9 8 2.95
30 5 0 1.52 6 8 2.03 8 0 2.44 10 0 3.05
31 5 2 1.58 6 10
11
16 2.1 8 3
3
16 2.52 10 4 3.15
32 5 4 1.63 7 1
5
16 2.17 8 6 2.6 10 8 3.25
33 5 6 1.68 7 4 2.24 8 9 2.68 11 0 3.35
34 5 8 1.73 7 6
11
16 2.3 9
13
16 2.76 11 4 3.45
35 5 10 1.78 7 9
5
16 2.37 9 4 2.85 11 8 3.56
36 6 0 1.83 8 0 2.44 9 7
3
16 2.93 12 0 3.66
37 6 2 1.88 8 2
11
16 2.51 9 10 3.01 12 4 3.76
38 6 4 1.93 8 5
5
16 2.57 10 1 3.09 12 8 3.86
39 6 6 1.98 8 8 2.64 10 4
13
16 3.17 13 0 3.96
40 6 8 2.03 8 10
11
16 2.71 10 8 3.25 13 4 4.06
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 8 of 11
2-8 (813 mm) M.O.
Approx. 2-8 (822 mm) actual
4 Nom. Brick Lengths + 1 Joint
3
8 /
Modular Brick,
inch (9.5 mm) Joints
3
8 /
5-4 (1.63 m) M.O.
Approx. 5-4 / (1.64 m) actual
24 Nom. Brick Courses + 1 Joint
3
8
Figure 4
Example of Determining Dimensions
for a Masonry Opening
Masonry Openings
The edges of masonry openings are defined by brick
units rather than mortar joints. Vertical dimensions
are based on the number of courses plus an extra
bed joint thickness. Figure 4 shows an example of
dimensions for a punched window opening for modular
sized brick units. Note that the height is for the
opening before the installation of the sill and extends
up to the bottom of the brick above, not to the bottom
of the lintel supporting the brick.
ESTIMATING BRICK MASONRY
There are various methods to estimate material
quantities on a project. Hand calculations and
computer programs have been used depending on
the complexity of the building. Because of its simplicity
and accuracy, the most widely used estimating
procedure is the wall-area method. It consists simply
of multiplying the net wall area (gross areas less areas
of openings) by known quantities of material required
per square foot (square meter).
Determining the area of brick and mortar within each
unit area of wall depends on both brick size and joint
width. For non-modular masonry, both dimensions
must be known to make accurate estimates. In
modular masonry, mortar joint sizes are dictated
by the size of the brick, simplifying the estimating
process.
TABLE 3
Vertical Coursing
No. of
Courses
Vertical Coursing of Unit
2C = 4 in.
(102 mm)
3C = 8 in.
(203 mm)
5C = 16 in.
(406 mm)
1C = 4 in.
(102 mm)
ft in. m ft in. m ft in. m ft in. m
41 6 10 2.08 9 1
5
16 2.78 10 11
3
16 3.33 13 8 4.17
42 7 0 2.13 9 4 2.85 11 2 3.41 14 0 4.27
43 7 2 2.18 9 6
11
16 2.91 11 5 3.5 14 4 4.37
44 7 4 2.24 9 9
5
16 2.98 11 8
3
16 3.58 14 8 4.47
45 7 6 2.29 10 0 3.05 12 0 3.66 15 0 4.57
46 7 8 2.34 10 2
11
16 3.12 12 3
3
16 3.74 15 4 4.67
47 7 10 2.39 10 5
5
16 3.18 12 6 3.82 15 8 4.78
48 8 0 2.44 10 8 3.25 12 9 3.9 16 0 4.88
49 8 2 2.49 10 10
11
16 3.32 13
13
16 3.98 16 4 4.98
50 8 4 2.54 11 1
5
16 3.39 13 4 4.06 16 8 5.08
100 16 8 5.08 22 2
11
16 6.77 26 8 8.13 33 4 10.2
TABLE 3 (continued)
Vertical Coursing
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 9 of 11
Brick and Mortar Quantities
Masons frequently use a rule of thumb that eight bags of masonry cement will lay 1000 modular brick. This is a
very rough estimate and includes an unspecified amount of waste. Table 4 presents the estimated quantities of
brick and mortar (not including waste) required for brick masonry according to the size of brick used in the wall.
The mortar quantities are based on theoretical dimensions of the mortar in the wall. Estimates made using Table 4
should also include applicable correction factors listed in the Correction Factor section. For guidance on the
volume of each solid material required for a specific mortar type, refer to ASTM C270, Standard Specification for
Mortar for Unit Masonry [Ref. 1]. The commonly used rule of thumb is appropriate: 1 cubic foot (cubic meter) of
loose, damp sand will yield about one cubic foot (cubic meter) of mortar.
TABLE 4
Quantity Estimates for Brick Masonry
MODULAR BRICK SIZES
Brick
Designation
Nominal Dimensions, in. (mm)
Joint
Thickness,
in. (mm)
Number of
Brick per
100 sq ft
(per 10 m)
Cubic Feet (Cubic
Meters) of Mortar
W H L
Per 100 sq ft
(10 m)
Per 1000
Brick
Modular 4 (102) 2 (68) 8 (203)
(9.5)
(12.7)
675 (727)
5.5 (1.7)
6.9 (2.1)
8.1 (0.23)
10.3 (0.29)
Engineer
Modular
4 (102) 3
1
5 (81) 8 (203)
(9.5)
(12.7)
563 (605)
4.8 (1.5)
6.1 (1.9)
8.5 (0.24)
10.8 (0.31)
Closure Modular 4 (102) 4 (102) 8 (203)
(9.5)
(12.7)
450 (484)
4.1 (1.3)
5.2 (1.6)
9.1 (0.26)
11.6 (0.33)
4 (102) 6 (152) 8 (203)
(9.5)
(12.7)
300 (323)
3.2 (0.98)
4.1 (1.3)
10.7 (0.30)
13.7 (0.39)
4 (102) 8 (203) 8 (203)
(9.5)
(12.7)
225 (242)
2.8 (0.84)
3.5 (1.1)
12.3 (0.35)
15.7 (0.44)
Roman 4 (102) 2 (51) 12 (305)
(9.5)
(12.7)
600 (646)
6.4 (2.0)
8.2 (2.5)
10.7 (0.30)
13.7 (0.39)
Norman 4 (102) 2 (68) 12 (305)
(9.5)
(12.7)
450 (484)
5.1 (1.5)
6.5 (2.0)
11.2 (0.32)
14.3 (0.41)
Engineer
Norman
4 (102) 3
1
5 (81) 12 (305)
(9.5)
(12.7)
375 (404)
4.4 (1.3)
5.6 (1.7)
11.7 (0.33)
14.9 (0.42)
Utility 4 (102) 4 (102) 12 (305)
(9.5)
(12.7)
300 (323)
3.7 (1.1)
4.7 (1.4)
12.3 (0.35)
15.7 (0.44)
6 (152) 3
1
5 (81) 12 (305)
(9.5)
(12.7)
375 (404)
6.8 (2.1)
8.8 (2.7)
18.1 (0.51)
23.4 (0.66)
6 (152) 4 (102) 12 (305)
(9.5)
(12.7)
300 (323)
5.7 (1.7)
7.4 (2.3)
19.1 (0.54)
24.7 (0.70)
8 (203) 4 (102) 12 (305)
(9.5)
(12.7)
300 (323)
7.8 (2.4)
10.1 (3.1)
25.9 (0.73)
33.6 (0.95)
4 (102) 2 (68) 16 (406)
(9.5)
(12.7)
338 (363)
4.9 (1.6)
6.5 (2.1)
14.5 (4.7)
19.2 (6.2)
Meridian 4 (102) 4 (102) 16 (406)
(9.5)
(12.7)
225 (242)
3.5 (1.1)
4.4 (1.4)
15.4 (0.44)
19.7 (0.56)
Double Meridian 4 (102) 8 (203) 16 (406)
(9.5)
(12.7)
113 (121)
2.1 (0.64)
2.7 (0.82)
18.6 (0.53)
23.8 (0.67)
6-in. Through-
Wall Meridian
6 (152) 4 (102) 16 (406)
(9.5)
(12.7)
225 (242)
5.4 (1.6)
7.0 (2.1)
24.0 (0.68)
31.0 (0.88)
8-in. Through-
Wall Meridian
8 (203) 4 (102) 16 (406)
(9.5)
(12.7)
225 (242)
7.3 (2.2)
9.5 (2.9)
32.5 (0.92)
42.3 (1.2)
Double Through-
Wall Meridian
8 (203) 8 (203) 16 (406)
(9.5)
(12.7)
113 (121)
4.4 (1.3)
5.7 (1.8)
39.1 (1.1)
51.0 (1.4)
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 10 of 11
NON-MODULAR BRICK SIZES
Brick
Designation
Specified Dimensions, in. (mm)
Joint
Thickness,
in. (mm)
Number of
Brick per
100 sq ft
(per 10 m)
Cubic Feet (Cubic
Meters) of Mortar
W H L
Per 100 sq ft
(10 m)
Per 1000
Brick
Queen
2 (70)
3 (76)
2 (70)
7 (194)
8 (203)
(9.5)
(12.7)
550 (592)
6.7 (2.1)
7.4 (2.3)
12.2 (0.35)
13.5 (0.38)
King
2 (70)
3 (76)
2 (67)
2 (70)
9 (244)
9 (248)
(9.5)
(12.7)
455 (490)
6.5 (2.0)
7.3 (2.2)
14.2 (0.40)
16.0 (0.45)
3 (76)
2 (67)
2 (70)
8 (219)
(9.5)
(12.7)
512 (551)
6.6 (2.0)
8.2 (2.5)
13.0 (0.37)
16.1 (0.46)
Standard
3 (92)
3 (89)
2 (57) 8 (203)
(9.5)
(12.7)
655 (705)
9.5 (2.9)
11.0 (3.3)
14.5 (0.41)
17.8 (0.50)
Engineer
Standard
3 (92)
3 (89)
2
13
16 (71)
2 (70)
8 (203)
(9.5)
(12.7)
539 (581)
8.2 (2.5)
9.7 (3.0)
15.3 (0.43)
18.7 (0.53)
Closure
Standard
3 (92)
3 (89)
3 (92)
3 (89)
8 (203)
(9.5)
(12.7)
430 (463)
7.0 (2.2)
8.5 (2.6)
16.4 (0.46)
20.0 (0.57)
Collar Joints. For multi-wythe construction, where the vertical joint between wythes is designed to be mortared
solid, the values in Table 5 can be used to estimate the quantity of mortar within the collar joint.
TABLE 4 (continued)
Quantity Estimates for Brick Masonry
Correction Factors
Hollow Brick. The mortar quantities in Table 4 are based on fully bedded, solid masonry units (coring up to 25
percent of the bedded area). In veneer applications, hollow brick should also be laid in full mortar beds. Field
testing has demonstrated that a veneer constructed of hollow brick units with a nominal thickness of 3 to 4 in. (76
to 102 mm) and cores or cells between 25 and 35 percent of the bedded area and laid in a full mortar bed does
not significantly increase mortar usage compared to the same veneer constructed of solid brick units. Care should
be taken to avoid using excessively plastic mortar or placement methods that would force excessive amounts
of mortar into the cells or cores of the brick below. If these steps are taken, the estimates of Table 4 are valid for
most hollow brick veneer applications. For hollow units laid with face shell bedding (as typically done in structural
applications), the estimated quantities can be reduced by a percentage equal to the percentage of voids in the
brick. This reduction will typically be between 25 and 35 percent.
Bond Pattern. The values in Table 4 are based on running or stack bond patterns. For patterns that incorporate
headers, the correction factors in Table 6 can be applied. The factor is a net increase for the number of brick
and the mortar quantity, not including waste. For definitions of the patterns cited, refer to Technical Note 30. For
example, for a standard-size brick laid with a in. (9.5 mm) joint thickness in a common bond with full headers
every fifth course, the following estimates would apply:
Number of brick per 100 sq ft (9.30 m) of brickwork: 655 + (
1
5 655) = 786 brick
Cubic feet (0.028 m) of mortar per 1000 brick: 14.5 + (
1

15
14.5) = 15.5
Cubic Feet of Mortar per 100 sq ft (m per 10 m) of Wall
-in. (6.4 mm) joint -in. (9.5 mm) joint -in. (12.7 mm) joint -in. (19.1 mm) joint
2.08 (0.064) 3.13 (0.095) 4.17 (0.13) 6.25 (0.19)
TABLE 5
Mortar Quantities in Collar Joints
www.gobrick.com | Brick Industry Association | TN 10 | Dimensioning and Estimating Brick Masonry | Page 11 of 11
Brick Breakage and Waste. In the estimating procedure, determine the net quantities of all brick, including all
correction factors above, before adding any allowances for waste. Allowances for waste and breakage vary, but
as a general rule, at least 5 percent is added to the net brick quantities delivered to the jobsite. Particular job
conditions or experience may warrant using a higher percentage for waste.
Mortar Waste. In the estimating procedure, determine the net quantities of all materials, including all correction
factors above, before adding any allowances for waste. Allowances for waste vary, but as a general rule, add 15 to
25 percent to the net mortar quantities. Particular job conditions, or experience, may dictate different factors.
SUMMARY
This Technical Note provides a discussion of brick sizes and modular masonry construction and a discussion
of the basic layout of brick masonry walls. Methods are presented for estimating quantities of brick and mortar
materials for a chosen brick size, mortar joint size and bond pattern.
The information and suggestions contained in this Technical Note are based on the available data
and the experience of the engineering staff and members of the Brick Industry Association. The
information contained herein must be used in conjunction with good technical judgment and a
basic understanding of the properties of brick masonry. Final decisions on the use of the informa-
tion discussed in this Technical Note are not within the purview of the Brick Industry Association
and must rest with the project architect, engineer and owner.
REFERENCES
1. Annual Volume of ASTM Standards, ASTM International, West Conshohoken, PA, 2008.
Volume 04.05
C216 Standard Specification for Facing Brick
C270 Standard Specification for Mortar for Unit Masonry
C652 Standard Specification for Hollow Brick
Volume 04.11
E835/ Standard Guide for Modular Coordination of Clay and Concrete Masonry Units
E835M
TABLE 6
Estimate Correction Factors for Bond Patterns
Bond
Brick Correction
Factor
1
Mortar Correction
Factor
2
Common Bond with full headers every fifth course only
1
5
1
15
Common Bond with full headers every sixth course only
1
6
1
18
Common Bond with full headers every seventh course only
1
7
1
21
English Bond (full headers every second course)
1
6
Flemish Bond (alternate full headers and stretchers every course)
1
9
Cross Bond with Flemish headers every sixth course
1
18
1
54
Flemish Cross Bond (Flemish headers every second course)
1
6
1
18
Double-stretcher, garden wall bond
1
5
1
15
Triple-stretcher, garden wall bond
1
7
1
21
1. The net increase for brick units may be less than that given when multiple headers can be made from a single brick.
2. Correction factors are applicable only to brick with lengths equal to twice the depth and 2 to three times the height.

Technical Notes 11 - Guide Specifications for Brick Masonry, Part 1
Rev. [Dec. 1971] (Reissued August 2001)
INTRODUCTION
Numerous methods are being explored to reduce constantly rising building costs. One means in which many
segments of the construction industry believe holds promise of lowering these costs is the use of specific, definitive
and concise specifications. They must convey to the contractor the exact requirements of the project and be
organized to facilitate take-off and estimating. Many general contractors have testified that the use of such
specifications results in lower contract bids.
During recent years, organizations, such as the American Institute of Architects (AIA), Producers' Council (PC),
Associated General Contractors of America (AGC), and the Construction Specifications Institute (CSI), have made
the improvement of construction specifications one of their major activities.
In accordance with the work of these agencies, the guide specifications in this series of Technical Notes are
written to follow the CSI format insofar as possible.
Use of Standards. It is recommended that, where suitable standards exist, such as those developed by the
American Society for Testing and Materials (ASTM), American National Standards Institute (ANSI), American
Concrete Institute (ACI) and other similar nationally recognized organizations, they be used and included in the
project specifications by reference.
Use of Detailed Descriptive Requirements. While detailed descriptive requirements are generally necessary as a
means of specifying installation or workmanship, it is recommended that they be used only as a last resort in
specifying materials.
Use of Performance Specifications. Performance specifications are not, in general, considered suitable for
specifying architectural building products. It is recommended that, if performance specifications are used to specify
building materials, they should state results desired or properties desired, but not both.
Use of Trade Names. It is recommended that, if building products are specified by trade names, the "special
conditions" contain a clause providing that substitutes will be considered on a quality and price basis, and that the
phrase "or equal", frequently included in such specifications, be eliminated.
The following paragraph is suggested for substitutions:
Variation From Materials Specified: It is intended that materials or products specified by name of manufacturer,
brand, trade name or by catalog reference shall be the basis of the bid and furnished under the contract, unless
changed by mutual agreement. Where two or more materials are named, the choice of these shall be optional with
the contractor. Should the contractor wish to use any materials or products other than those specified, he shall so
state, naming the proposed substitutions and stating what difference, if any, will be made in the contract price for
such substitution should it be accented.
Use of Allowances. It is recommended that allowances be used only with discretion. In all cases of allowances,
there should be sufficient description to indicate to the contractor the extent of labor required to install the items for
which allowances are listed. Also, all allowances should be listed under special conditions or under a separate
section with cross references to the individual trade sections involved.
SPECIFICATIONS FOR STRUCTURAL CLAY PRODUCTS
http://www.gobrick.com/BIA/technotes/t11.htm
1 of 4 9/13/2009 12:42 PM
Standard specifications for the various types and grades of brick and tile have been developed by technical
committees of the American Society for Testing and Materials. Membership of these committees is balanced
among consumers, manufacturers and a general interest group made up of engineers, scientists, educators, testing
experts and representatives of research organizations. Because of this balance of committee membership, ASTM
specifications are widely accepted and it is recommended that the appropriate ASTM specifications be included by
reference in all specifications for solid brick, hollow brick, structural facing tile (glazed or unglazed) and structural
clay tile.
ASTM standards are under continuous review by the stands committees having jurisdiction over them. From time to
time these standards are revised as a result of new developments. The ASTM designation of a standard consists
of a letter and a number permanently assigned to the standard, a dash and a number indicating the year the
standard was approved: as for example, C 216-69 which designates the Standard Specifications for Facing Brick
approved in 1969. If the letter T follows the year designation, it indicates a tentative standard.
When ASTM specifications are included by reference in project specifications, the full designation, including the
year of approval, should be given, since, obviously, after a contract has been awarded, a revision of specifications
by ASTM does not alter the contract. Similarly, the dates of any other specifications or codes included by
reference should be given.
Solid Masonry Units. ASTM Specifications C 216, C 62, and C 126 cover solid building brick, facing brick and
ceramic glazed units made from clay and/or shale. Under these specifications, a solid masonry unit may be cored
not in excess of 25 per cent; consequently, the term "solid brick" is not confined to those units which have no cores,
unless so stated in the project specifications.
Hollow Masonry Units. ASTM Specification C 652 covers hollow building brick, facing brick or hollow masonry units
made from clay, shale, fire clay or mixtures thereof, and fired. The term "hollow" in this specification is defined to
mean any unit cored in excess of 25 per cent, but not more than 40 per cent, in every plane parallel to the bearing
surface.
Supplementary Requirements. ASTM specifications for brick and tile do not fix the size or color and texture of the
units. They do, however, include requirements for several grades and types of products, and some of them contain
optional requirements which are applicable to specific projects, if so specified.
When ASTM specifications are included in project specifications by reference, it is essential that they be
supplemented with project requirements covering size, color, grade, type, etc. Without these supplementary
provisions, the specifications are incomplete and inadequate as a basis for estimating.
Size. Size of units required should be included in the project specifications. Without this information, a contractor
cannot accurately estimate quantity of materials or the labor required to construct the masonry.
It is recommended that the specified size be the manufactured size. Individual unit dimensions may vary from the
specified or manufactured size by the allowable tolerances included in the appropriate ASTM specifications for the
particular type or grade.
Specifying nominal sizes of clay masonry units is not recommended, due to the ambiguity of the term "nominal". In
some fields, it is understood to mean approximate and actual dimensions may vary from the nominal only by
permissible variations in dimensions included in the specifications. However, in modular design, the nominal
dimension of a masonry unit is understood to mean the specified or manufactured dimension plus the thickness of
the mortar joint with which the unit is designed to be laid; that is, modular brick, whose nominal length is 8 in.,
would have a specified (manufactured) length of 7 1/2 in. if designed to be laid with a 1/2 - in. joint, or 7 5/8 in. if
designed to be laid with a 3/8 - in joint.
Color and Texture. Generally, the color and texture of the brick or structural facing tile in a masonry wall vary
slightly. These variations, which prevent monotony in the appearance of the finished wall, are one of the most
attractive features of brick and tile. Because of these variations and of the wide variety of colors and textures
produced by the industry, it is impossible to write descriptions of either color or texture which will accurately identify
the products required.
For this reason, ASTM specifications for brick and structural clay facing tile provide that texture and color shall
http://www.gobrick.com/BIA/technotes/t11.htm
2 of 4 9/13/2009 12:42 PM
conform to an approved sample showing the full range of color and texture that will be acceptable. The number of
units required in the sample should be stated in the project specifications and will depend upon the range of color
and texture. In general, it will be from three to five.
Grade and Type. Most ASTM specifications for brick or structural clay tile cover two or more grades, and
specifications for facing brick, hollow brick and ceramic glazed structural facing tile include requirements for two or
more types. Specifications for structural clay facing tile cover two types and two classes.
When these specifications are included in project specifications by reference, it is essential that the grade and type
or type and class of product required be specified. Failure to do so makes it difficult for the contractor to estimate
the project and frequently results in a demand for extras after the contract is awarded.
Cell Arrangement. Structural clay tile are produced with either vertical cells or horizontal cells. Furring tile, nominal
thickness 2 in., in ceramic glaze often referred to as "soaps", are produced with either solid backs or open (ribbed)
backs. If either vertical-cell or horizontal-cell units are required for specific locations, this should be stated in the
project specifications. Similarly, if solid-back soaps or furring are required, it should be so stated. Otherwise,
product specifications make the selection optional with the supplier.
Plaster Base Finish. Specifications for structural clay facing tile and structural clay tile contain requirements for the
finish of surfaces suitable for the application of plaster. When such surfaces are required, they should be specified
in the project specifications; otherwise, the finish of the unexposed (back) of the unit is optional with the supplier.
Tests. Most ASTM specifications for structural clay products provide that the cost of tests of units furnished for any
particular project "shall be borne by the purchaser", unless the tests indicate that the units do not conform to the
requirements of the specifications, in which case "the cost shall be borne by the seller". Project specifications
should state the number of tests that will be required and should indicate who is responsible for selecting the
samples and who pays the cost of testing.
PROJECT SPECIFICATIONS FOR STRUCTURAL CLAY PRODUCTS
As previously indicated, it is recommended that ASTM specifications, supplemented to meet project requirements,
be used in specifying brick and structural clay tile. These specifications are suitable for use in any of the following
forms:
Open Specifications. This type of specification, frequently required in public work, makes no reference to product
trade names. In such a specification, ASTM specifications should be included by reference, supplemented with
project requirements, and an "approved sample" of the required color and texture should be available for inspection
by bidders prior to submission of bids.
Trade Names. For private work, specifying facing brick and structural facing tile by trade or manufacturer's names
gives the contractor definite information as to the product required and provides the architect with assurance that
the quality desired will be furnished.
In general, when this method is used, three or more acceptable products are named and the contractor is given the
option of selecting among them.
When trade names are used for specifying brick or tile, it is recommended that the units be required to comply with
applicable ASTM specifications and that samples of acceptable units be available for inspection of bidders prior to
bidding; also, that a provision for substitution, similar to that previously recommended, be included in the
specifications.
Allowances. The use of allowances for cost of facing brick and facing tile has been used successfully for many
years and, in general, this method is recommended by the Structural Clay Products Institute. Allowances place all
contractors on an equal basis and permit the owner to select products that he considers most desirable. However,
when this method is employed, the specifications should state the size and texture of the units that will be selected,
the tests that will be required and the responsibility for payment of tests.
GUIDE SPECIFICATIONS
http://www.gobrick.com/BIA/technotes/t11.htm
3 of 4 9/13/2009 12:42 PM
The guide specifications in Technical Notes 11A Revised and 11B Revised are written for both reinforced and
non-reinforced brick masonry, designed to comply with ANSI A41.1-1953 (R1970), "Building Code Requirements
for Masonry", ANSI A41.2-1960 (R 1970), "Building Code Requirements for Reinforced Masonry", or equivalent
sections in the Model Building Codes.
The guide specifications in these Technical Notes can be used for engineered brick masonry designed to comply
with Building Code Requirements for Engineered Brick Masonry, BIA, August 1969, or equivalent sections in the
Model Building Codes, when additional quality assurance requirements are incorporated into the specification. See
Technical Notes 11C Revised.
The specifications do not cover requirements for structural clay tile, concrete masonry units, glass block or stone.
Where these materials and design procedures are included in the masonry section, the specifications should be
supplemented or revised. It will be found, however, that many of the requirements pertaining to brick masonry are
also applicable to other types of masonry construction.
"Guide Specifications for Masonry Mortar" will be included as a separate Technical Notes 11E to comply with CSI
format.
Metric numbers listed are conversions from the current customary system and are not industry agreed-upon
standards; i.e., a typical modular 3 1/2 x 2 1/4 x 7 1/2 - in.. (actual size) brick may be produced at some
dimensions other than 89 x 57 x 191 mm when metric dimensions are adopted within the industry.
The cold weather protection requirements contained in paragraph 1.05.C are those recommended by the
International Masonry Industry All-Weather Council, published December 1, 1970.
In using these specifications, the specification writer should cheek each section to insure compliance with project
requirements and modify the paragraphs or delete those not needed.
REFERENCES
1. Brick and Tile Engineering, Harry C. Plummer, Brick Institute of America (BIA), November 1967.
2. Building Code Requirements for Engineered Brick Masonry, BIA, August 1969.
3. Recommended Practice for Engineered Brick Masonry, J. G. Gross, R. D. Dikkers and J. C.
Grogan, BIA, November 1969.
4. Specifications for Clay Masonry Construction, BIA, February 1962.
5. Technical Notes on Brick Construction, BIA, published monthly.
6. Building Code Requirements for Masonry, ANSI - A41.1-1953 (R 1970).
7. Building Code Requirements for Reinforced Masonry, ANSI - A41.2-1960 (R 1970).
http://www.gobrick.com/BIA/technotes/t11.htm
4 of 4 9/13/2009 12:42 PM

Technical Notes 11A - Guide Specifications for Brick Masonry, Part 2
Rev [June 1978] (Reissued Sept. 1988)
INTRODUCTION
This Technical Notes contains the guide specifications in CSI format for Division 4, Section 04210, Part I - General,
and Part II - Products. Part III - Execution is in Technical Notes 11B Revised.
The specifications are applicable to ANSI A41.1 - 1953 (R1970), ''Building Code Requirements for Masonry,'' ANSI
A41.2 - 1960 (R 1970), "Building Code Requirements for Reinforced Masonry,'' or equivalent sections in the Model
Building Codes.
The guide specifications in Technical Notes 11A Revised and 11B Revised can be used for engineered brick
masonry designed to comply with Building Code Requirements for Engineered Brick Masonry, BIA, August 1969,
or equivalent sections in the Model Building Codes, when additional quality assurance requirements are incorporated
into the specifications. See Technical Notes 11C Revised.
Guide Specification & Notes
PART I - GENERAL
1.01 DESCRIPTION:
A. Related Work Specified Elsewhere:
1. Concrete work: Section 03__________.
2. Rough carpentry: Section 06__________.
3. Structural steel and metals: Section 05__________.
4. Waterproofing: Section 07__________.
B. Material Installed but Furnished by Others:
1. Bolts.
2. Anchors.
3. Nailing blocks.
4. Inserts.
5. Flashing.
6. Lintels.
7. Doors.
8. Window frames.
9. Vents.
10. Conduits.
http://www.gobrick.com/BIA/technotes/t11a.htm
1 of 9 9/13/2009 12:43 PM
11. Expansion joints.
1.02 QUALITY ASSURANCE:
A. Brick Tests:
1. Test in accordance with ASTM C 67-__________with the following additional requirements:
a. If the coefficient of variation of the compression samples tested exceeds 12%, obtain
compressive strength by multiplying average compressive strength of specimens by
where v is the coefficient of variation of sample tested.
b. Cost of tests of units after delivery shall be borne by the purchaser, unless tests
indicate that units do not conform to the requirements of the specifications, in which case
cost shall be borne by the seller.
NOTE:
1.02.A This section can be deleted if Architect/Engineer has sufficient experience and confidence in the brick manufacturer to
accept compliance with project specifications based on certification section 1.03.C.
1.02.A. 1.a. To be applied only for engineered brick masonry.
1.02.A. 1.b. To be used only in a case of dispute.
B. Furnish Sample Panel:
1. Approximately 4 ft. (1.2 m) long by 3 ft. (1 m) high, showing the proposed color range,
texture, bond, mortar and workmanship. All brick shipped for the sample shall be included in the
panel.
2. Erect panel in the presence of the Architect/Engineer before installation of materials.
3. When required, provide a separate panel for each type of brick or mortar.
4. Do not start work until Architect/Engineer has accepted sample panel.
5. Use panel as standard of comparison for all masonry work built of same material.
6. Do not destroy or move panel until work is completed and accepted by Owner.
NOTE:
1.02. B. 1. The sample panel, when accepted, shall become the project standard for: bond, mortar, workmanship and
appearance.
1.02.B.3. Brick for sample panels are usually furnished at no cost. If additional panels are needed, care must be exercised not to
burden the supplier with excessive costs.
1.03 SUBMITTALS:
A. Samples: Furnish not less than five individual brick as samples, showing extreme variations in color
and texture.
B. Test Reports:
1. Test reports for each type of building and facing brick are to be submitted to the Architect
Engineer for approval.
2. Testing and reports are to be completed by an independent laboratory.
3. Test reports shall show:
a. Compressive strength.
b. 24 - hr. cold water absorption.
c. 5 - hr. boil absorption.
http://www.gobrick.com/BIA/technotes/t11a.htm
2 of 9 9/13/2009 12:43 PM
d. Saturation coefficient.
e. Initial rate of absorption (suction).
C. Certificates: Prior to delivery, submit to Architect/Engineer certificates attesting compliance with the
applicable specifications for grades, types or classes included in these specifications.
NOTE:
1.03.A and B Sections can be deleted if Architect/Engineer has sufficient experience and confidence in the brick manufacturer to
accept compliance with project specifications based on certification section 1.03.C.
1.03.B.3. This section should be altered to meet the requirements of the project. Brick are not required to meet the 5-hr boil
absorption and/or saturation coefficient requirements of ASTM C 216, ASTM C 62 and ASTM C 652 if they meet the physical
property requirements of Sections 5.1 and 5.2 of ASTM C 216, Sections 3A, 3.5 and 3.6 of ASTM C 62 and Sections 5.1 and 5.2 of
ASTM C 652.
No limit is placed on initial rate of absorption (suction). Units having initial rates of absorption exceeding 30 g./min./30 sq. in. (194
cm
2
) should be wetted prior to laying. For cold weather masonry construction, higher suctions may be tolerated (up to 30-40 g.)
than for normal construction. Note Sections 1.05.C.2.a and 3.01.A. 1.
1.03.C. List materials for which certificates of compliance are required.
1.04 PRODUCT DELIVERY, STORAGE AND HANDLING:
A. Store brick off ground to prevent contamination by mud, dust or materials likely to cause staining or
other defects.
B. Cover materials when necessary to protect from elements.
C. Protect reinforcement from elements
1.05 JOB CONDITIONS:
A. Protection of Work:
1. Wall covering:
a. During erection, cover top of wall with strong waterproof membrane at end of each
day or shutdown.
b. Cover partially completed walls when work is not in progress.
c. Extend cover minimum of 24 in. (610 mm) down both sides.
d. Hold cover securely in place.
2. Load application:
a. Do not apply uniform floor or roof loading for at least 12 hr. after building masonry
columns or walls.
b. Do not apply concentrated loads for at least 3 days after building masonry columns or
walls.
B. Staining:
1. Prevent grout or mortar from staining the face of masonry to be left exposed or painted:
a. Remove immediately grout or mortar in contact with face of such masonry.
b. Protect all sills, ledges and projections from droppings of mortar, protect door jambs
and corners from damage during construction.
Protection:
http://www.gobrick.com/BIA/technotes/t11a.htm
3 of 9 9/13/2009 12:43 PM
1. Preparation:
a. If ice or snow has formed on masonry bed, remove by carefully applying heat until top
surface is dry to the touch.
b. Remove all masonry deemed frozen or damaged.
2. Products:
a. When brick suction exceeds recommendations of Section 1.03.B.3, sprinkle with
heated water:
(1) When units are above 32
o
F. (0
o
C.), heat water above 70
o
F. (21
o
C.).
(2) When units are below 32
o
F. (0
o
C.), heat water above 130
o
F. (54
o
C.).
b. Use dry masonry units.
c. Do not use wet or frozen units.
3. Construction requirements while work is progressing:
a. Air temperature 40
o
F. (4
o
C.) to 32
o
F. (0
o
C.):
(1) Heat sand or mixing water to produce mortar temperatures between 40
o
F. (4
o
C.) and 120
o
F. (49
o
C.).
b. Air temperature 32
o
F. (0
o
C.) to 25
o
F. (-4
o
C.):
(1) Heat sand and mixing water to produce mortar temperatures between 40
o
F.
(4
o
C.) and 120
o
F. (49
o
C.).
(2) Maintain temperatures of mortar on boards above freezing.
c. Air temperatures 25
o
F. (-4
o
C.) to 20
o
F. (-7
o
C.):
(1) Heat sand and mixing water to produce mortar temperatures between 40
o
F.
(4
o
C.) and 120
o
F. (49
o
C.).
(2) Maintain mortar temperatures on boards above freezing.
(3) Use salamanders or other heat sources on both sides of walls under
construction.
(4) Use windbreaks when wind is in excess of 15 mph.
d. Air temperature 20
o
F. (-7
o
C.) and below:
(1) Heat sand and mixing water to produce mortar temperatures between 40
o
F.
(4
o
C.) and 120
o
F. (49
o
C.).
(2) Provide enclosures and auxiliary heat to maintain air temperature above 32
o
F.
(0
o
C.).
(3) Minimum temperature of units when laid: 20
o
F. (-7
o
C.).
http://www.gobrick.com/BIA/technotes/t11a.htm
4 of 9 9/13/2009 12:43 PM
4. Protection requirements for completed masonry and masonry not being worked on:
a. Mean daily air temperature 40
o
F. (4
o
C.) to 32
o
F. (0
o
C.):
(1) Protect masonry from rain or snow for 24 hr. by covering with weather-
resistive membrane.
b. Mean daily air temperature 32
o
F. (0
o
C.) to 25
o
F. (-4
o
C.):
(1) Completely cover masonry with weather-resistive membrane for 24 hr.
c Mean daily air temperature 25
o
F. (-4
o
C.) to 20
o
F. (-7
o
C.):
(1) Completely cover masonry with insulating blankets or equal protection for 24
hr.
d. Mean daily air temperature 20
o
F. (-7
o
C.) and below:
(1) Maintain masonry temperature above 32
o
F. (0
o
C.) for 24 hr. by:
(a) Enclosure and supplementary heat.
**OR**
(a) Electric heating blankets.
**OR**
(a) Infrared lamps.
**OR**
(a) Other approved methods.
NOTE:
1.05.C.3 Ideal mortar temperature is 70
o
F. 10
o
F. (21
o
C. 6
o
C.). The mixing temperature should be maintained within 10
o
F.
(6
o
C.).
1.05.C.4 The following options may be used in cold weather construction:
1. Change to a higher type of mortar required in ASTM C 270. (Example: If ASTM type N mortar is specified for normal
temperature, change to type S or type M.)
2. Increase the protection time where required in Section 1.05.C.4 to 48 hr. with no change being made in the type of
mortar.
3. Without changing the mortar type and maintaining 24-hr. protection in Section 1.05.C.4, replace type I Portland cement
in the mortar with type III, ASTM C 150.
1.05.C.4.d This section may be written to allow the contractor to select means of protection.
PART II-PRODUCTS
2.01 BRICK:
A. Facing Brick:
1. ASTM C 216-__________, Grade__________, Type__________.
2. Dimensions:__________ x __________ x __________.
(t) (h) (l )
3. Minimum compressive strength: _____________.
4. Provide brick similar in texture and physical properties to those available for inspection at the
http://www.gobrick.com/BIA/technotes/t11a.htm
5 of 9 9/13/2009 12:43 PM
Architect/Engineer's office.
5. Do not exceed variations in color and texture of samples accepted by the Architect/Engineer.
NOTE:
2.01.A.1 Grade: SW for brick in contact with earth or where weathering index is greater than 50, MW elsewhere. Type: FBS, FBX,
FBA.
2.01.A.2 Determine availability. Typical actual sizes for use with 3/8 - in. mortar joints: 3 5/8 x 2 5/16 x 7 5/8 or 11 5/8 in. (92 x 59 x
194 or 295 mm); 3 5/8 x 2 13/16 x 7 5/8 or 11 5/8 in. (92 x 74 x 194 or 295 mm): 3 5/8 x 3 5/8 x 7 5/8 or 11 5/8 in. (92 x 92 x 194 or
295 mm);3 5/8 x 5 x 7 5/8 or 11 5/8 - in (92x 127x 194 or 295 mm);3 5/8 x 1 5/8 x 11 5/8 in. (92 x 41 x 295 mm); 5 5/8 x 2 5/16 x 11
5/8 in. (143 x 59 x 295 mm); 5 5/8 x 2 13/16 x 11 5/8 in. (l43 x 74 x 295 mm); 5 5/8 x 3 5/8 x 11 5/8 in. (143 x 92 x 295 mm).
2.01.A.3 Required only for structural masonry. Range: 2000 psi to 14,000 psi (13.8 MPa to 96.5 MPa).
**OR**
A. Facing Brick: Provide a cash allowance of__________per thousand.
B. Glazed Brick:
1. ASTM C 126-__________, Grade__________, Type__________.
2. Dimensions:__________ x __________ x __________.
(t) (h) (l )
3. Minimum compressive strength:__________.
NOTE:
2.01.B.1 Grade: S for narrow mortar joints; SS where face dimension variation must be very small. Type: I, II.
2.01.B.2 See 2.01.A.2.
2.01. B.3 See 2.01.A.3.
C. Building Brick:
1. ASTM C 62-__________, Grade__________.
2. Dimensions:__________ x __________ x __________.
(t) (h) (l )
3. Minimum compressive strength:__________.
NOTE:
2.01.C.1 Grade: SW for brick in contact with earth or where weathering index is greater than 50, MW elsewhere, NW in interior
and backup areas.
2.01.C.2 See 2.01.A.2.
2.01.C.3 See 2.01.A.3.
D. Hollow Brick:
1. ASTM C 652-__________, Grade__________, Type__________.
2. Dimensions:__________ x __________ x __________.
(t) (h) (l )
3. Minimum compressive strength:__________.
4. Provide brick similar in texture and physical properties to those available for inspection at the
Architect/Engineer's office.
http://www.gobrick.com/BIA/technotes/t11a.htm
6 of 9 9/13/2009 12:43 PM
5. Do not exceed variation in color and texture of samples accepted by the Architect/Engineer.
NOTE:
2.01.D.1 Grade: SW for brick in contact with earth or where weathering index is greater than 50, MW elsewhere. Type: HBS, HBX,
HBA, HBB.
2.01.D.2 See 2.01.A.2.
2.01.D.3 See 2.01.A.3.
2.02 REINFORCEMENT:
A. Cold-drawn steel wire: ASTM A 82-__________.
B. Welded steel wire fabric: ASTM A 185-__________.
C. Billet steel deformed bars: ASTM A 615-_________, Grade__________.
D. Rail steel deformed bars: ASTM A 616-__________, Grade_________.
E. Axle steel deformed bars: ASTM A 617-__________, Grade__________.
NOTE:
2.02.C Grade 40, 50, 60.
2.02.D Grade 50, 60.
2.02.E Grade 40, 60.
2.03 ANCHORS AND TIES:
A. Coated or corrosion-resistant metal meeting or exceeding applicable standard:
1. Zinc-coating flat metal: ASTM A 153-__________, Class__________.
2. Zinc-coating of wire, ASTM A 116-__________, Class 3.
3. Copper-coated wire: ASTM B 227-__________ , Grade 30HS.
4. Stainless steel: ASTM A 167-__________, Type 304.
NOTE:
2.03.A.1 Class B-1, B-2, B-3.
B. Types:
1. Wire mesh:
a. Minimum gage: 20.
b. Mesh: 1/2 in. (12.7 mm).
c. Galvanized wire.
d. Width: 1 in. (25 mm) less than width of masonry.
2. Corrugated veneer ties:
a. Minimum gage: 22.
b. Minimum width: 7/8 in. (22 mm).
c. Length: 6 in. (152 mm)
**OR**
http://www.gobrick.com/BIA/technotes/t11a.htm
7 of 9 9/13/2009 12:43 PM
2. Wire ties: Use two 10-gage.
3. Cavity wall ties:
a. Wire diameter: 3/16 in. (4.7 mm).
b. Shape: Rectangular, at least 2 in. (51 mm) wide with ends overlapped or "Z" with 2 -
in. (51 mm) legs.
c. Length: Select length to allow 1 - in. (25 mm) minimum mortar cover of ends or legs.
4. Multi-wythe wall ties:
a. Prefabricated welded joint reinforcement.
b. Longitudinal cross tie wire:
(1) 9 gage.
(2) Spaced 16 in. (406 mm) o.c.
NOTE:
2.03.B.4 Cavity wall ties may be used.
5. Dovetail flat bar or wire anchors:
a. Flat bar:
(1) Minimum gage: 16.
(2) Minimum width: 7/8 in. (22 mm).
(3) Fabrication: Corrugated, turned up 1/4 in. (6.4 mm) at end or with 1/2-in. (12.7
mm) hole within 1/2 in. (12.7 mm) of end of bar.
b. Wire:
(1) Wire gage: 6.
(2) Minimum width: 7/8 in. (22 mm).
(3) Fabrication: Wire looped and closed.
6. Rigid anchors for intersecting bearing walls:
a. Dimensions: 1 1/2 in. (38 mm) wide by 1/4 in. (6.4 mm) thick by minimum 24 in. (610
mm) long.
b. Fabrication: Turn up ends minimum 2 in. (51 mm) or provide cross pins.
7. Wire ties for high-lift grout reinforced brick masonry:
a. Minimum gage: 9.
b. Fabrication:
(1) Bend into stirrups 4 in. (102 mm) wide and 2 in. (51 mm) shorter than overall
wall thickness.
(2) Form so that tie ends meet in center of one embedded end of stirrup.
2.04 CLEANING AGENTS:
http://www.gobrick.com/BIA/technotes/t11a.htm
8 of 9 9/13/2009 12:43 PM
A. Do not use cleaning agent other than water on brick, except with concurrence of Architect/Engineer.
B. Acceptable cleaner for dark brick: _____________.
C. Acceptable cleaner for light colored brick: _______________.
NOTE:
2.04.B Specify cleaner recommended by brick manufacturer.
2.04.C Proper cleaning agent is more critical for light colored brick.


http://www.gobrick.com/BIA/technotes/t11a.htm
9 of 9 9/13/2009 12:43 PM

Technical Notes 11B - Guide Specifications for Brick Masonry, Part 3
Rev [Feb. 1972] (Reissued Sept. 1988)

INTRODUCTION
This Technical Notes contains the guide specifications in CSI format for Part III - Execution. Part I - General, and
Part II - Products are in Technical Notes 11A Revised.
Guide Specification and Notes
PART III - EXECUTION
3.01 PREPARATION:
A. Wetting Brick:
1. Wet brick with absorption rates in excess of 30 g./30 sq. in./min. (30 g./194 cm
2
/min.)
determined by ASTM C 67-__________, so that rate of absorption when laid does not exceed
this amount.
2. Recommended procedure to insure that brick are nearly saturated, surface dry when laid is
to place a hose on the pile of brick until the water runs from the pile. This should be done one
day before brick are to be used. In extremely warm weather, place hose on pile several hours
before brick are to be used.
B. Cleaning Reinforcement: Before being placed, remove loose rust, ice and other coatings from
reinforcement.
NOTE:
3.01.A.1 Note requirements for cold weather, section 1.05.C.2.a, and section 1.03.B.3 for testing requirements.
3.02 GENERAL ERECTION REQUIREMENTS:
A. Pattern Bond:
1. Lay exposed masonry in running bond.
2. Bond unexposed masonry units in a wythe by lapping at least 2 in. ( 51 mm).
NOTE:
3.02.A.1 Alter if other than running bond required.
B. Joining of Work:
1. Where fresh masonry joins partially set masonry:
a. Remove loose brick and mortar.
b. Clean and lightly wet exposed surface of set masonry.
2. Stop off horizontal run of masonry by racking back 1/2 length of unit in each course.
3. Toothing is not permitted except upon written acceptance of the Architect/Engineer.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
1 of 9 9/13/2009 12:44 PM
C. Tooling and Tuck Pointing:
1. Tooling:
a. Tool exposed joints when "thumb-print" hard with a round jointer, slightly larger than
width of joint.
b. Trowel-point or concave-tool exterior joints below grade.
c. Flush cut all joints not tooled.
2. Tuck pointing:
a. Rake mortar joints to a depth of not less than 1/2 in.(12.7 mm) nor more than 3/4 in.
(19 mm).
b. Saturate joints with clean water.
c. Fill solidly with ______________ pointing mortar.
d. Tool joints.
NOTE:
3.02.C Alter to allow other joints to meet architectural requirements.
3.02.C.2 Delete if not required.
3.02.C.2.c Specify proportions. Pointing mortar should be of same proportions as mortar in main part of wall, if known; if not, type
N.
D. Flashing:
1. Clean surface of masonry smooth and free from projections which might puncture flashing
material.
a. Place through-wall flashing on bed of mortar.
b. Cover flashing with mortar.
E. Weep Holes:
1. Provide weep holes in head joints in first course immediately above all flashing by: (a)
Leaving head joint free and clean of mortar
**OR**
(a) Placing and leaving sash cord in joint.
********
2. Maximum spacing: 24 in. (610 mm) o.c.
3. Keep weep holes and area above flashing free of mortar droppings.
F. Sealant Recesses:
1. Leave joints around outside perimeters of exterior doors, window frames and other wall
openings:
a. Depth: uniform 3/4 in. (19 mm).
b. Width: 1/4 in. (6.4 mm) to 3/8 in. (9.5 mm).
G. Movement Joints:
1. Keep clean from all mortar and debris.
2. Locate as shown on drawings.
H. Cutting Brick:
1. Cut exposed brick with motor-driven saw.
**OR**
1. By other methods which provide cuts that are straight and true.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
2 of 9 9/13/2009 12:44 PM
********
I. Mortar Joint Thickness:
1. Lay all brick with__________in. joint.
NOTE:
3.02.I Coordinate joint thickness with brick specified in 2.01.A.2.
3.03 NON-REINFORCED BRICK MASONRY
A. Brick Installation:
1. Lay brick plumb and true to lines.
2. Lay with completely filled mortar joints.
3. Do not furrow bed joints.
4. Butter ends of brick with sufficient mortar to fill head joints.
5. Rock closures into place with head joints thrown against two adjacent brick in place.
6. Fill vertical, longitudinal joints, except in cavity walls:
a. By parging either face of backing or back of facing.
**OR**
a. By pouring the vertical joint full of grout.
**OR**
a. Shoving alone.
7. Do not pound corners and jambs to fit stretcher units after they are set in position. Where an
adjustment must be made after mortar has started to harden, remove mortar and replace with
fresh mortar.
NOTE:
3.03.A If hollow units are specified, alter to conform to requirements of the units.
B. Cavity Walls:
1. Keep cavity in cavity walls clean by:
a. Slightly beveling mortar bed to incline toward cavity.
**OR**
a. Placing wood strips with attached wire pulls on metal ties.
b. Before placing next row of metal ties, remove and clean wood strips.
********
2. As work progresses, trowel protruding mortar fins in cavity flat on to inner face of wythe.
C. Non-Bearing Partitions:
1. Extend from top of structural floor to bottom surface of floor construction above.
2. Wedge with small pieces of tile, slate or metal.
3. Fill topmost joint with mortar.
NOTE:
3.03.C Alter to local code requirements if suspended ceilings are used.
D. Structural Bonding:
1. Bond or anchor corners and intersections of loadbearing brick walls.
2. Structural bond multi-wythe non-reinforced brick walls with__________.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
3 of 9 9/13/2009 12:44 PM
a. Extend headers not less than 3 in. (76 mm) into backing.
b. Maximum distance between adjacent headers: 24 in. (610 mm) either vertically or
horizontally.
c. When a single header does not extend through wall, overlap headers from opposite
sides of wall at least 3 in. (76 mm).
d. Minimum headers: 4%.
**OR**
a. Provide minimum of one cavity wall tie for each 4 1/2 sq. ft. (0.42 m
2
) of wall surface.
b. Stagger ties in alternate courses.
c. Maximum distance between adjacent ties:
(1) Vertically: 24 in. (610 mm).
(2) Horizontally: 36 in. (920 mm).
d. Embed ties in horizontal joints of facing and backing.
e. Provide additional ties at openings:
(1) Maximum spacing around perimeter: 36 in. (920 mm).
(2) Install within 12 in. (305 mm) of opening.
**OR**
a. Use continuous prefabricated joint reinforcement to bond multi-wythe walls; spaced
not more than 16 in. (406 mm) vertically.
********
3. Stack bond:
a. Embed continuous No. 2 steel reinforcement or No. 9 gage wire in horizontal joints at
vertical intervals not to exceed 16 in. (406 mm).
b. Provide not less than one longitudinal bar or wire for each 6 in. (152 mm) of wall
thickness or fraction thereof.
NOTE:
3.03.D Note special bonding requirements for high-lift grout, section 3.05.B.
3.03.D.2 Masonry headers, metal ties or continuous joint reinforcement.
E. Anchoring:
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
4 of 9 9/13/2009 12:44 PM
1. Anchor exterior brick walls facing or abutting concrete members with dovetail, flat-bar or wire
anchors inserted in slots built into concrete.
a. Maximum anchor spacing:
(1) Vertically: 24 in. (610 mm).
(2) Horizontally: 36 in. (920 mm).
b. Maintain a space not less than 1/2 in. (12.7 mm) wide between masonry wall and
concrete members.
c. Keep space free of mortar or other rigid material to permit differential movement
between concrete and masonry .
2. For intersecting bearing or shear walls carried up separately:
a. Regularly block vertical joint with 8-in. (203 mm) maximum offsets.
b. Provide joints with rigid steel anchors.
c. Space anchors not more than 4 ft. (1.2 m) apart vertically.
**OR**
a. When acceptable to the Architect/Engineer, eliminate blocking and provide rigid steel
anchors spaced not more than 24 in. (610 mm) apart vertically.
********
3. Anchor non-bearing partitions abutting or intersecting other walls or partitions with:
a. Cavity wall ties at vertical intervals of not more than 24 in. ( 610 mm).
**OR**
a. Masonry bonders in alternate courses
********
4. Attach brick veneer to backing with metal veneer ties:
a. Use one tie for each 4 sq. ft. (0.37 m
2
) of wall area.
b. Maximum space between adjacent ties:
(1) Vertically and horizontally: 24 in. (610 mm).
c. Embed ties at least 2 in. (51 mm) in horizontal joint of facing.
d. Provide additional ties at openings:
(1) Maximum spacing around perimeter: 36 in. (914 mm).
(2) Install within 12 in. (305 mm) of opening.
NOTE:
3.03.E 4 Tie spacing is based on a design wind pressure of 20 psf (958 N/m
2
). Maximum spacing should be decreased for higher
wind pressures. Recommended spacing for:
30 psf (1436 N/m2):
Vertically: 24 in. (610 mm)
Horizontally: 16 in. (406 mm)
40 psf (1913 N/m2):
Vertically: 18 in. (457 mm)
Horizontally: 16 in. (406 mm)
3.04 REINFORCED BRICK MASONRY:
A. Brick Installation:
1. Lay brick plumb and true to lines.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
5 of 9 9/13/2009 12:44 PM
2. Lay with completely filled mortar joints.
3. Do not furrow bed joints.
4. Butter ends of brick with sufficient mortar to fill head joints.
5. Slightly bevel mortar bed to incline towards cavity.
6. Rock closures into place with head joints thrown against two adjacent brick in place.
7. Do not pound corners and jambs to fit stretcher units after they are set in position. Where an
adjustment must be made after mortar has started to harden, remove mortar and replace with
fresh mortar.
NOTE:
3.04.A If hollow units are specified, alter to conform to requirements of the units.
B. Forms and Shores:
1. Provide substantial and tight forms.
2. Leakage of mortar or grout is not permitted.
3. Brace or tie forms to maintain position and shape.
4. Do not remove forms and shores until masonry has hardened sufficiently to carry its own
weight and other temporary loads that may be placed on it during construction:
a. For girders and beams: Minimum 10 days.
b. Under brick slabs: Minimum 7 days.
C. Placing Reinforcement:
1. Position metal reinforcement accurately.
2. Secure against displacement:
a. Hold vertical reinforcement firmly in place by means of frames or other suitable
devices.
b. Horizontal reinforcement may be placed as brickwork progresses.
3. Spacing:
a. Minimum clear distance between longitudinal bars, except in columns: Nominal
diameter of bar or 1 in. (25 mm).
b. Minimum clear distance between bars in columns: Not less than 1/2 times bar
diameter or 1/2 in. (38 mm). 4
4. Minimum thickness of mortar or grout between brick and reinforcement: 1/4 in. (6.4 mm),
except:
a. 1/4 - in. (6.4 mm) bars may be laid in 1/2-in.(12.7 mm) horizontal mortar joints.
b. No. 6 gage or smaller wires may be laid in 3/8-in. (9.5 mm) mortar joints.
5. Minimum width of collar Joints containing both horizontal and vertical reinforcement: 1/2 in.
(12.7 mm) larger than sum of diameters of horizontal and vertical reinforcement.
6. Splice reinforcement or attach reinforcement to dowels by placing in contact and wiring.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
6 of 9 9/13/2009 12:44 PM
7. Do not splice reinforcement at points other than shown on drawings, unless approved by the
structural engineer.
8. Shape and dimension reinforcement as shown on drawings:
a. Cold bend all bars.
b. Do not straighten or repair in a manner that will injure material.
c. Do not use bars with kinks or bends not shown on drawings.
d. Reinforcement can be heated when entire operation is approved by structural
engineer.
3.05 GROUTING:
A. Low-Lift Grouting:
1. Keep grout core clean from mortar and drippings.
2. Grout spaces less than 2 in. (51 mm) in width at intervals of not more than 24 in. (610 mm) in
lifts of 6 to 8 in. (152 to 203 mm) as the wall is built.
3. In grout spaces more than 2 brick in thickness:
a. Place or float brick in grout.
b. Minimum grout between brick: 3/8 in. (9.5 mm).
4. Agitate or puddle grout during and after placement to insure complete filling.
5. Stop grout 1/2 in. (38 mm) below top of masonry:
a. If grouting is stopped for 1 hr. or more.
b. Except when completing grouting of finished wall.
6. If brick headers are used for ties in low-lift grouting space:
a. Maximum: 8% of wall area.
NOTE:
3.05.A.6 Using headers for tying wythes is not recommended; however, if selected, construction should conform to the
requirements of this section.
B. High-Lift Grouting:
1. For running bond, provide one metal tie for each 3 sq. ft. (0.28 ma) of wall with maximum
spacing:
a. Vertically: 16 in. (406 mm).
b. Horizontally: 24 in. (610 mm).
2. For stack bond, provide one metal tie for each 2 sq. ft. (0.19 m2) of wall with maximum
spacing:
a. Vertically: 12 in. (305 mm).
b. Horizontally: 24 in. (610 mm).
3. Keep grout core clean from mortar and droppings.
4. Provide cleanout holes by omitting every other brick in bottom course on one side of wall.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
7 of 9 9/13/2009 12:44 PM
5. Prior to closing cleanout holes and pouring grout, use high-pressure jet stream of water or
high-pressure air to remove excess mortar from grout space and to clean reinforcement.
6. Do not plug cleanout holes until condition of area to be grouted has been approved.
7. Before pouring grout, plug cleanout holes with masonry units and brace against grout
pressure.
8. Grout spaces 2 in. (51 mm) or more in width in lifts not exceeding 4 ft. (1.2 m) at intervals:
a. Coarse grout: Not more than 48 times the least clear dimension of grout space.
**OR**
a. Fine grout: Not more than 64 times the least clear dimension of grout space.
********
b. Not to exceed height of 12 ft. (3.7 m).
9. Do not place grout until the entire wall has been in place 3 days.
10. Vibrate or agitate grout during, and after placement to insure complete filling of grout space.
11. Stop grout 1 1/2 in. (38 mm) below top of masonry:
a. If grouting is stopped for 1 hr. or more.
b. Except when completing grouting of finished wall.
12. Provide grout blocks at convenient intervals to meet project requirements.
3.06 CLEANING:
A. Cut out any defective joints and holes in exposed masonry and repoint with mortar.
B. Clean all exposed unglazed masonry:
1. Apply cleaning agent to sample wall area of 20 sq. ft. (2 m2 ) in location acceptable to the
Architect/Engineer.
2. Do not proceed with cleaning until sample area is approved by Architect/Engineer.
3. Clean initially with stiff brushes and water.
4. When cleaning agent is required:
a. Follow brick manufacturer's recommendations.
b. Thoroughly wet surface of masonry on which no green efflorescence appears.
c. Scrub with acceptable cleaning agent.
d. Immediately rinse with clear water.
e. Do small sections at a time.
f. Work from top to bottom.
g. Protect all sash, metal lintels and other corrodible parts when masonry is cleaned with
acid solution.
h. Remove green efflorescence in accordance with brick manufacturer's
recommendations.
NOTE:
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
8 of 9 9/13/2009 12:44 PM
3.06 If care is taken during laying and the wall is acceptable, the requirements of this section can be deleted.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11b.htm
9 of 9 9/13/2009 12:44 PM

Technical Notes 11C - Guide Specifications for Brick Masonry, Part 4
Rev. [July 1972] (Reissued May 1998)
INTRODUCTION
This issue of Technical Notes and the following issue, Technical Notes 11D, contain the required additional sections
and statements to be incorporated into the "Guide Specifications for Brick Masonry", Technical Notes 11A Revised
and 11B Revised. This will make the guide specifications in those Technical Notes suitable for Engineered Brick
Masonry.
The sections contained in these Technical Notes deal primarily with the quality assurance, selection of units, strength
and construction tolerances to provide masonry that meets the minimum design requirements for Engineered Brick
Masonry.
In the construction of Engineered Brick Masonry, quality control may be maintained in either of two ways: (1) by
testing the brick and controlling the mortar which can be done by laboratory tests or by mixing proportions, or (2) by
periodic testing of masonry prisms. This Technical Notes covers quality control by method (1), testing brick and
control of mortar. Technical Notes 11D covers quality control by method (2), testing masonry prisms.
When quality control by materials testing (brick and mortar) is to be used, the design compressive strength (f'm) can
be assumed, using Table 2 of the BIA Standard, "Building Code Requirements for Engineered Brick Masonry", and
the quality control requirements of this Technical Notes should be incorporated into the guide specifications in
Technical Notes 11A Revised and 11B Revised.
All other sections of the Guide Specifications for Brick Masonry (Technical Notes 11A Revised and 11B Revised)
are appropriate for Engineered Brick Masonry.
QUALITY ASSURANCE BASED ON BRICK AND MORTAR TESTS
Guide Specifications and Notes
PART I - GENERAL
1.02 QUALITY ASSURANCE
Delete section and notes for 1.02.A in Technical Notes 11A Revised, and substitute the following quality
control requirements based on brick tests.
A. Brick Tests:
1. Preconstruction Tests:
a. Test five brick for compressive strength to determine acceptability of units for
compliance with specifications.
b. Use brick similar to those selected for use, matching color, texture, raw material,
moisture content and coring.
c. Cost of tests shall be borne by the General Contractor.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11c.htm
1 of 5 9/13/2009 12:45 PM
2. Site Control Tests:
a. Test units selected at random from units delivered to the project.
b. Cost of tests of units after delivery shall be borne by the General Contractor, unless
tests indicate that units do not conform to the requirements of the specifications, in which
case cost shall be borne by the seller.
3. Test in accordance with ASTM C 67-__________, with the following additional requirements:
a. If the coefficient of variation of the compression samples tested exceeds 12%, obtain
compressive strength by multiplying average compressive strength of specimens by
1 - 1.5 ( V - 0.12 )
100
where v is the coefficient of variation of sample tested.
Add the following to Section 1.02 in Technical Notes 11A Revised.
C. Preconstruction Requirements:
1. Prebid conference:
a. A prebid conference, directed by the Architect/Engineer, will be held one week prior to
the bid opening to discuss:
(1) Structural concept.
(2) Method and sequence of masonry construction.
(3) Special masonry details.
(4) Quality control requirements.
(5) Material requirements.
(6) Job organization.
(7) Workmanship.
b. Attendance is mandatory for all prospective:
(1) General contractors.
(2) Masonry subcontractors.
(3) Brick suppliers.
NOTE:
1.02.C This requirement may be deleted if not necessary for the project due to bidders being knowledgeable with engineered brick
masonry.
1.02.C.1.b Invitation to attend should be extended to others, such as the inspectors (local building department and other
government agencies) and Owner.
2. Preconstruction Testing and Certification:
a. After award of the contract, the General Contractor shall:
(1) Within 14 days, submit to the Architect/Engineer for approval the name of the
independent laboratory which will perform the site control tests and provide the
certificates and test reports required in Section 1.03.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11c.htm
2 of 5 9/13/2009 12:45 PM
(2) Upon approval of the laboratory, certificates and test reports, and prior to any
masonry construction, make arrangements for the following tests:
(a) Brick tests in accordance with preconstruction requirements, Section
1.02.A
(b) Mortar tests in accordance with mortar section.
b. Masonry work can begin only after approval of testing.
c. Testing is acceptable if test results indicate that materials meet the minimum
requirements of Part II - Products.
d. Cost of preconstruction testing shall be borne by the General Contractor, unless tests
indicate that units do not conform to the requirements of the specifications, in which case
cost shall be borne by the seller.
NOTE:
1.02.C.2 Inspection, laboratory and testing for quality control can be a responsibility of the Structural Engineer. If so, revise section.
3. Preconstruction Conference:
a. A preconstruction conference, directed by the Architect/Engineer, will be held after the
award of the General Contract, but prior to beginning of masonry work to discuss:
(1) Structural concept.
(2) Method and sequence of masonry construction.
(3) Special masonry details.
(4) Standard of workmanship.
(5) Quality control requirements.
(6) Job organization.
b. Attendance is mandatory for:
(1) General contractor job superintendent.
(2) Masonry subcontractor job superintendent.
(3) Masonry subcontractor foreman.
(4) At least two masons.
(5) Authorized representative of the brick supplier.
(6) Mortar material suppliers.
NOTE:
1.02.C.3.b Invitations to attend should be extended to others, such as inspectors (local building department and other government
agencies) and Owner.
D. Job Site Quality Control:
1. Site control brick and mortar tests:
a. Use compressive strength of brick units and compressive strength of mortar cubes to
control quality.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11c.htm
3 of 5 9/13/2009 12:45 PM
b. Test brick in accordance with site control requirements, Section 1.02.A.
c. Test mortar cubes in accordance with mortar section.
d. Test five brick and three mortar cubes for each 100,000 brick or fraction thereof.
e. Brick and mortar to be selected at random by the Architect/ Engineer.
f. Site control data shall be acceptable if material exceeds specified strength.
g. Cost of tests shall be borne by the General Contractor.
NOTE:
1.02.D.1.c Type M, S or N as specified in the mortar section. More than one mortar type may be specified. If so, provide sections
to cover all requirements. Mortar design compressive strengths should be based on laboratory tests of mortar made from
materials mixed to the proportion specification as required by the mortar section.
1.02.D.1.e As calculated by the Structural Engineer. May vary for different parts of the building. If so, provide sections to cover all
design strengths.
**OR**
1. Site control brick and mortar batching:
a. Use compressive strength of brick units and control on material proportions used in
batching mortar to control quality.
b. Test brick in accordance with site control requirements, Section 1.02.A.
c. Control mortar batches to conform to proportion specification specified in mortar
section.
d. Test five brick for each 100,000 brick or fraction thereof.
e. Site control data shall be acceptable if brick compressive strengths meet the
requirements of Part II and mortar is batched to proportion specification.
f. Cost of tests shall be borne by the General Contractor.
PART II - PRODUCTS
2.01 BRICK
A. Facing Brick:
1. Delete Note and replace with:
NOTE:
2.01.A.1 Grades and Types. Brick subject to the action of weather or soil, but not subject to frost action when permeated with
water, shall be of grade MW or grade SW, and where subject to temperature below freezing while in contact with soil shall be
grade SW. Brick used in loadbearing or shear wall construction shall comply with the dimensional and distortion tolerances
specified for type FBS of ASTM C 216-__________. Where such brick do not comply with these tolerance requirements, the
compressive strength of brick masonry shall be determined by prism tests.
PART Ill - EXECUTION
3.02 GENERAL ERECTION REQUIREMENTS
Delete Section 3.02.I in Technical Notes 11B Revised, and replace with the following Section 3. 02.I and add
Section 3. 02.J.
I. Mortar Joint Thickness
1. Lay brick with __________-in. mortar joints, not to exceed 1/2 in. (12.7 mm).
NOTE:
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11c.htm
4 of 5 9/13/2009 12:45 PM
3.02.I Coordinate joint thickness with brick specified in 2.01.A.2.
J. Construction Tolerances:
1. Maximum variation from plumb in vertical lines and surfaces of columns, walls and arrises:
a. 1/4 in. (6.4 mm) in 10 ft. (3 m).
b. 3/8 in. (9.6 mm) in a story height not to exceed 20 ft. (6 m).
c. 1/2 in. (12.7 mm) in 40 ft. (12 m) or more.
2. Maximum variation from plumb for external corners, expansion joints and other conspicuous
lines:
a. 1/4 in. (6.4 mm) in any story or 20 ft. (6 m) maximum.
b. 1/2 in. (12.7 mm) in 40 ft. (12 m) or more.
3. Maximum variation from level of grades for exposed lintels, sills, parapets, horizontal grooves
and other conspicuous lines:
a. 1/4 in. (6.4 mm) in any bay or 20 ft. (6 m).
b. 1/2 in. (12.7 mm) in 40 ft. (12 m) or more.
4. Maximum variation from plan location of related portions of columns, walls and partitions:
a. 1/2 in. (12.7 mm) in any bay or 20 ft. (6 m).
b. 3/4 in. (19 mm) in 40 ft. (12 m) or more.
5. Maximum variation in cross-sectional dimensions of columns and thicknesses of walls from
dimensions shown on drawings:
a. Minus 1/4 in. (6.4 mm).
b. Plus 1/2 in. (12.7 mm).
NOTE:
3.02.J These construction tolerances are for engineered brick masonry only, and are based on actual dimensions. They are
intended for the sole purpose of protecting the structural integrity of engineered brick masonry elements and may not be adequate
for establishing construction tolerances associated with esthetics or visual requirements.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11c.htm
5 of 5 9/13/2009 12:45 PM

Technical Notes 11D - Guide Specifications for Brick Masonry, Part 4 Continued
[Aug. 1972] (Reissued Sept. 1988)
INTRODUCTION
This issue of Technical Notes is a continuation of Technical Notes 11C Revised and contains additional sections and
statements to be incorporated into the "Guide Specifications for Brick Masonry", Technical Notes 11A Revised and
11B Revised. This will make the guide specifications in those Technical Notes suitable for Engineered Brick
Masonry.
The sections contained in these Technical Notes deal primarily with the quality assurance, selection of units, strength
and construction tolerances to provide masonry that meets the minimum design requirements for Engineered Brick
Masonry.
In the construction of Engineered Brick Masonry, quality control may be maintained in either of two ways: (1) by
testing the brick and controlling the mortar which can be done by laboratory tests or by mixing proportions, or (2) by
periodic testing of masonry prisms. This Technical Notes covers quality control by method (2), prism testing.
Technical Notes 11C covers quality control by method (1), testing brick and control of mortar.
When quality control is maintained by prism tests, the brick masonry strength is determined in accordance with
paragraph 4.2.2.1 of the BIA Standard, "Building Code Requirements for Engineered Brick Masonry". Test prisms
are built as the walls are constructed and tested in compression at 7 days or 28 days. If prism tests are used, the
quality control requirements of this Technical Notes should be incorporated into the guide specifications in Technical
Notes 11A Revised and 11B Revised.
All other sections of the Guide Specifications for Brick Masonry (Technical Notes 11A Revised and 11B Revised)
are appropriate for Engineered Brick Masonry.
QUALITY ASSURANCE BASED ON PRISM TESTS
Guide Specification and Notes
PART I - GENERAL
1.02 QUALITY ASSURANCE
Delete sections and notes for 1.02 in Technical Notes 11A Revised, and substitute the following quality
control requirements based on prism tests.
A. Prism Tests:
1. Preconstruction Prisms:
a. Build ten prisms:
(1) Of site materials insofar as possible.
(2) Use brick units similar as to color, texture, raw materials, moisture content and
coring.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11d.htm
1 of 7 9/13/2009 12:46 PM
(3) Under same conditions, insofar as possible, as the structure.
(4) With same bonding, insofar as possible, as for structure.
(5) With same mortar as for the structure.
(6) With same joint thickness.
(7) With same workmanship.
2. Site Control Prisms:
a. Build prisms as required by Section 1.02.D.1 at the direction of the Architect/Engineer.
(1) Of site materials.
(2) Of brick units selected at random from units delivered to the project.
(3) At the project site.
(4) With same bonding, insofar as possible, as the structure.
(5) With site mortar.
(6) With same joint thickness as for the structure.
(7) With same workmanship.
3. Dimensions
a. Minimum height: 12 in. (305 mm).
b. Height-to-thickness ratio (h/t) range:
(1) Minimum: 2
(2) Maximum: 5
4. Mark each specimen for identification.
5. Store prisms:
a. Preconstruction prisms:
(1) In air at temperatures not less than 65
o
; F. (18.3
o
C.).
b. Site control prisms:
(1) At site for not less than 24 hr.
(2) Thereafter, in air at temperatures not less than 65
o
F.(18.
o
C.).
6. Test prisms:
a. Preconstruction prisms:
(1) Five after aging 7 days.
(2) Five after aging 28 days.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11d.htm
2 of 7 9/13/2009 12:46 PM
b. Site control prisms:
(1) After aging 7 days.
c. Cap each prism with suitable material to provide bearing surfaces on each end:
(1) Plane within 0.003 in. (0.076 mm).
(2) Approximately perpendicular to the axis of the prism.
NOTE:
1.02.A.6.c It is suggested that calcined gypsum be used for the capping material.
7. Test in accordance with relevant provisions of ASTM E 447-__________, with the following
provisions:
a. For h/t less than 5, reduce specimen compressive strength by correction factors as
follows:
alnterpolate to obtain intermediate values.
b. If the coefficient of variation of the sample tested exceeds 10%, obtain the
compressive strength by multiplying the average compressive strength of the specimens
by
1 - 1.5 ( V - 0.10 )
100
where v is the coefficient of variation of the sample tested.
B. Brick Tests:
1. Preconstruction Tests:
a. Test five brick for compressive strength to determine acceptability of units for
compliance with specifications.
b. Use brick similar to those selected for use, matching color, texture, raw material,
moisture content and coring.
c. Cost of tests shall be borne by the General Contractor.
2. Test in accordance with ASTM C 67-__________, with the following additional requirements:
a. If the coefficient of variation of the compression samples tested exceeds 12%, obtain
compressive strength by multiplying average compressive strength of specimens by
1 - 1.5 ( V - 0.12 )
100
where v is the coefficient of variation of sample tested.
C. Preconstruction Requirements:
1. Prebid conference:
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11d.htm
3 of 7 9/13/2009 12:46 PM
a. A prebid conference, directed by the Architect/ Engineer, will be held one week prior
to the bid opening to discuss:
(1) Structural concept.
(2) Method and sequence of masonry construction.
(3) Special masonry details.
(4) Quality control requirements.
(5) Material requirements.
(6) Job organization.
(7) Workmanship.
b. Attendance is mandatory for all prospective:
(1) General contractors.
(2) Masonry subcontractors.
(3) Brick suppliers.
NOTE:
1.02.C This requirement may be deleted if not necessary for the project due to bidders being knowledgeable with engineered brick
masonry.
1.02.C.1.b Invitation to attend should be extended to others, such as the inspectors (local building department and other
government agencies) and Owner.
2. Preconstruction Testing and Certification:
a. After award of the contract, the General Contractor shall:
(1) Within 14 days, submit to the Architect/Engineer for approval the name of the
independent laboratory which will perform the site control tests and provide the
certificates and test reports required in Section 1.03.
(2) Upon approval of the laboratory, certificates and test reports, and prior to any
masonry construction, make arrangements for the following tests for each
combination of brick and mortar:
(a) Tests of ten prisms in accordance with preconstruction requirements,
Section 1.02.A.
(b) Test five brick in accordance with Section 1.02.B.
b. Masonry work can begin only after approval of testing.
c. Testing is acceptable if test results indicate that materials meet the minimum
requirements of Part II - Products, or Section 3.02.K.
d. Cost of preconstruction testing shall be borne by the General Contractor.
NOTE:
1.02.C.2 Inspection, laboratory and testing for quality control can be a responsibility of the Structural Engineer. If so, revise section.
3. Preconstruction Conference:
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11d.htm
4 of 7 9/13/2009 12:46 PM
a. A preconstruction conference, directed by the Architect/Engineer, will be held after the
award of the General Contract, but prior to beginning of masonry work to discuss:
(1) Structural concept.
(2) Method and sequence of masonry construction.
(3) Special masonry details.
(4) Standard of workmanship.
(5) Quality control requirements.
(6) Job organization.
b. Attendance is mandatory for:
(1) General contractor job superintendent.
(2) Masonry subcontractor job superintendent.
(3) Masonry subcontractor foreman.
(4) At least two masons.
(5) Authorized representative of the brick supplier.
(6) Mortar material suppliers.
NOTE:
1.02.C.3.b Invitations to attend should be extended to others, such as inspectors (local building department and other government
agencies) and Owner.
D. Job Site Quality Control:
1. Site control prism tests:
a. Use 7-day compressive strength of brick prisms to control quality.
b. Build, store and test prisms in accordance with site control requirements, Section
1.02.A.
c. Build three prisms for each 5000 sq. ft. (465 m
2
) of wall area as directed by the
Architect/Engineer.
**OR**
c. Provide three prisms for each story height.
d. Site control test data shall be acceptable if the 7-day prism strength indicates that the
28-day strength will be equal to or greater than the required minimum ultimate
compressive strength. See Section 3.02.K.
e. Cost of control prisms to be borne by the General Contractor
NOTE:
1.02.D.1.c Select, depending upon whichever is more frequent.
E. Furnish Sample Panel:
1. 4 ft. (1.2 m) long by 3 ft. (1 m) high, of the proposed color range, texture, bond, mortar and
workmanship.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11d.htm
5 of 7 9/13/2009 12:46 PM
2. Erect panel in the presence of the Architect/Engineer before installation of materials.
3. Provide separate panels for each type of brick or mortar.
4. Do not start work until Architect/Engineer has accepted sample panel.
5. Use panel as standard of comparison for all masonry work built of same material.
6. Do not destroy or move panel until work is completed and accepted by Owner.
1.03 SUBMITTALS
Add the following section to 1.03 in Technical Notes 11A Revised:
D. Prism Test Reports:
1. Test reports are to be submitted to Architect/Engineer for approval.
2. Testing and reports are to be completed by an independent laboratory.
3. Test reports shall show:
a. Age at test.
b. Storage conditions.
c. Dimensions (h/t).
d. Compressive strength of individual prisms.
e. Coefficient of variation (v).
f. Ultimate compressive strength of masonry (f 'm ) which has been corrected for the
coefficient of variation and the hit of the prisms tested.
PART II -PRODUCTS
2.01 BRICK
A. Facing Brick:
1. Delete Note and replace with:
NOTE:
2.01.A.1 Grades and Types. Brick subject to the action of weather or soil, but not subject to frost action when permeated with
water, shall be of grade MW or grade SW and where subject to temperature below freezing while in contact with soil shall be
grade SW. Brick used in loadbearing or shear wall construction shall comply with the dimensional and distortion tolerances
specified for type FBS of ASTM C 216-__________. Where such brick do not comply with these tolerance requirements, the
compressive strength of brick masonry shall be determined by prism tests.
PART III -EXECUTION
3.02 GENERAL ERECTION REQUIREMENTS
Delete Section 3.02.I in Technical Notes 11B Revised, and replace with the following Section 3.02.I and add
Sections 3.02.J And 3.02.K
I. Lay brick with __________-in. mortar joints, not to exceed 1/2 in. (12.7 mm).
J. Construction Tolerances:
1. Maximum variation from plumb in vertical lines and surfaces of columns, walls and arrises:
a. 1/4 in. (6.4 mm) in 10 ft. (3 m).
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11d.htm
6 of 7 9/13/2009 12:46 PM
b. 3/8 in. (9.6 mm) in a story height not to exceed 20 ft. (6 m)
c. 1/2 in. (12.7 mm) in 40 ft. (12 m) or more.
2. Maximum variation from plumb for external corners, expansion joints and other conspicuous
lines:
a. 1/4 in. (6.4 mm) in any story or 20 ft. (6 m) maximum.
b. 1/2 in. (12.7 mm) in 40 ft. (12 m) or more.
3. Maximum variation from level of grades for exposed lintels, sills, parapets, horizontal grooves
and other conspicuous lines:
a. 1/4 in. (6.4 mm) in any bay or 20 ft. (6 m).
b. 1/2 in. (12.7 mm) in 40 ft. (12 m) or more.
4. Maximum variation from plan location of related portions of columns, walls and partitions:
a. 1/2 in. (12.7 mm) in any bay or 20 ft. (6 m).
b. 3/4 in. (19 mm) in 40 ft. (12 m) or more.
5. Maximum variation in cross-sectional dimensions of columns and thicknesses of walls from
dimensions shown on drawings:
a. Minus 1/4 in. (6.4 mm).
b. Plus 1/2 in. (12.7 mm).
K. Minimum Ultimate Compressive Strength of Masonry (f'm)__________psi ( __________kgf/cm2).
NOTE:
3.02.K Ultimate compressive strength as determined by the Structural Engineer may vary for different parts and walls of the
building. If so, provide sections to cover all design requirements.
INTRODUCTION http://www.gobrick.com/BIA/technotes/t11d.htm
7 of 7 9/13/2009 12:46 PM

Technical Notes 11E - Guide Specifications for Brick Masonry, Part 5, Mortar and Grout
September 1991
Abstract: This Technical Notes is a guide specification for mortar and grout used in brick masonry. Using this
Technical Notes, a specifier can prepare a job specification for Section 04100. Notes are provided to help the
specifier understand certain decisions that affect the project specifications. The guide specification is in
accordance with the Construction Specifications Institute's (CSI) Masterformat.
Key Words: brick masonry, grout, guide specification, mortar.
INTRODUCTION
This Technical Notes is a continuation of Technical Notes 11 Series on "Guide Specifications for Brick
Masonry" and contains the requirements for mortar and grout for brick masonry. This Technical Notes is
appropriate for both empirically designed and rationally designed brick masonry.
The guide specification in this Technical Notes is in accordance with the Construction Specifications Institute's
Masterformat and is based on the requirements of BIA M1 Standard Specification for Portland Cement-Lime
Mortar for Brick Masonry contained in Technical Notes 8A and ASTM C 270 Mortar for Unit Masonry. Mortar
conforming to the requirements of BIA M1 will meet all of the requirements of portland cement-lime mortars of
ASTM C 270. A complete discussion of mortar properties is contained in Technical Notes 8.

GENERAL
Mortar requirements differ from concrete requirements because the primary function of mortar is to bond
masonry units into an integral element. The basic mortar ingredients include portland cement, hydrated lime,
sand and water. Masonry cements, proprietary mortar mixes, are sometimes used to replace portland
cement and hydrated lime or combined with portland cement to make mortar. BIA M1, ASTM C 270 and
ASTM C 1142 Ready-Mixed Mortar for Unit Masonry are the recommended standards for mortar to be used
with brick masonry.
Grout is different from both concrete and mortar. Grout is a high slump mixture used to fill cells of masonry
units or between wythes of masonry to resist stresses and develop bond with reinforcement. Grout can
consist of portland cement, hydrated lime, fine or coarse aggregate and water. Grout should be specified by
ASTM C 476 Grout for Masonry.

RECOMMENDED MORTAR USES
Selection of a particular mortar type is usually a function of the needs of the finished structural element. For
example, where high wind loads are expected, high lateral strength may be required and, hence, mortar with
high flexural bond strength should be considered. For loadbearing walls and reinforced brick masonry, high
compressive strength may be the governing factor. In some projects considerations of durability, color,
flexibility, etc., may be of most concern. No single type of mortar is best for all purposes. Factors which
improve one property of mortar may do so at the expense of others. For this reason, when selecting a
mortar, evaluate properties of each mortar type and choose that type and materials which will best meet all
requirements. Technical Notes 8B discusses the selection of mortar types in depth. The following sections
Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
1 of 8 9/13/2009 12:46 PM
briefly discuss selection of mortar.

Type N Mortar
Type N mortar is suitable for general use in exposed masonry above grade. It is recommended for use in
parapet walls, chimneys and exterior walls when subject to severe exposure.

Type S Mortar
Type S mortar is recommended for use in reinforced and unreinforced masonry where higher flexural
strengths than Type N are required.

Type M Mortar
Type M mortar is recommended for use in masonry in contact with earth such as foundations, retaining walls,
paving, sewers and manholes, and in reinforced masonry.

Type O Mortar
Type O mortar is suitable for interior use in non-loadbearing applications.

SPECIFYING MORTAR
Mortars are specified in one of two ways: proportions or properties, but not both. Mortar prepared by the
proportion requirements should not be compared to mortar prepared by the property requirements.
The proportion specification requires that mortar materials be mixed according to given volumetric proportions
or weight. If mortar is specified by this method, no laboratory testing of the mortar is required.
If mortar is specified by the property specifications, compressive strength, water retention and air content
tests must be performed on mortar mixed in the laboratory. Field mortar is then mixed to the proportions
selected from these laboratory tests.
When neither proportion nor property is specified, the proportion specifications govern.

SPECIFYING GROUT
Grout is specified by proportion using ASTM C 476. Either fine or fine and coarse aggregate can be used in
grout. Experience has shown that grout mixed to the specified proportions performs well with brick masonry
since the grout compressive strength closely matches the compressive strength of the brick masonry. Grout
must have adequate compressive strength, bonding with reinforcement and for embedment of anchor bolts.
There is usually no need to specify compressive strength of grout unless required by design. Grout strength
can be verified by field testing using ASTM C 1019. Slump of the grout is usually specified to be between 8
and 11 in. (203.2 and 279.4 mm).

Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
2 of 8 9/13/2009 12:46 PM
CONCLUSION
This Technical Notes is a guide specification for mortar and grout for brick masonry. Notes are provided to
assist in editing the specification.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Institute of America. The information contained herein must be
used in conjunction with good technical judgment and a basic understanding of the properties of masonry.
Final decisions on the use of information contained in this Technical Notes are not within the purview of the
Brick Institute of America and must rest with the project architect, engineer, owner or all.

04100 MORTAR AND GROUT

Guide Specification & Notes
PART 1 GENERAL
1.01 SECTION INCLUDES
A. Mortar for masonry.
B. Grout for masonry.
C. Repointing mortar.
1.02 RELATED SECTIONS
A Concrete: Section 03__________.
B. Masonry: Section 04__________.
C. Masonry Cleaning: Section 04500.
D. Structural Metal Framing: Section 05100.
E. Rough Carpentry: Section 06100.
F. Waterproofing: Section 07100.

NOTE:
1.02 Some of these broadscope sections may not be included. Other narrow scope sections under these broadscope sections may be
added.

1.03 PRODUCTS INSTALLED BUT NOT FURNISHED UNDER THIS SECTION
A. Reinforcing Steel: Section 03210.
B. Metal Accessories: Section 04150.
C. Masonry Units: Section O4200.
Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
3 of 8 9/13/2009 12:46 PM
D. Flashing and Sheet Metal: Section 07600.
1.04 REFERENCES
A. ACI 530.1/ASCE 6-__________ - Specifications for Masonry Structures.
B. ASTM C 91 - __________, [UBC Standard No. 24-16] - Masonry Cement.
C. ASTM C 144 - __________ - Aggregate for Masonry Mortar.
D. ASTM C 150 - __________, [UBC Standard No. 26-1] - Portland Cement.
E. ASTM C 207-__________, [UBC Standard No. 24-18] - Hydrated Lime for Masonry
Purposes.
F. ASTM C 270-__________, [UBC Standard No. 24-20] - Mortar for Unit Masonry.
G. ASTM C 404-__________ - Aggregates for Masonry Grout.
H. ASTM C 476-__________, [UBC Standard No. 24-29] - Grout for Masonry.
I. ASTM C 780-__________ - Preconstruction and Construction Evaluation of Mortars for Plain
and Reinforced Unit Masonry.
J. ASTM C 979-__________ - Pigments for Integrally Colored Concrete.
K. ASTM C 1019-__________, [UBC Standard No. 24-28] - Sampling and Testing Grout.
L. ASTM C 1142-__________ - Ready-Mixed Mortar for Unit Masonry.
M. BIA Technical Notes 8A - "Specifications for Portland Cement-Lime Mortar for Brick
Masonry" BIA M1-88).

NOTE:
1.04 The applicable date for each reference can be given here or in Section 01090-Reference Standards. Alternate standards are
given for Uniform Building Code specifications.

1.05 SUBMITTALS
A. Submit data indicating proportion or property specifications used for mortar.
B. Submit test reports for mortar materials indicating conformance to ASTM C 270 [UBC
Standard No. 24-20] property specifications. Report proportions resulting from laboratory
testing used to select mortar mix.
C. Submit test reports for field sampling and testing mortar in conformance to ASTM C 780.
D. Submit test reports for grout materials indicating conformance to ASTM C 476 [UBC
Standard No. 24-29].
E. Submit test reports for field sampling and testing grout in conformance to ASTM C 1019
[UBC Standard No. 24-28].
F. Samples: Submit two ribbons of mortar for conformance with color.
Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
4 of 8 9/13/2009 12:46 PM

NOTE:
1.05.A ASTM C 270 and UBC Standard No. 24-20 require that mortar be specified by proportion or property, not both.
1.05.B ASTM C 270 and UBC Standard No. 24-20 require comparison of laboratory prepared mortars to establish
proportions for field-mixed mortar when the property specifications are used.
1.05.C ASTM C 780 allows preconstruction evaluation of mortar and comparison of field prepared mortars. Mortar
prepared in the field should not be compared to values found in the property specifications of ASTM C 270 or UBC
Standard No. 24-20.
1.05.E ASTM C 1019 or UBC Standard No. 24-28 is used to test uniformity of grout preparation during construction.

1.06 DELIVERY, STORAGE AND HANDLING
A. Store materials in dry location and protected from dampness and freezing.
B. Stockpile and handle aggregates to prevent contamination from foreign materials.
1.07 ENVIRONMENTAL REQUIREMENTS
A. Follow requirements for cold and hot weather construction in ACI 530.1/ASCE 6 [Uniform
Building Code].
PART 2 PRODUCTS
2.01 MORTAR MATERIALS
A. Cementitious materials:
1. Portland Cement: ASTM C 150 [UBC Standard No. 26-1], Type__________.
2. Hydrated Lime: ASTM C 207 [UBC Standard No. 24-18], Type S__________.
3. Masonry Cements: ASTM C 91 [UBC Standard No. 24-16], Type__________.
B. Sand: ASTM C 144.
C. Admixtures:
1. No air-entraining admixtures or material containing air-entraining admixtures.
2. No antifreeze compounds shall be added to mortar.
3. No admixtures containing chlorides shall be added to mortar.
D. Water: Clean and potable.
E. Mortar pigment:
1. ASTM C 979: Pigment shall not exceed 10% of the weight of portland cement.
2. Carbon black shall not exceed 2% of the weight of portland cement.

NOTE:
2.01.A Allowable flexural tensile stresses for masonry built with air-entrained portland cement-lime mortars, or with
Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
5 of 8 9/13/2009 12:46 PM
masonry cement mortars are lower than those built with portland cement-lime mortars.
2.01.A.1 Only Types I, II or III.
2.01.A.3 Types M, S, or N.
2.01.B Sand not conforming to ASTM C 144 must have mortar meet the property specification requirements.
2.01.E Limits on amount of pigments should be halved when using masonry cement mortars.

2.02 GROUT MATERIALS
A. Cementitious materials:
1. Portland Cement: ASTM C 150 [UBC Standard No. 26-1], Type__________.
2. Hydrated Lime: ASTM C 207 [UBC Standard No. 24-18], Type S__________.
B. Aggregates:
1. Fine aggregate: ASTM C404.
2. Coarse aggregate: ASTM C 404.
C. Water: Clean and potable.
D. Admixtures.

NOTE:
2.02.A.1 Only Types 1, II or III.
2.02.D Grout admixtures are used to decrease grout shrinkage, aid in pumping grout, or for other reasons. The use of
such admixtures should not adversely affect the performance of the grout.

2.03 MORTAR AND GROUT MIXES
A. Mortar - ASTM C 270 [UBC Standard No. 24-20] or BIA M1:
1. Type based on proportion specifications.
**OR**
1. Type__________based on property specifications to achieve __________ psi
strength, __________% air content, __________% water retention.

NOTE:
2.03.A Mortar mixes can be specified separately by specifying ASTM C 270, UBC Standard No. 24-20 or BIA
M1; or by specifying ASTM C 1142 alone.
2.03.A.1 Type M, S, N or O depending on design requirements.

Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
6 of 8 9/13/2009 12:46 PM
**OR**
A. Mortar - ASTM C 1142: Type__________.

NOTE:
2.03.A Ready mixed mortar can be mixed on-site or off-site. Types RM, RS, RN or RO.

B. Grout: ASTM C 476 [UBC Standard No. 24-29]
1. Fine grout.
2. Coarse grout.
3. Slump: __________inches (__________mm).

NOTE:
2.03.B The use of fine or coarse grout is based on the size of the grout space and the height of the grout pour.
2.03.B.3 Specify desired slump between 8 and 11 inches (203.2 and 279.4 mm). Higher slump is necessary for smaller
dimensioned grout spaces and with higher unit/grout volume ratios.

PART 3 EXECUTION
3.01 FIELD MORTAR MIXING
A. All cementitious materials and aggregate shall be mixed between 3 and 5 min. in a
mechanical batch mixer with the maximum amount of water to produce a workable consistency.
B. Control batching procedure to ensure proper proportions by measuring materials by volume.
Sand measurement by shovel count shall not be permitted.
C. If water is lost by evaporation within 2 1/2 hours after initial mixing, retemper with water.
D. Discard all mortar which is more than 2 1/2 hours old.

NOTE:
3.01 ASTM C 270 or BIA M1 can be referenced for field mortar mixing.
3.01.B Materials can be specified by weight if volume proportions are converted to weight proportions
3.02 FIELD GROUT MIXING
A. Control batching procedure to ensure proper proportions by measuring materials by volume.

NOTE:
3.02 ASTM C 476 can be referenced for field grout mixing.
3.02.A Materials can be measured by weight if volume proportions are converted to weight proportions.
Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
7 of 8 9/13/2009 12:46 PM

3.03 INSTALLATION
A. Install mortar and grout in accordance with ACI 530.1/ASCE 6.
3.04 REPOINTING MORTAR
A. Use mortar materials listed in 2.01, Type N.
B. Prehydrate the mortar by the following method. Mix dry ingredients together. Then add only
enough water to make a damp, stiff mix which will retain its form when pressed in a ball. After 1
to 2 hours, add sufficient water to bring it to the proper consistency.

NOTE:
3.04.A If materials and proportions of existing mortar are known, use those instead of Type N mortar if the existing mortar
provided sufficient durability.
Abstract: This Technical Notes is a guide specification for mortar and grou... http://www.gobrick.com/BIA/technotes/t11e.htm
8 of 8 9/13/2009 12:46 PM
TECHNICAL NOTES on Brick Construction
13
December
2005
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
INTRODUCTION
Glazed brick can be used in both interior and exterior
applications, as accent brick or as the field brick covering
the entire facade, as shown in Photos 1 and 2. Glazed
units have been integral parts of buildings for decades
and have performed well under all climatic conditions.
Glazed brick are often selected for use because of the
many characteristics that make them distinct among brick
products. One of these is the wide variety of colors that
are not available in standard brick production. These may
be applied to special shapes or brick of different sizes
to further enhance visual interest. It is even possible to
apply multiple glazes to a single brick unit, as shown in
Photo 3. Glazes may be clear, translucent or opaque, and
are available in almost any color with a glossy, satin or
matte finish.
Glazed brick also provide an impervious surface that is
extremely durable and resistant to staining which results
in easy maintenance. Resistance to scratching and abra-
sion, as well as fire resistance of the glaze, also enhance
the durability of glazed brick units.
Successful performance of exterior glazed brick walls can
Page 1 of 6
Photo 1
Glazed Brick Used for Entire Facade
General:
Consult manufacturers for assistance with special shapes
and to determine the property requirements of double-fired
glazed units
Specify surfaces other than stretcher faces to be glazed
Wall System Design:
Use vented drainage walls to ensure the most rapid
removal of moisture that enters the wall
Specify concave, "V", or grapevine mortar joint profiles
Air Space:
2 in. (51 mm) minimum air space recommended, required
to be no less than 1 in. (25.4 mm)
When prescriptive anchor spacings are used, air space
may not exceed 4
1
/2 in. (114 mm)
Flashing:
Extend flashing to the face of the brickwork or beyond
Install at all horizontal interruptions to the air space
Turn flashing ends into head joint a minimum of 1 in. (25.4
mm) to form end dam
Weeps:
Open head joint weeps spaced no more than 24 in. (610
mm) o.c. recommended
Most building codes permit weeps no less than
3
/16 in. (4.8
mm) diameter and spaced no more than 33 in. (838 mm)
o.c.
Wick and tube weep spacing recommended at no more
than 16 in. (406 mm) o.c.
Vents:
Place vents at the tops of walls and below horizontal inter-
ruptions such as shelf angles and flashing locations
Use open head joint weeps as vents; If weeps are not
open head joints, vents are needed one or two courses
above weeps
Space vents 24 to 48 in. (610 mm to 1.22 m) o.c.
Stagger vents in relation to overlying weeps
Ceramic Glazed Brick Exterior Walls
Abstract: Buildings and other structures employ glazed brick in a variety of uses, from decorative bands to entire wall sys-
tems. Due to the imperviousness of its ceramic glazed surface, a vented air space is recommended behind the glazed brick
wythe. Proper wall design, detailing and material selection, along with quality construction will result in attractive glazed brick
applications exhibiting durability, structural stability and virtually maintenance free aesthetics.
Key Words: ceramic, condensation, drainage, expansion joints, flashing, glaze, moisture, movement, vents, weeps.
SUMMARY OF RECOMMENDATIONS
www.gobrick.com | Brick Industry Association | TN 13 | Ceramic Glazed Brick Exterior Walls | Page 2 of 6
be ensured through the use of vented drainage wall sys-
tems that allow water to evaporate from the unglazed back
surface of the brick as well as through mortar joints. Other
wall systems utilizing glazed brick exist and are serviceable,
but are outside the scope of this Technical Note.
Attention must be given to proper material selection, detail-
ing and construction practices to ensure successful perfor-
mance. Proper design and installation of exterior glazed
brick walls allows water drainage and minimizes the pos-
sibility of water being trapped behind the glazed surface
which may lead to efflorescence or spalling. As in all brick
construction, stress concentrations due to restrained move-
ment must also be minimized. This Technical Note address-
es these concerns and offers recommendations to ensure
proper performance.
WALL SYSTEM DESIGN
Moisture Resistance
It is recommended that exterior glazed brick walls be
designed to drain water that enters the wall system and
to allow moisture from wind-driven rain or condensation
to evaporate from the behind the brickwork. Therefore, a
vented drainage wall system is recommended. Drainage
walls must be designed, detailed and constructed properly
to accommodate the flow of water collected within the wall.
Common examples of drainage walls include brick and
block cavity walls, brick veneer, and rain screen walls. See
Technical Note 7 for more information on water penetration
resistance.
In drainage wall design, penetrant water is intended to drain
down the back of the brick, which is separated from interior
wall elements by an air space. While a minimum 1 in. (25.4
mm) air space is required, 2 in. (51 mm) is recommended.
Flashing and weeps are needed at horizontal interruptions
in the air space to collect water and direct it out of the wall
system, refer to Figures 1a, 1b and 1c. They are typically
provided above lintels and shelf angles, beneath sills, under
copings and masonry or stone caps, and at the wall base.
Discontinuous flashing, such as at window sills and loose
lintels, should be constructed with end dams to ensure that
collected water is directed out of the brickwork. End dams
are also recommended where stepped flashings are used,
such as at sloped grades, above arches, and above sloped
roofs. Weeps must be provided in head joints directly above
the flashing. Open head joint weeps are recommended
with a spacing of no more than 24 in. (610 mm) on center.
Spacing of wick and tube weeps is recommended at no
more than 16 in. (406 mm) on center. Most building codes
require weeps to have a minimum diameter of
3
/16 in. and
permit weeps to be spaced up to 33 inches (838 mm) on
center.
Mortar joints affect the moisture resistance of brickwork
since they can account for up to twenty percent of the
brickwork surface. Selecting mortar joint profiles that are
Figure 1
Glazed Brick Wall Sections
b) Glazed Brick at Shelf Angle
a) Glazed Brick at Top of Wall
c) Glazed Brick at Foundation
www.gobrick.com | Brick Industry Association | TN 13 | Ceramic Glazed Brick Exterior Walls | Page 3 of 6
most resistant to water penetration and cover the bed surface of the brick unit further minimize water intrusion and
the possibility of water being trapped behind the glazed surface. Thus, concave, "V" and grapevine tooled mortar
joints are recommended.
Vents. Vents placed at the top of glazed brick wall segments, as shown in Figures 1a and 1b, encourage air cir-
culation and help to dissipate moisture within the air space. These vents should be located a course or two below
horizontal interruptions of the air space such as shelf angles and flashing locations and be spaced 24 to 48 inches
(610 mm to 1.22 m) on center. The vent should be an open head joint and may include a weep vent or louvered
insert to deter insect access. In multi-story construction, the horizontal placement of vents and weeps should be
staggered, or louvered inserts may be utilized to prevent draining water from entering vents. When wick or tube
weeps are used at the base of a wall, additional vents should be added no more than two courses above weeps
to best assure air movement through the air space.
Caps, Copings and Sills. Glazed brick should not be used in locations where they are likely to be saturated.
Rowlock courses of brick used as caps, copings or sills are vulnerable to water penetration, especially when the
slope is not sufficient to drain water away quickly. Therefore, glazed brick should be avoided in favor of concrete,
stone, or metal elements that reduce the potential for water penetration at these locations. More information about
caps and copings can be found in Technical Note 36A.
Movement
Brick masonry walls expand or contract with changes in temperature and moisture content. Brick expansion and
other building movements are typically accommodated by expansion joints, placed vertically and horizontally,
which divide the wall into rectangular segments and limit cumulative movement. Segment lengths and heights will
vary with the building and wall design; however, expansion joints must be placed beneath all shelf angles.
Expansion joints are typically needed near corners, at changes in wall height, at offsets in the wall plane and
at the ends of elements rigidly anchored to the backing or structure. The segments formed by expansion joints
should be limited to a maximum length of approximately 25 feet (7.62 m). Segment lengths in building parapets
should be limited to approximately 15 feet (4.57 m). The building geometry will also dictate locations for vertical
expansion joints in glazed brick walls. Vertical expansion joints should extend full height from the foundation to the
roof, or between locations of horizontal support.
Brick masonry expansion joints must be formed with
highly compressible materials and be free of mortar
and obstructions. Expansion joints are typically
3
/8 in.
(9.5 mm) or
1
/2 in. (12.7 mm) wide with a foam backer
rod and elastic sealant at the wall face to prevent
air and water penetration from the exterior. See the
Technical Notes 18 Series for more discussion regard-
ing building movements and expansion joints.
Structural Design
Glazed brick can be used in loadbearing, cavity or
veneer walls and should be designed in accordance
with the appropriate chapter of ACI 530/ ASCE 5/
TMS 402, Building Code Requirements for Masonry
Structures, also known as the Masonry Standards
J oint Committee (MSJ C) Code. [Ref. 11] Design can
be based on either the requirements for veneer in
Chapter 6, or the rational design approach of Chapters
2 or 3. In either case, the preferred design should
be based on minimizing the potential for cracking of
the glazed brick wythe under applied loading. More
detailed information on structural design of veneer
walls and cavity walls are included in the Technical
Notes 28 Series and Technical Notes 21 Series,
respectively.
Photo 2
Glazed Brick Used as Accents
www.gobrick.com | Brick Industry Association | TN 13 | Ceramic Glazed Brick Exterior Walls | Page 4 of 6
The MSJ C Code and Specification also contain material, size and spacing requirements for wall ties in cavity
walls and anchors in veneer walls. Anchor and wall tie spacing in the MSJ C Code depends on the wall design
method and anchor type. See the Technical Notes 28 Series for anchor spacing in veneer walls and the Technical
Notes 21 Series for spacing of wall ties in cavity walls.
MATERIALS
Glazed Brick
Two methods are used to apply glazes to brick bodies: a
single-firing process and a double-firing process. In the
single-firing process, the glaze is applied to the unfired
brick body and is fused to the body when fired. In the
double-firing process, brick that have been fired previ-
ously have a glaze applied and are fired again to fuse
the glaze onto the brick. For some glazes with certain
compositions or color pigments, double-firing is neces-
sary to ensure the proper firing of the brick body at a
higher temperature and the proper color and finish of
the glaze at a lower temperature. Both methods result
in quality glazed brick. ASTM standards for glazed brick
include requirements for both brick body and the glazes.
Body Properties. The physical properties of brick vary
depending on raw material, method of forming, and the
degree of firing. ASTM standards establish indicators
of durability based only upon physical property require-
ments that correlate with freeze thaw testing.
Single-fired glazed brick must meet the requirements of ASTM C 1405, Standard Specification for Glazed Brick
(Single Fired, Brick Units). This standard establishes minimum criteria for the glaze as well as for solid and hollow
brick bodies. Single-fired glazed brick intended for exterior exposure should meet the property requirements for
Class Exterior. These include prescriptive requirements for minimum compressive strength, maximum cold water
absorption and maximum saturation coefficient as shown in Table 1. The saturation coefficient requirement does
not apply provided the average compressive strength of a random sample of five brick equals or exceeds 8000 psi
(55.2 MPa) with no individual strength less than 7500 psi (51.8 MPa) and the 24 hr cold water absorption of each
unit does not exceed 6.0%. The saturation coefficient and water absorption requirements do not apply if a sample
of five brick pass the freezing and thawing test in ASTM C 67.
Currently, proper specification of double-fired brick units requires the designer to adopt two separate ASTM
standards: ASTM C 126, Standard Specification for Ceramic Glazed Structural Clay Facing Tile, Facing
Brick, and Solid Masonry Units to cover applicable properties of the ceramic glaze finish, and ASTM C 216,
Standard Specification for Facing Brick (Solid Masonry Units Made from Clay or Shale); ASTM C 652, Standard
Specification for Hollow Brick (Hollow Masonry Units Made from Clay or Shale) or ASTM C 1088, Standard
Specification for Thin Veneer Brick Units Made From Clay or Shale to cover requirements for the brick body.
Photo 3
Multiple Glazes, Shapes and Sizes Add Variety
TABLE 1
ASTM C 1405 Physical Requirements of Clay Bodies for Glazed Units
Designation
Maximum
Water
Absorption by
24-hr Cold
1
, %
Average of 5
Brick
Individual Individual Average of 5
Brick
Individual
Class Exterior 6000 (41.4) 5000 (34.8) 7.0 0.78 0.80
Class Interior 3000 (20.7) 2500 (17.2)
2. The saturation coefficient is the ratio of absorption by 24 hr submersion in cold water to that after 5 hr submersion in boiling water.
1. The saturation coefficient and/or cold water absorption requirement(s) may not apply when other criteria are met. See Body Properties text
for more information.
Minimum Compressive Strength,
psi (MPa), Gross Area
Maximum Saturation Coefficient
1,2
www.gobrick.com | Brick Industry Association | TN 13 | Ceramic Glazed Brick Exterior Walls | Page 5 of 6
Specifying glazed brick in this manner addresses material concerns necessary for exterior use.
Conformity to the requirements of the appropriate standard is an indication of the ability of the brick to withstand
internal stresses caused by freezing of moisture within its body.
Glaze Properties. Ceramic glazes produce a durable, aesthetically pleasing surface feature on brick. ASTM
standards C 126 and C 1405 set minimum property requirements for glaze finishes. Both standards cover finish
requirements for glazes applied to the body before the brick unit is fired (single-firing). ASTM C 126 also covers
double-fired glazed brick when the glaze is fused to the brick unit at temperatures over 1500 F (816 C).
The ASTM C 126 and C 1405 property requirements for glaze finishes are listed below. One or more of the prop-
erties listed may not be applicable to some special decorative and textured glazes. Manufacturers should be con-
sulted for the property requirements of these.
Imperviousness - A wet cloth and water must be able to remove permanent blue black ink that has been
allowed to dwell on the finish for five minutes, with no stain remaining on or beneath the surface.
Hardness and Abrasion Resistance - Glazes must be rated above five on the Mohs hardness scale and
resist scratching from ordinary glass or steel in addition to being subjected to an abrasion test.
Resistance to Crazing - Glazes may not craze, spall or crack when subjected to one cycle of autoclaving.
Fire Resistance - The brick body and glaze are rated "noncombustible" and must withstand temperatures
up to 1900 F (878 C) without melting, distorting or releasing toxic fumes. They must also measure "0"
flame spread, fuel contribution and smoke density when tested in accordance ASTM E 84.
Resistance to Fading/Chemical Resistance - Glaze colors must not change from the approved sample
after a 3 hr submersion in prescribed acidic and basic solutions.
Opacity - When specified, ink applied to the brick body must not be visible through the glaze.
Appearance. Aesthetic characteristics of glazed brick are specified by grade and type in ASTM C 126 and ASTM
C 1405. The two grades, S and SS, limit dimensional variations, distortion and set squareness criteria of the
exposed face. The requirements of Grade S are utilized in most glazed brick projects. Where a higher degree of
precision is necessary the more precise Grade SS units should be specified.
The type of a glazed brick indicates the number of glazed faces on the brick. Type I units have one glazed face,
and Type II units are glazed on two opposite faces. Unless specified otherwise, the stretcher face (or exposed
face of shapes) is coated with the glaze finish. When glazed surfaces other than those identified by Type I or Type
II are required, the additional surface(s) should be specified. Brick which will be exposed on their ends, or on their
bed surfaces as in recessed courses or quoins, should also be explicitly specified. Consultation with the brick
manufacturer is advised to determine if the proposed glazed brick can be made.
Mortar
Mortar should conform to ASTM C 270 Standard Specification for Mortar for Unit Masonry. Type N mortar is typi-
cally recommended for exterior walls above grade. Type S may provide better flexural bond strength to brick hav-
ing initial rates of absorption (IRA) under 5 g/min30 in.
2
(5 g/min194 cm
2
). Use of admixtures and additives is not
usually recommended unless their effect on the masonry, masonry units and items embedded in the brickwork is
known, and they do not detrimentally affect plastic or hardened mortar. ASTM C 1384, Standard Specification for
Admixtures for Masonry Mortars provides methods to evaluate the effect of admixtures on mortar properties. See
Technical Note 8B for more information on mortar selection.
Anchors and Wall Ties
Acceptable connectors for anchored masonry veneer and cavity wall applications include adjustable ties, unit ties,
and ladder-type or tab-type joint reinforcement. Connectors may be of stainless steel conforming to ASTM A 580,
carbon steel protected from corrosion by hot-dipped galvanizing conforming to ASTM A 153 or epoxy coatings
conforming to ASTM A 884, Class A, Type 1, minimum 7 mils (175 m) for joint reinforcement and ASTM A 899,
Class C - 20 mils (508 m) for wire items. For more information on selection of anchors and ties see Technical
Note 44B.
www.gobrick.com | Brick Industry Association | TN 13 | Ceramic Glazed Brick Exterior Walls | Page 6 of 6
Flashing
Flashing materials should be sufficiently tough and flexible to resist puncture and cracking. In addition, flashing
should not degrade when exposed to ultraviolet light or when placed in contact with metal, mortar or sealants.
Flashing materials are generally formed from sheet metals, bituminous-coated membranes, rubber, or combina-
tions thereof. The selection is largely determined by cost and suitability. Asphalt-impregnated felt is not acceptable
as a flashing material. The cost of flashing materials varies widely. It is suggested, however, that only superior
materials be selected, since replacement in the event of failure is difficult. See Technical Note 7A for a more
detailed discussion of flashing materials.
SUMMARY
As with any brick masonry wall system, performance is the result of successful material selection, design detail-
ing and construction practices. While the recommendations contained in this Technical Note are similar to those
for non-glazed brick, it is important to consider the imperviousness of the glazed brick surface. Consequently,
attention to each aspect of design and construction is essential to obtain the intended service life of the structure.
Therefore, some glazed brick designs may entail more thorough detailing, as they may be less forgiving of detail-
ing and construction deficiencies than non-glazed brick. Ceramic glazed brick can present a bright, bold, colorful
statement with a durable brick surface.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment and a
basic understanding of the properties of brick masonry. Final decisions on the use of the informa-
tion contained in this Technical Note are not within the purview of the Brick Industry Association
and must rest with the project architect, engineer and owner.
REFERENCES
1. Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2005:
Volume 1.03 - A 580/A 580M, Standard Specification for Stainless Steel Wire
A 899, Standard Specification for Steel Wire, Epoxy-Coated
Volume 1.04 - A 884/A 884M, Standard Specification for Epoxy-Coated Steel Wire and Welded Wire
Reinforcement
Volume 1.06 - A 153/A 153M, Standard Specification for Zinc Coating (Hot-Dip) on Iron and Steel
Hardware
Volume 4.05 - C 126, Standard Specification for Ceramic Glazed Structural Clay Facing Tile, Facing
Brick, and Solid Masonry Units
C 216, Standard Specification for Facing Brick (Solid Masonry Units Made from Clay or
Shale)
C 652, Standard Specification for Hollow Brick (Hollow Masonry Units Made From Clay or
Shale)
C 1088, Standard Specification for Thin Veneer Brick Units Made From Clay or Shale
C 1384, Standard Specification for Admixtures for Masonry Mortars
C 1405, Standard Specification for Glazed Brick (Single Fired, Brick Units)
2. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
3. Specification for Masonry Structures (ACI 530.1-05/ASCE 6-05/TMS 602-05), The Masonry Society,
Boulder, CO, 2005.
2007 Brick Industry Association, Reston, Virginia Page 1 of 19
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
14
March
2007
Paving Systems Using Clay Pavers
Abstract: This Technical Note presents an overview of paving systems made with clay pavers used in pedestrian and
vehicular, residential and nonresidential projects. Commonly used systems that include clay pavers are discussed, and
guidance is given in selecting the appropriate clay paver, setting bed and base. Site conditions and project requirements that
may affect choice are discussed, including subgrade soil, pedestrian and vehicular traffic, accessibility requirements, drainage,
and appearance.
Key Words: base, design, flexible, mortared paving, mortarless paving, paving, permeable paving, rigid, subbase.
Select Paving System
Use Table 1 to determine paving system based on
application
Use Table 2 to evaluate clay paving systems based on
their general advantages and disadvantages
Use Table 3 to verify choice of the clay paving system for
specific site conditions and project requirements
Design Paving System
Use Technical Note 14 for design considerations and
general specification of clay pavers, base and subbase
Use appropriate Technical Note in this series to provide
design and construction information specific to the setting
bed of the paving system selected as follows:
- Sand Setting Bed Technical Note 14A
- Bituminous Setting Bed Technical Note 14B
- Mortar Setting Bed Technical Note 14C
Use a design professional as necessary to verify
suitability of a paving system design
INTRODUCTION
Technical Note 14 is the first in a series discussing the use of clay pavers for pedestrian and vehicular, residential
and nonresidential applications (see Photo 1). It provides guidance in selecting a paving system (see Figure 1) and
the appropriate clay paver, setting bed and base. Once these are determined, other Technical Notes in this series
provide additional information specific to the setting bed chosen, including common construction for particular
applications, typical details, installation practices and maintenance.
Paving systems exposed to more than 251 daily equivalent single axle loads (ESAL) from trucks or combination
vehicles having three or more loaded axles are considered heavy duty vehicular applications. Such paving
SUMMARY OF RECOMMENDATIONS:
Photo 1
Pedestrian Plaza with Clay Pavers
Figure 1
Typical Pavement Section
Geotextile (If Required)
Compacted Subgrade
Compacted Base
Setting Bed
Wearing Course
J oint with Sand or Mortar
Clay Pavers
Specified Edge Restraint
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 2 of 19
systems are beyond the scope of this Technical Note series. For more information on paving systems for heavy
duty vehicular use, refer to Flexible Vehicular Brick Paving A Heavy Duty Applications Guide [Ref. 14].
Table 1 lists acceptable paving systems for typical paving applications. Table 2 is a comparison of paving systems
listing the general advantages and disadvantages for each system. Table 3 indicates which paving systems are
appropriate for specific site conditions and project requirements.
TABLE 1
Acceptable Paving Systems
NOTES: KEY:
1. For a paving system that uses existing asphalt or concrete as base, A =Acceptable
verify that the condition of the base is acceptable. NA =Not Acceptable
2. For a definition of high volume of heavy vehicles, see Introduction.
3. For these applications, a design professional should design the
paving system.
Application
Typical
Examples
Sand Setting Bed
Bituminous
Setting Bed
Mortar Setting Bed
Bonded Unbonded
Aggregate
Base
Asphalt
Base
1
Cement-
Treated
Aggregate
Base
Concrete
Base
1
Asphalt
Base
1
Concrete
Base
1
Concrete
Base
1
Concrete
Base
1
Residential
Patios and
walks on
property of a
one- or two-
family house
or townhouse
A A A A A A A A
Driveways on
property of a
one- or two-
family house
or townhouse
A A A A A A A NA
Commercial/
Pedestrian
Public plazas,
courtyards or
sidewalks
A A A A A A A A
Light Duty
Vehicular
3
Paving with
low volume
2

of heavy
vehicles such
as streets,
parking
areas, turn-
arounds or
passenger
drop-offs
A A A A A A A NA
Heavy Duty
Vehicular
3
Paving with a
high volume
2

of heavy
vehicles such
as streets,
commercial
driveways or
crosswalks
across them
Refer to Flexible Vehicular Brick Paving - A Heavy Duty Applications Guide
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 3 of 19
Clay Pavers On: Advantages Disadvantages
Sand Setting Bed
on Aggregate Base
Most durable
Cost-effective
Easy access to repair underground utilities
Good as overlay to existing asphalt or concrete
pavement
Allows use of semi-skilled labor
Can be designed as a permeable pavement
Intensive cleaning may erode joint sand
May require a thicker base
Sand Setting Bed
on Asphalt Base
Good as overlay to existing asphalt pavement Intensive cleaning may erode joint sand
Sand Setting Bed
on Cement-Treated
Aggregate Base
Good over poor soils or in small, confined areas
Good as overlay to existing concrete pavement
Intensive cleaning may erode joint sand
Sand Setting Bed
on Concrete Base
Good over poor soils or in small, confined areas
Good as overlay to existing concrete pavement
Intensive cleaning may erode joint sand
Requires good drainage above base
Susceptible to greater offset with subgrade
movement
Bituminous Setting
Bed on Asphalt
Base
Reduced horizontal movement and uplift
Enhanced water penetration resistance
Repairs are more difficult and expensive
Little tolerance for paver thickness variations
or inaccurate base elevations
Bituminous Setting
Bed on Concrete
Base
Reduced horizontal movement and uplift
Enhanced water penetration resistance
Good over poor soils or in small, confined areas
Repairs are more difficult and expensive
Little tolerance for paver thickness variations
or inaccurate base elevations
Mortar Setting Bed
Bonded to Concrete
Base
Greater tolerance for paver thickness variations
or inaccurate base elevations
Can be used on steeper slopes and greater
vehicle speeds
Drainage occurs on the surface
Movement joints must align through entire
paving system
Least cost-effective
Mortar joint maintenance required
Repairs are most difficult and expensive
Mortar Setting
Bed Unbonded to
Concrete Base
Greater tolerance for paver thickness variations
or inaccurate base elevations
Movement joints in setting bed and base are not
required to align
Preferred when used over elevated structural slab
Bond break must be used to avoid stresses
caused by horizontal movement between
layers
Least cost-effective
Mortar joint maintenance required
Repairs are most difficult and expensive
TABLE 2
Comparison of Pavements Made with Clay Pavers
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 4 of 19
TABLE 3
Selection of Setting Bed and Base
NOTES: KEY:
1. Use stabilized joint sand R =Recommended
2. When snow melt system is in sand setting bed, A =Acceptable
use stabilized sand in setting bed. NA =Not Acceptable
3. Use Application PS or PX pavers
Site Condition
or Project Requirement
Sand Setting Bed
Bituminous
Setting Bed
Mortar Setting Bed
Bonded Unbonded
Aggregate
Base
Asphalt
Base
Cement-
Treated
Aggregate
Base
Concrete
Base
Asphalt
Base
Concrete
Base
Concrete
Base
Concrete
Base
Soft Soil in Subgrade R R A A R A A A
Tree Roots in/near
Subgrade R A NA NA A NA NA NA
Expansive Soil in
Subgrade A
1
R A NA R NA NA NA
Snow Melt System A
2
A
2
A
2
R
2
A
1
NA R R
Suspended Structural Slab A
1
NA A
1
R
1
NA R
1
R R
Good Surface Drainage R R R R R R R R
Poor Surface Drainage R R R R R R NA NA
Permeable Pavement R NA NA NA NA NA NA NA
Deep Frost Line R
1
R
1
R
1
R
1
A
1
A
1
A A
Freeze/Thaw R
1
R
1
R
1
R
1
A
1
A
1
A NA
Minimal Frosts R R R R R R R R
Pressure Washing R
1
R
1
R
1
R
1
R
1
R
1
R R
Vacuuming R
1
R
1
R
1
R
1
R
1
R
1
R R
Minimal Cleaning R R R R R R R R
ADA Compliance R R R R R R A A
Pedestrians Only R R R R R R R R
Light Vehicular Traffic R
3
R
3
R
3
R
3
R
3
R
3
R NA
DESIGN CONSIDERATIONS
Aesthetics
The relatively small size of clay pavers creates a
pavement surface with a human scale. As many
pavers can be observed simultaneously, the nuances
of different colors, textures and patterns can be clearly
seen when standing on the pavement. Single colors
can present a monolithic appearance. Multiple colors
can break down the scale of the pavement (see
Photo 2). Borders laid in a different color can add
interest to the pavement. In larger areas, it may be
desirable to introduce different colors in the form of
bands or panels. Some highly decorative pavements
have introduced patterns that flow, repeat and
intertwine (see Figure 2).
Color. Clay pavers are available in a wide range of
colors. The most common are red and brown earth
tones, but buff and gray colors also are produced
Photo 2
Multiple Colors Affect Pavement Scale
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 5 of 19
Double Basket Weave Herringbone Running Bond
Single Basket Weave Spanish Bond Stack Bond
Figure 2
Brick Paving Bond Patterns
(see Photo 3). Single colors as well as variegated
pavers can also be mixed together to form blends that
expand the palette of available colors.
The color of clay pavers is typically consistent
through the body of the paver and is highly resistant
to weathering and fading because of its vitrified
composition. Since clay pavers are made from natural
materials, there may be some inherent color variations
between different production runs from the same
manufacturer. This is most evident in large paved
areas of a single color. Using a field panel to establish
acceptable color variations and laying pavers taken
from different cubes of pavers helps avoid this issue.
Texture. Clay pavers are available with a range
of surface textures, such as wire cut and molded.
Viewed at a flat angle from a distance, a variation in
paver texture can be more obvious than a variation
in color. Designers may find it advantageous to
change the surface texture in different areas or bands
to exaggerate the contrast. The texture also has an
impact on slip and skid resistance.
Some pavers are manufactured with a more pronounced texture or surface pattern. Surface features including
a grid of dimples or domes also can be imprinted into the surface of the paver before firing. Pavers also can be
manufactured and installed to provide a tactile/detectable warning surface. In addition, patterns and words can be
engraved or laser etched into the surface of fired pavers.
Photo 3
Clay Paver Colors
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 6 of 19
Edge Treatment. Pavement texture is created not
only by the character of the texture of each paver,
but also by the treatment of the edges. Pavers can
have square edges, rounded edges or beveled edges
formed during the extrusion, molding or pressing
processes. These can be uniform along the entire
edge of the paver, which enhances the uniformity of
the surface, or they can be made to be variable or
irregular to create the feel of a historic pavement.
Additionally, fired pavers can be tumbled to create
distressed edges.
Pavement use and maintenance should be considered
when selecting the edge treatment of pavers, as
they may affect the appearance or smoothness of
the paving surface. When square-edge pavers are
laid with sand joints, care should be taken to ensure
that they do not make direct contact with or lip under
adjacent pavers. A minimum of 1/16 in. sand-filled
joint should separate each clay paver. Maintaining
full sand joints and taking care not to distress paver
edges during snow removal procedures helps minimize
potential chippage of a paver's edges. Using clay
pavers with chamfers enhances drainage by channeling water away from the surface, which can improve skid
resistance.
Bond Patterns. Many installation patterns can be used when laying clay pavers. Some of the most popular are
herringbone bond, running bond, stack bond and basket weave, as shown in Photo 4. When choosing a pattern,
considerations should include the setting bed of the pavement and the horizontal loads. Vehicle loads typically
generate the largest horizontal load on a pavement. Sand and bituminous setting beds are more prone to paver
creep, or horizontal movement. A herringbone bond best distributes horizontal forces across a pavement, reducing
the potential for creep. Running bond and other patterns with continuous joints do not distribute horizontal loads as
well as herringbone bond. If these bond patterns are used, continuous joints should be oriented perpendicular to
the direction of traffic.
In some projects, different-colored pavers are arranged to create a pattern that aligns with adjacent features, such
as building columns or trees. The size of different colored clay pavers may vary within permissible tolerances.
Pavers supplied to a project may be slightly smaller or slightly larger than the specified sizes assumed in design.
As such, the exact number of pavers that can be laid within a set dimension will vary unless the joint widths are
slightly adjustable. Paving systems with sand or bituminous setting beds that are subject to vehicular applications
can have their structural integrity reduced if joints are too wide. Therefore, the paver layout should be designed
with a degree of flexibility to accommodate slight variations in the pattern. As necessary, cutting individual pavers
also may be used to solve alignment and structural integrity issues.
Pedestrian Traffic
Paving systems using clay pavers exposed to pedestrian traffic for residential and nonresidential applications are
common. Many residential patios and walks can be constructed with only a base layer between the subgrade and
the setting bed. For more public pedestrian applications such as sidewalks and plazas, a more substantial paving
system may be required.
Vehicular Traffic
Light vehicular traffic includes general access for cars and for trucks, but in smaller volumes. As stated in ASTM C
1272, high volumes of traffic are considered traffic with over 251 daily equivalent single loads (ESAL), a standard
term used by pavement engineers. For further information about clay pavements subject to heavy vehicular traffic,
refer to Flexible Vehicular Brick Paving A Heavy Duty Applications Guide [Ref. 14].
Photo 4
Clay Paver Sidewalk in Basket Weave Pattern
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 7 of 19
The load capacity of a clay paving system with a sand setting bed and aggregate base is dependent on the total
pavement section rather than just the clay paver layer. Most individual clay pavers have a high compressive
strength and, with sufficient thickness, can develop significant interlock with surrounding pavers to support light
vehicular loads when properly constructed. Sufficient thickness and compaction of subbase, base and paver layers
virtually eliminates pavement deformation under loading.
For light duty vehicular paving systems, a maximum traffic speed of 30 mph (50 kph) is considered appropriate for
pavers in a sand setting bed. As vehicle speeds increase, the horizontal loading caused by accelerating, braking
and turning increases. Light duty vehicular clay paving systems with sand setting beds where a herringbone bond
is used, where joint width is maintained between 1/16 to 3/16 in. (1.6 to 4.8 mm), where an appropriate jointing
sand is properly installed and maintained and where sufficient edge restraint is provided can perform well and
substantially reduce the potential for movement of the pavers from horizontal creep.
Slip Resistance, Skid Resistance and Hydroplaning
Each of these issues relates to the slipperiness of the pavement surface. Slip resistance generally refers to the
slipperiness of a pavement as experienced by pedestrians. Skid resistance and hydroplaning are related to the
slipperiness of a pavement as experienced by vehicles.
The slip resistance is determined as the static coefficient of friction of a surface. A number of test procedures are
available for laboratory and field testing, but they may provide different values. Slip resistance can be measured
in the laboratory and the field using ASTM C 1028, Test Method for Determining the Static Coefficient of Friction
of Ceramic Tile and Other Like Surfaces by the Horizontal Dynamometer Pull-Meter Method [Ref. 4]. For surfaces
in an accessible route, the United States Access Board historically recommended, but did not mandate, a value of
0.6 for level surfaces and 0.8 for ramps when measured by the portable NBS-Brungraber machine using a silastic
sensor shoe. Most clay pavers exceed these values.
Skid resistance is typically determined on the basis of a materials dynamic coefficient of friction, which generally
decreases as speed increases. Testing usually involves either a specialized test vehicle moving at more than
30 mph (50 kph) or a portable British Pendulum Tester used in accordance with ASTM E 303, Test Method for
Measuring Surface Frictional Properties Using the British Pendulum Tester [Ref. 11]. For paving systems exposed
to light duty application pavements covered in this Technical Note 14 series, skid resistance is not an issue.
Hydroplaning also is associated with speed, but in conjunction with standing water on the pavement surface. Due
to the speed restrictions imposed on clay pavements subject to light duty vehicle traffic, hydroplaning should not
be a concern for clay pavements.
Slope
Paving systems can be successfully used on slopes with up to a 10 percent gradient. For projects where site
conditions involve slopes exceeding 10 percent, a design professional and local codes should be consulted.
Drainage
Adequate drainage is important to the performance and durability of any clay paving system. Water should be
drained from the paving system as quickly as possible. A minimum slope of 1/4 in. per ft of slope (2 percent grade)
is recommended. Adequate drainage should be provided to ensure the integrity of all layers in a paving system.
Three types of drainage potentially exist in clay paving systems: surface restricted, subsurface restricted and
unrestricted. Surface restricted drainage occurs on the surface of the paving system. This type of drainage is
typical of clay paving systems with a mortar setting bed. Subsurface restricted drainage occurs when water drains
over the surface and immediately below the paving course. This type of drainage is typical of paving systems
installed with a bituminous setting bed. Unrestricted drainage involves draining water from the surface, the
subsurface and through the subgrade. This type of drainage requires a sand setting bed on an aggregate base.
Drains should be selected and placed to adequately handle anticipated water flow. Drains serving paving systems
should have openings not only on the surface but also on the sides. Such drains should be used for all paving
systems to drain water from adjacent materials and to prevent capillary rise. Side openings should extend below
the top of any impervious layer or membrane in the paving system. Drains placed in pavements with sand setting
beds should have screens to prevent sand from entering the drain. Pavement edges that restrict water flow at the
lowest point in the paving system where water is anticipated should have weeps at 16 in. (406 mm) on center.
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 8 of 19
Accessibility
The Americans with Disabilities Act Accessibility
Guidelines (ADAAG) [Ref. 1] establish minimum
design requirements that cover access for people with
disabilities to public and private buildings and facilities.
The Public Rights-Of-Way Accessibility Guidelines
(Draft PROWAG) [Ref. 13] in draft form cover disability
access provisions for pedestrian areas along public
rights-of-way. Research has documented that clay
paving systems can comply with the accessible
provisions within these guidelines [Ref. 12 and 15].
The ADAAG and Draft PROWAG mandate several
surface profile requirements applicable to all pavement
systems. The designer should be aware of maximum
permissible gradients and other requirements that
often are overlooked (see Photo 5).
In addition to planning and designing in accordance
with these guidelines, it is important to implement
regular maintenance programs to maintain these
routes in a safe and serviceable condition. Specific
requirements especially pertinent to clay pavers
include surface, changes in level, joints and detectable
warning surfaces.
Surface. The ADAAG and Draft PROWAG require
an accessible surface to be firm, stable and slip-
resistant. Smoothness also may be an important
criterion, because a pedestrian in a wheelchair
may be more sensitive to vibration or trip hazards.
Properly designed, installed and maintained clay
paver surfaces achieve these properties. Besides
inadequate design, installation or maintenance, all
pavement systems may be subject to heaving and
settlement of underlying soils that result in changes in
level. Research has shown that the vibration on clay
paver surfaces is comparable to or less than that of
poured concrete and other common paving materials
[Refs. 12 and 15].
Changes in Level. Both the ADAAG and Draft
PROWAG allow a change in level (surface
discontinuity) up to 1/4 in. (6.4 mm) (see Figure 3a).
Both the ADAAG and Draft PROWAG allow a change
in level between 1/4 in. (6.4 mm) minimum and 1/2
in. (12.7 mm) maximum. The ADAAG requires this
change in level to be sloped (beveled) not steeper
than 1:2 (see Figure 3b). The Draft PROWAG also
requires a maximum slope (bevel) of 1:2 for this
change in level, but further mandates that the slope
(bevel) be applied across the entire change in level
(see Figure 3c).
With respect to pavers, sudden changes in level
(differences in elevation of the top surfaces of adjacent
pavers) should be kept to a minimum through careful
a) ADAAG & Draft PROWAG
Change in Level
up to 1/4 in. (6.4 mm)
Max. 1/4 in.
(6.4 mm)
1
2
Max. 1/2 in.
(12.7 mm)
c) Draft PROWAG Vertical Change
between 1/4 and 1/2 in.
(6.4 and 12.7 mm)
b) ADAAG Vertical Change
between 1/4 and 1/2 in.
(6.4 and 12.7 mm)
Max. 1/4 in.
(6.4 mm)
Max. 1/2 in.
(12.7 mm)
1
2
For Vertical Changes Greater Than
1/2 in. (12.7 mm), Use Ramp
Figure 3
Requirements for Making
Changes in Elevation
Photo 5
At Grade Street Crossing with ADA-Compliant
Surface Texture Changes
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 9 of 19
design and installation and should be maintained as part of a regular maintenance program. Changes in level can
result from heaving or settling of the pavement; uneven joints or can occur at frames and manhole covers.
Joints. The ADAAG does not specifically cover joints, but it does have requirements for openings in gratings,
which could be considered as being similar. The Draft PROWAG ADAAG has requirements for horizontal
openings in walkway joints and gratings. Both guidelines allow openings up to 1/2 in. (12.7 mm) wide, more than
twice the typical width of joints between pavers in pavements with sand and bituminous setting beds that are
typically 1/16 in. (1.6 mm) to 3/16 in. (4.7 mm) wide. J oints between pavers in a mortar setting bed are generally
3/8 in. (9.4 mm) to 1/2 in. (12.7 mm) wide, but would not be considered an opening.
Detectable Warning Surfaces. Both the ADAAG and the Draft PROWAG require detectable warning surfaces
consisting of truncated domes sized to have a base diameter of 0.9 in. (23 mm) minimum and 1.4 in. (36 mm)
maximum, a top diameter of a minimum of 50 percent to a maximum of 65 percent of the base diameter, and a
height of 0.2 in. (5.1 mm). Clay pavers can be made with truncated domes.
The ADAAG requires truncated domes to be placed on a square grid with a center-to-center spacing of 1.6 in.
(41 mm) minimum and 2.4 in. (61 mm) maximum, and a base-to-base spacing of 0.65 in. (17 mm) minimum,
measured between the most adjacent domes. The Draft PROWAG requires truncated domes to be placed in
either a square or a radial grid pattern meeting the same dimensional layout requirements as set forth in the
ADAAG.
Both guidelines require detectable warning surfaces to extend 24 in. (610 mm) from rail platform boarding edges.
The Draft PROWAG also covers curb ramps and blended transitions that are not covered in the ADAAG. Curb
ramps and blended transitions require detectable warning surfaces to extend 24 in. (610 mm) minimum in the
direction of travel for their full width. Flares of curb ramps are not required to have a detectable warning surface.
CLAY PAVERS
Manufacturing
Clay pavers are manufactured in much the same
way as face brick, as discussed in Technical Note
9. Extrusion (stiff-mud), molding (soft-mud) or dry-
pressing processes are used to produce pavers (see
Figure 4). Extruded clay pavers have a wire-cut texture
or smooth die-skin wearing surface. Lugs (spacer bars)
and chamfers may be formed on the sides and edges
of the pavers during the extrusion or cutting process.
Clay pavers produced by the molding or dry-pressing
processes have a smooth or textured surface. Lugs
and chamfers also may be formed by the sides and
edges of the molds. Pavers from any of the production
methods may have aesthetic features such as irregular
or textured edges. Clay pavers made by the molding
or dry-pressed process may have frogs or cavities on
one bed surface, although they would not be exposed.
Pavers generally are manufactured with their length equal to a module of their width. Two commonly specified
clay paver sizes are 4 in. wide by 8 in. long (102 mm by 203 mm) and 3
5
/8 in. wide by 7
5
/8 in. (92 mm by
194 mm) long. Other similar sizes are available, such as 3 in. (95 mm) wide by 7 in. (190 mm) long, and
several manufacturers are able to provide custom sizes. Common specified thicknesses are 1 in. (38 mm),
2 in. (57 mm) and 2
5
/8 in. (67 mm).
Standards
Clay pavers can be used as a wearing course in many exterior pavement and interior floors. Most pavers in the
United States are manufactured to comply with consensus standards published by ASTM International (ASTM).
Two ASTM standards define requirements for clay pavers for exterior use: ASTM C 902, Standard Specification
for Pedestrian and Light Traffic Paving Brick [Ref. 3], and ASTM C 1272, Standard Specification for Heavy
Paver with
Square Edges
Paver with
Textured Edges
Extruded or Re-Pressed Paver
with Chamfer and Lugs
Chamfer
Lug
Molded Paver with
Rounded Edges
Figure 4
Clay Pavers
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 10 of 19
Vehicular Paving Brick. [Ref. 5] For light duty applications addressed by the Technical Notes 14 series, clay
pavers complying with ASTM C 902 are normally used. Clay pavers manufactured to meet ASTM C 1272 may
be used in light duty or heavy vehicular applications and may provide longer pavement service life especially
where the pavement is subject to higher volumes of vehicular traffic. Only clay pavers meeting the requirements of
ASTM C 1272 are suitable for heavy vehicular applications, which are covered in Flexible Vehicular Brick Paving
A Heavy Duty Applications Guide.
ASTM C 902. This specification covers clay pavers suitable for patios, walkways, floors, plazas, residential
driveways and commercial driveways (passenger drop-offs). It describes three Classes and three Types of clay
pavers according to severity of their exposure to weather and to traffic, respectively. Three Applications also are
defined, based upon the pavers intended use, and limit their dimensional tolerances, distortion and extent of
chipping.
Class - A pavers Class relates to its resistance to damage from exposure to weather and is based on
compressive strength and absorption properties. Class SX pavers are intended for use where the pavers may
be frozen while saturated with water. Class MX pavers are intended for exterior use where the pavers will not
be exposed to freezing conditions. Class NX pavers are not acceptable for exterior use but may be used for
interior areas where the pavers are protected from freezing when wet. For most exterior residential or light duty
applications, Class SX pavers are used.
Type - A pavers Type relates to its resistance to abrasion. Type I pavers are intended for use where the pavers
are exposed to extensive abrasion, such as sidewalks and driveways in publicly occupied spaces. Type II pavers
are intended for use where the pavers are exposed to intermediate pedestrian traffic, such as heavily traveled
residential walkways and residential driveways. Type III pavers are intended for use in low pedestrian traffic,
residential areas such as floors and patios of single-family homes. For most exterior residential or light duty
applications, Type I or II pavers are used.
Application - A pavers Application relates to its aesthetics and use. Application PS pavers are intended for
general use and can be installed in any bond pattern with mortar or with sand-filled joints when not exposed to
vehicular traffic. When Application PS pavers are installed with sand-filled joints for light duty vehicular applications,
they should be laid in running bond or other bonds not requiring extremely close dimensional tolerances. Any
bond pattern can be used when Application PS pavers are installed with mortar joints. Application PX pavers have
tighter dimensional tolerances that allow consistently narrow joints between pavers. Such uses include pavements
without mortar joints between pavers where exceptionally close dimensional tolerances are required as a result of
special bond patterns or unusual construction requirements. Application PA pavers are characterized by aesthetic
effects such as variability in size, color and texture. Such pavers have performed successfully in many historic clay
paving applications and are generally used where a distinctive architectural character is desired. Such applications
are often installed with mortar joints between pavers, but can be successful in sand-filled joint applications that
are laid by workers with experience installing Application PA pavers in this manner. Using stabilized joint sand or
applying stabilizer to joint sand will help prevent sand loss from wider sand-filled joints.
Pavers complying with ASTM C 902 are not required to have a minimum thickness. However, they are commonly
manufactured to a specified thickness of 2 in. (57 mm) and 1 in. (38 mm). Except for patios or walks for one-
or two-family homes in southern climates with limited frost exposure, clay pavers 1 in. (38 mm) thick are usually
installed only over a rigid base.
ASTM C 1272. This standard addresses heavy vehicular pavers generally used in streets, commercial driveways
and industrial applications. ASTM C 1272 designates two Types of pavers depending on their method of
installation. Three Applications limit dimensional tolerances, distortion and extent of chipping.
The paver Type is based upon the compressive strength, breaking load and absorption properties of the pavers.
Type F pavers are intended to be set in a sand setting bed with sand-filled joints. The minimum paver thickness
is required to be 2
5
/8 in. (67 mm). They also can be installed over flexible or rigid bases. Type R pavers are
intended to be set in a mortar setting bed with mortar joints over a concrete base. Type R pavers also can be set
on a bituminous setting bed with sand-filled joints and supported by an asphalt or concrete base. The minimum
thickness for Type R pavers is required to be 2 in. (57 mm).
Applications PS, PX and PA are common to both ASTM standards and denote similar requirements.
Pavers complying with ASTM C 1272 may contain frogs but must be without cores or perforations.
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 11 of 19
ASTM Properties for Clay Pavers. The Class, Type and Application designations within ASTM clay paver
standards are based upon physical properties and characteristics, including compressive strength, breaking
load, absorption, abrasion, dimensional tolerances and extent of chipping. Pavers must be resistant to damage
from the effects of traffic and the environment. In many regions of the United States, clay pavers will be exposed
to severe environmental conditions. Pavers often are in a saturated condition and can experience numerous
freeze/thaw cycles. Application of deicers can cause additional thermal shock to pavers. Compliance with property
requirements of ASTM C 902 and C 1272 provides the required durability.
Compressive Strength, Breaking Load and Absorption - The strength and absorption requirements of pavers
from the ASTM standards are shown in Table 4. Some pavers are durable, but cannot be classified under the
physical requirements shown in Table 4. Using alternatives in the specifications allows pavers that are known to
perform well to meet the durability requirement. It does not signify that the pavers are of a lower quality.
TABLE 5
Maximum Abrasion Requirements
ASTM Standard Abrasion
Index
Volume Abrasion
Loss (cm
3
/cm
2
)
C 902 Type I 0.11 1.7
Type II 0.25 2.7
Type III 0.50 4.0
C 1272 Type R & F 0.11 1.7
For pavers complying with ASTM C 902 or C 1272, several alternatives are allowed. The freezing and thawing
test alternative allows the cold water absorption and the saturation coefficient to be waived if a sample of five brick
that meet all other requirements passes the freezing and thawing test of ASTM C 67 without breaking and with no
greater than 0.5 percent loss in dry weight of any individual unit. The sulfate soundness alternative allows the cold
water absorption and saturation coefficient to be waived if five brick survive 15 cycles of the sulfate soundness test
with no visible damage. The performance alternative allows specifiers to waive all property requirements for pavers
if they are satisfied with information furnished by the manufacturer on the performance of the pavers in a similar
application subject to similar exposure and traffic.
For pavers complying with ASTM C 902, the absorption alternative allows the saturation coefficient to be waived
for pavers that absorb less than 6.0% after 24 hours of submersion in room-temperature water.
Abrasion - The Abrasion Index is the ratio of the absorption divided by the compressive strength, multiplied by
100. The compressive specimen must be half pavers that are without core holes, frogs or other perforations, and
the full height of the paver no less than 2 in. (57 mm). The volume abrasion loss is used if the height requirement
cannot be met. The volume abrasion loss is determined by the loss of material created by sandblasting the surface
of the paver. The abrasion requirements of pavers from the ASTM standards are shown in Table 5.
TABLE 4
Property Requirements
ASTM Standard Minimum Compressive
Strength, psi (Mpa)
Maximum Cold Water
Absorption, %
Maximum Saturation
Coefficient
Minimum Breaking
Load, lb/in. (kN/mm)
Avg of 5
Brick
Individual Avg of 5
Brick
Individual Avg of 5
Brick
Individual Avg of 5
Brick
Individual
C 902 Class SX 8,000 (55.2) 7,000 (48.3) 8.0 11.0 0.78 0.80 ---- ----
Class SX (molded) 4,000 (27.6) 3,500 (24.1) 16.0 18.0 0.78 0.80 ---- ----
Class MX 3,000 (20.7) 2,500 (17.2) 14.0 17.0 No Limit No Limit ---- ----
Class NX 3,000 (20.7) 2,500 (17.2) No Limit No Limit No Limit No Limit ---- ----
C 1272 Type R 8,000 (55.2) 7,000 (48.3) 6.0 7.0 ---- ---- ---- ----
Type F 10,000 (69.0) 8,800 (60.7) 6.0 7.0 ---- ---- 475 (83) 333 (58)
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 12 of 19
Chippage - Clay pavers may chip in transit or during construction. Table 7 shows the extent of chippage allowed
by prescribing the maximum distance that chips may extend into the surface of a paver from an edge or a corner.
The sum of the length of chips on a single paver must not exceed 10 percent of the perimeter of the exposed
face of the paver. Cobbled or tumbled pavers that are intentionally distressed after production are classified as
Application PA pavers.
Distortion - Both ASTM C 902 and C 1272 limit distortion and warpage of surfaces and edges intended to be
exposed in use. The distortion must not exceed the maximum for the Application specified as noted in Table 8.
TABLE 8
Tolerances on Distortion
Specified Dimension, in. (mm) ASTM C 902 & C 1272
1
Maximum Permissable Distortion, in. (mm)
Application PX Application PS Application PA
8 (203) and under 1/16 (1.6) 3/32 (2.4) no limit
Over 8 (203) to 12 (305) 3/32 (2.4) 1/8 (3.2) no limit
Over 12 (305) to 16 (406) 1/8 (3.2) 5/32 (4.0) no limit
1
ASTM C 1272 Type F clay paver required to meet Application PX
Dimensional Tolerances - The dimensional tolerances for pavers are based upon the dimension width, height
or length considered. The actual dimensions may vary from the specified dimension by no more than plus or
minus the dimensional tolerance. The tolerances for both C 902 and C 1272 pavers are shown in Table 6.
TABLE 6
Dimensional Tolerance Requirements
Dimension, in. (mm) ASTM C 902 and C 1272
Application PS,
in. (mm)
Application PX,
in. (mm)
Application PA
3 (76) and under 1/8 (3.2) 1/16 (1.6) no limit
over 3 to 5 (76 to 127) 3/16 (4.7) 3/32 (2.4) no limit
over 5 to 8 (127 to 203) 1/4 (6.4) 1/8 (3.2) no limit
over 8 (203) 5/16 (7.9) 7/32 (5.6) no limit
TABLE 7
Maximum Chippage Requirements
ASTM Standard Edge, in. (mm) Corner, in. (mm)
C 902 Application PS 5/16 (7.9) 1/2 (12.7)
Application PX 1/4 (6.4) 3/8 (9.5)
Application PA As specifed by purchaser As specifed by purchaser
C 1272 Application PS & PX 5/16 (7.9) 1/2 (12.7)
Application PA No Limit No Limit
SETTING BEDS
Setting beds provide a means to adjust for dimensional variations in the height of a paver. They also support the
clay pavers and transfer load to the base.
Sand Setting Bed
Individual pavers in sand setting beds are held in position by the frictional interlock that is developed in each sand-
filled joint between adjacent pavers. The joints transfer vertical and horizontal forces, but can absorb expansion
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 13 of 19
and contraction of the individual pavers. If the pavement deflects slightly, the pavers will realign themselves to the
new profile without significant loss in structural capacity. Interlock is developed by properly sized joints filled with
consolidated joint sand. Sand setting beds may be installed directly on an aggregate base, asphalt base, cement-
treated aggregate base or concrete base. For further information about pavements with sand setting beds, refer to
Technical Note 14A.
Bituminous Setting Bed
In pavements with a bituminous setting bed, less interlock is developed by the joint material than in a pavement
with a sand setting bed. However, additional restraint is provided by the adhesive nature of the tack coat.
Bituminous setting beds can be are set on an asphalt base or concrete base. For further information about
pavements with bituminous setting beds, refer to Technical Note 14B.
Mortar Setting Bed
Pavers in a mortar setting bed are bonded to the underlying mortar bed and transfer most of the vertical load
through direct bearing. Mortar setting beds should be used only with a concrete base and may be bonded or
unbonded to it. The joints between pavers are filled with mortar that transfers horizontal load. However, mortar will
not absorb expansion and contraction of individual pavers. If the pavement deflects significantly, the pavement
may crack along mortar lines or across pavers. For further information about pavements with mortar setting beds,
refer to Technical Note 14C.
BASES
The base layer in the pavement is the primary structural layer. It is subjected to the compressive, tensile and
shearing stresses transmitted through the wearing course. Materials in the base layer need to be capable of
resisting these stresses. Pedestrian loading is sufficiently light that a base thickness of only 4 in. (102 mm) is
required when no specific site conditions dictate a thicker base. Vehicular loading requires a thicker base.
Including a subbase often provides economic benefits when the subgrade is of low strength or is susceptible to
frost. Because it is lower in the pavement section, the subbase is subjected to lower stresses than the base course
(see Figure 1). A subbase also can serve as a working platform to prevent subgrade damage from construction
equipment. Subbase material also may be added to increase the depth of the pavement section in frost-
susceptible soils. A subbase is not usually required for light duty vehicular pavements. Pedestrian-only pavements
generally do not include a subbase.
Aggregate Subbase and Base
Aggregate subbase materials are typically medium-quality graded aggregates or clean sand-and-gravel mixtures.
They should not be susceptible to deterioration from moisture or freezing. Subbase materials are covered by
ASTM D 2940, Specification for Graded Aggregate Material for Bases or Sub-bases for Highways or Airports
[Ref. 9]. Typical gradation envelopes are prescribed, along with other properties such as durability and plasticity.
Aggregate subbase materials generally are graded from 1 in. (38 mm) to No. 200 (0.075 mm) sizes. Aggregate
subbase materials may be used directly over the subgrade soil or on top of a geotextile.
Aggregate base materials are typically high-quality, crushed, dense-graded aggregates. They usually are specified
in ASTM D 2940. Aggregate base materials generally are graded from 3/4 in. (19.1 mm) to No. 200 (0.075 mm)
sizes. An aggregate base may be placed directly on the subgrade or over an aggregate subbase. A sand setting
bed may be installed directly on an aggregate base.
It is important to compact aggregate subbase and base layers. Each layer should be compacted in accordance
with ASTM D 698 to 95 percent maximum density.
Asphalt Base
Asphalt base materials consist of mixtures of aggregates and asphalt cement that are produced at a central hot-
mix plant. The materials are proportioned to comply with a mix design, and the materials usually are specified in
state or local standards and in ASTM D 3515, Specification for Hot-Mixed, Hot-Laid Bituminous Paving Mixtures
[Ref. 10]. Asphalt aggregates usually are blended to achieve a gradation from 1/2 in. (12.7 mm) or 3/8 in. (9.5 mm)
to No. 200 (0.075 mm). An asphalt base may be placed directly on the subgrade but is more commonly laid over
an aggregate subbase or base. It creates a relatively stiff and impermeable base layer.
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 14 of 19
Cement-Treated Aggregate Base
A cement-treated aggregate base material is a relatively dry, lean mixture of aggregate and portland cement
that creates a stiff and impermeable base layer. These materials should be mixed at a concrete plant and laid
by machine. Cement contents vary between 5 and 12 percent with sufficient water added to achieve required
compaction and full hydration of cement. Compressive strengths typically are around 750 psi (5.17 MPa). A
cement-treated aggregate base may be placed directly on the subgrade but is more commonly laid over an
aggregate subbase. This type of base does not include reinforcement, and because of the low water and cement
content, can be laid without movement joints.
Concrete Base
The compressive strength of a concrete base should be at least 4,000 psi (27.6 MPa). Concrete bases may be
plain or reinforced, incorporating a grid of movement joints with load transfer devices, such as dowels. Layouts of
movement joints require careful consideration of the overlying pavement system. Movement joints placed more
than 12 ft ( 3.66 m) apart should extend through the entire pavement to prevent damage to the pavers unless
using an unbonded system. A concrete base should be placed over an aggregate subbase or base.
SUBGRADE
The subgrade is classified by the existing soil conditions, the environment and drainage. For vehicular
applications, the existing soil conditions for the project should be determined by a geotechnical engineer before
design of the paving system. For pedestrian and residential applications, a geotechnical engineer should be used
as necessary to verify suitability of existing soil for the proposed paving system.
Environmental conditions and the quality of drainage can affect the support provided by the subgrade. In wet
climates, poorly drained areas or those that experience freezing conditions, the support from the subgrade is
likely to be reduced during certain periods of the paving systems life. Conversely, in arid climates or well-drained
areas, it is likely that a higher degree of subgrade support will be experienced during part of the paving systems
life. Where water can penetrate the subgrade, it is important to drain water quickly to alleviate any potential
fluctuations in soil moisture content.
Soils are typically classified into different groups to represent their engineering properties. In general, soils
consisting primarily of gravel and sand can be used to support most paving systems. In general, soil consisting
of clay can usually be used to support a paving system as long as it is located in a dry environment or is drained.
Soils classified as organic are not suitable for subgrade and should be removed and replaced. For further
guidance regarding soil capacities, refer to Flexible Vehicular Brick Paving A Heavy Duty Applications Guide
[Ref. 14].
GEOTEXTILE
Geotextiles are formed from plastic yarns or filaments such as polypropylene and polyester. They may be woven
or nonwoven fabrics supplied in rolls. A geotextile may be used between fine-grained subgrade materials and
base or subbase layers, particularly where moist conditions are anticipated. This separates the two layers,
preventing the intrusion of fine soil particles into the overlying granular layer and preventing larger aggregates from
punching down into the subgrade. This enables the base to retain its strength over a longer period. Geotextiles
also can provide limited reinforcement to the overlying pavement layer. As the subgrade begins to deform, the
geotextile is put into tension, which reduces the loading on the subgrade, slowing rut development. The geotextile
manufacturers recommendations should be sought during selection of the appropriate geotextile for particular soil
conditions.
PAVEMENT LAYER CONSTRUCTION
Subgrade Preparation
The subgrade should be excavated to achieve a uniform pavement thickness, and any substandard or soft
materials should be undercut and replaced with acceptable backfill. A subsurface drainage system may be
installed as perforated pipes or fin drains if necessary. All utility trenches should be properly backfilled and each
layer thoroughly compacted to prevent settlement. The subgrade should be scarified and moisture conditioned
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 15 of 19
to within 2 percent of optimum moisture content as determined by ASTM D 698, Test Methods for Laboratory
Compaction Characteristics of Soil Using Standard Effort (12,400 ft-lbf/ft
3
(600 kN-m/m
3
)) [Ref. 7], to a depth of
6 in. (152 mm). Moisture conditioning clay subgrades can be more complicated, because the clay absorbs water
more slowly. It should then be graded to the appropriate profile and compacted by rolling with appropriate static or
vibratory rollers. The subgrade should be compacted in accordance with ASTM D 698 to 95 percent maximum dry
density for clay and 100 percent maximum dry density for sand/gravel.
Geotextile
When a geotextile is used, it should be placed immediately before spreading the aggregate subbase or aggregate
base. Geotextiles are not used when other base types are constructed directly on the subgrade. Care should be
taken to stretch the material as it is unrolled to remove any wrinkles. A minimum lap of 12 in. (305 mm) should be
provided at the sides and ends of rolls. Construction equipment should not be allowed to operate directly on the
geotextile.
Aggregate Subbase and Base
Aggregate subbase and base courses are spread in layers of up to 6 in. (152 mm) in compacted thickness,
dependent upon the proposed compaction process. Material may be end-dumped from the delivery trucks
and spread by grader spreaders or by hand with care to avoid segregation. The material should be moisture
conditioned to within 2 percent of the optimum moisture content from ASTM D 698. It should then be compacted
by rolling with appropriate static or vibratory rollers, or with a plate vibrator. When using a plate vibrator, the layer
thickness must be 3 in. (76 mm) or less, and more than one layer may be required. The subbase and base
layers should be compacted according to ASTM D 698 to 95 percent maximum dry density. Limited regrading is
permissible to achieve correct surface profile and elevations. The maximum variation under the setting bed should
be +/- 3/16 in. (4.8 mm) when tested with a 10 ft (3.05 m) straightedge laid on the surface. The minimum slope of
the aggregate base should be 1 in. (25.4 mm) in 4 ft (1.22 m) to allow for drainage.
Asphalt Base
Asphalt materials are produced at a hot-mix plant. They are mixed at temperatures up to 300 F (149 C) and
should be installed before they cool to temperatures below 200 F (93 C). Asphalt base layers can be spread
by machine or by hand. Asphalt can be laid in lifts from 1 to 3 in. (38 to 76 mm) in thickness depending on the
aggregate size and compaction equipment. Hand spreading requires adequate compaction of the base. Machine
installation using a paving machine provides initial compaction, enabling more accurate placement and elevations
to be achieved. Compaction of the asphalt is accomplished by an initial breakdown rolling and then by a finish
rolling with steel- or rubber-tired rollers. Compaction is continued until the required density is achieved. This
normally is a minimum of 96 percent of the density of samples of the same material compacted in a laboratory.
Once materials have cooled to the ambient temperature, the layer can receive traffic, although the asphalt
continues to stiffen over several months. The maximum variation under the setting bed should be +/- 3/16 in. (4.8
mm) when a 10 ft (3.05 m) straightedge is laid on the surface. The minimum slope of the asphalt base surface
should be 1 in. (25.4 mm) in 4 ft (1.22 m) to allow for drainage.
Cement-Treated Aggregate Base
Plant-mixed cement-treated aggregate bases are transported to the site for spreading by machine or by hand.
When spread by a paving machine, the base should be compacted to the appropriate thickness. When spread
by a grader or by hand, adequate compaction is required. A cement-treated aggregate base also can be mixed
in place using special equipment. A granular subgrade or imported aggregate is thoroughly mixed with cement
and water to achieve the required thickness. Materials should be placed and compacted within two hours of
adding water and before initial set of the cement. The base should be compacted according to ASTM D 1557,
Test Methods for Laboratory Compaction Characteristics of Soil Using Modified Effort (56,000 ft-lbf/ft
3
(2,700 kN-
m/m
3
)) [Ref. 8] to at least 95 percent of the maximum dry density. The cement-treated layer should be cured by
water misting or by applying an asphalt emulsion cure coat. Traffic should not be allowed on the base for at least
seven days, but paver installation may commence after three days. The maximum variation under the setting bed
should be +/- 3/16 in. (4.7 mm) when a 10 ft (3.05 m) straightedge is laid on the surface. The minimum slope of
the base surface should be 1 in. (25.4 mm) in 4 ft (1.22 m) to allow for drainage.
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 16 of 19
Concrete Base
Concrete usually is plant-mixed and delivered to the site in ready-mix trucks. It is discharged between forms,
where it is spread and consolidated. The formwork is set to the correct elevations, and a vibrating screed is
drawn between the forms to achieve the appropriate surface elevations. Movement joints containing load-transfer
devices may be formed at the edges of each pour, or the devices can be cast into the concrete between forms.
Saw cutting may be undertaken to induce cracking at the desired locations. A concrete base may be finished with
a broom, brush or wood float. A polished surface finish should be avoided. Care should be taken to follow proper
curing procedures for at least 14 days. Vehicular loads should not be permitted for at least 7 days, but paver
installation may commence after 3 days. The maximum variation under the setting bed should be +/- 3/16 in. (4.7
mm) when a 10 ft (3.05 m) straightedge is laid on the surface. The minimum slope of the concrete base surface
should be 1 in. (25.4 mm) in 4 ft (1.22 m) to allow for drainage.
CLEANING AND MAINTENANCE
Clay pavers are highly resistant to absorption of stains and can be kept clean in most environments by regular
sweeping. Otherwise, cleaning of brick pavements essentially is the same as cleaning vertical brickwork, as
discussed in Technical Note 20. Mortar-filled joints generally are more resistant to aggressive cleaning methods
(i.e. pressure washers). Sand-filled joints subjected to aggressive cleaning methods should contain stabilized joint
sand or should be treated with a joint sand stabilizer.
Efflorescence
Efflorescence is a white, powdery substance that may occasionally appear on the surface of pavers. It is the
product of soluble compounds normally found in other pavement components or underlying soils, which are
deposited on the surface of the paver as absorbed water evaporates from the pavement surface. Soluble
compounds absorbed by the pavement from deicing chemicals also may cause efflorescence. Efflorescence often
can be vacuumed or brushed off the surface and removed from dry pavers. Washing downhill with water may
temporarily dissipate soluble compounds by dissolving them. However, care must be taken to ensure that the
contaminated water drains away from and does not re-enter the paving system.
In many cases, efflorescence will be minimal and will wear away naturally with traffic and weathering during
the early life of the pavement. If the salts are the result of groundwater or other more persistent water ingress,
proprietary cleaners are available to assist in their removal. Proper surface and subsurface drainage are critical in
these situations. For further information on efflorescence, refer to Technical Notes 23 and 23A.
Ice Removal
Several proprietary chemical products are available for preventing and removing ice from paved surfaces that
perform well and reduce potential staining of pavers. Among these are calcium magnesium acetate and urea. The
former is preferred because it is more effective at lower temperatures. Deicing of pavements has been undertaken
for many years using rock salt. This material contains calcium chloride and can cause efflorescence. Sand or grit
used to provide traction on ice should be swept up after the freezing cycle to minimize grinding of the pavers.
Snow Removal
Clearing snow from clay pavements can be undertaken using plows, snow blowers, shovels and brushes as used
for other pavements. Care must be taken to ensure that the blades of the equipment do not scrape the pavement
surface in a manner that might cause chipping. Rubber or urethane blade edges can be used, or proper blade
height can be maintained above the pavement surface using guide wheels. Any residual snow can be cleared with
brushes. Some snow-clearing procedures use heavy equipment to stockpile and subsequently remove the snow
from the property. If such equipment is used, the load capacity of the pavement should be adequately designed.
SPECIAL APPLICATIONS AND CONDITIONS
Clay pavers can be used in a number of special applications that require consideration of additional aspects.
The following sections cover the design of clay paver wearing surfaces for suspended decks, permeable paving
systems and hydronic snowmelt systems.
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 17 of 19
Suspended Structural Slabs
The design of pavement surfaces on suspended decks presents a special series of challenges, particularly when
constructed over habitable space (see Figure 5). These include prevention of water penetration into the structure,
reduction in heat loss/gain and dealing with elastic deflections.
Waterproofing. A pavement constructed over a structural concrete slab often requires a waterproof membrane.
Several sheet and liquid-applied membranes are available. In most applications, a protection board is required
over the waterproof membrane.
Drainage. Water inevitably will penetrate the paver system, and drainage is required to prevent it from collecting
on top of the waterproof membrane. Horizontal drainage mats consisting of a dimpled three-dimensional plastic
core covered with filter fabric frequently are used. A 2 percent slope should be provided toward drains to assist
drainage of water. Although the core material has a high compressive strength, the filter fabric can be compressed.
Consequently, horizontal drainage mats in pavements subject to vehicular traffic should not be positioned
immediately below the setting bed.
Insulation. When a paved surface is located over a habitable space, it may be necessary to incorporate insulation
into the section. The most common type of insulation is extruded polystyrene, available in boards of various
compressive strengths and thicknesses. However, compressive strength values are measured when the insulation
thickness is compressed 5 percent. As such, the material is resilient under load and should not be placed
immediately under the setting bed when vehicular traffic will use the pavement.
An alternative insulating material that can be used in pavement systems on suspended structural slabs subject to
vehicular traffic is foamed concrete. It is more rigid than extruded polystyrene but is less thermally efficient. This
material also is available in a range of compressive strengths and insulation values.
Loading. Pavers and setting bed materials can be considered to apply a dead load of 10 lb/sq ft per inch (190 Pa
per cm) of thickness.
Deflections. A maximum deflection of 1/360 of the span is recommended for flexible pavement systems installed
over a suspended structural slab. If vehicular loads are anticipated, flexible pavement deflection should be limited
to 1/480. When rigid paving systems are installed, the deflections should be limited to 1/600.
Permeable Pavements
Many urban development regulations require that the surface-water runoff from a new project should not exceed
the original values. This may be expressed as a peak flow rate or as a total quantity of water. Permeable
pavements (see Photo 6) can be used to reduce or delay entry of runoff from a pavement surface into stormwater
systems or environmentally sensitive areas. In pavements with clay pavers, this can be achieved by creating wide
Photo 6
Permeable Clay Pavement
Suspended Structural Deck
Topping Slab
Waterproof Membrane
Drainage Mat
Rigid Insulation
Base
Setting Bed
Clay Pavers
Figure 5
Typical Suspended Deck Paving Section
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 18 of 19
joints that are filled with a permeable aggregate rather
than sand. The pavers are also laid on a permeable
setting bed. This allows the water that falls on a
pavement to filter through the surface into a permeable
base. The water will be temporarily stored in the
base, or it may soak into the subgrade if this is also
permeable (see Figure 6).
Subgrades. If the subgrade is permeable, water that
infiltrates the pavement through the surface voids
can drain away over time, after a rain event. Good
practice usually requires that water completely drains
within three days of entering the pavement. However,
compaction in preparation for placing the base
material may result in significant reduction in subgrade
permeability. As such, there are few permeable
pavements that can rely completely on exfiltration
through the subgrade. If the water will not drain,
provision should be made to release the water stored
in the base material through drainage pipes.
Bases. Permeable bases are constructed using single size or open graded aggregate materials. These materials
typically have a void content of 15 percent to 40 percent to accommodate the water that needs to be detained.
Typical single-number aggregate sizes No. 4, No. 5 and No. 6 from ASTM C 33, Specification for Concrete
Aggregate [Ref. 2] or ASTM D 448, Classification for Sizes of Aggregate for Road and Bridge Construction [Ref. 6]
have a high void content and are frequently used. There are several double-number size options such as No. 57
and No. 67. For these aggregate materials, the void content is less because a broader grading envelope is used,
but the material may be more readily available.
Setting Bed and Joints. Similar aggregate is commonly used for the setting bed and joints. Size No. 8, No. 9 or
No. 89 aggregates complying with ASTM C 33 or ASTM D 448 are most frequently used. J oints ranging from 1/4
to 3/8 in. (6.4 to 9.5 mm) are typical. There also are several systems that use plastic spacers to create consistent
width joints of 1/2 to 3/4 in. (12.7 to 19.1 mm). However, the interlock between pavers is greatly reduced when joint
sizes are greater than 1/4 in. (6.4 mm) or when plastic spacers are used.
Hydronic Snow Melt Systems
Hydronic snow melt systems consist of a network of plastic tubing incorporated into the pavement system,
typically at 6 to 8 in. (150 to 200 mm) centers. Heated liquid is pumped around the system during near- and
subfreezing conditions so that the pavement temperature is maintained slightly above freezing, thus preventing the
accumulation of snow or the development of ice on the pavement surface. Continuous loops of 3/4 to 1 in. (19.1
to 25.4 mm) diameter tubing are made from cross-linked polyethylene. Tubing usually is secured to welded wire
fabric during construction to establish and maintain the designed layout.
There are two common approaches to positioning the tubing in the pavement. The first is to cast the tubing into
a concrete subslab, where it will be protected by the concrete. The second is to incorporate it within the bedding
material under the pavers. The latter option is not recommended for pavements with frequent vehicular traffic
but can be used for pavements under pedestrian loading. Adequate cover is required over the tubing, typically a
minimum of 1/2 in. (12.7 mm) after compaction. Bituminous bedding materials are not appropriate for this approach,
in part because of the installation temperature, but also because of the layer thickness. When a sand setting bed
is used, pre-compaction of the sand before screeding is recommended to minimize the occurrence of hard spots
under the pavers. This is achieved by providing approximately 1/2 in. (12.7 mm) additional cover when spreading
the sand, followed by several passes of the plate vibrator to compact the sand. The top surface then is loosened
slightly with a hoe or rake and screeded to the appropriate level, leaving a smaller surcharge than normal.
Permeable Subgrade
Geotextile
Permeable Setting Bed
Wide J oints with Permeable Filling
Perforated Drainage Pipe to Outfall
As Necessary for Impermeable Subgrade
Infiltration
Exfiltration Permeable Base
Outflow
Clay Pavers
Figure 6
Typical Permeable Pavement Section
www.gobrick.com | Brick Industry Association | TN 14 | Paving Systems Using Clay Pavers | Page 19 of 19
SUMMARY
Pedestrian and light duty vehicular pavements made with clay pavers can serve in a wide variety of applications,
including plazas, sidewalks and residential driveways and commercial driveways (passenger drop-offs). Many
paver sizes and colors are available, as are special shapes. Proper design and construction of a pavements base,
setting bed and pavers ensure a structurally stable, durable pavement able to meet site and project requirements.
Lending intrinsic character and sophistication to any space, clay pavers can be a structurally stable, economically
viable pavement option.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. Americans with Disabilities Act Accessibility Guidelines, United States Access Board, Washington, D.C., J uly
2004.
2. ASTM C 33, Standard Specification for Concrete Aggregate, Annual Book of Standards, Vol. 04.02, ASTM
International, West Conshohocken, PA, 2006.
3. ASTM C 902, Standard Specification for Pedestrian and Light Traffic Paving Brick, Annual Book of Standards,
Vol. 04.05, ASTM International, West Conshohocken, PA, 2006.
4. ASTM C 1028, Standard Test Method for Determining the Static Coefficient of Friction of Ceramic Tile and
Other Like Surfaces by the Horizontal Dynamometer Pull-Meter Method, Annual Book of Standards, Vol.
04.03, ASTM International, West Conshohocken, PA, 2006.
5. ASTM C1272, Standard Specification for Heavy Vehicular Paving Brick, Annual Book of Standards, Vol. 04.05,
ASTM International, West Conshohocken, PA, 2006.
6. ASTM D 448, Standard Classification for Sizes of Aggregate for Road and Bridge Construction, Annual Book
of Standards, Vol. 04.03, ASTM International, West Conshohocken, PA, 2006.
7. ASTM D 698, Standard Test Methods for Laboratory Compaction Characteristics of Soil Using Standard
Effort (12,400 ft-lbf/ft
3
(600 kN-m/m
3
)), Annual Book of Standards, Vol. 04.08, ASTM International, West
Conshohocken, PA, 2006.
8. ASTM D 1557, Standard Test Methods for Laboratory Compaction Characteristics of Soil Using Modified
Effort (56,000 ft-lbf/ft
3
(2,700 kN-m/m3)), Annual Book of Standards, Vol. 04.08, ASTM International, West
Conshohocken, PA, 2006.
9. ASTM D 2940, Standard Specification for Graded Aggregate Material for Bases or Sub-bases for Highways or
Airports, Annual Book of Standards, Vol. 04.03, ASTM International, West Conshohocken, PA, 2006.
10. ASTM D 3515, Standard Specification for Hot-Mixed, Hot-Laid Bituminous Paving Mixtures, Annual Book of
Standards, Vol. 04.03, ASTM International, West Conshohocken, PA, 2006.
11. ASTM E 303, Standard Test Method for Measuring Surface Frictional Properties Using the British Pendulum
Tester, Annual Book of Standards, Vol. 04.03, ASTM International, West Conshohocken, PA, 2006.
12. Cooper, R.A., Wolf, E., Fitzgerald, S.G., Dobson, A., and Ammer, W., Interaction of Wheelchairs and
Segmental Pavement Surfaces, Proceedings of the Seventh International Conference on Concrete Block
Paving, Cape Town, South Africa, Concrete Manufacturers Association of South Africa, October 2003.
13. Draft Public Right-of-Way Accessibility Guidelines, United States Access Board, Washington, D.C., 2005.
14. Flexible Vehicular Brick Paving A Heavy Duty Applications Guide, Brick Industry Association, Reston, VA,
2004.
15. Wolf, E., Pearlman, J ., Cooper, R.A., Fitzgerald, S.G., Kelleher, A., Collins, D.M., Boninger, M.L., Cooper,
R., Smith, D.R., Vibration Exposure of Individuals using Wheelchairs over Concrete Paver Surfaces,
Proceedings of the Eighth International Conference on Concrete Block Paving, San Francisco, CA,
International Concrete Pavement Institute, November 2006.
2007 Brick Industry Association, Reston, Virginia Page 1 of 13
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
14A
October
2007
Paving Systems Using Clay Pavers
on a Sand Setting Bed
Abstract: This Technical Note describes the proper design and construction of pavements made with clay pavers on a sand
setting bed in pedestrian and vehicular, residential and nonresidential projects.
Key Words: flexible, mortarless paving, paving, rigid, sand setting bed.
General
Determine if application is pedestrian, light duty vehicular
or heavy duty vehicular
Implement regular maintenance program to maintain
pavers in a safe and serviceable condition
Patterns
Use herringbone pattern for pavements subject to
vehicular traffic
Design flexibility into layout to accommodate field
conditions
Drainage
Provide a minimum slope of 1/4 in. per foot (2 percent
grade)
For concrete and impermeable bases, provide weeps
through base
Edge Restraints
For pavements subject to vehicular traffic, use concrete
or stone curbs or steel angles anchored to a concrete
base or foundation or a proprietary system rated for
traffic
For all other pavements, use any of the above or clay
pavers in a concrete foundation, proprietary plastic or
metal edge restraint systems spiked into aggregate
Use edge restraint with vertical face at paver interface
Clay Pavers
For most residential, pedestrian and light duty vehicular
applications, such as driveways, entranceways and
passenger drop-offs, use clay pavers complying with
ASTM C 902
For heavy duty vehicular applications, such as streets,
commercial driveways and industrial applications, use
clay pavers complying with ASTM C 1272.
Refer to Technical Note 14 for additional
recommendations
Joint and Setting Bed Sand
Use concrete sand complying with ASTM C 33
Stabilized Joint Sand
Use where potential sand loss or high water permeability
is anticipated and not desired
Follow paver manufacturers recommendation regarding
the use of stabilized joint sand or joint sand stabilizer
Use performance history as a basis for selection
Concrete Base
For concrete base on ground, provide control joints
spaced a maximum of 12 ft (3.66 m) o.c.
For elevated concrete slab, provide control joints through
concrete slab and expansion joints through pavement
above aligned with control joints
Provide weeps through base for drainage
Base, Subbase and Subgrade
Refer to Technical Note 14
INTRODUCTION
This Technical Note covers the design, detailing and
specification of clay pavers when laid on a sand
setting bed (see Figure 1). Refer to Technical Note
14 for clay paver design considerations, including
traffic, site conditions, drainage and appearance.
Sand-set pavers are the most cost-effective method of
constructing a pavement made with clay pavers. The
system relies upon developing interlock in the paving
course, which is generated by friction between the
pavers and the jointing sand. This enables the pavers
to function as part of the structural pavement system.
SUMMARY OF RECOMMENDATIONS:
Edge Restraint
Clay Pavers
Setting Bed
Compacted
Base
Compacted
Subgrade or Subbase
Figure 1
Typical Brick Pavement
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 2 of 13
Applications
Clay pavers set on a sand setting bed are
appropriate for virtually any paver application,
ranging from pedestrian to heavy duty vehicular
traffic. At a minimum, the system requires clay brick
pavers and a sand setting bed, compacted after
paver placement. Depending on subgrade conditions,
additional layers, base and subbase may be
required.
Residential Patios and Walkways. These
applications are the most common and handle the
lightest loads. The sand setting bed thickness should
be 1 to 1 in. (25 to 38 mm). The sand setting
bed should be separated from the subgrade by a
compacted aggregate base (see Figure 2). This base
typically consists of coarse aggregate (gravel) of
varying gradation, compacted to a minimum thickness
of 4 in. (102 mm) using mechanical tamping or
vibration.
Residential Driveways. The heavier and more
localized loads of vehicles on driveways serving one-
or two-family houses result in a thicker paving system
requiring a minimum 4 in. (102 mm) compacted
aggregate subbase. The base should consist of a
minimum 4 in. (102 mm) layer of coarse aggregate,
cast-in-place concrete or asphalt (see Figure 3).
The sand setting bed thickness should be 1 to 1 in.
(25 to 38 mm). The base typically consists of coarse
aggregate (gravel) of varying gradation, compacted
to a minimum thickness of 4 in. (102 mm) using
mechanical tamping or vibration.
Commercial/Public Plazas and Walkways. With
increased pedestrian traffic and increased risk of injury
from any localized differential displacements, these
types of applications require a firm pavement, similar
to that of residential driveways. For plazas, however, a
minimum 4 in. (102 mm) compacted aggregate base
and subbase typically are used (see Figure 4). Note
that for these applications on sites consisting of silty
or clayey soils, geotextile should be placed on the
compacted subgrade, below the subbase.
The sand setting bed thickness should be 1 to 1 in.
(25 to 38 mm). The base typically consists of coarse
aggregate (gravel) of varying gradation, compacted
to a minimum thickness of 4 in. (102 mm) using
mechanical tamping or vibration.
Min. 4 in. (102mm) Compacted Aggregate Base
Clay Pavers
Min. 1 to Max.
Sand Setting Bed
1

in. (25 to
38 mm)
1
2 /
Compacted Subgrade
Min. in. (1.6 mm) to Max.
in. (4.8 mm) Sand Filled Joints
1
16
3
16
/
/
Geotextile (If Required)
Figure 2
Typical Residential Patio or Walkway
Min. 4 in. (102 mm) Compacted Aggregate Subbase
Clay Pavers
Compacted Subgrade
Min. 4 in. (102 mm) Concrete, Compacted
Aggregate or Asphalt Base
Min. in. (1.6 mm) to Max.
in. (4.8 mm) Sand Filled Joints
1
16
3
16
/
/
Geotextile (If Required)
Min. 1 to Max.
Sand Setting Bed
1

in. (25 to
38 mm)
1
2 /
Figure 3
Typical Residential Driveway
Geotextile (If Required)
Min. 4 in. (102 mm) Compacted Aggregate Subbase
Clay Pavers
Compacted Subgrade
Min. in. (1.6 mm) to Max.
in. (4.8 mm) Sand Filled Joints
1
16
3
16
/
/
Min. 4 in. (102 mm) Compacted Aggregate Base
Min. 1 to Max.
Sand Setting Bed
1

in. (25 to
38 mm)
1
2 /
Figure 4
Typical Commercial/Pedestrian
Public Plaza/Sidewalk
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 3 of 13
Light Duty Vehicular. For parking areas and
neighborhood streets serving light duty vehicles, the
brick pavement section should be similar to that of
a residential driveway, but with a more substantial
base. A pavement with a concrete base as depicted
in Figure 5 or a thicker aggregate or asphalt base is
required.
Heavy Duty Vehicular. Paving systems exposed
to more than 251 daily equivalent single axle loads
(ESAL) from trucks or combination vehicles having
three or more loaded axles are considered heavy
duty vehicular applications. Such paving systems are
beyond the scope of this Technical Note series. For
further information about heavy vehicular applications,
refer to Flexible Vehicular Brick Paving A Heavy
Duty Applications Guide [Ref. 6].
GENERAL DESIGN AND DETAILING CONSIDERATIONS
Interlock
Sand-set pavers interlock with one another by generating friction across the joints. This is the result of tightly
packing sand into the joints during the vibration process. The interlock improves as the pavement is subjected to
traffic. There are three types of interlock present in a sand-set paver pavement when properly constructed: vertical,
horizontal and rotational interlock. Interlocked pavers cannot be readily extracted from the pavement.
Vertical interlock allows load transfer across joints between pavers. When a load is applied to one paver, a portion
is transferred through sand in the joints to adjacent pavers, as shown in Figure 6, distributing the load to a greater
area and reducing the stress on the sand bed and the underlying layers. Vertical interlock allows a paving layer to
act as a structural layer. Without vertical interlock, the pavers do not act as a structural layer, and localized stress
on the setting bed directly under a loaded paver is increased. Pavers installed on a sand setting bed should not be
laid with 1/4 in. (6.4 mm) joints, because this is too wide to achieve interlock, making the pavers unable to transfer
load to adjacent pavers. The proper joint width is 1/16 to 3/16 in. (1.6 to 4.8 mm).
Rotational interlock is the result of lateral resistance from adjacent pavers and adequate edge restraints, as shown
in Figure 7. It is improved with full joints that support the top of the paver. Without adequate restraint, the pavers
can roll in the direction of lateral loading, which may result in an irregular surface profile.
Min. 4 in. (102 mm) Concrete or 8 in.
(204 mm) Compacted Aggregate Base
Min. 4 in. (102 mm) Compacted Aggregate Subbase
Clay Pavers
Compacted Subgrade
Min. in. (1.6 mm) to Max.
in. (4.8 mm) Sand Filled Joints
1
16
3
16
/
/
Min. 1 to Max.
Sand Setting Bed
1

in. (25 to
38 mm)
1
2 /
Figure 5
Typical Light Duty Vehicular
Narrow Joints
Vertical Interlock
Wide Joints
No Vertical Interlock
Load
Load
Figure 6
Vertical Interlock of Pavers
No Rotational Interlock
Load
Figure 7
Rotational Interlock of Pavers
Load
Rotational Interlock
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 4 of 13
The extent of horizontal interlock depends upon the
laying (bond) pattern of the pavers and the edge
restraint. Patterns that have staggered joint lines allow
the load to be distributed to a larger number of pavers,
as shown in Figure 8. This reduces joint compressive
stress and potential for horizontal creep of pavers.
Continuous joints result in minimal load distribution
and increased joint compressive stress, which may
produce horizontal movement.
Pavement Section
Clay pavers over a sand setting bed can be installed
over a flexible or rigid base, including aggregate,
asphalt, cement-treated aggregate or concrete
bases. For further information on bases, refer to
Technical Note 14.
The design of the base is beyond the scope of this
Technical Note, and the advice of a qualified and
experienced pavement designer should be sought.
For preliminary design, it is reasonable to assume
that a minimum of 4 in. (102 mm) of concrete,
cement-treated aggregate, asphalt or aggregate base
will be needed for sand and gravel subgrades. For
residential driveway, commercial/pedestrian and light
duty vehicular applications with clay or silt subgrades,
an additional 4 in. (102 mm) of aggregate subbase
or base should be added to each option. Additional
thickness may be required when the subgrade is
susceptible to frost heave or when the pavement must
support heavy axle loads from trucks.
Concrete bases should be reinforced with welded wire
fabric or reinforcement bars and should have control
joints spaced at 12 ft (3.66 m) intervals to control
expansion and contraction. To minimize movement of
slabs, detail movement joints as shown in Figure 9.
Control joints in suspended structural slabs should
extend through the entire slab and align with an
expansion joint through the pavement above. Control
joints should have dowels or a keyway to limit vertical
separation across the joint.
Vehicular Traffic
For light duty vehicular paving systems, a maximum
traffic speed of 30 mph (50 kph) is considered
appropriate for pavers in a sand setting bed. When
frequent vehicular traffic is anticipated, additional
attention is required to ensure that joints between
pavers remain filled with sand. Higher speed
applications require more vigilance, as the interlock
between pavers is reduced with sand loss. Paving
systems for vehicular traffic applications usually will
include a compacted subbase to distribute loads (see
Figure 5).
Figure 8
Horizontal Interlock of Pavers
No Horizontal Interlock
Horizontal Interlock
Load
Load
Clay Pavers
Concrete Base
Sand Setting Bed
Control Joint
with Dowels
Compacted Subbase
Compacted Subgrade
Figure 9
Control Joints
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 5 of 13
The designer should consider the bond pattern for vehicular traffic applications. Any pattern may be used under
foot traffic. When vehicles operate on a pavement, patterns that distribute horizontal loads (i.e., loads from turning,
accelerating or braking vehicles) across multiple pavers, such as herringbone, are recommended. Patterns with
continuous joints, such as stack bond or running bond, are more susceptible to creep from horizontal loading.
Where such patterns are used in vehicular pavements, continuous joint lines should be oriented perpendicular to
the direction of vehicle travel.
Bond Patterns/Layout
The size of pavers may influence the selection of a suitable bond pattern. Pavers for use on a sand setting bed
typically are manufactured in sizes that accommodate a joint width of approximately 1/8 in. (3.2 mm) to encourage
optimal interlock.
Bond patterns such as herringbone, basket weave and others make use of the 1:2 or 1:3 ratios between the
pavers length and width to maintain the pattern and joint alignment. Pavers sized to accommodate joint widths
of approximately 3/8 in. (9.5 mm) do not achieve these ratios. Such pavers typically are used in pavements with
mortar joints. When they are laid on a sand setting bed, only a running bond, stack bond or chevron pattern should
be used, since these patterns do not depend on these ratios.
An individual clay pavers dimensions may be slightly different from the dimensions of another clay paver from the
same run. The inherent variability of their dimensions is a result of their manufacturing process. Pavers may be
larger or smaller within allowable tolerances of their specified size. This variability may not be consistent, because
actual dimensions may be greater or smaller than the specified dimensions. As such, the pavers may not be
able to be placed in a standard modular pattern. Blending of pavers from multiple cubes during installation can
overcome this issue. The installer should constantly monitor paver size during installation to ensure that the bond
pattern and joint size are maintained.
When designing an installation pattern with changes in bond and color, incorporating some tolerance in the
placement of certain paver features is recommended. This can be achieved by using saw cut pavers at junctions
of colored areas or by allowing approximate dimensions and realistic tolerances when placing certain paver
features. Two examples are depicted in Figure 10.
Field Pattern Brick
Cut as Necessary
to Ensure Good Fit
6 to 8 in. (152 to 204 mm)
Header Course Cut as
Necessary to Ensure
Good Fit
Edge of Band Aligned with
Building Column +/-
1
2 in.
(12.7 mm) to Ensure Good Fit
/
Figure 10
Pattern Options to Maintain Specified Joint Widths
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 6 of 13
Compacted Aggregate Base
Heavy Plastic or
Metal Edging
Secured with Spikes
Clay Pavers on
Sand Setting Bed
Compacted Subgrade
(a) Proprietary Edge Restraint System
Compacted Aggregate Base
Clay Paver,
Precast Concrete
or Cut Stone
Extend Base Beyond Restraint
Compacted Subgrade
Clay Pavers on
Sand Setting Bed
Concrete or Mortar as Required
Figure 11
Edge Restraints
(b) Clay Paver, Precast Concrete
or Cut Stone Edge Restraint
Edge Restraints
Edge restraints are critical in a pavement with a
sand setting bed to enable consistent interlock and
resist horizontal loads transferred from pavers.
Selection of edge restraint will depend on pavement
section and use. Figure 11 (pages 6 and 7) presents
various options, in increasing order of load capacity.
Concrete curbs or steel angles attached to a concrete
foundation or concrete base layer are the most
robust edge restraints. They are recommended for all
pavements subject to regular vehicular traffic. Edge
restraints for other applications may include pavers
bonded to a concrete foundation, and a range of
proprietary plastic and metal edge restraint systems
that are typically spiked into aggregate bases. Timber
edging and concrete backing poured to restrain edge
pavers may not be effective over the long term. It is
important that all edge restraints have a vertical rather
than inclined face for the pavers to butt against.
Base as Selected
Poured Concrete,
Precast, or Cut
Stone Curb
Clay Pavers on
Sand Setting Bed
Compacted Subgrade
(d) Curb Edge
Base as Selected
Clay Paver Bonded
to Concrete
Concrete Foundation
Compacted Subgrade
Clay Pavers on
Sand Setting Bed
(c) Bonded Clay Paver Edge Restraint
Galvanized Steel or Aluminum Angle
Compacted Subgrade
Concrete Base
Clay Pavers on
Sand Setting Bed
(e) Steel Angle Edge Restraint
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 7 of 13
Compacted Aggregate
Subbase (If Required)
Clay Paver,
Precast Concrete
or Cut Stone
Concrete Foundation
Compacted Subgrade
Concrete or
Asphalt Base
Clay Pavers on
Sand Setting Bed
Min. 1 in. (25.4 mm)
dia. Weep @ 24 in.
(610 mm) o.c.
Screen as
Necessary
(f) Clay Paver, Precast Concrete
or Stone Edge Restraint
Base as Selected
Concrete
Curb & Gutter
Compacted Subgrade
Pavers Set
1
8 in.
(3.2 mm) Higher than
Concrete After
Vibration of Pavers
/
(g) Poured Concrete Curb and Gutter Edge Restraint
Figure 11 (continued)
Edge Restraints
Drainage
Adequate drainage is important to the performance and durability of any clay paving system. Water should be
drained from the paving system as quickly as possible. A minimum slope of 1/4 in. per foot (2 percent grade) is
recommended. Adequate drainage should be provided to ensure the integrity of all layers in a paving system.
A sand setting bed will continue to consolidate slightly after construction is complete. Pavers should be finished
slightly higher than drainage inlets and other low edges of a pavement. This will minimize water puddling at these
locations. Typically 1/8 in. (3.2 mm) will be adequate and will not present a short-term tripping hazard.
Over time, small amounts of water will migrate through sand joints. Consequently, a sand-set paving system with
an impermeable base will require weep openings at low points in the pavement. Weep openings permit moisture
to seep out of the pavement rather than saturating the setting bed. Even a well-compacted aggregate base
may benefit from installing weep openings. Sand is less durable in a saturated state than when dry or slightly
damp.
Several weep opening options are available. A small-diameter (1 to 2 in. [38 to 51 mm]) pipe with ends wrapped
in geotextile may be placed through the side wall of drain inlets or through edge restraints. Such weeps should be
installed at spacings of 2 to 6 ft (0.60 to 1.83 m) depending on pavement geometry and profiles, environmental
conditions and pavement use. As an alternative, a drainage mat may be placed vertically through the base. This
may be used in conjunction with small pipes at drain inlets. For a concrete base, holes may be drilled or formed
through the slab to weep water to the subbase. Locating holes away from the impact of wheel loads is necessary
since subbase materials may be moisture-sensitive.
Penetrations
Large and small features that penetrate through the paver layer should be properly detailed. These features
include utility covers, tree pits, light pole bases, signposts and street furniture. Features may either penetrate the
entire pavement section to an independent structure or foundation, or be anchored to a concrete subslab. Such
features can present some issues in cutting the pavers to form a uniform joint around them.
Some utility covers and other frames are relatively shallow, or have buttresses, inclined faces, anchor bolts
or other features that may interfere with the bottom of a paver. Where possible, features should be specified,
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 8 of 13
Compacted Subgrade
Anchor Bolts
Concrete Plinth
on Foundation
Pole with Base Plate
Sand
Setting Bed
Clay Pavers
Cover Plate
Base as
Selected
Figure 13
Small Penetration
Compacted Subgrade
Soldier Course
in Thinset Mortar
Compacted Base
Concrete Surround
Manhole Wall
Frame with Inclined
Outer Face
Sand Setting Bed
Clay Pavers Set in. (3.2 mm) Higher
than Thinset Pavers After Vibration
1
8 /
Figure 12
Large Penetration
designed and installed deeper than the setting bed. Where this is not possible, casting a concrete collar around
the frame and thin-setting a header course of pavers on the concrete may clear obstructions to the sand setting
bed interface, as shown in Figure 12.
Accurately cutting and placing pavers against small features may prove difficult. An alternative is to construct a
concrete plinth up to the pavement surface and to install a cover plate to conceal the anchorage of the feature, as
shown in Figure 13. This also allows easy access for repairs, without removing pavers.
MATERIALS
Subgrade
For design purposes, the subgrade is considered to be either sand/gravel or clay/silt. The latter are more sensitive
to moisture and frost and may require the use of subbase layers and proper drainage to protect against shrinkage,
swelling and frost heave. The advice of a properly qualified and experienced pavement designer should be sought
in regard to the preparation of the subgrade.
Base and Subbase
Base materials for pavers laid in a sand setting bed may be of aggregate, cement-treated aggregate, asphalt
or concrete. When a subbase is required, aggregate generally is used. Aggregate materials should comply with
ASTM D 2940 and be compacted in accordance with ASTM D 698 to 95 percent maximum density. Asphalt should
meet ASTM D 3515. Concrete should have a minimum compressive strength of 4,000 psi (27.6 MPa) and should
have control joints spaced a maximum of every 12 feet (3.66 m). For a more detailed discussion of base and
subbase materials, refer to Technical Note 14.
Geotextiles
Geotextiles are used on top of silt or clay soils to help stabilize subgrades and under sand setting beds to prevent
loss of sand through weep openings and other gaps in the pavement base or at edge restraints or penetrations.
The preferred type of geotextile is a woven, polypropylene fabric complying with ASTM D 4751, Test Method for
Determining Apparent Opening Size of a Geotextile [Ref. 5], with an approximate opening size from a No. 70 to
No. 100 sieve size opening. Nonwoven geotextiles can be used for light-traffic applications. Geotextiles should
be lapped at the sides and ends of rolls a minimum of 12 in. (305 mm). Care should be taken to not locate laps
directly under anticipated wheel paths. Geotextiles should extend 6 in. (152 mm) beyond potential areas of sand
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 9 of 13
loss. These may be adhered in place, but generally will stay in position once covered by the sand setting bed.
Geotextiles should not be allowed to span over unfilled holes or pits in the surface of the base that are greater
than 1 in. (25.4 mm).
Setting Bed Sand
A sand setting bed provides a strong support layer under pavers and accommodates variations in paver thickness
to produce a smooth surface profile. A portion of setting bed sand penetrates the joints during vibration and
initializes the development of interlock between the pavers. Sand for the setting bed should be clean, naturally
occurring material with angular and subangular shaped particles, with a maximum size of about 3/16 in. (4.8 mm).
Concrete sand conforming to the requirements of ASTM C 33, Specification for Concrete Aggregate [Ref. 1], or
local department of transportation standards is recommended for use as setting bed material. This provides a
more stable and durable setting bed than mason sand or screenings, which have a more rounded shape and
should not be used. Sand rich in silica-based minerals is desirable, because carbonate-based minerals are softer
and can break down when saturated. Manufactured limestone sand usually causes efflorescence and should be
avoided unless it has a proven track record on similar projects.
Clay Pavers
A wide selection of colors and textures is available in clay pavers. Further information on clay pavers can be found
in Technical Note 14.
Pavers generally are manufactured with their length equal to a module of their width. Two commonly specified
clay paver sizes are 4 in. wide by 8 in. long (102 by 203 mm) and 3 in. wide by 7 in. long (95 by 190 mm).
Other similar sizes are available, such as 3 in. wide by 7 in. long (92 by 194 mm), and several manufacturers
are able to provide custom sizes. Common specified thicknesses are 1 in. (38 mm), 2 in. (57 mm) and
2 in. (70 mm) [2 in. (67 mm) excluding chamfered edge].
All clay pavers covered by ASTM C 902, Specification for Pedestrian and Light Traffic Paving Brick [Ref. 3], and
ASTM C 1272, Specification for Heavy Vehicular Paving Brick [Ref. 4], can be installed on a sand setting bed.
The designer should select the appropriate Application, Type and Class of the paver for the project based on
aesthetics, use, abrasion resistance and the required resistance to damage from weather exposure. For more
detailed information on specifying clay pavers, refer to Technical Note 14.
When square-edged pavers or pavers without lugs are laid with sand joints, care should be taken to ensure that
they do not make direct contact with or lip under adjacent pavers. A minimum 1/16 in. (1.6 mm) wide sand-filled
joint should separate each clay paver to minimize potential chipping. However, the maximum joint width should be
no more than 3/16 in. (4.8 mm) to minimize the potential for horizontal movement under vehicular traffic. If pavers
with spacers and/or a rounded or chamfered edge are installed, there is less potential for direct paver contact.
When lugs are used, the potential for creep is reduced.
Jointing Sand
Sand within pavement joints creates interlock between pavers by generating friction across the joint. Larger
particles present in joints reduce the potential for lateral movement. Finer particles act to reduce contact stresses
around the larger particles, reducing the potential of the particles breaking down. The sand also accommodates
the variations in paver size and reduces the potential for contact between pavers that can lead to chipping. ASTM
C 33 concrete sand should be placed in joints before vibration to maximize interlock at the bottom portion of
joints. However, coarse particles that do not fall into joints should be brushed off the pavement surface rather than
worked in. After vibration, finer jointing sand may be placed so that it penetrates to the bottom of the joints and
achieves better filling. When the typical joint dimension exceeds 3/16 in. (4.8 mm), stabilized sand or joint sand
stabilizer should be used.
Joint Sand Stabilizers
In conditions where potential sand loss or high joint permeability may not be desirable, a joint sand stabilizer is
recommended. These conditions include intensive cleaning practices, high surface water flows and flat areas with
moisture-sensitive subgrades. There are several different types of joint sand stabilizers. These include breathable
polymeric liquids that can be sprayed onto the pavement surface and squeezed into the joints with a squeegee, as
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 10 of 13
well as dry products that can be mixed with the joint sand before installation. Pretreated sands also are available
for joint filling. Strict adherence to the stabilizer manufacturers recommendations is required to achieve successful
installations. When selecting a stabilizer, it is important to choose one with a proven history that does not discolor
the surface or peel over time. The paver manufacturer's recommendation regarding joint sand stabilizers should
be followed. Joint sand stabilizers should be applied to the completed paver surface. Stabilizers should be applied
to the pavement surface before the application of other coatings to enhance the appearance of the pavers or to
protect against staining. For further guidance on selecting coatings for use on brick pavements, refer to Technical
Note 6A.
INSTALLATION AND WORKMANSHIP
Subgrade
The subgrade should be brought to the proper level and cleared of organic material. Compaction should comply
with ASTM D 698 to 95 percent maximum dry density for clay and 100 percent maximum dry density for
sand/gravel. For a more detailed discussion of subgrade preparation, refer to Technical Note 14.
Base and Subbase
Base and subbase materials should be placed per the design. Aggregate should be compacted in accordance with
ASTM D 698 to 95 percent maximum density. The maximum variation under the setting bed should be +/- 3/16
in. (4.8 mm) when a 10 ft (3.05 m) straightedge is laid on the surface. The minimum slope of the concrete base
surface should be 1 in. (25.4 mm) in 4 ft (1.22 m) to allow for drainage. For a more detailed discussion on the
installation of base and subbase materials, refer to Technical Note 14.
Setting Bed
Whenever possible, the direction of installation should be planned to protect the paving against premature use or
damage by rain or other construction activities. The surface of the underlying base material should be thoroughly
clean and dry before installation of the bedding sand. Elevations should be verified to ensure that the sand setting
bed will be a consistent thickness after compaction. The setting bed should not be used to bring the pavement
to the correct grade. Isolated high and low spots should be corrected before sand placement to avoid an uneven
pavement surface resulting from variable sand setting bed thicknesses. Lines should be established for setting out
the pattern. The contractor should become aware of size variations in the pavers to maintain the pattern without
localized opening or closing of joints to meet a fixed edge. All areas of potential sand loss should be covered with
geotextile.
Screed rails should be set on the surface of the base to proper line and level. They are typically placed 8 to 12 ft
(2.44 to 3.66 m) apart, or closer when working on a grade. An allowance should be made in the thickness of the
setting bed for compaction of bedding sand as pavers are installed, as well as additional consolidation in service.
An experienced contractor will be aware of the proper thickness for different conditions to achieve the correct long-
term surface profiles. The bed thickness should be established so that when the pavers are compacted, their top
surface will be 1/8 in. (3.2 mm) above the required grades to allow for limited settling in service.
To prevent disturbance, setting bed sand should not be spread too far ahead of the paver laying face. Voids left
after removing the screed rails should be filled. The screeded bedding sand may be affected by wind or rain as
well as by wayward construction operations. If sand is disturbed, it should be loosened and rescreeded. Extensive
areas of screeded sand should not be left overnight unless they are properly protected from disturbance and
moisture. Moisture content of setting bed sand should be kept as uniform as possible to minimize undulations in
the pavement surface. The sand should be kept in a damp condition conducive to packing. Water should not be
applied except by very light misting. Stockpiled sand should be covered to protect it from wind and rain.
Paver Installation
The pavers are laid on the setting bed working away from an edge restraint or the existing laying face while
following the pattern lines that have been established. Full pavers should be laid to the required pattern with
1/16 to 3/16 in. (1.6 to 4.8 mm) wide joints. The optimum joint width for vehicular traffic is between 1/16 and
1/8 in. (1.6 and 3.2 mm), but some wider joints may be required with Application PS pavers, and particularly with
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 11 of 13
Application PA pavers. Lugs enable the correct joint width to be achieved when the pavers are placed in contact
with one another. Pavers should not be forced together, resulting in excessive contact, because this may cause the
pavers to chip during installation or compaction. At least two cubes of each color of pavers should be drawn from
at one time, and the manufacturers recommendations on color blending should be followed. The pavers should be
adjusted to form straight pattern lines while maintaining the correct joint widths.
Several feet of pavers should be installed before beginning to add cut pavers as infill against edge conditions.
Bench-mounted masonry saws are the best means of cutting the pavers to achieve a neat edge and a vertical cut
face. Use of a wet saw or dust collection system is recommended to control dust. Guillotine cutters also may be
used, but their cuts typically are not as straight and neat. Convex curves can be formed using multiple cuts, but
this requires a skilled craftsman to meet allowable joint tolerances. Concave curves are very difficult to form and
should be avoided when possible.
Pavers should be compacted at the end of each day to prevent any damage while left unattended. The pavement
surface should be compacted using a plate compactor. These typically have a plate area of 2 to 3 sq ft (0.23 to
0.28 sq m) and operate at a frequency of 80 to 100 Hz. To prevent pavers from chipping during vibration, a little
bedding sand material can be swept into the joints, or the underside of the plate compactor can be fitted with
a rubber mat. Pavers also can be covered with a sheet of geotextile or sheets of plywood during vibration For
molded pavers, vibration is especially important since irregularities and dimensional variations on the underside
could lead to air gaps or improper support if not properly compacted into the sand setting bed. Compaction should
not be carried out within 4 ft (1.2 m) of unfinished edges.
The vibrated surface should be slightly above adjacent pavement surfaces, drainage inlets and channels to allow
for secondary compaction of the bedding layer under traffic. The maximum variation in surface profile should be
less than 3/16 in. (4.8 mm) in 10 ft (3.05 m). Water should drain freely from the surface and not form puddles.
Lipping between adjacent pavers should not be greater than 1/8 in. (3.2 mm) if the pavers have chamfers, or
1/16 in. (1.59 mm) if they have square edges.
After vibration of the pavers to finished elevations, dry fine-grained jointing sand is brushed over the surface of
the pavement and additional vibration is undertaken until all of the joints are completely filled with sand. Surplus
jointing sand should be maintained on the surface to enhance the process of joint filling. Typically the sand should
be level with the bottom of the chamfer or approximately 1/8 in. (3.2 mm) below the top of square edge pavers.
Joint Sand Stabilizers
The paver manufacturer's recommendation regarding joint sand stabilizers should be followed. Jointing sand that
is pretreated with a stabilizer product should be brushed or blown off the pavement surface as soon as possible
and not be allowed to become stuck in the surface texture of the pavers. If pretreated sand or a joint sand additive
is used, the stabilizer should be activated by lightly misting the surface with water. If a liquid joint sand stabilizer
is used, it should be sprayed onto the pavement surface and forced into the joints with squeegees. It may be
necessary to fill the tops of the joints with the liquid several times before it sets to achieve adequate penetration.
The stabilizer manufacturers instructions should be followed closely, because each stabilizer is slightly different.
Probing several joints to verify that the sand is stabilized to an adequate depth of approximately two times the joint
width rather than just forming a crust is recommended.
MAINTENANCE
Cleaning
Sand-set pavers can be kept clean in most environments by regular sweeping. In situations that lead to a greater
degree of buildup of grease, tire marks or other stains, the pavers can be cleaned by pressure-washing. The sand-
filled joints generally are resistant to this treatment if the nozzle surface is clear and the water jet is not directed
along the joints. Aggressive pressure-washing can cause localized removal of the joint filling material and can
even undermine the pavers. More stubborn stains, including paint and gum, can be cleaned by scraping off the
hard residue and then scrubbing with a stiff-bristled brush and a proprietary cleaner or scouring powder. In damp
or shady areas where moss and lichen have grown in the joints, these can be killed with a bleach-water mixture or
with proprietary treatments.
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 12 of 13
Snow Removal
Snow prevention and removal can be carried out by hand, by machine or by chemicals. Hand methods include
shovels and brooms. Mechanical methods include snow blowers, snowplows, and buckets or brushes attached to
tractors. Shovel and machine removal methods can chip the edges of the pavers, particularly if excessive lipping is
present. This equipment should be properly adjusted so that it does not damage the pavement surface.
Skid-steer snow removal equipment also may move pavers, causing distortion of pattern lines and some chipping
of the pavers if the equipment is driven aggressively. When tractor and particularly skid-steer mounted equipment
is used, the pavement must be able to support the wheel loads without damage.
A range of anti-icing and deicing chemicals are used for pavements. Deicing chemicals can cause thermal shock
in a pavement by supercooling the pavement surface. This can lead to spalling or surface damage on pavers
of Class NX or MX pavers. Deicing agents should be used with care, as chemical residue left on the surface can
penetrate into the joints and result in staining and efflorescence. Class NX should not be used where subject to
freezing.
Resanding
Over time, due to wind, rain and other means, the sand within the top portions of joints can be eroded. Therefore,
the joints should be periodically resanded using the same methods described above for applying jointing sand after
vibration of the pavers.
Repairs
Underground utilities frequently pass beneath paved areas on congested sites. Access to these utilities frequently
is required for repair or to install new lines. Sand set pavers readily accommodate such work, as they can be
removed and reinstated with little evidence of the work having been carried out. Repairs to the paving also can be
made if they are overloaded or otherwise damaged.
Removal can be undertaken by prying or breaking out the initial paver so that it can be removed without damaging
adjacent units. It is then possible to work the adjacent pavers loose using a hammer and chisel or pry bars in
the joints and under the paver. Some chipping of the pavers should be expected, and a few spare pavers will be
required for reinstatement. The bedding sand can be removed as necessary. Traffic should be kept at least 4 ft
(1.2 m) from the unrestrained edge. If a trench is open for a significant amount of time, the adjacent pavers should
be temporarily restrained to stop them from moving laterally. Trenches should be filled with proper care paid to
compaction of the backfill. The base should be replaced to match the original section.
To reinstall the pavers, the bedding sand should be replaced with an adequate pressure to allow for compaction.
The pavers should be replaced in the appropriate pattern and fresh sand spread into the joints. The repair area
should be leveled by hammering on a wooden pack if the area is small or with a plate vibrator if it is large enough.
The joints should be refilled with sand and new stabilizer applied if necessary.
SUMMARY
Pedestrian and light duty vehicular pavements of clay pavers laid on a sand setting bed provide the most cost-
effective system for pedestrian and light duty vehicular pavement. When properly constructed, the interlock of the
pavers provides the necessary stability for the desired service life of the pavement. This Technical Note provides
the basic information required to properly select materials, design, detail and construct brick pavements over sand
setting beds. Further information about the properties of other brick pavements and concepts not unique to sand
setting beds is discussed in the Technical Note 14 series.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
www.gobrick.com | Brick Industry Association | TN 14A | Paving Systems Using Clay Pavers on a Sand Setting Bed | Page 13 of 13
REFERENCES
1. ASTM C 33, Standard Specification for Concrete Aggregate, Annual Book of Standards, Vol. 04.02,
ASTM International, West Conshohocken, PA, 2006.
2. ASTM C 144, Standard Specification for Aggregate for Masonry Mortar, Annual Book of Standards, Vol.
04.05, ASTM International, West Conshohocken, PA, 2007.
3. ASTM C 902, Standard Specification for Pedestrian and Light Traffic Paving Brick, Annual Book of
Standards, Vol. 04.05, ASTM International, West Conshohocken, PA, 2007.
4. ASTM C 1272, Standard Specification for Heavy Vehicular Paving Brick, Annual Book of Standards, Vol.
04.05, ASTM International, West Conshohocken, PA, 2007.
5. ASTM D 4751, Standard Test Method for Determining Apparent Opening Size of a Geotextile, Annual
Book of Standards, Vol. 04.13, ASTM International, West Conshohocken, PA, 2007.
6. Flexible Vehicular Brick Paving A Heavy Duty Applications Guide, Brick Industry Association, Reston,
VA, 2004.

Technical Notes 15 - Salvaged Brick
May 1988
Abstract: The use of salvaged brick in new building construction is discussed. Factors affecting the selection include
altered physical properties (durability), esthetics, economics, building code requirements and experimental testing.
Key Words: brick, masonry, mortar bond, salmon brick, salvaged brick.
INTRODUCTION
Selecting building material requires the consideration of four factors: esthetics, design properties, economics and
required level of performance. Salvaged brick are occasionally selected for their "rugged appearance" and
sometimes for their low initial cost. Rare is the case when salvaged brick are chosen for their design properties. In
general, walls using salvaged brick are weaker and less durable than walls constructed of new brick masonry units.
Most salvaged brick are obtained from demolished buildings which stood 40 to 50 yr, or more. In fact, it may be next
to impossible to salvage brick from modern structures which use brick set in portland cement mortars. When brick
are initially placed in contact with mortar, they absorb some particles of the cementitious materials. It is virtually
impossible to completely clean these absorbed particles from the surfaces of the brick units. This may greatly affect
the bond between brick and mortar when reused.
MANUFACTURING METHODS
In the early 1900's, manufacturing methods were markedly different from those of today. De-aired brick were
unknown; coal- and wood-fired periodic and scove kilns were commonplace. The modern solid, liquid or gas-fired
tunnel kilns with accurate temperature controls throughout were also unknown. Manufacturing conditions years ago
were generally such that large volumes of brick were fired under greater kiln-temperature variations than could be
tolerated today. These conditions resulted in a wide variance in finished products. Brick from the high-temperature
zones were hard-burned, high-strength, durable products; those from low-temperature zones were under-burned,
low-strength products of low durability. These temperature variations also resulted in a wide range in absorption
properties and color. The under-burned brick were more porous, slightly larger, and lighter colored than the harder-
burned brick. (It is the nature of ceramic products to shrink during firing. Generally, for a given raw clay, the greater
the firing temperature, the greater the shrinkage and the darker the color.) Their usual pinkish-orange color resulted
in the name salmon brick.
During these bygone years, prevalent methods of construction made production of both hard-burned and salmon
brick economically feasible. Most masonry buildings had loadbearing brick walls which were a minimum of 12 in. in
thickness. The hardest, most durable units were used in exterior wythes; the salmons (and others) were used for the
interior wythes and were not exposed to the exterior elements. Much sorting and grading of brick was performed at
the construction site by the mason, although the brick manufacturers eventually assumed this responsibility.
The advent of skeleton frames marked the beginning of high-rise construction and the gradual demise of thick
loadbearing masonry wall construction. (Despite the reduction in its use, loadbearing remains a very economical
method for constructing low- and mid-rise buildings). Architects and engineers began to design non-loadbearing
walls, and gradually decreased wall thicknesses. This evolution in design and construction techniques necessitated a
change in brick manufacturing procedures. Slowly but surely, the demand for salmon brick dwindled. After the use of
hollow backup units became prevalent, the need for salmon brick became practically non-existent. At the same time,
having invented the thinner, lighter weight panel wall, designers focused their attention on wall strength which they
equated to compressive strength of the individual brick.
Because the principal demand was for high compressive strength and durability, manufacturers had to produce a
http://www.gobrick.com/BIA/technotes/t15.htm
1 of 6 9/13/2009 12:50 PM
high proportion of well-burned brick. This demand necessitated a change in manufacturing methods. Thus, an
evolution in design and construction techniques brought on a significant and beneficial evolution in the production of
brick. (For a synopsis of present day manufacturing methods, see Technical Notes 9 Revised, "Manufacturing,
Classification and Selection of Brick; Manufacturing - Part 1").
MATERIAL SELECTION
Physical Properties. Several arguments are often advanced in favor of the use of salvaged brick. Among these are:
1. Because brick are extremely durable, they can be salvaged and used again.
2. If the brick were satisfactory at the time they were first used, they are satisfactory at present.
Both arguments are fallacious.
When brick are initially placed in contact with mortar, they absorb some water and some particles of cementitious
materials. The initial rate of absorption (suction) is an important factor which greatly affects the bond between brick
and mortar. Brick with extremely high or extremely low suctions do not develop good bond. (For discussion of bond
strength between mortar and new clay masonry units, see Technical Notes 8 Revised, "Mortars for Brick Masonry").
With salvaged brick, more factors influence bond. Pores in brick are filled with particles of lime, dirt and other
deleterious matter. Many bedding surfaces of salvaged brick will not be thoroughly clean, but will instead be covered
with mortar. The bond between new mortar and old mortar is not very strong. If the original mortar bond was weak,
the new bond will be adversely affected. The bond to salvaged brick is considerably less than to similar new brick
and has been demonstrated many times by comparative tests (see Experimental Tests Section in this Technical
Notes).
Most authorities agree that water penetration through masonry results from incompletely filled joints and incomplete
bond between brick and mortar. That is, water penetrates through flaws at joints rather than directly through the
brick and mortar. Thus, masonry walls of salvaged brick, with their inferior mortar bond, are likely to be more
susceptible to water penetration and weaker under lateral loading than similar masonry of walls constructed of new
units. The ultimate compressive strength of the walls will also be lower if salmon brick are present.
The durability of masonry depends upon the quality of materials and mortar bond. Generally, salmon brick do not
provide the same durability as new brick when exposed to weathering. With the thinner masonry walls of today, brick
are used primarily as a facing material to provide a weather resistant barrier of protection. Thus, many salmon brick
are eventually placed in exposed faces of walls constructed of salvaged brick. Even where solid brick walls are
used, many salmons are likely to be exposed to weathering, because it is impossible to accurately sort and grade
salvaged brick. With soft, highly absorptive salmon brick exposed to the weather, and with poor mortar bond
permitting excessive water penetration, it is quite likely that masonry of salvaged brick will spall, flake, pit, and crack
due to freezing and thawing in the presence of excessive moisture.
One common characteristic of most manufactured building materials is a reasonable degree of uniformity within a
particular grade or within a given manufactured lot. Salvaged brick lack this distinction. Hard-burned and soft-burned
brick, hopelessly mixed during wrecking operations, effectively create a material stockpile of two widely differing
grades of materials (see Figure 1). A sample of the material will contain specimens of each grade. If tested for
absorption or compression strength properties, the sample will show widely diversified characteristics. The average
absorption or strength will not approximate the true values for either grade, but will lie somewhere between. In
effect, it is difficult to determine whether salvaged brick will meet present day material specifications or building code
requirements.

http://www.gobrick.com/BIA/technotes/t15.htm
2 of 6 9/13/2009 12:50 PM

When existing walls are demolished, hard-burned brick and salmons are hopelessly mixed. It is virtually
impossible to distinguish between durable and non-durable units.
FIG. 1

Esthetics. Salvaged brick may satisfy the desire for a rugged, colorful masonry surface. Architects often desire the
extreme range of colors from dark-red to the whites and grays of units still partially covered with mortar. But most
frequently the light pink color of the salmon creates the desired effect. Unfortunately, the pink in salmons results
from under-burning which produces units that must not be exposed to weathering. Excessive disintegration due to
weathering can soon drastically alter the appearance originally desired (See Figs. 2 - 4).


A chimney of salvaged brick which has spalled considerably
within a relatively short time after construction (Knoxville, Tennessee).
FIG. 2

http://www.gobrick.com/BIA/technotes/t15.htm
3 of 6 9/13/2009 12:50 PM

A close-up of a wall indicating the excessive spalling that is likely to occur
where salvaged brick are exposed to weathering.
FIG. 3



Because of the greater likelihood that moisture will be present,
salvaged brick should not be used for exterior patios, walks, pavements, etc.
FIG. 4

All pink brick are not necessarily under-burned. During recent years the architectural demand for a variety in colors
has led to the extensive use of raw clays which burn other than dark red when fired to maturity. Today, among other
colors, many hard-burned, pink brick are available. These units may conform to the requirements for highest quality
under applicable ASTM specifications. (Many pink brick conform to Grades SW or MW (severe or moderate
weathering) under ASTM C 216 or C 62.
Many manufacturers blend different colored brick to provide a rustic appearance similar to salvaged brick. There are
advantages to using new brick: the architect may specify any desired color blending and may specify the desired
grade under ASTM specifications. Thus, he can obtain the desired esthetic effect without sacrificing durability or
strength, a feat which is nearly impossible to accomplish when using salvaged brick.
Economics. Although in many instances salvaged brick have sold for more money than new brick, a principal reason
for their use is their low prevailing initial cost. But initial economy often proves to be false economy. For example,
http://www.gobrick.com/BIA/technotes/t15.htm
4 of 6 9/13/2009 12:50 PM
labor costs are usually higher for salvaged brick due to the required sorting and cleaning of the units. Maintenance
costs for salvaged-brick masonry are very likely to exceed this initial cost considering: 1) cutting out and replacing
disintegrated units; 2) tuck pointing mortar joints to reduce leaks and repair cracks; and 3) repeated attempts at
waterproofing. (The dangers of coating masonry of under-burned units are discussed in Technical Notes 6A,
"Colorless Coatings For Brick Masonry". Under-burned units may undergo accelerated disintegration if impermeable
coatings are applied to the exterior wall face). In many cases, the initial economics of salvaged brick prove false and
result in higher total expenditures.
BUILDING CODE REQUIREMENTS (See references)
American Standard Building Code Requirements for Masonry, ANSI A41.1, Section 2.1.1 (appendix
commentary):
"Irrespective of the original grading of masonry units, compliance with code requirements of material which has been
exposed to weather for a term of years cannot be assumed in the absence of test. Much salvaged brick comes from
the demolition of old buildings constructed of solid brick masonry in which hard-burned bricks were used on the
exterior and salmon brick as back-up, and, since the color differences which guided the original brick masons in their
sorting and selecting of bricks become obscured with exposure and contact with mortar, there is a definite danger
that these salmon bricks may be used for exterior exposure with consequent rapid and excessive disintegration.
Before permitting their use, the building official should satisfy himself that second-hand materials are suitable for the
proposed location and conditions of use. The use of masonry units salvaged from chimneys is not recommended,
since such units may be impregnated with oils or tarry material."
National Building Code, Section 1401.2:
"Second-hand units: Brick and other second-hand masonry units which are to be reused, shall be approved as to
quality, condition and compliance with the requirements for new masonry units. The unit shall be of whole, sound
material, free from cracks and other defects that would interfere with its proper laying or use, and shall be cleaned
free from old mortar before reuse."
Standard Building Code, Section 1401.2:
"1401.2.1 Masonry units may be reused when clean, whole and conforming to the other requirements of this
chapter, except that the allowable working stresses shall be 50% of those permitted for new masonry units.
1401.2.2 Masonry units to be reused as structural units in areas subject to the action of the weather or soil shall not
be permitted unless representative samples are tested for compliance with the applicable requirements of 1402."
Uniform Building Code, Section 2406 (k):
"Reuse of Masonry Units. Masonry units may be reused when clean, whole and conforming to the other
requirements of this section. All structural properties of masonry of reclaimed units, especially adhesion bond, shall
be determined by approved test. The allowable working stresses shall not exceed 50 percent of that permitted for
new masonry units of the same properties."
EXPERIMENTAL TESTS
At various times interested parties have conducted tests to compare salvaged-brick masonry to masonry of similar
new brick. One of the more comprehensive series of tests was conducted many years ago by the Engineering
Experiment Station, University of New Hampshire. The following statements are from this test report: (Project No.
98, "Relative Adhesion of Mortars to New and Used Brick", for Star Brick Yard, Epping, NY (1934-1935)).
"The object of this study was to determine by laboratory methods the relative adhesion of different standard mortars
to new and used or reclaimed brick . . . (using only) those materials . . . that would generally be employed . . .
". . . as far as materials are concerned . . . a wall laid up with used or reclaimed bricks . . . differ(s) from one laid up
with new bricks . . . (only) in the adhesion of the mortar to the brick surfaces. It is this quality with which this study is
concerned."
http://www.gobrick.com/BIA/technotes/t15.htm
5 of 6 9/13/2009 12:50 PM
Four types of used brick were tested and compared to the same four types of new brick (a total of eight types).
(The report describes the eight brick types tested as including hard and soft, water-struck and sand-struck, new and
old brick). Seven different standard mortars were employed. In describing the testing procedures, the report states,
in part:
"The brick to be tested were selected and were cleaned of all loose particles of mortar which could be removed by
means of a hammer and wire brush. No attempt, however, was made to remove any particle of mortar, etc. which
adhered so firmly to the brick surface that pounding and wire brushing would not release it."
The bulk of the report is too large to reproduce in its entirety. However, the following excerpts from the conclusions
to the tests are of interest:
" . . . The adhesion of mortar to new (hard) bricks was materially greater than to second hand bricks."
"With but few exceptions the adhesion of mortar to hard bricks was far greater than . . . to soft bricks of the same
type."
"Without exception . . . failure of the mortar to adhere to the surface of used brick far exceeded the failures of the
joint between mortars and new brick. In other words, it appears that the capillary pores of the second hand brick
were so plugged . . . that the new mortar could not gain any appreciable hold on the surface of the brick."
" . . . (The tests indicate) that the adhesive strength of mortar to the hard brick exceeded its cohesive strength..."
"With (all) used brick . . . cohesive strength of the mortars exceeded many times the adhesive strength of the same
mortars to the surfaces of the brick."
" . . . within the limits of the test . . . relative adhesion of mortars to . . . reclaimed brick . . . (is) less than half what
can be expected if the same mortars are used with new brick of the same type and degree of hardness."
SUMMARY
This Technical Notes has discussed the use of salvaged brick in new brick masonry wall construction. The
considerations are based on existing knowledge and experience. No effort is made or implied that this is a total
discussion of the subject matter, since conditions vary widely throughout the country. However, it is a basis from
which the designer can decide on the use of salvaged brick in new masonry structures.
Final decisions on the use of the information and suggestions discussed in this Technical Notes are not within the
purview of the Brick Institute of America and must rest with the project designer, owner or both.
REFERENCES (Building codes undergo continual revision. The editions listed are those current as the publication
date of this Technical Notes).
1. American Standard Building Code Requirements for Masonry; ANSI A41.1; National Bureau of Standards
(Miscellaneous Publications 211); Washington, D.C.; July 15, 1954 (Reaffirmed 1970).
2. National Building Code, 1987 Edition; Building Officials and Code Administrators International; 4051 W.
Flossmoor Road, Country Club Hills, Illinois.
3. Standard Building Code, 1985 Edition; Southern Building Code Congress International; 900 Montclair Road,
Birmingham, Alabama.
4. Uniform Building Code, 1985 Edition; International Conference of Building Officials; 5360 South Workman
Mill Road, Wittier, California.
http://www.gobrick.com/BIA/technotes/t15.htm
6 of 6 9/13/2009 12:50 PM
2008 Brick Industry Association, Reston, Virginia Page 1 of 16
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
16
March
2008
Fire Resistance of Brick Masonry
Abstract: This Technical Note presents information about the fire resistance of brick masonry assemblies in loadbearing and
veneer applications. Fire resistance ratings of several brick masonry wall assemblies tested using ASTM E119 procedures are
listed. For untested wall assemblies, procedures are presented for calculating a fire resistance rating.
Key Words: balanced design, building codes, equivalent thickness, fire, fire resistance period, fire resistance rating, fire test.
Fire Resistance Requirements
Use the building code to determine the fire resistance
rating required for separations, corridors, exterior walls
and other building features
Use fire control systems, compartmentalization of space
or other balanced design approaches to lower required
fire resistance ratings
Determine whether fire resistance is needed for one side
or two sides of fire exposure
Assembly with Tested Fire Resistance Rating
Use wall construction prescribed by the building code or
testing agency to achieve fire resistance rating
For wall construction not prescribed by the building code,
include reference for test results in design documents
Assembly with Calculated Fire Resistance Rating
Determine minimum equivalent thickness required of
brick unit from tables in the building code or ACI 216.1/
TMS 0216 [Ref. 5]
Specify brick standard, brick size and void area to meet
the minimum equivalent thickness requirements
For multi-wythe masonry walls, determine contributions
from other wall components such as concrete, concrete
masonry, air spaces and plaster
Construction Details
Where assemblies with a fire resistance rating are
supported by other assemblies, specify that the support
assembly have an equal or greater fire resistance rating
Seal penetrations through assemblies with a fire
resistance rating with appropriate sealants or details to
maintain fire resistance rating
INTRODUCTION
Building codes and other local ordinances require critical building components to have a certain level of fire
resistance to protect occupants and to allow a means of escape. Several factors contribute to the level of fire
resistance required of a wall, floor or roof assembly, including whether combustible (wood) or noncombustible
(steel, concrete and masonry) construction is used. Other factors include the buildings use, floor area and height,
the location of the assembly, and whether a fire suppression system such as stand pipes or sprinklers is installed.
Definitions
Fire Resistance. The property of a building element, component or assembly that prevents or retards the passage
of excessive heat, hot gases or flames under conditions of use.
Fire Resistance Period. A duration of time determined by a fire test or method based on a fire test that a building
element, component or assembly maintains the ability to confine a fire, continues to perform a given structural
function or both.
Fire Resistance Rating. A duration of time not exceeding 4 hours (as established by the building code) that a
building element, component or assembly maintains the ability to confine a fire, continues to perform a given
structural function or both. A legal term defined in building codes for various types of construction and occupancies.
A fire resistance rating is based on a fire resistance period and usually given in half-hour or hourly increments. As
an example, a wall with a fire resistance period of 2 hours and 25 minutes may only attain a fire resistance rating
of 2 hours. It is also referred to as a fire rating, fire resistance classification or hourly rating.
SUMMARY OF RECOMMENDATIONS:
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 2 of 16
Determining a Fire Resistance Rating
Traditionally, a fire resistance rating has been established by testing. The most common test method used is ASTM
E119, Standard Test Methods for Fire Tests of Building Construction and Materials [Ref. 3]. In this test, a sample of
the wall must perform successfully during exposure to a controlled fire for the specified period of time, followed by
the impact of a stream of water from a hose.
This standard test, along with other ASTM fire test standards, is used to measure and describe the response of
materials, products or assemblies to heat and flame under controlled conditions, but does not by itself replicate
actual fire conditions in a building. Rather, the intent of the test is to provide comparative performance to specific
fire-test conditions during the period of exposure. Further, the test is valid only for the specific assembly tested.
Fire testing is expensive because each specific assembly must be tested by constructing a large specimen,
placing multiple monitoring devices on that specimen and subjecting the specimen to both a fire and a hose
stream. As a result, a calculated fire resistance method developed jointly by The Masonry Society and the
American Concrete Institute and based on past ASTM E119 tests has largely replaced further fire resistance
testing for masonry and concrete materials [Ref. 5].
FIRE RESISTANCE TESTING
ASTM E119 Test Method
The test methods described in ASTM E119 are
applicable to assemblies of masonry units and to
composite assemblies of structural materials for
buildings, including bearing and other walls and
partitions, columns, girders, beams, slabs and
composite slab and beam assemblies for floors and
roofs.
When fire testing a wall assembly according to ASTM
E119, a sample of the wall is built using the materials
and details of the assembly to be used in construction.
The specimen is then subjected to a controlled fire
until a failure occurs (termination point is reached) or a designated extent of time passes. ASTM E119 requires that
the air temperature at a distance of 6 in. (152 mm) from the exposed (fire) side of the specimen conform to the
standard time-temperature curve, as shown in Figure 1.
Wall Specimens. The area exposed to the fire must be at least 100 sq ft (9.3 m
2
) with no dimension less than 9 ft
(2.7 m). Non-bearing walls and partitions are restrained at all four sides, but bearing walls and partitions are not
restrained at the vertical edges. Nine thermocouples are placed on the side of the wall unexposed to the fire to
measure temperature rise.
Protected Steel Column Specimens. If the fire resistant material protecting the column is structural, the column
specimen must be at least 9 ft (2.7 m) tall, and acceptance is based on its ability to carry an axial load for the
duration of the fire test. If the fire resistant material is not structural, the minimum column height is 8 ft (2.4 m), and
acceptance is based on temperature rise on the surface of the column. Temperature rise is measured by placing a
minimum of three thermocouples on the column surface (behind the fire resistant material) at each of four levels.
Hose Stream Test. For most fire resistance ratings ASTM E119 requires that walls be subjected to both a fire
endurance test and a hose stream test. The hose stream test subjects a specimen to impact, erosion and cooling
effects over the entire surface area that has been exposed to the fire. The procedure stipulates nozzle size,
distance, duration of application and water pressure at the base of the nozzle. Some of these requirements vary
with the fire resistance rating. The hose stream test may be performed on a duplicate wall specimen that has
been subjected to a fire endurance test for one-half of the period determined by the fire test (but not more than 1
hour); or the hose stream test may be performed on the wall specimen immediately after the full duration of fire
exposure. The latter option is typically used to test brick walls because the test termination point is almost always a
temperature rise rather than a failure by passage of hot gases or collapse where there is a degradation of the brick
wythe from the hose stream test. Some other materials rely on the duplicate specimen to meet certain fire ratings.
0
T
e
m
p
e
r
a
t
u
r
e
,

D
e
g
.

F
400
800
1200
1600
2000
2400
T
e
m
p
e
r
a
t
u
r
e
,

D
e
g
.

C
0
200
400
600
800
1000
1200
Figure 1
Time-Temperature Curve for ASTM Standard E119
1 2 3 4 5 6 7 8
Time, hours
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 3 of 16
Loading. Throughout the fire endurance and hose stream tests, a superimposed load is applied to bearing
specimens. The applied load is required to be the maximum load condition allowed by nationally recognized
structural design criteria or by limited design criteria for a reduced load.
Columns are loaded to simulate the maximum load condition allowed by nationally recognized structural design
criteria or by limited design criteria for a reduced load. The column is then subjected to the standard fire on all
sides. Where the fire protection is not designed to carry loads, an alternate test method in which the column is not
loaded may be used.
Conditions of Acceptance
The number of criteria considered as termination points for a fire test on an assembly depends on whether the
assembly is loadbearing or not.
Non-Bearing Walls and Partitions. The test is successful and a fire resistance rating is assigned to the
construction if all of the following criteria are met:
1. The assembly withstands the fire endurance test without passage of flame or gases hot enough to
ignite cotton waste for a period equal to that for which classification is desired.
2. The assembly withstands the fire endurance test without passage of flame and the hose stream test
without passage of water from the hose stream. If an opening develops in the wall specimen that
permits a projection of water beyond the surface of the unexposed side during the hose stream test,
the assembly is considered to have failed the test.
3. The average rise in temperature of nine thermocouples on the unexposed surface is not more
than 250 F (139 C) above their average initial temperature, and the temperature rise of a single
thermocouple is not more than 325 F (181 C) above its initial temperature.
Bearing Walls. The conditions of acceptance for bearing walls are the same as for non-bearing walls and
partitions (above), with the following addition:
4. The specimen must also sustain the applied load during the fire endurance and hose stream tests.
The first three criteria relate to providing a barrier against the spread of fire by penetration of the assembly; the
fourth relates to structural integrity. The termination point for fire tests of brick masonry walls is almost invariably
due to temperature rise (heat transmission) of the unexposed surface. Brick masonry walls successfully withstand
the load during the fire endurance test and the hose stream test conducted immediately after the wall has been
subjected to the fire exposure. This structural integrity of brick masonry walls is attested to in many fires where the
masonry walls have remained standing when other parts of the building have been destroyed or consumed during
the fire.
Columns. Columns with integral structural fireproofing are assigned a fire resistance rating when they support
the superimposed load during the fire endurance test. For columns with fireproofing not designed to carry loads,
a fire resistance rating is assigned when the average temperature rise does not exceed 1000 F (556 C) or the
maximum temperature rise does not exceed 1200 F (667 C) at any one point.
FIRE RESISTANCE RATINGS OF WALLS
There are several sources of fire resistance ratings for brick masonry assemblies that will typically satisfy
the requirements of the local building official. Model building codes contain results based on testing. Private
laboratories report fire test results. Individual associations and companies sponsor fire tests and make the results
available.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 4 of 16
Building Codes
Table 1 presents fire resistance ratings for various masonry wall assemblies, as taken from the 2006 International
Building Code Table 720.1(2) [Ref. 1]. Note that for item numbers 1.1-1 through 1.1-3, 1-2.1, and 2-1.1 through
2-1.2, the required thickness of clay brick masonry is the equivalent thickness, i.e. the thickness is the volume
of clay in a unit divided by the face area. Table 2 presents fire resistance ratings for brick veneer/steel stud wall
assemblies as taken from Table 721.4.1(2) of the same code.
TABLE 1
Fire Resistance Ratings (Periods) for Various Walls and Partitions
Material
Item
Number
Construction
Minimum Finished
Thickness,
Face-to-Face,

in. (mm)
4 hr 3 hr 2 hr 1 hr
1. Brick of clay
or shale
2
1-1.1 Solid brick of clay or shale
1
6.0
(152)
4.9
(124)
3.8
(97)
2.7
(69)
1-1.2 Hollow brick, not filled
5.0
(127)
4.3
(109)
3.4
(86)
2.3
(58)
1-1.3
Hollow brick unit wall, grouted solid or filled with perlite vermiculite
or expanded shale aggregate
6.6
(168)
5.5
(140)
4.4
(112)
3.0
(76)
1-2.1
4 in. (102 mm) nominal thick units at least 75 percent solid backed
with hat-shaped metal furring channel in. (76 mm) thick formed
from 0.021 in. (0.53 mm) sheet metal attached to the brick wall at
24 in. (610 mm) o.c. with approved fasteners, and in. (12.7 mm)
Type X gypsum wallboard attached to the metal furring strips with
1 in. (25.4 mm) long Type S screws spaced at 8 in. (203 mm) o.c.

5
3
(127)

2. Combination
of clay brick
and loadbearing
hollow clay tile
2
2-1.1
4 in. (102 mm) solid brick and 4 in. (102 mm) tile (at least
40 percent solid)

8
(203)

2-1.2
4 in. (102 mm) solid brick and 8 in. (203 mm) tile (at least
40 percent solid)
12
(305)

15. Exterior or
interior walls
4,5,6
15-1.5
7
2 3 in. (57 95 mm) clay face brick with cored holes over
in. (12.7 mm) gypsum sheathing on exterior surface of 2 4 in.
(51 102 mm) wood studs at 16 in. (406 mm) o.c. and two layers
in. (15.9 mm) Type X gypsum wallboard on interior surface.
Sheathing placed horizontally or vertically with vertical joints over
studs nailed 6 in. (152 mm) on center with 1 in. (44 mm) by
No. 11 gage by
7
16 in. (11.1 mm) head galvanized nails. Inner
layer of wallboard placed horizontally or vertically and nailed
8 in. (203 mm) on center with 6d cooler or wallboard nails. Outer
layer of wallboard placed horizontally or vertically and nailed
8 in. (203 mm) on center with 8d cooler or wallboard nails. All
joints staggered with vertical joints over studs. Outer layer joints
taped and finished with compound. Nail heads covered with joint
compound. 0.035 in. (0.89 mm) (No. 20 galvanized sheet gage)
corrugated galvanized steel wall ties 6 in. (19.1 168 mm)
attached to each stud with two 8d cooler or wallboard nails every
sixth course of bricks.

10
(254)

1. For units in which the net cross-sectional area of cored brick in any plane parallel to the surface containing the cores is at least 75 percent of
the gross cross-sectional area measured in the same plane.
2. Thickness shown for brick and clay tile are nominal thicknesses unless plastered, in which case thicknesses are net. Thickness shown
for clay masonry is equivalent thickness defined by Equation 3. Where all cells are solid grouted or filled with silicone-treated perlite loose-fill
insulation; vermiculite loose-fill insulation; or expanded clay, shale or slate lightweight aggregate, the equivalent thickness shall be the thickness
of the brick using specified dimensions. Equivalent thickness may also include the thickness of applied plaster and lath or gypsum wallboard,
where specified.
3. Shall be used for non-bearing purposes only.
4. Staples with equivalent holding power and penetration shall be permitted to be used as alternate fasteners to nails for attachment to wood
framing.
5. For all of the construction with gypsum wallboard described in this table, gypsum base for veneer plaster of the same size, thickness and
core type shall be permitted to be substituted for gypsum wallboard, provided attachment is identical to that specified for the wallboard, and the
joints on the face layer are reinforced and the entire surface is covered with a minimum of
1
16 in. (1.6 mm) gypsum veneer plaster.
6. For properties of cooler or wallboard nails, see ASTM C514, ASTM C547 or ASTM F1667.
7. The design stress of studs shall be reduced to 78 percent of allowable F
c with the maximum not greater than 78 percent of the calculated
stress with studs having a slenderness ratio l/d of 33.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 5 of 16
TABLE 2
Fire Resistance Ratings for Brick Veneer/Steel Stud Assemblies
Wall or Partition Assembly
Plaster
Side
Exposed
(hours)
Brick
Faced Side
Exposed
(hours)
Outside facing of steel studs:
in. (12.7 mm) wood fiberboard sheathing next to studs, in. (19.1 mm) air space formed
with 1 in. (19.1 41 mm) wood strips placed over the fiberboard and secured to the
studs; metal or wire lath nailed to such strips, 3 in. (95 mm) brick veneer held in place by
filling in. (19.1 mm) air space between the brick and lath with mortar.
Inside facing of studs:
in. (19.1 mm) unsanded gypsum plaster on metal or wire lath attached to
5
16 in. (7.9 mm)
wood strips secured to edges of the studs.
1.5 4
Outside facing of steel studs:
1 in. (25.4 mm) insulation board sheathing attached to studs, 1 in. (25.4 mm) air space, and
3 in. (95 mm) brick veneer attached to steel frame with metal ties every fifth course.
Inside facing of studs:
in. (22.2 mm) sanded gypsum plaster (1:2 mix) applied on metal or wire lath attached
directly to the studs.
1.5 4
Same as above except use in. (22.2 mm) vermiculite gypsum plaster or 1 in. (25.4 mm)
sanded gypsum plaster (1:2 mix) applied to metal or wire.
2 4
Outside facing of steel studs:
in. (12.7 mm) gypsum sheathing board, attached to studs, and 3 in. (95 mm) brick veneer
attached to steel frame with metal ties every fifth course.
Inside facing of studs:
in. (12.7 mm) sanded gypsum plaster (1:2 mix) applied to in. (12.7 mm) perforated
gypsum lath securely attached to studs and having strips of metal lath 3 in. (76 mm) wide
applied to all horizontal joints of gypsum lath.
2 4
UL Listings
Underwriters Laboratories is a resource recognized throughout the building industry that has thousands of
published fire resistance rated designs and product certifications that appear in the UL Fire Resistance Directory
[Ref. 7] and are typically accepted without modification by building officials. The UL certification is based on an
assembly complying with the ASTM E119 test, as described previously. The directory lists several masonry wall
assemblies with various potential alternates in materials as shown in Table 3.
TABLE 3
UL Fire Resistance Ratings for Brick Masonry Walls
Design
Number
Rating
1
Assembly
Brick Veneer/Wood Stud, Loadbearing
U302 2 hr
(2) layers in. (15.9 mm) thick gypsum wallboard or nominal
3
32 in. (2.4 mm) thick gypsum
veneer plaster on Classified veneer baseboard
(1) layer in. (12.7 mm) thick exterior gypsum sheathing
1 (25.4 mm) in. (51 102 mm) air space
nominal 2 4 in. wood studs spaced at 16 in. (406 mm) o.c.
nominal 4 in. (102 mm) clay facing brick laid in mortar with in. (19.1 mm) wide 6 in. 168
mm) long 20 MSG corrugated wall ties spaced at 16 in. (406 mm) o.c. each way
1. Unless noted otherwise, fire resistance rating applies to both sides of assembly.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 6 of 16
Design
Number
Rating
1
Assembly
Brick Veneer/Wood Stud, Loadbearing (continued)
U356 1 hr
(1) layer in. (15.9 mm) thick gypsum board
nominal 2 4 in. (51 102 mm) wood studs spaced at 16 in. (406 mm) o.c., with 3 in. (89 mm)
thick glass fiber batt or spray applied cellulose insulation

7
16 in. (11.1 mm) min. thick wood structural panels or min. in. (12.7 mm) thick mineral and fiber
boards
1 in. (25.4 mm) air space
nominal 4 in. (102 mm) brick veneer with corrugated metal wall ties spaced not more than each
sixth course of brick and max. 32 in. (813 mm) o.c. horizontally
U371 1 hr
(2) layer in. (15.9 mm) thick gypsum board
nominal 2 4 in. (51 102 mm) wood studs spaced at 16 in. (406 mm) o.c. with min. 3 in.
(76 mm) mineral wool batt insulation
(1) layer in. (15.9 mm) thick gypsum board
1 in. (25.4 mm) air space
nominal 4 in. (102 mm) brick veneer with corrugated metal wall ties attached with screws and
spaced not more than each fourth course and a max. 24 in. (610 mm) o.c. horizontally
Brick Veneer/Steel Stud, Loadbearing
U418
45 min
1 hr
2 hr
(45 min): (1) layer in. (15.9 mm) thick gypsum wallboard
(1 hr): (2) layers in. (12.7 mm) thick gypsum wallboard
(2 hr): (3) layers in. (12.7 mm) thick gypsum wallboard
3 or 5 in., (89 or 140 mm) 18 gage, steel studs, spaced at 24 in. (610 mm) o.c., with 3 in.
(89 mm) thick glass fiber batt insulation
(1) layer in. (12.7 mm) thick exterior gypsum sheathing
1 in. (25.4 mm) air space
4 in. (102 mm) nominal clay facing brick laid in mortar with metal ties at 24 in. (610 mm) o.c.
horizontally and 16 in. (406 mm) o.c. vertically
U424
45 min
1 hr
1 hr
2 hr
(45 min): (1) layer in. (15.9 mm) thick gypsum wallboard
(1 hr): (2) layers in. (12.7 mm) thick gypsum wallboard
(1 hr): (2) layers in. (15.9 mm) thick gypsum wallboard
(2 hr): (3) layers in. (12.7 mm) or (2) layers in. (19.1 mm) thick gypsum wallboard
3 in. (89 mm), 20 gage steel studs, spaced up to 24 in. (610 mm) o.c., with 3 in. (89 mm) thick
glass fiber or mineral wool batt or blanket insulation
(1) layer or in. (12.7 or 15.9 mm) thick exterior gypsum sheathing
Air space thickness not specified
3 in. (95 mm) min. brick veneer with corrugated metal wall ties attached to each stud with steel
screws, not more than each sixth course of brick
U425
45 min
1 hr
1 hr
2 hr
(45 min): (1) layer in. (15.9 mm) thick gypsum wallboard
(1 hr): (2) layers in. (12.7 mm) thick gypsum wallboard
(1 hr): (2) layers in. (15.9 mm) thick gypsum wallboard
(2 hr): (3) layers in. (12.7 mm) or (2) layers in. (19.1 mm) thick gypsum wallboard
3 in. (89 mm), 20 gage steel studs, spaced up to 24 in. (610 mm) o.c., with 3 in. (89 mm) thick
glass fiber or mineral wool batt or blanket insulation
(1) layer or in. (12.7 or 15.9 mm) thick exterior gypsum sheathing
Air space thickness not specified
3 in. (95 mm) brick veneer with corrugated metal wall ties attached to each stud with steel
screws, not more than each sixth course of brick
TABLE 3 (continued)
UL Fire Resistance Ratings for Brick Masonry Walls
1. Unless noted otherwise, fire resistance rating applies to both sides of assembly.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 7 of 16
TABLE 3 (continued)
UL Fire Resistance Ratings for Brick Masonry Walls
1. Unless noted otherwise, fire resistance rating applies to both sides of assembly.
Design
Number
Rating
1
Assembly
Brick Veneer/Steel Stud, Loadbearing (continued)
V434 1 hr
(1) layer in. (15.9 mm) thick gypsum wallboard
3 in. (89 mm), 20 gage, steel studs with max. spacing at 24 in. (610 mm) o.c., with 3 in.
(89 mm) thick glass fiber batt insulation
2 in. (51 mm) max. thick foamed plastic
1 in. (25.4 mm) min. air space
4 in. (102 mm) nominal brick veneer with wall anchor ties attached to studs at max. 24 in.
(610 mm) o.c.
V454 1 hr
(1) layer in. (15.9 mm) thick gypsum wallboard
3 in. (89 mm), 20 gage, steel studs at max. spacing of 24 in. (610 mm) o.c.
(1) layer in. (15.9 mm) thick gypsum wallboard
4 in. (102 mm) max. thick rigid polystyrene insulation
1 in. (25.4 mm) min. air space
4 in. (102 mm) nominal brick veneer with wall anchor ties attached to studs at max. 24 in.
(610 mm) o.c.
V458 45 min
(1) layer in. (15.9 mm) thick gypsum wallboard bearing UL Classification Mark
3 in. (92 mm) 18 gage steel studs at max. spacing of 24 in. (610 mm) o.c. with nominal 3.5 pcf
mineral wool batt
(1) layer in. (15.9 mm) thick gypsum wallboard
1 in. (25.4 mm) min. air space
3 in. (95 mm) min. thick brick veneer with corrugated metal wall ties attached to each stud with
steel screws, not more than each sixth course of brick
Brick Veneer/Steel Stud, Non-Loadbearing
V414
3 hr,
interior
1 hr,
exterior
(1) layer in. (15.9 mm) thick gypsum wallboard
3 in. (92 mm) wide, 1 in. (41 mm) legs, 20 gage steel studs, spaced 16 in. (406 mm) o.c.,
studs cut in. (19.1 mm) less than assembly height
2 in. (51 mm) thick foamed plastic (rigid insulation)
2 in. (51 mm) air space
4 in. (102 mm) nominal clay facing brick laid in mortar with metal ties at 16 in. (406 mm) o.c. max.
each way
Brick/Concrete Masonry, Loadbearing
U902 4 hr
4 in. (102 mm) nominal loadbearing concrete masonry unit laid with full mortar beds and with
9 gage joint reinforcement at 16 in. (406 mm) o.c. vertically
min. 1 in. (25.4 mm) air space with up to 4 in. (102 mm) foamed plastic (rigid insulation) as option
in. (19.1 mm) wide, 7 in. (178 mm) long, 26 gage corrugated metal ties spaced at 8 in.
(203 mm) o.c. horizontally and 16 in. (406 mm) o.c. vertically or truss or ladder type joint
reinforcement of 9 gage wire for full width of wall assembly, cross wires at 16 in. (406 mm) o.c.,
spaced at 16 in. (406 mm) o.c. vertically
4 in. (102 mm) nominal clay facing brick laid in mortar
Other
In addition to assemblies listed above, there are several other assemblies previously tested with results published
in past building codes or other publications. A selection of these appears in Table 4.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 8 of 16
CALCULATED FIRE RESISTANCE
Theory and Derivation
The extent of fire resistance provided by a clay masonry wall is a function of the walls mass or thickness. This
well-established fact is based on the results of many fire resistance tests conducted on walls of solid and hollow
clay units. During the ASTM E119 fire test, the fire resistance period of clay masonry walls is usually established
by the temperature rise on the unexposed side of the wall specimen. Few masonry walls have failed due to
loading or thermal shock of the hose stream.
The method for calculating a fire resistance period is described in NBS BMS 92, Fire-Resistance Classifications
of Building Construction [Ref. 6]. The construction must be similar to others for which the fire resistance periods
are known or of composite construction for which the fire resistance period of each component is known. The
calculated fire resistance formulas are based on the temperature rise on the unexposed side of the wall.
Heat transmission theory states that when a wall made of a given material is exposed to a heat source that maintains
a constant temperature at the surface of the exposed side and the unexposed side is protected against heat loss, the
unexposed side will attain a given temperature rise inversely proportional to the square of the walls thickness.
In the standard fire test, the time required to attain a given temperature rise on the unexposed side will be different
than when the temperature on the exposed side remains constant. This is because the fire in the standard fire
test increases the temperature at the exposed surface of the wall as the test proceeds. Based on fire test data
collected from many fire tests, the following formula has been derived to express the fire resistance period of a
wall based on its thickness:
1. As tested by the Southwest Research Institute [Ref. 4].
2. Fire resistance rating applies to brick (exterior) side only. Test stopped at 1 hour.
3. Width not in compliance with 2006 IBC veneer requirements; however, complies with 2006 IRC [Ref. 2] veneer requirements.
Test Rating
2
Assembly
Brick Veneer/Wood Stud
1 1 hr
(1) layer in. (12.7 mm) thick gypsum wallboard
2 4 in. (51 102 mm) wood studs spaced at 16 in. (406 mm) o.c. with 3 in. (89 mm) glass fiber batt
insulation between studs
(1) layer in. (12.7 mm) thick wood fiberboard sheathing
(1) layer No. 15 asphalt felt paper
1 in. (25.4 mm) air space
3 in. (89 mm) actual width hollow clay brick with void area of 34.5% (equivalent thickness of 2.3
in. (58 mm)), laid in mortar with in. (22.2 mm) wide, 22 gage corrugated wall ties spaced at 24 in.
(610 mm) o.c. horizontally and 16 in. (406 mm) o.c. vertically
2 1 hr
(1) layer in. (12.7 mm) thick gypsum wallboard
2 4 in. (51 102 mm) wood studs spaced at 16 in. (406 mm) o.c. with 3 in. (89 mm) glass fiber batt
insulation between studs
(1) layer in. (12.7 mm) thick wood fiberboard sheathing
(1) layer No. 15 asphalt felt paper
1 in. (25.4 mm) air space
2 in. (73 mm) actual width hollow clay brick with void area of 36% (equivalent thickness of 1.8 in.
(46 mm)), laid in mortar with in. (22.2 mm) wide, 22 gage corrugated wall ties spaced at 24 in.
(610 mm) o.c. horizontally and 16 in. (406 mm) o.c. vertically
3 1 hr
(1) layer in. (12.7 mm) thick gypsum wallboard
2 4 in. (51 102 mm) wood studs spaced at 16 in. (406 mm) o.c. with 3 in. (89 mm) glass fiber batt
insulation between studs
(1) layer in. (12.7 mm) thick wood fiberboard sheathing
(1) layer No. 15 asphalt felt paper
1 in. (25.4 mm) air space
1 in. (44 mm) actual width
3
hollow clay brick with void area of 26.9% (equivalent thickness of 1.3
in. (32 mm)), laid in mortar with in. (22.2 mm) wide, 22 gage corrugated wall ties spaced at 24 in.
(610 mm) o.c. horizontally and 16 in. (406 mm) o.c. vertically
TABLE 4
Fire Resistance Ratings for Other Brick Masonry Wall Assemblies
1
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 9 of 16
R = (cV)
n
Eq. 1
where:
R = fire resistance period, hr
c = coefficient depending on the material, design of the wall, and the units of measurement of R and V
V = volume of solid material per unit area of wall surface, and
n = exponent depending on the rate of increase of temperature at the exposed face of the wall
For walls of a given material and design, an increase of 50 percent in volume of solid material per unit area of
wall surface results in a 100 percent increase in the fire resistance period. This relationship results in a value of
1.7 for n. The lower value for n compared with 2 for the theoretical condition should be anticipated since a rising
temperature at the exposed surface will shorten the fire resistance period of a wall.
For a wall composed of layers of multiple materials, the fire resistance period may be expressed as follows:
R = (c1V1 + c2V2 + c3V3)
n
= (R1
1/n
+ R2
1/n
+ R3
1/n
)
n
Substituting 1.7 for n and 0.59 for 1/n, the general formula for calculating a fire resistance period becomes:
R = (R1
0.59
+ R2
0.59
+ R3
0.59
+ Ri
0.59
)
1.7
Eq. 2
where:
R1, R2, R3, Ri = known fire resistance period of each component layer, hr
Where available, the fire resistance period (the full duration of the fire test before a termination point is reached)
should be used. Where this period is not available (many brick wall tests are stopped after the desired rating time
period elapses), the fire resistance rating (typically truncated to be the highest full hour of fire test duration) can be
used. However, using the fire resistance rating for a component layer will generally result in a lower calculated fire
resistance period for the overall assembly than using the fire resistance period for each component layer.
The calculated fire resistance, calculated using either the fire resistance period or fire resistance rating of each layer, can
then be used to verify that the wall assembly equals or exceeds the fire resistance rating required by the building code.
The theory proposed and derived in NBS BMS 92 has been incorporated into Code Requirements for Determining
Fire Resistance of Concrete and Masonry Assemblies (ACI 216.1/TMS-0216) [Ref. 5].
Calculations
The 2006 International Building Code (IBC) [Ref. 1] permits the fire resistance of masonry assemblies to be
calculated in accordance with TMS-0216. In addition, the IBC also includes methods for calculating the fire
resistance of a masonry assembly that are based on and very similar to those in TMS-0216. The methods
discussed below are taken from TMS-0216 unless noted otherwise.
Equivalent Thickness of a Single Wythe. The average thickness of the solid material (i.e., minus cores or cells)
in a masonry unit as placed in the wall is the equivalent thickness of the masonry unit. This is determined by
measuring the total volume of the masonry unit, subtracting the volume of the core or cell spaces and dividing by
the area of the exposed face of the masonry unit, which is expressed as follows:
Te=Vn/LH Eq. 3
where:
Te = equivalent thickness of the masonry unit, in.
Vn = net volume of the masonry unit, in.
3
L = specified length of the masonry unit, in.
H = specified height of the masonry unit, in.
Equation 3 can be simplified as follows:
Te = [WLH (1 Pv)] / LH
= (1 Pv) W Eq. 4
= Ps W Eq. 5
where:
W = specified width of the masonry unit, in.
Pv = percent void of the masonry unit
Ps = percent solid of the masonry unit
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 10 of 16
Multiple Wythe Walls. For walls with multiple wythes of brick, concrete masonry or concrete, the calculated fire
resistance formula is:
R = (R1
0.59
+ R2
0.59
+ Rn
0.59
+ A1 + A2 + An)
1.7
Eq. 6
where:
R = calculated fire resistance period of the assembly, hr
R1, R2 Rn = fire resistance periods of the individual wythes, hr
A1, A2 An = 0.30; the air factor for each continuous air space having a distance of to 3 in. (12.7 to
89 mm) between wythes
The fire resistance period used in Equation 6 for each individual wythe or layer is determined from Table 5 for a
wythe made of clay units, from Table 6 for a wythe made of concrete masonry units and from Table 7 and Figure 2
for a concrete layer.
TABLE 6
Fire Resistance Periods of Concrete Masonry Walls
Aggregate Type Minimum Equivalent Thickness for Fire Resistance Rating in. (mm)
1,2
hr hr 1 hr 1 hr 2 hr 3 hr 4 hr
Calcareous or siliceous gravel (other than
limestone)
2.0 (51) 2.4 (61) 2.8 (71) 3.6 (91) 4.2 (107) 5.3 (135) 6.2 (157)
Limestone, cinders, or air-cooled slag 1.9 (48) 2.3 (58) 2.7 (69) 3.4 (86) 4.0 (102) 5.0 (127) 5.9 (150)
Expanded clay, expanded shale or
expanded slate
1.8 (46) 2.2 (56) 2.6 (66) 3.3 (84) 3.6 (91) 4.4 (112) 5.1 (130)
Expanded slag or pumice 1.5 (38) 1.9 (48) 2.1 (53) 2.7 (69) 3.2 (81) 4.0 (102) 4.7 (119)
1. Fire resistance periods between the hourly fire resistance rating listed shall be determined by linear interpolation based on the equivalent
thickness value of the concrete masonry assembly.
2. Minimum required equivalent thickness corresponding to the fire resistance rating for units made with a combination of aggregates shall be
determined by linear interpolation based on the percent by dry-rodded volume of each aggregate used in manufacturing the units.
For ungrouted and partially grouted construction, the equivalent thickness should be determined according to Equation
3. The equivalent thickness should be taken as the actual thickness of the masonry unit for solid grouted construction
or brickwork constructed of hollow brick units complying with ASTM C652 and filled with one of the following:
Sand, pea gravel, crushed stone or slag complying with ASTM C33
Pumice, scoria, expanded shale, clay, slate, slag or fly ash; or cinders complying with ASTM C331
Perlite complying with ASTM C549
Vermiculite complying with ASTM C516
Fire Resistance of a Single Wythe. The minimum equivalent thickness required to achieve a given fire resistance
rating with a clay masonry wythe is listed in Table 5. The table is organized by material type and hourly fire
resistance ratings. For fire resistance periods that are between the hourly increments listed in the table, the
minimum equivalent thickness may be determined by linear interpolation. Where combustible members such as
wood floor joists are framed into the wall, the thickness of solid material between the end of each member and the
opposite face of the wall, or between members set in from opposite sides is allowed to be no less than 93 percent
of the thickness shown in Table 5.
Material Type
Minimum Equivalent Thickness for Fire
Resistance, in. (mm)
1,2,3

1 hr 2 hr 3 hr 4 hr
Solid brick of clay or shale
4
2.7 (69) 3.8 (97) 4.9 (124) 6.0 (152)
Hollow brick or tile of clay or shale, unfilled 2.3 (58) 3.4 (86) 4.3 (109) 5.0 (127)
Hollow brick or tile of clay or shale, grouted or filled with materials specified 3.0 (76) 4.4 (112) 5.5 (140) 6.6 (168)
1. Equivalent thickness as determined from Equations 3, 4 or 5.
2. Calculated fire resistance between the hourly increments listed shall be determined by linear interpolation.
3. Where combustible members are framed into the wall, the thickness of solid material between the end of each member and the opposite
face of the wall, or between members set in from opposite sides, shall not be less than 93% percent of the thickness shown.
4. Units in which the net cross-sectional area of cored or deep frogged brick in any plane parallel to the surface containing the cores or deep
frogs is at least 75 percent of the gross cross-sectional area measured in the same plane.
TABLE 5
Fire Resistance Ratings of Clay Masonry Walls
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 11 of 16
Finish Materials. When drywall, stucco or plaster finishes are applied to a masonry wall, the fire resistance of
the wall is increased. Where finish materials are used to attain a required fire resistance rating, the fire resistance
provided by the masonry alone must be a minimum of half the required fire resistance rating to ensure the
structural integrity of the wall.
For finishes applied to the non-fire exposed side of a wall, the finish is converted to an equivalent thickness of
brickwork. This adjusted thickness is then calculated by multiplying the thickness of the finish by the applicable
factor from Table 8 established by the durability of the finish and the wall material. The adjusted finish thickness is
then added to the base equivalent thickness of the wall used in Table 5.
Type of Material Used in Slab or Wall Type of Finish Applied to Slab or Wall
Portland
Cement-Sand
Plaster
1
or
Terrazzo
Gypsum-Sand
Plaster
Gypsum-
Vermiculite or
Perlite Plaster
Gypsum
Wallboard
Clay masonry solid brick of clay or shale 1.00 1.25 1.75 3.00
Clay masonry hollow brick or tile of clay or shale 0.75 1.00 1.50 2.25
Concrete masonry siliceous, calcareous, lime-
stone, cinders, sir-cooled blast-furnace slag
1.00 1.25 1.75 3.00
Concrete masonry made with 80% of more by
volume of expanded shale, slate or clay, expanded
slag, or pumice
0.75 1.00 1.25 2.25
Concrete siliceous, carbonate, air-cooled blast-
furnace slag
1.00 1.25 1.75 3.00
Concrete semi-lightweight 0.75 1.00 1.50 2.25
Concrete lightweight, insulating concrete 0.75 1.00 1.25 2.25
1. For portland cement-sand plaster in. (15.9 mm) or less in thickness and applied directly to clay masonry on the non-fire exposed side of
the wall, the multiplying factor shall be 1.0.
TABLE 8
Multiplying Factor for Finishes on Non-Fire Exposed Side of Masonry and Concrete Walls
0
2 3 4 5 6 7
Panel Thickness, inches
1
2
3
4
5
Insulating Concrete
35 pcf (560 kg/m )
3
F
i
r
e

R
e
s
i
s
t
a
n
c
e

P
e
r
i
o
d

(
R
)
,

h
o
u
r
s
n
Panel Thickness, mm
50 75 100 125 150 175
Lightweight
100 pcf
(1600 kg/m )
3
Semi-Lightweight
115 pcf
(1850 kg/m )
3
Air-Cooled Blast
Furnace Slag
Carbonate
Aggregate
Silicaceous
Aggregate
Figure 2
Fire Resistance Periods for Other Concrete Panels
Aggregate
Type
Minimum Equivalent Thickness
for Fire Resistance Rating,
in. (mm)
1 hr 1 hr 2 hr 3 hr 4 hr
Siliceous
3.5
(89)
4.3
(109)
5.0
(127)
6.2
(157)
7.0
(178)
Carbonate
3.2
(81)
4.0
(102)
4.6
(117)
5.7
(145)
6.6
(168)
Semi-
lightweight
2.7
(69)
3.3
(84)
3.8
(97)
4.6
(117)
5.4
137)
Lightweight
2.5
(64)
3.1
(79)
3.6
(91)
4.4
(112)
5.1
(130)
TABLE 7
Fire Resistance Periods of
Normal-Weight Concrete Panels
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 12 of 16
Examples
Example 1, Cavity Wall with Air Space. A multi-wythe cavity wall consists of a wythe of 4 in. (102 mm) nominal
solid brick units complying with ASTM C216 and cored at 25 percent, a 2 in. (51 mm) air space and a wythe of 8
in. (203 mm) nominal concrete masonry unit made of calcareous gravel. The concrete masonry unit has actual
dimensions of 7 7 15 inches (194 194 397 mm) and is 53 percent solid. The fire resistance rating is
determined by Equation 6 as follows:
a. From Equation 4, the equivalent thickness of the solid brick is:
Te = (1 0.25) 3.625 in. (92 mm) = 2.71 in. (69 mm)
b. From Table 5, the fire resistance period for the clay unit is:
R1 = 1.0 hr
c. For a 2 in. (51 mm) air space:
A = 0.30
d. From Equation 5, the equivalent thickness of the concrete masonry unit is:
Te
= 0.53 7.625 in. (194 mm) = 4.0 in. (102 mm)
e. Interpolating from Table 7, the fire resistance period of the concrete masonry is:
R
2 = 1.5 hr + 0.5 hr [(4.0 3.6) / (4.2 3.6)] = 1.5 hr + 0.5 hr (0.67) = 1.8 hr
f. From Equation 6, the fire resistance rating of the entire wall assembly is:
R = [(1.0)
0.59
+ (1.8)
0.59
+ 0.3]
1.7
= 5.5 hr Fire Resistance Rating = 4 hr
For finishes on the fire exposed side of the wall, a time is assigned to the finish according to Table 9, which is the
length of time the finish will contribute toward the fire resistance rating of the fire exposed side of the wall. This
time is added to the fire resistance rating determined for the base wall and non-fire exposed finish.
TABLE 9
Time Assigned to Finish Materials on Fire Exposed Side of Wall
Finish Thickness Time (minutes)
Gypsum wallboard
in. (9.5 mm) 10
in. (12.7 mm) 15
in. (15.9 mm) 20
Two layers of in. (9.5 mm) 25
One layer of in. (9.5 mm) and one layer of
in. (12.7 mm)
35
Two layers of in. (12.7 mm) 40
Type X gypsum wallboard
in. (12.7 mm) 25
in. (15.9 mm) 40
Direct-applied portland cement-sand plaster See Note 1
Portland cement-sand plaster on metal lath
in. (19.1 mm) 20
in. (22.2 mm) 25
1 in. (25.4 mm) 30
Gypsum-sand plaster on in. (9.5 mm) gypsum lath
in. (12.7 mm) 35
in. (15.9 mm) 40
in. (19.1 mm) 50
Gypsum-sand plaster on metal lath
in. (19.1 mm) 50
in. (22.2 mm) 60
1 in. (25.4 mm) 80
1. For purposes of determining the contribution of portland cement-sand plaster to the equivalent thickness of concrete or masonry for use in
Tables 5, 6 or 7, it shall be permitted to use the actual thickness of the plaster or in. (15.9 mm), whichever is smaller.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 13 of 16
Example 2, Composite Wall. A multi-wythe composite wall consists of 4 in. (102 mm) nominal hollow brick
complying with ASTM C652 with a mortared collar joint and 4 in. (102 mm) siliceous aggregate concrete wall. The
gross volume of the hollow brick includes 36 percent void. The fire resistance rating is determined as follows:
a. From Equation 4, the equivalent thickness of the hollow brick is:
Te = (1 0.36) 3.625 in. (92 mm) = 2.3 in. (58 mm)
b. From Table 5, the fire resistance period for the hollow brick is:
R1 = 1.0 hr
c. From Figure 2, the fire resistance period for the concrete wall with siliceous aggregate is:
R2 = 1.3 hr
d. Using Equation 6, the fire resistance rating of the wall assembly is:
R = [(1.0)
0.59
+ (1.3)
0.59
]
1.7
= 3.7 hr Fire Resistance Rating = 3 hr
Example 3, Composite Wall with Gypsum Wallboard. Determine the fire resistance rating for the composite
wall of Example 2 when in. (12.7 mm) thick gypsum wallboard is applied to the interior side of the wall. The fire
resistance rating will apply to only the exterior side of the wall.
a. Using Table 8, the multiplying factor for a siliceous concrete wall and gypsum wallboard is 3.00. The
equivalent thickness of concrete for the gypsum wallboard on the unexposed side is:
T
e (finish) = 3.00 0.5 (12.7 mm) in. = 1.5 in. (38 mm)
b. The equivalent thickness of concrete for the gypsum wallboard finish and concrete is:
Te (finish) + Te (concrete) = 1.5 in. (38 mm) + 4 in. (102 mm) = 5.5 in. (140 mm)
c. From Figure 2, the fire resistance period for a 5.5 in. (140 mm) thick concrete wall of siliceous aggregate is:
R2 = 2.4 hr
d. Using Equation 6, the fire resistance rating of the wall assembly is:
R = [(1.00)
0.59
+ (2.4)
0.59
]
1.7
= 5.3 hr Fire Resistance Rating = 4 hr
Example 4, Composite Wall with Gypsum Wallboard. Determine the fire resistance rating for the composite wall
of Example 3 when the fire resistance rating will apply to both sides of the wall. Since the fire resistance rating will
be applied to both sides, a calculation for fire exposure on each side of the wall must be performed.
Gypsum Board Side (Interior) Exposed to Fire
a. Using Table 9, the contribution of the in. (12.7 mm) thick gypsum wallboard to the fire resistance is:
Rf = 15 min / 60 min/hr = 0.25 hr
b. Using the fire resistance period determined in Example 2, the fire resistance period for the interior side of
the wall assembly is:
R (interior) = 3.7 hr + 0.25 hr = 3.9 hr
Brick Side (Exterior) Exposed to Fire
c. The fire resistance period determined in Example 3 for exposing the exterior side of the wall assembly to
fire is:
R (exterior) = 5.3 hr
Wall Assembly
d. The fire resistance rating for the wall assembly is the lower of the fire resistance periods calculated from
the interior and the exterior:
3.9 hr < 5.3 hr Fire Resistance Rating = 3 hr
DESIGN AND DETAILING
Support of Brick Masonry Rated for Fire Resistance
Walls with a fire resistance rating should be supported by assemblies with a similar or better fire resistance rating.
This prevents collapse of a rated assembly by an unrated support that burns through long before the required
fire resistance period. Thus for a second-story wall with a 2-hour fire resistance rating, the floor-ceiling assembly
providing support for the wall must also have a 2-hour fire resistance rating.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 14 of 16
Where steel columns provide the support, clay masonry can be used as fireproofing of those columns. Table 10,
taken from TMS 0216, gives the equivalent thickness required to provide various levels of column protection,
based on the steel shape used for the column.
W Shapes
Column Size
Clay Masonry
Density, lb/ft
3
(kg/m
3
)
Minimum equivalent thickness for fire-resistance rating of clay masonry
protection assembly, in. (mm)
1 hour 2 hours 3 hours 4 hours
W14 82
120 (1926) 1.23 (31) 2.42 (61) 3.41 (87) 4.29 (109)
130 (2087) 1.40 (36) 2.70 (69) 3.78 (96) 4.74 (120)
W14 68
120 (1926) 1.34 (34) 2.54 (65) 3.54 (90) 4.43 (113)
130 (2087) 1.51 (38) 2.82 (72) 3.91 (99) 4.87 (124)
W14 53
120 (1926) 1.43 (36) 2.65 (67) 3.65 (93) 4.54 (115)
130 (2087) 1.61 (41) 2.93 (74) 4.02 (102) 4.98 (126)
W14 43
120 (1926) 1.54 (39) 2.76 (70) 3.77 (96) 4.66 (118)
130 (2087) 1.72 (44) 3.04 (77) 4.13 (105) 5.09 (129)
W12 72
120 (1926) 1.32 (34) 2.52 (64) 3.51 (89) 4.40 (112)
130 (2087) 1.50 (38) 2.80 (71) 3.88 (99) 4.84 (123)
W12 58
120 (1926) 1.40 (36) 2.61 (66) 3.61 (92) 4.50 (114)
130 (2087) 1.57 (40) 2.89 (73) 3.98 (101) 4.94 (125)
W12 50
120 (1926) 1.43 (36) 2.65 (67) 3.66 (93) 4.55 (116)
130 (2087) 1.61 (41) 2.93 (74) 4.02 (102) 4.99 (127)
W12 40
120 (1926) 1.54 (39) 2.77 (70) 3.78 (96) 4.67 (119)
130 (2087) 1.72 (44) 3.05 (77) 4.14 (105) 5.10 (130)
W10 68
120 (1926) 1.27 (32) 2.46 (62) 3.46 (88) 4.35 (110)
130 (2087) 1.44 (37) 2.75 (70) 3.83 (97) 4.80 (122)
W10 54
120 (1926) 1.40 (36) 2.61 (66) 3.62 (92) 4.51 (115)
130 (2087) 1.58 (40) 2.89 (73) 3.98 (101) 4.95 (126)
W10 45
120 (1926) 1.44 (37) 2.66 (68) 3.67 (93) 4.57 (116)
130 (2087) 1.62 (41) 2.95 (75) 4.04 (103) 5.01 (127)
W10 33
120 (1926) 1.59 (40) 2.82 (72) 3.84 (98) 4.73 (120)
130 (2087) 1.77 (45) 3.10 (79) 4.20 (107) 5.13 (130)
W8 40
120 (1926) 1.47 (37) 2.70 (69) 3.71 (94) 4.61 (117)
130 (2087) 1.65 (42) 2.98 (76) 4.08 (104) 5.04 (128)
W8 31
120 (1926) 1.59 (40) 2.82 (72) 3.84 (98) 4.73 (120)
130 (2087) 1.77 (45) 3.10 (79) 4.20 (107) 5.17 (131)
W8 24
120 (1926) 1.66 (42) 2.90 (74) 3.92 (100) 4.82 (122)
130 (2087) 1.84 (47) 3.18 (81) 4.82 (122) 5.25 (133)
W8 18
120 (1926) 1.75 (44) 3.00 (76) 4.01 (102) 4.91 (125)
130 (2087) 1.93 (49) 3.27 (83) 4.37 (111) 5.34 (136)
TABLE 10
Fire Resistance of Clay-Masonry-Protected Steel Columns
1
1. Tabulated values assume a 1 in. (25.4 mm) air gap between masonry and steel section.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 15 of 16
Square structural tubing
Nominal tube
size, in. (mm)
Clay masonry
density, lb/ft
3

(kg/m
3
)
Minimum equivalent thickness for fire-resistance rating of clay masonry
protection assembly, in. (mm)
1 hour 2 hours 3 hours 4 hours
4 4
(102 102 12.7)
120 (1926) 1.44 (37) 2.72 (69) 3.76 (96) 4.68 (119)
130 (2087) 1.62 (41) 3.00 (76) 4.12 (105) 5.11 (130)
4 4
(102 102 9.5)
120 (1926) 1.56 (40) 2.84 (72) 3.88 (99) 4.78 (121)
130 (2087) 1.74 (44) 3.12 (79) 4.23 (107) 5.21 (132)
4 4
(102 102 6.4)
120 (1926) 1.72 (44) 2.99 (76) 4.02 (102) 4.92 (125)
130 (2087) 1.89 (48) 3.26 (83) 4.37 (111) 5.34 (136)
6 6
(152 152 12.7)
120 (1926) 1.33 (34) 2.58 (66) 3.62 (92) 4.52 (115)
130 (2087) 1.50 (38) 2.86 (73) 3.98 (101) 4.96 (126)
6 6
(152 152 9.5)
120 (1926) 1.48 (38) 2.74 (70) 3.76 (96) 4.67 (119)
130 (2087) 1.65 (42) 3.01 (76) 4.13 (105) 5.10 (130)
6 6
(152 152 6.4)
120 (1926) 1.66 (42) 2.91 (74) 3.94 (100) 4.84 (123)
130 (2087) 1.83 (46) 3.19 (81) 4.30 (109) 5.27 (134)
8 8
(203 203 12.7)
120 (1926) 1.27 (32) 2.50 (64) 3.52 (89) 4.42 (112)
130 (2087) 1.44 (37) 2.78 (71) 3.89 (99) 4.86 (123)
8 8
(203 203 9.5)
120 (1926) 1.43 (36) 2.67 (68) 3.69 (94) 4.59 (117)
130 (2087) 1.60 (41) 2.95 (75) 4.05 (103) 5.02 (128)
8 8
(203 203 6.4)
120 (1926) 1.62 (41) 2.87 (73) 3.89 (99) 4.78 (121)
130 (2087) 1.79 (45) 3.14 (80) 4.24 (108) 5.21 (132)
Steel pipe
Column size,
diameter
thickness,
in. (mm)
Clay masonry
density, lb/ft
3

(kg/m
3
)
Minimum equivalent thickness for fire-resistance rating of clay masonry
protection assembly, in. (mm)
1 hour 2 hours 3 hours 4 hours
4 0.674
(102 17.1)
120 (1926) 1.26 (32) 2.55 (65) 3.60 (91) 4.52 (115)
130 (2087) 1.42 (36) 2.82 (72) 3.96 (101) 4.95 (126)
4 0.337
(102 8.6)
120 (1926) 1.60 (41) 2.89 (73) 3.92 (100) 4.83 (123)
130 (2087) 1.77 (45) 3.16 (80) 4.28 (109) 5.25 (133)
4 0.237
(102 6.0)
120 (1926) 1.74 (44) 3.02 (77) 4.05 (103) 4.95 (126)
130 (2087) 1.92 (49) 3.29 (84) 4.40 (112) 5.37 (136)
5 0.750
(127 19.1)
120 (1926) 1.17 (30) 2.44 (62) 3.48 (88) 4.40 (112)
130 (2087) 1.33 (34) 2.72 (69) 3.84 (98) 4.83 (123)
5 0.375
(127 9.5)
120 (1926) 1.55 (39) 2.82 (72) 3.85 (98) 4.76 (121)
130 (2087) 1.72 (44) 3.09 (78) 4.21 (107) 5.18 (132)
5 0.258
(127 6.6)
120 (1926) 1.71 (43) 2.97 (75) 4.00 (102) 4.90 (124)
130 (2087) 1.88 (48) 3.24 (82) 4.35 (110) 5.32 (135)
6 0.864
(152 21.9)
120 (1926) 1.04 (26) 2.28 (58) 3.32 (84) 4.23 (107)
130 (2087) 1.19 (30) 2.60 (66) 3.68 (93) 4.67 (119)
6 0.432
(152 11.0)
120 (1926) 1.45 (37) 2.71 (69) 3.75 (95) 4.67 (119)
130 (2087) 1.62 (41) 2.99 (76) 4.10 (104) 5.08 (129)
6 0.280
(152 7.1)
120 (1926) 1.65 (42) 2.91 (74) 3.94 (100) 4.84 (123)
130 (2087) 1.82 (46) 3.19 (81) 4.30 (109) 5.27 (134)
TABLE 10 (continued)
Fire Resistance of Clay-Masonry-Protected Steel Columns
1
1. Tabulated values assume a 1 in. (25.4 mm) air gap between masonry and steel section.
www.gobrick.com | Brick Industry Association | TN 16 | Fire Resistance of Brick Masonry | Page 16 of 16
Penetrations
All penetrations through an assembly with a fire resistance rating must conform to building code requirements for
those penetrations. Codes typically require doors and windows to have a fire resistance rating, though depending
on application, the fire resistance ratings of the doors and windows may be less than that of the surrounding
wall construction. For mechanical ducts, fire dampers are typically required. For smaller penetrations, such as
for drainage pipes and conduits, the space around the pipes is typically required to be filled with a fire resistant
material and sealed to the surrounding masonry with a sealant rated for a specific fire resistance. In all cases,
products used to seal penetrations should be carefully researched, selected and installed to ensure that the fire
resistance rating of a wall is not compromised.
Other Details
In fire resistant construction, the intent is to provide fire resistance that surrounds a three-dimensional, occupied
space. Where a wall with a fire resistance rating meets an interior wall or floor/ceiling assembly, the integrity of the
walls fire resistance rating should be maintained. In closely spaced buildings, a brick veneer wall assembly with a
1-hour fire resistance rating on the exterior side is typically unaffected by the intersection of interior partition walls.
However, in a brick veneer wall assembly with a 2-hour fire resistance rating, the interior wallboard is a required
component and may have to be installed prior to framing the interior partitions. Each project and circumstance may
require specific details to maintain the fire resistance rating of the masonry assemblies described above.
SUMMARY
Brick masonry has traditionally been used to provide superior fire resistance and safety for occupants. The
fire resistance rating required by the building code for a wall will depend on many factors including the type
of construction, the use of building, and the location of the wall within the building. Brick wall assemblies of
varying styles have been tested to provide designers with standard wall sections and details that comply with the
various fire resistance ratings. Alternatively, for non-standard assemblies, calculation methods presented herein
can provide the fire resistance rating of the proposed wall sections based on the results of previously tested
assemblies.
The information and suggestions contained in this Technical Note are based on the available data
and the experience of the engineering staff and members of the Brick Industry Association. The
information contained herein must be used in conjunction with good technical judgment and a
basic understanding of the properties of brick masonry. Final decisions on the use of the informa-
tion discussed in this Technical Note are not within the purview of the Brick Industry Association
and must rest with the project architect, engineer and owner.
REFERENCES
1. 2006 International Building Code, International Code Council, Inc., Country Club Hills, IL, 2006.
2 2006 International Residential Code, International Code Council, Inc. Country Club Hills, IL, 2006.
3. ASTM E119-07, Standard Test Methods for Fire Tests of Building Construction and Materials, Annual Book
of Standards, Vol. 04.07, ASTM International, West Conshohocken, PA, 2007.
4. Borchelt, J.G., and Swink, J.E., Fire Resistance Tests of Brick Veneer/Wood Frame Walls, Proceedings
of the 14th International Brick and Block Masonry Conference, University of Newcastle, Callaghan,
Australia, 2008.
5. Code Requirements for Determining Fire Resistance of Concrete and Masonry Construction Assemblies,
(ACI 216.1-07 / TMS-0216-07), The Masonry Society, Boulder, CO, 2007.
6. Fire-Resistance Classifications of Building Constructions, BMS92, National Bureau of Standards,
Washington, D.C., 1942.
7. UL Fire Resistance Directory, Underwriters Laboratories, Northbrook, IL, 2007.

Technical Notes 17 - Reinforced Brick Masonry - Introduction
Reissued Oct. 1996
Abstract: The concept and use of reinforced brick masonry (RBM) has a long history. This Technical Notes
documents the history of RBM. Recent and current code provisions are enumerated. Several applications of RBM
show the variety of possible uses.
Key Words: applications, brick, constructions, history, reinforced brick masonry, reinforcement, research.
INTRODUCTION
Reinforced brick masonry (RBM) consists of brick masonry which incorporates steel reinforcement embedded in mortar
or grout. This masonry has greatly increased resistance to forces that produce tensile and shear stresses. The
reinforcement provides additional tensile strength, allowing better use of brick masonry's inherent compressive strength.
The two materials complement each other, resulting in an excellent structural material. The principles of reinforced brick
masonry design are the same as those commonly accepted for reinforced concrete, and similar formulae are used.
Brick masonry is one of the oldest forms of building construction, and reinforcement has been used to strengthen
masonry since 1813. In the modern sense reinforced brick masonry in the United States is a relatively new type of
construction, with specific design procedures and construction methods. These have been developed from experimental
investigations beginning in the 1920's and with the experience of the performance of thousands of reinforced masonry
buildings. These structures demonstrate the practicality and economy of the construction, and their performance
confirms the soundness of the design principles. Figure 1 shows the Los Angeles Police Department, Devonshire
Station, a reinforced brick structure, located 3 miles (4.8 km) from the epicenter of the Northridge earthquake. There
was no structural damage and the building reportedly functioned as an emergency services coordination center
following the 6.7 magnitude earthquake.
Los Angeles Police Department, Devonshire Station
FIG. 1
This Technical Notes presents the history of reinforced brick masonry with a review of recent research and
applications. Other Technical Notes in this series provide information on the design of reinforced brick masonry
including applications such as beams, lintels, and retaining walls.
HISTORY
Marc Isambard Brunel is credited with the discovery of reinforcd masonry. He first proposed the use of reinforced brick
masonry in 1813 as a means of strengthening a chimney then under construction. However, it was in connection with
the building of the Thames Tunnel in 1825 that he made his first major application of reinforced brick masonry. As a
part of the construction of this tunnel, two brick shafts were built, each 30 in.(760 mm) thick, 50 ft (15 m) in diameter
and 70 ft (21m) deep.
The shafts were reinforced vertically with wrought iron rods 1 in. (25 mm) in diameter, built into the brickwork. Iron
hoops, 9 in. (230 mm) wide and 1/2 in.(13 mm) in thickness, were laid in the brickwork as building progressed. The first
shaft was built to a height of 42 ft (13 m) and then sunk by excavating soil from the interior, using what is now
commonly known as the open method of cassion construction. The remaining 28 ft (8.5 m) of its height was added to
t17 http://www.gobrick.com/BIA/technotes/t17.htm
1 of 7 9/13/2009 12:51 PM
the top of the shaft as it settled and was stabilized by underpinning.

In spite of unequal settlement of the shaft no cracks developed in the brick masonry. As a result, these cond shaft was
built to its entire height of 70 ft (21 m) before it was lowered. Richard Beamish, in his Memoirs of the Life of Sir Marc
Isambard Brunel [1], describes this construction and states that, after an unequal settlement of 7 in. (180 mm) on one
side and 3 in. (76mm) on the other, "the surge was alarming, but so admirably was the structure bound together that no
injury was sustained." Brunel continued the use of reinforced masonry and in 1836 constructed test structures in an
effort to determine the additional strength imparted to the masonry by the reinforcement.
Other engineers became interested in this type of construction and in 1837 Colonel Pasley of the Corps of Royal
Engineers conducted a series of tests on reinforced brick masonry beams and reported results comparable to those
obtained by Brunel. Pasley's tests were designed to settle the prevailing argument as to whether the flat hoop iron used
as reinforcement really strenthened brick beams.
Three beams were built, each 18 in. (460 mm) wide and 12 in (305 mm) (4 brick courses) deep, with a 10 ft (3 m)
span. One beam was built without reinforcement, with the brick laid in neat cement. The second beam was also laid in
neat cement, but this beam was reinforced with 5 pieces of hoop iron; two placed in the top mortar joint, one in the
middle joint and two in the bottom mortar joint. The latter of these obviously carried most of the tensile stress. The third
beam was reinforced in the same manner as the second beam, but the brick were laid in a mortar composed of 1 part
lime and 3 parts sand. The first beam failed at a load of 498 lb (2.2 kN); the second beam carried 4723 lb (21.0 kN);
and the third beam failed at between 400 and 500 lb (1.8 and 2.2 kN); thus settling the dispute. The results point out
that bond between the brick, mortar and reinforcement develops when cement-based mortars are used.
As indicated by the placement of the reinforcementin Pasley's beams, the manner in which steel and masonry act
together to resist forces was not completely understood at the time. The empirical formulae dereved from such tests
could not be used to determine dimensions and reinforcement of structural members varying in cross section or span
from those tested. However, the interest in reinforced masonry construction continued and, with the increased use of
cement in mortar, additional tests were conducted.
One such test that received widespread publicity was a reinforced brick beam tested at the Great Exposition in London
in 1851. The "new cement," comercially known as Portland Cement was used in the construction. This test was highly
successful, and the publicity which it received resulted in the more widespread use of portland cement in several
European countries and, to a lesser degree, in the United States.
N. B. Corson published an article in the July 19, 1872 issue of Engineering [5] in which he reviewed the data obtained
from the Exposition's test beam, Brunel's test structures, tests of unreinforced masonry beams and arches, and the
performance of a large number of masonry structures. From these data, Corson computed tensile stresses of
unreinforced masonry and recommended an allowable tensile stress for use in the design of masonry lintels. This
appears to be the first recorded technical discussion of the relation of tensile strength of masonry to mortar strength.
However, it did not recognize the full effect of the metal reinforcement in increasing the tensile strength of a member.
The use of reinforced brick masonry continued to spread. The benefits of combining the tensile strength of iron or steel
with the compressive strength of masonry was evident to those familiar with the potential damage of earthquakes. The
Palace Hotel opened in San Francisco in 1875, covering a full city block, rising seven stories in height. The 3 ft (0.9 m)
thick solid brick walls were reinforced by iron bands every few feet. These formed a "basket" that completely encircled
the building. This is one of the few large structures that endured the 1906 San Francisco earthquake [2].
During the period 1880 to 1920, there was little recorded use of reinforced brick masonry and experimental
investigations of this type of construction appear to have been practically discontinued.
In 1923, the Public Works Department of the Government of India published Technical Paper No. 38 [3], a
comprehensive report by Undersecretary A. Brebner of extensive tests of reinforced brick masonry structures extending
over a period of about two years. A total of 282 specimens were tested, including reinforced brick masonry slabs of
varying thickness, reinforced brick beams, both reinforced and unreinforced columns, and reinforced brick arches. The
tests reported by Brebner appear to be the first organized research program on inforced brick masonry and the data
obtained provided answers to questions raised regarding this type of construction. This research marks the initial stage
of the modern development of reinforced masonry.
Following Brebner's report and his statement of a rational design theory for reinforced brick masonry, its use increased,
particularly in India and Japan. Both countries are subject to severe earthquakes, and buildings expected to withstand
such shocks must be designed with relatively high resistance to lateral forces. Since structural steel and suitable lumber
for concrete formwork were relatively expensive in these countries, engineers turned to reinforced brick masonry. It be-
t17 http://www.gobrick.com/BIA/technotes/t17.htm
2 of 7 9/13/2009 12:51 PM
came standard construction for public and important private buildings, as well as for many types of engineering
structures, such as retaining walls, bridges, storage bins and chimneys.
Brebner wrote in 1923 of reinforced brick masonry, "In all, nearly 3,000,000 ft
2
(279,000 m
2
) have been laid in the last
three years." Skigeyuki Kanamori, Civil Engineer, Department of Home Affairs, Imperial Japanese Government, is
reported in the July 15, 1930 issue of Brick and Clay Record [7] as stating, "There is no question that reinforced
brickwork should be used instead of (unreinforced) brickwork when any tensile stress would be incurred in the structure.
We can make them more safe and stronger, saving much cost. Further, I have found that reinforced brickwork is more
convenient and economical in building than reinforced concrete and, what is still more important, there is always a very
appreciable saving in time." Structures described by Kanamori include sea walls, culverts and railway retaining walls, as
well as buildings.
Research in the United States, sponsored by the Brick Manufacturers Association of America and continued by the
Structural Clay Products Institute and the Structural Clay Products Research Foundation contributed much valuable
material to the literature on reinforced brick ma-
sonry. Since 1924, numerous field and laboratory tests have been made on reinforced brick beams, slabs and columns,
and on full size structures. Fig. 2 is an example of a 1936 test to demonstrate the structural capabilities of reinforced
brick masonry elements.
Early Test of RBM Element
FIG. 2
During this period, research was conducted on both reinforced and unreinforced brick masonry at the National Bureau
of Standards, now the National Institute of Standards and Technology, and at practically all of the principal engineering
colleges of the United States. As new data was developed through research, the er ratic performance of some of the
earlier reinforced brick test specimens could be explained and, one by one, the principal variables affecting the strength
of reinforced brick masonry have been identified and, to large degree, evaluated.
In 1933 the Brick Manufacturers Association of America published Brick Engineering, Vol. 111, Reinforced Brick
Masonry, by Hugo Filippi [6]. Regarding the uses of reinforced brick masonry, the author states, "Reinforced brick
masonry is well adapted for use in the following types of structures, either wholly or in part: Buildings, Culverts and
Bridges; Retaining Walls and Dams; Reservoirs; Sewers and Conduits; Tanks and Storage Bins; Chimneys and Circular
Constructions; Abutments, Piers, Trestle Bents, etc.
"In the United States alone, during the past year and one-half, more than 40 individual jobs of reinforced brick masonry
have been built, consisting of such distinctive types of construction as highway bridges, storage bins, industry track
trestle piers, floor and roof slabs, beams, girders and long lintels. At the present time approximately 50 additional jobs
are either under construction or under consideration in various parts of the country."
During the period referred to by Filippi, the development and use of reinforced brick masonry in the United States were
in their early stages. A significant change in the use of RBM came after the 1933 Long Beach earthquake. It was
realized that unreinforced structures were susceptible to major damage from earthquakes and that RBM could be used
to save lives. Codes were developed that promoted the use of reinforced structures. Since that time thousands of such
t17 http://www.gobrick.com/BIA/technotes/t17.htm
3 of 7 9/13/2009 12:51 PM
structures have been built and reinforced brick masonry construction has been adopted as standard practice for various
types of structures in many areas.
RECENT RESEARCH
Research on reinforced brick masonry has continued. In 1984, the Technical Coordinating Committee for Masonry
Research (TCCMAR) was formed for the purpose of defining and performing both experimental and analytical research
and development necesary to improve structural masonry technology [9]. A unique aspect of this research was a
phased step-by-step program of sepate, but coordinated research tasks. Initial research on materials was used in later
tests on assemblies. These led to tests of building elements and then the combination of wall and floor elements. The
research culminated in a full-scale, five story structure subjected to dynamic loading in 1993. Much of the research led
to the development of a limit states design procedure for masonry.
Interest in better utilization of brick masonry's high compressive strength has led to research in prestressed brick
masonry. Knowledge about this form of reinforced brick masonry was increased by research in Great Britain. Research
is currently underway in the United States, as is the development of design procedures.
BUILDING CODE PROVISIONS
Building codes first covered reinforced brick masonry in 1953 in the American National Standards Institute's A41.2
document [4]. Since that first code on RBM, other codes such as the Uniform Building Code and the Masonry
Standards Joint Committee Code (ACI
530/ASCE5/TMS 402) have adopted provisions.
Most code provisions on reinforced masonry arebased on allowable stress design (ASD). In ASD, the reinforcement in
masonry is designed to resist all tensile forces. The reinforcement increases the masonry's shear resistance and may
contribute to the compressive strength. The stress-strain relationship is linear at working loads and the strain is
proportional to the distance from the neutral axis. Code requirements cover axial compression, flexure, and shear.
The Uniform Building Code has provisions for slender wall design, which is loosely based on strength design. A more
comprehensive design method, known as limit states design is in development. Limit states design considers the actual
performance of the materials as they undergo load and deformation. Significant changes in the state of stress, such as
cracking of the masonry and yielding of the steel, are identified. The capacity, or strength, of the element at these limit
states is compared to that required to resist the applied load. These code provisions are expected to provide a
complement to ASD.
BASIC CONSTRUCTION PROCEDURES
The earliest method of placing reinforcement into brick masonry was simply to place iron or steel bars in mortar joints
as the bricks were laid. Later the reinforcement was placed in collar joints between two masonry wythes and
surrounded by mortar or fine grout.
Eventually the space between wythes was increased in width and filled with grout. Horizontal reinforcement and grout
were placed as the outer wythes were completed. The next development was the "High Lift Grouting System" in which
the brick masonry wythes are built up around the reinforcement and allowed to set for a minimum period of three days.
Then grout is pumped into the space containing the reinforcement. This method was developed in the San Francisco
area during the late 1950s. This double wythe reinforced brick masonry is shown in Fig. 3.
t17 http://www.gobrick.com/BIA/technotes/t17.htm
4 of 7 9/13/2009 12:51 PM
Double Wythe Reinforced Brick Masonry
FIG. 3
The most recent means of constructing reinforced masonry incorporates hollow brick. These units are manufactured
with large open cells which align vertically when the units are laid. Vertical reinforcement is placed in the cells by laying
the brick over or around the bars, or by threading the bar in after the brick are laid. Horizontal reinforcement is placed
in bed joints or in continuous bond beams made by removing portions of the webs that connect the face shells. Spaces
containing reinforcement are grouted in lefts of up to 5 ft (1.5m) to make grout pours of up to 24 ft (7.3m). Construction
of reinforced hollow brick masonry is shown in Fig. 4
Reinforced Hollow Brick Masonry
FIG. 4
APPLICATIONS AND EXAMPLES
During the past 60 years, reinforced brick masonry has been used for the construction of a variety of structures. In those
countries where labor costs are low, one of its principal uses has been for the construction of floor and roof slabs.
However, in the United States, its most extensive use has been in the construction of vertical members, such as walls and
columns. Since no forms are required for these members, reinforced brick masonry is competitive with reinforced
concrete, and walls of minimum thickness and light structural members can be constructed at substantially less cost in
reinforced brick masonry than in reinforced concrete.
Reinforced brick beams and lintels allow the designer to achieve exposed brick on the underside of these elements as in
Fig. 5.
t17 http://www.gobrick.com/BIA/technotes/t17.htm
5 of 7 9/13/2009 12:51 PM
Reinforced Brick Beams
FIG. 5
This provides a hoizontal finished surface that matches the vertical surface. The idea of brick hanging upside down must
be disconcerting. Some designers seem reluctant to use RBM construction for brick lintels or soffits. As demonstrated by
tests since 1837, the bond of the mortar and grout to the brick holds the brick in place.
Structures of all sizes, from single story residences to 23 story buildings have been constructed of reinforced brick
masonry as shown in Figs. 6 and 7.
Reinforced Brick Masonry Single Family Residence, Ashbrun VA
FIG. 6
Reinforced Brick Masonry High Rise, Cleveland, OH
FIG. 7
The applications range from retaining walls to exterior cladding. The added tensile strength of the reinforcing steel opens
the possibility for prefabricated brick panels. This method of design and construction is utilized frequently to achieve
unusual shapes and bond patterns in brick masonry. See Fig. 8.
t17 http://www.gobrick.com/BIA/technotes/t17.htm
6 of 7 9/13/2009 12:51 PM
Reinforced Brick Masonry Panels
FIG. 8
SUMMARY
The use of reinforced brick masonry has been recorded for over 175 years. RBM construction has been adapted to a
wide variety of applications throughout its history. Beams, column, pilasters, arches, and other RBM elements have been
used in buildings, culverts, retaining walls, silos, chimneys, pavements and bridges. Continuing research on RBM results in
more economical structures able to withstand all types of loading.
The information and suggestions contained in this Techical Notes are based on the available data and the experience of
the engineering staff of the Brick Institute of America. The information contained herein must be used in conjunction with
good technical judgment and a basic understanding of the properties of brick masonry. Final decisions on the use of the
information contained in this Technical Notes are not within the purview of the Brick Institute of America and must rest with
the project architect, engineer and owner.
REFERENCES
1. Beamish, R., Memoirs of the Life of Sire Marc Isambard Brunel, Longmans, London, England, 1862.
2. Berger, Molly W., "The Old High-Tech Hotel," Invention and Technology, Fall 1995, pp. 46-52.
3. Brebner, A., Notes on Reinforced Brickwork, Technical Paper No. 38, Government of India, Public Works Deparment,
India, 1923.
4. "Building Code Requirements for Reinforced Masonry," American Standards A 41.2-1960, American Standards
Association, New York, NY, 1960.
5. Corson, N. B., "Article on Brick Masonry," Engineering, London, July 19, 1872.
6. Filippi, Hugo, Brick Engineering, Volume III, Reinforced Brick Masonry, Brick Manufacturers Association of America,
Cleveland, OH, 1933.
7. Kanamori, S., "Reinforced Brickwork Opens Greater Possibilities," Brick and Clay Record, Chicago, IL, Vol. 77, #2,
July 1930, pp.96-100.
8. Plummer, H. C. and Blume, J. A., "Reinforced Brick Masonry and Lateral Force Design," Structural Clay Products
Institute, Washington, D.C., 1953.
9. "Status Report, U.S. Coordinated Program for Masonry Building Research," Technical Coordinating Committee for
Masonry Research, Nov. 1988.
t17 http://www.gobrick.com/BIA/technotes/t17.htm
7 of 7 9/13/2009 12:51 PM

Technical Notes 17A - Reinforced Brick Masonry - Materials and Construction
Reissued Aug. 1997
Abstract: This Technical Notes provides a discussion of the proper methods of constructing reinforced brick
masonry. Materials used in reinforced brick masonry are included Construction of brick masonry, placement of steel
reinforcement and grouting are addressed. Recommendations are provided to ensure that the completed masonry
will provide adequate performance. Particular empahasis is placed on those aspects of construction that are unique
to reinforced brick masonry. Various quality assurance procedures and tests are also explained.
Key Words: bracing, brick ,construction, grouting, inspection, reinforced brick masonry reinforcement, shoring.
INTRODUCTION
Reinforced brick masonry (RBM) is different from more conventional brick veneer in many ways. Key to those
differences is the concept of grouting the brick masonry. Ground brick masonry is defined as construction made with
clay or shale units in which cavities or pockets in elements of solid units, or cells of hollow units are filled with grout.
Common examples of RBM elements are beams, columns, pilasters, multi-wythe brick walls with grouted collar joints
and hollow brick walls. This Technical Notes reviews the materials and construction practices used to build RBM
elements. The different techniques are discussed with particular emphasis on the concepts of grouting and the
placement of reinforcement. Quality assurance and minimum standards of workmanship to ensure a high level of
consistency and adequate masonry performance are addressed.
The information in this Technical Notes should be carefully reviewed by the mason contractor prior to constructing
reinforced brick masonry. It should also be studied by the masonry inspector. Other Technical Notes in this series
provide design theories and design aids for RBM elements such as beams, walls, columns and pilasters.
RBM MATERIALS
The materials used to construct RBM elements should comply with applicable ASTM standards. Brick should meet
the requirements of ASTM C 62 Specification for Building Brick, C 216 Specification for Facing Brick, or C 652
Specification for Hollow Brick. Mortar should comply with the requirements of ASTM C 270 Specification for Mortar
for Unit Masonry. Grout should comply with ASTM C 476 Specification for Grout for Masonry. Metal wall ties, bar
positioners, and reinforcing bars and wires should comply with the applicable ASTM standards as required by the
Specification for Masonry Structures (ACI 530.1/ASCE 6/TMS 602)[2], also known as the MSJC Specification. All
metal wall ties, positioners and joint reinforcement should be corrision resistant or protected from corrosion by
appropriate coatings. Refer to Technical Notes 3A for a discussion of the material properties of brick, mortar, grout
and reinforcement.
The materials in both fine and coarse grout should comply with the requirements of ASTM C 476 Specification for
Grout for Masonry. Both fine grout and coarse grout should comply with the volume proportions given in ASTM C
476. Specifying grout by proportions is preferred over specifying a minimum grout strength. Typically, the maximum
aggregate size should be 3/8 in. (9.5 mm) for coarse grout. While larger size aggregate can be used when filling
large grout spaces, it must be noted that such grout likely cannot be pumped and will require placement by pouring
from a hopper.
It must be remembered that grout is different from concrete. Concrete is placed with a minimum of water into
nonporous forms. Grout is poured with considerably more water, as the brick masonry creates absorptive forms.
Grout should be sufficiently fluid to flow into the space to be filled, and surround the steel reinforcement, leaving no
voids. It should be wet enough to flow without separation of the constituents. Whereas good mortar should stick to
a trowel, it should be impossible for grout to do so. The water cement ratio as mixed, highly important in concrete
work, is less important for grout in brick masonry. Although excessive water is detrimental to the strength and
durability of the grout, when introduced into the brick masonry the water cement ratio rapidly changes from a high to
a low value. Grout is often mixed too dry and stiff for proper placement.
Home http://www.gobrick.com/BIA/technotes/t17a.htm
1 of 10 9/13/2009 12:51 PM
One concern with the use of a very fluid grout mixture is excessive shrinkage. Shrinkage can create voids in the
grout space, which are to be avoided. For this reason, plasticizers and shrinkage-compensating admixtures are
recommended for grout in brick masonry. Such admixtures will provide the necessary fluidity while also providing a
hardened grout mixture with minimal voids.
There is a temptation to fill the grout space with the mortar that is used to lay up the brickwork, especially when
simultaneously laying brick and grouting. This is not recommended, but may be permitted by local building codes. It
is common to find excessive voids in the grout space with this practice. Proper placement and consolidaton of grout
or grout mixture with a shrinkage-compensating admixture and poured in a continuous process is much more likely to
form a solid grout fill.
RBM CONSTRUCTION
The construction of RBM elements can be separated into three parts: brick masonry construction, placement of the
steel reinforcement and grouting. Each of these steps is critical to the end result. Following is a review of the three
construction procedures in the order of their execution.
There are two key points to remember when laying the brick. First, the brick masonry is the permanent formwork for
the grout. This masonry formwork must be built in a manner that facilitates placement and positioning of the steel
reinforcement and installation of grout. Second, the quality of workmanship will have a significant impact on the
strength of the RBM. Unfilled mortar joints and elements that are out-of-plumb will not provide the performance
assumed by the designer.
All RBM elements constructed of solid brick should be laid with full head and bed joints. The ends of brick should be
buttered with sufficient mortar to fill the head joints. Furrowing of bed joints should not be deep enough to result in
voids. Years ago, it was believed by some that the head joints in solid brick masonry could be made only half full and
that the grout would flow into the remainder of the head joint and fill the voids. It was felt that the grout would form a
shear key and make the brick masonry bond more strongly to the grout core. This is not the case. In fact, creation
of voids is more likely with this practice, which reduces the masonry's strength and can promote efflorescence due to
entrapped water.
Hollow brick are normally laid with face shell bedding. That is, the unit's face shells are filled solidly with mortar and
head joints are filled with mortar to a depth equal to the face shell thickness. In some instances, bed joints of cross
webs are covered with mortar to confine grout or to increase net area. Head joints may be filled solid for similar
reasons.
Cleanouts and Maintaining a Clear Grout Space
Cleanouts are used to remove all mortar droppings and debris from the bottom of a grout space and also to ensure
proper placement of reinforcement prior to grouting. Cleanouts should be provided in the bottom course of all spaces
to be grouted when the grout pour exceeds 5 ft (1.5 m) in height. In partially grouted masonry, a cleanout is
recommended at each vertical bar. In fully grouted masonry, the spacing of cleanouts should not exceed 32 in. (813
mm) on center according to the MSJC Specification. For reinforced brick masonry elements constructed with solid
brick, cleanouts should be formed by omitting brick in the bottom course periodically along the base of the element.
For hollow brick masonry, cleanouts should be provided in the bottom course of masonry by removing the face shell
of the cells to be grouted. Examples of cleanouts in brick masonry walls are shown in Figure 1.
Example of Grout Space Cleanouts
Home http://www.gobrick.com/BIA/technotes/t17a.htm
2 of 10 9/13/2009 12:51 PM
FIG. 1
The minimum cleanout opening dimension should be 3 in. (76 mm). However, smaller spaces can be used if it is
shown with a demonstration panel that the spaces can be cleaned.
The grout spaces should be cleaned prior to grouting. It is good practice to clean out grout spaces at the end of
each work day so that mortar droppings can be easily removed. A high pressure water spray, compressed air or
industrial vacuum cleaner should be used for this purpose. Many contractors have found that cleaning of the grout
space is facilitated by placing a layer of sand or sheets of plastic film at the bottom of the cleanout to catch mortar
droppings. After cleaning and prior to grouting, cleanouts should be closed with masonry units or sealed with a
blocker to resist grout pressure. A minimum curing time of two days is recommended for the cleanout plugs or they
should be adequately braced against the grout pressure. Bracing is discussed further in the section on Shoring and
Bracing.
For solid brick masonry, the top of the mortar bed joint should be beveled outward from the center of the grout space
to minimize the amount of mortar extruded into the grout space when the brick are laid, as illustrated in Fig. 2.
Beveling Mortar Bed Joints
FIG. 2
Mortar protruding from bed or head joints into the grout space should be struck flush with the surface or removed
prior to grouting. The maximum protrusion of a mortar fin should be 1/2 in. (13 mm). The spaces to be grouted
should also be kept free of mortar droppings. One method of keeping collar joints clear consists of laying wood
strips on the metal ties as the two wythes of brick masonry are built. The strips catch mortar droppings during
construction and are removed by means of attached heavy strings or wires as the wall is built. To keep the cells of
hollow brick clear for grouting, sponges are typically used, as shown in Fig. 3.
Sponges to Keep Cell Clear of Mortar Droppings
FIG. 3
Erection Tolerances
All RBM elements should be laid within the permitted dimensional tolerances found in the MSJC Specifications.
Masonry elements that are not constructed within these limits are not as strong in compression as those that are.
The thickness of mortar joints will also influence the masonry's strength. Excessively thin or thick mortar joints will
reduce brick masonry's tensile and compressive strength. The erection tolerances stated in the MSJC Specification
are given in Table 1.
Home http://www.gobrick.com/BIA/technotes/t17a.htm
3 of 10 9/13/2009 12:51 PM
This specification and its accompanying Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS
402) [1] [also known as the MSJC Code] stipulate minimum size of grout spaces that are dependent on the height of
grout pour and the grout type. The limits given in Table 2 are to ensure adequate access of grout to the space.
Shoring and Bracing
RBM elements typically require temporary support during construction provided by shoring and bracing. These
supporting members are typically of wood or steel construction. Temporary support is required for two reasons.
First, grout is very fluid when placed and exerts considerable pressure on the surrounding brick masonry. Second,
RBM elements gain strength over time as the mortar and grout cure and harden. RBM walls, columns and pilasters
are often braced along their height. RBM beams and arches may require both shoring for vertical support and
bracing for lateral load resistance and grout pressure resistance.
Shoring and bracing should be left in place until it is certain that the masonry has gained sufficient strength to carry its
own weight and all other imposed loads including temporary loads that occur during construction. The most common
problem related to temporary supports for masonry elements is inadequate lateral bracing to resist wind pressures
during construction until the masonry has gained sufficient strength to resist these loads. This is especially true when
the roof and floor diaphragms have not been installed and anchored to the top of the masonry wall. Without proper
bracing, the wall is a free-standing cantilever element and is more vulnerable to collapse.
Appropriate time for removal of shoring and bracing depends on many factors. For example, proper curing of the
mortar and grout may take considerably longer under cold weather conditions. The results of suitable compression
tests of prisms or grout may be necessary as evidence that the masonry has attained sufficient strength to permit
removal of shoring or bracing. Rules-of-thumb for the minimum time which should elapse before removal of shoring
or bracing that have been recommended for many years include the following:
Home http://www.gobrick.com/BIA/technotes/t17a.htm
4 of 10 9/13/2009 12:51 PM
1. For RBM beams, 10 days after completion of the element
2. For RBM arches, 7 days after completion
3. Lateral bracing for walls, columns and pilasters, 7 days after placement of the grout.

Longer time periods will be necessary with inadequate curing conditions. It is always a good idea to consult the
project engineer for a recommended bracing scheme and the length of time required for bracing to remain in place.
Curing Time Prior to Grouting. If grouting is performed too rapidly after construction the hydrostatic pressure of
the grout can cause "blowout" of mortar joints or even entire sections of brickwork. This is especially true when the
grout pour is high. Blowout of the grout can be avoided by a combination of proper curing time, adequate wall ties or
joint reinforcement across the grout space and bracing.
Recommended duration of curing prior to grouting depends upon the method of grouting and the extent of bracing to
resist the grout pressure. If no bracing against grout pressure is provided, the masonry should be permitted to cure
for at least 3 days to gain strength before placement of grout in lifts greater than 5 ft (1.5 m) in height. For shorter
grout lift heights, grout may be poured relatively soon after the brick are laid. Since grout lift heights are very short,
the mason contractor should adjust the speed of construction as needed to avoid blowout of the wall.
Wall Ties Across Grout Spaces. Freshly placed grout exerts a hydrostatic pressure on the surrounding masonry
formwork. This pressure increases with increasing pour height. To resist the grout pressure, wall ties are used
across the grout space to tie the brick wythes together. For multi-wythe masonry walls, a minimum number of wall
ties will already be provided to tie the wythes together in accordance with the building code. The wall ties resist the
grout pressure by their tensile capacity. The ties provide the additional benefit of a positive mechanical anchorage
between the grout core and the surrounding masonry. Ties may not be required across small grout spaces such as
in columns or pilasters.
Wall ties across grout spaces should be at least W 1.7 (9 gage) wire. For masonry elements laid in running bond,
ties should be spaced not more than 24 in (610 mm) o.c. horizontally and not more than 16 in. (406 mm) o.c.
vertically. If stack bond is used, the vertical spacing should be reduced to 12 in. (305 mm) o.c. All ties should be
placed in the same line vertically to facilitate the grout consolidation process. Ties should be embedded at least
one-half the thickness of the masonry wythe.
Bracing Against Grout Pressure. For grout pour heights less than approximately 5 ft (1.5 m), bracing of the brick
masonry may not be necessary. If the grout pour height is greater, consideration should be given to bracing the
masonry. This is especially true when a longer curing time for the brick masonry prior to grouting is not feasible.
Bracing members are typically externally applied wood construction. The bracing members should be designed by an
engineer, based on the grout pour height.
PLACEMENT OF STEEL REINFORCEMENT
Steel reinforcement should be placed in accordance with the size, type and location indicated on the project
drawings, and as specified. Dissimilar metals should not be placed in contact with each other because this can
promote corrosion of the reinforcement. Nonmetallic flashing should be used when the flashing will come in contact
with the reinforcement. If it is possible, all vertical steel reinforcement should be placed after completion of the
masonry surrounding the grout space. This keeps the reinforcement out of the mason's way during construction and
makes cleaning of the grout space easier. It also prevents contamination of the reinforcement by mortar droppings
or protrusions that can adversely affect grout bond to the reinforcement.
Applicable building codes should be consulted regarding placement requirements for reinforcement in masonry
elements. A summary of the placement requirements for reinforcement in masonry stated in the MSJC Code is given
in Table 3.

Home http://www.gobrick.com/BIA/technotes/t17a.htm
5 of 10 9/13/2009 12:51 PM
1
In Flexual members, the "d" dimension is the distance from the extreme compression face to the
centroid of the tensile reinforcement.
These requirements are to ensure proper bond to the grout, corrosion protection and fire resistance of the
reinforcement. Table 3 also identifies the tolerance limits on positioning reinforcement in masonry elements.
Reinforcement should only be spliced where indicated on the project drawings. Reinforcement should not be bent or
disturbed after placement of the grout. Vertical reinforcement should be accurately placed and secured prior to the
grouting process. Reinforcement can be secured by wire ties or other spacing devices. Some examples of common
bar spacing devices are shown in Fig. 4.
Bar Spacing Devices
FIG. 4
Vertical reinforcement should be braced at the top and bottom of the element. Additional positioners may be
necessary to facilitate proper placement of the bars. When reinforcement is spliced in a grout space between
wythes or within an individual cell of hollow brick masonry, the two bars should be placed in contact and wired
together. Vertical reinforcement in hollow brick masonry may be spliced by placing the bars in adjacent cells,
provided the distance between the bars does not exceed 8 in. (204 mm).
Home http://www.gobrick.com/BIA/technotes/t17a.htm
6 of 10 9/13/2009 12:51 PM
Horizontal reinforcement is usually placed in the mortar joints as the work progresses or in bond beams at the
completion of the bond beam course. In partially grouted walls, the bond beam should be grouted prior to further
construction of brick masonry on top of the bond beam. For two-wythe, solid brick masonry walls, the horizontal
reinforcement may be placed in the grouted collar joint. All horizontal bars should be on the same side of the vertical
reinforcement to facilitate consolidation of the grout.
GROUTING
The most crucial aspect of constructing RBM elements is the grouting process. While grouting may seem a simple
matter of filling cavities or cells of masonry, it is the one aspect of RBM construction that can cause the most
problems. The most common problem is the creation of voids in the grout space due to stiff grout, excessive pour
height, grout shrinkage, or blocked grout spaces. To ensure proper grouting, four sequential steps should be
properly executed: preparation of the grout space, grout batching, grout placement and consolidation, and curing
and protection.
Preparation of the Grout Space
The configuration and condition of the grout space can vary considerably. Common grout spaces for RBM elements
are the cells of hollow brick, the collar joint between multi-wythe brick walls, the core of columns or pilasters and the
depth of a beam.
For a multi-wythe brick wall with a grouted collar joint, vertical grout barriers, or dams, should be built across the
grout space for the entire height of the wall at intervals of not more than 25 ft (7.6 m). Grout barriers control the
horizontal flow of grout and reduce segregation. With hollow brick, mortar is placed on the cross webs to confine
grout to certain vertical cells. Wire mesh is installed beneath a bond beam to prevent the flow of grout into the
masonry below the bond beam. Examples of common grout barrier techniques are shown in Fig. 5.
Vertical Grout Barriers
FIG. 5
Grout spaces should be checked to see that all foreign materials and debris have been removed prior to grouting.
The reinforcement should be clean and properly positioned in the grout space. If cleanouts are used, they should be
sealed and braced if needed. All grout barriers should be secured and braced, if necessary.
The absorption rate of brick masonry will vary considerably with different units and weather conditions. To make the
absorption more consistent, the grout space may be wetted prior to grouting. No free water should be on the units
when the grout is placed.
Grout Batching
The quantities of solid materials in the grout mix should be determined by accurate volume measurement at the time
of placing in the mixer. All materials for grout should be mixed in a mechanical mixer. Grout is most often supplied in
bulk by ready-mix trucks and pumped into place because of the volume and speed of placement required. Batching
Home http://www.gobrick.com/BIA/technotes/t17a.htm
7 of 10 9/13/2009 12:51 PM
on site is more common for smaller projects. When prepared on site, the grout mix should be batched in multiples of
a bag of portland cement as a quality control measure. If less than a single bag of portland cement is used, extreme
care should be used to accurately measure all parts.
Water, sand, aggregate and portland cement should be mixed for a minimum of 2 minutes, then the hydrated lime (if
any) and additional water should be added and mixed for an additional 5 to 10 minutes. Make and maintain as high a
flow as possible, consistent with good workability. This means that the grout should be wet enough to pour without
segregation of the constituent materials or excessive bleeding. Grout should be a plastic mix that is suitable for
pumping. The grout slump should be tested in accordance with ASTM C 143 Test Method for Slump of Hydraulic
Cement Concrete and should be between 8 and 11 in. (203 and 279 mm).
Grout Placement and Consolidation
Grout should be placed within 1 1/2 hours after the water is first added to the mix and prior to the initial set. Grout
slump should be maintained during placement. The grout pour should be done in one or more lifts and the total height
of each pour should be from the center of one course to the center of another course of brick masonry. When
grouting is stopped for 1 hour or longer, the grout pour should be stopped approximately 1 1/2 in. (38 mm) below the
top of the masonry to create a shear key.
Whenever possible, grouting should be done from the unexposed face of the masonry element. Extreme care should
be expanded to avoid grout staining on the exposed face or faces of the masonry. If grout does contact the face, it
should be cleaned off immediately with water and a bristle brush. Waiting until after curing has occurred will make
removal difficult.
Grout in contact with brick solidifies more rapidly than that in the center of the grout space. It is, therefore, important
to consolidate the grout immediately after pouring to completely fill all voids. The best procedure is to have two
people performing the operation jointly; one to pour the grout and the other to consolidate it. A mechanical vibrator
or pudding stick is used for this purpose, depending on the construction method used.
There are two methods of RBM construction: simultaneous brick construction and grouting, and grouting after brick
construction. These are sometimes referred to as "low lift" and "high lift" grouted masonry. In the first method, grout
is placed in the masonry as the courses are laid. The grout is consolidated with a pudding stick or a mechanical
vibrator. This method is typically used with narrow grout spaces. In the second method, the masonry is built to the
story height or its full height, after which grout is poured from a hopper or pumped by mechanical means. The grout
is consolidated with a low velocity vibrator with a 3/4 in. (19 mm) head. When grouting between wythes, the vibrator
should be placed in the grout at points spaced 12 to 16 in. (305 to 406 mm) apart. The grout pour height restrictions
given in Table 2 will limit the method of grout placement permitted in some instances. The mason contractor should
give consideration to the advantages and disadvantages of each method.
Simultaneous Brick Construction and Grouting. The main benefits of simultaneous construction and grouting are
elimination of cleanouts, reduction of grout pressures, and simplicity of construction. With this method of grouting,
the entire grout space can be kept entirely clear of blockage and can be easily inspected prior to grouting.
Consolidation of the entire grout pour is also easier and bracing may be lessened or eliminated.
For a multi-wythe brick wall with grouted collar joint, one wythe should be built up not more than 16 in. (406 mm)
ahead of the other wythe. Typically, the grout pour height will not exceed 12 in. (305 mm) for such walls and a
pudding stick may be used for consolidation purposes. If the grout is carried up too rapidly, there is a chance
blowout will occur. If a wythe does move, even as little as 1/8 in. (3 mm) out-of-plumb, the work should be torn down
rebuilt. This is because the bed joint bond has been broken and cannot be repaired merely by shoving the wall back
into plumb.
The grout should be placed to a uniform height between grout barriers and should be consolidated with a mechanical
vibrator or pudding stick immediately after placement. Extreme care should be exercised during grout placement and
consolidation to avoid displacement of the brick masonry.
Grouting After Brick Construction. Grouting after construction of the masonry has become the industry standard
due to its speed and the fact that grout is often supplied by ready-mix trucks. These trucks deliver large quantities of
grout, but cannot remain on site indefinitely during construction. Grout delivery must be coordinated with brick
Home http://www.gobrick.com/BIA/technotes/t17a.htm
8 of 10 9/13/2009 12:51 PM
construction and preparation of grout spaces. The grout spaces in RBM elements can be very small and become
crowded with reinforcing bars and wall ties. In addition, some contractors have commented that the highly absorptive
nature of some brick masonry causes the grout to dry out and not flow properly to the bottom of grout spaces if the
lift height is too great.
The first lift of grout should be placed to a uniform height between the grout barriers or the surrounding brick
masonry, and should be mechanically vibrated to fill all voids. This first vibration should be done within 10 minutes
after pouring the grout, while the grout is still plastic and before it has set. Grout pours in excess of 12 in. (305 mm)
should be reconsolidated by mechanical vibration after initial water loss and settlement has occurred. The
succeeding lift should be poured, vibrated and reconsolidated in a similar manner. In the first vibraton, the vibrator
should extend 6 to 12 in. (152 to 305 mm) into the preceding lift. This further reconsolidates the first lift and closes
any shrinkage cracks or separations that may have formed. The work should be planned for a single, continuous
grout pour to the top of the wall in 5 ft (1.5 m) lifts. Under normal weather conditions, in the range of 40 to 90
degrees F of (4 to 32 degrees C), the waiting period between lifts should be between 30 and 60 minutes.

Curing and Protection
The masonry work, particularly the top of the grout pour, should be kept covered and damp to prevent excessive
drying. The newly grouted masonry should be fog sprayed three times each day for a period of three days following
construction when the ambient temperature exceeds 100 degrees F (38 degrees C) or 90 degrees F (32 degrees C)
with a wind speed in excess of 8 mph (13 km/hr). The exposed faces of brickwork should be cleaned prior to the fog
spraying. Cleaning will be much more difficult if it is postponed until after this curing. The water from cleaning will
also aid in the curing process. Refer to the information in Technical Notes 1 Revised for proper construction and
protection methods during excessive cold or hot weather conditions.
For walls, columns and pilasters, at least 12 hours should elapse after construction before application of floor or roof
members, except that 72 hours should elapse prior to application of heavy, concentrated loads such as truss, girder
or beam members.
QUALITY ASSURANCE MEASURES
Not all masonry projects wil involve testing or inspection. However, the MSJC Code states that, "A quality assurance
program shall be used to ensure that the constructed masonry is in conformance with the Contract Documents." A
quality assurance program typically includes inspection of the work by an owner's representative and periodic
sampling and testing of masonry materials.
Inspection
The masonry inspector's job is to obtain good quality masonry construction and workmanship according to plans and
specifications. The inspector should be able to explain the reasons for the specified procedures and know the
important aspects of quality workmanship that will produce RBM elements with the properties assumed in the
structural design. The inspector should verify clean grout spaces prior to grouting. Type and positioning of wall ties,
bar positioners, joint reinforcement, and reinforcement should be verified against the project drawings and
specifications. During the grouting process, the inspector should verify that: the grout is proportioned properly, the
proper grouting technique is used, and all grout spaces are completely filled with grout. The inspector should look for
darkening of the masonry due to water absorption from the grout as evidence of proper grout placement. Bracing
and shoring should be inspected for proper installation. Protection measures such as covering the tops of
uncompleted work, heated enclosures, and insulation blankets should be verified
.
Testing
On some RBM projects, it may be necessary to conduct various quality control tests to ensure that the masonry has
been constructed properly. The frequency of testing should be stated in the project specifications. Testing may be
conducted prior to or during construction on the individual materials, e.g. brick, mortar, and grout. This is the most
common form of quality control testing. Brick are typically tested for compressive strength prior to construction.
Mortar may be tested in compression prior to construction in order to establish proportions of ingredients to be
measured at the jobsite. The same is true for grout, which should also be tested to verify the slump. Prism
compression tests are one example of such testing. The MSJC Specification stipulates the type, method and
frequency of material and assemblage quality control tests required for masonry elements. This document states
that prisms and grout will be tested for each 5000 sq.ft. (465 m2) of wall area or portion thereof when testing is
Home http://www.gobrick.com/BIA/technotes/t17a.htm
9 of 10 9/13/2009 12:51 PM
required. Finally, tests of samples extracted from the constructed masonry may be necessary to verify the strength
of elements when this is in question. A prism cut out of a masonry element to be used for compression testing is an
example of such a test. For a review of the common quality control tests for brick masonry, refer to the Technical
Notes 39 Series.
Quality control tests can seem an onerous and unwanted expense, but they are provided for two very important
reasons. First and foremost, tests can indicate consistency during construction. Dramatic changes in strength
properties of elements as the work progresses can indicate a problem and should be explained. The second reason
for testing is to monitor the strength gain of the masonry elements upon curing. The strength gain is monitored to
indicate when shores or bracing can be removed, when loads can be applied to an element, and to verify that the
strength assumed in the design has been achieved by the constructed masonry.
SUMMARY
Reinforced brick masonry is constructed in a manner that is different in many ways from conventional brick veneer
construction. Proper materials and construction practices as dicussed in this Technical Notes should be followed to
ensure that RBM elements achieve adequate strength and meet the applicable building code requirements.

The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Institute of America. The information contained herein must be used
in conjunction with good technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the purview of the Brick
Institute of America and must rest with the project architect, engineer and owner.
REFERENCES
1. Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402-95), American Society of Civil
Engineers, New York, NY, 1996
2. Specification for Masonry Structures (ACI 530.1/ASCE 6/TMS 602-95), American Society of Civil Engineers, New
York, NY, 1996

Home http://www.gobrick.com/BIA/technotes/t17a.htm
10 of 10 9/13/2009 12:51 PM

Technical Notes 17B -REINFORCED BRICK MASONRY - BEAMS
March 1999
Abstract: Reinforced brick masonry (RBM) beams are an efficient and attractive means of
spanning building openings. The addition of steel reinforcement and grout permits brick
masonry to span considerable distances while maintaining continuity of the building
facade. Attractive brick soffits and elimination of steel support members are two of the
advantages of reinforced brick masonry beams. This Technical Notes addresses the
design of reinforced brick masonry beams. Building code requirements are reviewed and
design aids are provided to simplify the design process. Illustrations indicate the proper
detailing and typical construction of reinforced brick masonry beams.
Key Words: beam, deflection, girder, lintel, reinforced brick masonry, reinforcement.
INTRODUCTION
Reinforced brick masonry (RBM) beams are widely used as flexural members. Common
applications of RBM beams include girders supporting floor and roof systems, and arches
and lintels spanning openings for windows and doors. Girder is the term applied to a large
beam with a long span that usually supports smaller framing members. A lintel is a beam
over a wall opening, typically simply supported with no framing members. The main
advantage of RBM beams is that the structural element and the architectural finish are one
and the same. In some cases, however, they provide economical solutions without
considering the savings due to a built-in finish. They are often built as an integral part of a
masonry wall as illustrated in Figure 1. RBM beams are designed to carry all
superimposed loads, including that portion of the wall weight above supported by the
beam. While steel lintels are more common, RBM beams provide distinct advantages over
steel lintels. Among the advantages are:
1. More efficient use of materials. The masonry serves as a structural element with a
relatively small amount of steel reinforcement added.
2.Elimination of differential movement. This movement is often the cause of cracks in
masonry.
3. Inherent fire resistance.
4. Reduced maintenance. Periodic painting of exposed steel is eliminated.
5. Lower cost.
This Technical Notes provides a review of the design of RBM beams. Factors influencing
design and performance are reviewed. Design recommendations and aids are provided
and their use illustrated with an example. For additional information about RBM beams
and design calculations, refer to the Masonry Designers' Guide (MDG) [2]. The MDG also
provides an extensive review of the requirements of the Building Code Requirements for
Masonry Structures (ACI 530/ASCE 5/TMS 402-95)[1], hereafter termed the MSJC Code.
Other Technical Notes in this series provide the history of RBM, material and construction
requirements, and design of other RBM elements.
This Technical Notes does not address the design of deep beams (wall beams) or bond
beams. A deep beam is one with a depth-to-span ratio exceeding 0.8. Assumptions made
in this Technical Notes regarding the distribution of stress in beams under flexure and the
loading conditions do not apply to deep beams. Bond beams are formed by placing
horizontal reinforcement in a wall without an opening underneath.
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
1 of 14 9/13/2009 12:52 PM
NOTATION
Following are notations used in the text, figures, and table in this Technical Notes.
A
v
= Area of shear reinforcement, in.
2
(mm
2
)
b = Length of bearing plate, ft (m)
d = Effective depth of beam, in. (mm)
d
b
= Nominal diameter of reinforcement, in. (mm)
Fs = Allowable steel stress, psi (MPa)
f
'
m = Specified compressive strength of masonry, psi (MPa)
H = Height of beam, in. (mm)
l
d
= Embedment length of reinforcement, in. (mm)
M
G
= Design moment due to gravity loads, in.-lb (N-m)
M
s
= Design moment due to in-plane shear, in.-lb (N-m)
M
w
= Design moment due to out-of-plane wind or seismic load, in.-lb (N-m)
P = Design concentrated load, lb (kg)
s = Spacing of shear reinforcement, in. (mm)
V = Design shear force, lb (kg)
W = Width of beam, in. (mm)
wp
= Design uniform distributed load, lb/ft (kg/m)
y = Distance from top of beam to bearing plate, ft (m)
DETERMINATION OF LOADING
The basic concept of a beam is as a pure flexural member. A flexural member spans an
opening and transfers vertical gravity loads to its supports, as illustrated in Fig. 2(a). RBM
beams act in this manner to support their own weight and other applied gravity loads.
However, it is also common for RBM beams to be part of a masonry wall. As such, RBM
beams are often subjected to out-of-plane wind and seismic forces, as depicted in Fig.
2(b). This causes bending of the RBM beam in the out-of-plane direction, which is often
about the weak axis of the beam. In addition, reinforced masonry walls may be shear-
resisting members, or "shear walls", which are part of the lateral load-resisting system of
a building. In such a structural system, RBM beams may be used as connections between
shear walls or piers, as illustrated in Fig. 2(c). Such beams are called coupling beams
because they "couple" the shear walls or piers. If the relative sizes of the two piers being
coupled are similar, the RBM beam is subject to considerable load when an in-plane shear
force is applied to the wall. This is why damage to masonry shear walls is often
concentrated at coupling beams following an earthquake or high-wind event.
The designer should consider all aspects of loading for an RBM beam. It is difficult to
predict the loading condition that will produce the critical design condition. For example, a
RBM beam that is part of a wall will be subject to a combination of gravity loads and
out-of-plane wind or seismic loads. Many factors influence the loading conditions for RBM
beams.
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
2 of 14 9/13/2009 12:52 PM
Arching Action
Arching action is a property of all masonry walls which are laid in an overlapping bond
pattern. Brick masonry will span, in a step-like manner similar to a corbel, over a wall
opening when laid in running bond pattern. Vertical gravity loads above the openings are
transferred to the wall elements on each side.
This is the reason why sizable holes can be created in masonry walls without causing
collapse. Arching action will occur provided that the following conditions are met:
1.An overlapping bond pattern is used in the masonry surrounding the opening.
2.The masonry above the apex of a 45 degree isosceles triangle above the beam exceeds
12 in. (300 mm).
3.There are no movement joints or adjacent wall openings that hinder the load path of
arching action.
4.The abutments are sufficiently strong and rigid to resist the horizontal thrust due to
arching action.
These concepts are illustrated in Fig. 3.
Provided arching action occurs, the self weight of masonry wall carried by the beam may
be safely assumed as the weight within a triangular area above the beam formed by 45
degree angles, as shown in Fig. 3. The self weight of the wall must be added to the live
and dead loads of floors and roofs which bear on the wall above the opening. If a stack
bond pattern is used, the full area of brick masonry above the wall opening should be
considered in the RBM beam design with no assumption of arching action.
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
3 of 14 9/13/2009 12:52 PM
Conditions for Arching Action
FIG. 3
Concentrated Loads
Loads from beams, girders, trusses and other concentrated loads that frame into the wall
must be applied to the RBM beam in the appropriate manner. Concentrated loads may be
assumed to be distributed over a wall length equal to the base of a trapezoid whose top is
at the point of load application and whose sides make an angle of 60 degrees with the
horizontal. In FIG. 4, the portion of the concentrated load carried by the beam is
distributed over the length indicated as a uniform load. The distributed load, wp, on the
RBM beam is computed by the following equation:
wp = P/(b + 2ytan 30) Eq. 1
where:
wp= design uniform distributed load, 1b/ft (kg/m)
P= design concentrated load, 1b (kg)
b= length of bearing plate, ft (m)
y= distance from top of beam to bearing plate, ft (m)
This is approximately 0.866 times P divided by y. Because the apex of the 45 degree
triangle is above the top of the wall in this example, the RBM beam should be designed
assuming no arching action occurs.
The designer should check the stress condition at bearing points for RBM beams. This
applies to loads on the beam and to the beam's reaction on the wall. The MSJC Code
limits the bearing stress to 0.25 f
'
m, where f
'
m is the specified compressive strength of
masonry. A rule-of-thumb recommended for many years is to provide a minimum of 4 in.
(100 mm) of bearing length for masonry beams. The masonry directly beneath a bearing
point should be constructed with solid brick or with solidly grouted hollow brick.
Concentrated loads should not bear directly on ungrouted hollow brick masonry because
of the potential for localized cracking or crushing of the face shells.
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
4 of 14 9/13/2009 12:52 PM
Loads on RBM Beam
FIG. 4
Construction Loads
When designing a RBM beam that is prefabricated or built on the ground and lifted into
place, it is important to consider the loads during transport and handling. To address
these loads, the beam may require reinforcement at both the top and bottom of the beam.
Beams built in place are constructed on shores. These must be designed for the dead
weight of the beam plus any superimposed load prior to adequate curing of the reinforced
brickwork.
Movement Joints
Movement joints are a necessity in masonry walls to accommodate differential movement
and avoid cracking. It is common to place vertical expansion joints at or near the jamb of
wall openings. In RBM buildings there is a reduced need for expansion joints and such
joints may be spaced farther apart. Refer to Technical Notes 18 Series for a discussion of
the placement of movement joints. The presence of a movement joint near a RBM beam
will influence the loads and support conditions for the beam. For example, a simple
support condition should be assumed since arching action will not occur if a movement joint
is at or near the jamb of the opening. Furthermore, the beam will not act as a coupling
beam between shear walls. This is, in fact, one means of simplifying the design and
function of a RBM beam by eliminating loads due to in-plane shear.
DESIGN OF RBM BEAMS
RBM beam design should not be relegated to "rule-of-thumb" methods or arbitrary
selection of beam configuration and steel reinforcement. In any beam design, a careful
analysis of the loads to be carried and a calculation of the resultant stresses should be
incorporated to provide adequate strength and to prevent excessive cracking and
deflection.
In addition to adequate strength, it is preferred that beams exhibit ductile behavior when
overloaded. If the beam is overloaded, it should deform (deflect) a considerable amount
prior to collapse. Deformation allows redistribution of loads to other members and
provides visual indication that the beam is overloaded. Some building codes stipulate a
maximum reinforcement ratio for RBM beams for this purpose.
Another aspect is the relation between the RBM beam's strength and its cracking moment.
Failure of unreinforced masonry in flexure is brittle, exhibiting sudden cracking and often
collapse. Consequently, a reinforced beam should provide a moment strength in excess of
its cracking moment. The amount of this overstrength is somewhat arbitrary, but a factor
of 1.3 is required by the Uniform Building Code[3]. This means that the moment strength
of a cracked-section, RBM beam should exceed 1.3 times the cracking moment of the
beam. This is not a requirement of the MSJC Code, but is considered good engineering
practice.
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
5 of 14 9/13/2009 12:52 PM
Beam Sizing
In the design of an RBM beam, the required cross-sectional area of masonry is based
primarily on the maximum bending moment. However, there are other factors to consider
when sizing an RBM beam. For example, it is often desirable to have the width of the
RBM beam coincide with the specified wall thickness. RBM beams are sometimes formed
with special U-shaped, hollow brick for this reason. These brick may be manufactured
specially for this purpose or they may be cut from full-size units at the site. Manufactured
special shapes may not be readily available in many localities, so it is best to contact the
brick manufacturer as early as possible before proceeding with a design based on their
use. The beam's depth will be determined by the appropriate number of courses of
masonry units present. The beam's depth should be taken as only those courses of solid
brick or that are solidly grouted. The beam's depth may be limited by the height of the
wall above an opening. In such cases, compression steel may be necessary when
sufficient masonry area is not provided.
Lateral Bracing
With short spans and relatively deep beams, there is little likelihood of excessive cracking,
deflection or rotation. This may not be the case, however, for beams that are relatively
long span, shallow or highly loaded. Such beams may be vulnerable to lateral torsional
buckling. The designer should consider the lateral bracing conditions to ensure that the
beam is laterally braced. The MSJC Code requires that the compression face of beams
be laterally supported at a maximum spacing of 32 times the beam thickness. A brick
veneer wall is laterally braced by wall ties to the backup system. A RBM beam that is
part of a load-bearing wall system may not be laterally braced along its span length. In
addition, movement joints at the jambs of a wall opening may result in a lack of lateral
bracing for the beam at its supports. In such cases, attachment of the wall to the floor or
roof diaphragm is the common means of providing lateral bracing for the beam.
RBM Arches

Design of RBM arches should begin with an analysis assuming the arch is unreinforced, in
accordance with Technical Notes 31A or the ARCH computer program available from the
Brick Industry Association. Such an analysis will indicate the locations of highest moment
and shear, and the horizontal thrust at the abutments. Should the analysis so indicate, the
arch should be designed as a reinforced beam. Further, if the conditions shown in Fig. 3
are not met, or if movement joints are provided at the abutments so that the arch may
spread under load, the arch should be designed as if it were a straight, simply supported
beam as a conservative measure. Alternately, a finite element analysis of the arch may be
conducted to determine design moment, shear, and thrust values.
RBM arches cause both a vertical bearing stress and a horizontal thrust on their
abutments. The designer has the option of resisting the horizontal thrust of the arch by the
abutments or providing room for movement as the RBM arch deforms under load.
Judicious placement of vertical expansion joints and flashing will permit horizontal
movement and simplify the arch design. This is recommended for longer span arches
because providing adequate thrust resistance is difficult and movement joint spacing is
limited. In this case, it is very important to provide adequate bearing at the abutments.
STEEL REINFORCEMENT AND TIES
The quantity of reinforcement required for an RBM beam is typically determined by the
applied loads. However, the applicable building code may prescribe a minimum amount of
reinforcement and this may dictate the amount of reinforcement required in a RBM beam.
For example, all building codes now stipulate a minimum amount of reinforcement for
masonry members in areas prone to earthquakes. Some building codes require that
reinforcement in masonry coupling beams be uniformly distributed throughout the beam's
height. This may require additional reinforcement and grouting of the masonry above wall
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
6 of 14 9/13/2009 12:52 PM
openings in RBM beams.
Bond and Hooks

Typically, reinforcement is inserted in masonry beams to resist tension. The tension must
be transferred from the masonry to the reinforcement. This is achieved through adequate
bond between the steel reinforcement and the masonry. The bond stress along the length
of the reinforcement should not exceed an allowable bond stress of 160 psi (1.1 MPa),
according to the MSJC Code Commentary. A minimum embedment length must be
provided in order to not exceed this bond stress. Consequently, the MSJC Code
stipulates a required bond length for reinforcement in tension, called the minimum
embedment length. The minimum embedment length is computed by the following
equation:
ld = 0.0015dbFs Eq. 2
where:
ld= embedment length of reinforcement, in. (mm)
db= nominal diameter of reinforcement, in. (mm)
Fs= allowable steel stress, psi (MPa)
Table 1 provides the minimum development lengths for various bar and wire sizes, based
on Grade 60 ksi (414 MPa) reinforcing bars and 70 ksi (483 MPa) steel wire.
The ends of reinforcing bars and wires may require a standard hook to properly secure
the reinforcement and to achieve its strength. In simply-supported beams, the peak
moment is often at midspan. For this case, the reinforcement in RBM beams can likely be
developed by the bond between the bar or wire and the surrounding masonry with no need
for hooks at the ends of the beam. However, a cantilever RBM beam may require a hook
at the support end. In addition, shear reinforcement should always be terminated with a
hook. Standard hooks for principal reinforcement may be either a 90 degree or 180
degree turn. Often, the designated space for grout and reinforcement in RBM beams is
very small. It can be difficult for a contractor to execute a reinforcement detail properly.
Consider that a 180 degree hook doubles the number of bars at a given cross section.
The designer should always consider the reinforcement placement, tolerances, and cover
restrictions stated in the building codes. Technical Notes 17A Revised provides further
information on bar sizes, placement requirements and construction tolerances.
Shear Reinforcement
Where shear reinforcement is required, it should be spaced so that every potential crack is
crossed by shear reinforcement. Shear cracks are assumed to be oriented at a 45
degree angle to the longitudinal axis of the RBM beam. This restricts the spacing of shear
reinforcement to one-half the beam's effective depth, d. The spacing of shear
reinforcement may be computed by the following equation:
s = AvFsd/V Eq. 3
where:
s= spacing of shear reinforcement, in. (mm)
Av= area of shear reinforcement, in.
2
(mm
2
)
Fs= allowable stress for shear reinforcement, psi (MPa)
d= effective depth of beam, in. (mm)
V= design shear force, 1b (kg)
When shear reinforcement is required, it should be designed to resist the entire shear
force. Shear reinforcement should always be placed parallel to the shear force. For RBM
beams the shear reinforcement should be placed vertically. It can be difficult to provide
shear reinforcement in RBM beams due to the limited size of grout spaces. This is
especially the case with hollow brick units 6 in. (150 mm) or less in thickness and grout
spaces between wythes less than approximately 2 in. (50 mm) in width. Consequently, it
may be advantageous to increase the beam's depth so that shear reinforcement is not
necessary. In fact, this is often the method used by designers to determine the minimum
depth of a RBM beam required for a given loading.
Ties
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
7 of 14 9/13/2009 12:52 PM
There are two instances when it may be necessary to include ties in reinforced brick
beams. These instances occur only when the beam is formed by grouting between
wythes. If the beam has sufficient depth, ties may be required between the wythes. The
grout exerts a hydrostatic pressure that must be resisted during construction. The MSJC
requires wall ties between wythes as follows:
Wire size W1.7 (3.8 mm), one tie per 2 2/3 ft
2
(0.25 m
2
)
Wire size W2.8 (4.8 mm), one tie per 4 1/2 ft
2
(0.42 m
2
)
Maximum spacing of 36 in. (914 mm) horizontally
and 24 in. (610 mm) vertically
Rectangular or Z ties may be used.
In beams that form deep soffits (large beam widths) it may be advisable to tie the soffit
brickwork to the grout. Although the grout does bond to the brick, the metal ties should
provide additional capacity and safety. Such ties are placed in the mortar joint and extend
into the grout.

DEFLECTION

Deflection of RBM beams is considered a serviceability issue. Excessive deflection might
cause damage to interior finishes, functional problems with doors or windows, and
cracking of masonry supported by the beam. The MSJC Code requires that the deflection
of RBM beams that support unreinforced or empirically-designed masonry should not
exceed the lesser of 0.3 in. (7.6 mm) or span length divided by 600. Deflection of RBM
beams may be computed based on uncracked or cracked section properties. Use of
uncracked sections results in underestimating the deflection. Deflection based on cracked
sections only are over-estimated and are more difficult to calculate. Use of uncracked
section is recommended.
Creep is a time-dependent property of brick masonry that will cause the deflection of RBM
beams to increase over time. An accurate formula for the estimation of long-term
deflections of RBM beams due to creep, that is applicable for all cases and easy to use,
does not currently exist. A rule-of-thumb is that the long-term deflection of RBM beams
due to creep will be approximately 50 percent greater than their instantaneous deflection.
This means that a beam that deflects 1.0 in. (25 mm) when it is fully loaded will creep over
time such that its final deflection will be approximately 1.5 in. (38 mm).
DESIGN CURVES
Maximum efficiency and safety dictate the need for a rational design of all RBM beams
according to the applicable building code. However, it is often helpful for the designer to
have design aids that can be used to quickly develop a preliminary beam design. The
design curves in Figs. 5-9 are provided for that purpose. The size and configuration of
masonry and quantity of reinforcement can be quickly determined from these curves based
on the span of the beam and the uniform gravity load supported by the beam, including the
beam's self-weight. The curves are based on the following assumptions:
1.Compressive strength of masonry is not less than 2000 psi (14 MPa). For most brick
masonry, this value will be exceeded. This value was chosen so that beam capacity was
not limited by the masonry's compressive strength.
2.Elastic modulus of masonry is not less than 1600 ksi (11030 MPa).
3.The beam is simply supported and subject to uniform gravity loads only.
4.No compression or shear reinforcement is provided.
5.Deflection is calculated on uncracked section properties. The deflection limit of span
length divided by 600 does not govern for span lengths less than 14 ft. (4.3 m).
The effective depth, d, reflected in the design curves is based on the beam height, H,
minus a value for masonry cover. The cover value is based on a reasonable
approximation of brick, mortar and grout cover on the underside of reinforcement for the
beams shown. The actual effective depth should always be checked for each particular
RBM beam configuration.
DESIGN EXAMPLE
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
8 of 14 9/13/2009 12:52 PM
To illustrate the use of the Design Curves, consider the following example. A RBM beam
is to span over a garage door with a clear span of 9 ft (2.7 m). The beam supports its
own weight and the weight of the brick masonry wall above the beam, so that the uniform
load on the beam is 250 lbs/ft (372 kg/m) of span. The RBM beam and the wall above the
beam are nominal 6 in. (150 mm) wide and constructed with hollow brick. Determine the
beam depth and reinforcement required for these conditions. From Figs. 5(b) and 5(e),
one concludes that a 4 in. (100 mm) or 8 in. (200 mm) high by 6 in. (150 mm) wide RBM
beam is not adequate for the given span and loading. Therefore, the applicable Design
Curve is Fig. 6(b), which is for a full unit depth, RBM beam. For the given conditions, a
minimum depth of 12 in. (300 mm) and one No. 4 bar are required. At this point, any
deflection criteria should be considered and may require a greater beam depth.
SUMMARY
RBM beams are an attractive and efficient means of spanning openings. Attention to
detailing of reinforcement and proper design are the key aspects addressed in this
Technical Notes. The most common RBM beam configurations are shown with
consideration of the inter-connection of beam and wall elements. Design curves provided
in this Technical Notes can be used to develop preliminary beam designs for many
different applications and loading conditions.
The information and suggestions contained in this Technical Notes are based on the
available data and the experience of the engineering staff of the Brick Industry
Association. The information contained herein must be used in conjunction with good
technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the
purview of the Brick Industry Association and must rest with the project architect, engineer
and owner.
REFERENCES

1.Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402-95),
American Society of Civil Engineers, Reston, VA, 1996.
2.Masonry Designers' Guide, John Matthys, ed., The Masonry Society, Boulder, CO,
1993.
3.Uniform Building Code, 1997 Edition, International Conference of Building Officials,
Whittier, CA, 1997.
<
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
9 of 14 9/13/2009 12:52 PM
Design Curves for Partial Soldier Course Beams
FIG. 5

Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
10 of 14 9/13/2009 12:52 PM
Design Curves for Soldier Course Beams
FIG. 6

Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
11 of 14 9/13/2009 12:52 PM
Design Curves for 12 in. (305 mm) Wide Beams
FIG.7
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
12 of 14 9/13/2009 12:52 PM
Design Curves for 16 in. (406 mm) Wide Beams
FIG.8
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
13 of 14 9/13/2009 12:52 PM
Design Curves for 24 in. (610 mm) Wide Beams
FIG.9
Technical Note 17 http://www.gobrick.com/BIA/technotes/t17b.htm
14 of 14 9/13/2009 12:52 PM

Technical Notes 17L - Four Inch RBM Curtain and Panel Walls
Rev [Feb./Mar. 1973] (Reissued September 1988)
INTRODUCTION
Building Code Requirements for Engineered Brick Masonry, SCPI (BIA), August, 1969 defines a curtain wall as "an
exterior non-loadbearing wall not wholly supported at each story. Such walls may be anchored to columns, spandrel
beams, floors or bearing walls, but not necessarily built between structural elements." It further defines a panel wall
as "an exterior non-loadbearing wall wholly supported at each story". Curtain walls must be capable of supporting
their own weight for the height of the wall. Panel walls are required to be self-supporting between stories. Both walls
resist lateral forces such as wind pressures and must transfer these forces to adjacent structural members. This
Technical Notes presents the design and construction of 4-in. brick masonry curtain and panel walls which are
considered to span horizontally in resisting lateral forces.
Recent structural research and rational design methods have greatly aided the design of 4-in. brick walls. However,
exterior brick walls, 4 1/2 in. thick, were used prior to 1900 in the Church of St. Jean de Montmarte, Paris, France.
Designed by M. A. deBoudot, these walls were reinforced with vertical wires through holes in the brick and horizontal
wires in the mortar joints. One wall has an unsupported length of 29 ft 6 in. for the full height of 115 ft. Another wall
is approximately 38 ft high with an unsupported length of 65 ft. The walls discussed herein differ from those
designed by deBoudot in that they have a nominal thickness of 4 in. and contain only horizontal reinforcement.
In the United States, one of the first buildings to be constructed with 4-in. reinforced brick masonry walls was
completed in June, 1932. This building, a compressor house, was constructed at Wood River, Illinois, in the refinery
of the Standard Oil Company (Indiana). All the walls and columns of this building were of reinforced brick masonry.
The walls, with the exception of two bays, were nominally 4 in. thick. The wall height was 18 ft 6 in. with a (column)
spacing of 15 ft 4 in. The horizontal mortar joint above the first course of brick was reinforced with two 3/8-in. round
bars, placed about 1/2 in. from each face of the wall. Above this, the reinforcement was placed in every third
horizontal mortar joint, at a vertical spacing of about 9 5/8 in.
Two more recent examples of the 4-in. curtain wall located in the Denver, Colorado area are the Lakewood Brick
Company building (Fig. 1) and the Kistler Printing Plant (Fig. 2). The Lakewood walls are 22 ft high and span
continuously across steel and masonry composite columns spaced 20 ft on centers. The horizontal reinforcement
varied from two No. 2 bars every second course at points of maximum moment to two No. 2 bars every fourth
course at points of minimum moment. The Kistler Plant has similar spans, heights, and construction.


FIG. 1

17l http://www.gobrick.com/BIA/technotes/t17l.htm
1 of 11 9/13/2009 12:53 PM

FIG. 2

The application of the 4-in. wall in commercial and industrial buildings utilizes the many intrinsic properties normally
found in brick masonry. Brick presents a pleasing appearance without the problem of maintenance. Four-inch walls
built with face brick exhibit the following physical properties without insulation or plastering:
1. U value of 0.76 BTU per sq ft per hr per deg F.
2. Sound transmission class of 45.
3. Fire resistance rating of 1 hr.
4. Average weight of 40 psf.
Structural properties of 4-in. brick walls without reinforcement may be found in Research Report No 9,
"Compressive, Transverse, and Racking Strength Tests of Four-Inch Brick Walls," SCPI, August' 1965.

DESIGN CONSIDERATIONS

Major factors to be considered in the design of 4-in. curtain or panel walls are (1) Structural, (2) Moisture Control,
(3) Differential Movement, and (4) Special Considerations. The walls presented herein are designed in accordance
with the SCPI Standard, Building Code Requirements for Engineered Brick Masonry, August, 1969, and the
materials used should conform to the requirements for reinforced brick masonry as contained in this Standard.
Structural. Since the 4-in. walls do not meet the requirements specified in the SCPI Standard for reinforced walls,
they are designed as "partially reinforced", see Section 4.7.1 1. The horizontal reinforcement in the wall resists the
tensile stresses resulting from lateral pressures as the wall spans horizontally between structural elements such as
columns, pilasters or cross walls.
Since the lateral load due to wind can be either pressure or suction, the wall must be designed to resist the assumed
lateral load acting in either direction. The maximum moment in the wall is used to calculate the required area of
reinforcement, and to provide for stress reversals. This area of reinforcement is required in each wall face. It is
recommended that the maximum area of reinforcement be provided throughout the length of the wall. This
arrangement will provide more reinforcement than needed in portions of the wall near supports where the moment is
less than maximum. However, the savings in fabrication and placement time will, in most cases, offset the added
expense of the excess steel. Steel reinforcement should be lapped a minimum of 16 in. to insure the development of
17l http://www.gobrick.com/BIA/technotes/t17l.htm
2 of 11 9/13/2009 12:53 PM
the tensile stresses without exceeding the allowable bond stresses. Bond and shear stresses should also be
checked at points of maximum shear which would normally occur at the supports.
However, due to the large span-to-depth ratios employed, the bending stresses usually govern.
Where openings such as windows or doors occur, the effective section of the wall is, of course, reduced. This
reduced section will affect not only the required area of reinforcement and strength of masonry but by reducing the
stiffness of the wall will also affect the distribution of lateral loads in continuous walls. Since the tables contained
herein are based on solid walls, walls with openings must be investigated by the designer.
Figure 3 shows a typical application of a 4-in. curtain wall under construction. Care must be taken to assure proper
placement of the reinforcement during construction. The bars may be spread apart or pushed together as
successive brick are laid and pressed into position. Ladder or truss-type joint reinforcement provides two longitudinal
bars connected by cross wires, thus aiding proper positioning and reducing the time and effort involved in placement.
Care should be taken by the mason to insure that the reinforcement has mortar coverage both top and bottom as
well as proper horizontal placement within the wall. The reinforcement should not be laid directly on top of the brick.
Proper vertical placement can be attained by supporting reinforcement wires on small pads of mortar placed prior to
full bedjoint mortar placement.
The proposed usage of the building will determine the permissible deflection for the 4-in. brick wall. For the loads
and spans presented in Table 2, the maximum calculated horizontal deflection is on the order of L/200 for simply
supported walls. The actual maximum deflection will generally be less due to the restraint which may exist at the top
and bottom of the wall.
The general formula for the maximum deflection due to a uniformly distributed load on a simply supported beam is:


FIG. 3
The "Progress Report of the ASCE on Reinforced Masonry Design and Practice", Journal of the Structural Division,
Proceedings of the ASCE, Vol. 87, No. ST8, December, 1961, contains the following recommendation: "When
checking for deflection, it is recommended that the moment of inertia of the masonry cross section be computed
neglecting the effect of the reinforcement, and then use the standard deflection formulae for flexural members "
Moisture Control. The control of moisture through the 4-in. wall should be considered by the designer. Heavy rains
driven by high winds may result in water penetrating the wall. If this is objectionable, one method of handling the
problem is to provide drainage space on the inside of the wall to permit the water to flow to the flashing at the base
and to be conducted back to the outside through weep holes. Drainage wall types (Fig. 4) employ this principle. A
17l http://www.gobrick.com/BIA/technotes/t17l.htm
3 of 11 9/13/2009 12:53 PM
second method is to provide a water barrier (parging) to the inside of the wall. Barrier wall types (Fig. 5) employ this
principle. It is recommended that, where a maximum resistance to rain penetration is desired in areas of severe
exposure, drainage wall type designs be used. Barrier wall types may be sufficient in areas of moderate exposure.


Drainage Type Wall
FIG. 4


17l http://www.gobrick.com/BIA/technotes/t17l.htm
4 of 11 9/13/2009 12:53 PM

Barrier Type Wall
FIG. 5


Differential Movement. When masonry walls are used to enclose a structural frame building, consideration must be
given to the method of anchoring the walls to the framing elements in a manner which will permit each to move
relative to the other. The frame and the enclosing walls differ in their reactions to moisture in the magnitude of
thermal movements. These and other causes for differential movement are discussed in greater detail in Technical
Notes 18.

No single recommendation on the positioning and spacing of expansion joints can be applicable to all structures.
Each building design should be analyzed to determine the potential movements and provisions should be made to
relieve excessive stress which might be expected to result from such movement. Generally, expansion joints may be
located at or near corners, offsets, junctures and openings. Since 4-in. walls are considered to span horizontally,
expansion joints where required must be located at vertical supports. The maximum recommended spacing for
expansion joints for these 4-in. walls is 100 ft for a straight wall without openings. However, conditions may require
expansion joints as close together as 40 ft. Typical 4-in. wall details relating to expansion joints and methods of
anchoring are shown in Fig. 6. Table 1 presents the maximum vertical spacing of No. 6 gage, galvanized wire
anchors at the wall-to-support connections. The load distribution per anchor was assumed to be 600 lb. Maximum
spacing was limited to 24 in.

17l http://www.gobrick.com/BIA/technotes/t17l.htm
5 of 11 9/13/2009 12:53 PM

Typical Four Inch Wall Details
FIG. 6

(a) No. 6 gage galvanized wire anchors.
(b) Load distribution assumed to be 600 lb per anchor
(c) Maximum vertical spacing limited to 24 in.
(d) Spacing is for an interior support of a continuous wall. Spacing at an end support is 24 in.

Special Considerations. There are special design possibilities which the designer of structures incorporating 4-in.
reinforced walls may desire to consider. One such consideration is to design the wall as a plate supported on three
or four sides. Another consideration is to utilize the loadbearing capabilities of the 4-in. wall. The SCPI Standard may
be used to design the wall to carry vertical loads in addition to lateral loads.
17l http://www.gobrick.com/BIA/technotes/t17l.htm
6 of 11 9/13/2009 12:53 PM
The reinforced wall itself may be considered to act as a lintel over openings in the wall such as provided for windows
and doors. The wall reinforcement should be checked to determine if it is adequate for the particular opening. Where
the normal wall reinforcement is insufficient, supplemental steel should be added. Assuming a lintel to carry only the
dead load of the wall for a height equal to one-half the opening span, a 4-in. brick masonry lintel, with an ultimate
strength (f'm) of 2200 psi, an effective depth of 36 in. and two No 2 bars, will span an opening of 12 ft 6 in. This is
neglecting other reinforcement which is in the wall above the opening. Therefore, depending on the actual wall and
opening layout, wider openings may be spanned. Technical Notes 17B provides additional information on reinforced
brick masonry lintels.

DESIGN TABLES
Design Assumptions. As presented herein, the 4-in. curtain wall is analogous to a one-way reinforced slab. In
preparing the tables, a width of cross section "b" of 12 in. was selected. The effective depth "d" was assumed as 2
3/4 in. (Fig. 7). This dimension is based on the minimum thickness for a 4-in. nominal wall (3 1/2 in.) and the
requirement by the SCPI Standard for shut-in. mortar coverage for bars or wire to in. or less in diameter embedded
in the horizontal mortar joints. In order for the effective depth assumptions to remain as stated, raked joints could not
be used. If raked joints are desired on 4-in. reinforced curtain walls, the effective depth must be reduced by the
amount of raking of the joint and steel must be placed to maintain a minimum 5/8-in. coverage. It is not
recommended that raked joints be used on 4-in. curtain walls for the above reason. Additional assumptions are as
follows:
1. For stresses due to wind, the allowable stresses in brick masonry and reinforcing steel are increased by 1/3.
2. A brick masonry compressive strength (f'm) of 2200 psi resulting in an allowable brick masonry stress (fm) equal
to 0.32 (f'm) (1.33) or 936 psi.
3. An allowable steel stress (fs) equal to 20,000 (1.33) or 26,600 psi.
4. Only type M or type S portland cement-lime mortars are to be used.
5. Architectural or engineering inspection is provided during construction. A 1/3 reduction in the allowable stresses is
required if this inspection is not provided.
6. In a simple span or two continuous spans, the maximum shear is equal to 0.625 wL and the maximum moment is
equal to 0.125 wL
2
.
7. In three or more continuous spans, the maximum shear is equal to 0.60 wL and the maximum moment is equal to
0.10 wL
2
.
8. Coursing is assumed to be three brick masonry courses per 8 in. of wall height.
9. Four-inch walls are solid without openings.

17l http://www.gobrick.com/BIA/technotes/t17l.htm
7 of 11 9/13/2009 12:53 PM

Section Through Four Inch Brick Wall
FIG. 7


Use of Tables. Table 2 provides the reinforcement requirements for various support conditions, span lengths and
wind pressures. The steel areas tabulated are for one face of the wall. An equal amount must be provided in the
opposite face. It should be noted that an expansion joint interrupts the continuity of the wall. The condition at the joint
will therefore be one of simple support.

(a) See Design Assumptions

Table 3 gives the area of reinforcement as provided by various steel sizes and vertical spacings for one wall face.
After the required area is selected from Table 2, the bar size and spacing which provides an area of reinforcement
17l http://www.gobrick.com/BIA/technotes/t17l.htm
8 of 11 9/13/2009 12:53 PM
equal to or greater than what is required may be selected from Table 3.

(a) Based upon three brick courses per 8 in. of height.
In conjunction with the use of Table 2, a minimum value of 2200 psi for the compressive strength of brick masonry
(f'm) is required. Lesser values of f'm may be employed if the designer calculates the actual strength of masonry
required by his design. Table 4 can be used in selecting unit strength and mortar type as determined by the
calculated stresses
(a) Taken from Table 2, Building Code Requirements for Engineered Brick Masonry, SCPI (BIA), August, 1969.
(b) See Design Assumptions.

Design Example
Given: Three continuous 20-ft spans

Wind = 25 psf
f'm = 2200 psi
fm = 0.32(f'm)(1.33) = 936 psi
fs = 20,000(1.33) = 26,600 psi

17l http://www.gobrick.com/BIA/technotes/t17l.htm
9 of 11 9/13/2009 12:53 PM
Required: As

1. Design by calculation:
V(max) 0.60 wL = 0.60(25)(20) = 300 lb
M(max) 0.10 wL2 = 0.10(25)(20)2 = 12,000 in. -lb


Select one No. 2 bar (each face, every course):
As (each face) = 0.225 sq in. (Table 3)

Check masonry stress (fm):

Check steel stress (fs):


17l http://www.gobrick.com/BIA/technotes/t17l.htm
10 of 11 9/13/2009 12:53 PM
Check shear at interior support:


Check bond:


2. Selection from Table 2:
As = 0.188 in. 2 per face

Selection from Table 3:
One No. 2 bar in each wall face every course.
17l http://www.gobrick.com/BIA/technotes/t17l.htm
11 of 11 9/13/2009 12:53 PM

Technical Notes 17M - Reinforced Brick Masonry Girders - Examples
[July 1968] (Reissued September 1988)
INTRODUCTION
"Girder" is the name applied to a large size beam which usually has smaller beams framing into it. A Reinforced
Brick Masonry (RBM) girder consists of brick masonry in which steel reinforcement is embedded so that the
resulting horizontal member is capable of resisting loads which produce compressive, tensile and shearing stresses.
The principles of design for RBM girders and beams are the same as those commonly accepted for the working
stress design for reinforced concrete flexural members, and similar formulae may be used. These formulae may be
found in Technical Notes 17A, "Reinforced Brick Masonry Flexural Design", Technical Notes 17J, "Design Tables For
Reinforced Brick Masonry Flexural Members", may be consulted for design tables and illustrative examples.
APPLICATIONS
Where a design requires horizontal structural members, the RBM girder or beam is most advantageous on projects
where (1) the structural medium is brick, (2) the surrounding areas are brick, or (3) the appearance of brick is
desired. The main advantage of RBM girders is that the structural finish and the architectural finish are one and the
same. In some cases, however, they provide economical solutions without considering the savings due to a built-in
finish.
Additional advantages are that no forms are required for the erection of a RBM girder except for a work scaffold set
at the bottom elevation of the girder. Brick girders are inherently fire-resistant, weather-resistant and
maintenance-free. These properties, in addition to the high compressive strength of brick masonry, constitute
savings in construction time and costs.
The examples of RBM girders presented in this Technical Notes are representative of some uses for such
members. They are by no means exhaustive.
St. Hedwig's Church. Architect: J.T. Golabowski; Structural Engineer: William A. Herrmann. The design of St.
Hedwig's Church (Fig. 1 ) in St. Louis, Missouri, built in 1957, incorporated a modified clerestory that extends the
length of the church from front to rear (Fig. 2). In order to keep the nave and sanctuary clear of interior columns, the
sidewalls of the clerestory were supported only at the ends. Exposed brick were used extensively both outside and
inside the church. Structural members, 65 ft long, were required that would have a brick finish on both sides, and
carry the roof load from above and part of the loads from the side roofs.
St. Hedwig's Church - St. Louis, Missouri
http://www.gobrick.com/BIA/technotes/t17m.htm
1 of 8 9/13/2009 12:54 PM
FIG. 1
Plan of St. Hedwig's Church St. Louis Missouri
FIG. 2
Three types of girders were considered:
1. Prestressed concrete girders. These were eliminated because, with masonry on both sides, they would be
too thick for the desired appearance.
2. Brick-encased steel plate girders. These would have been end-supported on steel columns built into the
masonry walls.
3. RBM girders. These would be supported on RBM columns built integrally with the front and rear masonry
walls.
At that time, contractors in the St. Louis area were not familiar with RBM construction. Due to this, alternate bids
were taken on both the brick-encased plate girder and RBM girder systems. All ten bidders indicated savings would
be made with the RBM system.
The church contractor, C. Rallo Contracting Company, Inc., reported that the total cost of the two RBM girders and
supporting columns was only $8000; columns alone cost $600. The bid figures show two steel plate girders and four
steel columns would have cost $8200 without the brick encasing.
Figure 3 gives the overall dimensions of the RBM girder. The depth was dictated by the clerestory proportions.
During the construction of the two girders, the only form needed was a work scaffold set to the bottom elevation of
the girder. Soffit brick were laid out first (Fig. 4) with the bottom steel being placed next. Stirrups were then added
and tied near the tops for stability. After that, it was just a matter of laying brick and placing grout in increments of
three or four courses (Fig. 5).
A building permit was issued for the job with the provision that one of the girders be successfully load-tested after
completion. The specification required a test load equal to the design dead load plus twice the live load, in
accordance with Building Code Requirements For Reinforced Concrete of the American Concrete Institute.
http://www.gobrick.com/BIA/technotes/t17m.htm
2 of 8 9/13/2009 12:54 PM
Section of RBM Girder - St. Hedwig's Church
FIG. 3
Soffit Brick in Place on Scaffold
St. Hedwig's Church
http://www.gobrick.com/BIA/technotes/t17m.htm
3 of 8 9/13/2009 12:54 PM
FIG. 4
Completed RBM Girder - St. Hedwig's Church
FIG. 5
Test loads of 67,500 lb were applied at each third point of the girder. This was in addition to the dead load of both
the girder and the concrete roofs it supports. Pittsburgh Testing Laboratory performed the tests with calibrated
hydraulic jacks through a fulcrum system.
All deflection measurements checked out within the allowable limits. After 24 hr under full load, the maximum girder
deflection was only half the amount allowed by the test provisions. Recovery after load release was 94 percent.
Maryland City Shopping Center. Architect: Anthony F. Musolino & Associate; Structural Engineer: Lugi Iacono. A
project in Maryland City, Maryland which extensively uses RBM girders is the Maryland City Shopping Center now
nearing completion. The architect is well pleased with the use of RBM girders as handsome structural facades (Fig.
6). Details of two girders are shown in Figs. 7 and 8. The RBM girder in Fig. 7 is supported on RBM columns
spaced 23 ft 4 in. on center. It carries half of the arcade roof load transmitted to it by a beam located 6 ft 8 in. from
one support. Other girders are loaded similarly by beams at varying distances from the girder support.
RBM Girder as Structural Facade - Maryland City Shopping Center
FIG. 6
http://www.gobrick.com/BIA/technotes/t17m.htm
4 of 8 9/13/2009 12:54 PM
Section of RBM Girder Maryland City Shopping Center
FIG. 7
Another girder is utilized above a store front which is glazed for display purposes (Fig. 8). This girder supports a roof
area of 500 sq ft carried by bar joists 4 ft 2 in. on center in addition to a 4-ft masonry parapet 8 in. thick. The column
spacing for the girder is 25 ft.
Section of RBM Girder Maryland City Shopping Center
FIG. 8

Mt. Vernon Shopping Center. Architect: Anthony F. Musolino & Associate; Structural Engineer: Peter Dragan. The
architectural beauty of repeating arches with the structural capability of RBM girders is displayed at the Mt. Vernon
Shopping Center, Alexandria, Virginia (Fig. 9). The reinforced arches which form the bottom of the RBM girders are
not intended to be structural. The girders transmit half the covered promenade roof load to their supporting solid
brick masonry columns which are spaced 19 ft 9 in. on center (Figs. 10 and I 1). The girders were constructed of
8000 psi brick units and type M mortar.
http://www.gobrick.com/BIA/technotes/t17m.htm
5 of 8 9/13/2009 12:54 PM
Mt. Vernon Shopping Center - Alexandria, Virginia
FIG. 9
Elevation of RBM Girder and Arches - Mt. Vernon Shopping Center
FIG. 10
Sections of RBM Girder and Arch Mt. Vernon Shopping Center
FIG. 11
Alexander Art Center. Architect: Griffey and Stroll Associates; Structural Engineer: William Blanton. The designers
of the Alexander Art Center at Athens, West Virginia were confronted with the problem of spanning over the main
auditorium proscenium opening. A structural member having a fire-resistant brick finish was desired to support steel
http://www.gobrick.com/BIA/technotes/t17m.htm
6 of 8 9/13/2009 12:54 PM
trusses which, in turn, support the roof loads over a clear span of 39 ft and a monorail system spaced 1 ft on center
connected to the bottom chord of the trusses. The monorails will provide a flexible system for lighting and scenery
movement. The designers felt that deflections in a brick-encased steel girder may cause cracking of the enclosing
masonry. The encasing operation for the steel girder would require scaffolding similar to that used in the construction
of RBM girders. Also, RBM was being utilized on the project in retaining walls and in loadbearing walls. These
factors resulted in the final selection of a RBM girder spanning 50 ft 8 in. to solid brick masonry columns on either
side of the proscenium. The girder section detailed in Fig. 12 is to be constructed of 8000 psi brick units and type M
mortar. The total design load is 5820 lb per 1in ft of girder.
Section of RBM Girder Alexander Art Center
FIG. 12
The Art Center will also include a studio theater in which the 24-ft stage will be spanned by a RBM girder of the
dimensions shown in Fig. 13. The loading was the same as for the larger girder plus an additional load of a
prestressed concrete roof which spans 46 ft 6 in. This results in a total design load of 7064 lb per 1in ft. Figure 14
shows this girder.
http://www.gobrick.com/BIA/technotes/t17m.htm
7 of 8 9/13/2009 12:54 PM
Section of RBM Girder Alexander Art Center
FIG.13

Completed RBM Girder - Alexander Art Center
FIG. 14
http://www.gobrick.com/BIA/technotes/t17m.htm
8 of 8 9/13/2009 12:54 PM
Volume Changes - Analysis and Effects
of Movement
Abstract: This Technical Note describes the various movements that occur within buildings. Movements induced by changes
in temperature, moisture, elastic deformations, creep, and other factors develop stresses if the brickwork is restrained. Restraint
of these movements may result in cracking of the masonry. Typical crack patterns are shown and their causes identified.
Key Words: corrosion, cracks, creep, differential movement, elastic deformation, expansion.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
18
October
2006
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Use the following coefficients to calculate movements of
brick veneer:
- Thermal expansion: 4 x 10
-6
in./in./F
(7.2 x 10
-6
mm/mm/C)
- Moisture expansion: 5 x 10
-4
in./in. (mm/mm)
- Creep: 0.7 x 10
-7
in./in. per psi
(0.1 x 10
-4
mm/mm per MPa)
Consider coefficients of movements for other materials in
contact with brickwork
Consider elastic deformation and movement of structural
elements supporting and connected to brickwork
Page 1 of 8
INTRODUCTION
All building materials change in volume in response to changes in temperature or moisture. Changes in volume,
elastic deformations due to loads, creep and other factors result in movement. Restraint of these movements may
cause stresses within building elements that result in cracks.
To avoid cracks, masonry elements should be designed to minimize movement or accommodate differential move-
ment between materials and assemblies. A system of movement joints can reduce the potential for cracks and the
problems they cause. Movement joints can be designed by estimating the magnitude of the different movements
that occur in masonry and other building materials.
This Technical Note describes volume changes in brick masonry and other building materials. It also describes the
effects of volume change when materials are restrained. Technical Note 18A discusses the design and detailing of
movement joints and the types of anchorage that permit movement.
MOVEMENTS OF CONSTRUCTION MATERIALS
Most buildings do not allow exact prediction of building element movements. Volume changes are dependent on
material properties and are highly variable. The age of the material and temperature at installation also influence
expected movement. When mean values of material properties are used in design, the actual movement may
be underestimated or overestimated. The
designer should use discretion when select-
ing the applicable values. The types of move-
ment experienced by various building materi-
als are indicated in Table 1.
Brickwork will generally increase in size over
its service life. This is the net effect of a vari-
ety of conditions that causes the size of brick-
work to change, but is influenced primarily by
irreversible moisture expansion. Unrestrained
elements or sections of brickwork will expand
vertically from their support and horizontally
from the center as shown in Figure 1.
Rever sible Ir r ever sible Elast ic
Moist ur e Moist ur e Def or mat ion
Brick Masonry ---
Concrete Masonry ---
Concrete ---
Steel --- --- ---
Wood ---
Building Mat er ial Ther mal Cr eep
Table 1
Types of Movement of Building Materials
2006 Brick Industry Association, Reston, Virginia
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 2 of 9
Temperature Movement
All building materials expand and contract with varia-
tions in temperature. For unrestrained conditions,
these movements are theoretically reversible. Table
2 indicates the coefficients of thermal expansion for
brick and other common building materials.
Unrestrained thermal movement is the product of
temperature change multiplied by the coefficient of
thermal expansion and the length of the element.
The stresses developed by restrained thermal move-
ments are equal to the change in temperature multi-
plied by the coefficient of thermal expansion and by
the modulus of elasticity of the material.
The temperature change used for estimating thermal
movements should be based on mean temperatures
in the component. For solid masonry walls, tempera-
tures at the center of the wall should be used. In cav-
ity walls and veneers, the temperature at the center
of each wythe or component should be used. In dis-
continuous construction, the wythes will have different
temperatures due to the separation of the wythes by
an air space and perhaps insulation.
Surface temperatures of brick walls may be much
higher than the ambient air temperature. Wall orienta-
tion, color, brick wall type and presence and location
of insulation are governing factors. It is possible for
a dark, south facing wall to reach surface tempera-
tures as high as 140 F (60 C), while the ambient air
temperature is well below 100 F (38 C). The mean
temperature of a 4 in. (102 mm) thick insulated brick
veneer is very close to the surface temperature of the
brick. A thicker or uninsulated wall may have a lower
mean temperature than the outside surface. The tem-
perature range experienced by brickwork is the differ-
ence of the high and low mean temperatures of the
brickwork. In practice, this range is usually taken as
100 F (38 C) and is based on the annual high and
low temperature of the exterior ambient air.
Other building materials expand and contract at
rates different from that of brick masonry. These
differences are important when elements such as
window frames, railings, or copings are attached to
brick masonry. Distress may occur in either material.
Bowing may occur in composite walls that have con-
crete masonry interior wythes.
Moisture Movement
With the exception of metals, many building materials tend to expand with an increase in moisture content and
contract with a loss of moisture. For some building materials these movements are reversible; while for others
they are irreversible or only partially reversible.
Brick. Brick expand slowly over time upon exposure to water or humid air. This expansion is not reversible by
drying at normal temperatures. A brick is smallest in size when it cools after exiting the kiln. The brick will draw
Brick Support
Direction of
Expansion
Figure 1
Direction of Brick Expansion
x 10
-6
in./in.
per F
x 10
-6

in./in
per C
Brickwork 4.0 7.2
Concrete Masonry 4.5 8.1
Stone
Granite 4.4 7.9
Limestone 4.4 7.9
Marble 7.3 13.1
Concrete 5.5 9.9
Metal
Aluminum 12.8 23.1
Bronze 10.1 18.1
Stainless Steel 9.9 17.8
Structural Steel 6.5 11.7
Wood, Parallel to Fiber
Fir 2.1 3.7
Oak 2.7 4.9
Pine 3.0 5.4
Wood, Perpendicular to Fiber
Fir 32 58
Oak 30 54
Pine 19 34
Autoclaved Aerated Concrete 4.5 8.1
Mat er ial
Design Coef f icient s of
Linear Ther mal Expansion
Table 2
Thermal Expansion
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 3 of 9
moisture from its environment and increase in size from that time. Most of the expansion takes place over the first
few weeks, but will continue at a much lower rate for several years (see Figure 2). The amount of moisture expan-
sion depends primarily on the raw materials and to a lesser extent on the firing temperatures. Brick made from the
same raw materials that are fired at lower temperatures will expand more than those fired at higher temperatures.
In brickwork, moisture expansion of the brick is
somewhat offset by drying shrinkage of the mortar.
As brick with larger face dimensions cover more wall
area, the brickwork will experience more moisture
expansion.
Predicting the total moisture expansion of brickwork
is difficult; however, it can be estimated by multiply-
ing the coefficient of moisture expansion by the
length of the wall. The Building Code Requirements
for Masonry Structures (ACI 530/ASCE 5/TMS 402)
[Ref. 3] design coefficient of linear moisture expan-
sion for brickwork is 3 x 10
-4
in./in. (mm/mm). For
brick veneer, a design coefficient of linear expansion
of 5 x 10
-4
in./in. (mm/mm) is recommended.
Masonry materials expand due to freezing when saturated. Freezing expansion, when it occurs, has a small effect
on total expansion of masonry. Although limited, available data indicates that the coefficient of freezing expansion
for brick ranges from 0 to 10.3 x 10
-4
in./in. (mm/mm). A design value for brick masonry of 2 x 10
-4
in./in. (mm/mm)
is recommended. Freezing expansion is typically negligible, as it only occurs when fully saturated brick are sub-
jected to temperatures at or below 14 F (-10 C).
Concrete Masonry. Concrete masonry units experience shrinkage as a result of moisture loss and carbonation
and will expand as moisture content increases. These combined movements typically result in a net shrinkage
of concrete masonry that is also affected by the method of curing, aggregate type, change in moisture content,
cement content, temperature changes and wetting and drying cycles. The total potential linear drying shrinkage
due to changes in moisture content is determined using ASTM C 426, Test Method for Linear Drying Shrinkage of
Concrete Masonry Units [Ref. 2], which measures unit shrinkage from a saturated condition to a condition of equi-
librium at a relative humidity of 17 percent. Typical linear shrinkage values for concrete masonry units range from
2 x 10
-4
to 4.5 x 10
-4
in./in. (mm/mm). The coefficient of shrinkage for concrete masonry is assumed to be half of
the total linear shrinkage determined by ASTM C 426.
Carbonation is the chemical combination of hydrated portland cement with carbon dioxide present in air. Although
there is currently no standard test method to measure carbonation shrinkage, the National Concrete Masonry
Association recommends a value of 2.5 x 10
-4
in./in. (mm/mm) be used to estimate carbonation shrinkage in con-
crete masonry walls [Ref. 5].
Concrete. Concrete shrinks as it cures or dries and swells when it is wet. Shrinkage of concrete is influenced by
the water-cement ratio, composition of the cement, type of aggregate, size of concrete member, curing conditions,
and amount and distribution of reinforcing steel. Shrinkage values for ordinary concretes are generally range from
2 x 10
-4
to 7 x 10
-4
in./in. (mm/mm) depending on the factors listed.
Wood. Wood shrinks during the natural seasoning process as the moisture content drops from the fiber saturation
point (28 to 30 percent) until it reaches equilibrium with the environment. Shrinkage occurs differently in the radial,
tangential, and longitudinal directions of the wood. The American Softwood Lumber Standard PS 20 [Ref. 1] sug-
gests an average shrinkage value of one percent per each four percent drop in moisture content (a coefficient of
0.0025 in./in. (mm/mm) per percent change in moisture content) for typical softwoods. Longitudinal shrinkage (0.5
x 10
-4
in./in. (mm/mm) per percent change) is usually small enough to be neglected in design. Moisture expansion
and contraction continues with changes in moisture content of the wood.
Elastic Deformation
Elastic deformation is a reversible change in length, volume or shape produced by stress in a material. In the
structural design of a building, the designer must consider all forces imposed on the structure. These include dead
and live loads and such external lateral forces as wind, soil, snow loads, earthquake and blast. Each of these
Figure 2
Projected Moisture Expansion of Fired Brick vs. Time
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 4 of 9
forces creates stresses in the building materials which can result in deformations and deflections of the building
elements. If a material remains within its elastic range, it will return to its original shape once the applied forces
are removed.
There are several types of deformation to consider. Horizontal elements such as beams and lintels deflect verti-
cally due to their own weight and dead and live loads. Lateral deflections of walls and columns and reductions in
lengths of axially loaded structural elements due to design loads must be considered. Walls, beams, columns and
building frames move horizontally from lateral loads such as wind and seismic events. Columns and bearing walls
are shortened in length due to vertical dead and live loads.
Elastic deformation is most important when considering elements that support brickwork. The design of longer
lintels and shelf angles are typically controlled by deflection. Such deflection should be limited or accommodated
by the veneer design or cracking of the veneer may result.
Lateral Drift. The drift or side-sway of a structural frame may cause distress to brick masonry used as in-fill walls
or exterior cladding. Lateral loads from wind or earthquakes are transferred to brickwork if it is attached rigidly to
the frame. The same is true for deflection of floor slabs or spandrel beams. Masonry built in contact with these
elements will be loaded due to the movement of the member. Masonry intended to be non-loadbearing may
become loadbearing.
Creep
Creep, or plastic flow, is the continuing, irreversible deformation of materials under load or stress. The magnitude
of movement due to creep in masonry and concrete depends on the stress level, material age, duration of stress,
material quality, and environmental factors.
In frame structures, especially concrete frame buildings, vertical shortening due to creep or shrinkage of the struc-
tural frame may impose high stresses on the masonry. These stresses develop at window heads, shelf angles,
and other points where stresses are concentrated.
Brick. Creep in brick masonry primarily occurs in the mortar joints and is negligible. ACI 530/ASCE 5/TMS 402
stipulates a design coefficient of creep for clay masonry of 0.7 x 10
-7
in./in. per psi (0.1 x 10
-4
mm/mm per MPa).
Concrete Masonry. Concrete masonry exhibits more creep than brick masonry because of the cement content in
the units. ACI 530/ASCE 5/TMS 402 stipulates a value of 2.5 x 10
-7
in./in. per psi (0.36 x 10
-4
mm/mm per MPa).
Concrete. Creep is most significant in concrete frame structures. Creep in concrete begins after load is applied
and proceeds at a decreasing rate. High-strength concrete experiences less creep than low-strength concrete.
Creep is slightly greater in lightweight aggregate concretes than normal weight concretes. In high-rise buildings,
the total elastic and inelastic shortening of columns and walls due to gravity loads and shrinkage may be as high
as 1 in. (25 mm) for every 80 ft (24 m) of height.
Estimating Combined Movements
Equation 1 below, combines the effects of movements above that affect brickwork, and can be used estimate the
amount of expansion that would be experienced by an unrestrained brick wythe. Although typically neglible, local
conditions must be considered to determine if freezing expansion will occur.
m
u
= (k
e
+ k
f
+k
t
T) L Eq. 1
where:
m
u
=total unrestrained movement of the brickwork, in. (mm)
k
e
=coefficient of moisture expansion, in./in. (mm/mm)
k
f
=coefficient of freezing expansion, in./in./F (mm/mm/C)
k
t
=coefficient of thermal expansion, in./in./F (mm/mm/C)
T =temperature range experienced by brickwork, F (C)
L =length of wall, in. (mm)
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 5 of 9
Using the recommended values given previously for coefficients of expansion and temperature range experienced
by brickwork, the equation becomes:
m
u
= (0.0005 + 0 + (0.000004 x 100))L
m
u
= 0.0009L
In addition to the expansion of brickwork, other movements of building materials decribed herein, restraint condi-
tions, construction tolerances and wall orientation may affect the size and spacing of expansion joints.
Other Causes of Movement
Other causes of movement in building elements that may occur under given conditions include corrosion of steel,
drift of the building frame, and the action of unstable soils. It is beyond the scope of this Technical Note to discuss
these items in detail. However, they are briefly described below.
Corrosion of Steel. Excessive corrosion of steel embedded in masonry can cause cracking or spalling of mason-
ry. The volume of rust is greater than that of the steel from which it is formed. This increase causes pressure on
the surrounding masonry and may result in movement and cracking.
Anchors, ties and joint reinforcement are embedded in mortar and may be exposed in an air space or cavity. Thus
they may be susceptible to corrosion. Metals embedded in grout, such as reinforcing bars, are less susceptible to
corrosion since they are protected by the grout and not exposed. Other items in masonry susceptible to corrosion
are steel lintels, steel shelf angles, anchor bolts and other metal fasteners in masonry. To minimize corrosion, do
not use additives in mortar that accelerate corrosion, such as calcium chloride, and minimize the amount of water
within masonry through proper design, detailing and installation. See Technical Note 44B for more on corrosion
resistance of metal ties and anchors.
Unstable Soils. Unstable or expansive soils often cause movement or differential settlement in foundations that
support brick masonry. Proper foundation design will help ensure stable support and allow uniform settlement
within acceptable limits.
IDENTIFYING EFFECTS OF MOVEMENT
Changes in building materials and technology have affected the design and behavior of many building compo-
nents, including masonry walls. The increased use of thinner walls and the tendency to specify high compressive
strength mortars have become common. Although stronger units and mortars increase the compressive strength
of the masonry, they do so at the expense of other important properties. Thus, masonry walls today tend to be
thinner and more brittle than their massive ancestors. These thinner walls are more susceptible to cracking and
spalling if differential movement is not addressed in design. Technical Note 18A includes recommendations for
accommodating differential movement in new construction. Proper design and construction of brickwork can help
prevent the detrimental effects of movements. The following section demonstrates specific conditions in brickwork
and their underlying causes.
Cracking is perhaps the most frequent type of distress that
affects masonry walls. The shape, location and magnitude of
cracking will often indicate the cause. Conditions that occur
when movement is not accommodated are illustrated in the
following photographs. Technical Note 18A recommends
details that help prevent these conditions.
Long Walls. When expansion joints are too narrow or
spaced too far apart, the expansion of the brickwork may not
be adequately accommodated. This may force sealant mate-
rial out of an expansion joint as shown in Photo 1. If expan-
sion continues, then cracking occurs at other locations. In
walls with openings, diagonal cracks may occur in brickwork
between windows or doors. Such cracks usually extend from
the head or sill at the jamb of the opening, depending upon
the direction of movement and the path of least resistance.
Because the effects of expansion are cumulative, dividing
Photo 1
Sealant Forced Out of Expansion Joint Due
to Expansion of Long Wall
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 6 of 9
long walls into smaller segments reduces the amount of
movement that the expansion joint has to accommodate.
Corners. Brickwork will expand in the plane of the wall. At
a corner, the brickwork on each side will expand toward
the corner. Absence of an expansion joint near a corner or
an insufficient number of expansion joints in the wall can
result in cracking at the corner as shown in Photo 2. This
typically occurs at the first head joint on one side of the
corner.
Offsets and Setbacks. When parallel walls expand
toward an offset without an expansion joint, the movement
may produce rotation of the offset and vertical cracks as
shown in Photos 3 and 4.
Structural Frame Concerns. The brick veneer in Photo
5 is supported by a steel shelf angle on a concrete frame.
Over time, creep and shrinkage of the concrete frame
along with expansion of the brickwork has caused the
steel shelf angle to bear on the masonry below. Brickwork
between floors can bow if it is not adequately attached to
the backing, or the backing is not sufficiently rigid.
Steel frames typically have larger drifts and deflections
than concrete frames. This movement generally becomes
evident at shelf angles and may result in spalling if not
accommodated. An expansion joint below each shelf angle
alleviates this concern.
Movement of structural elements rigidly attached to
masonry is transferred to the masonry and may cause
cracks. These movements may be due to drift of the build-
ing frame or lateral expansion from creep. These cracks
may occur on the exterior as well as the interior of the
building. Space between the structural member and the
brickwork and use of flexible anchors will reduce the likeli-
hood of such cracking.
Photo 4
Rotation at Offset Without Expansion Joint
Photo 5
Spalling in Brickwork Without Horizontal
Expansion Joint Due to Shortening
of Structural Frame
Photo 2
Crack at Corner Without Expansion Joint
Photo 3
Crack at Offset Without Expansion Joint
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 7 of 9
Parapet Walls. Parapets are exposed to the elements
on three sides, as opposed to most walls which are
exposed on only one side. As a result, parapet walls
are subjected to extremes of moisture and temperature
which may be substantially different from those in the
wall below. Parapets also lack the dead load of overly-
ing masonry to help resist movement. Expansion can
cause parapets to bow if restrained at both corners
and move away from corners if restrained only at one
end as shown in Photo 6.
Foundations. Cracking of concrete foundations, as
shown in Photo 7 or movement of above grade brick-
work away from the foundation corner is often the
result of shear stress at the interface between the
brick and concrete. Because brick walls expand and
concrete foundations shrink, differential movement will
cause shearing stresses to develop when these mate-
rials are bonded together. A bond break or flashing
placed between the concrete and brickwork will permit
movement to occur.
Deflection and Settlement. Deflection and settlement
cracks are identified by a tapering shape. Photo 8
shows a deflection crack caused by supporting brick-
work on an undersized lintel. The crack is wider at
the steel angle and tapers to nothing. Technical Note
31B details the proper design of steel lintels support-
ing masonry. Deflection cracks may also occur at steel
shelf angles attached to spandrel beams that deflect.
Differential settlement may cause cracking when one
portion of a structure settles more than an adjacent
part, as shown in Photo 9.
Curling of Concrete. Masonry that is supported by
or bonded to cast-in-place concrete slabs may crack if
curling of the slab lifts the adjacent masonry. In some
cases cracking of the brickwork can be prevented
by separating it from the concrete slab with a bond
break. Curling of concrete is most often the result of
Photo 8
Crack Due to Deflection of Undersized Lintel
Photo 9
Crack Due to Differential Settlement
Photo 6
Movement of Parapet Away from Corner
Photo 7
Crack at Foundation Due to Lack of Bond
Break Between Brickwork and Concrete
slab deflection and differences in moisture or temperature between the top and bottom of the slab. The American
Concrete Institute or other concrete industry organizations should be consulted for recommended practices that
minimize slab curling.
Embedded Items. Items embedded in or attached to
masonry may cause spalling or cracking when they move
or expand. J oint reinforcement should not bridge expansion
joints. As the joint closes, the wire may buckle, pushing out
adjacent mortar, as shown in Photo 10. J oint reinforcement
may also transfer load across the expansion joint resulting
in additional cracking.
Corrosion of metal elements within masonry may cause
volume increases of such a magnitude as to crack or spall
the masonry; however, mortar, masonry units and grout are
considered to provide adequate protection when the mini-
mum cover and clearance requirements of Specification for
Masonry Structures (ACI 530.1/ASCE 6/TMS 602) [Ref. 13]
are met. Proper corrosion resistant coatings on the steel
item are also necessary.
SUMMARY
This Technical Note describes the various movements that occur within common building materials and construc-
tions. It also explains the effects of these movements. Cracking in brickwork can be minimized if all factors are
taken into consideration and the anticipated movement is accommodated.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. American Softwood Lumber Standard, Voluntary Product Standard PS 20, National Institute of Standards
and Technology, U.S. Department of Commerce, Washington, DC, 2005.
2. ASTM C 426, Test Method for Linear Drying Shrinkage of Concrete Masonry Units, Annual Book of ASTM
Standards, Vol. 04.05, ASTM International, West Conshohocken, PA, 2006.
3. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
4. Building Movements and J oints, Portland Cement Association, Skokie, IL, 1982.
5. Crack Control in Concrete Masonry Walls, NCMA TEK 10-1a, National Concrete Masonry Association,
Herndon, VA, 2005.
6. Fintel, M., Ghosh, S.K. and Iyengar, H., Column Shortening in Tall Structures-Prediction and
Compensation, Portland Cement Association, Skokie, IL, 1987.
7. Grimm, C.T., Masonry Cracks: A Review of the Literature, Masonry: Materials, Design, Construction, and
Maintenance, ASTM STP 992, H.A. Harris, Ed., ASTM, Philadelphia, PA, 1988.
8. Grimm, C.T., Probabilistic Design of Expansion J oints in Brick Cladding, Proceedings of the 4th
Canadian Masonry Symposium, University of New Brunswick, Canada, 1986.
9. Load and Resistance Factor Design Manual of Steel Construction, Third Edition, American Institute of
Steel Construction, Inc., Chicago, IL, 2001.
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 8 of 9
Photo 10
Spalling Due to Buckling of Joint
Reinforcement Bridging an Expansion Joint
10. Nilson, A.H., Darwin, D. and Dolan, C.W., Design of Concrete Structures, Thirteenth Edition, The
McGraw-Hill Companies, New York, NY, 2003.
11. Robinson, G.C., The Reversibility of Moisture Expansion, American Ceramic Society Bulletin, Vol. 64,
No. 5, 1985.
12. Scheffler, M.J ., Chin, I.R. and Slaton, D., Moisture Expansion of Fired Bricks, Proceedings of the Fifth
North American Masonry Conference, The Masonry Society, Boulder, CO, J une 1990.
13. Specification for Masonry Structures (ACI 530.1-05/ASCE 6-05/TMS 602-05), The Masonry Society,
Boulder, CO, 2005.
14. Wood Handbook - Wood as an Engineering Material, General Technical Report FPL-GTR-113, U.S.
Department of Agriculture, Forest Service, Forest Products Laboratory, Madison, WI, 1999.
15. Young, J .E. and Brownell, W.E., Moisture Expansion of Clay Products, Journal of the American Ceramic
Society, Vol. 42, No. 12, 1959.
16. Davidson, J .I., "Linear Expansion Due to Freezing and Other Properties of Bricks," Proceedings of the
Second Canadian Masonry Conference, Carleton University, Ottawa, Ontario, J une 1980.
www.gobrick.com | Brick Industry Association | TN 18 | Volume Changes - Analysis and Effects of Movement | Page 9 of 9
Accommodating Expansion
of Brickwork
Abstract: Expansion joints are used in brickwork to accommodate movement and to avoid cracking. This Technical Note
describes typical movement joints used in building construction and gives guidance regarding their placement. The theory and
rationale for the guidelines are presented. Examples are given showing proper placement of expansion joints to avoid cracking
of brickwork and methods to improve the aesthetic impact of expansion joints. Also included is information about bond breaks,
bond beams and flexible anchorage.
Key Words: differential movement, expansion joints, flexible anchorage, movement, sealants.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
18A
November
2006
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Vertical Expansion Joints in Brick Veneer:
For brickwork without openings, space no more than 25 ft
(7.6 m) o.c.
For brickwork with multiple openings, consider symmetrical
placement of expansion joints and reduced spacing of no
more than 20 ft (6.1 m) o.c.
When spacing between vertical expansion joints in para-
pets is more than 15 ft (4.6 m), make expansion joints
wider or place additional expansion joints halfway between
full-height expansion joints
Place as follows:
- at or near corners
- at offsets and setbacks
- at wall intersections
- at changes in wall height
- where wall backing system changes
- where support of brick veneer changes
- where wall function or climatic exposure changes
Extend to top of brickwork, including parapets
Horizontal Expansion Joints in Brick Veneer:
Locate immediately below shelf angles
Minimum in. (6.4 mm) space or compressible material
recommended below shelf angle
For brick infill, place between the top of brickwork and
structural frame
Brickwork Without Shelf Angles:
Accommodate brickwork movement by:
- placing expansion joints around elements that are rigidly
attached to the frame and project into the veneer, such
as windows and door frames
- installing metal caps or copings that allow independent
vertical movement of wythes
- installing jamb receptors that allow independent
movement between the brick and window frame
- installing adjustable anchors or ties
Expansion Joint Sealants:
Comply with ASTM C 920, Grade NS, Use M
Class 50 minimum extensibility recommended; Class 25
alternate
Consult sealant manufacturers literature for guidance
regarding use of primer and backing materials
Bond Breaks:
Use building paper or flashing to separate brickwork from
dissimilar materials, foundations and slabs
Loadbearing Masonry:
Use reinforcement to accommodate stress concentrations,
particularly in parapets, at applied loading points and
around openings
Consider effect of vertical expansion joints on brickwork
stability
2006 Brick Industry Association, Reston, Virginia Page 1 of 11
INTRODUCTION
A system of movement joints is necessary to accommodate the changes in volume that all building materials
experience. Failure to permit the movements caused by these changes may result in cracks in brickwork, as
discussed in Technical Note 18. The type, size and placement of movement joints are critical to the proper
performance of a building. This Technical Note defines the types of movement joints and discusses the proper
design of expansion joints within brickwork. Details of expansion joints are provided for loadbearing and
nonloadbearing applications. While most examples are for commercial structures, movement joints, although rare,
also must be considered for residential structures.
TYPES OF MOVEMENT JOINTS
The primary type of movement joint used in brick construction is the expansion joint. Other types of movement
joints in buildings that may be needed include control joints, building expansion joints and construction joints. Each
of these is designed to perform a specific task, and they should not be used interchangeably.
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 2 of 11
An expansion joint separates brick masonry into segments to prevent cracking caused by changes in
temperature, moisture expansion, elastic deformation, settlement and creep. Expansion joints may be horizontal
or vertical. The joints are formed by leaving a continuous unobstructed opening through the brick wythe that may
be filled with a highly compressible material. This allows the joints to partially close as the brickwork expands.
Expansion joints must be located so that the structural integrity of the brickwork is not compromised.
A control joint determines the location of cracks in concrete or concrete masonry construction due to volume
changes resulting from shrinkage. It creates a plane of weakness that, in conjunction with reinforcement or joint
reinforcement, causes cracks to occur at a predetermined location. A control joint is usually a vertical gap through
the concrete or concrete masonry wythe and may be filled with inelastic materials. A control joint will tend to
open rather than close. Control joints must be located so that the structural integrity of the concrete or concrete
masonry is not affected.
A building expansion joint is used to separate a building into discrete sections so that stresses developed in one
section will not affect the integrity of the entire structure. The building expansion joint is a through-the-building joint
and is typically wider than an expansion or control joint.
A construction joint (cold joint) occurs primarily in concrete construction when construction work is interrupted.
Construction joints should be located where they will least impair the strength of the structure.
EXPANSION JOINT
CONSTRUCTION
Although the primary purpose of expansion joints is to
accommodate expansive movement, the joint also must resist
water penetration and air infiltration. A premolded foam or
neoprene pad that extends through the full wythe thickness
aids in keeping mortar or other debris from clogging the joint
and increases water penetration resistance. Fiberboard and
similar materials are not suitable for this purpose because
they are not as compressible.
Mortar, ties or wire reinforcement should not extend into
or bridge the expansion joint. If this occurs, movement will
be restricted and the expansion joint will not perform as
intended. Expansion joints should be formed as the wall is
built, as shown in Photo 1. However, vertical expansion joints
may be cut into existing brickwork as a remedial action.
Sealants
Sealants are used on the exterior side of expansion joints to prevent water and air penetration. Many different
types of sealants are available, although those that exhibit the highest expansion and compression capabilities
are best. Sealants should conform to ASTM C 920, Standard Specification for Elastomeric Joint Sealants [Ref. 1],
Grade NS, Use M, and be sufficiently compressible, resistant to weathering (ultraviolet light) and bond well to
adjacent materials. Sealant manufacturers should be consulted for the applicability of their sealants for expansion
joint applications. Compatibility of sealants with adjacent materials such as brick, flashings, metals, etc., also
must be taken into consideration. Manufacturers recommend three generic types of elastomeric sealants for use
on brickwork: polyurethanes, silicones and polysulfides. Most sealants suitable for use in brickwork expansion
joints meet an ASTM C 920 Class 25 or Class 50 rating that requires them to expand and contract by at least
25 percent or 50 percent of the initial joint width, respectively. Sealants meeting Class 50 are recommended to
minimize the number of joints. Many sealants require a primer to be applied to the masonry surface to ensure
adequate bond.
Use a circular foam backer rod behind sealants to keep the sealant at a constant depth and provide a surface
to tool the sealant against. The sealant must not adhere to the backer rod. The depth of the sealant should be
approximately one-half the width of the expansion joint, with a minimum sealant depth of
1
/4 in. (6.4 mm).
Photo 1
Vertical Expansion Joint Construction
VERTICAL EXPANSION JOINTS
Figure 1 shows typical methods of forming vertical
expansion joints with either a premolded foam pad, a
neoprene pad or a backer rod.
While generally limited to rain screen walls, a two-stage
joint as shown in Figure 2 can increase resistance to
water and air infiltration. This type of joint provides a
vented or pressure-equalized joint. The space between
the sealants must be vented toward the exterior to allow
drainage. This is typically achieved by leaving a hole or
gap in the exterior sealant joint at the top and bottom of
the joint.
Spacing
No single recommendation on the positioning and spacing
of expansion joints can be applicable to all structures.
Review each structure for the extent of movements
expected. Accommodate these movements with a
series of expansion joints. Determine the spacing of
expansion joints by considering the amount of expected
wall movement, the size of the expansion joint and the
compressibility of the expansion joint materials. In addition
to the amount of anticipated movement, other variables
that also may affect the size and spacing of expansion
joints include restraint conditions, elastic deformation
due to loads, shrinkage and creep of mortar, construction
tolerances and wall orientation.
The theory and equation for estimating the anticipated extent of unrestrained brick wythe movement are presented
in Technical Note 18. Estimated movement is based on the theoretical movement of the brickwork attributed to
each property and expressed as coefficients of moisture expansion (k
e
), thermal expansion (k
t
) and freezing
expansion (k
f
). As discussed in Technical Note 18, for most unrestrained brickwork, the total extent of movement
can be estimated as the length of the brickwork multiplied by 0.0009. A derivative of this equation can be written to
calculate the theoretical spacing between vertical expansion joints as follows:

S
e

=
wjej

0.09 Eq. 1
where:
Se =spacing between expansion joints, in. (mm)
wj =width of expansion joint, typically the mortar joint width, in. (mm)
ej =percent extensibility of expansion joint material
The expansion joint is typically sized to resemble a mortar joint, usually
3
/8 in. (10 mm) to
1
/2 in. (13 mm). The width
of an expansion joint may be limited by the sealant capabilities. Extensibility of sealants in the 25 percent to 50
percent range is typical for brickwork. Compressibility of filler materials may be up to 75 percent.
Example. Consider a typical brick veneer with a desired expansion joint size of
1
/2 in. (13 mm) and a sealant with
50 percent extensibility. Eq. 1 gives the following theoretical expansion joint spacing:

S
e
=
(0.5 in.)(50)
0.09
= 278 in. or 23 ft - 2 in. (7.06 m)
Therefore, the maximum theoretical spacing between vertical expansion joints in a straight wall would be 23 ft
- 2 in. (7.1 m). This spacing does not take into account window openings, corners or properties of other materials
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 3 of 11
Figure 2
Two-Stage Vertical Expansion Joint
Figure 1
Premolded Foam Pad Neoprene Pad
Sealant & Backer Rod
Backer Rod
and Sealant
Vented Cavity
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 4 of 11
that may require a reduction in expansion joint spacing.
In most instances it is desirable to be conservative,
but it may be economically desirable to exceed the
theoretical maximum spacing as a calculated risk. For
example, calculations may result in a theoretical spacing
of expansion joints every 23 ft 2 in. (7.06 m) but the
actual expansion joint spacing is set at 24 ft (7.3 m)
to match the structural column spacing or a specific
modular dimension. Vertical expansion joint spacing
should not exceed 25 ft (7.6 m) in brickwork without
openings.
Placement
The actual location of vertical expansion joints in a
structure is dependent upon the configuration of the
structure as well as the expected amount of movement.
In addition to placing an adequate number of expansion
joints within long walls, consider placing expansion joints
at corners, offsets, openings, wall intersections, changes
in wall heights and parapets.
Corners. Walls that intersect will expand toward their
juncture, typically causing distress on one or both sides
of a corner, as shown in Figure 3a. Place expansion
joints near corners to alleviate this stress. The best
location is at the first head joint on either side of the
corner; however, this may not be aesthetically pleasing.
Masons can typically reach about 2 ft (600 mm) around
the corner from the face where they are working. An
expansion joint should be placed within approximately
10 ft (3 m) of the corner in either wall, but not necessarily
both. The sum of the distance from a corner to the
adjacent vertical expansion joints should not exceed
the spacing of expansion joints in a straight wall,
as shown in Figure 3b. For example, if the spacing
between vertical expansion joints on a straight wall is
25 ft (7.6 m), then the spacing of expansion joints around
a corner could be 10 ft (3.0 m) on one side of the corner
and 15 ft (4.6 m) on the other side.
Offsets and Setbacks. Parallel walls will expand toward
an offset, rotating the shorter masonry leg, or causing
cracks within the offset, as shown in Figure 4a. Place
expansion joints at the offset to allow the parallel walls to
expand, as Figure 4b illustrates. Expansion joints placed
at inside corners are less visible.
Openings. When the spacing between expansion joints
is too large, cracks may develop at window and door
openings. In structures containing punched windows and
door openings, more movement occurs in the brickwork
above and below the openings than in the brickwork
between the openings. Less movement occurs along
the line of openings since there is less masonry. This
differential movement may cause cracks that emanate
from the corners of the opening, as in Figure 5. This
pattern of cracking does not exist in structures with
continuous ribbon windows.
Figure 5
Cracking in Structure with Punched
Windows, Without Proper Expansion Joints
+ <Typ. Spacing
Between Expansion J ts.
Exp. J t.
Exp. J t.
L
1
L
2
(b)
(a)
Movement at Corner Without Expansion J oints
L
1
L
2
or <
_
10 ft. Either L
1
L
2
Proper Expansion J oint Locations at Corner
Direction of Expansion
Figure 3
Vertical Expansion Joints at Corners
(b)
(a)
Exp. J t.
Exp. J t.
Exp. J t.
Movement at Offset Without Expansion J oints
Proper Expansion J oint Locations at Offset
Direction of Expansion
Extent of
Expansion
Expansion
Figure 4
Vertical Expansion Joints at Offsets
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 5 of 11
Window and door openings weaken the wall and act as
natural expansion joints. One alternative is to place
expansion joints halfway between the windows. This
requires a sufficiently wide section of masonry between
the openings, typically 4 ft (1.2 m). It is often desirable
to locate vertical expansion joints along the edge or
jamb of the opening. In cases where the masonry above
an opening is supported by shelf angles attached to
the structure, a vertical expansion joint can be placed
alongside the opening, continuing through the horizontal
support.
If a vertical expansion joint runs alongside an opening
spanned by a loose lintel as shown in Figure 6a, the loose
steel lintel must be allowed to expand independently of
the masonry. A slip plane should be formed by placing
flashing above and below the angle. Mortar placed in front
of the lintel is subject to cracking; thus, a backer rod and
sealant should be used, as shown in Figure 6b. Because
steel expands more than masonry, a
1
/8 to
1
/4 in. (3.2 to 6.4
mm) space should be left at each end of the lintel. These
measures form a pocket that allows movement of the
steel angle within the brickwork. Locating the expansion
joint adjacent to the window will influence the dead weight
of the masonry bearing on the lintel. Instead of the usual
triangular loading, the full weight of the masonry above
the angle should be assumed to bear on the lintel. See
Technical Note 31B for more information about steel lintel
design. If a vertical expansion joint cannot be built in this
manner, do not place it alongside the opening.
Junctions. Expansion joints should be located at
junctions of walls with different environmental exposures
or support conditions. Separate portions of brickwork
exposed to different climatic conditions should be
separated with expansion joints since each area will
move differently. An exterior wall containing brickwork
that extends through glazing into a buildings interior
should have an expansion joint separating the exterior
brickwork from the interior brickwork. You may need
to use expansion joints to separate adjacent walls of
different heights to avoid cracking caused by differential
movement, particularly when the height difference is very
large. Examples are shown in Figure 7.
Parapets. Parapets with masonry exposed on the back
side are exposed on three sides to extremes of moisture
and temperature and may experience substantially
different movement from that of the wall below. Parapets
also lack the dead load of masonry above to help resist
movement. Therefore, extend all vertical expansion joints
through parapets. Since parapets are subject to more
movement than the wall below, they must be treated
differently. When vertical expansion joints are spaced
more than 15 ft (4.6 m) apart, the placement and design of
expansion joints through parapets need to accommodate
this additional movement. In this situation, make
Figure 6
Expansion Joint at Loose Lintel
Figure 7
Expansion Joints at Junctions
Expansion J oint
Steel Lintel Beyond
Backer Rod
and Sealant
(a)
(b)
Compressible Material
Steel Lintel
Flashing
Flashing
Used for
Slip Plane
Backer Rod and
Sealant
Exp. J oint
Different Environmental Exposure
(a)
Fence
Different Support Conditions
(b)
Opening
Expansion J oint
Expansion J oint
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 6 of 11
expansion joints in the parapet wider or add expansion joints placed halfway between those running full height.
These additional expansion joints must continue down to a horizontal expansion joint. As a third alternative, install
joint reinforcement at 8 in. (203 mm) on center vertically in the parapet.
Aesthetic Effects
Although expansion joints are usually noticeable on flat
walls of masonry buildings, there are ways to reduce
their visual impact. Architectural features such as quoins,
recessed panels of brickwork or a change in bond pattern
reduce the visual impact of vertical expansion joints. In
some cases, it may be desirable to accentuate the location
of the expansion joint as a design detail. This is possible by
recessing the brickwork at the expansion joint, or by using
special-shaped brick units as shown in Photo 2.
Colored sealants that match the brick in running bond, or
the mortar in stack bond, help to hide vertical expansion
joints. Masons sand also can be rubbed into new
sealant to remove the sheen, making the joint blend in
more. Expansion joints also are less noticeable when located at inside corners. Hiding expansion joints behind
downspouts or other building elements can inhibit maintenance access and is not advised. Toothing of expansion
joints to follow the masonry bond pattern is not recommended. It is more difficult to keep debris out of the joint
during construction; such debris could interfere with movement. Further, most sealants do not perform well when
subjected to both shear and tension.
Symmetrical placement of expansion joints on the elevation of buildings is usually most aesthetically pleasing.
Further, placing the expansion joints in a pattern such that wall areas and openings are symmetrical between
expansion joints will reduce the likelihood of cracking.
Other Considerations
Location of vertical expansion joints will be influenced by additional factors. Spandrel sections of brickwork
supported by a beam or floor may crack because of deflection of the support. Reduced spacing of expansion
joints will permit deflection to occur without cracking the brickwork.
Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402) [Ref. 4] and most building codes
allow anchored masonry veneer with an installed weight not exceeding 40 lb/ft
2
(1,915 Pa) and a maximum height
of 12 ft (3.66 m) to be supported on wood construction, provided that a vertical expansion joint is used to isolate
the veneer supported by wood from the veneer supported by the foundation.
HORIZONTAL
EXPANSION JOINTS
Horizontal expansion joints are typically needed if the
brick wythe is supported on a shelf angle attached to the
frame or used as infill within the frame. Placing horizontal
expansion joints below shelf angles provides space for
vertical expansion of the brickwork below and deformation
of the shelf angle and the structure to which it is attached.
Structures that support the brick wythe on shelf angles,
usually done for each floor, must have horizontal expansion
joints under each shelf angle. Figure 8 shows a typical
detail of a horizontal expansion joint beneath a shelf angle.
If the shelf angle is not attached to the structure when the
brick below it are laid, any temporary shims that support the
angle must be removed after the shelf angle is connected.
The joint is formed by a clear space or highly compressible
material placed beneath the angle, and a backer rod and
Photo 2
Accentuated Expansion Joint
Horizontal Expansion Joint at Shelf Angle
Shelf Angle
Weep
Sealant and Backer Rod
Min. 1/4 in. (6 mm) Thick
Compressible Material
Flashing
Flashing Protection
on Bolt Heads
Figure 8
Expansion Joint at Shelf Angle
sealant at the toe of the angle to seal the joint. It is not necessary to interrupt shelf angles at vertical expansion
joint locations. However, shelf angles must be discontinuous to provide for their own thermal expansion. A space
of in. in 20 ft (6 mm in 6 m) of shelf angle length is typically sufficient. Bolt heads anchoring a shelf angle to the
structure should be covered to decrease the possibility of flashing puncture.
The size of the horizontal expansion joint should take into account movements of the brickwork and movements
of the frame. Frame movements include both material and load-induced movements, such as deflections of
the shelf angle, rotation of the horizontal leg of the shelf angle, and movement of the support from deflection,
temperature change, shrinkage, creep or other factors.
When a large horizontal expansion joint is necessary, a lipped brick course may be used to allow movement while
minimizing the aesthetic impact of the joint. To avoid problems with breakage, the height and depth of the lipped
portion of the brick should be at least in. (13 mm). Lipped brick should be made by the brick manufacturer for
quality assurance purposes.
Construction using lipped brick requires careful
consideration of the frame movements noted
previously. Allowance for adjacent material tolerances
including the building frame should also be considered.
Adequate space should be provided between the
lipped portion of the brick and the shelf angle to ensure
no contact. Contact should not occur between the
lipped brick and the brickwork below the shelf angle
or between the lip of the brick and the shelf angle, not
only during construction, but also throughout the life of
the building.
Lipped brick may be installed as the first course above
a shelf angle, as shown in Figure 9a. Flashing should
be placed between the shelf angle and the lipped brick
course. Proper installation of flashing is made more
difficult because the flashing must conform to the shape
of the lip. This shape may be achieved with stiffer
flashing materials such as sheet metal. If the specified
flashing materials are made of composite, plastic or
rubber, a sheet metal drip edge should be used. The
practice of placing flashing one course above the shelf
angle is not recommended, as this can increase the
potential for movement and moisture entry.
Lipped brick also may be inverted and placed on the
last course of brickwork below a shelf angle, as shown
in Figure 9b. While installing an inverted lipped brick
course allows the flashing of the brickwork above to
maintain a straight profile through the brickwork, it also
allows the lipped brick course to move independent
of the shelf angle. Thus, there is an increased
possibility of the shelf angle coming in contact with
the lipped brick course, resulting in cracking at the lip.
It is difficult, if not impossible, to install compressible
material below the shelf angle. Further, it is likely that
temporary shims may be left in place between the lipped
brick and the shelf angle.
Horizontal expansion joints are also recommended
when brick is used as an infill material within the frame of the structure. Expansion joints must be provided
between the top course of brickwork and the member above. Deflections of the frame should be considered when
sizing the expansion joint to avoid inadvertently loading the brickwork.
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 7 of 11
Figure 9
Alternate Expansion Joint Detail
Inverted Lipped Brick
(b)
Lipped Brick
(a)
Shelf Angle
Weep
Sealant and Backer Rod
Min. 1/4 (6 mm) Thick
Compressible Material
Flashing
Lipped Brick
Flashing Protection
on Bolt Heads
Shelf Angle
Weep
Sealant and Backer Rod
Min. 1/4 in. (6 mm) Thick
Compressible Material
Flashing
Lipped Brick
Flashing Protection
on Bolt Heads
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 8 of 11
STRUCTURES WITHOUT SHELF ANGLES
Some buildings with brick veneer construction do not support the brickwork on shelf angles. These typically include
low-rise buildings constructed with wood and steel stud framing and buildings with shear walls. Building Code
Requirements for Masonry Structures limits brick veneer with wood or steel stud backing to a height of 30 ft (9 m)
to the top plate and 38 ft (12 m) to the top of a gable. Brick veneer with a rigid backing of concrete or concrete
masonry has no such limitation in the code. Brick veneer with this rigid backing may be supported by the foundation
without intermediate shelf angles to a recommended maximum height of about 50 ft (15 m), provided the building is
Anchorage to Steel Beam (section)
(a)
Anchorage to Steel Column (plan)
(b)
Anchorage to Concrete Beam (section)
(c)
Anchorage to Concrete Column
or Wall (section)
(d)
Adjustable
Anchor
Dovetail Anchor
Compressible
Filler
Dovetail Slot Dovetail
Anchor
Dovetail Slot
Concrete
Slab
Welded
Anchor Rod
Wire Anchor
Insulation
CMU
J oint Reinf.
with Eye &
Pintle
Debonded
Shear
Anchor
J oint Reinforcement
with Eye & Pintle
Insulation
Dampproofing
Compressible Filler
Flashing Membrane
Adjustable Ties
Figure 10
Flexible Anchorage to Beams and Columns
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 9 of 11
detailed appropriately for the differential movement and the moisture drainage system is designed and constructed
properly. In these buildings, differential movement is accommodated by the anchor or tie system, window details,
detailing at top of the wall and where other building components pass through the brickwork. These details must
provide independent vertical movement between the brickwork and the backing. Building components that extend
into or through the brick veneer (e.g., windows, doors, vents, etc.) also must be detailed to allow independent
vertical movement of the brick veneer and the component. The structural frame or backing provides the brick
veneer with lateral support and carries all other vertical loads. The veneer is anchored by flexible connectors or
adjustable anchors that permit differential movement. Allowance for differential movement between the exterior
brickwork and the adjacent components should be provided at all openings and at the tops of walls. Vertical
expansion joints also must be incorporated, as discussed in previous sections of this Technical Note.
Connectors, anchors or ties that transfer load from the brick wythe to a structural frame or backing that
provides lateral support should resist movement perpendicular to the plane of the wall (tension and
compression) but allow movement parallel to the wall without becoming disengaged. This flexible anchorage
permits differential movements between the structure and the brickwork. Figure 10 shows typical methods
for anchoring masonry walls to columns and beams. Technical Note 44B provides detailed information about
masonry ties and anchors.
The size and spacing of anchors and ties are based on tensile and compressive loads induced by lateral loads
on the walls or on prescriptive anchor and tie spacing requirements in building codes. Technical Note 44B lists
recommended tie spacing based on application.
There must be sufficient clearance among the masonry elements and the beams and columns of the structural
frame to permit the expected differential movement. The masonry walls may be more rigid than the structural
frame. This clearance provides isolation between the brickwork and frame, allowing independent movement.
COMBINING MATERIALS
Movement joints must be provided in multi-wythe brick and
concrete masonry walls. Expansion joints are placed in the
brick wythe, and control joints are placed in the concrete
masonry, although they do not necessarily have to be
aligned through the wall.
Bond Breaks
Concrete and concrete masonry have moisture and
thermal movements that are considerably different from
those of brick masonry. Floor slabs and foundations also
experience different states of stress due to their loading
and support conditions. Therefore, it may be necessary
to separate brickwork from these elements using a bond
break such as building paper or flashing. Such bond
breaks should be provided between foundations and
walls; between slabs and walls; and between concrete
and clay masonry, to allow independent movement while
still providing gravity support. Typical methods of breaking
bond between walls and slabs, and between walls and
foundations are shown in Figure 11.
When bands of clay brick are used in concrete masonry
walls, or when bands of concrete masonry or cast stone are
used in clay brick walls, differences in material properties
may cause mortar joints or masonry units to crack. Such
problems can be easily avoided by using bands of brickwork featuring brick of a different color, size or texture or
a different bond pattern. If, however, a different material is used for the band, it may be prudent to install a bond
break between the two materials, provide additional movement joints in the wall, or place joint reinforcement in the
bed joints of the concrete masonry to reduce the potential for cracking.
Bond Break
Brick
Insulation
CMU
Flashing as
Bond Break
Concrete Roof
Weep
Figure 11
Bond Breaks in Loadbearing Cavity Wall
Breaking the bond in this way does not affect the
compressive strength of the wall and should not affect
the stability of the veneer wythe when anchored properly.
The weight of the masonry, additional anchorage and
the frictional properties at the interface provide stability.
Sealant at the face of the joints between the different
materials will reduce possible water entry. If the band is
concrete masonry or cast stone, additional control joints are
recommended in the band. If the band is a single course,
there is a likelihood of vertical cracks at all head joints.
These can be closed with a sealant. Bands of two or more
courses should include horizontal joint reinforcement in
the intervening bed joints, as shown in Figure 12.
LOADBEARING MASONRY
The potential for cracking in loadbearing masonry
members is less than in nonloadbearing masonry
members because compressive stresses from dead and
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 10 of 11
Sealant in Raked J oint
(Typical) (Optional)
J oint Reinforcement
and Adjustable Anchor
in CMU Veneer Band
Bond Breaker
(Typical)
Adjustable Anchor
(Typical)
Brick Veneer
CMU Veneer
Band
Figure 14
Bond Breaks
Bond Break Material
Foundation
Anchor Bolt
Steel Plate
Bond Break at Foundation
(b)
Anchor Bolt
Bond Break at Roof
(a)
Bond Beam
Lath or
Hardware Cloth
Figure 12
Multi-Course Concrete Masonry Band
in Brick Veneer
Figure 13
Bond Beams
Bond Beam
Hollow Brick Construction
(b)
Cavity Wall Construction
(a)
Lath or
Hardware Cloth
Bond Beam
live loads help offset the effects of any movement. Adding reinforcement at critical sections such as parapets,
points of load application and around openings to accommodate or distribute high stresses will also help control
the effects of movement. Reinforcement may be placed in bed joints or in bond beams, as shown in Figure 13.
Historic loadbearing structures were not constructed with expansion joints. However, these walls were made of
multi-wythe brick construction, unlike typical structures built today.
When it is necessary to anchor a masonry wall to a foundation or to a roof, it is still possible to detail the walls in a
manner that allows some differential movement, as shown in Figure 14a and Figure 14b. Such anchorage is often
required for loadbearing walls subjected to high winds or seismic forces.
SUMMARY
This Technical Note defines the types of movement joints used in building construction. Details of expansion joints
used in brickwork are shown. The recommended size, spacing and location of expansion joints are given. By using
the suggestions in this Technical Note, the potential for cracks in brickwork can be reduced.
Expansion joints are used in brick masonry to accommodate the movement experienced by materials as they
react to environmental conditions, adjacent materials and loads. In general, vertical expansion joints should be
used to break the brickwork into rectangular elements that have the same support conditions, climatic exposure
and through-wall construction. The maximum recommended spacing of vertical expansion joints is 25 ft (7.6 m).
Horizontal expansion joints must be placed below shelf angles supporting brick masonry.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. ASTM C 920, Standard Guide for Use of Elastomeric J oint Sealants, Annual Book of Standards, Vol.
04.07, ASTM International, West Conshohocken, PA, 2006.
2. Beall, C., Masonry Design and Detailing for Architects, Engineers and Contractors, Fifth Edition, McGraw
Hill, Inc., New York, NY, 2003.
3. Beall, C., Sealant J oint Design, Water on Exterior Building Walls: Problems and Solutions, ASTM STP
1107, T.A. Schwartz, Ed., ASTM, West Conshohocken, PA, 1991.
4. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
5. Building Movements and J oints, Portland Cement Association, Skokie, IL, 1982.
www.gobrick.com | Brick Industry Association | TN 18A | Accommodating Expansion of Brickwork | Page 11 of 11

Technical Notes 19 - Residential Fireplace Design
January 1993
Abstract: This Technical Notes covers the components, design and dimensions of residential wood-burning
fireplaces. The recommendations are limited to single-face fireplaces. Concepts for increased energy efficiency as a
supplemental heating unit are also addressed. Recommendations for the selection of materials as they relate to the
construction of fireplaces are included.
Key Words: brick, building codes, damper, design, energy efficiency, fireplace, heating, masonry, mortar.
INTRODUCTION
In past years, fireplaces were mostly decorative and seldom used for heating residences. Recently, there has been
a revival of the use of fireplaces as a supplemental heating source. This use requires that fireplaces be built as a
more functional unit.
The combustion process, the variety of firebox configurations and the varying rates at which fuel is consumed are
complex. Development of test methods to take into consideration the many variables affecting fireplace performance
has been conducted for several years. However, the comparative performance of different fireplace configurations is
not yet well documented. Thus the design suggestions in this Technical Notes are based on past proven
performance. Although governed by the laws of thermodynamics, fireplace design is not an exact science, instead it
is an empirical art to be applied with knowledge of basic principles and good judgment.
Successful concepts that were discarded when fireplaces became mostly decorative are now finding new relevance.
In addition, several new concepts have been developed which can increase the energy efficiency of conventional
fireplaces.
The purpose of this Technical Notes is to provide basic information for the design of successful, single-face,
wood-burning fireplaces and to introduce concepts that can increase their energy efficiency. Other Technical Notes
in this series discuss design, detailing and construction of other styles of residential fireplaces and the design of
residential chimneys and masonry heaters.
TYPES OF FIREPLACES
There are several distinct types of fireplaces currently in use for residential applications. There are many individual
variations within each general type, but most of the functional principles are similar.
Single-Face
Single-face fireplaces have been in use since early recorded history with developments in design through most of the
major architectural periods. Most of the available information on the proper opening sizes, dampers, and flue sizing
is based on empirical developments.
Single-face fireplaces can provide relatively efficient room heating. The amount of radiated and reflected heat
produced increases with the amount of brick masonry surrounding the fire. The amount of brick masonry surface
area exposed to the fire, its distance from the fire and the size of the fire determine the amount of reflected and
radiated heat. Also, the mass of the fireplace assembly stores heat and radiates the heat into the room after the fire
is extinguished. Key elements of a single-face fireplace are shown in Figures 1a through 1c.
19 http://www.gobrick.com/BIA/technotes/t19.htm
1 of 14 9/13/2009 12:59 PM
Typical Single-Face Fireplace (See Table 1 for Dimensions)
FIG. 1a
Typical Single-Face Fireplace (See Table 1 for Dimensions)
19 http://www.gobrick.com/BIA/technotes/t19.htm
2 of 14 9/13/2009 12:59 PM
FIG. 1b
Typical Single-Face Fireplace (See Table 1 for Dimensions)
FIG. 1c
Rumford Fireplaces. The Rumford fireplace is a single-face fireplace with a firebox which features widely splayed
sides, a shallow depth and a high opening. These features increase energy efficiency. Performance tests indicate
that the radiated and reflected heat output from a Rumford fireplace is higher than that from a conventional fireplace.
Information on the design and construction of Rumford fireplaces is provided in Technical Notes 19C Revised.
Rosin Fireplaces. The Rosin fireplace is a single-face fireplace with a specially curved back to the firebox, designed
to increase energy efficiency. The Rosin has a cast refractory firebox with widely splayed sides which increases
radiation and heat storage. The Rosin firebox can be retrofitted into an existing masonry fireplace or built into a new
fireplace.
Air-Circulating Fireplaces. Air-circulating fireplaces are so named because they circulate room air behind the
combustion chamber through a series of brick baffles or steel plates. As a result, additional heat output can be
distributed to the room or other areas of the residence by convection.
Examples of this type of single-face fireplace are the Heatilator and the Brick-O-Lator [7]. This type of fireplace can
be used as a supplemental heat source. Brick has the advantage of radiating the heat stored in the mass of the
fireplace after the fire is out. Additional heat can be distributed to the room by continuing to circulate air long after
the fire is out.
Multi-Face
Multi-face fireplaces have adjacent, opposite or all faces open to the room. Although generally associated with
contemporary design, the multi-face fireplace is also of ancient origin. For example, the so-called corner fireplace
which provides two adjacent open sides has been in use for several hundred years in Europe. Some multi-face
fireplaces have unique design requirements which have to be met before satisfactory performance can be reached.
These fireplace configurations are less energy efficient than single-face fireplaces. This is due to the lack of radiating
surfaces and increased use of room air. Multi-face fireplaces are usually selected for aesthetics rather than energy
efficiency. Multi-face fireplaces are discussed in Technical Notes 19C Revised.
19 http://www.gobrick.com/BIA/technotes/t19.htm
3 of 14 9/13/2009 12:59 PM
FIREPLACE DESIGN
The performance of a fireplace is primarily governed by three factors: fuel combustion, air pressure differential
between the firebox and the top of the chimney and temperature differential between air in the room of the fire and
that at the top of the chimney. All must be considered in order to achieve successful combustion and exhaust
performance. All fireplaces include the same four basic components. These are the base, firebox, smoke chamber
and the chimney. Of these, all but the base influence burning performance.
Base
The base consists of the foundation and hearth support, as shown in Figs. 1a and 2. It is not necessary that all of
the components shown be present. For slab-on-grade construction, the slab can provide both the foundation and the
hearth support, providing it is adequately designed to support the weight of the fireplace assembly.
Typical Base Assembly
FIG. 2
Foundation. Masonry fireplaces must be supported with an adequate foundation. The foundation consists of either
footings which support foundation walls or a structural slab. Local building codes should be reviewed for design soil
pressures for foundations. The minimum requirements contained in most building codes for the foundation
components are included in the following discussion. The foundation must be designed to carry the weight of the
fireplace without excessive or differential settlement.
Footings-Footings should be made of masonry or concrete and at least 12 in. (300 mm) thick, and extend at least 6
in. (150 mm) beyond the fireplace walls on all sides. The footings should penetrate below frost line unless they are
located within a space maintained above freezing. Footings should be placed on undisturbed or properly prepared
soils.
Foundation Walls-Foundation walls raise the fireplace to the desired level and should be constructed of masonry or
concrete with a minimum thickness of 8 in. (200 mm). There should be no voids except for the ash pit and external
combustion air ducts formed in the base assembly, as shown in Figs. 1a and 2. Typically the shape of foundation
walls matches the perimeter of the fireplace structure above.
Structural Slab-The structural slab must be properly designed to support the weight of the fireplace assembly.
When the fireplace is constructed on a slab-on-grade it is usually necessary to thicken the slab under the fireplace to
support the loads from the fireplace and chimney.
Hearth Support. Support for the hearth can be provided in a number of ways. These include the use of corbeled
brickwork, a structural concrete slab or cantilevered reinforced brick masonry. The maximum projection of each brick
in a corbel should not exceed one-half the height of the unit nor one-third its thickness. When corbeling from walls,
the overall horizontal projection should be limited to one-half of the wall thickness unless the corbel is reinforced.
These maximum horizontal individual and overall projections are consistent with current model building code
requirements. Hearth support featuring corbeled brickwork and a structural slab are shown in Figs. 1a and 2.
19 http://www.gobrick.com/BIA/technotes/t19.htm
4 of 14 9/13/2009 12:59 PM
A structural concrete slab or reinforced brick masonry is used to span the foundation walls and may cantilever to
support the hearth extension.
Firebox
The firebox consists of the hearth, fireplace opening, combustion chamber, throat and often a smoke shelf as shown
in Figs. 1 and 3. The thicknesses of the firebox walls are set by the model building codes. When refractory brick or
firebrick are used to line the walls the total thickness may be reduced.
Hearth. The hearth consists of two basic parts, the inner hearth and the extended hearth. The hearth can be raised
or flush with the floor surface. A fireplace hearth flush with the floor is shown in Figs. 1a and 3.
Inner Hearth-The inner hearth is within the firebox area and forms the floor of the combustion chamber. All model
building codes require that the inner hearth and the hearth support be noncombustible and a minimum of 4 in. (100
mm) thick.
Extended Hearth-The extended hearth is that portion of the hearth that projects out into the room beyond the face
of the fireplace and must be noncombustible. Model building codes require the extended hearth to be supported by
noncombustible materials with no combustible material against the underside. Wooden forms or centers used to
construct the hearth extension must be removed when construction is completed. The extended hearth may be a
reinforced brick masonry cantilever.
Typical Firebox Assembly
FIG. 3
Model building codes also require that the hearth extend a minimum of 8 in. (200 mm) on each side of the fireplace
opening and 16 in. (400 mm) in front of the fireplace opening. If the fireplace opening is greater than 6 ft
2
(0.55 m
2
),
building codes require hearth extensions of 12 in. (300 mm) on either side of the opening and 20 in. (500 mm) in
front of the fireplace opening.
Fireplace Opening. The fireplace opening is a very important element in fireplace design. The configuration and
dimensions of most other components of the fireplace and chimney are based primarily on the dimensions of the
fireplace opening selected. Figure 1 shows details and Table 1 provides the widths and heights of fireplace openings
found to be the most satisfactory for appearance and successful operation. These dimensions may be varied slightly
to allow for brick coursing.
Proper Sizing-Firebox dimensions should be selected so that the fire fills the combustion chamber during operation.
This provides greater heating efficiency. Careful consideration should be given to size of the fireplace opening best
suited for the room in which it is to be located. Location and size are important not only from the standpoint of
appearance, but also of operation. If the fireplace opening is too small, it may function properly but will not produce
19 http://www.gobrick.com/BIA/technotes/t19.htm
5 of 14 9/13/2009 12:59 PM
enough heat to warm the room. If the opening is too large, a fire that would fill the combustion chamber may
overwhelm the room. In such a case, the firebox opening would require a larger flue area and consume larger
amounts of interior air even if exterior combustion air is provided. Table 2 provides suggested widths of conventional
fireplace openings appropriate for various room sizes. For example, a room with 300 ft
2
(28 m
2
) of floor area is best
served by a fireplace with an opening 30 in. (750 mm) to 36 in. (900 mm) wide.
The shape of the fireplace opening is important aesthetically and functionally. Higher openings increase the radiant
heating, increase the demand for room air and require taller chimneys.

a
Adapted from Book of Successful Fireplaces, 20th Edition.
b
SI conversion: mm = in. x 25.4.
c
L and M are shown in Fig. 1 and are equal to outside dimensions of flue lining plus at least 1 in. (25 mm). Determine flue lining dimensions from Fig.
5. L is greater than or equal to M.
d
Angle sizes: A - 3 x 3 x 1/4 in., B-3 1/2 x 3 x 1/4 in., C-5 x 3 1/2 x 5/16 in.
Support Above Fireplace Opening-The brickwork above the fireplace opening must be adequately supported.
There are several alternatives for support. These include brick arches, reinforced brick masonry lintels, stone,
precast concrete and loose angle lintels.
Brick arches usually require no steel reinforcement and are an attractive option. When determining the height of a
fireplace opening which incorporates an arch use the maximum height to the arch soffit. Information on arch design
may be found in Technical Notes 31 Series.
Reinforced brick masonry (RBM) lintels may be built in place or prefabricated. The advantages of using RBM lintels
are numerous, but include more efficient use of materials and exposed brick rather than steel at the top of the
opening. RBM lintel design procedures are given in Technical Notes 17H. Loose steel angle lintels are the most
prevalent means of support. For this reason, Table 1 gives recommended steel angle dimensions. If opening sizes
other than those listed in Table 1 are used, information found in Technical Notes 31B Revised can be used for
design of the loose steel angle lintel.
19 http://www.gobrick.com/BIA/technotes/t19.htm
6 of 14 9/13/2009 12:59 PM
a
Reprinted with permission of Structures Publishing Company from Book of Successful
Fireplaces, 20th Edition.
b
SI conversions: mm = ft. x 0.305; mm = in. x 25.4.
General recommendations are: the steel angles should be at least 1/4 in. (6A mm) thick; the horizontal leg should be
at least 3-1/2 in. (89 mm) for use with nominal 4 in. (100 mm) thick brick and 3 in. (75 mm) for use with nominal 3 in.
(75 mm) thick brick. The minimum required bearing length on each end of the fireplace opening is 4 in. (100 mm).
Steel angle lintels should have a space at their ends to permit thermal expansion.
Combustion Chamber. The shape and depth of the combustion chamber will greatly influence draft, combustion air
requirements and the amount of heat reflected and radiated into the room. Figure 1 illustrates the shape and Table 1
provides recommended dimensions for the combustion chamber. These dimensions may be varied slightly, but the
information given is based on successful designs. Significant changes should not be made without consulting a
fireplace design consultant.
The sides and lower portion of the back of the combustion chamber should be vertical. Above the vertical portion of
the back, the brick should be sloped forward towards the fireplace opening to support the metal damper and the
clay flue lining. For the maximum amount of reflected heat into the room, the sloped portion of the back should be
plane rather than concave. If it is concave, more heat will be reflected back into the fire rather than into the room.
Greater splay of the sides also increases the amount of heat reflected into the room.
The combustion chamber should be constructed of nominal 4 in. (100 mm) thick brick. When refractory brick or
firebrick are used, model building codes permit the total wall thickness to be reduced. Thin mortar joints, not more
than 1/4 in.(6.4 mm), should be specified. A 1 in. (25 mm) air space should be provided between the combustion
chamber wall and the backup wall, although not required by building codes. This air space provides for thermal
expansion of the combustion chamber. A noncombustible, compressible, fibrous insulation or similar material should
be wrapped around the combustion chamber to ensure that this air space is maintained. The backup wall should be
no less than 4 in. (100 mm) in thickness around the back of the combustion chamber to support the loads from the
smoke chamber and chimney above.
Throat. The throat is a slot-like opening directly above the top of the firebox through which flames, smoke and
combustion gases pass into the smoke chamber and upward through the chimney. Because of its effect on draft, the
throat of the fireplace should be carefully designed. It should be a minimum of 8 in. (200 mm) above the highest point
of the fireplace opening. The throat is illustrated in Fig. 1a and appropriate dimensions are found in Table 1.
Cast refractory and formed clay throats are available, built to certain angles and dimensions to fit most conventional
fireplace dimensions. These elements are positioned on top of the firebox walls and eliminate the need of
constructing brick courses to form the throat. Once in place, brick masonry can be built around the throat to give the
appearance of conventionally built throats in the breastwork of the fireplace.
Damper-The damper closes the fireplace opening to exterior air infiltration and can be used to control the burning
rate of the fireplace. A metal damper may be placed in the throat, extending the full width of the throat opening, or at
the top of the chimney.
A throat damper should have an open area of approximately twice the area of the flue. The damper should have a
valve plate which opens toward the back of the fireplace. Such a plate when opened, forms a barrier to deflect any
19 http://www.gobrick.com/BIA/technotes/t19.htm
7 of 14 9/13/2009 12:59 PM
down draft which may occur. Many different damper shapes are available. A high formed damper is recommended
because it extends the throat with its construction and forms a critical portion of the smoke chamber. This damper
type reduces the possibility of masonry blocking the valve plate of the damper. The damper should be spot bedded
in mortar for a good fit and support, but not mortared in solidly at the ends because expansion could cause cracking
in masonry. A noncombustible, compressible, fibrous insulation or similar material should be placed between the
damper ends and adjacent masonry to allow differential movement.
A chimney top damper is an alternative to the damper installed at the top of the throat. The damper is operated by a
control chain which extends down into the firebox. This type of damper permits the chimney and flue to be heated
when the fireplace is not in use and may help reduce water penetration into the flue. Chimney top dampers must be
weighted or spring loaded to be in the open position if the operating mechanism fails. This is necessary so the
damper remains open during operation of the fireplace.
Smoke Shelf. The origin of and need for a smoke shelf is not clear. Some say its purpose is to provide a location
for chimney sweeps to work from when cleaning large chimneys. Others contend it deflects down drafts and
prevents direct access of water entering the top of the flue to the firebox. It also serves as a depository for ash
which does not clear the chimney. The smoke shelf, if used, should be designed so that a uniform air flow results.
The smoke shelf should be directly under the flue, be level across the face and in plane with the base of the damper.
The smoke shelf should also extend the full width of the throat. It can be flat, extending back to and perpendicular to
the rear wall of the smoke chamber, or curved to blend with the rear wall of the smoke chamber. Refer to Figs. 1a
and 3 for details and Table 1 for recommended dimensions.
Some designs, such as Rumford fireplaces, do not include a smoke shelf. These types of fireplace designs are often
referred to as having "the streamline effect". In this instance, the flue tile is vertically aligned with the top of the last
course of brick at the back of the firebox wall. Such a design provides a clear vertical passageway from the firebox
to the top of the last chimney flue liner.
Smoke Chamber
The smoke chamber forms the chimney flue support, as shown in Figs. 1a, 1b and 4, and conveys by-products of
the combustion process up to the chimney. The back wall of the chamber is built vertically and the side walls are
sloped uniformly toward the center. The front wall above the throat is also sloped to meet and provide support for
the bottom of the clay flue liner. Flue liners should be supported on all sides. The front wall above the throat should
be supported by reinforced brick masonry or a steel angle, not by the damper.
Typical Smoke Chamber Assembly
FIG. 4
The slope of the smoke chamber should be smooth, with each course of brick corbeled to achieve the required
angle. The inside of the smoke chamber should be parged with refractory mortar to reduce friction and prevent
smoke leakage. Figures 1a and 4 show the shape of the smoke chamber and Table 1 gives recommended
dimensions.
There are alternative means of building the inside surface of the smoke chamber. Cast refractory materials or cut
19 http://www.gobrick.com/BIA/technotes/t19.htm
8 of 14 9/13/2009 12:59 PM
pieces of clay flue liner may be used. Dimensional coordination is important so that all components are correctly
fitted without cracks or leaks of combustion products.
Chimney Flue
Draft of the fireplace is affected by the dimensions of the firebox opening, the shape and cross-sectional area of the
flue and the height of the chimney. Figure 5 provides a graphical determination of the appropriate flue size for
fireplace opening area and overall height [3]. For purposes of Fig. 5, the height is defined as the distance from the
combustion chamber floor to the top of the last chimney flue liner. When using Fig. 5 it is normally best to use the
smaller flue size when the opening and height selected intersect between standard flue sizes. Taller chimneys have a
better draw than shorter chimneys with the same flue size.
For fireplace openings greater than those given in Table 1 and Fig. 5, the area of the flue and height of the chimney
can be determined by methods in Technical Notes 19B Revised. Chimney design and construction are also covered
in Technical Notes 19B Revised.
Flue Size Nomograph
FIG. 5
Structural Considerations
Masonry fireplaces must withstand wind and seismic loads resulting from local conditions. In areas of high wind and
seismic activity, vertical and horizontal reinforcement may be required. Vertical reinforcement is located at least at
each corner of the fireplace. Such reinforcement must be anchored to the foundation and properly lapped to be
continuous for the entire chimney height. The size and spacing of reinforcement depends on design loads, overall
dimensions of the fireplace and chimney, location of the reinforcement and means of attachment to the structure.
Fireplaces and chimneys are typically attached to the structure by steel straps located at each floor or ceiling line.
Consult the local building code for design loads and prescriptive requirements. For more information on chimney
design see Technical Notes 19B Revised.
Aesthetic Considerations
The appearance of the fireplace has evolved through the centuries from the elaborately carved mantels of the
Georgian Period to the smaller, streamlined fireplaces found in contemporary style homes. The aesthetic design of a
fireplace is often based on the style of the house or room. The fireplace may project from adjacent walls to add
19 http://www.gobrick.com/BIA/technotes/t19.htm
9 of 14 9/13/2009 12:59 PM
emphasis to the fireplace or may be flush with its surroundings. The effect of a fireplace can be simple, just a
rectangular opening with a brick surround in an otherwise blank wall. Conversely, a focal point can be created with
an ornate brick area filling an entire wall. Functional aspects such as wood storage areas or seating can be
incorporated. Brickwork can be combined with materials in other locations.
The most prominent features of the fireplace are the fireplace surround, the mantel, and the hearth. Although certain
aspects of these features must conform to building code requirements, the resulting appearance is limitless.
Mantel. The mantel is a shelf or facing ornament above the fireplace opening. Depending on the architectural style
of the room, the mantel may be recessed into the wall or may project out from the wall. Mantels may be built
integrally with the fireplace or may be anchored to it. Specially carved mantels are sometimes used to surround the
fireplace. Projecting mantels are usually made of corbeled masonry, wood, stone or other materials. All combustible
materials used for the mantel must be at least 6 in. (150 mm) away from the fireplace opening. Combustible
materials projecting out more than 1-1/2 in. (38 mm) must be 12 in. (300 mm) away from the top of the fireplace
opening. Corbeled masonry must conform to the corbeling limitations listed in the Hearth Support section. All mantels
should be securely attached to the masonry. The wall above the mantel is an area which is often integrated with the
fireplace design. This may include patterned brickwork, brick sculptures or art work. Figure 6 is an example of a
mantel and the possibilities above the mantel.
FIG. 6
Fireplace Surround. The fireplace surround is the area immediately surrounding the fireplace opening. The first 6 in.
(150 mm) adjacent to the fireplace opening must be noncombustible material. The fireplace surround may be integral
with the mantel in the case of decorative tile, marble or other noncombustible material placed on either side of the
opening. Alternately, a wooden surround may be combined with the mantel. The lintel forms the top of the fireplace
opening and may also be made of various noncombustible materials. The shape of the lintel can be modified to add
a certain look to the fireplace. A semi-circular arch is one simple way of dressing up the fireplace. An example is
shown in Fig. 7. Other options include using a contrasting material such as cast stone for the lintel.
Hearth Extension. The hearth extension is necessary for the safe operation of a fireplace and may also be a focal
point of the fireplace. The hearth extension may be flush with the floor or may be raised. The extended hearth may
be made of brickwork, slate or any other noncombustible material. The hearth extension may be only as small as
allowed by the building code or may extend along the entire front face of the wall. The raised hearth may serve as
additional seating for the room. A raised hearth brings the level of the fireplace up to eye level when seated. Raised
19 http://www.gobrick.com/BIA/technotes/t19.htm
10 of 14 9/13/2009 12:59 PM
hearths are shown in Figs. 6 and 7.
FIG. 7
ENERGY EFFICIENCY WITH FIREPLACES
Energy efficient fireplaces may be used for supplemental heating and to decrease the consumption of non renewable
resources. Several modifications to conventional fireplace design make them more energy efficient. The
modifications discussed here are appropriate to most conventional fireplace designs. Other energy efficient
modifications to the shape and size of the firebox are the Rosin and Rumford fireplace designs.
Location
For maximum thermal benefit, the fireplace should be located entirely within the structure. This enables the mass of
the fireplace to store heat within the residence. Heat stored in the brickwork is then radiated into the room long after
the fire is extinguished.
By choosing a central location, a more even heating of the living area results. Fireplace walls can be exposed in
several rooms. Cold spots in areas away from the fire are kept to a minimum and, if the fireplace is an air-circulating
type, heat can be vented into adjacent rooms more efficiently.
Outside Air
One way to increase the efficiency of a fireplace is to use air from outside the structure for combustion and draft.
Conventional fireplaces draw air from the room, air that has already been heated to some extent. The drop in room
air pressure, caused by the air loss, may result in increased infiltration from other areas of the structure. In very
tightly built houses less air is available for proper combustion, so outside air must be intentionally provided. Even
when outside air is provided, some interior room air is always necessary for proper combustion.
There are many ways in which outside air can be brought into the firebox area. Each method requires three basic
parts: the intake, the air passageways and the inlet. One example is shown in Figs. 1 and 2. Thight-fitting inlet
dampers are tight-closing intake louvers and recommended to keepthe fireplace from becoming a source of air
infiltration when not in use.
Intake. The intake should be located on an ouside wall or on the back of the firplace. A screen-backed, closeable
louver is required. Preferably, this will be a type that can be operated from inside the structure. Many building codes
19 http://www.gobrick.com/BIA/technotes/t19.htm
11 of 14 9/13/2009 12:59 PM
will not permit the intake to be located within a garage because of the presence of fuel fumes. Other possible
locations for the intake are in a crawl space, attic or other unheated spaces. It is advisable to check local building
code regulations for the appropriateness of other intake locations.
Passageway. A passageway or duct connects the intake to the inlet. It must be formed of noncombustible material.
Ducts with cross-sectional area ragning from 6 in.
2
(3870 mm
2
) have been used successfully. The passageway can
be built integral withthe fireplace base assembly or channeled between floor joists. It can aslo enter throught inlets
located in the sides ofthe firebox. In any case, the passageway is usually insulated to reduce heat loss.
Inlet. The inlet brings the outside air into the firebox. A damper is required to control the volume and direction of the
air flow. This is necessary because cold outside air channeled into the fireplace expands and could possibly result in
more air than is needed for draft and combustion. this can create a spill over effect into the room proior to the air
being warmed. The inlet can be located in the sides or the floor of the combusiton chamber, preferably infront of the
grate for best performance. If the inlet is located toward the back of the combustion chamber, ashes may be blown
into the room by drafts for the inlet. As an option, the inlet can be located on or near the floor within 24 in. (600 mm)
of the firebox opening. Any inlet should be closeable and designed to prevent burning material from dropping into
concealed comustible spaces.
A potential problem due to oncreased velocity of the air coming throught the inle is that the temperature within the
combustion chamber can increase significantly. This can result in grates and inlet dampers being destroyed or
distorted by the higher temperatures. To help decrease the velocity of the air through the inlet, a space before the
inlet should be constucted as a stilling chamber, as shown in Figs. 1a and 2.
Glass Fireplace Screens
Glass screens can be used on both conventional fireplaces and fireplaces with an outside air supply. these screens
should be sealed around the edges and have tight-fittingdoors and vents so that the fireplace is not a source of air
infiltration or heat loss when not in use. The screens are normally closed when the fireplace is not being used.
During a fire, glass screens provide a barrier which reduces the amount of heated air being channeled up the
chimney, but still permit smoke and comustion gases to escape. The screens should be kept closed until it is safe to
close the damper.
Caution is necessary whenb fireplaces are operated with the glass screens in closed position. Increased
temperatures due to higher air velocities through intakes can warp grates or metal in conjunction with the glass
doors, cause expansion of the glass doors and the steel lintel above the fireplace opening and lead to early
disintegration ofthe firebox mortar joints.
SELECTION OF MATERIALS
The proper selection of quality materials is essential to the successful performance ofthe fireplace and chimney. No
amount of design, detailing and construction can compensate for the imporoper selection of materials.
Brick
Building codes require that solid masonry units, i.e. cored up to 25 percent, be used for fireplace construction. Brick
should conform to ASTM C 216 or C 62 for facing brick and building brick, respectively. In areas of hight seismic
activity, the option exists to use hollow brick conforming to ASTM C 652 which can be vertically reinforcedand fully
grouted. Grade SW should be specified for durability since the fireplace assembly is usually subject to severe
exposure conditions.
For the firebox, the use of refactory brick or firebrick which conform to ASTM C 27, low duty, permit a reduced wall
thickness. Refractories are more resistant to high temperatures and thermal shock. Grade SW building brick or
facing brick may be used as an alternative when exposure to wood-burning firesis anticipated. Currently, ASTM
Committee C-15 is working ona standard specification for firebox brick which will replacethe discontinued ASTM C
64 previously used in most model building codes.
Salvaged brick should not be used because they may not provide the strength and durablility necessary for
satisfactory performance. The use of salvaged brick is discussed in Technical Notes 15 Revised.
Mortar
Combustion Chamber, Smoke Chamber and Flue. Mortars used in these locations are subject to high surface
temperatures and possibly corrosive effects from combustion gases. The mortar joints at the top of the chimney flue
may be subjected to periodic wetting and freeze-thaw cycling. The mortar must withstand these conditions while
providing adequate support and a barrier to combustion gases. Mortars used for these three parts of the fireplace
19 http://www.gobrick.com/BIA/technotes/t19.htm
12 of 14 9/13/2009 12:59 PM
can be a refractory mortar or a conventional mortar.
Refractory mortar should conform to ASTM C 199, medium duty, and may be one of several types. The properties
of each should be evaluated for the intended use and exposure. Fireclay is the primary ingredient of refractory
mortars, often mixed with calcium aluminate or sodium silicate as a binder. Refractory mortars must be used with
thin joints.
High-lime mortars, such as ASTM C 270 Type O portland cement-lime mortar, have been found to be more resistant
to heat in the combustion chamber than high portland cement content mortars. The joint size also effects the
performance of the mortar. In any case, mortar joints in the combustion chamber should be no greater than 1/4 in.
(6.4 mm) thick to protect against the effects of cracking or deterioration through fireplace use.
Conventional Brickwork. It is often more convenient and economical to use only one type of mortar for all
components of the fireplace and chimney. Type N portland cement-lime mortar and Type S masonry cement mortar
conforming to ASTM C 270 are good all-purpose mortars for most residential fireplaces and chimneys. Chimney
wind loads in excess of 25 psf (1.2 kPa) may require Type S portland cement-lime mortar. Masonry in contact with
earth should be laid with a Type M mortar.
Clay Flue Liners
Flue liners should conform to ASTM C 315. They should be free from cracks or other damage that might contribute
to smoke or gas leakage. Clay flue liners come in rectangular, round and oval shapes. Rectangular flue liners are
either modular or nonmodular in cross-sectional dimensions. Sizes stated in ASTM C 315 for rectangular and oval
liners are outside dimensions. Modular sizes start at 3.5 in. (90 mm) and increase in 4 in. (100 mm) modules and
may be specified by nominal dimensions. Round clay flue liners are specified as nominal inside diameter. See
Technical Notes 19B Revised for a list of clay flue liner sizes.
Steel Lintels
Steel conforming to ASTM A 36 should be used for lintels supporting brick masonry in fireplace construction.
Ties and Reinforcement
Corrugated Metal Ties. Corrugated metal ties may be used to tie brick of the fireplace walls and the exterior
brickwork to wood frame backups. Ties should be corrosion resistant, at least 22 gage, 7/8 in. (22 mm) wide, and
long enough to be embedded at least half-way into each wythe thickness.
Wire Ties. Wire ties are recommended for tying brick construction together. They should be at least wire size W1.7
(9 gage) and corrosion resistant. Ties should be fabricated from wire which conforms to ASTM A 82 or A 185.
Prefabricated Joint Reinforcement. Prefabricated joint reinforcement should be corrosion resistant and fabricated
from wire which conforms to ASTM A 82 or A 185.
Bar Reinforcement. Reinforcement should conform to any of the following applicable standards: ASTM A 615, A
616 or A 617.
Corrosion Resistance. Corrosion resistance is usually provided by zinc coatings or by using stainless steel. To
ensure adequate resistance to corrosion, coatings or materials should conform to any of the following applicable
standards: Zinc-Coating of Flat Metal-ASTM A 153, Class B-2
Zinc-Coating of Wire-ASTM A 641, Class 3
Copper-Coated Wire-ASTM B 227, Grade 30 HS
Stainless Steel-ASTM A 167, Type 304
SUMMARY
This Technical Notes describes the components of masonry fireplaces and covers design and material selection.
Dimensions recommended for components of single-faced fireplaces are based on empirical data from field
performance of fireplaces and the expertise of the technical staff of the Brick Institute of America. The
recommendations contained herein will produce a functional and durable fireplace. The information and suggestions
contained in this Technical Notes are based on available data and the experience of the engineering staff of the Brick
Institute of America. The information contained herein must be used in conjunction with good technical judgment and
19 http://www.gobrick.com/BIA/technotes/t19.htm
13 of 14 9/13/2009 12:59 PM
a basic understanding of the properties of brick masonry. Final decisions on the use of the information contained in
this Technical Notes are not within the purview of the Brick Institute of America and must rest with the project
architect, engineer and owner.
REFERENCES
1. Chimneys, Fireplaces, Vents and Solid Fuel Burning Appliances, NFiPA211, National Fire Protection
Association, Quincy, MA, 1992.
2. Lytle, R.J. and Lytle, M.J., Book of Successful Fireplaces -How to Build, Decorate and Use Them, 20th
Edition, Structures Publishing Co., Farmington, MI, 1977.
3. Morstead, H. and Knudsen, O., Fireplace Report, a Guide for the Design and Construction of Fireplaces
and Chimneys, Alberta Masonry Institute, Calgary, Alberta, Canada.
4. Orten, V., The Forgotten Art of Building a Good Fireplace, Yankee, Dublin, NH, 1974.
5. Residential Masonry Fireplace and Chimney Handbook, Masonry Institute of America, Los Angeles, CA,
1989.
6. Shelton, J.W., The Measured Performance of Fireplaces and Fireplace Accessories, Williamstown, MA,
1978.
7. The Brick-O-Lator, Brick Association of North Carolina, Greensboro, NC, 1979.
19 http://www.gobrick.com/BIA/technotes/t19.htm
14 of 14 9/13/2009 12:59 PM

Technical Notes 19A - Residential Fireplaces, Details and Construction
Rev [May 1980] (Reissued August 2000)
Abstract: Brick masonry residential fireplaces can be made more energy efficient by providing a source of
combustion and draft air drawn from the exterior of the structure. Proper detailing and construction can also
contribute to the overall performance of the fireplace regarding both energy efficiency and structural integrity.
Building code requirements often control the configuration of the fireplace as well as component sizes.
Key Words: Bricks, combustion chamber, energy efficiency, firebrick, fireclay, fireplaces, hearths, masonry,
mortar.
INTRODUCTION
This Technical Notes contains recommended details and construction techniques which, when used to execute a
proper design, will yield a functional, energy-efficient fireplace. These same recommendations are applicable to
conventional fireplace construction when the provisions for the exterior air supply system are omitted.
This is the second in a series of Technical Notes dealing with the design and construction of fireplaces and
chimneys. Technical Notes 19 Revised contains a comprehensive discussion of fireplace design and materials
selection. Other Technical Notes in this series address design and construction of both residential and industrial
chimneys.
RECOMMENDATIONS
General
Energy-efficient fireplaces vary only slightly from conventional fireplaces. The recommendations to construct an
energy-efficient fireplace include properly sizing and locating an exterior air supply for combustion and draft air, and
tight-fitting dampers. Operation of the fireplace and other devices, such as glass screens, may also substantially
affect the performance of the fireplace. They are not, however, addressed in this Technical Notes. Regardless of
how much care is taken in the design and detailing process, workmanship remains a critical factor to the
performance of fireplaces. The designer should, therefore, be familiar with the fireplace construction techniques of
the locality in which the fireplace is to be built.
No matter which fireplace configuration is selected by the designer, there are several common features that should
be considered. Fireplaces, as discussed in this Technical Notes, are for burning wood and are not specifically
designed or constructed for fuels that generate temperatures in excess of those generated by combusting wood.
The primary function of the fireplace is to contain a fire safely and deliver heat to habitable spaces.
The fireplace assembly must be isolated from combustible materials. General requirements incorporated into many
building codes are: (1) All spaces between masonry fireplaces and wood or other combustible material should be
firestopped by placing 1 in. (25 mm) of noncombustible material in such spaces. (2) In the plane parallel to the front
wall of the fireplace, combustible material should not be placed within 6 in. (150 mm) of a fireplace opening. (3)
Combustible material within 12 in. (300 mm) of the fireplace opening should not project more than 1/8 in. (3.2 mm)
for each inch distance from the opening. A more specific discussion of these requirements or variations may be
provided by local building codes.
All void areas within the body of the fireplace from the foundation through the chimney should be solidly filled with
masonry mortared in place. The only exceptions are the air passageway, the ashpit, the 1-in. (25 mm) airspace
between the combustion chamber and the brickwork surrounding it, and the functional voids of the fireplace, such as
the smoke chamber.
Solidly filling the nonfunctional voids in the fireplace assembly increases its overall performance and durability as well
http://www.gobrick.com/BIA/technotes/t19a.htm
1 of 8 9/13/2009 1:01 PM
as its structural integrity and resistance to rain penetration. All exposed mortar joints should be properly tooled.
Concave jointing is preferred.
For a graphic definition of fireplace components and a section through an energy-efficient fireplace, refer to Fig. 1.
Single Face Fireplace Section
FIG. 1

FIREPLACE BASE ASSEMBLY
Foundation
The foundation supports the fireplace base and thus the entire fireplace and chimney assembly. It must, therefore,
be designed to carry these loads. However, most building codes disallow using the fireplace and chimney
assemblage as a structural element to support other building components. When designing the foundation, care
should be taken to account for soil condition and type. Undisturbed or well-compacted soil will generally be sufficient,
however, some types of soil or the condition of the soil may require additional analysis.
Building codes generally require that the foundation be at least 12 in. (300 mm) thick and, in plan view, extend a
minimum of 6 in. ( 150 mm) beyond every face of the masonry bearing on it. It should also penetrate the frost line to
reduce the possible "heaving" of the foundation, when the ground is frozen.
Exterior Air Supply System and Ashpit
There are many options to the construction methods and layout of the air passageway and ashpit discussed in this
section. When varying from the suggested details, keep in mind the function which these components serve, as well
as the manufacturer's recommendations for a specific device's installation.
The exterior air supply system is the component that is intended to increase the overall efficiency of the fireplace by
diminishing the amount of heated air drawn from the structure for combustion and draft. As discussed earlier in this
Technical Notes this provision may be deleted resulting in a conventional fireplace design.
http://www.gobrick.com/BIA/technotes/t19a.htm
2 of 8 9/13/2009 1:01 PM
The exterior air for combustion and draft may be drawn directly from the exterior or from unheated areas of the
building, such as crawl spaces. Local building codes may restrict the location of the air intake. For example, many
do not allow air from a garage to be vented into habitable spaces. This decreases the possibility of introducing
noxious gases from automobile exhausts into the house. When the fireplace configuration does not lend itself to
practical incorporation of the air passageway in the base, the intake may be located on any exterior wall. No matter
where the intake is located, it should be a screen-backed, closeable louver, preferably one that is operable from the
interior of the building.
The air passageway is generally incorporated into the base assembly. When this is not practical due to fireplace
configuration or the level of the exterior grade at the fireplace, the fireplace may be changed by raising the hearth or
the passageway may be formed of ductwork and attached to or incorporated into the floor system. In many
locations, the perimeter of the air passageway should be insulated, especially when the ductwork is adjacent to or
passing through heated areas. Many air passageway areas have been used successfully, usually varying from 6 to
60 sq in. (3870 to 38,700 mm
2
). The smaller areas may, however, yield high velocity air flow and more rapid
combustion, generally resulting in higher temperatures, which may produce negative results such as fireplace grate
or combustion chamber deterioration. Larger areas may also present a potential difficulty by delivering air in excess
of the combustion and draft requirements. The excess air, usually below room temperature, may be forced into the
room containing the fireplace. The volumetric expansion of this air due to the exterior to interior temperature
differential generally compounds the problem.
The air inlet should be located in the base or sidewalls of the combustion chamber. Inlets have performed
successfully, even when located in the rear wall of the combustion chamber. Care should be exercised when locating
the inlet in the combustion chamber walls since an air surge may force smoke and gases into the room. The air inlet
should be equipped with both directional and volume controls, so that the fire burns evenly and toward the rear of the
combustion chamber. Thus, the best performance is generally achieved when the inlet is located near the front of the
fireplace within the combustion chamber.
The air inlet damper assembly area, as shown in Fig. 2, generally ranges from 6 to 60 sq in. (3870 to 38,700 mm
2
)
depending on the other components in the exterior air supply system. The critical factor affecting the performance of
the fireplace is the proper operation of the air inlet damper. Opening the damper to a position that produces an area
of from 2 to 16 sq in. (1300 to 10,300 mm
2
) has proven to perform successfully in most areas.
Combustion Chamber Plan
FIG. 2
If corbeling is necessary to achieve proper size or location of the air inlet, it should be limited to a maximum
horizontal projection of one-half (1/2) the distance from the ashpit face to the exterior of the fireplace assembly, see
Fig. 3.
http://www.gobrick.com/BIA/technotes/t19a.htm
3 of 8 9/13/2009 1:01 PM
Corbeling Limitations
FIG. 3
The maximum projection for an individual unit should not exceed either one-half (1/2) the height of the unit, or
one-third (1/3) the bed depth.
The ash drop and ashpit are often not incorporated into the design due to either a configuration difficulty, such as
slab-on-grade construction, or the absence of a desire by the designer or owner to have such a component included
in the fireplace.
Figure 4 shows a plan view of an air passageway and ashpit. The ashpit cleanout door may be oriented toward the
interior of the building if sufficient space for cleanout exists. The ashpit cleanout door should be metal and fit tightly
to reduce air infiltration.
Hearth Support
When construction reaches the stage shown in Fig. 4, sturdy, noncombustible forming, such as metal, is set in place
to contain the slab pour. This is required since this forming is inaccessible for removal and is thus a permanent part
of the fireplace. In slab-on-grade construction, this requirement is not necessary unless a raised hearth is used. The
blackouts that form the opening in the slab for the ash drop and air inlet should be set so that they extend
approximately 1 in. (25 mm) above the top of the finished slab. This will facilitate removal of the forms. The slab
should be properly reinforced when it cantilevers from the fireplace wall to the floor system or spans across the air
passageway and ashpit, as shown in Fig. 1. This reinforcement is also beneficial in resisting the stresses induced by
the high temperatures the slab will be subjected to. Care should be taken to keep the top of the slab as nearly level
as possible to reduce the difficulty of laying the combustion chamber base.
http://www.gobrick.com/BIA/technotes/t19a.htm
4 of 8 9/13/2009 1:01 PM
Air Passageway and Ashpit Plan
FIG. 4
Similar results may be obtained by eliminating the concrete slab and supporting the hearth on masonry. This is
accomplished by solidly filling the base assembly with masonry mortared in place. The ash dump and air
passageway are formed by the judicious use of corbeling.
FIREBOX ASSEMBLY
Combustion Chamber and Firebox
Combustion chamber layout is critical since the chamber must be contained within the firebox assembly yet isolated
from it. Figure 2 shows one method of properly locating the combustion chamber. Once desired dimensions have
been selected from Table 1, in Technical Notes 19 Revised, the front wall (facing) of the fireplace is located and line
A-A struck at the inside face position. Next, locate Point I on this line. Point I corresponds to the centerline of the
combustion chamber in the direction perpendicular to the existing line. Squaring from Point I into the combustion
chamber for the chamber's depth, defines Point II. Striking a line connecting Points I and II results in Line B-B. From
Point II, squaring perpendiculars to line B-B in each direction for a distance of one-half (1/2) the rear chamber wall
dimension, thus locates Points c and d. On line A-A, measuring from Point I one-half (1/2) the fireplace opening
dimension in each direction, thereby defines points a and b. Connecting these four points (a, b, c and d) gives the
outline of the inside face of the combustion chamber.
Preferable combustion chamber construction consists of firebrick, in accordance with ASTM C 64 and fireclay
mortar, in accordance with ASTM C 105. Fireclay mortar joints should be 1/16 to 3/16 in. (1.6 to 4.8 mm) thick to
reduce thermal movements and mortar joint deterioration. When using fireclay mortar, extremely thin mortar joints
may be obtained by using the ''Pick and Dip'' method. This consists of dipping the unit into a soupy mix of fireclay
mortar and immediately placing it in its final position. The mortar joints need be only thick enough to provide for
dimensional irregularities in the unit being laid.
Acceptable construction includes the use of Grade SW brick, in accordance with ASTM C 62 or ASTM C 216, and
Type N or Type O. portland cement-lime mortar, in accordance with ASTM C 270 or BIA Designation M1-72. Type
N or Type O, portland cement-lime mortar may also be used with the firebrick option. Mortar joints should be limited
to 1/2 in. (12.7 mm) maximum and be properly tooled, resulting in a concave profile when using the portland
cement-lime mortar.
The brickwork surrounding the combustion chamber may be brought up, either at the same time as the combustion
chamber, or after it has been completed. In any case, there should be a full course of masonry surrounding the
combustion chamber, to allow for solidly filling the void created and maintaining a minimum 1-in. (25 mm) airspace
between the firebrick and surrounding brickwork, see Fig. 2. This airspace may be filled with a compressible,
http://www.gobrick.com/BIA/technotes/t19a.htm
5 of 8 9/13/2009 1:01 PM
noncombustible material, such as a fibrous insulation. The purpose of this material is to keep the space clear of
obstructions. Either the airspace or the compressible, noncombustible material reduces the stress from thermal
movements by isolating the combustion chamber.
The wall behind the firebrick at the rear of the firebox should be at least 8 in. (200 mm) thick. A greater thickness
may be required to support higher chimneys.
With the exception of the combustion chamber walls, wall ties should be used at all intersections where the wall is
not masonry bonded. These ties should be spaced a maximum of 16 in. (400 mm) vertically, and embedded at least
2 in. (50 mm) into bed joints of the brick masonry.
Horizontal joint reinforcement may also be beneficial, most especially at the corners of adjacent wythes and in the
wythe surrounding the combustion chamber walls. This precaution should help reduce cracking at these areas.
The combustion chamber walls should be firebrick. Firebrick may be laid with any face exposed, but they are
preferred as a stretcher course. However, if Grade SW brick are used, they should only be laid as a stretcher
course since they may not be as durable as the firebrick.
No fires should be built in the combustion chamber for thirty (30) days after construction. Fires before this time
period drive off the moisture necessary for proper curing of the mortar.
Lintels
When placing the lintel above the fireplace opening and the lintel above the damper, a compressible,
non-combustible material, such as insulation of a fibrous nature, should be placed at the end of the lintel where it is
embedded in masonry. This precaution is a means of dealing with the dissimilar expansion characteristics of
masonry and steel, which tend to induce stresses in the masonry, causing cracking.
The use of a lintel above the damper is highly recommended. his lintel is provided so that the masonry does not bear
directly on the metal damper which is subjected to extremely high temperatures, and high magnitude thermal
movements. All lintels used in the fireplace should bear on the brick masonry at least 4 in. (100 mm) at each end.
Damper
The damper is then seated using the same mortar that was used in the combustion chamber. This is accomplished
by spreading a mortar bed, just thick enough to ensure a level set of the damper and a seal that will prevent gas and
smoke leakage. The damper should not be embedded in mortar, but merely seated on the thin setting bed.
The damper assembly should only be in contact with masonry, on which it bears. To ensure this, once the damper
assembly is seated, it should be wrapped with a compressible, noncombustible material, such as fibrous insulation,
see Fig. 5. This material provides space for thermal expansion and movement of the damper during fireplace
operation.
http://www.gobrick.com/BIA/technotes/t19a.htm
6 of 8 9/13/2009 1:01 PM
Combustion Chamber and Smoke Chamber
FIG. 5a
Optional Smoke Shelf Configuration
FIG. 5b
SMOKE CHAMBER ASSEMBLY
Beginning at the level of the smoke shelf, the front and sides of the smoke chamber are corbeled in and the rear wall
is constructed vertically. This ensures total perimeter support for the flue liner. Corbeling limitations for this
component are determined by the fireplace configuration itself. The maximum corbel for each unit is the horizontal
distance to be corbeled divided by the number of courses from the bottom of the flue liner to the first corbeled
http://www.gobrick.com/BIA/technotes/t19a.htm
7 of 8 9/13/2009 1:01 PM
course. The usual limitations for corbeling walls are not applicable in this area of the fireplace since the corbels are
continuously laterally supported by adjacent masonry. The last two courses before the flue liner should be laid as
headers. These headers should be cut to a length that provides total perimeter support of the flue liner, without
obstructing the flue liner opening.
The smoke shelf may be a flat surface or curved, to assist flow through the smoke chamber, see Fig. 5a. The entire
smoke chamber should be parged. Care should be taken to ensure that the smoke shelf is kept free of mortar
tailings and debris for the same flow considerations. This may be accomplished by placing a material such as an
empty cement bag or plastic film on the smoke shelf during construction. When construction is completed, it can be
removed through the damper throat, bringing any foreign material with it.
SUMMARY
This Technical Notes has given suggested details and construction techniques for single-face residential fireplaces.
Other Technical Notes in this series address fireplace design, as well as residential and industrial chimney design
and construction. Emphasis has been placed on workmanship and proper construction methods.
It should be noted that all fireplace designs, no matter how sophisticated, are empirical and based on past
performance of specific configurations. Any variation from these configurations produces an "experimental" design.
While small deviations from the dimensions and proportions given may have little or no effect on performance, larger
magnitude changes should be carefully considered since they may have serious negative effects on the function of
the fireplace.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the Brick Institute of America's technical staff.
This information should be recognized as recommendations and suggestions for consideration by the designers,
specifiers, and owners of buildings when anticipating the design, detailing and construction of single-face residential
fireplaces. The final decision to use or not to use these recommendations and types of products in brick masonry
fireplaces is not within the purview of the Brick Institute of America and must rest with the project designer or owner.
REFERENCES
1. One and Two Family Dwelling Code, published by Interstate Printers and Publishers, Danville, Illinois,
1975
2. Book of Successful Fireplaces, How to Build, Decorate and Use Them, 20th Edition, by R. J. Lytle and
Marie-Jeanne Lytle, Structures Publishing Company, Farmington, Michigan, 1977.
3. How to Install A Fireplace, by Donald R. Brann, Direction Simplified, Inc., Briarcliff Manor, New York.
1976.
http://www.gobrick.com/BIA/technotes/t19a.htm
8 of 8 9/13/2009 1:01 PM

Technical Notes 19B - Residential Chimneys - Design and Construction
Rev [June 1980] (Reissued Apr. 1998)
Abstract: All residential chimneys. both for fireplaces and appliances, are designed and constructed to serve the
same basic functions. They must provide fire protection and safely convey combustion by-products to the exterior of
the structure at a rate that does not adversely affect the combustion process. Design, materials selection,
construction, and building code requirements all have a significant impact on the chimney's potential to fulfill these
functions. Chimney height and flue area are the two most critical factors in chimney desire.
Key Words: bricks, building codes, chimneys, draft, flashing, flues, masonry, mortar.
INTRODUCTION
This Technical Notes addresses the design and construction of residential chimneys. Other Technical Notes in this
Series deal with residential fireplaces and commercial chimneys. The design of residential chimneys is empirical and
based on successful prototypes. The function of residential chimneys is to allow combustion by-products to be
conducted away from the structure safely.
GENERAL
Residential chimneys generally fall into two categories: 1) chimneys serving fireplaces, and 2) chimneys serving
appliances. While there are dissimilarities between the two types, they both serve the same basic functions. It is
worthwhile, therefore, to consider their similarities. Both are constructed of similar materials and must meet the
same building code requirements. Even though they may convey different combustion by-products at different
velocities, they both must be designed and constructed to discharge these by-products at a rate that does not
adversely affect the combustion process and to release the discharged material at a height and location that
provides fire safety.
Flues may slope to join with other flues so as to discharge through a common flue, or to achieve the desired location
of the chimney. The maximum allowable slope is 30 deg from vertical. When combining flues the main discharge flue
should be sized for the maximum combined flow from the smaller flues. Combining flues of dissimilar systems or
fuels. i.e., appliances and fireplaces, is not allowed by many building codes. Separate flues may be incorporated into
one chimney so long as minimum wall thickness requirements are met and a full wythe of brick is laid between them
and bonded to the chimney walls.
Building Code Requirements
Building code requirements for chimneys may vary on a local basis. There are, however, several that are accepted
nearly everywhere. They include:
1. Chimney wall thickness should be a nominal 4 in. (100 mm) unless no flue liner is used, in which case a nominal 8
in. (200 mm) is required.
2. Neither chimney nor flue liner may change size or shape within 6 in. (150 mm) of either floor components, ceiling
components or rafters.
3. The minimum chimney height for fire safety is the greater of 3 ft (1.0 m) above the highest point where the
chimney penetrates the roofline, or 2 ft (600 mm) higher than any portion of the structure or adjoining structures
within 10 ft (3.0 m) of the chimney, see Fig. 1.
4. Chimney clearance from combustible material is a minimum of 2 in. (50 mm) except where the chimney is located
19b http://www.gobrick.com/BIA/technotes/t19b.htm
1 of 16 9/13/2009 1:02 PM
entirely outside the structure, in which case 1 in. (25 mm) is acceptable.
5. The spaces between a chimney and combustible material should be firestopped using a minimum of 1-in. (25 mm)
thick noncombustible material.
6. All exterior spaces between the chimney and adjacent components should be sealed. This is most commonly
accomplished by flashing and caulking.
7. Masonry chimneys should not be corbeled more than 6 in. (150 mm) from a wall or foundation nor should a
chimney be corbeled from a wall or foundation which is less than 12 in. (300 mm) in thickness unless it projects
equally on each side of the wall, except that on the second story of two-story dwellings corbeling of chimneys or the
exterior of the enclosing walls may equal the wall thickness. Corbeling may not exceed 1-in. (25 mm) projection for
each course of brick protected.

Building Code Dimension Requirements
FIG. 1
Recommendations
In many situations it may be desirable to use the chimney as a structural element. This may be accomplished within
most building codes by maintaining the chimney wall thickness and adding a structural wall around the chimney. This
structural wall may be built integrally with the chimney wall. Most building codes require a minimum of 4 in. (100 mm)
of bearing. Considering all the building code dimensional requirements, the minimum wall thickness of a lined chimney
to be used as a structural component is 10 in. (250 mm) consisting of: 1) 4-in. (100 mm) chimney wall (brick), 2)
2-in. (50 mm) of noncombustible material (brick), and 3) 4-in. (100 mm) bearing length (brick). An unlined chimney's
minimum wall thickness is 14-in. (350 mm) consisting of the same elements as the lined chimney except that the
chimney wall must be 8-in. (200 mm), see Fig. 2.
19b http://www.gobrick.com/BIA/technotes/t19b.htm
2 of 16 9/13/2009 1:02 PM
Chimney Used as Structural Support
FIG. 2
MATERIALS
Brick
The chimney, by the nature of its function, is at least partially exposed to weathering. The brick should conform to
ASTM C 216, Grade SW, or ASTM C 62, Grade SW, to assure sufficient durability. Paving brick should conform to
ASTM C 902, Class SX.
Mortar
To allow for both weathering and thermal considerations, Type N portland cement-lime mortar is recommended for
the chimney. Type S portland cement-lime mortar is acceptable, and may be necessary when the chimney is
subjected to high lateral forces such as wind loads in excess of 25 psf (1.2 kPa) or seismic loads. Where the
chimney is in contact with earth, Type M portland cement-lime mortar is recommended. The mortar used to bed the
flue liners should be able to perform well under high temperatures. Therefore, fireclay mortars are highly
recommended. Type N portland cement-lime mortar is an acceptable substitute. For a comprehensive discussion of
portland cement-lime mortar types and uses, see Technical Notes 8 Series.
Flue Liners
Flue liners should conform to ASTM C 315. They should be thoroughly inspected just prior to installation for cracks
or other damage that might contribute to smoke and flue gas leakage.
Flashing
Corrosion-resistant sheet metal flashing is required by most building codes. Quality materials should be specified
since replacement may be expensive and troublesome. See Technical Notes 7A Revised for selection of flashing
materials.
Chimney Caps
A prefabricated chimney cap similar to the one shown in Fig. 3 should be used. This type cap provides better
durability and is more easily made water-resistant than a cast-in-place cap. When a cast-in-place cap is used, it
should incorporate the same shape as the prefabricated. The thickened sides and overhangs will reduce the
potential for water penetration.
Chimney Cap Detail
FIG. 3
Rain Caps
Rain caps vary from sophisticated turbine type metal caps to simple slabs set above the termination point of the flue
19b http://www.gobrick.com/BIA/technotes/t19b.htm
3 of 16 9/13/2009 1:02 PM
liner. When specifying a manufactured rain cap, information regarding its effect on the gas flow through the chimney
should be obtained from the manufacturer. If the cap is metal, it should be corrosion-resistant.
Sealants
Caulking is frequently considered a means of correcting or hiding poor workmanship, rather than as an integral part
of construction. It should be detailed and installed with the same care as the other elements of the structure. In all
cases, the use of a good grade, polysulfide, butyl, or silicone rubber sealant is recommended. Oil-based sealants
should not be used. Regardless of the sealant used, proper priming and backing rope, are a must.
Ties and Reinforcement
Ties used in chimney construction should be corrosion-resistant metal ties. For a general discussion of ties and their
placement, refer to Technical Notes 28 Revised.
Reinforcing steel should conform to one of the following ASTM Standards:
1. Welded Wire-ASTM A 185
2. Steel Bar-ASTM A 615, ASTM A 616 or ASTM A 617
3. Wire-ASTM A 82
DESIGN
Design of fireplace and appliance chimneys is limited to the determination of height requirements that when used in
conjunction with proper flue sizes, detailing and construction will provide adequate draft. Building code requirements
for minimum chimney height remain in effect and must be met or exceeded.
Fireplace Chimneys
The design of residential fireplace chimneys is directly related to: 1) the area of the fireplace opening, 2) the area of
the flue liner, and 3) the height of the chimney. In most situations, the area of the fireplace opening is controlled by
considerations other than the performance of the system, such as aesthetics. The other components of the system
are usually designed based upon the desired fireplace opening.
A frontal face velocity of 0.80 ft per second (0.245 m/sec) at the fireplace opening has been accepted by the
American Society of Heating, Refrigerating and Air-Conditioning Engineers (ASHRAE) to be sufficient to prevent
smoke and gases from being discharged into habitable spaces. This is a minimum velocity and usually only
encountered while starting a fire. Flue liner size as a function of fireplace opening size may be obtained from
Technical Notes 19 Revised, Table 1.

With the opening and flue sizes known, Equation 4 may be used to calculate the minimum chimney height to provide
adequate draft in a properly designed, detailed and constructed assembly. The height calculated using Equation 4 is
19b http://www.gobrick.com/BIA/technotes/t19b.htm
4 of 16 9/13/2009 1:02 PM
measured from the top of the fireplace opening, and is the minimum required to produce an adequate draft. Building
code requirements previously discussed for minimum chimney heights are based solely on fire safety considerations
and must always be met or exceeded.
Procedure
Step 1. From Table 1 in Technical Notes 19 Revised select fireplace opening dimensions A and B and the
corresponding flue liner size, L and M. Using Equation 1 calculate the Fireplace Opening Area, Ao. From Table 1
determine the Minimum Flue Area, AF, using L and M.

Step 2. The Flue Friction Coefficient, KT must now be determined. The friction loss due to the acceleration of
ambient air to Flue Gas Velocity, K1 is always equal to 1.0 in residential applications.
Based on the size of the flue selected, a preliminary damper size must be assumed. At this point in the design
process, it is only necessary to decide if the damper throat area to be used will be equal to or twice the flue area.
Once this decision has been made the Inlet Loss Coefficient, K2, may be determined. If the damper throat area is
equal to the flue area, K2 is equal to 2.5. If the damper throat is twice the flue area, K2 is equal to 1.0.
At this time decisions concerning general fireplace configuration must be made. The designer must determine
whether or not to use a rain cap and if so, at what distance above the chimney termination point it will be placed.
The Termination Coefficient, K3, may be selected using this information. If no rain cap is used K3 equals 0.0. If a rain
cap is set at a distance of D/2 (see Table 1 for equivalent diameter, D) above the termination point of the flue liner,
K3 may vary from 0.0 to 4.0. This information may be obtained from the manufacturer.
Having determined K1 K2, and K3, the Flue Friction Coefficient, KT, may be calculated using Equation 2.

Step 3. From Table 1 determine the Minimum Flue Area, AF, using L and M.


Step 2. The Flue Friction Coefficient, KT must now be determined. The friction loss due to the acceleration of
ambient air to Flue Gas Velocity, K1 is always equal to 1.0 in residential applications.
Based on the size of the flue selected, a preliminary damper size must be assumed. At this point in the design
process, it is only necessary to decide if the damper throat area to be used will be equal to or twice the flue area.
Once this decision has been made the Inlet Loss Coefficient, K2, may be determined. If the damper throat area is
equal to the flue area, K2 is equal to 2.5. If the damper throat is twice the flue area, K2 is equal to 1.0.

At this time decisions concerning general fireplace configuration must be made. The designer must determine
whether or not to use a rain cap and if so, at what distance above the chimney termination point it will be placed.
19b http://www.gobrick.com/BIA/technotes/t19b.htm
5 of 16 9/13/2009 1:02 PM
The Termination Coefficient, K3, may be selected using this information. If no rain cap is used K3 equals 0.0. If a rain
cap is set at a distance of D/2 (see Table 1 for equivalent diameter, D) above the termination point of the flue liner,
K3 may vary from 0.0 to 4.0. This information may be obtained from the manufacturer.
Having determined K1 K2, and K3, the Flue Friction Coefficient, KT, may be calculated using Equation 2.


Step 3. From Table 1 determine the Inside Perimeter, PI, of the flue liner previously selected and using Equation 3
calculate the Hydraulic Radius, RH.


Step 4. The general design equation for chimneys with rectangular flues (Equation 4) may now be used to calculate
the Minimum Height, H. to produce adequate draft.

This is the height of the chimney from the lintel above the fireplace opening. To obtain the height from the combustion
chamber floor, add the opening height. Building code requirements for minimum chimney heights remain in effect.
The greater of the values obtained from calculations and building code requirements should be used.


Example
Step 1. From Table 1 in Technical Notes 19 Revised fireplace opening dimensions of 30 in. x 29 in. and a
corresponding flue liner size of 12 in. x 12 in. were selected. Using Equation 1 calculate Ao.


From Table 1

19b http://www.gobrick.com/BIA/technotes/t19b.htm
6 of 16 9/13/2009 1:02 PM

Step 2. Assuming that there is no rain cap and that the damper throat area is twice the flue area, Equation 2 may be
utilized.


Step 3. From Table 1 determine PI and solve Equation 3 for the Hydraulic Radius.


Step 4. Substitute these values into Equation 4 and calculate the Minimum Height, H, to provide adequate draft.


Appliance Chimneys

Appliance chimneys are divided into two types those venting one appliance, see Fig. 4, and those venting two or
more appliances, see Fig. 5. The two variables that are most commonly known to the designer are the input rating
and configuration of the system. Typical design criteria are shown in Tables 2 and 3. Building code requirements for
chimney heights should be considered as minimum heights for fire safety and should be strictly adhered to.

19b http://www.gobrick.com/BIA/technotes/t19b.htm
7 of 16 9/13/2009 1:02 PM

Masonry Chimney Serving a Single Appliance (See Table 2)
FIG. 4


19b http://www.gobrick.com/BIA/technotes/t19b.htm
8 of 16 9/13/2009 1:02 PM

Masonry Chimney Serving Two or More Appliances (See Table 3)
FIG. 5


19b http://www.gobrick.com/BIA/technotes/t19b.htm
9 of 16 9/13/2009 1:02 PM

aSI conversions: W = Btu/h x 0.293; m = ft x 0.3048; mm = in. x 25.4; mm 2 = in. 2 x 645
bKnot recommended.

19b http://www.gobrick.com/BIA/technotes/t19b.htm
10 of 16 9/13/2009 1:02 PM

aSI conversions: W = Btu/h x 0.293; m = ft x 0.3048; mm = in. x 25.4.
bNot recommended.

CONSTRUCTION AND DETAILS
General

Since both fireplace and appliance chimneys have an identical function, their construction methods and materials are
similar. Building code requirements insofar as construction is concerned are identical.

Fireplace Chimneys
General. The chimney of a fireplace is considered to be that portion of the fireplace from the base of the first flue
liner to the top of the last flue liner, or any rain cap above it.
Single-wythe chimneys should be attached to the structure. This is generally accomplished by using corrosion-
resistant metal ties spaced at a maximum of 24 in. (600 mm) on center. Multi-wythe chimneys that are not masonry
bonded should be bonded together using metal wire ties.
Racking. Chimneys are generally not as wide as the body of the fireplace below. When racking back to achieve the
desired dimensions or location of the chimney care must be exercised to insure that, since there is no limitation on
the distance each unit may be racked, cores of the units are not exposed. Preferred construction consists of a
setting bed over the racked face with uncored or paving brick set to provide a weather resistant surface. Mortar
washes may also be used. They may not, however, be as durable. When using a mortar wash it should not bridge
over the rack, but should fill each step individually. Both methods of racking are shown in Fig. 6.

19b http://www.gobrick.com/BIA/technotes/t19b.htm
11 of 16 9/13/2009 1:02 PM

Racking
FIG. 6


Flue Liners. The first flue liner should be supported along its entire perimeter by masonry. The liner should be
bedded in mortar with the joints cut flush and smoothed on the interior and the exterior joint area parged. The flue
liners should be set one section ahead of the chimney brickwork.

Flashing. Base flashing and counter flashing are installed at the chimney/roof interface, see Fig. 7. The base
flashing is installed first on the faces of the chimney perpendicular to the ridgeline with tabs at each corner. The
flashing should extend a minimum of 4 in. (100 mm) up the face of the chimney and along the roof. Counter flashing
is then installed over the base flashing. It is inserted into a mortar joint for 3/4 to 1 in. (19.1 mm to 25 mm) and
mortared solidly into the joint. The counter flashing should lap the base flashing by at least 3 in. (75 mm). If the
flashing is installed in sections, the flashing higher up the roofline should lap over the lower flashing a minimum of 2
in. (50 mm). All joints in the base flashing and counter flashing should be thoroughly sealed. The unexposed side of
any bends in the flashing should also be sealed.

19b http://www.gobrick.com/BIA/technotes/t19b.htm
12 of 16 9/13/2009 1:02 PM

Typical Section and Flashing Detail
FIG. 7


Cricket. If a cricket is desired, usually for chimneys whose dimension parallel to the ridgeline is greater than 30 in.
(750 mm) and do not intersect the ridgeline, it should be constructed similar to the one shown in Fig. 8. The
dimensions of the cricket are based on the chimney measurements parallel to the ridgeline. The intersection of the
cricket and the chimney should be flashed and counter flashed in the same manner as a normal chimney roof
intersection. The flashing at the roofline should extend to at least 4 in. (100 mm) under the roofing material. For
dimensions and construction details, see Table 4, and Fig. 9.


Typical Flashing Detail
FIG. 8




19b http://www.gobrick.com/BIA/technotes/t19b.htm
13 of 16 9/13/2009 1:02 PM

Typical Chimney Cricket Framing
FIG. 9

Chimney Caps. There are, as discussed in the materials section, two options regarding chimney caps: 1)
prefabricated, and 2) cast-in-place. Prefabricated caps generally provide superior performance as compared to the
cast-in-place type. Regardless of which type cap is used, it should be thoroughly primed, backed, and sealed at the
cap and flue liner interface to reduce the potential for water penetration.

Prefabricated caps are set in place on a mortar bed. There should be a bond break between the brickwork and the
setting bed to allow the cap to respond to the differential movement it will encounter without distressing the
brickwork. Figure 3 depicts a typical prefabricated cap. From this figure, general configurations and waterproofing
methods may be obtained.
Cast-in-place caps should conform to the shape and minimum dimensions shown in Fig. 3. Feathering the cap to the
edge should be avoided since this substantially reduces the thickness at the edge and therefore the potential for
deterioration is increased. Waterproofing requirements are different since shrinkage of the concrete as it cures is a
certainty. Flashing is highly recommended for cast-in-place caps. The flashing may also be considered as the bond
break material. Adequate reinforcement should be placed in the cap to help control cracking due to shrinkage and
thermal movements. Additional reinforcement may be necessary in the portion of the cap that overhangs the face of
the chimney. Figure 3 shows one method of forming a cast-in-place chimney cap.
When using a chimney cap that does not overhang the face of the chimney, the last two courses of the chimney
brickwork should be corbeled out to form a drip to help reduce the amount of water allowed to run down the face of
the chimney. The flue liner should extend a minimum of 2 in. (50 mm) above the top of the cap, see Fig. 3.

19b http://www.gobrick.com/BIA/technotes/t19b.htm
14 of 16 9/13/2009 1:02 PM
Appliance Chimneys
General. Fireplace and appliance chimneys have few dissimilarities. The general recommendations for the
construction of fireplace chimneys and the proper consideration of three additional components should produce a
functional appliance chimney. The three components, either not present in fireplace chimneys or incorporated into the
body of the fireplace are: 1) the foundation, 2) the cleanout door, and 3) the thimble.
Foundation. The foundation supports the chimney and must be sized to carry all superimposed loads. However,
most building codes disallow using the chimney walls as structural elements to support other building components.
When designing the foundation, care should be taken to account for soil conditions and type. Undisturbed or
well-compacted soil will generally be sufficient, however, some types of soil conditions may require additional
analysis.
Building codes generally require that the foundation be at least 12 in. (300 mm) thick, and, in plan view, extend a
minimum of 6 in. (150 mm) beyond each face of the masonry bearing on it. It should also penetrate the frost line to
reduce the possibility of "heaving" of the foundation while the ground is freezing.
Cleanout Door. A cleanout door may not be necessary when venting appliances that use clean burning fuels such
as natural gas, however other fuels may produce combustion by-products that will accumulate at the bottom of the
chimney and require periodic removal. The cleanout door should be of ferrous metal and set to provide as airtight a
seal as possible. If desired, the cleanout door may be oriented toward the interior of the structure, however. the
prime consideration in sizing and locating the door is the ease with which it can be used.
Thimble. A thimble is the lined opening through the chimney wall that receives the smoke pipe connector, as shown
in Fig. 10. A thimble should be set in the chimney at the location of the entrance of the pipe connector. It should be
built integrally with the chimney and made as airtight as possible, by using either boiler putty or asbestos cement.
The thimble should be set flush with the interior face of the flue liners, and at least 18 in. (460 mm) below the ceiling.
The thimble should have a minimum of 8 in. (200 mm) of flue liner extending below its lowest point, see Fig. 10.


Thimble Detail
FIG. 10


19b http://www.gobrick.com/BIA/technotes/t19b.htm
15 of 16 9/13/2009 1:02 PM

SUMMARY

This Technical Notes has given suggested design and construction methods for residential chimneys. Although there
are differences. both appliance and fireplace chimneys use similar construction techniques and materials. Since the
prime function of a chimney is fire safety both quality workmanship and materials should be used.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the Brick Institute of America's technical staff. The recommendations and suggestions are offered as a
guide for consideration by the designers, specifiers, and owners of buildings when anticipating the design, detailing
and construction of residential chimneys. The final decision to use or not to use these recommendations and
materials in brick masonry chimneys is not within the purview of the Brick Institute of America. and must rest with the
project designer, or owner.

REFERENCES

1. One and Two Family Dwelling Code, published by Building Officials and Code Administrators, Inc.,
Homewood, Illinois; International Conference of Building Officials, Whittier, California; and Southern Building
Code Congress, International, Inc., Birmingham, Alabama.
2. Book of Successful Fireplaces, How to Build, Decorate and Use Them, 20th Edition, by R. J. Lytle and
Marie-Jeanne Lytle, Structures Publishing Company, Farmington, Michigan, 1977.
3. How to Install a Fireplace, by Donald R. Brann, Direction Simplified, Inc., Briarcliff Manor, New York,
1976.
4. 1979 Equipment Volume, ASHRAE Handbook and Product Directory, by American Society of Heating,
Refrigerating and Air-Conditioning Engineers, Inc., New York, New York, 1979.
19b http://www.gobrick.com/BIA/technotes/t19b.htm
16 of 16 9/13/2009 1:02 PM

Technical Notes 19D - Brick Masonry Fireplaces, Part 1, Russian-Style Heaters
[Jan. 1983] (Reissued June 1987)
Abstract: Brick masonry heaters may be used instead of conventional fireplaces to provide efficient supplemental
heating for residential buildings. The design, detailing and construction of brick masonry fireplaces with baffle
systems for combustion gases are discussed. Information regarding building code compliance, operation and the
accessories required is presented with the basic principles by which these heaters provide supplemental heat for
buildings.
Key Words: brick, buildings (codes), design, energy, fireplace, heating, masonry, mortar.
INTRODUCTION
There are many ways of improving the energy efficiency of fireplaces. Fireplace energy efficiency may be increased
by providing glass screens or exterior air for draft and combustion; altering the shape of the firebox for increased
radiant heating; incorporating baffles within the mass of the fireplace through which room air may be circulated for
increased convective heating; or combinations of these features. These features, along with the proper design,
construction and operation are discussed in Technical Notes 19 Revised, 19A Revised, and 19C, and are used to
increase the efficiency of the wood-burning fireplace for heating the building or room. By altering the design of the
fireplace to change the firebox shape and to replace the smoke chamber with a baffle system, greater heating and
wood combustion efficiencies may be achieved. These alterations in the fireplace design result in a baffled brick
masonry fireplace or brick masonry heater.
The brick masonry heater is a concept that has been used for centuries in Northern and Eastern Europe. Various
styles of brick masonry heaters are often referred to as "Finnish" or "Russian" stoves, although they are used in
many other countries such as Belgium, Germany, Switzerland, Sweden, The Netherlands and Norway. The basic
principles used to obtain the high heating and combustion efficiencies are: 1) controlled air intake to the combustion
chamber or firebox; and 2) a baffle system through which hot combustion gases are circulated. The combustion
gases circulated through the baffle system heat the walls of the heater, which in turn heat the room. The basic
concepts of design, construction and operation are simple, but there are several concerns which must be addressed
to insure safety and durability. Two basic designs of brick masonry heaters are discussed in this Technical Notes:
the "Russian Stove" with a horizontal baffle system, and the "Russian Stove" with a vertical baffle system. Other
Technical Notes in this Series address the "Finnish" or "Fountain Style" brick masonry heater and modifications
which may be applied to result in contemporary designs of brick masonry heaters.
These heaters are often referred to as stoves because as originally designed and built, they had provisions which
permitted portions of the heater to be used for baking and cooking. This was done by circulating the hot combustion
gases through baffles surrounding a brick oven and under exposed metal plates which were used for cooking. Such
provisions are not practical for modern lifestyles because of the difficulty in controlling temperatures, thus the baking
and cooking features are not addressed in this Technical Notes. However, the cooking features may be a
consideration for designing and construction of an outside barbecue or cookstove. (See Reference 5).
GENERAL
Operation
Operation of a brick masonry heater is simple. The firebox is loaded with about 20 lb (9.1 kg) of wood. The fire is
ignited. Once good combustion of the wood begins, the firebox door is closed and the air intakes are adjusted to the
proper setting so that good combustion is maintained. The combustion gases exhaust from the rear of the firebox,
circulated through the baffle system, warm the entire mass of the brick masonry heater and then are exhausted to
the exterior of the building through a conventional chimney constructed on top of the heater. The fire usually burns
about 30 min when properly seasoned wood is used, and thus the firebox should be reloaded with 20 lb (9.1 kg) of
http://www.gobrick.com/BIA/technotes/t19d.htm
1 of 15 9/13/2009 1:04 PM
wood four times in a 2-hr period for maximum heating. This procedure usually results in the brick masonry heater
being sufficiently heated to keep a room, about 2400 ft
3
(68 m
3
) warm for 8 to 12 hr. The operation will vary slightly,
depending upon the size of the heater, the size of the room and the amount of heat needed to be comfortable.
During the coldest months in severe climates such as those of Scandinavia, the heater usually needs to be operated
twice a day, once in the morning and once in the evening. During the more moderate seasons, or in moderate
climates, the heater may need to be operated only once, usually in the early evening, to supplement the heating
requirements to maintain comfortable temperatures throughout the night. Other variations in operation, such as the
number of loadings or the amount of wood used per loading during each firing, may also result in increased comfort.
The operator should experiment with several variations of operation to determine the best performance for various
seasons.
Properly operated, these brick masonry heaters are very effective in supplying radiant heat to the area of the
building surrounding the heater. These heaters are not only good sources of heat, but have wood-burning efficiencies
of 80 to 90 percent. (See Reference 3).
Building Code Requirements
There are no major model building code requirements which specifically address brick masonry heaters. For the
most part, the building code requirements for fireplaces and chimneys are applicable to brick masonry heaters
except for those requirements which address the dimensioning of the firebox and smoke chamber. All building code
chimney requirements and clearances for combustibles are applicable.
There are, however, two concerns regarding safety which apply to the brick masonry heaters that are not presently
listed in the major model building codes. The first concern is the integrity of the heater's enclosing walls. The walls
forming the shell of the brick masonry heater should be at least two wythes of brick thick so that major cracks do not
occur in the brick masonry heater. It is advisable to provide a nominal 1-in. (25 mm) air space between the two
wythes. This will prevent cracks from penetrating through the interior to the exterior of the heater. Filling this 1-in. (25
mm) air space with compressible, non-combustible material, such as fiberglass insulation, will insure this separation.
The two wythes should be tied together with corrosion-resistant metal ties. The insulation used to maintain a
compressible space between the two wythes of brick should not affect the overall thermal performance of the brick
masonry heater. To add to the integrity of the exterior wythe of the brick masonry heater, horizontal joint
reinforcement should be placed in about every sixth course. Horizontal joint reinforcement should not be used on the
interior wythe because the extreme differential thermal movements may deteriorate the mortar joints.
The second concern is the temperature of the exterior surfaces of the brick masonry heater walls. The highest
surface temperatures of the walls are normally between 100
o
F (38
o
C) and 130
o
F (54
o
C), but temperatures as high
as 190
o
F (88
o
C) have been reported. Although these temperatures are much lower than those achieved with metal
wood-burning stoves, sufficient clearances to combustibles should be maintained. A minimum 36-in. (900 mm)
clearance is usually required between metal wood-burning stoves and combustibles. A minimum 12-in. (300 mm)
clearance is recommended between the sides and back of the brick masonry heater and combustibles. At the floor
line, this may be achieved by providing a 12-in. (300 mm) extended hearth. In front of the heater, a 20-in. (500 mm)
extended hearth should be used. This is easily achieved when the heater is properly positioned in the room for
maximum heating. This position is in the center of the room so that all four walls of the heater are providing radiant
heat to the room. The brick masonry heater may also be installed against interior brick masonry walls.
DESIGN AND CONSTRUCTION
General
The brick masonry heater should always be positioned entirely inside the building. It should never be located on an
exterior wall. When incorporated into an exterior wall, much of the radiant heat being supplied will be lost to the
exterior. In addition, the location on an exterior wall will usually result in at least one cold surface on which a
considerable amount of creosote may form. A creosote fire can reach temperatures which could result in cracks
within the heater that may become too large to allow safe operation because combustion gases may leak into the
room.
The actual dimensions of the masonry heater are limited by the available firebox door sizes, the number of baffles,
and the height of the heater. For best performance, the heaters should not be more than one story in height, nor
should they contain more than five baffle chambers. This is because the increased distances that the hot combustion
http://www.gobrick.com/BIA/technotes/t19d.htm
2 of 15 9/13/2009 1:04 PM
gases must flow will result in a cooling of the gases. This causes a reduction in their heating capacity and could
result in increased creosote deposits which may lead to potential fires.
There are two types of Russian-style brick masonry heaters: the vertically baffled heater shown in Figures 1 and 2,
and the horizontally baffled heater shown in Figures 3 and 4. Many of the features are similar for both.
Brick Masonry Heater With Vertical Baffles
FIG. 1
http://www.gobrick.com/BIA/technotes/t19d.htm
3 of 15 9/13/2009 1:04 PM
Brick Masonry Heater With Vertical Baffles
FIG. 2
http://www.gobrick.com/BIA/technotes/t19d.htm
4 of 15 9/13/2009 1:04 PM
Brick Masonry Heater With Horizontal Baffles
FIG. 3
http://www.gobrick.com/BIA/technotes/t19d.htm
5 of 15 9/13/2009 1:04 PM
Brick Masonry Heater With Horizontal Baffles
FIG. 4
The most critical factors for the proper performance of brick masonry heaters are the gas flow through the baffle
system and the draft of the chimney. All openings between the baffles and the area enclosed by the baffles should
be at least 64 sq in. (40,000 mm
2
). The chimney should be constructed with clay flue liners with two wythes of brick
surrounding the flue liners in such a way as to maintain a nominal 1-in. (25 mm) air space between the flue liners and
the interior wythe of brick.
Clearances between the brick masonry heater exterior walls and any combustible materials should be at least 12 in.
(300 mm). This will require the masonry heater to be either centrally located within a room, or located adjacent to a
brick masonry or other non-combustible interior wall with at least a 1-hr fire rating. To ensure clearances, it may be
beneficial to provide an extended hearth around the entire masonry heater. Construction using the proper clearances
also maximizes the use of the warmed brick masonry surface as radiant heat sources.
Base Assembly
The base assembly includes the foundation, extended hearth, and ash drop. These features are the same for both
the vertically and horizontally baffled Russian-style heaters.
Foundation. The foundation must be adequate to support the mass of the brick masonry heater and the masonry
chimney. When designing the foundation, care should be taken to account for soil types and foundation conditions.
Undisturbed or well-compacted soil will generally be sufficient, however, some types of soil or foundation conditions
may require additional analysis, which may result in the need for special soil treatment or a unique foundation design.
Building codes generally require that the foundation be at least 12 in. (300 mm) thick, and in plan view, extend a
minimum of 6 in. (150 mm) beyond each face of the masonry heater. The foundation should be positioned so that the
bottom of the footing is below the frost line to reduce the possibility of "heaving."
http://www.gobrick.com/BIA/technotes/t19d.htm
6 of 15 9/13/2009 1:04 PM
Unless the foundation is a thickened slab in a newly constructed slab-on-grade structure, masonry is usually used to
construct the base assembly to the height of the hearth support. The hearth support may be solid masonry
construction carried up from the foundation to support the entire hearth area. To conserve materials, the masonry is
usually brought up only equal to the dimensions of the masonry heater itself and the brick masonry of the base
assembly is corbeled out to form the support for the extended hearth. See Figures 1 through 4. An individual corbel
should not exceed one-half the unit height nor one-third of the unit thickness. The total projections of the corbel
should not exceed one-half the thickness of the base assembly, nor one-half of the thickness of the solid masonry
wall forming the base assembly. A further discussion of corbeling is provided in Technical Notes 19A Revised.
Extended Hearth. The extended hearth may be formed by placing a reinforced concrete slab on top of the corbeled
base assembly. Non-combustible or removable forming should be placed so that it spans from the corbeled masonry
assembly to the floor joists forming the opening for the brick masonry heater. Double joists should be used around
the entire perimeter of the opening with a nailer to support the edge of the non-combustible extended hearth, as
discussed in Technical Notes 19A Revised, except that the extended hearth for the brick masonry heater should be
at least 20 in. (500 mm) in front of the firebox and 12 in. (300 mm) around the remaining perimeter of the heater.
However, the extended hearth may be eliminated on one side, or the back of the heater if it is positioned against a
non-combustible wall with a minimum fire rating of 1 hr. Once the reinforced concrete slab is installed, it may be
finished with brick masonry pavers. If removable forming is used, the concrete slab used to support the extended
hearth must be designed as a cantilever.
Ashpit. Once the extended hearth is installed, the brick masonry heater is laid out. The dimensions of the brick
masonry heater are determined by the size of the available firebox doors and the number of baffles used in the
masonry heater. The masonry heater, because of its efficient combustion of wood, does not require a large ashpit.
The ashpit is usually formed by providing an opening, usually three courses of brick in height, and as wide as the
firebox. This results in an ashpit which may be accessed from the front face of the brick masonry heater. The ashpit
may be modified so that it is accessed from either side, or even from the rear of the masonry heater. Directly above
the ashpit will be the base for the brick masonry heater firebox. Thus, the ashpit should be formed with corbeled
brick masonry which will support the firebox base or a reinforced concrete slab may be used to serve as the top of
the ashpit and the support for the firebox base. Using the reinforced concrete slab requires a formed opening in the
slab for the ash drop.
Firebox Assembly
The dimensions of the firebox will depend on the size of the firebox doors, and either the length of the baffle
chambers in the horizontally baffled masonry heater or the number of baffle chambers in the vertically baffled
masonry heater. Thus, the baffle chambers need to be sized prior to laying out the masonry heater dimensions of
the firebox.
There are several alternatives for constructing the firebox. The firebox base, sides and back should be lined with
refractory units to obtain a thickness of at least 2 1/2 in. (63 mm). The refractory units on the rear wall of the firebox
should extend to the top of the first baffle, as shown in Figures 1 and 2.
The top of the firebox may be formed either by using a precast reinforced refractory concrete slab, or by using
refractory units which will span the width of the firebox. Both options require slabs or units wide enough to bear at
least 2 in. (50 mm) on the side walls of the firebox liner. Another alternative is to form the top of the firebox with a
masonry arch constructed of refractory brick units. This method of forming the top of the firebox is compatible with
using several types of Dutch oven doors for the front of the fireplace.
Vertically Baffled. The baffle system for the vertically baffled Russian-style brick masonry heater is the easiest
baffle system to build. There should be at least three vertical baffle chambers and usually no more than five,
although successful systems have been built with up to nine. The greater the number of baffles, the longer the
masonry heater needs to be fired to warm the entire mass of brickwork. This may decrease the efficiency of the
system for both heating and combustion of the wood. It will also result in much hotter fires, which may augment the
deterioration of the masonry heater. Another problem with using more baffles is that those portions of the heater
which remain cooler invite creosote problems.
The baffles should be formed by using a single wythe of brick masonry to separate the baffle chambers, which are
usually 64 to 144 sq in. (40,000-90,000 mm2) in cross-sectional area. These single wythe brick baffles should be
masonry bonded to the interior wythe of the enclosing 8-in. (200 mm) brick masonry. Thus, thermal expansion of the
http://www.gobrick.com/BIA/technotes/t19d.htm
7 of 15 9/13/2009 1:04 PM
baffles will impose a lateral load on the enclosing brick masonry which should be considered in the design. Major
cracks in the heater should be avoided by keeping the wythes of the 8-in. (200 mm) thick enclosing brick masonry
walls separated by a nominal 1-in. (25 mm) air space. To help insure that thermal expansion is provided for, this
space may be filled with a compressible, noncombustible material. If a filler material is used, it may be easiest to
construct the interior wythe with properly spaced ties, then wrap the heater with the compressible, non-combustible
material, prior to constructing the exterior wythe.
Another problem is that thermal movement may separate the baffles from the interior wythe and allow lateral
movement of the baffles. Thus, in addition to the masonry bond to the interior wythe of the enclosing walls, metal
ties should also be installed every 8 in. (200 mm) vertically. These metal ties should be the only metal inside the
interior wythe of the enclosing walls of the brick masonry heater, except for reinforcement in concrete slabs. The top
of the openings through the baffles may be formed by corbeling brick units, as shown in Figures 2 and 4. Other
alternatives for forming the baffle openings are shown in Figure 5. There should be at least 12 in. (300 mm) of brick
masonry covering the top of the baffle chambers. Baffles with openings at the bottom are again single-wythe brick
masonry walls. The openings in the bottom of baffles may be formed by using corbeled brick masonry, brick
masonry arches, or by masonry units, or reinforced, precast concrete long enough to span the width of the baffle
chamber and bear a minimum of 2 in. (50 mm) on each side of the interior wythe of brick masonry. At the front face
of the masonry heater, near the bottom of the baffle chamber, a clean-out door is recommended so that any ash
buildup may be removed from the baffle chamber.
Typical Vertical and Horizontal Baffle Constructions
FIG. 5 a & b
http://www.gobrick.com/BIA/technotes/t19d.htm
8 of 15 9/13/2009 1:04 PM
Typical Vertical and Horizontal Baffle Constructions
FIG. 5 c & d
Typical Vertical and Horizontal Baffle Constructions
FIG. 5 e
http://www.gobrick.com/BIA/technotes/t19d.htm
9 of 15 9/13/2009 1:04 PM
The last upward baffle chamber in the baffle system, i.e., the chamber at the front face of the heater, becomes the
support for the conventional flue liner. By corbeling the top course of the last baffle chamber, the support for a
conventional clay flue liner is obtained.
Horizontally Baffled. The baffles for the horizontally baffled Russian-style brick masonry heater may be formed by
using corbeled brick masonry, arches, a precast, reinforced concrete slab, or clay flue liners. When the slab is used,
it should be sufficiently wide to span across the width of the baffle chamber and bear at least 2 in. (50 mm) on the
interior wythe of each side wall of the baffle chamber. Arches or corbels used to form the air passageway may be
started from the interior wythe of the baffle chamber wall. These are shown in Figures 3 and 5. The horizontal baffle
system requires a clean-out at the bottom of each baffled area. The side and either the front or rear baffle chamber
walls (depending on the number of horizontal baffles) and the top of the last baffle should be used to support the
conventional flue liner for the chimney. There should be at least a 12-in. (300 mm) thickness of brick masonry
forming the top of the last horizontal baffle.
Clean-Outs. Clean-outs for the baffle chamber of either Russian-style brick masonry heater are optional. Usually,
when properly seasoned wood is used under adequate air intake conditions, and at high temperature, creosote
should not form in large quantities. In addition, because of the baffles, most soot and ash remain in the firebox.
However, the installation of clean-outs is recommended to observe any buildup. If a buildup is occurring, the
operation of the stove should be modified, so that the buildup no longer occurs. This may be accomplished by
increasing the amount of combustion air being supplied through the firebox doors.
Crown. The crown of the Russian-style brick masonry heater should terminate at least 12 in. (300 mm) below the
ceiling of the room. Multi-story heaters are not recommended because the distance the combustion gases must
flow, from the firebox through the baffles to the chimney, cools the gases and decreases performance. Typically, the
distance from the firebox through the baffle chambers to the chimney should be limited to no more than 16 ft (4.9 m).
The crown should be at least 12 in. (300 mm) thick, starting from the highest point of the baffle chamber. The crown
may be flat, but is often constructed as an arch for esthetics.
Chimney. The chimney for the Russian-style brick masonry heater is similar to those used for fireplaces. The
chimney should be constructed with clay flue liners and 8 in. (200 mm) of brick masonry surrounding the flue liner in
such a way that a nominal 1-in. (25 mm) air space is maintained between the flue liner and the surrounding brick
masonry. The 8-in. (200 mm) chimney wall is recommended to help keep the chimney at a higher temperature to
increase performance. Additional information on chimney design and construction is provided in Technical Notes 19B
Revised.
The chimney height required for draft is usually higher than that necessary for conventional fireplaces, but following
the building code requirements for fire safety will usually result in a sufficiently high chimney. Most codes require that
the chimney terminate at least 3 ft (1 m) above the roof at the highest point of exit and at least 2 ft (600 mm) above
any portion of the building or any adjacent structure within 10 ft (3 m) of the chimney. If the draft is determined to be
inadequate by a smoke test, the chimney height should be increased to provide adequate draft.
The chimney for the Russian-style masonry heater must be free to move vertically to allow for the vertical thermal
expansion of the masonry heater supporting it. This requires properly sealed flashing and counter-flashing where the
chimney penetrates the roof line.
Esthetics
An additional consideration in the design of a brick masonry heater is the esthetics. Figures 1 through 4 show the
basic heater design for function. This results in a rectangular mass of brick within the building, which may or may not
be esthetically pleasing. Incorporating arches, corbels, racks and mantels into the design may greatly increase the
esthetic value.
SELECTION OF MATERIALS
General
The design and construction of a brick masonry heater using products available in the United States is slightly
different than the construction of the heater in Europe. European heaters are usually constructed using a single
wythe of 5-in. (125 mm) thick brick for the exterior shell. The exterior brick is then often covered with glazed ceramic
http://www.gobrick.com/BIA/technotes/t19d.htm
10 of 15 9/13/2009 1:04 PM
tiles, set in high temperature-resistant epoxy grout. Using a single wythe around the firebox and baffles may result in
cracking. The positive draft through the firebox and baffle chambers results in little danger of toxic gases escaping
into the occupied areas of the building. By using multiple wythes for the exterior shell, the potential for a crack
penetrating completely through the heater is substantially reduced. Single-wythe construction of masonry heaters is
therefore not recommended.
In addition to the construction differences of the exterior shell, the accessories used in Europe are not usually
available in North America. There are methods to modify the design and construction so that products readily
available in North America may be used in the brick masonry heater. The options also exist to either import the
accessories or to fabricate accessories similar to those used in Europe. However, these options are usually
uneconomical. Additional information regarding accessories may be obtained from the cited references.
Brick
Most building codes require that solid masonry units be used for fireplace construction. Solid brick should conform to
ASTM C 216 or C 62 for facing brick or building brick, respectively. Hollow brick conforming to ASTM C 652 may be
used if vertical reinforcement is required.
If vertical reinforcement is to be used to provide resistance to cracking, the brick masonry heater may be
constructed using a single wythe of reinforced, grouted hollow brick. Reinforced hollow brick masonry should be
constructed using at least nominal 8-in. (200 mm) thick hollow brick units. The shell of the heater may also be
constructed of a vertically and horizontally reinforced, fully grouted, multi-wythe brick masonry wall. The grout core
should be at least 2 in. (50 mm) thick and the brick wythes must be properly tied.
When face brick or building brick is used, the walls of the heater should be at least two wythes thick, using nominal
4-in. (100 mm) or 3-in. (75 mm) thick brick. Grade SW brick should be used because of its greater durability.
Refractory brick, conforming to ASTM C 64, medium duty, should be used for the firebox. The lining for the back of
the firebox should extend to the top of the first baffle chamber. These areas are exposed to the greatest amount of
heat and the refractory units are more resistant to heat and thermal shock.
Salvaged or used brick should not be used because they usually will not bond well with the mortar and lack the
durability necessary for satisfactory performance. The use of salvaged brick is discussed in Technical Notes 15.
Mortar and Grout
It is most convenient and economical to use only one type of mortar for the entire brick masonry heater and chimney
construction. This becomes difficult when constructing a brick masonry heater because of the specific requirements
of each component. The portions of the heater consisting of building, face or hollow brick should be constructed
using a Type N, portland cement-lime mortar, conforming to the proportion specifications of ASTM C 270 or BIA M
1-72. The same mortar should be used for the chimney brickwork except when wind loads exceed 25 psf (1.2 kPa).
Where high wind loads exist, a Type S, portland cement-lime mortar should be used. It may be desirable to use high
temperature-resistant mortars, such as calcium aluminate mortars, for the interior wythes and baffles of the brick
masonry heater. Such mortars will increase the durability of the heater.
The firebox and all other components constructed of refractory units should be set using a fireclay mortar,
conforming to ASTM C 105, medium duty. Other refractory mortars have also been successfully used, and thus any
high temperature-resistant mortars that have performed well may be used. It is not within the purview of the Brick
Institute of America to recommend proprietary products. The selection of the proper mortars should be determined
by an experienced fireplace expert for the specific design being considered.
For reinforced brick masonry, all cores of hollow brick masonry construction and the grout space of hollow wall
construction must be fully grouted. The grout should conform to ASTM C 476.
Flue Liners
Clay flue liners used for the chimney or to form the baffle chambers should conform to ASTM C 315. They should be
thoroughly inspected just prior to installation for cracks or other damage which might contribute to smoke or flue gas
leakage. All flue liners should be set in fireclay mortar.
http://www.gobrick.com/BIA/technotes/t19d.htm
11 of 15 9/13/2009 1:04 PM
Steel Lintels and Dampers
Steel lintels should not be used inside the exterior 4-in. (100 mm) wythe of the brick masonry heaters because of the
high temperatures involved. The difference in thermal expansion characteristics could cause cracking of the brick
masonry heater. For the same reason, metal dampers are not used within the Russian-style fireplace. Lintels of
corrosion-resistant steel, conforming to ASTM A 36, should be used over the firebox door and clean-out door
openings.
Ties and Reinforcement
Corrugated Metal Ties. Corrugated metal ties may be used to attach the baffles to the interior wythe of the heater
walls and to tie the two wythes of the exterior walls of the heater together. Ties should be corrosion-resistant, and
at least 22 ga, 7/8 in. (22.2 mm) wide, and 6 in. (150 mm) long.
Wire Ties. Wire ties are preferred for tying the brick masonry together. Wire ties should be at least 9 ga and
corrosion-resistant. The ties should be fabricated from wire conforming to ASTM A 82 or ASTM A 185.
Prefabricated Joint Reinforcement. Prefabricated joint reinforcement should be used for the exterior wythe of the
heater walls. The joint reinforcement should be fabricated from wire which complies with ASTM A 82 or ASTM A
185, and should be corrosion-resistant.
Reinforcement. Reinforcement should conform to any of the following applicable standards:
Standard Specifications for Deformed and Plain Billet-Steel Bars for Concrete Reinforcement-ASTM A 615.
Standard Specifications for Rail-Steel Deformed and Plain Bars for Concrete Reinforcement-ASTM A 616
Standard Specifications for Axle-Steel Deformed and Plain Bars for Concrete Reinforcement-ASTM A 617
Corrosion Resistance. Corrosion resistance is usually provided by a copper or zinc coating, or by using stainless
steel. To ensure adequate resistance to corrosion, coatings or materials should conform to any of the following
applicable standards:
Zinc-Coating of Flat Metal-ASTM A 153, Class B-l, B-2, or B-3
Zinc-Coating of Wire-ASTM A 116, Class 3
Copper Coated Wire-ASTM B 227, Grade 30 HS
Stainless Steel-ASTM A 167, Type 304
Firebox Doors. The doors for the firebox opening may be fabricated locally, ordered from Europe or may be
conventional metal Dutch oven doors, which are the most economical. The brick masonry heater is not designed for
airtight combustion and thus the doors need to be equipped with operable vents to control air intake into the firebox.
The size of the firebox door is a major consideration in the design of the brick masonry heater. The height and width
of the firebox and the width of the baffle chambers are usually the same as, or just slightly larger than, the firebox
door. Other alternatives exist for the firebox design and firebox doors, and are discussed in other Technical Notes in
this Series. The typical European-style door is shown in Figure 6.
http://www.gobrick.com/BIA/technotes/t19d.htm
12 of 15 9/13/2009 1:04 PM
Typical Firebox Door
FIG. 6a

Typical Firebox Door With a Metal Screen
(Fire Box and Clean-Out Doors)
FIG. 6b
http://www.gobrick.com/BIA/technotes/t19d.htm
13 of 15 9/13/2009 1:04 PM
Typical Dutch Oven Style Door
FIG. 6c
European Style Clean-Out Door Assembly
(Fire Box and Clean-Out Doors)
http://www.gobrick.com/BIA/technotes/t19d.htm
14 of 15 9/13/2009 1:04 PM
FIG. 6d
Clean-out Doors. Clean-out doors used in Europe for the baffle system are tight-fitting doors which have tapered
latches to ensure tightness of fit. These doors are shown in Figure 6. Conventional clean-out doors may be used, but
to ensure tightness, refractory units should be placed within the door opening with a compressible, non-combustible
material or set in a sand-lime mortar. This is shown in Figure 7. The refractory units increase the resistance to
combustion gas leaks, provide protection to the metal door from high temperatures, and may easily be removed and
replaced when cleaning, if necessary. Clean-out doors for the ash drop may be conventional clean-out doors,
installed in the conventional manner.
Clean-Out Door Backed with Firebrick
FIG. 7
SUMMARY
The information and suggestions contained in this Technical Notes are an accumulation of the available information
within the Brick Institute of America on Russian-style fireplaces and brick masonry heaters. The information is based
on empirical data from actual performance of such heaters here in North America and in Europe. The information and
recommendations are provided for use with good technical judgment for the design and construction of a functional
brick masonry heater. Final decisions on the design and use of materials as discussed in this Technical Notes are
not within the purview of the Brick Institute of America, and must rest with the project designer, owner, or both.
REFERENCES
1. "Complete Plans and Instructions for Construction and Operation of a Masonry Stove, Finnish, or Russian
Fireplace," by Basilio Lepuschinko, Richmond, Maine, 1980.
2. "How to Build a Russian Fireplace," by Jay Jarpe, Los Lunas, New Mexico, 1980.
3. "A Russian-Type Fireplace Demonstration and Workshop, " New Mexico Energy Institute, The University of
New Mexico, Albuquerque, New Mexico, 1981.
4. "On Building Masonry Firestoves," by J. Patrick Manley, Farmstead Magazine, Fall Issue 1980, No. 34.
5. "What's So Hot About A Russian Fireplace?" by M. R. Allan, Yankee Magazine, Dublin, New Hampshire,
February 1978.
http://www.gobrick.com/BIA/technotes/t19d.htm
15 of 15 9/13/2009 1:04 PM

Technical Notes 19E - Brick Masonry Fireplaces, Part 2 - Fountain and Contemporary Style Heaters
[1983](Reissued Feb. 1988)
Abstract: Brick masonry heaters may be used instead of conventional fireplaces to provide efficient supplemental
heating for residential buildings. The design, detailing and construction of brick masonry fireplaces with baffle
systems through which combustion gases are circulated are discussed. Information regarding building code
compliance, operation and accessories is presented, along with the basic heating principles.
Key Words: brick, buildings (codes), design, energy, fireplaces, heating, masonry, mortar.
INTRODUCTION
The basic concepts of a fireplace designed with a baffle system replacing the conventional smoke chamber are
discussed in Technical Notes 19D. Such fireplaces are often referred to as Russian-style fireplaces. These
fireplaces, or brick masonry heaters, although capable of providing efficient heat, tend to eliminate the esthetic value
of the fireplace because the firebox is typically deep within a small opening. An alternate approach to the fireplace
designed to include a baffle system through which combustion gases are circulated is the Finnish or fountain-style
heater. This Technical Notes provides the information necessary to properly design and construct a fountain-style
brick masonry heater or to modify conventional fireplace designs to incorporate baffle systems to allow the
circulation of combustion gases.
The basic principles by which high efficiencies are obtained for heating the building interior and for the combustion of
wood are the same as for the Russian-style brick masonry heater. Hot combustion gases circulating through the
massive brick masonry heater, combined with properly controlled air intake for combustion, result in high efficiencies.
The hot combustion gases are circulated through baffle chambers within the heater. The massive brick masonry is
warmed and retains the heat. This warmed brick masonry radiates the heat long after the fire is extinguished. The
basic concepts of design, construction and operation are simple, but there are several concerns which must be
addressed to insure safety and durability.
GENERAL
Operation
The operation of both the fountain-style brick masonry heater and the contemporary-style heaters is quite similar.
Both systems have unique operational advantages which are directly related to their design and the way in which
combustion gases are circulated through the massive brick masonry assembly.
Typically, the combustion chamber or firebox is loaded with 10 lb (4.5 kg) to 20 lb (9.1 kg) of wood, after a fire with
kindling is ignited. Once good combustion starts, the firebox doors or glass screens are closed and the air intakes
adjusted to the proper setting so that good combustion continues. Unlike the Russian-style brick masonry heater,
because of the design of the firebox, glass screens may be used on the fountain-style heater and the modified
conventional fireplace. The glass screens or firebox doors used should be equipped with operable air inlets so that
the air intake to the combustion chamber can be controlled. The metal firebox doors or glass screens selected for
the firebox opening should be capable of withstanding the high temperatures in the combustion chamber.
Temperatures of combusting wood will usually range from 1000
o
F (540
o
C) to 1500
o
F (820
o
C), and the temperature
of the combustion gases near the fire usually ranges from about 800
o
F (430
o
C) to 1200
o
F (650
o
C). (See Reference
10). Thus, all components of the firebox should be capable of withstanding these temperatures.
The 10 lb (4.5 kg) to 20 lb (9.1 kg) of wood loaded in the firebox will burn for about 30 min when properly seasoned
wood is used with adequate draft and combustion air. For maximum heating, the firebox should be reloaded with
about 10 lb (4.5 kg) to 20 lb (9.1 kg) of wood every 30 min for a 2-hr period. This procedure usually results in
enough heat being supplied by the brick masonry heater to keep a 2400 ft
3
(68 m
3
) room warm for a period of 8 to
http://www.gobrick.com/BIA/technotes/t19e.htm
1 of 15 9/13/2009 1:05 PM
12 hr during the coldest months in severe climates, such as in Scandinavia. Under these severe climate conditions,
the heater is usually operated for a 2-hr period twice a day, once in the morning and once in the evening. During the
more moderate seasons, or in the more moderate climates, operating the heater once in the evening may be
adequate to supplement the mechanical heating system to maintain comfortable interior temperatures throughout the
night. Other variations in operation, such as the number of loadings and the amount of wood used for each firing may
also result in increased comfort. The operator should experiment with several operations to determine how to
achieve the best performance for the various seasons.
This method of operation may be used with any properly designed and constructed brick masonry heater. Because
of the variations in the design of the fountain style heater and the modified conventional fireplace, the actual method
of operation should be modified for the specific design. The contemporary-style brick masonry heater is usually
operated the same as a conventional fireplace.
Fountain-Style Heater. The fountain-style heater is shown in Figures 1 through 3. It consists of a relatively
conventional fireplace firebox. The smoke chamber is rectangular and the baffle chambers for circulating the
combustion are located on the sides of the smoke and combustion chambers. The baffle chambers meet under the
combustion chamber where combustion gases are vented to the chimney.
FIG. 1a
http://www.gobrick.com/BIA/technotes/t19e.htm
2 of 15 9/13/2009 1:05 PM
FIG. 1b
Fountain-Style Brick Masonry Heater, Isometric
FIG. 1c
The heating output of the fountain-style heater may be increased by using a specially designed air supply system
from the firebox doors to the smoke chamber. Providing the smoke chamber with this additional air ignites the
combustion gases in the smoke chamber. The combustion gases burn at temperatures of about 1800
o
F (1100
o
C) to
2100
o
F (1150
o
C). This method of operation, because of the higher temperatures of the circulating combustion
gases, greatly increases the heating capability of the heater. This second combustion within the heater also results
in a very clean and complete combustion of the wood and its by-products. This nearly complete combustion
maximizes the efficiency of the wood fuel and considerably reduces any creosote buildup within the brick masonry
heater and chimney.
http://www.gobrick.com/BIA/technotes/t19e.htm
3 of 15 9/13/2009 1:05 PM
Fountain-Style Brick Masonry Heater
FIG. 2
http://www.gobrick.com/BIA/technotes/t19e.htm
4 of 15 9/13/2009 1:05 PM
Fountain Style Brick Masonry Heater
FIG. 3
Contemporary-Style Heater. The contemporary-style heater, shown in Figs. 4 through 7, although much more
efficient than a conventional fireplace, will not achieve the high efficiency normally obtained in the fountain-style
heater. The fountain-style heater achieves efficiencies of about 80 to 95 percent and the contemporary-style heater
provides a heating efficiency of about 70 to 80 percents. (See Reference 8). The contemporary-style heater is
designed with a rectangular firebox with the throat located at the rear of the firebox. Combustion gases are
circulated through baffle chambers within the smoke chamber and exhausted to a conventional fireplace chimney
located on top of the smoke chamber. Although this type of heater does not provide as much heat, it is usually
preferred because it provides the esthetic appearance similar to a conventional fireplace.
http://www.gobrick.com/BIA/technotes/t19e.htm
5 of 15 9/13/2009 1:05 PM
Modified Conventional Fireplace
FIG. 4
http://www.gobrick.com/BIA/technotes/t19e.htm
6 of 15 9/13/2009 1:05 PM
Modified Conventional Fireplace
FIG. 5
Building Code Compliance
There are no major model building code requirements which specifically address brick masonry heaters or their
baffle chambers. For the most part, the building code requirements for fireplaces and chimneys are applicable to the
brick masonry heater except for those requirements which address the firebox and smoke chamber dimensions. All
building code requirements for chimneys and clearances for combustibles are applicable. Because the temperatures
normally achieved in the brick masonry heater are much higher than those obtained in conventional fireplaces,
additional consideration should be given to safety and durability.
The exterior walls of the brick masonry heater should consist of at least two wythes of brick masonry when
constructed of solid units. The two wythes should be separated by a nominal 1-in. (25 mm) air space. This
separation prevents any cracks from penetrating the interior wythe through the exterior wythe of the heater. Filling
this 1-in. (25 mm) air space with a compressible, non-combustible material will insure that a separation is provided.
The two wythes should be tied to each other with corrosion-resistant metal ties, placed 16 in. (400 mm) o.c.
vertically and a maximum of 24 in. (600 mm) o.c., horizontally. The exterior wythe should contain horizontal joint
reinforcement every 16 in. (400 mm) vertically to add to the integrity of the heater. The ties and joint reinforcement
should not occur in the same course. These provisions should prevent problems due to thermal expansion and
http://www.gobrick.com/BIA/technotes/t19e.htm
7 of 15 9/13/2009 1:05 PM
differential movement without affecting the overall thermal performance of the brick masonry heater.
When reinforced brick masonry construction is required, the enclosing walls or shell of the heater may be
constructed of fully grouted hollow wall construction or fully grouted hollow brick units. If the fully grouted hollow wall
construction is to be used, the minimum 2 in. (50 mm) grout core must be fully grouted and contain sufficient
horizontal and vertical reinforcement to resist structural and thermal stresses. The two wythes must also be
adequately tied together with corrosion-resistant metal ties. If grouted hollow brick masonry is used, the cores must
be fully grouted and sufficiently reinforced to resist structural and thermal stresses. Horizontal joint reinforcement
should have proper mortar coverage in the bed joints and the units should not be less than 8 in. (200 mm) in
thickness. For additional information on reinforced brick masonry construction and hollow brick, see Technical Notes
17 and 41.
In addition to the thermal movements, the exterior surface temperatures of the brick masonry heater also need to be
considered. These surface temperatures normally range between 100
o
F (38
o
C) and 130
o
F (54
o
C). Temperatures as
high as 190
o
F (88
o
C) have been reported on certain styles of brick masonry heaters and are typical when
fountain-style heaters are operated with gas combustion occurring in the smoke chamber.
A minimum 16-in. (400 mm) clearance is recommended between the sides and the back of the brick masonry heater
and combustibles. At the floor line, this may be achieved by providing a 16-in. (400 mm) extended hearth. In the front
of the heater, a minimum 20-in. (500 mm) extended hearth should be used. This is easily achieved when the heater
is properly positioned in the room for maximum heating. This position is in the center of the room so that all four brick
walls of the heater are providing radiant heat to the room.
The brick masonry heater should always be positioned entirely inside the building. It should never be incorporated
into an exterior wall, because much of the radiant heat would be lost to the exterior. In addition to this heat loss, the
location on the exterior wall will usually result in at least one cold surface on which a considerable amount of
creosote may form. A creosote fire may well result in sufficient damage to the heater so that it is no longer safe to
operate.
FOUNTAIN-STYLE HEATER
General
A typical design of a fountain-style heater is shown in Figs. 1 and 2. There are many variations to the design and
construction of the fountain-style heater. The fountain-style heater, often referred to as a Finnish stove or vertical
contra-flow masonry stove, is believed by most experts to be the most efficient brick masonry heater. For proper
combustion, the air intake to the combustion chamber should be approximately 0.80 cfm/lb (0.00083 m
3
/sec/kg) of
wood being burned in the combustion chamber. This requirement is for the typical size fountain-style heater, which is
usually loaded with 10 lb (4.5 kg) to 20 lb (9.1 kg) every 30 min for a 2-hr period.
Base Assembly
In the fountain-style heater, the baffle system provides for the circulation of combustion gases under the hearth of
the firebox where they are vented to the chimney. Thus, when designing and constructing this style heater, the baffle
system and chimney support must be incorporated into the base assembly.
Foundation. The foundation system must be adequate to support the mass of the brick masonry heater and the
masonry chimney. When designing the foundation, care should be taken to account for soil types and foundation
conditions. Undisturbed or well-compacted soil will generally be sufficient. However, some types of soils or
foundation conditions may require special analysis. This could result in the need for special soil treatments or a
unique foundation design.
Building codes generally require that the footing be at least 12 in. (300 mm) thick, and in plan view extend a minimum
of 6 in. (150 mm) beyond each face of the masonry heater and chimney assembly. The footing should be positioned
so that its base is below the frost line to reduce the possibility of "heaving."
Hearth Support. The brick masonry used to form the hearth support is constructed directly on the foundation
system. Even with a thickened slab in a newly constructed slab-on-grade structure, masonry is usually used to
construct the base assembly to the height of the hearth support because to properly construct a fountain-style
heater, a raised hearth is usually required. The hearth support may be solid masonry carried up from the footing to
http://www.gobrick.com/BIA/technotes/t19e.htm
8 of 15 9/13/2009 1:05 PM
support the entire hearth area. To conserve materials, the base assembly is usually constructed of masonry which
has the same dimensions in plan as the masonry heater and chimney assembly itself. The base assembly is
corbeled to form the support for the extended hearth.
A soot pocket should be formed at the base of the chimney assembly, approximately 20 in. (500 mm) to 24 in. (600
mm) below the firebox hearth. These dimensions will provide a soot pocket that is, approximately 8 in. (200 mm) in
height, which will accommodate a standard size clean-out door. The baffle chamber is started at the top of the soot
pocket. The baffle chamber should be at least 5 in. (125 mm) (about two courses of standard size brick) high. The
last course of brick forming the soot pocket should be corbeled to provide the support for the chimney liner. The first
chimney liner should be cut so that the entire front of the liner, for the height of the base baffle chamber, is removed.
This opening is used to exhaust the combustion gases to the chimney. The baffle system needs to be designed so
that the vertical baffle chamber, at least 3 in. (75 mm) wide and 16 in. (400 mm) long, may be continued up along
the sides of the firebox. The remaining portion of the baffle chamber should be constructed so that the combustion
gases will be directed to the rear of the base assembly and vented to the chimney. At the base of the baffle
chamber, clean-out doors should be installed so that soot and ash deposits may be removed. The shell of the
fountain-style heater is also started at the base of the baffle system. This shell should be constructed of two wythes
of brick masonry separated by a nominal 1-in. (25 mm) air space and tied with metal ties. The exterior wythe should
contain horizontal joint reinforcement. Because these vertical baffle chambers must extend up along the sides of the
firebox, the size of the base is directly dependent upon the firebox dimensions.
Extended Hearth. As the base assembly is constructed, the support for the extended hearth is usually formed so
that the top of the extended hearth may be constructed at or slightly below the clean-outs. The extended hearth is
usually formed by placing a reinforced concrete slab on top of the corbeled base assembly, as shown in Fig. 1.
Non-combustible, or removable, forming should be placed so that it spans from the corbeled masonry to the floor
joists forming the opening for the heater. Double joists should be used around the entire perimeter of the opening
with a nailer attached to the joists to support the edge of the forming material. This type of construction is discussed
in more detail in Technical Notes 19A Revised. To decrease the number of concrete slabs occurring within the
heater assembly, the extended hearths may be formed by using trimmer arches, as shown in Fig. 3.
The extended hearth should be at least 20 in. (500 mm) in front of the firebox opening and 16 in. (400 mm) around
the remaining perimeter of the heater. The extended hearth may be eliminated on the back of the heater if it is
positioned against a non-combustible wall with a minimum fire rating of 1 hr. Once the reinforced slab is installed, the
extended hearth may be finished with brick pavers and at least one course of brick masonry should be placed to
form the base of the fountain-style heater baffle chamber. If removable forming is used, the concrete slab used to
support the extended hearth must be designed as a cantilever
Ash Pit. The masonry heater, because of its efficient combustion of wood, does not require an ash pit. If an ash pit
is desired, it should be positioned so that the ash drop occurs in the center of the firebox width, toward the front of
the firebox. This location results in the ash pit being formed by brick masonry used to construct the portion of the
baffle system in the base assembly. The ash drop should extend through the firebox hearth and hearth support. If an
exterior combustion air system is desired, the optional ash pit should be eliminated and replaced with the vertical air
passageway which forms the exterior combustion air inlet to the firebox.
Firebox Assembly
The firebox assembly is constructed on a reinforced concrete slab, which spans from the front to the rear of the
heater. The reinforced concrete slab should be at least 2 1/2 in. (63 mm) thick and must be thick enough to satisfy
the structural and reinforcement coverage requirements. The slab should bear at least 2 in. (50 mm) on the interior
wythe of the shell of the brick masonry heater. The width of the slab must be limited to the outside dimensions of the
firebox width so that the baffle chamber may be continuous to the base assembly.
The typical dimensions of the firebox opening are: 12 in. (300 mm) to 18 in. (450 mm) wide, by 24 in. (600 mm) in
height. The firebox is usually 18 in. (450 mm) to 24 in. (600 mm) deep, and about as wide as the firebox opening.
The shape of the firebox is similar to that of a conventional fireplace without flared sides. The back wall of the
firebox is sloped forward to form the throat. This slope usually begins at about 12 in. (300 mm) from the firebox
hearth and should form a 4-in. (100 mm) throat that is as wide as the combustion chamber.
The combustion chamber should be constructed of refractory brick units to obtain a thickness of at least 2 1/2 in. (63
mm). A sliding damper may be installed at the throat. The throat damper is usually installed so that it may be
http://www.gobrick.com/BIA/technotes/t19e.htm
9 of 15 9/13/2009 1:05 PM
operated from the face of the fireplace. If a throat damper is used, sufficient provisions are necessary to allow for
the thermal expansion so that the damper does not bind or crack the surrounding brick masonry. Whenever possible,
a throat damper should not be used.
If air is to be provided to the smoke chamber, provisions must be made in the face of the heater to form an air
passageway. The air inlet is usually through specially fabricated firebox doors and placed so that the air enters the
smoke chamber at the throat where the velocity of the combustion gases is the greatest. A specially designed
firebox door is usually required to provide combustion air to the smoke chamber.
Smoke Chamber Assembly
The smoke chamber assembly should be constructed of refractory brick units. The base of the smoke chamber
should be constructed of fire brick. The front and rear walls of the smoke chamber should be 12 in. (300 mm) to 16
in. (400 mm) high. The side walls should be 8 in. (200 mm) to 12 in. (300 mm) high to provide an opening for the
combustion gases to enter the vertical baffle chambers. The top of the smoke chamber should be constructed of a 2
1/2 in. (63 mm) to 4 in. (100 mm) thick reinforced refractory-concrete slab. The slab thickness must be adequate for
structural and reinforcement coverage requirements. The rear of the slab should bear at least 2 in. (50 mm) on the
wythe of brick immediately behind the fire brick wall of the smoke chamber. The front and sides should bear at least
2 in. (50 mm) on the interior wythe of the shell of the heater. The distance from the bottom of this slab to the bottom
of the baffle chamber should not exceed 5 ft (1.5 m), thus controlling the distance combustion gases are circulated
and the height of the heater.
The smoke chamber is often constructed with an opening directly into the chimney. This opening is installed to
provide a by-pass for when little or no heating is required. It should be equipped with a rotating damper. This
by-pass is shown in Figs. 1 through 3.
Crown
The crown of the fountain-style heater should be at least 12 in. (300 mm) thick, including the refractory slab forming
the top of the smoke chamber. The crown should terminate at least 12 in. (300 mm) below the ceiling of the room.
This may require that the baffle chambers be extended below the floor line of the room.
Chimney
The chimney for the fountain-style heater is similar to those used for residential appliances. The chimney should be
constructed of fireclay flue liners and 8 in. (200 mm) of brick masonry surrounding the liner in such a way as to
maintain a nominal 1 in. (25 mm) space between the flue liner and the brick chimney walls. Additional information on
chimney design and construction is provided in Technical Notes 19B Revised.
The chimney height required for draft is usually higher than that necessary for a conventional fireplace, but following
building code requirements for fire safety will usually result in a sufficiently high chimney. The major model building
codes require that chimneys must terminate at least 3 ft (1 m) above the roof at the highest point of exit and at least
2 ft (600 mm) above any portion of the building or adjacent structures within 10 ft (3 m) of the chimney. If draft is
determined to be inadequate by a smoke test, the chimney height should be increased to provide adequate draft.
CONTEMPORARY-STYLE HEATERS
General
The brick masonry heater formed by modifying a conventional fireplace, as shown in Figs. 4 through 7, is one of the
easiest brick masonry heaters to construct and also retains most of the esthetic value of a conventional fireplace.
This style of heater has the baffle system, through which combustion gases are circulated, incorporated solely in the
smoke chamber assembly. Variations in the smoke chamber and the firebox are the only major differences from a
conventional fireplace. The front and side sections of the contemporary-style heater are shown in Figs. 4 and 5,
respectively. The heater under construction is shown in Fig. 6, and Fig. 7 shows a completed contemporary-style
heater.
Base Assembly
http://www.gobrick.com/BIA/technotes/t19e.htm
10 of 15 9/13/2009 1:05 PM
The requirements for the base assembly are essentially the same as for the other brick masonry heaters and
conventional fireplaces. The foundation system requirements discussed for fountain-style heaters are also applicable
to the modified conventional fireplace.
The hearth support is similar to that of a conventional fireplace, as discussed in Technical Notes 19 Revised and
19A Revised. The hearth support is usually constructed of a reinforced concrete slab, as previously discussed. For
this style heater, it is necessary to have only a 20-in. (500 mm) extended hearth at the front face of the heater. The
thickness of the side and rear walls at the floor line usually provides adequate fire protection. However, because the
higher portions of this heater are much hotter, a minimum 12-in. (300 mm) clearance should be maintained between
the back and sides of the heater and any combustible materials.
If an ash pit is desired, it may be constructed the same as for a conventional fireplace. External combustion and
draft air systems used with conventional fireplaces may also be incorporated into this style heater. Additional
information is provided in Technical Notes 19 Revised and 19A Revised.
Firebox Assembly
The firebox is constructed as a rectangular box that is usually about 20 in. (500 mm) deep and 30 in. (750 mm)
wide. The opening of the firebox is usually about 27 in. (675 mm) high. Immediately behind the top of the firebox
opening, which is constructed of face or building brick on a steel lintel, is the top of the firebox. The top of the firebox
is formed as a brick masonry arch, extending from the inside face of the heater and terminating about 6 in. (150 mm)
to 8 in. (200 mm) from the rear wall of the firebox. At the crown of the arch, above the opening, the first horizontal
baffle is constructed. This baffle extends about 8 in. (200 mm) to 12 in. (300 mm) on either side of the firebox, and is
constructed of refractory units.
Smoke Chamber Assembly
The conventional smoke chamber is replaced by a baffle system as shown in Figs. 4 and 5. The baffle system
begins at the top of the rear of the firebox. The baffle chambers should be about 7 1/2 in. (188 mm) by 7 1/2 in.
(188 mm) square. Immediately above the first horizontal baffle are three horizontal baffle chambers, each separated
by 5 in. (125 mm) of face or building brick. The baffle system is separated vertically into right and left chambers by 4
in. (100 mm) to 8 in. (200 mm) of brickwork. The second horizontal baffle above the firebox should be equipped with
a clean-out so that any soot or ash deposits may be removed.
Contemporary-Style Brick Masonry Heater Under Construction
FIG. 6
Crown
The crown of this modified conventional fireplace should be at least 12 in. (300 mm) thick, and terminate at least 12
in. (300 mm) from the ceiling of the room. A rotating damper should be installed at the center of the heater width.
Chimney
The chimney should be constructed as a lined chimney with 8 in. (200 mm) of brick masonry surrounding the flue
http://www.gobrick.com/BIA/technotes/t19e.htm
11 of 15 9/13/2009 1:05 PM
liner in such a way that a nominal 1-in. (25 mm) air space is maintained between the flue liner and the brick. The flue
liner support should be formed by the last course of brick masonry in the crown. The termination requirements as
previously discussed apply and additional information regarding chimney design is provided in Technical Notes 19B
Revised.
Contemporary-Style Brick Masonry Heater
FIG. 7
SELECTION OF MATERIALS
General
The accessories used in Europe are not usually available in the United States. The design and construction of the
brick masonry heater using products available in North America are slightly different from the construction of the
heater in Europe. There are methods to modify the design and construction so that products readily available in the
United States may be used with the brick masonry heater. The option also exists to import the accessories or to
fabricate accessories similar to those used in Europe. However, these options are usually uneconomical. Additional
information regarding accessories may be obtained from the cited references.
Brick
Most building codes require that solid masonry units be used for fireplace construction. Solid brick should be nominal
3 in. (75 mm) or 4 in. (100 mm) thick, conforming to ASTM C 216 or C 62, for facing brick or building brick,
respectively. If a single wythe of reinforced, grouted hollow brick is used, the hollow brick should be at least 8 in.
(200 mm) wide and should conform to ASTM C 652. Grade SW brick should be used because of its greater
durability.
Refractory brick, conforming to ASTM C 64, medium duty, should be used for the firebox. The lining for the first
baffle chamber of the contemporary-style heater and the smoke chamber of the fountain-style heater should also be
constructed of refractory units because these areas are exposed to the greatest amounts of heat. The refractory
units are more resistant to heat and thermal shock.
Salvaged or used brick should not be used, because they usually will not bond well with the mortar and lack the
durability necessary for satisfactory performance. The use of salvaged brick is discussed in Technical Notes 15.
Flue Liners
Flue liners should conform to ASTM C 315. They should be thoroughly inspected just prior to installation for cracks
or other damage that might contribute to smoke and flue gas leakage.
Mortar and Grout
It is most convenient and economical to use only one type of mortar for the entire brick masonry heater and chimney
assembly. This becomes difficult when constructing a brick masonry heater because of the specific requirements of
each component. The portions of the heater consisting of building, face or hollow brick should be constructed using a
http://www.gobrick.com/BIA/technotes/t19e.htm
12 of 15 9/13/2009 1:05 PM
Type N, portland cement-lime mortar, conforming to the proportion specifications of ASTM C 270 or BIA M1-72. The
same mortar should be used for the chimney brickwork except when wind loads exceed 25 psf (1.2 kPa). Where
high wind loads exist, a Type S, portland cement-lime mortar, conforming to the proportion specifications of ASTM C
270 or BIA M1-72, should be used. It may be desirable to use high temperature-resistant mortars, such as calcium
aluminate mortars, for the interior wythes and baffles of the brick masonry heater. Such mortars will increase the
durability of the heater.
The firebox and all other components constructed of refractory units should be placed using a fireclay mortar
conforming to ASTM C 105, medium duty. Other refractory mortars have also been used successfully, and thus any
high temperature-resistant mortars that have performed well may be used. It is not within the purview of the Brick
Institute of America to recommend proprietary products. The selection of the proper mortars should be determined
by an experienced fireplace expert for the specific design being considered.
Grout is required for reinforced hollow walls and reinforced hollow brick construction. Such assemblies must be fully
grouted. The grout should conform to ASTM C 476.
Steel Lintels and Dampers
Steel lintels should be used only above the firebox opening because of the high temperatures occurring within the
heater. The thermal expansion characteristics of the different materials could cause cracking of the brick masonry
heater. Lintels of corrosion-resistant steel, conforming to ASTM A 36, should be used to span over the firebox door
openings. Non-combustible, compressible material should be placed at the ends of all lintels to provide for differential
thermal movements.
Ties and Reinforcement
Corrugated Metal Ties. Corrugated metal ties may be used to anchor the baffles to the interior wythe of the heater
walls and to tie the two wythes of the exterior walls of the heater. Ties should be corrosion-resistant, approximately
22 ga, 7/8 in. (22.2 mm) wide, and 6 in. (150 mm) long. Stiffer ties should not be used, as they may transfer
stresses due to thermal expansion of the interior wythe to the exterior wythe of the brick masonry heater.
Prefabricated Joint Reinforcement. Prefabricated joint reinforcement should be used only in the exterior wythe of
the heater walls. T he joint reinforcement should be fabricated from wire which complies with ASTM A 82 or ASTM
A 185, and must be corrosion-resistant.
Reinforcement. Reinforcement should conform to any of the following applicable standards:
Standard Specifications for Deformed Billet-Steel Bars for Concrete Reinforcement-ASTM A 615
Standard Specifications for Rail-Steel Deformed Bars for Concrete Reinforcement-ASTM A 616
Standard Specifications for Axle-Steel Deformed Bars for Concrete Reinforcement-ASTM A 617
Corrosion Resistance. Corrosion resistance is usually provided by a copper or zinc coating, or by using stainless
steel. To ensure adequate resistance to corrosion, coatings or materials should conform to any of the following
applicable standards:
Zinc-Coating of Flat Metal-ASTM A 153, Class B-1, B-2, or B-3
Zinc-Coating of Wire-ASTM A 641, Class 3
Copper Coated Wire-ASTM B 227, Grade 30 HS
Stainless Steel-ASTM A 167, Type 304
Firebox Doors
The doors for the firebox opening may be fabricated locally, ordered from Europe, or may be conventional, metal
Dutch oven doors, which are the most economical. The brick masonry heater is not designed for airtight combustion
http://www.gobrick.com/BIA/technotes/t19e.htm
13 of 15 9/13/2009 1:05 PM
and thus the doors need to be equipped with operable vents to control air intake into the firebox. The size of the
firebox opening is determined by the size of the firebox door used. Other alternatives which exist for the firebox
design and firebox doors are discussed in other Technical Notes in this Series. Glass screens may be used, but
they must be capable of withstanding the high temperatures and severe exposure occurring within the firebox.
Clean-Out Doors
Conventional clean-out doors may be used, but to ensure tightness, refractory units should be placed within the door
opening with a compressible material or set in a sand-lime mortar. The refractory units increase the resistance to
combustion gas leaks, provide protection to the metal door from high temperatures, and may easily be removed and
replaced when cleaning. This is shown in Technical Notes 19D. Clean-out doors for the ash drop may be
conventional clean-out doors, installed in the conventional manner.
SUMMARY
The information and suggestions contained in this Technical Notes are an accumulation of the available information
within the Brick Institute of America on brick masonry heaters. The information is based on empirical data from
actual performance of such heaters here in North America and in Europe. The information provided in this Technical
Notes and Technical Notes 19D does not address all of the possible variations, or alterations, which may be
incorporated into a brick masonry heater by design or construction requirements. The information and
recommendations provided in this Technical Notes discuss the basic principles and guidelines by which fireplace
experts using good technical judgment may design and construct a functional brick masonry heater.
Final decisions on the design and use of materials, as discussed in this Technical Notes, are not within the purview
of the Brick Institute of America and must rest with the project designer, owner, or both.
ACKNOWLEDGMENTS
The assistance provided by Mira Wisniewska, of Technical Translation International, Limited, 500 Fifth Avenue, Suite
200, New York, New York, 10036, who translated the articles listed in References 1 through 5, is greatly
appreciated. The information provided by Heikki Hyytiainen to the members of the BIA Technical and Manpower
Development Staff at the ''Masonry Stoves Workshop,'' held by Albie Barden in Camden, Maine, July 12-15, 1981,
was most helpful, and is greatly appreciated.
REFERENCES
1. "Tulisija-ja hormirakenteet.'' by Kari Mkel, Tekn. Iis. Tiilikeskus Oy, Iso Roobertinkatu 20,00120 Helsinki
12, Finland, 1980 Rakennustaito. (Fireplaces and Flue Structures)
2. ''Tiilessa" on inhimillinen jlki, Keskusuuni,''Tiilikeskus Oy, Iso Roobertinkatu 20, 00120 Helsinki 12, Finland,
1980. (Brick Bears Traces of Human Touch).
3. ''Tiilessa" on inhimillinen jlki. Lmmittv Ja Lampoavaraava Takka,'' Tiilikeskus Oy, Iso Roobertinkatu
20,00120 Helsinki 12, Finland, (Fireplaces for Heating and Heat Storage, List of Materials and Supplies and
Drawings).
4. ''Tarvitaanko uuneja jalleen'' by Professor Olavi Vuorelainen, Tiili, March 1978. (Are Stoves Again
Required'?)
5. ''Tilisijan suunnittelu" by Heikki Hyytiainen, Architect, Tiili, March 1978. (Fireplace Design).
6. ''The Russian Fireplace,'' Colorado Masonry Institute, Suite 301, 3003 East Third Avenue at Milwaukee,
Denver, Colorado, 80206, 1982.
7. ''Complete plans and Instruction for Construction and Operation of a Masonry Stove, Finnish or Russian
Fireplace," by Basilio Lepuschinko, Richmond, Maine, 1980.
8. "A Russian-Type Fireplace Demonstration and Workshop,'' New Mexico Energy Institute, The University of
http://www.gobrick.com/BIA/technotes/t19e.htm
14 of 15 9/13/2009 1:05 PM
New Mexico, Albuquerque, New Mexico, May 1981.
9. Summit Pressed Brick and Tile Company, Thirteenth and Erie, Pueblo, Colorado, 81002.
10. The Woodburners Encyclopedia, Section One, by Jay Shelton, Vermont Crossroads Press, Waitsfield,
Vermont, 05672, February 1978.
http://www.gobrick.com/BIA/technotes/t19e.htm
15 of 15 9/13/2009 1:05 PM
Cleaning Brickwork
Abstract: This Technical Note addresses cleaning of brickwork and brick pavements. Methods for removal of efflorescence
and a variety of specific stains are discussed, that if followed, should result in the successful cleaning of brickwork.
Key Words: abrasive blasting, acid, bucket and brush cleaning, cleaning, efflorescence, poultice, pressurized water, stains.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
20
June
2006
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
During Construction:
Use bricklaying techniques that reduce mortar smears dur-
ing construction
Use construction practices that prevent debris from splash-
ing onto brickwork and minimize water penetration into
unfinished masonry
Prior to Cleaning:
Match the cleaning method and cleaning solution to the
type of brick
Protect adjacent materials that may be damaged by clean-
ing
Remove large mortar tags using wooden paddles or non-
metallic tools
Test the cleaning method and materials on a 20 ft
2
(2 m
2
)
sample area and allow wall to dry before evaluation
Determine the environmental impact and appropriate
removal method of cleaning effluent
For All Cleaning Methods:
Select the gentlest effective cleaning method
Follow the brick manufacturers recommended cleaning
procedure
Do not use unbuffered muriatic acid
Clean new masonry as soon as possible after mortar hard-
ens, typically 7 days. More aggressive cleaning methods,
such as abrasive blasting, may require a longer mortar
curing time prior to cleaning
Clean from the top of the wall section to the bottom
For consistent results, do not overlap areas being cleaned
Bucket and Brush Cleaning:
Saturate the area to be cleaned and brickwork below with
water prior to applying cleaning solution and keep wet until
final rinse
Mix and apply cleaning solution according to manufactur-
ers instructions
Do not allow cleaning solution to dry on brickwork
After cleaning, thoroughly rinse the area being cleaned
and the area below with water
Pressurized Water Cleaning:
Determine appropriate water pressure, nozzle type and
distance between wall and nozzle by trial cleaning; main-
tain consistently throughout cleaning
Saturate the area to be cleaned and brickwork below with
water prior to applying cleaning solution, and keep wet
until final rinse
Apply cleaning solution according to manufacturers
instructions with a low-pressure sprayer, 30 to 50 psi (200
to 350 kPa) using a 50 fan-shaped sprayer, or by brush
Do not use high pressure to apply cleaning solution
Do not allow cleaning solution to dry on brickwork
Thoroughly rinse using a maximum water pressure of
200 to 300 psi (1,400 to 2,100 kPa) with a 25 to 50 fan-
shaped tip
Abrasive Blasting:
Do not use abrasive blasting on brick with a sand finish or
decorative surface coating
Brickwork should be dry and well cured prior to abrasive
cleaning
Determine appropriate air pressure, abrasive, distance and
angle between wall and nozzle by trial cleaning; maintain
consistently throughout cleaning
Efflorescence Control:
Allow one year of weathering to naturally remove new
building bloom
Remove light efflorescence by dry brushing or with a stiff
fiber brush and water
Before attempting to clean recurring efflorescence, identify
and correct the source of water penetration and allow the
brickwork to dry
Remove stubborn accumulations with a proprietary cleaner
according to the manufacturers instructions
Page 1 of 10
INTRODUCTION
The final appearance of brickwork depends primarily on the attention given to masonry surfaces during construc-
tion and the cleaning process. Recommended cleaning methods and materials vary depending on the type of
brick, mortar, construction and reason for cleaning. For example, cleaning the newly constructed brickwork of an
entire building requires a different approach than removing stains from an isolated portion of an existing wall.
An effective general approach to ensuring clean brickwork includes the following steps before the cleaning opera-
tion begins:
- Reduce the need to clean through detailing and construction techniques that reduce water penetration
and staining.
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 2 of 10
- Account for any special considerations, such as decorative coatings or finishes, water repellents, mortar
type, mortar color or historic significance.
- Repair leaks contributing to staining.
- Identify the stain or discoloration and select appropriate cleaning materials and methods that will produce
desired results.
- Clean a sample test area or panel, carefully following brick manufacturers directions, and allow to dry
before evaluating and applying to larger areas.
The selection of effective cleaning solutions, as well as the use of consistent and appropriate cleaning procedures
throughout the job, are essential to successful cleaning and cannot be overemphasized. Improper cleaning prac-
tices can cause a host of problems that, in severe cases, cannot be repaired.
This Technical Note does not address specific safety issues related to various methods of cleaning brick masonry.
It should be noted however that cleaning agents and processes may be hazardous and may cause injury if used
carelessly. Cleaning operations should only be performed by persons familiar with and equipped to handle the
safety risks associated with the work.
CLEANING NEW BRICKWORK
Brickwork is often cleaned soon after construction is completed to remove mortar smears and construction dirt
that detract from the appearance of the masonry. With new construction, keeping the masonry clean as it is
erected can be very cost-effective by eliminating the need for extensive cleaning after construction. When it is
determined that brickwork needs to be cleaned, the brick cube identification card and other pertinent manufac-
turer information should be consulted first to ascertain the recommended cleaning procedures for the brick. As
discussed under Suggested Cleaning Methods below, recommended cleaning materials and methods vary with
the type of brick.
Keeping Brickwork Clean During Construction
Some general practices that can be used to construct a cleaner wall are:
- Protect site-stored brick from mud. Store brick off the ground under protective covering.
- Erect scaffolding far enough away from the wall to allow mortar droppings to fall to the ground. Scaffold
boards closest to the wall should be angled away from the wall or removed at the end of the day to
remove excess mortar droppings and prevent rain from splashing mortar and dirt directly onto the com-
pleted masonry.
- Protect the base of the wall from rain-splashed mud and mortar splatter. Use straw, sand, sawdust, plas-
tic sheeting or fabric spread out on the ground, extending 3 to 4 ft (0.9 to 1.2 m) from the wall surface
and 2 to 3 ft (0.6 to 0.9 m) up the wall. Keep this protection in place until final landscaping.
- Cover wall openings and tops of walls with a waterproof membrane at the end of the workday and at
other work stoppages to prevent mortar joint wash out and entry of water into the completed masonry.
- Protect newly constructed brickwork from adjacent construction practices that may cause staining, such
as placing concrete or spraying curing agent.
It is always advisable for masons to keep brickwork as free from mortar smears as possible. Masons should also
be careful to prevent excessive mortar droppings from contacting the face of the wall or falling into the air space.
In addition to the bricklaying techniques described in Technical Note 7B, the practices below should be followed:
- After spreading mortar, but before laying brick, the trowel edge should be used to cut mortar even with
the wall face, preventing excessive extrusion of mortar onto the face of the wall as the brick are laid.
- After tooling joints, excess mortar and dust should be brushed from the surface, preferably using a medi-
um-soft bristle or fiber brush. Brushes with steel bristles are not recommended as they may leave behind
small particles which can rust. Brushing is preferable to bagging or sacking. Avoid any motion that will
result in rubbing or pressing mortar particles into the brick faces.
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 3 of 10
- Large clumps of mortar that adhere to brickwork should be allowed to become firm, then removed by
hand with wooden paddles, a loose brick or nonmetallic tools.
Trial Cleaning
Before cleaning, it is beneficial to test potential cleaning procedures and solutions on a sample area of about
20 ft
2
(2 m
2
), or large enough to evaluate the selected cleaning procedure. Although not common for small resi-
dential projects, trial cleaning on larger, more complex projects not only serves as a means to determine whether
mortar or stains can be removed, but helps to identify the most effective procedures that cause the least damage
to the masonry. Optimal concentrations of cleaning products and unexpected problems can also be determined
through trial cleaning. Once approved, the test area can serve as a standard for the appearance of the brickwork
after cleaning.
Reactions between cleaning solutions and certain minerals found in some brick or their surface coatings may
cause stains. Thus, it is safer to test a small area before subjecting the entire project to the cleaning procedure.
Ideally, a portion of the sample panel can be tested, leaving the building and the rest of the sample panel undam-
aged in case the brickwork is adversely affected. If trial cleaning must be performed on a building, select an incon-
spicuous location. Trial cleaning should be performed at temperature and humidity conditions that closely approxi-
mate the conditions that will be experienced during cleaning.
J udge the effectiveness of a cleaning agent or procedure by inspecting both brick and mortar in the trial area after
it has dried sufficiently, usually about one week. Approval of the trial area should precede application to any addi-
tional areas.
Suggested Cleaning Methods
Generally, the cleaning method that effectively cleans the brickwork while being the gentlest, or least harmful to
the masonry is the most appropriate. Commonly used cleaning methods for new masonry include bucket and
brush hand cleaning and pressurized water cleaning.
Always consult brick manufacturers for recommendations on cleaning specific brick. When more than one type or
color of brick is used, the brick manufacturer can aid in identifying a cleaning method that will be safe for all of the
brickwork. Table 1 suggests appropriate cleaning methods for various brick types, which can be used when guide-
lines are not available from the brick manufacturer. These are general recommendations and may not be effective
on all brick described in each category. The use of colored mortars may require special consideration, as noted in
Table 1.
Special Considerations. Air temperature, temperature of masonry and wind conditions affect the drying time
and reaction rate of cleaning solutions. Chemical cleaning solutions are generally more effective when the out-
door temperature is 50 F (10 C) or above. To avoid harming the masonry or increasing the risk of efflorescence,
cleaning methods that involve water should not be used during freezing weather or when it is expected.
Do not allow cleaning solutions to dry on brickwork. In hot weather, the cleaning crew can avoid this by working
on small or shaded areas. The size of the work area should be determined after a trial run. For consistent results,
avoid overlapping work areas.
Some chemicals used to clean brickwork and their fumes may be harmful. Use protective clothing and accesso-
ries, proper ventilation and exercise safe handling procedures. Comply with federal, state or local laws regulating
the use and disposal of chemicals and cleaning wastewater. Strictly observe the cleaner manufacturers material
safety data sheet and recommended handling requirements.
Brick texture may also influence the effectiveness of cleaning operations. Mortar stains and smears are generally
easier to remove from brick with smooth textures because less surface area is exposed. These brick include die
skin extruded brick, glazed brick, water-struck molded brick and dry-pressed brick. They are easier to presoak and
rinse because their unbroken surfaces are more likely to display poor rinsing, acid staining and poor removal of
mortar smears. Mortar and dirt tend to penetrate deeper into textures. Brick that are wire-cut, coated or textured
extruded brick and sand-struck molded brick provide additional surface area for water and acid absorption. Use of
pressurized water may assist in complete rinsing of rough textured brick.
General Cleaning Procedure. The following general cleaning procedure is applicable to a variety of cleaning
methods and is commonly used for new brickwork as well as removing stains from existing masonry.
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 4 of 10
1. Decide when to clean. Mortar must harden prior to cleaning. It is generally best to schedule cleaning at
least seven days after brickwork is completed. In some cases it may be possible to clean earlier; how-
ever, effects on the masonry and influencing factors such as weather conditions and the type of brick and
mortar should be carefully considered. Prolonged time periods between the completion of the masonry
and cleaning should be avoided. After one month, mortar smears and splatters left on brickwork become
increasingly difficult to remove.
2. Remove larger clumps of mortar using wooden paddles or nonmetallic tools. Metal tools may damage the
brickwork or leave behind fragments that oxidize and cause rust stains.
3. Select the proper cleaning solution. There are many types of proprietary cleaners available that are for-
mulated to remove specific stains or for use with a particular type of brick. Be careful to select cleaning
products suitable for the brick, mortar and adjacent materials and follow the cleaner manufacturers rec-
ommended instructions. Each product being considered should be evaluated as discussed previously in
Trial Cleaning.
Do not use unbuffered muriatic acid. Use of unbuffered muriatic acid solutions tend to cause further stains
and damage mortar joints. Many proprietary cleaners contain acids, however, their formulations include
other chemicals that make them safer, easier to use properly and more environmentally responsible.
4. Protect adjacent materials and nearby plants. Mask or otherwise protect windows, doors, and materials
such as sealants, metal, glass, wood, limestone, cast stone, concrete masonry and ornamental trim from
cleaning solutions. Cleaning chemicals may also damage plants and grass. It may be necessary to pre-
vent the cleaning solution and run-off from contacting plants or the surrounding soil.
5. Saturate the area to be cleaned. Flush with water from the top down. Saturated brick masonry will not
absorb the cleaning solution or dissolved mortar particles. Areas below the area being cleaned should
also be saturated and kept wet until after the final rinse to prevent streaking and absorption of the run-off
from above.
Brick Category Cleaning Method Remarks
Pressurized Water
Abrasive Blasting
Pressurized Water
Abrasive Blasting
See Brick Category for additional remarks based on brick color.
Water, detergents, emulsifying agents, or suitable proprietary compounds may be
used.
Clean with water, detergents, emulsifying agents, or suitable proprietary
compounds. Unbuffered muriatic acid solutions tend to cause stains in brick
containing manganese and vanadium. Light colored brick are more susceptible to
"acid burn" and stains, compared to darker units.
Red and Red Flashed
White, Tan, Buff, Gray,
Pink, Brown, Black,
Specks and Spots
Sand Finish or
Surface Coating
Bucket and Brush
Hand Cleaning
Clean with water and scrub brush using light pressure. Stubborn mortar stains may
require use of cleaning solutions. Abrasive blasting is not recommended. Cleaning
may affect appearance.
Bucket and Brush
Hand Cleaning
Bucket and Brush
Hand Cleaning
Glazed Brick
Colored Mortars
Bucket and Brush
Hand Cleaning
Pressurized Water
Wipe glazed surface with soft cloth within a few minutes of laying units. Use a soft
sponge or brush plus ample water supply for final washing. Use detergents where
necessary and proprietary cleaners only for very difficult mortar stain. Consult brick
and cleaner manufacturer before use of proprietary cleaners on salt glazed or
metallic glazed brick. Do not use abrasive powders. Do not use metal cleaning
tools or brushes.
Many manufacturers of colored mortars do not recommend chemical cleaning
solutions. Unbuffered acids and some proprietary cleaners tend to bleach colored
mortars. Mild detergent solutions are generally recommended.
Method is generally
controlled by
Brick Category
TABLE 1
Quick Guide for Cleaning Brickwork
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 5 of 10
6. Apply the cleaning solution. For proprietary compounds, follow the manufacturers instructions for appli-
cation, dwell time and cleaning technique. Wooden paddles or other non-metallic tools may be used to
remove stubborn particles.
7. Rinse thoroughly. Flush walls with large amounts of clean water from top to bottom before cleaned sur-
faces can dry. Failure to completely flush the wall of cleaning solution and dissolved matter may result in
the formation of white scum.
Individual cleaning methods and procedures may vary slightly from this general procedure; where appropriate
such variations are noted in succeeding sections of this Technical Note.
Bucket and Brush Hand Cleaning. This is a popular but misunderstood method used to clean brick masonry. Its
popularity is due to the simplicity of execution and the availability of proprietary cleaning compounds. This clean-
ing method is applicable to virtually all brick types. The least aggressive method of cleaning is the bucket and
brush method with clean water only. If a cleaning solution is used, it should be matched to the specific brick. The
General Cleaning Procedure given above is applicable to bucket and brush cleaning with the following amend-
ments:
- In Step 1, cleaning can often begin 24 hours after the masonry is completed if only clean water without
chemicals is used.
- In Step 6, use a long handled stiff fiber brush or other type as recommended by the cleaning solution
manufacturer. Do not use metal brushes which may damage mortar joints or result in further staining.
Depend on the chemical reaction of the cleaner rather than the scrubbing action of the brush. If stubborn
mortar smears are not removed, reapplication is often more effective than hard scrubbing.
Pressurized Water Cleaning. Cleaning contractors often utilize pressurized water because it is less labor
intensive than bucket and brush cleaning and permits large areas to be cleaned much faster. Pressurized water
cleaning permits the operator to spray clean water on a wall over 100 ft (30 m) from the tank and compressor.
However, the method requires more skill than the bucket and brush method, as consistent results depend on
maintaining a consistent pressure, water flow rate, distance from the wall and angle between the water jet and
the wall. It is also important to use uniform horizontal strokes. The effects of pressurized water cleaning on each
project or type of brick should be carefully considered as excessive pressure may damage brick surfaces, erode
mortar joints and remove finishes or other surface coatings, resulting in a different appearance. Nozzle pressures
less than 300 psi (2,100 kPa) are typically recommended. The brick manufacturer should be consulted before use
of pressurized water to clean brick.
With the following modifications, the General Cleaning Procedure described previously is applicable to pressur-
ized water cleaning:
- In Step 3, when selecting a cleaning solution, verify its compatibility with the equipment to be used. Mix
proprietary cleaners in accordance with the manufacturers instructions.
- In Step 5, a maximum pressure of 30 to 50 psi (200 to 350 kPa) with a 25 to 50 fan-shaped nozzle is
recommended when using a sprayer to presoak the wall.
- In Step 6, the cleaning solution should be applied by a low-pressure sprayer, (30 to 50 psi [200 to 350
kPa]), with a 50 fan-shaped sprayer nozzle, or by brush. Cleaning solutions applied under high pressure
can be driven into the masonry and become the source of future staining.
- In Step 7, use a 25 to 50 fan-shaped nozzle and a maximum water pressure of 200 to 300 psi (1,400 to
2,100 kPa) to flush the cleaning solution from the brickwork. If trial cleaning or prior experience with the
selected brick has established that no damage will result, higher pressures may be used.
Improper Cleaning
Cleaning failures generally fall into one of the following categories:
- Failure to thoroughly saturate the brick masonry surface with water before and after application of
chemical or detergent cleaning solutions. Dry masonry permits absorption of the cleaning solution and
may result in white scum, efflorescence, manganese or vanadium stains. Saturating the surface prior to
cleaning reduces the masonrys absorption rate, permitting the cleaning solution to stay on the surface
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 6 of 10
of the brickwork rather than being absorbed. Likewise, thorough rinsing reduces the potential for stains
caused by cleaning solution residue.
- Use of improper chemical cleaning solutions. Improperly mixed or overly concentrated acid solutions
can etch the brick or dissolve cementitious materials from mortar joints. Unbuffered acid has a tendency
to discolor masonry units, particularly lighter shades, producing an appearance frequently termed acid
burn and can also promote the development of vanadium and manganese stains.
- Excessively aggressive cleaning methods. Cleaning methods such as abrasive blasting and high pres-
sure water cleaning, that remove stains from the masonry by abrasion, can etch mortar joints and remove
the outer surface of brick, resulting in permanent damage.
- Failure to protect windows, doors, and trim. Many cleaning agents, particularly acid solutions, have
a corrosive effect on metal. If permitted to come in contact with metal frames, the solutions may cause
pitting of the metal or staining of the masonry surface and trim materials such as limestone, concrete
masonry and cast stone.
CLEANING EXISTING MASONRY (STAIN REMOVAL)
Bucket and brush hand cleaning and pressurized water cleaning discussed above in Suggested Cleaning
Methods, are also used to remove stains from existing masonry. Besides these, poultices, additional proprietary
solutions and a variety of abrasive blasting methods are among the techniques typically used to remove dirt or
specific stains from existing masonry [Ref 3]. These are described briefly below.
It is always advisable to collect as much information as possible before attempting to clean existing masonry. In
some cases, water repellents may have been applied to the masonry or other unexpected treatments or condi-
tions may interfere with cleaning. In these instances, professional guidance should be sought in determining how
to address these conditions to achieve successful cleaning.
Using a Poultice
A poultice is a paste made with a solvent or reagent and an inert material. It works by dissolving a stain and
absorbing or pulling it into the poultice. Poultices tend to prevent stains from spreading during treatment and pull
stains out of the pores of brick. Poultices are normally used for stains affecting small areas of brickwork.
Poultices for cleaning masonry can be purchased commercially or made on site. The inert material used in the
poultice may be talc, whiting, fullers earth, diatomaceous earth, bentonite or other clay. The solution or solvent
used depends upon the nature of the stain to be removed. Enough of the solution or solvent is added to a small
quantity of the inert material to make a smooth paste. The paste is smeared onto the stained area with a trowel or
spatula, to make a coating at least
1
/8 in. (3 mm) thick. When dried, the remaining powder, which now contains the
staining material, is scraped, brushed or washed off. Repeated applications may be necessary.
If the solvent used in preparing a poultice is an acid, do not use whiting as the inert material. Whiting is a carbon-
ate which reacts with acids to give off carbon dioxide. While this is not dangerous, it will make a foamy mess and
destroy the power of the acid.
Abrasive Blasting
Abrasive methods are not generally recommended for cleaning brickwork. Attempting to remove dirt or stains by
abrasion is risky because the outer surface of the masonry may also be removed, resulting in permanent damage
and increased water penetration. Abrasive cleaning may also roughen the surface of masonry, which increases
its tendency to hold dirt and makes future cleaning more difficult. Sanded, coated, glazed and slurry-finished brick
should not be cleaned by abrasive blasting.
It is possible to safely clean brick masonry by abrasive blasting, however a gentler abrasive than sand and a
highly qualified operator are typically required, in conjunction with proper specifications and job inspection. In lim-
ited instances, abrasive blasting is the only method that will remove persistent stains. This method is sometimes
preferred over conventional wet cleaning since it eliminates the problem of chemical reactions with vanadium salts
and other materials used in manufacturing brick.
Abrasive blasting involves an air compressor, blasting tank, blasting hose, nozzle, and protective clothing, a hood
and a respirator for the operator. The air compressor should be capable of producing 60 to 100 psi (400 to 700
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 7 of 10
kPa) at a minimum air flow capacity of 125 ft
3
(3.5 m
3
) per minute. The inside orifice or bore of the nozzle may
vary from
3
/16 to
5
/16 in. (4.8 to 7.9 mm) in diameter. The sandblast machine (tank) should be equipped with con-
trols to regulate the flow of abrasive materials to the nozzle at a minimum rate of 300 lb/hr (0.004 kg/s).
Methods for cleaning masonry using abrasives may be executed at high or low pressures and with dry abrasives
or abrasives added to a stream of water. Abrasives should be selected based on the degree of cutting or cleaning
desired and the amount of change in the surface of the masonry that is permissible. Silica sands, crushed quartz,
crushed granite and white urn sand (round particles) are among the harder abrasives at approximately 6 on Mohs
Scale. Softer abrasives include crushed nut shells, dry ice, baking soda and others. If used these minerals should
have a gradation appropriate for the intended use [Ref. 2].
Dry abrasive blasting (sandblasting) at high pressure is perhaps the best known of these methods, but has a sig-
nificant potential to damage masonry. In addition, the dust it creates can be harmful if inhaled, which poses health
and safety concerns.
Wet sand cleaning depends on water-cushioned abrasive action for its effectiveness. It is similar to sandblasting,
with the addition of water into the air stream, which eliminates dust. It is often suggested when abrasion of the
surface is permissible. Such instances may include removal of paint or other surface coatings.
Wet aggregates delivered at low pressure through a special nozzle are sometimes used on soft brick and soft
stone materials, and are particularly effective on surfaces with flutings, carvings and other ornamentation. Wet
aggregate cleaning is a gentle but thorough process, employing a mixture of water and a friable aggregate free
from silica, with a scouring action that cleans effectively with less surface damage than sandblasting or wet sand
cleaning.
The General Cleaning Procedure can also be followed for abrasive blasting with the following modifications:
- In Step 3, select abrasives that are clean, dust free and sufficiently hard. Test clean several areas at vary-
ing distances from the wall and several angles that afford the best cleaning job without damaging brick
and mortar joints. Workers should be instructed to direct abrasive at the brick and not directly at the mor-
tar joints.
- Omit Steps 5 through 7.
REMOVING EFFLORESCENCE
The removal of efflorescing salts is relatively easy compared to some other stains. Refer to Technical Notes 23
Series for a detailed discussion on efflorescence. Efflorescing salts are water soluble and generally will disappear
of their own accord with normal weathering. This is particularly true of new building bloom, which tends to occur
shortly after construction is completed (or during construction) due to normal water loss during post-construction
drying.
Before efflorescence is removed, any leaks should be repaired and the brickwork should be allowed to dry. White
efflorescence can often be removed by dry brushing or with a stiff fiber brush and water. Heavy accumulations
or stubborn deposits of white efflorescence may be removed with a proprietary cleaner. It is imperative that the
manufacturers instructions are carefully followed.
REMOVING SPECIFIC STAINS
Whether a stain results from chemical reactions within a brick, or external materials being spilled, splattered on, or
absorbed by brickwork, each is an individual case and must be treated accordingly. When using any cleaner, it is
advisable to consult the brick manufacturer for cleaning advice, follow the instructions of the cleaner manufacturer,
and trial clean in an inconspicuous area before using on an entire project.
There are a variety of proprietary cleaners that effectively remove most of the common substances that stain
brickwork, including bronze and copper stains, efflorescence, graffiti, iron stains (rust), lime run, manganese stain,
moss, oil and tar stains, paint, smoke and vanadium stain. When available, these are preferred over site-mixed or
homemade cleaning solutions because they are generally safer, easier to control and more consistent, resulting
in successful cleaning. In some cases these cleaners have been developed in conjunction with brick manufactur-
ers.
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 8 of 10
In addition to proprietary cleaners, many stains can be removed by scrubbing with kitchen cleansers, bleach or
other household chemicals. A combination, such as is found in some kitchen cleansers, may prove most effective.
The sections below list some non-proprietary alternatives for removal of common stains. Further information on
causes and prevention of stains is contained in the Technical Notes 23 Series.
Brick Dust
Dust produced from the cutting of brick sometimes adheres to the surface of brickwork. Compressed air, such as
from a portable cylinder, has been found effective in removing this dust.
Dirt and Mud
Dirt can be difficult to remove, particularly from a textured brick. In addition to proprietary cleaners, scouring pow-
der and a stiff bristle brush are effective if the texture is not too rough. For very rough textures, pressurized water
cleaning can be effective.
Egg Splatter
Brickwork vandalized with raw eggs has been successfully cleaned by prewetting the stain, applying a saturated
solution of oxalic acid crystals dissolved in water and rinsing with water. Mix the solution in a non-metallic con-
tainer and apply with a brush.
If the egg splatter is to be removed from brick that contain vanadium (typically light colored units), a solution of
1.5 oz (10 g) washing soda (sodium carbonate) per gal (1 L) of water should be applied to the brickwork following
the oxalic acid solution. Without this neutralizing solution, cleaning with oxalic acid may cause more severe stain-
ing.
Manganese (Brown) Stain
Besides specially formulated proprietary compounds, manganese stains have been effectively removed and their
return prevented by carefully mixing a solution of acetic acid (80 percent or stronger), hydrogen peroxide (30 to
35 percent) and water in the following proportions by volume: 1 part acetic acid, 1 part hydrogen peroxide, and 6
parts water. After wetting the brickwork, brush or spray on the solution. Do not scrub. The reaction is usually very
rapid and the stain quickly disappears. After the reaction is complete, rinse the wall thoroughly with water.
Caution: Although this solution is very effective, it is a dangerous solution to mix and use. Acetic acid-hydrogen
peroxide may also be available in a premixed form known as peracetic acid.
An alternate treatment sometimes suggested for new and mild manganese stains is oxalic acid crystals and water.
Mix 1 lb of crystals (0.45 kg) to 1 gal (3.79 L) of water. The neutralizing wash mentioned above in Egg Splatter
should be considered when oxalic acid is applied to brown or light colored brick.
Oil and Tar Stains
Oil and tar stains may be effectively removed by commercially available oil and tar removers. For heavy tar stains,
mix the agents with kerosene to remove the tar, and then water to remove the kerosene. After application, the
stains can be hosed off. When used in a steam cleaning apparatus, cleaners have been known to remove tar
without the use of kerosene.
Where the area to be cleaned is small, or minimal cleanup is desired, a poultice using naphtha or trichloroethylene
is most effective in removing oil stains.
Dry ice or compressed carbon dioxide may be applied to make tar brittle. Then, light tapping with a small hammer
and prying with a putty knife generally will be enough to remove thick tar splatters.
Organic Growth
Occasionally, an exterior masonry surface remains in a constantly damp condition, thus encouraging moss, algae,
lichen or other organic growth. Applications of household bleach, ammonium sulfate or weed killer, in accordance
with furnished directions, have been used successfully for the removal of such growths.
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 9 of 10
Paint and Graffiti
Commercial and proprietary paint removers and organic solvents are most effective at softening or dissolving
paint so that it can be removed with a scraper and a stiff bristle brush or rinsed away with water. For very old dried
paint, organic solvents may not be effective, in which case the paint must be removed by sandblasting or scrub-
bing with a non-metallic abrasive pad. Graffiti that has penetrated into masonry is best removed by a poultice,
paste or gel that can cling to the masonry, extending its working time on the stain.
Smoke
Scrubbing with scouring powder (particularly one containing bleach) and a stiff bristle brush is often effective.
Vanadium (Green) Stain
Applying a solution of potassium or sodium hydroxide, consisting of 0.5 lb (0.23 kg) hydroxide to 1 qt (0.95 L)
water or 2 lb (0.91 kg) per gal (3.79 L) to brickwork is an alternative treatment for vanadium stains. The solution
should be allowed to remain for two or three days and then washed off. Use a hose to wash off any white residue
remaining on the brickwork after this treatment.
Sodium hypochlorite, the active ingredient in household bleaches, can also be used to remove mild vanadium
stains. Spray or brush onto the stain, and then rise off after the stain disappears.
Oxalic acid is another chemical known to remove vanadium stains. A mixture of 3 to 6 oz (20 to 40 g) oxalic acid
per gal (1 L) of water (preferably warm) should be applied to the brickwork, followed by the neutralizing wash
described in Egg Splatter. More severe staining may result if the oxalic acid solution is applied without the neu-
tralizing wash.
Welding Splatter
When metal is welded too close to brick stored on site or completed brickwork, molten metal may splash onto the
brick and melt into the surface. A mixture of 1 lb (0.45 kg) oxalic crystals and 0.5 lb (0.23 kg) of ammonium biflou-
ride per gal (3.79 L) of water is particularly effective in removing welding splatters. This mixture should be used
with caution as it generates dangerous hydroflouric acid, which can also etch brick and glass.
Scrape as much of the metal as possible from the brick. Apply the mixture in a poultice, and remove when it is
dried. If the stain has not disappeared, use sandpaper to remove as much as possible and apply a fresh poultice.
For stubborn stains, several applications may be necessary.
Stains of Unknown Origin
Stains of unknown origin can be a real challenge. Laboratory tests of unknown stains maybe necessary to deter-
mine their composition. Then the appropriate method may be implemented to clean the brickwork. The application
of a cleaning agent without identifying the initial stain may result in stains that are more difficult to remove. The
visual characteristics of a stain may be the first clues as to its source. Identification of stains is discussed further in
Technical Note 23.
CLEANING HISTORIC STRUCTURES
Improper cleaning can cause irreparable damage to historic brickwork. Therefore, cleaning of structures with his-
toric significance should be overseen by a restoration specialist. Before a historic structure is cleaned, consider
the purpose of cleaning: to improve the appearance; to slow deterioration; or to provide a clean surface for evalu-
ation or further treatments. With historic structures, it is imperative to use the least harmful cleaning method that
will achieve the desired results. Cleaning methods and materials must be carefully matched to the substance to
be cleaned, the type of soiling/staining to be removed and the desired results.
These issues are discussed in detail in Assessing Cleaning and Water-Repellent Treatments for Historic Masonry
Buildings [Ref. 4].
CLEANING BRICK PAVING
Cleaning of brick pavements is essentially the same as for brickwork in walls and other applications. The methods
described above can be used successfully to remove stains that also affect pavements such as efflorescence,
hardened mortar, plant life, oil and tar, etc. However, after construction is complete, dirt and stains resulting from
deicing salts or materials tracked onto or spilled on pavements typically build up more quickly than other brick
applications. Frequent sweeping and washing with clean water will help reduce the need for more aggressive
cleaning methods and solutions.
Fresh mortar stains can be removed from existing or mortarless pavements before they set by covering the pave-
ment with clean, slightly damp, washed sand and sweeping toward the edges. When the surface is almost clean,
sweeping with dry sand should remove remaining residue.
Chewing gum can usually be removed from brick pavements by wire brushes, carefully applied high pressure
water or freezing each piece of gum with compressed carbon dioxide or dry ice, and then scraping or chiseling it
off the pavement. Food stains and tire marks are typically removed by scrubbing with a detergent or proprietary
cleaner.
SUMMARY
Testing of cleaning procedures and chemicals as suggested in this Technical Note is strongly recommended. Such
testing should be performed under conditions of temperature and humidity that closely approximate the conditions
under which the brick masonry will be cleaned. Cleaning solutions recommended by the brick or cleaning agent
manufacturer should also be trial tested before being committed to an entire project. The effects of any cleaning
process on people and the environment should be carefully evaluated before cleaning begins.
The recommendations in this Technical Note should be used as a guide for successful cleaning of brick masonry.
Due to the diverse nature of cleaning solutions, procedures and problems, the Brick Industry Association cannot
accept responsibility for the final success or effectiveness of these procedures.
In conclusion, nothing is quite as effective as careful attention exercised during construction to keep brickwork
relatively clean. If this is successful, it will eliminate the need for costly cleaning procedures.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. Clay Paving Design and Construction, Clay Brick and Paver Institute, Baulkham Hills BC, Australia, 2003.
2. "Good Practice for Cleaning New Brickwork," Brick SouthEast, Charlotte, NC, 2003.
3. Grimm, C.T., Cleaning Masonry - A Review of the Literature, Construction Research Center, University
of Texas at Arlington, Arlington, TX, 1988.
4. Mack, R.C. and Anne Grimmer, Assessing Cleaning and Water-Repellent Treatments for Historic
Masonry Buildings, Preservation Briefs No. 1, National Park Service, Washington, DC, 2000.
www.gobrick.com | Brick Industry Association | TN 20 | Cleaning Brickwork | Page 10 of 10

Technical Notes 21 - BRICK MASONRY CAVITY WALLS
August 1998
Abstract: This Technical Notes covers brick masonry cavity walls. Description of the properties of
cavity walls, including structural properties, water penetration resistance, fire resistance, and thermal
and sound transmission properties are included. Theories for structural design are introduced.
Recommendations for cavities, flashing, expansion joints, ties and other related subjects are covered.
Key Words: brick, cavity wall, design, expansion joints, fire resistance, flashing, structural, thermal
resistance, ties, veneer.
INTRODUCTION
Brick masonry cavity walls consist of two wythes of masonry separated by an air space connected
by corrosion-resistant metal ties (see Fig.1). The exterior masonry wythe can be solid or hollow
brick, while the interior masonry wythe can be solid brick, hollow brick, structural clay tile, or hollow
or solid concrete masonry units. The selection for each wythe depends on the required wall
properties and features. A cavity of 2 to 4 1/2 in. (50 to 114 mm) between the two wythes may be
either insulated or left as an air space. The interior surface of the cavity wall may be left exposed or
finished in conventional ways.
Typical Cavity Wall
FIG. 1
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
1 of 15 9/13/2009 1:07 PM
Cavity walls, long common in Europe, were first built in the United States as early as 1850.
However, it was not until 1937 that this type of construction gained official acceptance by any building
code or construction agency in the U.S. Since then, interest in and use of cavity walls in this country
has increased rapidly. Extensive testing and research and empirical evidence of existing cavity wall
construction has been used to determine cavity wall properties and performance. Cavity walls are
often regarded as the premier masonry wall system.
The early use of cavity walls in this country was limited primarily to exterior load-bearing walls, one
and two-stories in height. In the 1940's, designers of high-rise buildings began to recognize the
advantages of cavity walls and used them as curtain and panel walls in structural frame buildings.
Today, masonry cavity walls are used extensively throughout the United States in all types of
buildings. See Fig. 2. The primary reasons for their popularity are: excellent rain penetration
resistance, fire resistance, thermal capabilities and sound transmission resistance.
A brick cavity wall is differentiated from a brick veneer with masonry backing by how the designer
considers load resistance by the exterior wythe. The exterior wythe of a cavity wall is designed to
resist loads by stresses developed in that wythe. Further, both wythes resist out-of-plane loads by
stresses in each wythe. These stresses, whether axial, flexural or shear, must be less than the
corresponding allowable stresses. The exterior wythe of a brick veneer wall transfers out-of plane
loads to the backing and is not subject to limitations of the allowable stress values. No axial loads
are applied to the veneer wythe. Out-of plane lateral loads are transferred by metal ties to the
backing which is designed for the full load. Shear stresses generated by the veneer's weight are
ignored. Other design issues, such as water penetration resistance, fire resistance, thermal, and
sound transmission, are the same for either brick masonry cavity walls or brick veneer over a
masonry backing; therefore, such information in this Technical Notes is appropriate for both types of
wall systems.
Some parts of the country use the term "reinforced cavity walls" to denote a multi-wythe masonry
wall with grout placed between the wythes. This should actually be considered a multiwythe grouted
masonry wall. Since the definition of a cavity wall includes an air space, this type of wall is not truly a
cavity wall.
This Technical Notes discusses the properties of cavity walls, and the proper design to achieve
these properties. Other issues in this Technical Notes series deal with materials, detailing and
proper construction practices for cavity walls.
Carnegie Hall Tower,
New York City
Fig. 2
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
2 of 15 9/13/2009 1:07 PM

PROPERTIES OF CAVITY WALLS
Structural Properties
Properly designed, detailed and constructed cavity walls may be used in any building requiring
load-bearing or non-loadbearing walls. The increased flexibility by the separation of the wythes and
the use of metal ties permits more freedom from differential movement between the wythes. This is
extremely important in today's construction which makes use of many combinations of dissimilar
materials.
The structural behavior of cavity walls is complex because of the interaction of the wythes, ties and
support conditions. Typically, the inner wythe of a cavity wall is designed to support the weight of
floors, roofs and live loads. The outer wythe is mainly non-loadbearing. Out-of plane loads are
shared by the wythes in proportion to their stiffnesses and the stiffness of the connecting ties. Walls
tied together by brick headers (masonry bonded hollow walls or utility walls) behave differently from
walls tied together by metal ties. Therefore, this Technical Notes addresses only metal-tied walls.
Information on masonry bonded hollow walls (utility walls) can be found in other technical literature
[1].
Resistance to Moisture Penetration
One of the major functions of an exterior wall is to resist moisture penetration. A brick masonry cavity
wall, properly designed and built, is virtually resistant to water penetration through the entire wall
assembly. The outside wythe may permit some moisture penetration, but the overall design of the
cavity wall assembly accommodates this expected infiltration. It should be assumed that wind-driven
rain will penetrate the exterior wythe of brick masonry. A cavity wall is designed as a drainage wall
system, so that any moisture which does penetrate the exterior wythe will run down the back face of
the exterior wythe to the bottom of the cavity where it is diverted to the outside by flashing and weep
holes. For further discussion of moisture penetration, refer to the Technical Notes 7 Series.
Rain Screen Walls
Cavity walls can also be designed as pressure-equalized rain screen walls. This wall system
provides compartmented air spaces with vents at the top and bottom of the cavity allowing wind
pressures to equalize between the cavity and the exterior. In theory, the outer wythe is essentially
an open rain screen that eliminates water penetration due to air pressure differences. Although
pressure-equalized rain screen walls can provide increased resistance to water penetration, they are
more difficult to design, detail and construct. They are typically used in projects located in areas
which receive high volumes of wind-driven rain and when resistance to water penetration is of prime
concern. Refer to Technical Notes 27 for more information on pressure-equalized rain screen walls.
Condensation
Although moisture penetration due to wind-driven rain may be a major concern, condensation may
also be a problem in certain climates and occupancies. Differences in humidity between inside and
outside air will cause vapor flow within the wall and, unless controlled, this vapor may condense
within the wall under certain temperature conditions. This condensation may contribute to
efflorescence when soluble salts are present, corrosion of metal ties or disintegration of the masonry
units. A condensation analysis, as described in Technical Notes 7D, will help determine the points in
a wall system where condensation may occur. Information found in the Air Barriers and Vapor
Retarders section of this Technical Notes and Technical Notes 7C should be used for the design
against condensation.
Thermal Properties
Heat losses and heat gains through masonry walls can be minimized by the use of cavity wall
construction. The separation of the exterior and interior wythes by the cavity eliminates or reduces
thermal bridging and allows a large amount of heat to be absorbed and dissipated in the outer wythe
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
3 of 15 9/13/2009 1:07 PM
and cavity before reaching the inner wythe and the building interior. The cavity provides an excellent
location to incorporate insulation in the wall assembly. Insulation can be placed in the air space by
using a rigid board attached to the backing or by completely filling the air space with granular fill or
foam. Insulation may also be placed in the cores of the backing wythe, such as granular fill in
concrete block. Finally, the interior side of the cavity wall may be finished with insulation and gypsum
board.
Steady-state U-values can range from 0.33 BTU/ (hr
o
F
o
ft
2
) [1.9 W / m
2

o
K ] for an 8-in. (200 mm)
uninsulated cavity wall to 0.06 BTU/ (hr
o
F
o
ft
2
) [0.34 W / m
2

o
K) for a 16 in. (400 mm) cavity wall
with 2-in. (50 mm) polyisocyanurate board insulation. Table 1 lists R-values for typical cavity walls.
Table 2 lists R-values of the materials used in brick masonry cavity walls. Table 3 lists thermal
properties of interior finishes used in combination with cavity walls. For example, using Table 3,
combine a 12 in. (300 mm) brick and block cavity wall with extruded polystyrene in the cavity
(R-value of 9.5), with 11/2 in. (38 mm) extruded polystyrene insulation between metal furring strips
at 16 in. (400 mm) o.c. (R-value of 4.6) resulting in a total R-value of 14.1. Thermal properties are
further enhanced when considering the mass effect of masonry. Thus, considerable energy savings
can be realized by the use of cavity walls. Refer to Technical Notes 4 Series for U-values of
materials and assemblies.
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
4 of 15 9/13/2009 1:07 PM

Thermal Mass
Brick masonry exhibits superior thermal mass, that is, the ability to store and slowly release heat at a
later time. These properties help shift the peak heating or cooling loads to off-peak times and
reduce the peak temperatures. Current energy codes take thermal mass into account by requiring a
lower R-value for mass walls, such as brick cavity walls (see Technical Notes 4B).
Passive solar design can be used in conjunction with cavity wall construction to take full advantage of
masonry's thermal mass properties. Solar design techniques, such as the use of building orientation,
daylighting and thermal mass, can provide both comfort and energy savings for the owner and
occupants. It is best to incorporate passive solar techniques in a building during the preliminary
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
5 of 15 9/13/2009 1:07 PM
design phase. Technical Notes 43 Series has information on passive solar design techniques. Other
references are available to assist in design [4, 8].
Fire Resistance
The results of fire resistance tests clearly show that brick masonry cavity walls have excellent fire
resistance. Fire resistance ratings of brick masonry cavity walls range from 2 to 4 hr, depending
upon the wall thickness and other factors (see Table 4). Due to their high fire resistance properties,
brick walls make excellent fire walls or building separation walls for compartmentation in buildings.
By using compartmentation, the spread of fire can be halted. Technical Notes 16 Series describes
fire ratings and applicable design conditions.
An alternative way of determining the fire resistance of a cavity wall assembly is by using the
calculated fire resistance method. This approach is approved by the model building codes for
determining fire ratings of walls that are not physically tested by ASTM E 119 Test Methods for Fire
Tests of Building Construction and Materials. The fire rating of cavity walls can be calculated using
Technical Notes 16B.
Sound Transmission
Resistance to transmission of sound in masonry construction is accomplished in two ways: the use of
heavy massive walls or the use of discontinuous construction. The cavity wall employs both
techniques, i.e., the massiveness of the two masonry wythes plus the partial discontinuity of the
cavity.
In cavity wall construction, the air space provides a partial isolation of the two wythes. Sound on one
side of a cavity wall causes vibration of a wythe. Because of the separation and cushioning effect of
the air space and the massiveness of the masonry, the vibration is dampened and greatly reduced.
A 10-in. (250 mm) brick cavity wall with brick backing has a Sound Transmission Class (STC) rating
of 50, which is usually sufficient for substantially reducing typical outside noises entering the building
through the wall. For more information on sound transmission, see Technical Notes 5A.
DESIGN OF CAVITY WALLS
The successful design of cavity walls depends on proper attention to four elements: appropriate
design, proper detailing, selection of quality materials and execution of good workmanship. All four
elements must be satisfied to produce a successful cavity wall.
Structural Design
There are many various design procedures for cavity walls. Both rational and empirical design
methods are used. Rational design methods currently use working stress analysis. Limit states or
strength design methods are now being written. Other methods such as a modified yield line
approach, fracture line approach and elastic performance of cracked panels have been proposed for
the design of unreinforced walls. However, insufficient data and lack of recognition by model building
codes relegate these methods to use as special systems for design.
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
6 of 15 9/13/2009 1:07 PM
The structural design of cavity walls should follow either rational or empirical methods. The rational
design method is based on the properties of the wall materials and engineering analysis. This
method may be used for any structure where high loads are likely or where tall walls are a
necessity. The empirical method is generally satisfactory for one and two-story buildings consisting
of light construction with limited floor spans and wall heights; and for multistory buildings where
unsupported wall heights are not excessive.
Rational Design
The design of cavity walls is governed by model building codes. Most of these reference the ACI
530/ASCE 5/TMS 402 Building Code Requirements for Masonry Structures, also known as the
Masonry Standards Joint Committee (MSJC) Code [2]. The Uniform Building Code incorporates
similar design requirements for cavity walls [9]. Seismic design requirements are also stipulated in
these building codes based on site location. A detailed design of cavity walls is not covered in this
Technical Notes. For design aids and examples of design procedures, several books listed in the
REFERENCES section should be consulted [5, 7].
The MSJC Code requirements differentiate between multiwythe walls as those with composite or
non-composite action. Composite action requires a transfer of shear stress between wythes so that
the wythes act as a single element in resisting loads. The wythes must be bonded with a filled collar
joint and metal ties or with masonry headers. Therefore, all cavity walls are classified as
non-composite walls and must be designed using those requirements. When non-composite action
occurs, each wythe is designed to individually resist the effects of imposed loads. In order for both
wythes to carry part of the vertical load, the floor or roof system must bear on both wythes or on a
spreader beam bearing on both wythes. Typically, the wythe closest to the center of the span of
horizontal members resists the resulting vertical load and the associated shear forces. Out-of plane
loads are apportioned to wythes based upon their relative stiffnesses.
The rational design method used in the MSJC Code is allowable stress design. Individual wythes
may be either reinforced or unreinforced. See the Technical Notes 3 Series for more information on
the MSJC Code.
Empirical Design
Chapter 9 of the MSJC Code presents empirical requirements for masonry structures. These
requirements are based on past proven performance and pre-dates rational design methods.
According to the 1995 edition of the MSJC Code, the empirical requirements in Chapter 9 may be
applied to the following masonry elements:
1.The main lateral force resisting system for buildings in Seismic Performance Category (SPC) A and
all other masonry elements for buildings in SPC A, B and C.
2.Buildings subject to a wind velocity pressure not exceeding 25 psf (1.2 kPa).
3.Buildings not exceeding 35 ft (10.7 m) in height when the masonry walls are part of the main
lateral load resisting system.
The empirical requirements may not be applied to structures resisting horizontal loads other than
those due to wind or seismic events.
The MSJC Code contains limits on the ratios of wall thickness to distance between lateral supports.
These limits provide controls on the flexural tensile stresses within the wall and limit possible buckling
under axial loads. Maximum h/t or l/t ratios and minimum thicknesses used for determining distances
between lateral supports are consistent with past masonry standards. Definitions for height (h),
length (l) and thickness (t) for use in the lateral support ratios for cavity walls are as follows:
h =the vertical distance or height between lateral supports;
l =the horizontal distance or length between lateral supports; and
t =the sum of the nominal thicknesses of the inner and outer wythes.
Further information on empirical design of brick masonry is provided in Technical Notes 42 Revised.
Seismic Issues
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
7 of 15 9/13/2009 1:07 PM
As with other loading, the seismic loads from other elements applied to the wythes are resisted by
the loaded wythes. This requires each loaded wythe to meet prescriptive seismic reinforcement
requirements. For seismic loading due to wall weight, if the tie spacing and stiffnesses are
adequate, the wythes will share the out-of-plane load in proportion to their relative rigidities. In such
instances, the seismic force can be resisted by just one wythe. In many cases, the inner wythe is
reinforced according to the minimum reinforcement requirements prescribed by the code and the
brick exterior wythe is treated as a veneer. Proper ties, reinforcement or isolation joints may be
required in specific seismic performance categories (zones).
Architectural Design
Cavity
The cavity or air space between wythes should be between 2 in. (50 mm) and 4 1/2 in. (114 mm).
Air spaces less than 2 in. (50 mm) can not practically be kept free from mortar bridging. Air spaces
greater than 41/2 in. (114 mm) do not allow the normally prescribed ties to properly transfer lateral
loads. Air spaces different from these can be used, but care in design and construction would be
required. If larger air spaces are used additional ties and/or thicker ties may be necessary. When
rigid board insulation is placed in the air space, the clear distance from the back side of the brick to
the exterior side of the insulation must be no less than 1 in. (25 mm). This allows the mason to lay
the brick properly and for the wall to still function as a drainage wall. Mortar protrusions should not
contact the insulation as this is a direct path for water. These air space requirements also allow for
tolerances in construction.
Since it is assumed that water never crosses the air space, parging of the cavity face of either wythe
is neither necessary, nor recommended. In cases where mortar could bridge the cavity, where
mortar droppings fill the bottom of the cavity due to poor workmanship, or when a pressure-equalized
rain screen wall is designed, a coating or membrane on the exterior face of the masonry backing
may be prudent. To be effective, the coating or membrane must be continuous over the backing and
be sealed at interfaces with other elements or materials. Consideration must be given to the effects
of movement on the coating or membrane.
Flashing and Weep Holes
Flashing and weep holes collect water that enters a wall and directs it back to the exterior. Flashing
should be provided at the base of the wall, above and below all wall openings, at the tops of walls,
beneath copings, and any other discontinuities in the air space, such as recessed courses or shelf
angles. Fig. 3 shows typical flashing locations.
Since flashing is a bond break, the tensile strength of the wall at that location is assumed to be zero.
The shear stress will also be reduced. The structural design of the building should address these
issues appropriately.
Flashing and weep holes should be located above grade level. If the wall continues below the
flashing at the base of the wall, the space between the exterior wythe and the interior wythe should
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
8 of 15 9/13/2009 1:07 PM
be filled with mortar or grout to the elevation of the flashing.
Flashing should be securely fastened to the interior wythe and extend through the face of the exterior
brick wythe. The flashing should be turned up at least 8 in. (200 mm) and embedded in the inner
wythe. Flashing should be carefully installed with no punctures or tears. Where flashing is required
to be lapped, the ends of the flashing should be overlapped a minimum of 6 in. (150 mm) and the
laps properly sealed to avoid water running between the sections. Where the flashing is not
continuous, such as over and under openings in the wall, the ends of the flashing should be turned up
approximately 1 in. (25 mm) into a head joint in the exterior wythe to form an end dam.
Prefabricated flashing pieces may be available to help form the detail. Typical flashing details are
shown in Technical Notes 21B Revised and 7 Revised.
Weep holes must be located immediately above all flashing. Open head joint or vent weep holes
should be spaced no more than 24 in. (600 mm) o.c. Weep holes formed with wick materials or with
round tubes should be spaced at a maximum of 16 in. (400 mm) o.c.
Drainage materials, such as pea gravel or plastic mesh, can be used at the base of the cavity when
mortar droppings are a high probability. Although these drainage materials are not always
necessary, they may help keep the cavity and the weep holes from being totally blocked by mortar.
However, the use of these materials does not negate the importance of good workmanship and
keeping the cavity clean from mortar droppings. When pea gravel is used, its depth should not
exceed 2 to 3 in. (50 to 75 mm). Consideration should be given to the weight of the gravel on the
flashing, the presence of salts in the gravel and the size of the pea gravel to keep it from flowing out
of open weep holes. These types of materials may interfere with walls that are designed for air
circulation in the cavity.
Ties
Wall ties provide a connection between the inner and outer wythes of a cavity wall, and may
accommodate differential movements between the wythes. Tie spacing requirements for cavity walls
differ slightly from those of brick veneer with masonry backing. Non-composite walls (cavity walls)
designed in accordance with the MSJC Code permit a maximum spacing of 36 in. (910 mm) o.c.
horizontally and 24 in. (610 mm) o.c. vertically. In addition, the spacing for W1.7 size ties must be no
more than one tie for every 2.67 ft
2
(0.25 m
2
); and for W2.8 size ties, no more than one tie for every
4.5 ft
2
(0.42 m
2
). Adjustable ties must be spaced at no more than 16 in. (400 mm) o.c. in each
direction and no more than one tie for every 1.77 ft
2
(0.16 m
2
).
When designing the exterior wythe as a veneer, the tie spacing depends on the type of tie. For
adjustable two-piece ties and ties with a wire size of W 1.7, the spacing should be no more than one
tie for every 2.67 ft
2
(0.25 m
2
); and for all other ties, no more that one tie for every 3.5 ft
2
(0.33
m
2
). The maximum spacing for ties in a brick veneer are 32 in. (810 mm) o.c. horizontally and 18 in.
(460 mm) o.c. vertically. Drips in the ties reduce the strength of the tie, therefore, the wall area per
tie must be reduced by 50 percent. More information on ties can be found in Technical Notes 44B.
Foundations
Foundations of brick, concrete masonry or concrete are used to support brick masonry cavity walls.
It is recommended that the thickness of the foundation or foundation wall supporting the cavity wall
be at least equal to the total thickness of the cavity wall less 2 in. (50 mm). The brick exterior wythe
may be corbeled out over the top of the foundation; however, the overhang may not be more than
one-third of the brick's thickness nor more than one-half its height per course.
A bond break is recommended between the exterior brick wythe and the foundation. Differential
movement between the brick and concrete or concrete masonry foundation and cracks emanating
from the foundation can be avoided by placing a bond break, such as a flashing material, on top of
the foundation. However, it must also be remembered that the flashing or bond break changes the
end condition of the wall. The wythes separated from the foundation by flashing should be designed
as a simply-supported or pinned end walls, not as a fixed end condition. If the exterior brick wythe is
designed as a veneer then this is not a concern.
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
9 of 15 9/13/2009 1:07 PM
The exterior brick wythe should start above grade. Brick used below grade must be properly
waterproofed with a waterproof membrane or coating. Weep holes should not be placed below
grade.
Shelf Angles
Shelf angles are used in buildings to accommodate vertical movement by allowing an expansion joint
to be placed beneath the angle. Shelf angles must be accompanied with horizontal expansion (soft)
joints, as shown in Fig. 4. The need for shelf angles in a building depends on the type of structure,
height of building, location of windows, window size and other factors. The decision to use shelf
angles should be based on the effects of differential movement, type of tie system and connection of
the exterior brick wythe to other building components. The MSJC Code gives no limits for the
placement of shelf angles in cavity walls or in brick veneer with backings of concrete or concrete
masonry. For low-rise structures, less than 30 ft (9.1 m), it is advisable to have the brick wythe bear
on the foundation wall. For multi-story structures the effects of differential movement should be
considered. The local building code may dictate the placement of shelf angles and their required
height from the foundation.
Care should be taken to ensure proper anchorage and shimming of the angles to prevent excessive
deflection or rotation. These movements may create problems in construction and induce
concentrated loads in the masonry below. The maximum deflection and rotation should be no more
than the size of the space below the angle.
Shelf angles should not be installed as one continuous member. Spaces should be provided at
intervals to permit thermal expansion and contraction of the steel to occur without causing distress to
the masonry. Shelf angles must provide continuous support around corners. As at the foundation, at
least 2/3 of the brick wythe thickness must bear on the shelf angle.
Expansion Joints
All building materials change dimension with changes in temperature. Most building materials, with
the exception of glass and metals, change dimension with changes in their moisture content. Brick
undergoes a permanent moisture expansion. All materials will deform when subjected to loads. And,
some materials, notably those with cement matrices, will deform plastically (creep) when loaded.
Building frames are subject to movements from applied vertical and horizontal loads. Concrete and
concrete masonry frames and members, in addition to the above, are subject to drying shrinkage.
Since many cavity walls are built of dissimilar materials, there will be differential movement between
these materials. All of these movements must be considered in the design and construction of the
wall system. Problems, such as cracking, can arise if these movements are not recognized and
accommodated.
In order to prevent distress in the masonry, vertical and horizontal expansion joints are used to
accommodate movements. No single recommendation for the positioning and spacing of vertical and
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
10 of 15 9/13/2009 1:07 PM
horizontal expansion joints can be applicable to all structures. Each building must be analyzed to
determine the potential movements, and provisions must be made to permit such movement or to
resist stresses resulting from such movements. Expansion joints must also be designed, located and
constructed so not to impair the integrity of the wall. The movement of the exterior brick wythe due
to thermal and moisture expansion may be greater than the movement in solid or composite walls
exposed to the same environment. This is due to the greater temperature differences between
wythes and the absence of vertical loads. For further information, see Technical Notes 18 Series.
A distinction should be made between the use of the terms expansion joint and control joint. An
expansion joint is used to separate brick masonry into segments to prevent cracking due to an
increase in size. The joints are formed of highly elastic materials placed in a continuous,
unobstructed opening through the brick wythe, see Fig. 5. This allows the joints to close as a result
of an increase in size of the brickwork. A control joint is used in concrete or concrete masonry to
create a plane of weakness which, used in conjunction with reinforcement or joint reinforcement,
controls the location of cracks due to a reduction in volume resulting from shrinkage and creep. A
control joint is usually a vertical opening through the concrete masonry wythe and may be formed of
inelastic materials. A control joint will open rather than close. This distinction is made so that the
wrong type of joint is not used in either material. These vertical movement joints do not have to be
placed at the same location in a wall.
Horizontal Expansion Joints
If the brick is supported on shelf angles attached to the structural frame, horizontal expansion (soft)
joints should be placed immediately beneath each angle. This is particularly important in reinforced
concrete frame buildings because of frame shortening. Horizontal expansion joints may be
constructed by leaving an air space or by placing a compressible material under the shelf angle.
Using a compressible pad, such as a preformed foam pad may assist in keeping debris out of that
space. Mortar should never be placed in this space. In either case, the joint must be sealed at the
exterior face with a suitable elastic sealant and backer rod.
Parapets
Of all the masonry elements used in buildings, probably the most difficult to adequately detail is the
parapet wall. Designers have tried many different ways to design parapets to minimize cracking,
leaking and displacement. Some experts believe that the only sure way to avoid parapet problems is
to eliminate the parapet altogether. However, they are frequently required by building codes or for
architectural or fire safety considerations.
The detail shown in Fig. 6 is suggested as one method of building a parapet. For cavity wall
construction, it is recommended that the cavity continue up through the parapet, thereby maintaining
the separation between the outside wythe and the inner wythe. In addition, the backing wall of the
parapet may need to be reinforced and attached to the structural frame. Expansion joints should
extend up through the parapet. Expansion joints should also be placed near corners to avoid
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
11 of 15 9/13/2009 1:07 PM
displacement of the parapet.

Cavity Wall Parapet
FIG. 6
Parapet copings should provide a drip on at least one side of the wall. Metal, stone, and fired clay
copings of various designs usually provide this feature. The back side of the parapet should be
constructed of durable materials, preferably the same material that is used in the front side of the
parapet. They should not be painted or coated, but must be left free to "breathe." Unless copings
are impervious with watertight joints, place through-wall flashing in the mortar bed immediately
beneath them and firmly attach the coping to the wall below with mechanical anchors. The joints
between precast concrete and stone copings should be raked out to two times the width of the joint
and filled with a backer rod and sealant.
Thermal Design
Since energy costs to heat and cool a building over its life are greater than the initial construction
cost of the building, good energy design is important. Local climate and building codes may dictate
the level of thermal performance necessary for a particular building. Depending on these
requirements, insulation may need to be included in the cavity wall. Heavy-weight walls, such as
cavity walls, have good thermal mass properties; therefore, required R-values are less than those
required for lighter-weight walls.
When insulation is necessary in brick masonry cavity walls, several locations can be used. Insulation
can be placed: 1) in the cavity; 2) in the cells of hollow unit backing; or 3) on the inside of the interior
wythe using studs or furring strips. The location of the insulation influences the energy performance
of the building and the drainability and constructability of the wall.
Newer types of insulation strategies, such as foam-filled cavities, drainage mats attached to
insulation and drainable mineral fiber insulation used in the air space should have a good track record
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
12 of 15 9/13/2009 1:07 PM
of in-place performance before they are considered. Manufacturer's technical data should be
carefully reviewed.
Thermal Bridges
A thermal bridge is a component or assembly through which heat (or cold) is transferred at a
substantially higher rate than through the surrounding wall. A thermal bridge can degrade a wall's
thermal performance or allow condensation to occur more easily. Examples of thermal bridges in
cavity wall construction are slab edge details, corner columns and roof parapets (see Fig. 7). Metal
ties can be considered a thermal bridge; however, the effect of ties are negligible. Insulation
placement is usually the easiest way to avoid thermal bridging.

Thermal Bridges in Cavity Walls
FIG. 7
Air Barriers and Vapor Retarders.
Air leakage through the building envelope can severely degrade the thermal performance of a wall,
and increase space conditioning loads and the chance for condensation. Air leakage carries heat
and moisture between the inside and outside of the building. Infiltrating air is not filtered or
conditioned, and its rate can not be controlled. The control of air leakage is important to the thermal
performance of the building. The need for air barriers and vapor retarders is dependent upon
climate, building use and the construction assembly. What works in one situation may not be
appropriate for another. Therefore, each building must be examined on a case-by-case basis.
Typically, extreme climates that have very cold winters or very humid summers are candidates for air
retarders and vapor barriers. Moderate climates are less likely to require these membranes. In
many cases, an air barrier may be more effective than a vapor retarder since air leakage can carry
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
13 of 15 9/13/2009 1:07 PM
several hundred times more water vapor than vapor movement.
Air barriers must provide continuity, airtightness, structural integrity and durability. Since the
placement of an air barrier is usually not critical, it is often located on the exterior face of the inner
wythe. This allows for a solid surface on which to attach (adhere) the air barrier, although there may
be discontinuities at the structural frame or penetrations through the wall. For walls with insulation
placed in the air space, the insulation protects the air barrier from temperature extremes. Continuity
of the air barrier is of utmost importance. The air barrier must be kept free from punctures and
tears, and it must be continuous over the entire wall assembly by sealing any penetrations. Materials
include parging, elastomeric membranes or insulation. More information on materials is found in
Technical Notes 21A Revised.
Vapor retarders are used to slow or stop the rate of water vapor transmission through a wall. An
effective vapor retarder decreases the potential for condensation in a wall. A low vapor permeance
classifies the material as a vapor barrier. Although this value is usually taken as less than 1 perm (57
ng/Pa
o
s
o
m
2
), the perm rating should be based on the entire wall design and the vapor pressure
difference across the wall. A vapor retarder should have the same characteristics as an air barrier,
i.e. continuity, low permeance, structural integrity and durability. The preferred location of the vapor
retarder is on the high vapor pressure side of the wall. Most people know the saying "on the warm
side of the wall." In climates dominated by heating, the vapor retarder would be located on the
inside of the wall and in cooling climates near the exterior. These rules are often too simplistic and
do not cover many parts of the country. Therefore, it is suggested that a condensation analysis be
run on the wall in question based on the local climate. This will provide an accurate depiction of any
potential problem. Technical Notes 7C and 7D discuss condensation and analysis procedures.
A potential problem can occur where the construction materials and their location may create a
double vapor retarder. This may trap water between the two barriers with no chance for
evaporation. When specifying materials, such as interior wall coverings, make sure the permeance is
high enough to avoid the entrapment of vapor. A membrane may serve as both an air barrier and
vapor retarder. This is true of many elastomeric coatings used in cavity wall construction. In this
case, use the more critical design requirement and apply that to that material.
SUMMARY
This Technical Notes provides an introduction to brick masonry cavity walls and discusses their
various properties. Information is provided for the proper design of cavity walls. Cavity wall use will
continue and recommendations found in this Technical Notes will improve building performance.
The information and suggestions contained in this Technical Notes are based on the available data
and the experience of the engineering staff of the Brick Industry Association. The information
contained herein must be used in conjunction with good technical judgment and a basic understanding
of the properties of brick masonry. Final decisions on the use of the information contained in this
Technical Notes are not within the purview of the Brick Industry Association and must rest with the
project architect, engineer and owner.
REFERENCES
An Inside Look at the Utility Brick Wall, Brick Association of the Carolinas, Charlotte, NC. 1.
Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402), The
Masonry Society, Boulder, CO, 1995.
2.
Catalog of Thermal Bridges in Commercial and Multi-Family Residential Construction, Oak
Ridge National Laboratory, Oak Ridge, TN, December 1989.
3.
Designing Low-Energy Buildings: Passive Solar Strategies & Energy 10 Software, Passive
Solar Industries Council, Washington, D.C., June 1996.
4.
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
14 of 15 9/13/2009 1:07 PM
Drysdale, R.G., Hamid, A.A., Baker, L.R., Masonry Structures: Behavior and Design,
Prentice Hall, Englewood Cliffs, NJ, 1994.
5.
Handbook of Fundamentals, American Society of Heating, Refrigerating and Air Conditioning
Engineers, Inc., Atlanta, GA, 1997.
6.
Masonry Designer's Guide, The Masonry Society, Boulder, CO, 1993. 7.
Thermal Mass Handbook, Concrete and Masonry Design Provisions Using ASHRAE/IES
90.1-1989, Eley Associates, San Francisco, CA, 1994.
8.
Uniform Building Code, International Council of Building Officials, Whittier, CA, 1997. 9.
Technical Notes 21 http://www.gobrick.com/BIA/technotes/t21.htm
15 of 15 9/13/2009 1:07 PM

Technical Notes 21A -BRICK MASONRY CAVITY WALLS -SELECTION OF MATERIALS
Feb. 1999
Abstract: The selection of quality materials is essential to the successful performance of brick masonry cavity walls.
Careful evaluation of materials is required in order to obtain the high performance level associated with these wall
systems. This Technical Notes addresses selection of appropriate materials, referencing ASTM standards when
applicable.
Key Words: air retarder, brick, cavity wall, expansion joint, flashing, insulation, materials, mortar, ties, vapor retarder,
weep holes.
INTRODUCTION
Brick masonry cavity walls have always been a popular choice with structural frames and as bearing walls, particularly
for use in commercial construction. These wall types are selected for their superior in-service performance resulting
from such properties as excellent moisture penetration resistance, thermal capabilities, sound transmission resistance
and fire resistance.
All of these qualities of brick masonry cavity walls depend on four key elements: design, material selection, detailing
and construction. Proper design will not compensate for inadequate material selection or detailing. Conversely,
superior design, material selection, or detailing will not compensate for poor construction practices. The use of quality
materials in the construction of brick masonry cavity walls is of prime importance. The selection of materials is even
more critical now that masonry design standards and model codes put greater emphasis on material properties for
design requirements. This Technical Notes addresses proper material selection. Properties of brick masonry cavity
walls, adequate detailing and construction are addressed in other Technical Notes in this series.
GENERAL
The standards of choice for quality construction materials are those developed by the American Society for Testing and
Materials (ASTM). ASTM has developed standard specifications for virtually all construction-related building materials.
These specifications are based on laboratory tests, field experience and, in the case of brick masonry units, are the
result of the legacy of masonry's long successful performance. ASTM standards are consensus documents that set
minimum acceptable requirements for materials of construction. ASTM standards allow for different performance levels
and it may be necessary to choose a level in the project specifications to meet the particular requirements for an
individual project. ASTM standards should be reviewed and understood before they are incorporated into a project
specification. It must also be understood that the use of ASTM standards does not guarantee the desired
performance, even though all materials may meet the specifications.
There are additional materials available that are not addressed in this Technical Notes that would accomplish the goal
of providing proper material selection. Lack of specific reference to a material does not preclude its use for brick
masonry cavity walls.
MASONRY UNITS
The exterior wythe of a brick masonry cavity wall may be of solid or hollow brick. The interior wythe can be solid brick,
hollow brick, structural clay tile or hollow or solid concrete masonry units.
Brick
Solid brick units should meet the requirements of ASTM C 216 Specification for Facing Brick when appearance is a
factor or C 62 Specification for Building Brick when appearance is not important. Hollow brick units should meet the
requirements of ASTM C 652 Specification for Hollow Brick. For these specifications, Grade SW should be specified
for all exterior exposures in areas that experience freeze/thaw cycling. Grade MW units may be used for the interior
wythe. Ceramic glazed solid brick used for both the exterior and interior wythe of cavity walls should meet the
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
1 of 13 9/13/2009 1:08 PM
requirements of ASTM C 1405 Specification for Glazed Brick (Single Fired, Solid Brick). Other ceramic glazed units
should conform to ASTM C 126 Specification for Ceramic Glazed Structural Clay Facing Tile, Facing Brick and Solid
Masonry Units. Information on classification and selection of brick can be found in Technical Notes 9A and 9B,
respectively. Further information on brick masonry material selection for adequate strength and compliance with the
Masonry Standards Joint Committee (MSJC) Code and Specification can be found in Technical Notes 3 Series.
Concrete Masonry Units
Concrete masonry units are usually used for the interior wythe of a cavity wall in combination with a brick masonry
exterior. They may also be used as accent bands in the exterior brick wythe. Hollow or solid concrete masonry units
should conform to ASTM C 55 Specification for Concrete Building Brick, ASTM C 90 Specification for Loadbearing
Concrete Masonry Units or ASTM C 129 Specification for Non-Loadbearing Concrete Masonry Units. Non load-bearing
concrete masonry units are usually specified for the interior wythe when the brick and block cavity wall is used as infill
walls for concrete or steel frame structural systems if shear loads are not transmitted to that wythe.
Structural Clay Tile
Structural clay tile has been used as a backing material for cavity wall construction. Structural clay tile for this purpose
should conform to ASTM C 126, ASTM C 34 Specification for Structural Clay Load-Bearing Wall Tile or ASTM C 212
for Structural Clay Facing Tile. Clay tile units under ASTM C 126 and ASTM C 212 are commonly used for the interior
wythe of a cavity wall when left exposed for architectural appearance reasons.
MORTAR
The strength and moisture penetration resistance of a brick masonry cavity wall are affected by the mortar selection
and compatibility with the brick units. Portland cement-lime, mortar cement, or masonry cement mortars can be used.
However, mortars with an air content less than 12 percent are recommended for their superior bond strength and
resistance to moisture penetration. The MSJC Code requires that the allowable flexural tensile stresses be reduced by
approximately 50 percent for assemblies constructed with masonry cement mortars or portland cement-lime mortars
that have air entrainment. In addition, some building codes prohibit the use of masonry cement mortars and all Type N
mortars in Seismic Performance Categories D and E (formerly Seismic Zones 3 and 4).
Mortar should meet the proportion requirements of ASTM C 270 Specification for Mortar for Unit Masonry as shown in
Table 1. Mortar type selection should be based on project requirements, such as strength and compatibility with a
particular brick unit. Types N or S mortars are typically used in brick masonry cavity walls. See Technical Notes 8
Series for more detailed information on mortar types and selection.
WALL TIES
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
2 of 13 9/13/2009 1:08 PM
Wall ties must provide two important functions: 1) distribute lateral loads between wythes; and 2) accommodate
differential movement between wythes. For a wall tie system to provide these functions, it must:
1)be securely attached and embedded in the masonry wythes;
2) be placed at the appropriate spacing;
3)have sufficient strength to transfer lateral loads with minimal deformations;
4)have a minimal amount of mechanical play, if an adjustable tie;
5)have sufficient corrosion resistance;
6)be easily installed without damage to the tie system or other wall components; and
7)not compromise expected wall performance.
Although cost of the wall tie should be considered, this should not be the controlling factor since the cost of the wall tie
system is small as compared to the total cost of the wall assembly.
There are many different wall tie types available for use in brick masonry cavity walls. These include unit ties,
horizontal joint reinforcement and adjustable ties.
Unit Ties
Fig.1

Unit Ties
Unit ties can be rectangular ties or "Z" ties as shown in Figure 1. These tie types are usually fabricated from
cold-drawn steel wire in accordance with ASTM A 82 Specification for Steel Wire, Plain, for Concrete Reinforcement.
They can also be fabricated from stainless steel conforming to ASTM A 167 Specification for Stainless and
Heat-Resisting Steel Plate, Sheet and Strip for use in more corrosive environments. Corrugated sheet metal ties are
not recommended for brick masonry cavity walls because they do not usually transfer loads properly between wythes.
Wall ties with drips should not be used since they reduce the capacity of the ties significantly.
Metal "Z" ties should only be used between wythes of solid masonry units in brick masonry cavity walls. Rectangular
ties can be used for all brick masonry cavity walls are therefore recommended instead. Wire ties should be either wire
size W1.7, [No. 9 gage, (0.148 in.) (3.8 mm)] or W2.8, [3/16 in. (4.8 mm)] in diameter.
Horizontal Joint Reinforcement
Continuous horizontal joint reinforcement should meet the requirements of ASTM A 951 Specification for Masonry Joint
Reinforcements. Horizontal joint reinforcement is typically produced in 10 to 12 ft (3 to 4 m) lengths. Longitudinal wires
are typically W1.7 [No. 9 gage, (0.148 in.) (3.8 mm)] or W2.8 [3/16 in. (4.8 mm)] diameter wire. Cross wires are W1.7
or W2.8 diameter wire and should be spaced at a maximum of 16 in. (400 mm) on center horizontally. Cross wires
without drips should be used. The total thickness of the wires should not exceed one-half the joint thickness. Horizontal
joint reinforcement configurations available are the ladder, truss and tab types as shown in Fig. 2.
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
3 of 13 9/13/2009 1:08 PM
Tests indicate that brick masonry cavity walls tied by the use of horizontal joint reinforcement perform similar to that of
the same wall systems tied together with metal unit ties. These tests also indicate that truss-type joint reinforcement
used in brick cavity wall construction helped to develop a degree of composite action in the horizontal span, but did not
contribute to any composite action in the vertical span. This restraint in the horizontal direction will reduce the amount
of in-plane movement and possibly result in bowing of the masonry walls. Thus, truss-type joint reinforcement is not
recommended for brick and block cavity walls.
Joint Reinforcement
Fig. 2
Adjustable Ties
Use of adjustable ties has increased for several reasons: 1) the structure can be enclosed faster since the exterior
brick wythe can be erected later in the construction process; 2) adjustable ties compensate for the height differences
between wythes and construction tolerances; and 3) they can accommodate larger differential movement between
wythes.
When specifying adjustable ties, there are certain conditions which must be considered. The model building codes limit
the vertical offset between the eye and pintle components to 11/4 in. (31.8 mm). Maximum play within the connecting
pieces is limited to 3/16 in. (1.6 mm). Once engaged, the pieces should not be able to separate. The strength and
stiffness of adjustable ties are generally less than that of unit ties or horizontal joint reinforcement. Thus, more ties are
required.
Adjustable Unit Ties. Such ties available for brick masonry cavity walls are shown in Fig. 3. These ties typically have
two-pieces consisting of a double eye and pintle configuration. Adjustable unit ties should be at least W2.8 [3/16 in.
(4.8 mm)] in diameter and meet the conditions previously stated.
Adjustable Joint Reinforcement Assemblies. Adjustable ties with ladder- and truss- type joint reinforcement are also
produced. This type of tie consists of rectangular tie extensions connected to standard joint reinforcement, see Fig. 4.
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
4 of 13 9/13/2009 1:08 PM
Adjustable Unit Ties
Fig.3
Corrosion Resistance
Corrosion potential can be affected by the function of the structure, geographic location, presence of insulation or vapor
retarders (alters the dew point within the wall), compatibility of construction materials, design and detailing of the wall
system and workmanship used in construction. Resistance to corrosion can be provided by control of environmental
factors or by selection of the tie material.
There are three types of materials used for corrosion protection of wall ties: galvanizing (zinc coatings), stainless steel
and epoxy coatings. Galvanizing provides resistance to corrosion in two ways. First, the zinc coating acts as a barrier
to corrosion by shielding the underlying metal. Second, the zinc coating acts as a sacrificial element that is consumed
by the corrosive environment before the base metal is attacked. Generally, as the thickness of the zinc coating
increases, the period of protection increases.
Two methods of galvanizing are available: mill galvanizing and hot-dip galvanizing. Mill galvanizing takes place after the
steel wire has been drawn, but prior to fabrication of the tie. Therefore, the cut end of the tie will not be protected.
Hot-dip galvanizing is performed by dipping the completed assembly into molten zinc until a specified amount of zinc is
bonded to the base metal. Hot-dip coatings are typically thicker than mill galvanizing, and therefore, provide longer
periods of protection. Minimum corrosion protection for galvanized ties should be according to ASTM A 153, Class B-2
(1.5 oz/ft2) (458 g/m2). Zinc coating requirements are related to the size of the element to be coated.
Stainless steel can be used for unit ties or joint reinforcement to resist corrosion. Stainless steel materials should
conform to ASTM A 167, Type 304.
Another method recently developed for corrosion protection of metal ties is epoxy coating. The application of an epoxy
coating is similar to that for reinforcing bars used in reinforced concrete or masonry construction. Epoxy coatings
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
5 of 13 9/13/2009 1:08 PM
provide protection by acting as an impervious barrier. These coatings are applied to the base steel by an electrostatic
spray method. The coating bonds to the steel by a heat-induced chemical reaction in which chemical and mechanical
bonds form. This aids in preventing cracking of the coating due to handling and installation. This coating type is not
sacrificial like zinc coatings. At present, an ASTM Standard governing this type of corrosion protection has not been
developed. Some manufacturers have been using an adaption of ASTM A 775 Specification for Epoxy-Coated
Reinforcing Steel Bars for metal tie purposes. At this time, hot-dip galvanizing is preferred over epoxy-coated ties
since any nicks, voids or cuts of the epoxy coating could lead to corrosion of the base steel.
THROUGH-WALL FLASHING
Flashing is a necessary component of the wall assembly to assure the excellent moisture penetration resistance of
brick masonry cavity walls. Its main purpose is to collect any moisture that penetrates the exterior wythe and divert it
to the outside of the wall system.
There are a variety of flashing materials that can be used with brick masonry cavity walls. Flashing for masonry
construction generally fall into three categories: sheet metals, composite materials (combination flashing) and fabrics
(plastic or rubber compounds). Materials such as polyethylene sheeting and asphalt-impregnated building felt should
not be used as flashing materials. These materials can be easily torn or punctured during installation.
Flashing must possess certain physical properties. Water resistance is the main attribute. However, flashing should
also be durable and resistant to damage during installation. Resistance to puncturing or tearing and resistance to
ultraviolet light must be evaluated in flashing selection. Flashing material should not be susceptible to corrosion in fresh
mortar or react with adjacent materials such as rigid insulation.
Flashing should be easily formed into the desired shape. Compatibility with materials such as sealants or adhesives
should be reviewed. Expected life of the flashing materials should be the expected life of the structure, at a minimum.
All of these qualities of the flashing material must be taken into consideration in selection because replacement is
expensive. Table 2 describes the thickness and advantages and disadvantages of flashing materials that can be
installed in brick masonry cavity walls.
Sheet Metals
Sheet metals used for flashing include stainless steel, copper, lead-coated copper and galvanized steel. Aluminum
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
6 of 13 9/13/2009 1:08 PM
should not be used since it can corrode in fresh mortar. Stainless steel and copper materials are the most durable, but
are also the most expensive. Galvanized steel has also been used in limited applications. These flashing materials
have a greater life expectancy than composite or fabric flashing. Sheet metal flashing is bent and formed on site and
sealed by soldering or with adhesives and rivets. This additional installation time can result in additional construction
costs. Stainless steel materials should conform the ASTM A 167 Specification for Stainless and Heat-Resisting Steel
Plate, Sheet and Strip. Copper flashing should comply with ASTM B 370 Specification for Copper Sheet and Strip for
Building Construction. Solder should conform to ASTM B 32 Specification for Solder Metal.
Composites
Composite or combination flashings are typically less expensive than sheet metal and are easier to install. The most
predominant type is a thin layer of metal sandwiched between one or two layers of another material. The metal layer is
usually of aluminum, copper or lead. It is covered on one or both sides with various materials, such as asphalt
coatings, kraft paper, fiberglass fabric or plastic films. Product literature should be reviewed to determine whether
these materials are appropriate for use in brick masonry cavity walls. Consideration should be given to compatibility
with adjacent sealants, delamination due to moisture penetration and movement of the wall system.
Plastic and Rubber Compounds
These flashing materials are usually the least expensive flashing suitable for brick masonry cavity walls. They are
flexible and can make complex shapes, except in the heavier thicknesses. Polyvinyl chloride (PVC), Ethylene Propylene
Diene Monomer (EPDM) and rubberized asphalt are the most common plastic flashings available.
PVC deteriorates and breaks down when exposed to ultraviolet (UV) light. The material also becomes brittle and
shrinks over time due to loss of plasticizers. Some PVC flashing is not compatible with polystyrene insulation and can
cause the insulation to degrade. However, not all PVC flashings have experienced such problems. Appropriate quality,
density and thickness of PVC flashings are paramount to successful performance and should be obtained from
well-recognized manufacturers.
EPDM was originally developed as a roofing membrane. However, this type of material has been gaining widespread
use as through-wall flashing for masonry walls. It has increased resistance to weathering and performs better in low
temperature environments than PVC. EPDM should be in a cured state. Junctures and laps of plastic flashing are
usually sealed with plastics or adhesives. These materials may require a metal drip edge adhered to the flashing when
exposed to exterior elements. EPDM should be talc free or the talc should be removed where laps are formed.
Another type of fabric flashing is a self-adhering, rubberized asphalt. This flashing material easily adheres to itself at
junctures, laps and interior wall surfaces and can be self healing to some extent. However, this material cannot be
placed on damp, dirty or dusty surfaces. The adhesion properties may be reduced during cold weather. Rubberized
asphalt can also degrade in the presence of UV light and therefore, requires the same type of drip edge as other
plastic flashing materials.
WEEP HOLES
Weep holes channel moisture collected on the flashing to the exterior of the wall assembly. For best performance,
weep holes should always be located directly on the flashing. If weep holes are installed one to two masonry courses
above the flashing, they will not perform their intended function. Some types of weep holes may aid in drying out the
wall system, although this is not their primary purpose.
Weep holes can be formed in a number of ways. Some of the most common are: 1) omitting mortar in all or part of the
head joint; 2) use of removable rods or ropes; 3) plastic or metal tubes; and 4) use of a wicking material. There are
also plastic and metal vents that cover weep holes used in lieu of mortar in vertical head joints. Open head joints are
recommended as weep holes; however, as long as weep holes remain open for drainage and positioned at the required
flashing locations with the appropriate size and spacing, the specific type of weep hole selected is not critical.
DRAINAGE MATERIALS
For brick masonry cavity walls to retard moisture infiltration, the air space separating the masonry wythes must be kept
clean of mortar droppings or mortar that may bridge the air space. These obstructions can render flashing and weep
holes ineffective.
While it is important to keep cavities clear, there are a number of approaches available to keep open a drainage path to
the weep holes. Drainage materials can be installed above flashing consisting of materials configured and installed to
allow water to flow around any mortar droppings. The use of these materials does not negate the importance of good
workmanship and may not be required in all situations. Use of drainage materials may result in the mortar bridging the
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
7 of 13 9/13/2009 1:08 PM
air space at certain locations, possibly above the flashing level. This may lead to isolated spots of dampness on the
interior or exterior wythe. At worst, it could lead to a path for water penetration. In all cases, the cavity size should be
maintained.
A layer of pea gravel with an approximate diameter of 3?8 in. (10 mm) can be placed on top of flashing installations
within the wall system. This layer should be approximately 2 to 3 in. (50 to 75 mm) deep. This will help to keep mortar
droppings from clogging weep holes. The smallest gravel size should be larger than the weep hole opening so as not
to interfere with drainage. It is recommended that a bed of mortar, conforming to the curve of the flashing, be placed
under the flashing for additional support of the pea gravel. Care must be exercised when installing pea gravel at bolted
shelf angle locations. The weight of the gravel on the flashing may cause tearing or puncturing at the bolt head. Pea
gravel at loose lintels needs to be contained so it does not flow off the end. Fig. 5 shows a typical detail for the use of
pea gravel although it would apply to many of the drainage materials.
Mesh materials are gaining popularity. These materials are usually manufactured from high density polyethylene or
nylon strand and are available in 1 in. to 2 in. (25 mm to 50 mm) thicknesses. They keep weep holes open and
permanently suspend the mortar above the flashing level. Many have unique patterns such as dovetail shapes which
break up mortar droppings so moisture has open flow paths to flashing and weep holes.
Another mortar dropping control device consists of staggered shelves inserted in the cavity such that they overlap each
other. The individual pieces are fixed in place by tabs which are mortared into the outer wythe of masonry. This may
collect mortar droppings above the flashing level and create a bridge for moisture to cross the cavity.
Some manufacturers of rigid board insulation provide a grooved face with an adhered filter fabric that is positioned in
the air space towards the exterior. The aligned grooves act as a channel which allow moisture to drain down to the
flashing and weep holes in the wall system. The grooves are intended to provide drainage capabilities in the event of
the air space being clogged during construction of the masonry wythes; however, the grooves must align vertically for
proper drainage. Other rigid insulation boards have an adhered mesh material which behaves in a similar manner.
EXPANSION JOINT MATERIALS
Building materials and elements used in construction react differently to loading and environmental conditions. Thus,
they are constantly moving. A system of movement joints is necessary to accommodate these expected movements.
For wythes of brick masonry, placement of expansion joints is necessary to permit these movements without concern.
Typical expansion joint details are shown in Fig. 6. A backer rod and sealant must always be installed to resist
moisture penetration at the expansion joint location. A filler material may be used in the expansion joint behind the
backer rod to keep debris out of the joint during construction. Common expansion joint filler materials are pre-molded
foam pads and neoprene pads. The sealant, backer rod and filler material, if present, must be designed to allow the
expected movement; therefore, the movement potential of these materials must be included when determining
expansion joint size and spacing. Expansion joint fillers should conform to ASTM D 1056 Specification for Flexible
Cellular Materials - Sponge or Expanded Rubber, Class 2A1.
Sealants come in many varieties and usually consist of polyurethanes, polysulfides or silicone. Sealants which exhibit
the highest movement capabilities should be specified. Sealants should conform to ASTM C 920 Specification for
Elastomeric Joint Sealants. For adequate sealing of expansion joints the sealant must bond well to the adjacent
brickwork, flashing, windows, etc. Sealant manufacturers should be consulted to determine whether priming is
necessary based on the sealant material selected. Oil-based caulking compounds should not be used as sealants for
masonry wall assemblies be cause most lack the necessary flexibility and durability needed for in-service use.
Backer rods are used to keep the sealant at the appropriate depth, provide a suitable shape for the sealant and act as
a bond breaker to prevent back-adhesion. The backer rod should be slightly larger than the joint. Closed cell
polyethylene rods are recommended and should be free from punctures.
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
8 of 13 9/13/2009 1:08 PM
STEEL SHELF ANGLES AND LINTELS
Several different types of steel members can support either the exterior or interior wythe of a brick masonry cavity
wall. They generally fall into two categories: shelf angles and lintels. Shelf angles are used in panel wall systems to
support the exterior wythe of brick masonry at floor levels. Shelf angles are anchored to floor slabs or beams and
break the exterior wall facade into sections. Loose angle lintels are generally used over wall openings to support the
masonry above. Typical loose angle lintels bear on the masonry at each end of the opening and are not attached to the
frame or backing.
Steel for shelf angles and lintels should conform to ASTM A 36 Specification for Carbon Structural Steel. Steel angles
should be at least 1?4 in. (6 mm) thick with a horizontal leg of at least 31?2 in. (89 mm) for use with nominal 4 in. (100
mm) brick wythes and 3 in. (75 mm) for use with nominal 3 in. (75 mm) thick brick wythes. Further information on the
design of steel lintels can be found in Technical Notes 31B Revised.
The use of galvanized steel shelf angles or lintels may be necessary in areas subject to severe corrosion. Steel lintels,
if not galvanized, should be painted before installation. These steel members support the weight of the masonry and
must remain in good condition. Repair or replacement of shelf angles or lintels is time consuming and costly so the
selection of adequate materials is critical to long service life.
INSULATION MATERIALS
In cavity wall construction, the prescribed air space acts as an insulating layer in addition to the masonry units.
Thermal performance of the system can be further enhanced by placing insulation materials in the cavity. Insulation
materials used in brick masonry cavity walls include inorganic cellular materials such as perlite and vermiculite, and
organic cellular materials such as polystyrene, polyurethane, polyisocyanurate and foams. These insulation types are
manufactured in the form of rigid boards, granular fills and foams. Each of these types, if properly used, will result in a
more thermally efficient wall system.
Although the most important characteristic for insulation is its thermal resistance, other properties should be considered
including water absorption, combustibility, density, insect resistance and ease of installation. The following criteria can
be used for the selection of insulation materials for brick masonry cavity walls:
1.The insulation must permit the air space to perform its function as a barrier to moisture penetration by allowing
moisture to drain without passage to the interior wythe.
2.Thermal insulating efficiency must not be impaired nor degrade over time due to retained moisture from any source,
i.e., wind-driven rain or vapor condensation.
3.Insulating materials must be long lasting, resisting rot due to moisture or dryness, offering no food value to vermin
and meeting the building code requirements for flame resistance.
4.Granular fill materials must be capable of supporting their own weight without settlement to assure that no portion of
the wall is without insulation and allow moisture to drain from the cavity.
5.Foam insulation materials must not shrink with age to assure that no portion of the wall is without insulation and that
moisture does not have a path to migrate to the interior wythe.
6.Rigid boards must be firmly attached to the backing so as not to become dislodged in the cavity and allow air or
water movement around the insulation.
7.Consider environmental concerns regarding off gassing and recycling of insulation materials.
Properties of insulation materials vary widely. Table 3 shows various properties of insulation materials used in brick
masonry cavity walls. The thermal conductivity (k) and thermal resistance (R) provide a means of comparing the
insulating properties of insulation materials. These are determined in accordance with ASTM C 177 Test Method for
Steady-State Heat Flux Measurements and Thermal Transmission Properties by Means of the Guarded-Hot-Plate
Apparatus measured at various temperatures. Due to the large number of types of insulation, and even larger number
of manufacturers, Table 3 lists only a few representative values of physical properties. For some materials, an aged or
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
9 of 13 9/13/2009 1:08 PM
stabilized value is given. In all cases, the aged or stabilized value is the one that should be used in design. Individual
manufacturers should be consulted for design values and other properties of their specific materials and their
applications.
Rigid Boards
There are many rigid board insulation materials that can be installed in the air space of brick masonry cavity walls.
Among the most common are: expanded and molded polystyrene, extruded polystyrene, expanded polyurethane,
polyisocyanurate, mineral fibers and perlite board.
Composition. Rigid board insulations are many and varied. They include the various mineral fiber boards and cellular
insulation including polystyrenes, polyurethanes and polyisocyanurates. Air, or other gases, introduced into the material
expands the material by as much as 40 times. Cells are formed in various patternsopen (interconnected) or closed
(unconnected). Most rigid insulation is expanded with hydrogenated chlorfluorocarbons (HCFC), pentane or other
hydrogenated gases used as blowing agents. Gradual air leakage into the cells may replace some of the original gas
and eventually reduce the thermal insulating quality. Some types of insulation use foil facers. These facers keep air
leakage to a minimum and must not be punctured during construction. Aged R-values should be used when comparing
different types of insulation.
Fibrous insulation materials include wood, cane, or vegetable fibers bonded with plastic binders. To make them
moisture resistant, they are sometimes impregnated with asphalt. Fibrous glass insulation consists of a core board of
non-absorbent fibers held together by phenolic binders and a surface coating of asphalt-saturated organic material
reinforced with glass-fiber.
Properties. Water entrapped in insulation can destroy its thermal insulating value. Water vapor can flow wherever air
can flowbetween fibers, through interconnected open cells, or where a closed cell structure breaks down. Wherever
water replaces air, the insulating value drops drastically since water's thermal conductivity exceeds that of air by 20
times.
Fibrous organic insulations are especially vulnerable to moisture damage. Free water will eventually damage any
fibrous organic material or organic binder. Fiberboard exposed to moisture for long periods of time may warp or buckle
and eventually decay. The expansion and contraction that accompanies the changing moisture content may lead to
problems.
Though less vulnerable to moisture, inorganic materials are not immune. Water penetrating into fiberglass insulation not
only impairs the insulating value, but may also dissolve the binder. Some types of cellular insulation may also break
down under repeated freeze/thaw cycles.
Rigid board insulation should conform to one of the following: ASTM C 208 Specification for Cellulosic Fiber Insulating
Board, ASTM C 552 Specification for Cellular Glass Thermal Insulation, ASTM C 578 Specification for Rigid, Cellular
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
10 of 13 9/13/2009 1:08 PM
Polystyrene Thermal Insulation, ASTM C 1224 Specification for Reflective Insulation for Building Applications, ASTM C
1289 Specification for Faced Rigid Cellular Polyisocyanurate Thermal Insulation Board, ASTM D 3490 Specification for
Flexible Cellular Materials - Bonded Urethane Foam and ASTM D 3770 Specification for Flexible Cellular Materials -
High Resilience Polyurethane Foam. There may be other products available for rigid foam products; therefore,
manufacturer's literature should be reviewed before selection.
Granular Fills
Granular fills are typically used in the hollow cells of concrete block backing. However, they can be used in the cavity.
One advantage of granular fills is that they can be installed after wall sections have been completed. This permits the
mason to work uninterrupted thus shortening construction time and lowering costs. However, the water resistance of
the fill is of utmost importance. Settling of the fill material could lead to thermal bridging in the wall system.
Two types of granular fill insulation have been found to meet the objectives of cavity wall insulation. These are water-
repellent vermiculite and silicone-treated perlite insulation.
Vermiculite is an inert, lightweight, granular insulating material manufactured by expanding an aluminum magnesium
silicate mineral, which is a form of mica. The raw material is made up of approximately one million separate layers per
inch, with a minute amount of water between each layer. When particles of the mineral are suddenly exposed to
temperatures in the range of 1800?F to 2000?F (982?C to 1093?C), the water changes to steam, causing the
vermiculite to expand into cellular granules of vermiculite insulation about 15 times their original size.
Perlite is a white, inert, lightweight, granular insulation material made from volcanic siliceous rock. When the crushed
stone is heated to approximately 1800?F (982?C), it expands or pops much like popcorn as the combined water
vaporizes and creates countless, tiny bubbles in the heat-softened, glassy particles. Perlite can be expanded up to 20
times its original volume.
Water-repellent vermiculite and silicone-treated perlite should conform to ASTM C 516 Specification for Vermiculite
Loose Fill Thermal Insulation and C 549 Specification for Perlite Loose Fill Insulation, respectively. Each of these
specifications contains limits on density, grading, thermal conductivity and water repellency. Some properties of these
materials is given in Table 3.
Foamed-in-place Insulation
Foamed-in-place insulations have traditionally been used to fill the cells of hollow masonry units. However, new
technology has now produced foamed-in-place insulations that can completely fill the air space between the two wythes
of a brick masonry cavity wall. One advantage of these foamed-in-place insulations is that they can be installed after
wall sections have been completed. However, the water resistance of brick cavity walls with these types of insulation
materials are unproven in the field. Shrinkage of the foam material could lead to moisture drainage paths through the
wall system. In addition, mortar bridging in the cavity may not allow the foam to flow around it properly. These
materials used in the cavity contradict the drainage system and should be considered a barrier wall system.
Most foams are a two component system. They are formed by a chemical reaction of two liquids. They are placed
under pressure, so newly constructed walls must be cured long enough to resist the pressures. Typical compositions
of these foams are urethane, polyicynene, magnesium oxychloride cement and ceramic talc and amino-plast resin.
Since most of these products are proprietary, there are no ASTM standards at this time to verify quality. Therefore,
manufacturers literature should be consulted for technical information on applications and usage.
AIR AND VAPOR RETARDERS
Sheets or layers of materials which effectively retard or reduce the flow of air and water vapor are called air retarders
and vapor retarders, respectively. In recent years, there has been much confusion pertaining to the functions of each
within a masonry wall system. Air retarders limit the amount of air flow through the wall system. Vapor retarders are
intended to control transmission of water vapor through building assemblies. A vapor retarder can also serve as an air
retarder. An air retarder may or may not serve as a vapor retarder. It is sometimes difficult to ensure that either
retarder performs only one function. For example, polyethylene films will function as a vapor retarder, but it will also
resist the passage of air.
To provide effective air and vapor retarders, it is necessary to seal joints in these materials so that continuity is
provided. It is also necessary to seal around the edges of wall openings such as windows, doors and access for utility
services. Adhering or taping of the joints should be specified. The adhesive or tape used must be compatible with the
material composition of the retarder in addition to performing adequately when exposed to moisture.
Air Retarders
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
11 of 13 9/13/2009 1:08 PM
Masonry cavity walls can appear to be relatively air tight, but may experience high air leakage rates. Parging the
exterior or interior face of the interior masonry wythe with mortar is one method to reduce air leakage, but it will crack if
the wythe cracks. Membranes or liquid-applied materials usually provide superior performance. Special attention must
be given to adjacent materials or structural members intersecting the interior wythe and the membrane. Permanent
fixtures in the interior wythe and movement joints at the top and sides of the wythe must provide a continuous air seal to
perform successfully.
For improved air retarder performance, the air and vapor retarders can be included in combination with mortar parging.
The parging provides a base for the application of the vapor retarder. The vapor retarder must be able to span
possible movement cracks in the interior wythe.
Interior or exterior insulation boards applied to the interior wythe can also form part of the air retarder system. For
insulation to work effectively as a retarder, proper adhesion to the interior wythe by the use of full grid of adhesive will
eliminate air spaces between the insulation and the interior wythe. Mechanical anchorage may provide a tight fit of the
insulation to the interior wythe. Joints between insulation boards must be sealed with a moisture resistant tape or
sealant.
Interior gypsum board finish materials can be used as air retarders when joints are properly sealed. Even when the
interior wall covering is not intended to serve as the main air retarder, it is suggested that steps be taken to provide
good air tightness in order to reduce possible air circulation in air spaces in the interior wythe.
Vapor Retarders
There are many materials that can be used to effectively provide vapor transmission resistance. However, the
installation methods vary just as much as the materials themselves. Selection should consider the material and ease of
application for the best results.
The vapor retarder selection depends on the type and location of insulation in the brick masonry cavity wall and the
interior and exterior climate. While some materials and methods of application may be successful in the cavity or on
the interior side of the interior wythe, there are certain vapor retarders which perform better on one side that the other.
Suitable vapor retarders consist of two coats of oil-based or alkyd emulsion paint on the interior side of the wall finish;
a 15-mil thick polyethylene sheet over the insulation; or the insulation itself if it is highly impermeable to water vapor
transmission and has taped or sealed joints. Foil-faced insulation and extruded polystyrene insulation boards meet
these criteria.
When the vapor retarder is installed within the cavity space, other material options are available. Paints or suitable
coatings can be applied to the outside face of the interior wythe, but must have the ability to span or bridge small
cracks. The vapor retarder can also consist of torched-on or spray-applied bituminous coatings, self-adhesive plastic
sheets or plastic materials.
The material must be continuous and be able to accommodate possible movement cracks that may form in the
masonry. For long term performance, vapor retarder materials must remain firmly affixed as vapor pressure
differentials occur. These materials must be chemically compatible with any insulation placed in the cavity.
If the insulation is used as the vapor retarder, it must be held firmly in place by adhesives or mechanical anchorage. In
addition, the joints between the insulation boards must be fully sealed with moisture resistant sealant or tape.
SUMMARY
This Technical Notes is the second in a series dealing with brick masonry cavity walls. It is concerned primarily with
recommended properties and selection of materials. Other Technical Notes in this series discuss cavity walls in
general, design, detailing and construction.
The information and suggestions contained in this Technical Notes are based on the available data and the experience
of the engineering staff of the Brick Industry Association. The information contained herein must be used in conjunction
with good technical judgment and a basic understanding of the properties of brick masonry. Final decisions on the use
of the information contained in this Technical Notes are not within the purview of the Brick Industry Association and
must rest with the project architect, engineer and owner.
REFERENCES
1."Brick Masonry Cavity Walls," Technical Notes 21 Revised, Brick Industry Association, Reston, VA, August 1998.
2.Handbook of Fundamentals, American Society of Heating, Refrigeration and Air Conditioning Engineers, Inc., Atlanta,
GA, 1997 Edition.
3."Thermal Insulation, Environmental Acoustics," Volume 04.06, Annual Book of ASTM Standards, American Society for
Testing and Materials, West Conshohocken, PA, November 1998.
4."Through-Wall Flashing," Engineering and Research Digest, Brick Industry Association, Reston, VA, 1996.
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
12 of 13 9/13/2009 1:08 PM
5."Wall Ties for Brick Masonry," Technical Notes 44B, Brick Industry Association, Reissued Sept. 1988.
6."Water Resistance of Brick Masonry, Design and Detailing - Part I of III," Technical Notes 7 Revised, Brick Industry
Association, Reston, VA, Reissued Feb. 1998.
7."Water Resistance of Brick Masonry, Materials - Part II of III," Technical Notes 7A Revised, Brick Industry
Association, Reston, VA, Reissued Dec. 1995.
Technical Note 21 http://www.gobrick.com/BIA/technotes/t21a.htm
13 of 13 9/13/2009 1:08 PM

Technical Notes 21B - Brick Masonry Cavity Walls - Detailing
April 2002
INTRODUCTION
Materials and workmanship alone are not sufficient to ensure adequate cavity wall performance. Unless properly
detailed, cavity walls constructed of the finest materials by the most talented masons will suffer the consequences of
poor detailing. This Technical Notes promotes quality cavity walls by discussing and depicting pertinent details.
This is the third in a series of Technical Notes devoted to brick masonry cavity walls. Other Technical Notes in this
series discuss cavity walls in general, including; properties, design, material selection, and construction. This
Technical Notes addresses proper detailing for brick cavity walls.
FLASHING AND WEEP HOLES
Through-wall flashing and weep holes are installed in exterior masonry wall construction to collect and divert
moisture that penetrates the exterior wythe of masonry to the outside of the wall. Through-wall flashing must be
provided at the base of the wall, at roof and chimney intersections, at roof and wall intersections, and at the top of
parapets. Flashing is also needed over and under door and window openings, at shelf angles, and at other horizontal
discontinuities in the cavity. Any penetrations in the flashing membrane should be sealed to prevent leakage.
Sealants and flashing used together must be compatible so that staining does not occur and long-term performance
is maintained.
For flashing and weep holes to perform as intended, the air space separating the masonry wythes must be kept
clear of mortar droppings and other obstructions that may bridge the air space. To achieve this end, primary
importance should be placed on good workmanship. A secondary method of keeping the cavity clean is through the
use of drainage materials specifically designed to stop any mortar droppings from blocking the cavity and allow
water to flow around them. Use of drainage materials is not required, and in some cases may contribute to water
penetration problems within the wall. More information on drainage materials is presented in Technical Notes 21A of
this series.
Drip Edge/Flashing Extension
Water that collects on flashing can re-enter the wall below if flashing terminates behind the face of the wall. For best
performance, flashing should be extended 14 in. (6 mm) beyond the wall plane and turned down at an angle of 45
degrees to form a drip. This forces water away from the wall surface. A protruding flashing is neither possible nor
desirable in some cases. For example, exposure to ultraviolet radiation may cause some flexible flashings to
deteriorate. In these cases, flashing should be cut flush with the face of the wall. In cases where the flashing itself
cannot be exposed, a non-corrosive metal drip edge may be used, see Fig. 1. Before specifying metal drips to be
used in conjunction with steel shelf angles or lintels, the potential for galvanic action between the metals should be
considered. Flashing materials should overlap the metal drip edge by a minimum of 1 in. (25 mm), and be fully
bonded to the top surface of the drip edge with a mastic or manufacturer-approved sealant. Metal drip edges should
be sealed at all laps and penetrations.
http://www.gobrick.com/BIA/technotes/t21b.htm
1 of 15 9/13/2009 1:09 PM
Flashing with Metal Drip Edge
FIG. 1
Spacing of Weep Holes
FIG. 2

Weep Holes
Structural Steel
Weep holes should be placed immediately above all flashing and be spaced no more than 24 in. (600 mm) on center
when open head joints are used, and no more than 16 in. (400 mm) on center when wick materials or round tubes
are used, see Fig. 2. Open head joints are preferred because they allow water to drain more quickly. Insects do not
commonly enter working weep holes, but they can be prevented from entering open head joints by using louvered,
vent-type weep inserts, stainless steel wool, or other drainage material. End Dams
Detailing and installation of end dams cannot be overemphasized. Their purpose is to ensure that collected water is
directed toward the weep holes where flashing is not continuous. Such areas occur above windows, doors and other
openings and under sills beneath windows. End dams are also used in conjunction with tray flashing at arches, when
flashing is stepped, and where a lower sloping roof line intersects a wall. Without end dams the collected water may
run off the ends of the flashing into the air space and saturate the brickwork below. To prevent this, each end of the
flashing should extend beyond the opening and turn up into the head joint a minimum of 1 in. (25 mm,) or a
prefabricated end dam may be used, see Fig. 3.
End Dams
Detailing and installation of end dams cannot be overemphasized. Their purpose is to ensure that collected water is
directed toward the weep holes where flashing is not continuous. Such areas occur above windows, doors and other
openings and under sills beneath windows. End dams are also used in conjunction with tray flashing at arches, when
flashing is stepped, and where a lower sloping roof line intersects a wall. Without end dams the collected water may
run off the ends of the flashing into the air space and saturate the brickwork below. To prevent this, each end of the
flashing should extend beyond the opening and turn up into the head joint a minimum of 1 in. (25 mm,) or a
prefabricated end dam may be used, see Fig. 3.
http://www.gobrick.com/BIA/technotes/t21b.htm
2 of 15 9/13/2009 1:09 PM
Foundation Detail
FIG. 4
Prefabricated Corners
(From Masonry Design and Detailing:
For Architects, Engineers, and Contractors,
4th Edition, Christine Beall)
FIG. 6
End Dam
FIG. 3
Foundations
To prevent moisture penetration and promote cavity drainage, place the bottom of the wall so that flashing is above
the finished grade. Care should to taken to ensure that flashing and weep holes are placed far enough above grade,
typically 8 in. (200 mm), so that they will not be covered by future grading or landscaping, see Fig. 4. With basement
construction, it is important to use through-wall flashing at the bottom of the cavity to prevent moisture from
penetrating to the basement wall, see Fig. 5. Below the flashing, any cavity should be filled solid with mortar or
grout. The flashing also prevents rising damp, ground water drawn up into the brickwork by capillary action. This
reduces the potential for staining and efflorescence. The flashing also serves as a bond break between the brick and
the concrete foundation. This permits differential movement between the materials and reduces the likelihood of
cracking. In construction without basements, the flashing may also serve as a termite shield.
http://www.gobrick.com/BIA/technotes/t21b.htm
3 of 15 9/13/2009 1:09 PM
Basement Detail
FIG. 5
Flashing at Steel Column
(From Masonry and Steel detailing Handbook, W.Laska)
FIG. 7

Corners
It is important to make sure flashing is continuous around corners. Forming corners with conventional flashing is a
complicated process involving folding and/or cutting, which increases the potential for flashing failure. Specifying
prefabricated corners eliminates the need to cut, patch and fold flashing, thereby reducing some of the potential for
water penetration, see Fig. 6. Whether field-formed or prefabricated, all corners should overlap at least 6 in. (150
mm), be sealed with mastic or an adhesive compatible with the flashing material, and conform to the shape of the
structure.

Columns
In some cases, vertical supports may make it necessary to cut, puncture or otherwise interrupt the flashing. When
http://www.gobrick.com/BIA/technotes/t21b.htm
4 of 15 9/13/2009 1:09 PM
this occurs, it is important to make sure that all openings in the flashing are tightly sealed, and that the flashing is
attached to these supports with mechanical means or approved adhesives. A common problem condition exists
when the inside wythe of a cavity wall spans between steel columns, and the column flanges are perpendicular to
the masonry. Fig. 7 illustrates one way that this problem can be addressed. The flashing is formed into a tray and
adhered to the column. Cut brick or concrete masonry units may be placed at the column base to provide support for
the flashing. Otherwise, the flashing can continue in front of the column if it is fully supported from behind.
Shelf Angle Detail
Fig. 8
Window Jamb Detail with Cavity Seal
Fig. 10
Shelf Angles
Shelf angles are used to support brick masonry, transfer the weight of brickwork to the structural frame and create
horizontal expansion joints. Flashing should be installed over all shelf angles, extended beyond the edge of the angle
when possible. Flashing may be self adhered to the shelf angle or have sealant at the edge to prevent wind driven
rain from penetrating underneath the flashing. The flashing material should extend back to the inside wythe and turn
up a minimum of 8 in. (200 mm). All shelf angles should have a horizontal expansion joint underneath, (see Fig. 8).
More information about expansion joints is found later in this Technical Notes.
DOORS AND WINDOWS
Detailing openings in masonry cavity walls requires special attention because any air or water which bridges the
cavity may cause problems. When the lintel used above any masonry opening is not continuous, flashing should
extend beyond the ends of the lintel and turn up to form end dams. Air leakage around window and door frames to
the cavity can be minimized by placing a pre-compressed pad or sealant in the cavity at the perimeter of the
opening. Such a seal is placed at the head of the opening between the back of the lintel and interior masonry wythe,
see Fig. 9, and below the window sill. The seal must be integrated with a similar seal at the jamb between the two
http://www.gobrick.com/BIA/technotes/t21b.htm
5 of 15 9/13/2009 1:09 PM
masonry wythes, see Fig. 10. A sealant joint at the exterior window/masonry interface is the primary defense
against the ingress of both air and water. Jamb flashing is not required, but may be placed between the interior
masonry wythe and the window frame as an additional barrier. In the case of masonry jambs, flashing between the
two masonry wythes prevents contact and transmission of water to the interior, see Fig. 11. Jamb flashing should be
fully adhered to the interior masonry wythe and extend down to lap over the through-wall flashing at the sill. Flashing
is also needed below window sills, see Figs. 12 and 13. When one-piece continuous masonry sills are specified, the
flashing should extend into the exterior masonry wythe at the jambs, see Fig. 10. When detailed in this manner, sill
flashing can manage water that bypasses sealant joints at both the sill and jambs. If not so detailed, the jamb
flashing must be placed to direct water to the sill flashing. Sealing the pre-compressed cavity seal at the jambs to
the sill flashing completes a continuous barrier between the window frame and the cavity. End dams must be formed
at the ends of the sill flashing. Self-flashing windows only handle water which makes its way inside the window frame
and do not negate the need for the flashing installation described above.
Window Head Detail and Cavity Seal
FIG. 9
Window Jamb Detail
FIG. 11

Sealant Joints
Sealant joints are the primary defense against moisture penetration through joints in exterior elements. Sealant joints
at masonry openings in exterior walls, such as door and window frames, and expansion joints should be designed,
detailed and installed with the same care as other building components, not applied as an afterthought. Too
frequently, sealants are used to correct or hide poor workmanship, rather than being included as an integral part of
the construction.
Sealants perform best when they are bonded to two opposing surfaces. When a sealant is bonded to three
surfaces, its ability to accommodate movement is significantly reduced. Backer rods are recommended and provide
support for the sealant within the joint and prevent three-sided adhesion. A bond-breaking tape may be required with
some types of backer rods. The sealant should be installed in accordance with the manufacturers
recommendations. Additionally, maintenance programs should be implemented to inspect and replace sealants that
http://www.gobrick.com/BIA/technotes/t21b.htm
6 of 15 9/13/2009 1:09 PM
may have dried out, split, or separated from the substrate.

PARAPETS
Of all the masonry elements used in buildings, probably the most difficult to properly detail is a parapet wall.
Designers have tried many different ways to minimize cracking, leaking, and displacement. Generally, the only
guarantee against parapet problems is to eliminate the parapet. However, they are frequently required by building
codes or included for aesthetic reasons.
For cavity wall construction, it is recommended that the cavity continue to the top of the parapet, thereby permitting
differential movement between the outer and inner wythes. Expansion joints should extend to the top of the parapet
as well. In addition, the inner wythe of the parapet may need to be reinforced and attached to the structural frame.
Additional vertical expansion joints should be placed in the parapet, located between those in the wall below, and
near corners to avoid displacement of the parapet.
Copings on parapets should provide a drip on at least one side of the wall and slope towards the drip. Metal, stone,
and fired clay copings of various designs usually provide this feature. Place through-wall flashing in the mortar joint
immediately beneath the coping and firmly attach the coping to the wall below with mechanical anchors. Sealant
should be applied where the anchors penetrate the flashing. More information about caps and copings can be found
in Technical Notes 36A. Parapets should not be painted or coated, they must be able to breathe. Roofing
membrane should not extend up the back side of the parapet without consideration of moisture vapor transmission.
Window Sill Detail
FIG. 12
http://www.gobrick.com/BIA/technotes/t21b.htm
7 of 15 9/13/2009 1:09 PM
Window Sill Detail
FIG. 13

MOVEMENT JOINTS
Vertical Expansion Joints
The exterior of each building must be analyzed to determine the potential for horizontal movement, and provisions
must be made to relieve the stress that results from such movement. Typical locations for vertical expansion joints
include at intervals in long walls, interior corners, returns, and the jambs of large openings. See Technical Notes 18
and 18A for a thorough discussion of movements and detailing to address its effects, formulas for spacing of vertical
expansion joints, and specific applications. Details of typical expansion joints and their locations are shown in Figs.
14 and 15. Vertical expansion joints are usually 38 to 12 (9 mm to13 mm) wide to match typical mortar joints
and filled with a backer rod and sealant. Toothed expansion joints are difficult to construct and do not perform as
well as straight expansion joints. For aesthetic reasons, consideration may be given to hiding vertical expansion joints
in locations such as interior corners. It may be desirable to accentuate vertical expansion joints, making them
attractive design elements, see Fig 16. Joint reinforcement should not span vertical expansion joints.
Vertical Expansion Joint
FIG. 12
http://www.gobrick.com/BIA/technotes/t21b.htm
8 of 15 9/13/2009 1:09 PM
Vertical Expansion Joint at Interior Corner
FIG. 15

Horizontal Expansion Joints
Cavity walls are successfully used as curtain walls in concrete and steel-frame buildings. In such instances, the inner
wythe is usually supported by the building frame at each floor level. The outer wythe, often supported by shelf
angles attached to the frame, is tied by metal ties to the inner wythe and the building frame. Care should be taken to
ensure that the shelf angle is properly anchored and shimmed to prevent rotations and deflections that may induce
high concentrated stresses in the masonry. Shelf angles should be designed so that total deflections are less than
116 in. (1.6 mm). Shelf angles should be segmented to permit thermal expansion and contraction of the steel
without damage to the wall. Spaces between segments of the shelf angle do not have to align with brick expansion
joints. Continuous flashing should be installed above all shelf angles. A horizontal expansion joint is formed by
providing a minimum space of 18 (4 mm) below the shelf angle. This space is filled with a compressible material
and sealed with an elastic sealant, typically of a color that closely matches the mortar joint, see Fig. 8. In some
instances, the width of the expansion joint below the shelf angle may be larger than is desirable. An alternate detail
shown in Fig. 17 allows room for movement while providing a narrower joint. Horizontal expansion joints may be
eliminated by having the weight of the exterior wythe supported on the foundation. One alternative has both wythes
of the cavity wall bypass the structural frame, with the weight supported on the foundation. The inner wythe is
anchored to the structure with flexible anchors for lateral support. One method of such anchoring is shown in Fig. 18.
Alternately, only the interior wythe can be supported on the structural frame. If this second support condition is used,
it is imperative that adjustable ties be used between the wythes and that all components that penetrate or are
attached to both wythes be detailed to permit differential movement.
accentuated Vertical Expansion Joint
FIG. 16
http://www.gobrick.com/BIA/technotes/t21b.htm
9 of 15 9/13/2009 1:09 PM
Lipped Brick Detail
FIG. 17
ANCHORAGE AND TIES
Masonry walls used to enclose frame structures must be carefully designed and detailed to permit the transfer of
loads applied to the wythes to the frame in a manner that accommodates relative movement. Frame structures are
more flexible than masonry cavity walls and can undergo greater deflections. The frame and enclosing cavity wall
also differ in their exposure and reaction to temperature and moisture conditions. Anchors that provide lateral
support to a wall should resist loads perpendicular to the plane of the wall, but should not transfer in-plane loads.
This allows the vertical and horizontal movement of the brick in the plane of the wall while restricting its out-of-plane
movement, and permits differential movement between the frame and the wall without cracking or distressing the
masonry. Figures 18 through 20 show typical methods for anchoring masonry walls to columns and beams with
metal anchors. Anchors must be detailed to accommodate construction tolerance differences and movement
resulting from loads applied to the frame and floor elements. Anchors should not be located in the same bed joint as
flashing. A more complete discussion and examples of ties appropriate for connecting the wythes in cavity wall
construction are given in Technical Notes 21A. As noted there, frequency and spacing of cavity wall ties depends on
wire size and tie type. Table 1 contains such prescriptive requirements from Building Code Requirements for
Masonry Structures, ACI 530/ASCE 5/TMS 402, [3] also known as the Masonry Standards Joint Committee (MSJC)
Code. In addition, ties must be located within 12 in. (305 mm) of openings larger than 16 in. (406 mm) in either
direction. These ties are spaced at a maximum of 3 ft (0.91 m) on center around the perimeter. Ties should also be
located within 12 in. (305 mm) of free edges of each wythe and at the horizontal or vertical spacing indicated.
Anchorage to Steel Column
(From Masonry and Steel Detailing Handbook, W,Laska) (Fireproofing omitted for clarity)
FIG. 18
TABLE 1
Wall Tie Area and Spacing Requirements
Tie Wire Size
in.(mm)
Wall Area per Tie
Number per ft
2
(m
2
)
Maximum Tie Spacing
Horizontal by Vertical,
in.(mm)
W1.7 (MW11)
0.125 (3.06)
2
2/3
(0.25)
36 BY 24
(914 BY 610)
http://www.gobrick.com/BIA/technotes/t21b.htm
10 of 15 9/13/2009 1:09 PM
W2.8 (MW18)
0.188 (4.76)
4
1/2
(0.42)
36 BY 24
(914 BY 610)
Adjustable, with 2
W2.8 (MW18)
1.88 (4.76) legs
1.77 (0.16)
16 BY 16
(406 BY 406)
Anchorage to Steel Column
(From Masonry and Steel Detailing Handbook, W,Laska) (Fireproofing omitted for clarity)
FIG. 19
Steel Joist Bearing
From Masonry and Steel Detailing Handbook,
W.Laska)
FIG. 21
BEARING
Structural Steel
The coefficient of thermal expansion of steel is approximately twice that of brick masonry. When the temperature
difference between the materials is large, and the steel is firmly anchored to or confined within the masonry, the
potential for cracking the masonry wall increases. Masonry bond beams are usually placed in the interior wythe
below the steel joists to distribute the load on the bearing wall. The common practice is to mechanically attach floor
and roof joists to steel anchors embedded in the masonry. This detail can be improved by lubricating the bearing
surfaces, or providing bearing pads with low coefficients of friction, and providing slotted holes in joist ends.
Surrounding the joist end with a layer of building paper prior to grouting the bearing wall creates a slip joint,
permitting movement. Anchor bolts should be hand-tightened or friction will prevent the necessary movement.
Minimum bearing requirements are established by the Steel Joist Institute and depend on the span and depth of the
joist. Figure 21 illustrates a structural system using steel joists bearing on a masonry wall.
Concrete Planks
http://www.gobrick.com/BIA/technotes/t21b.htm
11 of 15 9/13/2009 1:09 PM
Precast hollow-core concrete planks generally bear on the interior wythe of a cavity wall. The plank rests a minimum
of 3 in. (75 mm) on a bearing pad that separates the plank from a concrete masonry bond beam below. Anchorage
to the wall may be achieved with reinforcing steel (see Fig. 22) or, when lateral loads are small, planks may be
solidly grouted to the wall. In either case, connections should be designed by a structural engineer.
Wood Floor Joists
Wood floor joists normally have a 3 in. (75 mm) fire cut end and bear only on the interior wythe of a cavity wall. The
ends of the joist must not project into the cavity; they can form a ledge, which may create a moisture bridge across
the cavity. Wood floor joists may also be fastened to metal joist hangers attached to a ledger bolted to the inner
wythe of the cavity wall, or joist hangers may be embedded in the inner wythe of the cavity wall. See Fig. 23.
Wood Floor Joist Connection Detail
FIG. 23
Anchorage to Steel Column
FIG. 20
Steel reinforced Concrete Plank Bearing
FIG. 22
Wood Roof Framing
Wood roof framing can be anchored to cavity walls by many methods, one of these is shown in Fig. 24. The detail
illustrates a method in which the bearing plate is secured by anchor bolts grouted into the top of the cavity. The roof
framing can then be attached to the bearing plate with strap anchors as shown. Anchor bolts holding roof plates
http://www.gobrick.com/BIA/technotes/t21b.htm
12 of 15 9/13/2009 1:09 PM
should extend into the masonry a minimum of four bolt diameters or 2 in. (50 mm). Resistance to uplift of roof
members may require longer bolt embedment or vertical reinforcement. After the wood plate is installed, the nut
should be hand-tightened. Occasionally, a wind driven rain may cause a difference in pressure sufficient to drive
water up over the top course of the exterior masonry wythe and into the cavity. This can be prevented by adding a
frieze board that extends a few inches down from the top of the wall and is sealed against the masonry at the
bottom.
Bond Breaks
Foundations
Foundation movement may cause cracking in masonry walls rigidly attached to the foundation. Walls not bonded to
the foundation tend to span the low points and thus reduce the potential for cracking. Figure 25 illustrates a typical
foundation detail. In this case, the bond is broken between the base of the cavity wall and the top of the concrete
foundation by building paper. The transfer of movement in the foundation to the wall is thus minimized. In many
instances the brick wythe is separated from the foundation by flashing. Bond breaks also permit differential thermal
and moisture movements without distress to either the brickwork or the concrete foundation.
In locations with high winds or seismic activity, it is necessary to anchor at least one wythe of the cavity wall to the
foundation. This is typically achieved by bond between the mortar or grout, but may require the use of shear keys on
reinforcing bars from the foundation into hollow masonry units or into a grouted cavity.
Wood Roof Framing Detail
FIG. 24
Foundation Detail
FIG. 25
Concrete Slabs
Thermal strains or other movements are often blamed for cracking in masonry walls that is caused by the shrinkage
or curling of concrete slabs that bear on the walls and are bonded to them. Curling of a concrete slab is caused by
deflection of the slab when the forms are removed and response to dead and live loads. Unfortunately, this behavior
of concrete is frequently overlooked by the designer in detailing the structure. Figure 26 illustrates a typical detail
that will relieve this condition. Installation of a bond break between the concrete slab and the concrete masonry
permits the slab to have some freedom of movement with respect to the plane of the wall. The slab is thickened into
a beam over the interior wythe to help stiffen the slab and minimize curling. The weight of the wall above the plate
must be greater than the uplift force. Slab curling may also be reduced by placing diagonal reinforcement in the slab
corners.
http://www.gobrick.com/BIA/technotes/t21b.htm
13 of 15 9/13/2009 1:09 PM
Concrete Roof Slab Detail
FIG. 26
SEISMIC DETAILING
Masonry walls in buildings in locations with seismic activity must be designed to resist the lateral loads imposed by
seismic events. The requirements are determined by the Seismic Design Category (SDC) into which the structure
fits, as outlined in Minimum Design Loads for Buildings and Other Structures, ASCE 7 [13].
Seismic requirements for masonry walls are found in Section 1.13 of the 2002 MSJC Code, as well as in model
building codes. Prescriptive amounts of horizontal and vertical reinforcement are required, based on Seismic Design
Category and if the wall is part of the lateral force-resisting system. However, the requirements will not be discussed
here as the focus of this Technical Notes is detailing. Such reinforcement must be placed with the restrictions for
size, cover, and tolerances that are included in the building code.
SUMMARY
This Technical Notes has discussed and illustrated the general principles that are involved in the proper detailing of
brick masonry cavity walls. It is not possible to cover all of potential conditions and variations in a single Technical
Notes. However, the intent is to address the general principles and considerations for detailing. The information and
suggestions contained in this Technical Notes are based on the available data and the experience of the engineering
staff of the Brick Industry Association. The information contained herein must be used in conjunction with good
technical judgment and a basic understanding of the properties of brick masonry. Final decisions on the use of the
information contained in this Technical Notes are not within the purview of the Brick Industry Association and must
rest with the project architect, engineer and owner.
REFERENCES
1. Wallace, M.A., Anatomy of a Cavity Wall, Magazine of Masonry Construction, Vol. 3, No. 7, July 1990, pp.
296-304.
2. Hoke, J.R., Editor, Architectural Graphics Standards, Ninth Edition, John Wiley & Sons Inc., New York, NY, 1994.
3. Building Code Requirements for Masonry Structures and Commentary (ACI 530/ASCE 5/TMS 402-02) and
Specification for Masonry Structures and Commentary (ACI 530.1/ASCE 6/TMS 602-02), American Concrete
Institute, Structural Engineering Institute of the American Society of Civil Engineers, and The Masonry Society,
2002.
4. Laska, W. and Ostrander, C, Cavity Walls: A Case of High Performance, Masonry Advisory Council, Park
Ridge, IL.
5. Krogstad, N., Weber, R., and Johnson, D., Common Problems at the Interface Between Masonry Drainage
Walls and Windows, American Society for Testing and Materials STP 1246, 1996, pp. 29-43.
6. Laska, W., Connecting Precast Planks to Concrete Masonry Walls, Magazine of Masonry Construction, Vol. 4,
No. 7, July 1991, pp. 250-251. 7. Hooker, K., Corners in a Flash, Magazine of Masonry Construction, Vol. 4
No. 12, December 1991, p. 466.
8. Krogstad, N., Kozoil, R., and Weber, R., Detailing Critical Interfaces Between Masonry Walls and Roofing
Systems, Seventh Canadian Masonry Symposium, 1995, pp. 43-62.
9. Laska, W., Detailing Shelf Angles, Magazine of Masonry Construction, Vol. 3, No. 7, January 1989, pp. 16-17.
10. Laska, W., Masonry and Steel Detailing Handbook, The Aberdeen Group, 1993
http://www.gobrick.com/BIA/technotes/t21b.htm
14 of 15 9/13/2009 1:09 PM
11. Masonry: The Cavity Wall Controversy, Proceedings of the British Masonry Society, No. 5, May 1993.
12. Beall, C., Masonry Design and Detailing: for Architects, Engineers and Contractors, Fourth Edition, McGraw-Hill,
1997.
13. Minimum Design Loads for Buildings and Other Structures (ASCE 7), Structural Engineering Institute of the
American Society of Civil Engineers, Reston, VA, 2002.
14. Standard for Hurricane Resistant Residential Construction, (SSTD 10), Southern Building Code Congress
International, March 1999.
15. Subasic, C., Seismic Reinforcement for Masonry, Magazine of Masonry Construction, Vol. 13, No. 4, April
2000, pp. 20-26.
16. Zinter, K., Technical Guide-Selection and Use of Sealants, Bostik Inc., 1999.
17. Uniform Building Code, International Council of Building Officials, Council of Building Officials, Whittier, CA, 1997.
18. Catani, M.J., Where Do You Need Joints?, Magazine of Masonry Construction, Vol. 1, No. 10, October 1988.
http://www.gobrick.com/BIA/technotes/t21b.htm
15 of 15 9/13/2009 1:09 PM

Technical Notes 21C - Brick Masonry Cavity Walls - Construction
October 1989
Abstract: This Technical Notes describes proper techniques which should be used during the construction of brick
masonry cavity walls. These techniques cover: storing materials, completely filling all mortar joints, placing wall ties,
flashing and weep holes, keeping the cavity clean and protecting the wall from weather during construction.
Key Words: brick, cavity walls, flashing, joints, mortar, ties, weep holes, workmanship.
INTRODUCTION
This fourth in the series of Technical Notes devoted to brick masonry cavity walls covers good construction
practices. Other Technical Notes in this series are concerned with cavity walls in general, how to insulate them, and
how to properly detail them.
The proper construction of a brick masonry cavity wall is as important to proper performance as are the design, the
use of quality materials, and proper detailing. Proper construction may not be achieved if it is considered of
secondary importance by the designer. Adequate supervision may be necessary to ensure proper construction.
GENERAL
In the construction of a cavity wall there are no changes required in basic bricklaying techniques, only modifications
of practices commonly used in the construction of any brick masonry wall. The fundamental principle in a cavity wall
is that there shall be no bridge of solid material capable of carrying water across the minimum 2-in. (50 mm) cavity
space. Therefore, the construction of two separate wythes, with a clean cavity, is of prime importance. This
Technical Notes will discuss certain construction practices which are necessary for brick masonry cavity walls to
perform successfully.
WORKMANSHIP
The importance of the workmanship used in constructing masonry has been stressed by many, sometimes to the
point that it may appear that workmanship alone is responsible for the performance of masonry walls, regardless of
the wall design, detailing, or the materials used. While this is by no means true, good workmanship is a very
important factor in the construction of high performance masonry. See Technical Notes 7 Series for more
information on moisture resistance of masonry walls.
Complete Filling of All Mortar Joints
Extensive laboratory tests at the National Institute of Standards and Technology, formerly the National Bureau of
Standards, and elsewhere, as well as hundreds of observations of masonry buildings, indicate that to obtain good
masonry performance, there is no substitute for the complete filling of all mortar joints that are intended to receive
mortar. Partially filled mortar joints result in leaky walls, reduce the strength of masonry, and may contribute to
spalling due to freezing and thawing in the presence of excessive moisture. Therefore, all joints intended to receive
mortar in both the exterior and interior wythes should be completely filled as the brick are laid.
Keeping the Cavity Clean
It is vital that the cavity be kept clean of mortar droppings and other foreign materials. If mortar falls into the cavity,
it may form "bridges" for moisture passage, or it may fall to the flashing, blocking the weep holes.
Over the years many methods have been developed and considerable time and discussion have been devoted to the
proper method to use in keeping the cavity clean. One method is to take a wooden or metal strip, slightly smaller
than the cavity width, and place it in the air space. This strip rests on the wall ties as the wall is built. Wire or rope is
http://www.gobrick.com/BIA/technotes/t21c.htm
1 of 7 9/13/2009 1:10 PM
attached to the strip. Then, as the brickmason builds the wall, this strip is easily lifted out. Before the next row of ties
is placed, any mortar which may have fallen into the cavity is removed (Figure 1).
Keeping the Cavity Clean
FIG. 1
Another method is to place every third brick or so in the course above the flashing of the exterior wythe dry and
wedge it into proper position so that it can be removed for final cleaning of the cavity. Mortar droppings at the base
of the cavity can be easily removed and weep holes provided when the brick are mortared in the wall.
In addition to the above mentioned methods of cleaning the cavity, the brickmason can use techniques that, if
properly applied, should eliminate a considerable amount of mortar falling into the cavity in the first place (Figures 2
through 7).
1. After spreading the mortar bed, the brickmason should bevel the cavity edge with the flat of the trowel (Figure 2).
When mortar is spread in this manner, very little will be squeezed out of the bed joints into the cavity when the units
are laid (Figure 3).
2. The brick units are next rolled into place, keeping most of the mortar on the outside (Figures 4, 5 and 6).
3. After the brickmason has placed the unit on the bed joint any mortar fins protruding into the cavity should be
flattened over the backs of the unit, not cut off (Figure 7). This prevents the mortar from falling into the cavity and
provides a smooth surface which will not interfere with insulation materials which may be placed in the cavity.
http://www.gobrick.com/BIA/technotes/t21c.htm
2 of 7 9/13/2009 1:10 PM
FIG. 2
FIG. 3
FIG. 4
FIG. 5
http://www.gobrick.com/BIA/technotes/t21c.htm
3 of 7 9/13/2009 1:10 PM
FIG. 6
FIG. 7

Tooling
Weather tightness and textural effect are the basic considerations of mortar joint finish selection and execution.
Properly "striking" or "tooling" the joint helps the mortar and brick units bond together and seal the wall against
moisture. Nine common joint finishes are shown in Figure 8 in order of their decreasing weather tightness.
Compression of the mortar makes the concave, V, and grapevine joints the most weather tight and acceptable to
use. The remaining six joint types are not recommended for exterior use. All holes in the mortar joints should be
filled. Joints should be tooled when the mortar is "thumbprint" hard.
http://www.gobrick.com/BIA/technotes/t21c.htm
4 of 7 9/13/2009 1:10 PM
Typical Mortar Joints
FIG. 8
Weep holes
Weep holes must be placed at the base of the cavity and at all other flashing levels. They provide a means of
draining away any moisture that may have found its way into the cavity. Weep holes must provide a clear access to
the cavity and must be placed directly on the flashing for proper drainage.
Weep holes can be easily created or installed by various methods. In order of effectiveness these are:
1. Eliminating each second or third head joint.
2. Inserting oiled rods, rope or pins in the head joint at a maximum of 16 in. (410 mm) o.c. and removing before final
set of the mortar.
3. Placing metal or plastic tubing in the head joint at a maximum of 16 in. (410 mm) o.c.
4. Placing sash cord or other suitable wicking material in the head joint at a maximum of 16 in. (410 mm) o.c.
Insulation Placement
The installation of insulation in a cavity wall is covered in Technical Notes 21A Revised. In addition, the literature of
the insulation manufacturer should be consulted before beginning construction.
Tie Placement
In a properly constructed cavity wall, both wythes of masonry must be adequately and properly tied together. The
main concern to the designer is assurance that all of the ties are in place and remain operative, firmly embedded in
and bonded to the mortar. To achieve this, the two wythes of the cavity must be laid with completely filled bed joints,
and the ties must be in the correct position so that later disturbance of the wall assembly is unnecessary. There
exists extensive data showing that wall ties have excellent tying capacity if they are well embedded in the masonry,
and also that the cavity wall is a weak structural member if few working ties must do the job of many (Figure 9).
Cavity wall metal ties are embedded in bed joints as units are laid.
FIG. 9
There must be at least one 3/16-in. (4.76 mm) diameter steel wall tie in every 4 1/2 sq ft (0.4 m
2
) or #9 gauge wire
for each 2 2/3 sq ft (0.25 m
2
) for a cavity wall whose cavity is no greater than 4 in. (100 mm). If the cavity width is
greater than 4 in. (100 mm), a wall tie analysis should be performed. The most common type of wall tie for brick
masonry construction is the Z tie. In addition, there are the rectangular and U-shaped ties which are to be used
when the backup units are hollow masonry units with cells laid vertically.
From a performance standpoint, the most important factors for wall ties are:
http://www.gobrick.com/BIA/technotes/t21c.htm
5 of 7 9/13/2009 1:10 PM
1. Being corrosion resistant.
2. Placing ties at proper spacing. Spacing of ties should be reduced by one half for ties with drips. Crimping of the
metal ties to form a drip is not necessary, and will decrease the strength of the tie.
3. Full bedding of the bed joint and placing the wall tie in the mortar 5/8 in. (16 mm) from either edge of the brick.
Horizontal Joint Reinforcement. Prefabricated horizontal joint reinforcement may be used to tie the interior and
exterior wythes. Truss type joint reinforcement should never be used to tie the wythes of a brick and block cavity
wall together. Instead, ladder type reinforcement which allows for the in-plane movement between the wythes is
recommended.
Horizontal joint reinforcement is not usually required in brick masonry walls since they are not subject to shrinkage
stresses. The use of horizontal joint reinforcement makes the placing of the cavity wall ties more convenient and
there is less concern over omitting them.
PROTECTION
Storage of Materials
The manner in which materials are stored at the construction site may have an influence on their future performance.
Materials should be stored to avoid wetting by rain or snow, and also avoid contamination by salts or other matter
which may contribute to efflorescence and staining.
Masonry Units. Masonry units should be stored off the ground to avoid contamination by dirt and by ground water
which may contain soluble salts. They should also be covered by a water-resistant membrane to keep them dry.
Cementitious Materials. Cementitious materials for mortar should be stored off the ground and under cover.
Sand. Sand for mortar should also be stored on high ground, or ideally, off the ground to prevent contamination from
dirt, organic materials and ground water, any of which may contribute to efflorescence and may be deleterious to
mortar performance. In addition, it is advisable to store sand and other aggregates under a protective cover. This
will avoid saturation and freezing in cold weather.
Flashing. Flashing materials should be stored in places where they will not be punctured or damaged. Plastic and
asphalt coated flashing materials should not be stored in areas exposed to sunlight. Ultraviolet rays from the sun
break down these materials, causing them to become brittle with time. Plastic flashing exposed to the weather at the
site for months before installation should not be used. During installation, flashing must be pliable so that no cracks
occur at corners or bends.
Protection of Walls
Rain. Masonry walls exposed to weather and unprotected during construction can become so saturated with water
that they may require weeks, or even months (depending upon climatic conditions), to dry out. This prolonged
saturation may cause many of the slightly soluble salts to go into solution, thus raising the possibility of
efflorescence. Such conditions may also contribute to the contamination of the masonry with soluble salts from
elsewhere in the construction (concrete, concrete block, plaster, trim, etc.).
During construction, all walls should be kept dry by covering the top of the wall with a strong, water-resistant
membrane at the end of each day or shutdown period. The covering should overhang the wall by at least 24 in. (610
mm) on each side, and should be secured against wind. The covering should remain in place until the top of the
cavity wall is completed or protected by adjacent materials.
Freezing. Leaky walls can sometimes be attributed to the freezing of mortar before it has set, or the lack of
protection of materials and walls during cold weather construction. Therefore, when building in cold weather, all
materials and walls should be properly protected against freezing. This involves the following items: storing of
materials, preparation of mortar, heating of masonry units, laying precautions, and protection of work. Technical
Notes 1 Series, "Cold Weather Masonry Construction," contains recommendations for construction and protection of
masonry during freezing weather. ACI-ASCE 530.1 Specifications for Masonry Structures also has requirements for
cold weather construction.
SUMMARY
http://www.gobrick.com/BIA/technotes/t21c.htm
6 of 7 9/13/2009 1:10 PM
This Technical Notes provides the basic information required for good construction of brick masonry cavity walls.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the technical staff of the Brick Institute of America. The information and recommendations contained in
this publication must be used in conjunction with good engineering judgment and a basic understanding of the
properties of brick masonry and related construction materials. Final decisions on the use of the materials and
recommendations contained in this publication are not within the purview of the Brick Institute of America and must
rest with the project architect, engineer, owner or all.
http://www.gobrick.com/BIA/technotes/t21c.htm
7 of 7 9/13/2009 1:10 PM
Stains - Identification and Prevention
Abstract: This Technical Note provides descriptions and photographs that aid in identifying efflorescence and stains on brick-
work. It includes information on stain composition, factors that influence their occurrence, and stain prevention.
Key Words: carbonates, chlorides, efflorescence, manganese, rain water, silicates, stains, soluble salts, sulfates, vanadium.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
23
June
2006
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Identification:
Use photos and descriptions for preliminary identification
of efflorescence or stains
When uncertain of correct identification of efflorescence or
stain, have experienced brick personnel or professionals
verify prior to cleaning
Prevention:
Do not clean brickwork with unbuffered hydrochloric (muri-
atic) acid
Use cleaning agent or procedure recommended by brick
manufacturer to prevent cleaning-related stains
Store brick off ground and cover with non-staining water-
proof material
Protect top of unfinished brickwork from weather
Page 1 of 5
INTRODUCTION
Brick has been used to create beautiful buildings for centuries. Most of these structures have a substantial history
of outstanding performance. In some instances the appearance is affected by the development of efflorescence or
stains. These may originate from materials in the brick or mortar, from adjacent materials, and from outside sourc-
es such as cleaning agents. Each has a particular chemical composition and a unique means of removal.
Identification of the origin of the efflorescence, stain or foreign material is the first step in returning brickwork to its
proper appearance. Some stains are often misidentified or are mistaken for efflorescence. Since correctly identify-
ing efflorescence or a stain can be difficult, it is recommended that experienced brick personnel or professionals
verify the efflorescence or type of stain. Misidentification may result in application of an inappropriate correction
method. When correctly identified, efflorescence and stains can generally be removed. Inappropriate correction
methods may result in further staining or damage of the brickwork.
Further information on the formation and prevention of efflorescence is discussed in Technical Note 23A. Once
final identification of efflorescence or a stain is made, refer to Technical Note 20 for removal recommendations.
Efflorescence
Efflorescence is not considered a stain but will be
discussed here for identification purposes. Refer to
Technical Note 23A for causes and prevention.
Efflorescence is normally a harmless deposit of water-
soluble, white salt crystals, as shown in Photo 1. In
some instances efflorescence may appear on mortar
joints as shown in Photo 2. Efflorescence may appear
in limited areas on the surface of brickwork as shown in
Photo 3 or, in extreme cases, cover the entire brickwork
surface.
Efflorescence is usually white in color; however, all white
stains on brick masonry are not necessarily efflores-
cence. Lime run and white scum, as discussed below,
are occasionally mistaken for efflorescence.
Photo 1
Efflorescence
Lime Run (Carbonate Deposits)
Calcium carbonate deposits, sometimes referred to
as lime run, usually appear as white or gray, crusty
formations originating from a spot and running down
the face of the wall, as shown in Photo 4. However,
the term lime run is misleading since the stain is not
a direct result of the lime component in the mortar. In
fact, hydrated lime actually helps to reduce the risk
of lime run. Lime run nearly always occurs at a small
hole or opening or hairline crack in the face of the brick
masonry, as shown in Photo 5.
The source of the calcium compounds that contribute
to lime run can be trim, mortar, backing, or other con-
struction materials. Lime run requires large quantities
of water that follow the same path over an extended
period of time, similar to the formation of stalactites in
limestone caves. The water takes any of several cal-
cium compounds into solution and brings them to the
surface of the masonry through an opening. At the sur-
face, the solution reacts with carbon dioxide in the air,
thus forming the crusty calcium carbonate deposit.
Materials containing cement are sources of calcium
compounds and are an integral component of or may
be in contact with the brickwork. To reduce the pos-
sibility of lime run, excess water must be eliminated or
the path must be disrupted. Once lime run begins, it
is likely to continue until difficult to the water source is
stopped.
White Scum (Silicate Deposits)
Silicate deposits, sometimes referred to as white
scum, usually appear as white or gray discolorations
on the face of brick masonry, as shown in Photo 6. The
discoloration may be present over the entire face of the
brickwork or in smaller, irregularly shaped areas. White
Photo 2
Efflorescence on Mortar Joints
Photo 3
Efflorescence in Limited Areas
Photo 4
Lime Run
www.gobrick.com | Brick Industry Association | TN 23 | Stains - Identification and Prevention | Page 2 of 5
Photo 5
Lime Run
www.gobrick.com | Brick Industry Association | TN 23 | Stains - Identification and Prevention | Page 3 of 5
scum may also occur adjacent to trim elements,
precast concrete and, occasionally, large expanses
of glass.
A number of mechanisms may precipitate white
scum on brickwork. White scum is typically related
to the cleaning of brickwork with unbuffered hydro-
chloric (muriatic) acid solutions or inadequate pre-
wetting or rinsing of the brickwork during cleaning.
Silicate deposits on brick masonry should not be
confused with scumming that sometimes occurs on
brick during the manufacturing process. This type
of scumming will be evident on brick before they
are placed in the wall.
Vanadium (Green or Yellow)
Stain
Some brick develop yellow or green salt depos-
its as shown in Photos 7 (yellow) and 8 (green)
when they come in contact with water or unbuf-
fered hydrochloric (muriatic) acid. These stains
are usually vanadium salts. They may be found
on red, buff or white brick; however, they are more
conspicuous on lighter-colored brick. The vana-
dium salts responsible for these stains originate
in the raw materials used for the manufacture of
the brick. The yellow and green stains are usually
vanadyl salts, consisting of sulfates and chlorides,
or hydrates of these salts.
Vanadium stains occur in a manner similar to efflo-
rescence, except that vanadium oxide and sulfates
are dissolved and result in a solution that may be
quite acidic. As water evaporates from this solution
at the surface of the brickwork, vanadyl salts are
deposited. The chloride salts of vanadium, such
as vanadyl chloride, may form as a result of wash-
ing with unbuffered hydrochloric (muriatic) acid or
excessive moisture exposure. [Ref. 2]
Preventing vanadium stains is important, since
they can be difficult to remove and improper clean-
ing efforts may result in a brown, insoluble deposit.
To minimize the potential for vanadium stains, the
following steps are recommended:
- Store brick off the ground and under non-
staining protective covers.
- Never use or permit the use of highly-con-
centrated, unbuffered hydrochloric (muri-
atic) acid solutions to clean light-colored
brick.
- Seek and follow the cleaning recommen-
dations of the brick manufacturer.
Photo 6
White Scum
Photo 7
Yellow Vanadium Stain
Photo 8
Green Vanadium Stain
www.gobrick.com | Brick Industry Association | TN 23 | Stains - Identification and Prevention | Page 4 of 5
Manganese (Brown) Stain
Under certain conditions, tan, brown, or occasionally
gray staining may occur on the mortar joints of brick-
work as shown in Photo 9. Occasionally, a brown stain
will streak down onto the faces of the brick as shown
in Photo 10. This type of stain is the result of having a
manganese oxide as a coloring agent in tan, brown,
black or gray brick and its reaction to an acid.
During the brick firing process, the manganese coloring
agents undergo several chemical changes, resulting
in manganese compounds that are insoluble in water.
They have varying degrees of solubility in weak acids.
Once dissolved, these compounds may migrate in
solution toward the surface of brickwork. As previously
discussed, acid solutions can occur in brickwork under
certain conditions. Brick may also absorb unbuffered
hydrochloric (muriatic) acid during cleaning. It is also
possible that some geographical areas may be subject
to acid rain. [Ref. 4]
Manganese staining is closely related to efflorescence,
since it is the sulfate and chloride salts of manganese
that travel to the surface of the brickwork. When the
solution reaches the mortar joints, the salts are neu-
tralized by the cement or lime in the mortar produc-
ing insoluble manganese hydroxide. The manganese
hydroxide precipitate is deposited on the mortar joint
and when dry converts to brown manganese tetroxide
resulting in the stain. [Ref. 2]
Unbuffered hydrochloric (muriatic) acid should not
be used to clean tan, brown, black or gray brick.
Proprietary cleaning compounds are available for
cleaning brick containing manganese. Test for effective-
ness and follow the advice of the brick manufacturer.
Stains from External Sources
Other stains affecting brickwork are generally caused
by external sources such as pollution, organic growth,
runoff or others. Usually, the source or composition
of these stains is obvious. Organic stains, as shown
in Photo 11, can include algae, mold, or other organ-
isms. Certain materials above or adjacent to brickwork
such as copper (Photo 12), bronze, aluminum, syn-
thetic stucco or paint (Photo 13) can stain brickwork
surfaces. In addition, externally caused stains such
as hard water from sprinkler systems can effect brick-
work (Photo 14). The color and appearance should
be considered during identification. Laboratory or field
tests can determine the stain composition and assist in
proper identification. Once the correct identification is
made, the appropriate method to clean the brickwork
can be implemented.
Rust-colored stains may actually be corrosion, as
shown in Photo 15. Such stains can be the result of
Photo 9
Manganese Stain
Photo 10
Manganese Stain
Photo 11
Organic Stain
www.gobrick.com | Brick Industry Association | TN 23 | Stains - Identification and Prevention | Page 5 of 5
corrosion of wall ties or joint reinforcement in or
adjacent to the brickwork. The use of improper mor-
tar additives or ingredients, placement of wall ties
or joint reinforcement with inadequate cover, welding
splatter on the brick, or the corrosion of a material
placed on the brick cube or pile prior to being laid in
the brickwork can all attribute to these stains.
SUMMARY
The proper identification of efflorescence and
stains on brickwork is essential to stain removal.
Photographs and laboratory or field testing can
assist in this effort. When uncertain of the composi-
tion or origin of efflorescence or stains, verification
by trained, experienced brick personnel or profes-
sionals is recommended. The use of an incorrect
cleaning agent or method on a stain could result in
further staining that is more difficult to remove than
the original stain. Understanding the mechanisms
involved with the formation of efflorescence and
stains on brickwork is useful in design and construc-
tion to minimize their occurrence.
The information and suggestions contained in this
Technical Note are based on the available data and
the combined experience of engineering staff and
members of the Brick Industry Association. The
information contained herein must be used in con-
junction with good technical judgment and a basic
understanding of the properties of brick masonry.
Final decisions on the use of the information con-
tained in this Technical Note are not within the pur-
view of the Brick Industry Association and must rest
with the project architect, engineer and owner.
REFERENCES
1. Brownell, W.F., Kenna, J .L., and Wilko, J r., P. P.
Staining of Mortar by Manganese Colored Brick,
Bulletin, American Ceramic Society, Vol. 45, No.
12, 1966.
2. Chin, I.R., and Petry, L., Design and Testing
to Reduce Efflorescence Potential in New Brick
Masonry Walls Masonry: Design and Construction,
Problems and Repair, ASTM STP 1180, J .M.
Melander and L.R. Lauersdorf, Eds., American
Society for Testing and Materials, Philadelphia, PA,
1993.
3. Minnick,T.J ., Effect of Lime on Characteristics of
Mortar in Masonry Construction, Bulletin, American
Ceramic Society, Vol. 38, No. 5, 1959.
4. Grimm, C.T., Cleaning Masonry A Review
of the Literature, Construction Research Center,
University of Texas at Arlington, Arlington, TX, 1988.
Photo 12
Stain from Copper
Photo 14
Hard Water Stain
Photo 13
Runoff Stain from Paint
Photo 15
Rust Stain
Efflorescence - Causes and Prevention
Abstract: This Technical Note describes the mechanisms leading to the formation of efflorescence, including probable
sources of soluble salts and moisture. Conditions necessary to cause efflorescence to appear are presented, along with design
recommendations and practices that reduce the potential for efflorescence.
Key Words: admixtures, chlorides, condensation, efflorescence, masonry, mortar, rain water, soluble salts, sulfates, trim.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
23A
June
2006
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
General
Positively identify efflorescence (see Technical Note 23)
Determine if the efflorescence is "new building bloom"
Identify causes of efflorescence using the included trouble-
shooting list and correct the causes before attempting
removal
Design and construct brickwork to maximize water pen-
etration resistance
Consider materials that contain fewer soluble salts
Coatings are not recommended as the sole treatment to
control efflorescence
Design
Isolate exterior brick wythe with an air space
Waterproof the exterior of walls that extend below grade
For walls without an air space, separate the brick wythe
from the backing with a dampproof coating
Construction Practices
Store masonry materials off the ground and cover with
waterproof materials to protect them from groundwater
and precipitation
Protect unfinished masonry from weather during construc-
tion
Page 1 of 9
INTRODUCTION
Brickwork provides aesthetically pleasing structures. Occasionally, a white crystalline deposit will appear on the
surface. These deposits may be efflorescence, water-soluble salts that occasionally occur on the surface of
masonry. Although undesirable, efflorescence is usually not harmful to brick masonry.
This Technical Note describes the mechanisms of efflorescence, including possible sources of salts and water,
providing a basic understanding of the phenomenon of efflorescence that is essential to minimize its occurrence.
Although similar, sources of salts and the development of efflorescence on paving surfaces are discussed in the
Technical Notes 14 Series on brick pavements.
A thorough program of efflorescence prevention or control may include identification of the type of efflorescence
and investigation of its causes before options for correction are considered. This Technical Note is intended to
serve as a guide in this process. Procedures for removal of efflorescence are discussed in Technical Note 20.
CAUSES OF EFFLORESCENCE
There are many, often complicated, mechanisms of efflorescence. Simply stated, efflorescence occurs when water
containing dissolved salts is brought to the surface of masonry, the water evaporates and the salts are left on the
surface of the masonry. The salt solutions may migrate across surfaces of masonry units, between the mortar and
units, or through the pores of the mortar or units.
There are certain simultaneous conditions that must exist in order for efflorescence to occur:
- Soluble salts must be present within or in contact with the brickwork. These salts may be present in brick,
backing materials, mortar ingredients, trim, adjacent soil, etc.
- There must be a source of water in contact with the salts for a period of time sufficient to dissolve them.
- The masonry must have a pore structure that allows the migration of salt solutions to the surface or other
locations where evaporation of water can occur.
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 2 of 9
Efflorescence on brickwork less than one year old can often be attributed to new building bloom. When efflores-
cence occurs more than a year after construction is complete, it can generally be attributed to excessive water
penetration or poor drainage and is often associated with or is most severe in winter.
Under certain specific circumstances and conditions, it is possible for the crystals of efflorescence to form within
the bodies of brick. When this occurs, it is possible that the growth of crystals and the resulting pressure may
cause cracking and distress to masonry.
It is not practical to attempt to preclude all soluble salts from masonry materials or possible to prevent moisture
from coming in contact with masonry exposed to weather. However, reducing each contributing factor is a realistic
approach that will usually prevent the occurrence or reduce the severity of efflorescence.
The most effective means of preventing efflorescence is to minimize the amount of water that penetrates brick-
work. This, as well as separating brickwork from sources of salts, is primarily accomplished through careful design
and construction. To the extent possible, the amount of salts available is reduced through material selection.
MATERIAL SELECTION
Material selection provides the first opportunity to reduce efflorescence potential, however the ability to eliminate
these salts is somewhat limited and less practical than minimizing water penetration. The following recommenda-
tions are presented to assist the designer in the selection of materials to limit the occurrence of efflorescence.
The chemical composition of efflorescing salts is usually alkali and alkaline earth sulfates and carbonates,
although chlorides may also be present. The most common salts found in efflorescence are sulfate and carbonate
compounds of sodium, potassium, calcium, magnesium and aluminum. When chlorides occur as efflorescence, it
is usually a result of the use of calcium chloride as a mortar accelerator, contamination of masonry units or mortar
sand by salt water, or the improper use of hydrochloric acids in cleaning solutions.
Since efflorescence appears on the face of the brickwork, it is often erroneously assumed to come from the brick.
Usually however, soluble salts originate in other masonry units and materials that make up the assembly.
Brick
It is possible for brick to contain soluble salts, however brick that do not contribute to efflorescence are readily
available throughout the United States and Canada. Soluble salts contained within a brick may be dissolved by
any water the brick absorbs, contributing to the development of efflorescence.
The potential for brick to effloresce can be assessed using the efflorescence test in ASTM C 67, Standard Test
Methods for Sampling and Testing Brick and Structural Clay Tile [Ref. 1]. This test consists of partially immersing
brick in distilled water for a period of 7 days. At the end of this period, the brick are allowed to dry, examined for
efflorescence and compared to brick that were not immersed. The individual performing the test, based on visual
observation, rates the brick as either not effloresced or effloresced.
The ASTM C 67 efflorescence test is more severe than normal environmental exposures. It is important to under-
stand that this is a product test that only establishes whether the brick contain enough soluble material to efflo-
resce under laboratory conditions. Brick rated effloresced" may not exhibit efflorescence in service; brick rated
"not effloresced" may exhibit efflorescence in service. Migration of salts from other sources may also cause brick
to effloresce. Construction practices that reduce the potential for contamination of brick are discussed later.
Mortar and Grout
Mortar can be a significant contributor to efflorescence. W.E. Brownell, the author of a research report on efflores-
cence in brickwork [Ref. 2] states:
The primary and most obvious source of contamination of otherwise efflorescence-free brick is
the mortar used in wall construction. The mortar is in intimate contact with the brick on at least
four and sometimes five sides. It is applied to the brick in a wet, paste-like condition which pro-
vides ample moisture for the transfer of soluble salts from the mortar to the brick. If any appre-
ciable soluble material is present in the mortar, it will be carried into the brick proportionately to
the amount of moisture transferred.
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 3 of 9
This would also apply to grout, which is made with the same constituent materials as mortar, and contains a great-
er amount of water. Brownell continues:
The simplest case of soluble salt contamination of efflorescence-free brick is the migration of
free-alkali solutions from the mortar to the brick. This situation is not only the simplest mecha-
nism, but it is also the most common. In the trade, it is known as new building bloom.
New building bloom is typically fairly uniform across the wall surface. New-building bloom, tends to occur shortly
after construction ends, due to normal water loss during post-construction drying. Efflorescence that reappears
after this initial period is likely due to water entering and remaining in the wall.
Mortar types and proportions should be selected on the basis of structural and exposure requirements for the par-
ticular project. Recommendations for mortar are contained in Technical Notes 8 and 8B.
Cement. The principal contributor to efflorescence in mortar and grout is the alkali content of portland cement.
The tendency of cement to effloresce may be predicted with reasonable accuracy from a chemical analysis of the
cement. Cements high in alkalies are more prone to produce efflorescence than cements of lower alkali content.
All cements contain some water-soluble alkalies. Those common in mortar and grout are sodium and potassium.
It is suspected that the sulfate content of the cement may be as significant as the alkali content in contributing to
efflorescence.
ASTM C 150, Standard Specification for Portland Cement [Ref. 1], includes provisions for specifying low-alkali
cement, when required. In these cases, the optional compositional requirements limit equivalent alkalis (Na
2
O +
0.658K
2
O) to a maximum of 0.60 percent as measured using ASTM C 114, Standard Test Methods for Chemical
Analysis of Hydraulic Cement [Ref. 1]. The alkalies referred to are the total of acid-soluble and water-soluble frac-
tion alkalies. In general, the water-soluble alkali content will be of the order of 60 percent of the total. Brownell
[Ref. 2] states:
Experience has shown that 0.1 percent free alkali in a portland cement used in common mortars
will cause new building bloom; therefore, if such efflorescence is to be avoided, the free alkali of
the cement should be less than this and should be specified as low as possible.
This severe limitation on water-soluble alkali content is only met by a few cements other than portland blast-fur-
nace slag cement and masonry cements made with slag cement. It should be stated, however, that all investiga-
tors do not agree. Many believe that Brownell is extremely conservative.
Other ingredients for mortar and grout, i.e., lime, sand and water, should also be selected with care, although their
contribution to efflorescence is typically less than that of cement.
Lime. Various investigators disagree as to the possible contribution of lime to efflorescence. It has been demon-
strated that lime, clay or sand additions to a mortar do not generally contribute to efflorescence [Ref. 6]. In fact,
these ingredients tend to dilute the deleterious effects of high alkali cement.
On the other hand, lime is somewhat water-soluble. Its presence may serve to neutralize sulfuric acids generated
within the masonry. Under certain conditions, lime reacts with unbuffered hydrochloric acid to produce very soluble
calcium chloride, which can migrate to the surface. Nevertheless, lime improves the extent of bond between mor-
tar and brick, and thereby increases the water resistance of the masonry.
Sand. Sand used in mortar and grout is not water-soluble; however, it may be contaminated with material that will
contribute to efflorescence. Sand for use in mortar or grout should be taken from sources free of contamination
from salt water, soil runoff, plant life and decomposed organic compounds, among others. Using clean, washed
sand will eliminate any efflorescing contribution.
Admixtures. A wide variety of admixtures for masonry mortar and grout is available. Most of these products are
proprietary and their compositions are not disclosed. In general, they are classified as workability enhancers, bond
enhancers, water repellents, set retarders or set accelerators.
Compressive strength, water retention and air content are generally the only properties included in reports on
mortar. However, an admixtures effect on extent of bond and bond strength may also influence the efflorescence
potential, as a reduction in either may make masonry walls more vulnerable to water penetration. Therefore,
admixtures with unknown compositions are not recommended for use in mortars unless it has been established by
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 4 of 9
experience or laboratory tests that they will neither materially impair mortar bond nor contribute to efflorescence.
Chlorides that occur as efflorescence may result from the use of calcium chloride and compounds containing cal-
cium chloride added to mortar as set accelerators. Because calcium chloride causes corrosion of metal embedded
in mortar under certain conditions, Specification for Masonry Structures, ACI 530.1/ASCE 6/TMS 602 [Ref. 5], lim-
its chloride ions to 0.2 percent for mortar admixtures. Normally, this amount of calcium chloride will not contribute
materially to efflorescence.
Backing
Materials commonly used as backings or inner wythes of masonry walls, such as concrete and concrete masonry
units may contain large quantities of soluble salts. As with mortar, the cement in concrete and concrete products is
the principal contributor to efflorescence. Measurements of the soluble salt contents of clay and concrete masonry
units and their efflorescing tendencies have shown that concrete products can contain two to seven times as
much soluble material as the fired clay material [Ref. 3]. Such salts found in backing materials may contribute to
efflorescence on the face of the brickwork if sufficient water is present to dissolve the salts and there is direct con-
tact between the brick and backing or a pathway for the solution to reach the masonry surface.
As with mortars, concrete and concrete masonry units made with low alkali cements decrease the quantity of salts
available for efflorescence. Design suggestions and construction practices that prevent absorption of these salts
by brick are presented later.
Trim
Building trim, such as caps, coping, sills, lintels, keystones, etc., are often built into the brick wythe. These items
may be natural stone, cast stone, precast concrete, etc. any of which may contain soluble salts. Such materials
may contribute significantly to efflorescence on the face of adjacent brickwork. As the amount of soluble salts con-
tained in these materials cannot usually be controlled, their effects on brickwork should be carefully considered.
The contribution of these materials to efflorescence can be minimized as discussed in Design and Detailing.
DESIGN AND DETAILING
The most meticulous design and detailing may be thwarted by the selection of inappropriate materials or by poor
workmanship. The converse is also true: the use of the best possible materials and craftsmanship will not in them-
selves ensure a successful and permanent structure if the design is improper.
As previously discussed, the mechanism of efflorescence is dependent upon the presence of free water in the
masonry to dissolve the available soluble salts. The primary source of moisture for the occurrence of efflores-
cence is rain water that penetrates or comes in contact with masonry. Rain water will penetrate all masonry walls
to some degree, more so if they are improperly designed or detailed.
Limiting avialable moisture helps to suppress the devel-
opment of efflorescence. Therefore, much depends on
the design and attention to certain critical details, particu-
larly those that are associated with preventing moisture
entry into the brickwork and directing water away from
wall tops and horizontal surfaces, see Photo 1.
A general discussion of water penetration resistance is
provided in the Technical Notes 7 Series. The follow-
ing paragraphs present design measures that minimize
the potential for efflorescence by improving the water
resistance of brickwork. The craftsmanship employed in
the construction of a masonry wall also has a significant
effect on the amount of water penetrating the wall and is
discussed later.
Air Space
Drainage walls are recommended for maximum resis-
tance to rain penetration and minimum efflorescence.
Photo 1
Efflorescence from Water Penetrating Wall Top
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 5 of 9
The drainage wall system provides a cavity or air space
that separates the exterior brickwork from other elements
in a wall assembly, as shown in Figure 1. This separation
impedes the development of efflorescence in two ways. It
allows water to drain down the back of the brick wythe and
out of the wall without wetting other materials and prevents
the migration of salts from backing materials by isolating
the brick wythe.
Photo 2 illustrates the transfer of soluble salts from backing
units to brick when the two materials are in direct contact.
This result was obtained by placing the concrete masonry
units in pans of water with five brick on top of each block.
The brick had previously been subjected to the efflores-
cence test and rated "not effloresced." The potential for
backing materials to effloresce can be determined by using
the same efflorescence test method used for brick.
The barrier wall strategy of filling a collar joint between
masonry wythes with mortar or grout to provide a continu-
ous barrier to water penetration is not recommended.
Because mortar and grout contain cement, collar joints
increase the potential for efflorescence by placing a source
of salts in direct contact with the brickwork.
Providing a dampproof coating on the exterior surface of
concrete and concrete masonry backings of drainage and
barrier walls may reduce the potential for brick veneer
to effloresce. These coatings limit the passage of salts
and moisture between backing materials and the brick
wythe. Dampproof coating materials for concrete masonry
walls are typically bituminous and applied by spraying or
brushing.
Trim
Because trim materials are often in contact with brickwork,
any salts they contain may contribute to efflorescence
as stated earlier in Material Selection. These materials
are frequently used in locations that are most vulnerable
to water penetration, such as caps, copings and sills.
Therefore, all three of the conditions necessary for efflores-
cence are met almost anywhere these elements are used.
However, placing flashing or other materials that can act
as a capillary break between trim materials and brickwork
can reduce the contribution of trim materials to efflores-
cence by preventing contact and the migration of salts
from trim.
Below Grade Walls
Soluble salts in soil are dissolved by water that percolates
through the ground. Consequently, most ground water
contains a high concentration of these salts. When
masonry is in contact with the earth, ground water may
be absorbed by the masonry and may rise through capil-
lary action several feet above the ground. An accumula-
tion of salts in the masonry is then possible, as shown in
Photo 3.
Air Space
Water Exits
at Flashing
Through Weep
Flashing
Water Runs
Down Back
of Brick
Water Enters
Brick Wythe
Figure 1
Drainage Wall System
Photo 2
Migration of Soluble Salts from Backing to Brick
Photo 3
Migration of Soluble Salts from Soil
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 6 of 9
To eliminate these salts as sources of efflorescence, waterproof the masonry that extends below grade and install
flashing a few courses above grade. The waterproofing prevents the masonry from absorbing dissolved salts from
the adjacent soil and the flashing acts as a capillary break, preventing any absorbed salts from rising higher in the
wall.
Condensation
In addition to rain water and ground water, water may accumulate within a wall as a result of condensation of
water vapor. On rare occasions, accumulation of condensed water within a wall may provide enough moisture to
cause brickwork to effloresce.
Condensation typically contributes to efflorescence when the moisture originates inside buildings. Indoor air may
have a particularly high moisture content due to a buildings use (e.g., swimming pool) or certain activities that
take place inside such as cooking, bathing, washing and other operations employing water or steam. During cold
weather, building interiors conditioned to normal comfort settings may contain enough moisture to cause conden-
sation at lower temperatures.
If this water vapor reaches a vapor-resistant surface at a temperature below the dew point temperature, the vapor
will condense and can provide the moisture necessary to trigger the mechanism for efflorescence. This con-
densed moisture can contribute to efflorescence on the wall surface.
Condensation is typically controlled through design that considers the possibility of its occurrence. Using materi-
als with the appropriate water vapor permeance and preventing the transfer of air through the building envelope
reduce the possibility of condensation. See Technical Note 47 for detailed guidance on avoiding condensation.
CONSTRUCTION
Construction practices and workmanship employed in building masonry walls can significantly affect the amount
of water within masonry materials and the completed construction. Some discussion and recommendations for
proper construction practices follow.
Storage of Materials
The method of storing materials at a construction project site may influence future occurrence of efflorescence.
Masonry units, cementitious materials, sand and water should be stored off the ground, in such a manner as to
avoid contact by rain or snow, or contamination by dirt, plant life, organic materials or ground water, any of which
may contribute to efflorescence. These materials should also be covered by a waterproof membrane to keep them
dry.
Water
Water used in mortar or grout, for wetting brick, cleaning or other applications can also be a source of contami-
nants. Clean, potable water free of salts, deleterious acids, alkalies or organic materials should be used at all
times.
Workmanship
Workmanship characterized by the complete filling of all mortar joints intended to receive mortar is critical, as is
the need to keep all cavities and air spaces clean and free of mortar droppings. Attention to both of these items is
of primary importance in preventing efflorescence. Full mortar joints are more effective at reducing water penetra-
tion and clean cavities help masonry dry faster and separate brickwork from backings containing salts. Technical
Note 7B discusses recommended workmanship practices for masonry construction. Brick masonry with workman-
ship characterized by partially filled joints, deep furrowing of the mortar beds and improper execution of flashing
and sealant joint details is subject to increased rain penetration.
Unfinished Brickwork
Partially completed masonry walls exposed to rain and other elements during construction may become saturated
with water and depending on climatic conditions can require weeks, or even months to dry after the completion of
the building. This prolonged saturation may cause many slightly soluble salts, as well as the highly soluble salts
within masonry, to go into solution. Such conditions may also contribute to the contamination of the masonry with
soluble salts from elsewhere in the construction (concrete, plaster, trim, etc.).
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 7 of 9
During construction, Specification for Masonry Structures, ACI 530.1/
ASCE 6/TMS 602, requires the tops of all unfinished masonry to be
protected from the weather. At the end of each workday, shutdown
period or times of inclement weather, unfinished walls should be cov-
ered with water-resistant membranes or tarpaulins, securely fastened
or weighted in position. Mortar boards, scaffold planks and light plastic
sheets weighted with brick should not be accepted as suitable cover.
Additional recommendations and construction procedures for protect-
ing walls during cold weather are discussed in Technical Note 1.
Sealant Joints
Sealant joints should be integral elements of construction, designed
and installed with the same care as other elements of the structure.
Sealants and caulk should not be used as a means of correcting or
hiding poor workmanship that would permit water penetration.
Joints between masonry and door and window frames, expansion
joints and other locations where sealants are required are the most fre-
quent sources of rain penetration into masonry. These vulnerable loca-
tions should be given careful attention during design and construction.
Sealants and caulk do not have service lives as long as buildings. Maintenance programs should be established
to identify and replace sealants and caulk that have dried or become ineffective, as shown in Photo 4.
POST CONSTRUCTION
Troubleshooting Efflorescence
In addition to the information given above, the following measures can be used to help determine the cause and
extent of the efflorescence problem, and to suggest methods for repair and alleviation.
1. Determine the age of the structure at the time when the efflorescence first appeared. If new building
bloom is involved (structures less than one year old), the source of the salts is often the cement in the
mortar, and water that was used for or otherwise entered during construction is the source of moisture.
If the structure is over one year old, construction details should be examined to identify possible leaks in
the wall or in the surrounding construction. The appearance of efflorescence on an established building,
which had previously been free of efflorescence, is usually attributable to a new source of water in the
masonry.
2. Observe the location of the efflorescence, both on the structure and on the individual units or mortar
joints. The location on the building may offer some information as to where the water is entering. The
recent use or occupancy of the building should also be noted. For example, has it been vacant for some
time or has there been new construction? In short, what has occurred that might cause, or trigger, the
appearance of the efflorescence?
3. Examine the condition of the masonry. The profile of the mortar joints, the condition of the mortar, the
quality of workmanship employed, the condition of caulking and sealant joints, the condition of flashing
and drips, obstructed weeps, any deterioration or eroding of mortar joints in copings or in sills should all
be carefully noted. This information should offer clues as to the entry paths of moisture into the construc-
tion.
4. Review wall sections and details of construction for an indication of possible paths of moisture travel, and
for possible sources of contamination by soluble salts. A careful examination of roof and wall juncture and
flashing details should be made. If available, a comparison of contract drawings with as built drawings
may be helpful. This examination will also be useful for determining steps for repair or alleviation of the
efflorescence.
5. Review laboratory test reports on the materials of construction if they are available. This may help deter-
mine the source of the soluble salts, and may be of use in analyzing and making repair judgments.
Photo 4
Sealant Joint in Need of Repair
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 8 of 9
6. Identify the efflorescence. Knowing the composition of efflorescing salts is sometimes of use. Table 1,
taken from Brownells report lists the most probable sources of efflorescing salts. Commercial testing
laboratories can use X-ray diffraction, petrographic analysis or chemical analysis can, in some instances,
determine both the type of salts present and their relative quantity.
7. Consider miscellaneous sources of water when all obvious sources have been eliminated. Some of
these sources are: condensation within the wall, poorly directed landscape sprinklers, leaky pipes, faulty
drains and condensation on heating or plumbing pipes. If necessary, condensation analysis methods are
described in Technical Note 47.
When the mechanisms causing the efflorescing salts to appear have been established and the sources of salts or
moisture are identified (usually the latter), suitable corrections must be addressed. Solutions to efflorescence prob-
lems usually involve reducing water penetration into the masonry and removing the efflorescence from the wall.
Recommendations for the correction of water penetration in masonry walls are contained in Technical Note 46.
Removal of Efflorescence
The removal of efflorescence from the face of masonry is generally a relatively easy operation. Most efflorescing
salts are water-soluble and many will disappear of their own accord with normal weathering. This is especially true
of new building bloom.
It is usually not advisable to wash efflores-
cence off of the brickwork except in warm, dry
weather, since this results in the availability of
considerably more moisture which may bring
more salts to the surface. Many efflorescing
salts can be removed by dry brushing.
For recommendations concerning removing
efflorescence and other stains on masonry
walls, see Technical Note 20. Special care
should be exercised in cleaning new masonry,
since improper procedures and errors can
contribute to or cause efflorescence or stains.
Coatings
Clear water repellents, silicone and acrylic
coatings are among the solutions often sug-
gested for preventing efflorescence. A coating
may prevent efflorescence from recurring by
reducing the amount of water absorbed by an
exposed masonry surface. However, applying
a coating to brickwork that has a tendency to
effloresce, without stopping the mechanisms
causing that efflorescence, may lead to deg-
radation of the masonry. Further, applying a
coating that inhibits evaporation of water from
the masonry may also lead to degradation.
As water and dissolved salts within masonry travel toward a coated exterior surface, they may be stopped at the
inner margin of the coating (usually
1
/
8
to
1
/
4
in. [3 to 6 mm] below the surface). At this point, the water will evapo-
rate, passing through the treated area as vapor and soluble salts in the water will be deposited within the mason-
ry. Known as cryptoflorescnece or subflorescence, salts that crystallize within masonry can develop tremendous
pressures which may result in brick spalling. For this reason, coatings are not recommended as the sole treatment
for efflorescence problems.
Calcium sulf ate CaSO
4
2H
2
O Brick
Sodium sulf ate Na
2
SO
4
10H
2
O Cement-brick reactions
Potassium sulf ate K
2
SO
4
Cement-brick reactions
Calcium carbonate CaCO
3
Mortar or concrete
backing
Sodium carbonate Na
2
CO
3
Mortar
Potassium carbonate K
2
CO
3
Mortar
Potassium chloride KCl Acid Cleaning
Sodium chloride NaCl Sea Water
Vanadyl sulf ate VOSO
4
Brick
Vanadyl chloride VOCl
2
Acid Cleaning
Manganese oxide Mn
3
O
4
Brick
Iron oxide Fe
2
O
3
or
Fe(OH)
3
Iron in contact or brick
with "black core" or
"black heart"
Calcium hydroxide Ca(OH)
2
Cement
Pr incipal Ef f lor escing Salt Most Pr obable Sour ce
TABLE 1
Common Sources of Efflorescence
www.gobrick.com | Brick Industry Association | TN 23A | Efflorescence - Causes and Prevention | Page 9 of 9
OTHER STAINS
A number of stains that affect brick masonry are white in color and may be confused with efflorescence. There are
still other stains that are different in color, but have a mechanism of development similar to that of efflorescence.
Discussion of these stains is outside the scope of this Technical Note. Identification of these stains, as well as
their causes and prevention, is addressed in Technical Note 23.
SUMMARY
The development of efflorescence requires the presence of soluble salts and moisture. To prevent or minimize
efflorescence, the elimination of either will suffice. Design, detailing and construction practices that promote
resistance to water penetration are the most effective methods of preventing efflorescence. When efflorescence
does occur, careful observation is valuable in determining and correcting sources of water entry or salts prior to
removal.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. Annual Book of Standards, Vol. 04.05, ASTM International, West Conshohocken, PA, 2006.
- ASTM C 67, Standard Test Methods for Sampling and Testing Brick and Structural Clay Tile
- ASTM C 150, Standard Specification for Portland Cement
- ASTM C 1400, Standard Guide for Reduction of Efflorescence Potential in New Masonry Walls
2. Brownell, W.E., The Causes and Control of Efflorescence on Brickwork, Research Report No. 15,
Structural Clay Products Institute, 1969.
3. Merrigan, M.W., "Efflorescence: Cause and Control," The Masonry Society Journal, Vol. 5, No. 1, The
Masonry Society, Boulder, CO, 1986.
4. Miller, F. and Melander, J.M.,Efflorescence A Synopsis of the Literature, Proceedings of the Ninth
North American Masonry Conference, The Masonry Society, Boulder, CO, 2003.
5. Specification for Masonry Structures (ACI 530.1-05/ASCE 6-05/TMS 602-05), The Masonry Society,
Boulder, CO, 2005.
6. Young, J.E., Backup Materials as a Source of Efflorescence, Journal, American Ceramic Society, 40 (7),
1957.

Technical Notes 24 - The Contemporary Bearing Wall
(June 2002)
INTRODUCTION
Historically, the structural design of masonry buildings was based on the empirical requirements of building codes for
minimum wall thickness and maximum height. Bearing wall construction for buildings higher than three to five stories
was uneconomical and other methods of support (steel or concrete skeleton frame) were generally used. In 1965,
there was a renewed interest on the part of the design professional, architect and engineer, in modern bearing wall
construction, wherein the design is based on a rational structural analysis rather than on outmoded arbitrary
requirements. This interest was first stimulated by the work in Europe, where many loadbearing brick buildings
exceeding ten stories in height have been constructed during the past two decades.
EXAMPLES
One of the world's tallest thin-brick bearing wall structures built in 1957, is located near Zurich, Switzerland (Fig. 1).
This 18-story apartment structure utilizes interior loadbearing brick walls of 5 to 10 inches. in thickness. The exterior
walls (Fig. 2) are 15-1/4 inches. in thickness; the thickness in this instance being determined by the requirements for
thermal insulation rather than by structural requirements. By using cavity walls, the Swiss have found a way to
provide the required thermal insulation and still maintain relatively thin exterior walls. Figure 3 shows a 16-story
apartment building in Grenchen, Switzerland which utilizes cavity wall construction. The exterior walls of this building
(Fig. 4) are comprised of a 6-inches. brick inner bearing wythe, a 1-1/2-inches. cavity and a 5-inches. exterior brick
wythe. The floor loads are carried by the 6-inches. inner wythe and 6-inches. interior brick bearing partitions. The
5-inches. exterior wythe of the exterior wall is self-supporting on the foundation and is tied only at each floor level,
with 1/4-inches. stainless steel wire anchors embedded in the edge of the slab at approximately 20 inches. on
center.
FIG. 1
http://www.gobrick.com/BIA/technotes/t24.htm
1 of 9 9/13/2009 1:13 PM
Bearing Wall @ Floor Slab
FIG. 2
FIG. 3
http://www.gobrick.com/BIA/technotes/t24.htm
2 of 9 9/13/2009 1:13 PM
Cavity Wall @ Floor Slab
FIG. 4
Many of the high-rise brick bearing wall buildings in Switzerland are stuccoed on the exterior; this being their
traditional method of building. However, there is an increasing use of exposed clay masonry units. Another example
(Fig. 5) is a 14-story structure in Lucerne. The exterior bearing walls of this structure are only 7-1/4 inches. in
thickness (Fig. 6). The resistance to water penetration is provided by parging on the interior of the exterior wall
rather than stucco on the outside.
FIG. 5
http://www.gobrick.com/BIA/technotes/t24.htm
3 of 9 9/13/2009 1:13 PM
Brick & Clay Tile Cavity Wall @ Floor Slab
FIG. 6
Recent American examples of brick bearing wall construction are shown in Figs. 7, 9 and 11. Figure 7 is the 17-story
Park Mayfair East building, in Denver, Colorado. The structural system in this building utilizes 11-inches. reinforced
brick masonry (RBM) walls for the full building height (165 feet). Fourteen-inch prestressed concrete twin-tee slabs
were used for the floor spans, which vary from 32 to 37 feet (Fig. 8).
FIG. 7
http://www.gobrick.com/BIA/technotes/t24.htm
4 of 9 9/13/2009 1:13 PM
Cavity Wall @ Precast Tee Floor
FIG. 8
Figure 9 illustrates the 8-story Oakcrest Towers apartment building in Prince Georges County, Maryland near
Washington, D. C. This building utilizes 8-inches. brick exterior walls for the full height of the building, and 6-inches.
brick bearing corridor walls. The floor system, in this instance is steel joints with 2-1/2-inches. concrete topping over
metal deck for the 24-feet. apartment span, and a 5-inches. flat concrete slab or the 6-feet corridor span. Figure 10
is the typical detail for the 8-inches. exterior and the 6-inches. corridor walls.
FIG. 9
http://www.gobrick.com/BIA/technotes/t24.htm
5 of 9 9/13/2009 1:13 PM
Interior Bearing Wall @ Floor Truss & Concrete Slab
FIG. 10a
Exterior Bearing Wall @ Floor Truss & Concrete Slab
FIG. 10b
The Muskegon Retirement Apartments (Fig. 11) is an 11-story, 194-unit structure in Muskegon, Michigan. The
structure consists of 8-inches. solid brick bearing cross walls which support the 8-inches. precast hollow core
concrete plank floor system. Non-bearing exterior walls are 10-inches. brick cavity walls, insulated with water-
repellent vermiculite (Figs. 12 and 13). Exposed brick forms the interior finish on the bearing cross walls, which are
http://www.gobrick.com/BIA/technotes/t24.htm
6 of 9 9/13/2009 1:13 PM
also separation walls between units. These walls provide attractive low maintenance surfaces, fire resistance,
structure, and excellent sound separation.
FIG. 11
Interior Bearing Wall
FIG. 12
http://www.gobrick.com/BIA/technotes/t24.htm
7 of 9 9/13/2009 1:13 PM
Exterior Non-Bearing Wall
FIG. 13
BUILDING CODES
The Structural Clay Products Institute published the first edition of a rational design standard for engineered brick
masonry in May of 1966. This first Standard was based on laboratory research and historical performance data. In
August 1969, the Institute (now the Brick Institute of America), developed and published a second generation
standard, Building Code Requirements for Engineered Brick Masonry,
For easy reference, the BIA Standard has been accepted in substance or by reference in the following model
building codes:
Model Building Codes
1. BOCA Basic/National Building Code, 1984 Edition.
2. Standard Building Code, 1985 Edition.
3. Uniform Building Code, 1985 Edition; Chapter 24 - Masonry.
In addition to the model building codes, many local municipal sub-divisions - cities, counties and townships - have
http://www.gobrick.com/BIA/technotes/t24.htm
8 of 9 9/13/2009 1:13 PM
also adopted, by reference or in principle, these engineered brick masonry provisions and requirements.
ECONOMICS
Experience to date indicates that brick masonry bearing wall construction is competitive with structural frame, and
for many types of buildings can be constructed at less cost. This economic advantage is due in part to the increased
efficiency of the use of materials. For example, in a brick bearing wall structure, the brick walls become enclosure,
separation, structure, finish and fire protection. The economic position is also the result of the simplified and faster
construction process.
Mr. Harold Simpson, General Contractor and part owner of the Park Mayfair East project in Denver (see Figs. 7 and
8), stated these reasons for using the brick bearing wall structural system:
"...As an owner, the number one consideration is lower initial and future costs, which means we can charge lower
rents and have lower vacancy rates. The next important single criterion of a good apartment is sound control. For
this, the 11 - inches. wall of brick and grout gave us a sound resistance of 58 decibels which is excellent. Another
important aspect is the shorter construction time which allows earlier occupancy. This means lower interest
payments on construction loans and earlier rent payments. In fact as the building was being topped out, some of the
lower floor apartments were complete with carpets and draperies, and future tenants were visiting the fully furnished
display apartments.
"As a contractor, I look at this brick bearing wall structural system much the same as an owner, but with some
differences. The faster erection time is very important because time is money. Brick bearing walls give me lower
costs and faster erection by being the structural system as well as the enclosure walls. When the bearing walls are
complete and the floor slabs have been placed, we have only to insert the windows and (1) the floor is closed in, (2)
the exterior walls are complete, (3) the structure is fireproof. In addition, we have fire walls between apartments and
many of the interior walls are also complete. This is not true with a steel or concrete structure. This finished building,
including appliances, draperies, carpeting and landscaping, cost $12.25 per sq. feet. of floor area. This total includes
some costs that would not have to be duplicated if we constructed the buildings again, but, with labor and material
cost increases expected in the future, a new building similar to this one would probably cost about the same.'
REFERENCES
1. Monk, Clarence B., Jr. and Gross, James G., European Clay Masonry Loadbearing Buildings, SCPI,
1964.
2. Proceedings - The First National Brick and Tile Bearing Wall Conference, SCPI, 1965.
3. "Contemporary Brick Bearing Wall" Case Study, BIA .
4. Technical Notes on Brick and Tile Construction, BIA (a monthly series).
5. Building Code Requirements for Engineered Brick Masonry, BIA, August 1969
6. Gross, James G.; Dikkers, Robert D.; and Grogan, John C., Recommended Practice for Engineered Brick
Masonry, BIA. November 1969.
http://www.gobrick.com/BIA/technotes/t24.htm
9 of 9 9/13/2009 1:13 PM

Technical Notes 24C - The Contemporary Bearing Wall - Introduction to Shear Wall Design
[Sept./Oct. 1970] (Reissued May 1988)
INTRODUCTION

The general design concept of the contemporary bearing wall building system depends upon the combined structural
action of the floor and roof systems with the walls. The floor system carries vertical loads and, acting as a
diaphragm, lateral loads to the walls for transfer to the foundation. Lateral forces of wind and earthquake are usually
resisted by shear walls which are parallel to the direction of the lateral load. These shear walls, by their shearing
resistance and resistance to overturning, transfer the lateral loads to the foundation. See Fig. 1.


Shear Wall Action
FIG. 1



It is the purpose of this Technical Notes to discuss some of the factors involved in the design of brick masonry shear
24c http://www.gobrick.com/BIA/technotes/t24c.htm
1 of 26 9/13/2009 1:14 PM
walls and to present some of the available test data regarding their strength. Other issues of Technical Notes will
contain examples relating to the design of brick masonry shear walls.

LATERAL FORCES

The principal lateral forces to be considered in the design of shear walls are wind pressure and earthquake. Most
building codes and engineering practice standards specify that wind and earthquake may be assumed never to occur
simultaneously.
Wind Pressure. Building codes usually specify design wind load requirements which should be considered minimum.
For additional information on wind pressure, the designer may refer to the American Standard Building Code
Requirements for Minimum Design Loads in Buildings and Other Structures, A58.1 - 1955, and "Wind Forces on
Structures", ASCE Transactions, Vol. 126, Part II, 1961.
Earthquake. Unlike wind pressure, earthquake forces on a structure are considered a function of the mass and
stiffness of the structure. Generally speaking, subject to dynamic phenomena, the greater the weight and rigidity of
the structure - the greater are the forces which must be resisted by the structure. It is, therefore, current engineering
practice to design "box systems" (structures without complete vertical load-carrying space frames) to resist the
greater lateral forces. In this type of structural system the lateral forces are resisted by the shear walls. The
Uniform Building Code, 1970 edition, contains the following requirements pertaining to minimum earthquake forces
for structures:
Every structure shall be designed and constructed to withstand minimum total lateral seismic forces assumed to act
nonconcurrently in the direction of each of the main axes of the structure in accordance with the following formula:
V = Z K C W
where:

V = total lateral load or shear at the base
Z = numerical coefficient dependent upon the zone of seismic activity:
Z = 1/4 for Zone 1 (minor damage);
Z = 1/2 for Zone 2 (moderate damage);
Z = 1 for Zone 3 (major damage)

K = numerical coefficient from Table 1
C = numerical coefficient dependent on fundamental period of vibration of
the structure (in seconds) in the direction considered
W = total dead load of the structure.

24c http://www.gobrick.com/BIA/technotes/t24c.htm
2 of 26 9/13/2009 1:14 PM


For additional information on earthquake forces and design, the designer may refer to the following publications:
1. Uniform Building Code, 1970 edition, International Conference of Building Officials.
2. Reinforced Brick Masonry and Lateral Force Design, Harry C. Plummer and John C. Blume, Structural Clay
Products Institute, 1953.
3. Seismic Design for Buildings, (Department of the Army, TM 5-809-10; Department of the Navy, NAVDOCKS
P-355; Department of the Air Force, AFM 88-3, Chapter 13), March 1966.
4. Earthquake Engineering Research, The Committee on Earthquake Engineering Research - Division of
Engineering - National Research Council, National Academy of Engineering, National Academy of Sciences, 1969.

DISTRIBUTION OF LATERAL FORCES
Diaphragms. Horizontal distribution of lateral forces to shear walls is achieved by the floor and roof systems acting
as diaphragms (see Fig. 2).

24c http://www.gobrick.com/BIA/technotes/t24c.htm
3 of 26 9/13/2009 1:14 PM

Diaphragm Action
FIG. 2



To qualify as a diaphragm, a floor and roof system must be able to transmit the lateral forces to the shear walls
without exceeding a deflection which would cause distress to any vertical element. The successful action of a
diaphragm also requires that it be properly tied into the supporting shear walls. The designer should insure this
action by appropriate detailing at the juncture between horizontal and vertical structural elements of the building.
Diaphragms may be considered as analogous to horizontal (or inclined, in the case of some roofs) plate girders. The
roof or floor slab constitutes the web; the joists, beams and girders function as stiffeners; and the walls or bond
beams act as flanges.
Diaphragms may be constructed of materials such as concrete, wood or metal in various forms. Combinations of
such materials are also possible. Where a diaphragm is made up of units such as plywood, precast concrete planks
or steel deck units, its characteristics are, to a large degree, dependent upon the attachments of one unit to another
and to the supporting members. Such attachments must resist shearing stresses due to internal translational and
rotational actions.
The stiffness of a horizontal diaphragm affects the distribution of the lateral forces into the shear walls. No
diaphragm is infinitely rigid or flexible. However, for the purpose of analysis, diaphragms may be classified into three
groups: rigid, semirigid or semiflexible, and flexible.
A rigid diaphragm is assumed to distribute horizontal forces to the vertical resisting elements in proportion to their
relative rigidities (see Fig. 3).

24c http://www.gobrick.com/BIA/technotes/t24c.htm
4 of 26 9/13/2009 1:14 PM

FIG. 3



Semirigid or semiflexible diaphragms are those which have significant deflections under load, but which also have
sufficient stiffness to distribute a portion of the load to the vertical elements in proportion to the rigidities of the
vertical resisting elements. The action is analogous to a continuous beam system of appreciable stiffness on yielding
supports (see Fig. 4). The support reactions are dependent upon the relative stiffness of both diaphragm and the
vertical resisting element.


24c http://www.gobrick.com/BIA/technotes/t24c.htm
5 of 26 9/13/2009 1:14 PM
FIG. 4



A flexible diaphragm is analogous to a shear deflecting continuous beam or series of beams spanning between
supports. The supports are considered non-yielding, and the relative stiffness of the vertical resisting elements
compared to that of the diaphragm is great. Thus, a flexible diaphragm is considered to distribute the lateral forces
to the vertical resisting elements on a tributary area basis (see Fig. 5).


FIG. 5



Where the center of rigidity of a shear wall system does not coincide with the center of application of the lateral
force, the distribution of the rotational forces due to a torsional moment on the system must also be considered.
Where rigid or semirigid diaphragms are used, it may be assumed that the torsional forces are distributed to the
shear walls in direct proportion to their relative rigidities and their distance from the center of rigidity (see Fig. 6). In
the design provisions for earthquake forces of the 1970 Uniform Building Code, shear resisting elements are
required to resist an arbitrary torsional moment equivalent to the story shear acting with an eccentricity of not less
than five per cent of the maximum building dimension at that level. A flexible diaphragm is not considered capable of
distributing torsional stresses.

24c http://www.gobrick.com/BIA/technotes/t24c.htm
6 of 26 9/13/2009 1:14 PM

Torsional Movement on a Shear Wall System
FIG. 6

24c http://www.gobrick.com/BIA/technotes/t24c.htm
7 of 26 9/13/2009 1:14 PM


When dealing with a rigid diaphragm and distributing the horizontal forces to vertical resisting elements in proportion
to the relative rigidities, the relative rigidity of the shear wall is dependent upon the shear and flexural deflections.
However, for the proportions of the shear walls in most high-rise buildings, the flexural deflection greatly exceeds the
shear deflection, in which case only flexural rigidity need be considered in determining the relative stiffness of the
shear walls. For determination of relative shear and flexural deflections, see Fig. 7.


Relative Shear and Flexural Deflections Determined
for a Uniformly Loaded Cantilever Member of Rectangular Section
FIG. 7
24c http://www.gobrick.com/BIA/technotes/t24c.htm
8 of 26 9/13/2009 1:14 PM



A rigorous analysis of the lateral load distribution to the shear wall is sometimes very time-consuming and frequently
unjustified by the results. Therefore, in many cases a design based on reasonable limits may be used. For example,
the load may be distributed by first considering the diaphragm as rigid and then by considering it flexible. If the
difference is not great, the shear wall can then be safely designed for the maximum applied load.
Diaphragm Deflection. As previously indicated, deflection is another factor that must be considered in designing a
horizontal diaphragm. As shown in Fig. 8, diaphragm deflection should be limited to prevent excessive stresses in the
walls which are perpendicular to the shear walls. The following formula has been suggested by the Structural
Engineers Association of Southern California for allowable deflection of horizontal diaphragms in buildings having
masonry or concrete walls:


Diaphragm Deflection Limitation
FIG. 8




where:D = allowable deflection between adjacent supports of wall, in inches

24c http://www.gobrick.com/BIA/technotes/t24c.htm
9 of 26 9/13/2009 1:14 PM
h = height of wall between adjacent horizontal supports, in feet
t = thickness of wall, in inches
f = allowable flexural compressive stress of wall material, in pounds per square
inch
E = modulus of elasticity of wall material, in pounds per square inch
The application of these limits on deflection must be used with engineering judgment. For example, continuity at floor
level is assumed, which in many cases is not present due to through-wall flashing. In this situation the deflection may
be based on the allowable compressive stress in the masonry, assuming a reduced cross section of wall. The effect
of reinforcement which may be present in a reinforced brick masonry wall or as a tie to the floor system in a
non-reinforced or partially reinforced masonry wall is not considered. It should also be pointed out that the limit on
deflection is actually a limit on differential deflection between two successive floor or diaphragm levels.
Maximum span-to-width or depth ratios for diaphragms are usually used to indirectly control diaphragm deflection.
Normally, if the diaphragm is designed with the proper ratio, the diaphragm deflection will not be critical. As a guide
for horizontal diaphragm proportions, Table 2, taken from the State of California Administrative Code, Title 21, Public
Works, may prove useful.

1
The use of diagonal sheathed or unblocked plywood diaphragms for buildings having masonry or reinforced concrete walls shall be limited to one-story
buildings or to the roof of a top story.

Rigidity of Shear Walls. Where shear walls are connected by a rigid diaphragm so that they must deflect equally
under horizontal load, the proportion of total horizontal load at any story or level carried by a perpendicular shear
wall is based on its relative rigidity or stiffness. The rigidity of a shear wall is inversely proportional to its deflection
under unit horizontal load. The total deflection of the shear wall can be determined from the sum of the shear and
moment deflections. Equations for the deflection of fixed and cantilevered walls or piers are shown in Fig. 9.

24c http://www.gobrick.com/BIA/technotes/t24c.htm
10 of 26 9/13/2009 1:14 PM

Calculation of Wall Deflections
FIG. 9

24c http://www.gobrick.com/BIA/technotes/t24c.htm
11 of 26 9/13/2009 1:14 PM


Where a shear wall contains no openings, the computations for deflection and rigidity are quite simple. In Fig. 10(a),
the shear walls are of equal length and rigidity, and each takes one half of the total load. In Fig. 10(b), wall C is one
half the length of wall D and it, therefore, receives less than one eighth of the total load. Where shear walls contain
openings such as doors and windows, the computations for deflection and rigidity are more complex. However,
approximate methods have been developed which may be used. See Fig. 11.


Distribution of Wind Load
FIG. 10



24c http://www.gobrick.com/BIA/technotes/t24c.htm
12 of 26 9/13/2009 1:14 PM

Calculation of Wall Rigidity
FIG. 11
24c http://www.gobrick.com/BIA/technotes/t24c.htm
13 of 26 9/13/2009 1:14 PM



To increase the stiffness of shear walls as well as their resistance to bending, intersecting walls or flanges may be
used. Very often in the design of buildings, Z, T, U, and I-shape sections develop as natural parts of the design. See
Figs. 12 and 13. Shear walls with these shapes, of course, have better flexural resistance. The 1969 BIA Standard,
Building Code Requirements for Engineered Brick Masonry, (Sec. 4.7.12A) limits the effective flange width that
may be used in calculating flexural stresses. In the case of symmetrical T or I sections, the effective flange width
may not exceed one sixth of the total wall height above the level being analyzed. In the case of unsymmetrical L or C
sections, the width considered effective may not exceed one sixteenth of the total wall height above the level being
analyzed. In either case, the overhang for any section may not exceed six times the flange thickness (see Figs. 14
and 15). It is, of course, necessary to insure that the shear stress at the intersection of the walls does not exceed
the permissible shear stress. This will depend on the method used in bonding the two walls together.


Shear Walls of Equivalent Stiffness
FIG. 12



24c http://www.gobrick.com/BIA/technotes/t24c.htm
14 of 26 9/13/2009 1:14 PM

Shear Walls With Flanges
FIG. 13




Effective Flange Width
FIG. 14



24c http://www.gobrick.com/BIA/technotes/t24c.htm
15 of 26 9/13/2009 1:14 PM

Effective Flange Width
FIG. 15



Coupled Shear Walls. Another method that may be used to increase the stiffness of a bearing wall structure and
reduce the possibility of tension developing in shear walls due to wind parallel to the wall is the coupling of collinear
shear walls. Figures 16 and 17 indicate the effect of coupling on the stress distribution in the wall due to parallel
forces. A flexible connection between the walls is assumed in Figs. 16(a) and 17(a), so that the walls act as
independent vertical cantilevers in resisting the lateral loads. Figures 16(b) and 17(b) assume the walls to be
connected with a more rigid member which is capable of shear and moment transfer so that a frame-type action
results. This can be accomplished with a steel, reinforced concrete or reinforced brick masonry section. The plate
type action, which is indicated in Figs. 16(c) and 17(c), assumes an extremely rigid connection between walls, such
as full story height walls or deep rigid spandrels.



24c http://www.gobrick.com/BIA/technotes/t24c.htm
16 of 26 9/13/2009 1:14 PM

End Shear Walls
FIG. 16




Interior Shear Walls
FIG. 17

SHEAR STRENGTH
Test Data. The present standard racking test, described in ASTM E 72-68, Method of Conducting Strength Tests of
Panels for Building Construction, provides only a relative measure of the shearing or diagonal tension resistance of
a wall. Results of this test method are consequently valid only for comparison purposes and are not suggested for
determination of design values.

In this method of test, horizontal movement of the wall specimen (8 by 8 ft), due to the horizontal racking load at the
top of one end, is prevented by a stop block at the bottom of the other end. To counteract rotation of the specimen
due to this overturning couple, tie rods are used near the loaded edge of the wall specimen. Under racking load
these rods superimpose an indeterminate compressive force which suppresses the critical diagonal tensile stresses
and increases the load required to rack the specimen. A summary of the racking data for non-reinforced brick
24c http://www.gobrick.com/BIA/technotes/t24c.htm
17 of 26 9/13/2009 1:14 PM
masonry walls tested in accordance with ASTM E 72 is given in Table 3. A typical mode of failure for a 4-in. brick
wall subject to racking is shown in Fig. 18.


1
Tested in accordance with ASTM E 72.
2
Walls not loaded to failure.
3
"Structural Properties of Six Masonry Wall Constructions", H. L. Whittemore, A. H. Stang and D. E. Parsons,
National Bureau of Standards Report BMS5, 1938.
4
"Structural Properties of a Brick Cavity-Wall Construction". H. L. Whittemore. A. H. Stang, D. E. Parsons,
National Bureau of Standards Report BMS23, 1939.
5
"SCR brick'* Wall Tests", C. R. Monk, Jr., Structural Clay Products Research Foundation, Research Report
No. 1, June 1953.
6
"Compressive, Transverse and Racking Strength Tests of Four-Inch Brick Walls", Structural Clay Products
Research Foundation, Research Report No. 9, August 1965.
*Reg. U.S. Pat. Off., SCPI



24c http://www.gobrick.com/BIA/technotes/t24c.htm
18 of 26 9/13/2009 1:14 PM

Typical Mode of Failure in Racking

FIG. 18

Circular Shear Specimens. Based on experimental work done at the Balcones Research Center of the University of
Texas by Professors Neils Thompson and Frank Johnson, the Research Division of the Structural Clay Products
Institute conducted a series of diagonal tensile tests on circular brick masonry specimens. In these tests, a 15-in.
diameter specimen is tested in compression with the line of load at 45 deg to the bed joints. As shown in Fig. 19 the
diametrical stresses are largely tensile over the central 80 per cent of the specimen. The tensile stress is
approximately constant for about 60 per cent of the diameter and may be calculated by the following equation:




Stress Distribution in Tensile Splitting Test
24c http://www.gobrick.com/BIA/technotes/t24c.htm
19 of 26 9/13/2009 1:14 PM
FIG. 19


where: P = load at rupture, in pounds

D = diameter of specimen, in inches
t = thickness of specimen, in inches

The test results for 133 specimens built with 27 types of brick and type S portland cement-lime mortar of ASTM C
270, Specifications for Mortar for Unit Masonry, are summarized in Table 4. A typical mode of failure for a circular
brick specimen is shown in Fig. 20. While there was no consistent relationship between diagonal tensile strength and
brick properties, such as initial rate of absorption, it appeared that brick with the weakest bond characteristics as
shown by flexural strength values also yielded the lowest diagonal tensile strengths.


1
"Small Scale Specimen Testing", Progress Report No. 1, SCPI-SCPRF, October 1964.
2
All specimens built with type S mortar.



24c http://www.gobrick.com/BIA/technotes/t24c.htm
20 of 26 9/13/2009 1:14 PM

Small-Scale Diagonal Tension Test
FIG. 20

The test results for 20 circular specimens built with one type of brick and four types of mortars are summarized in
Table 5. As indicated, the mortar type had a marked effect on the diagonal tensile strength of circular back masonry
specimens.


1
"Small Scale Specimen Testing", Progress Report No. 1. SCPI-SCPRF, October 1964.
2
C = portland cement; L = hydrated lime (type S); S = sand.
3
28-day briquets.
4
All specimens built with 3/8-in. joints and brick having an average compressive strength of 11,771 psi and an initial rate of absorption of 10.6 g per
min per 30 sq in.

Square Shear Specimens. The Brick Institute of America has continued to study the diagonal tensile or shear
strength of brick masonry in an effort to develop both test methods and design information. Working with the
National Bureau of Standards and the Research Division of John A. Blume and Associates (San Francisco), a test
method has been developed that has several advantages over the ASTM E 72 racking test procedure. As previously
stated, in the E 72 procedure the hold-down tie rods required to prevent overturning of the specimen under load
produce an indeterminate bearing condition at the bottom edge of the specimen, thus preventing an analytic
determination of the stress within the wall specimen itself.
24c http://www.gobrick.com/BIA/technotes/t24c.htm
21 of 26 9/13/2009 1:14 PM

In the alternate test procedure, which will be submitted to ASTM as a proposed alternate to the ASTM E 72
procedure, the specimens are nominally 4 ft by 4 ft as opposed to the 8-ft square specimen in the E 72 procedure.
In this method, the test results are susceptible to stress analysis. In addition, they are more reproducible and thus
more reliable for comparison and design data purposes.
The square specimen is placed in the testing frame so as to be loaded in compression along a diagonal, thus
producing a diagonal tension failure with the specimen splitting apart along the loaded diagonal (see Figs. 21 and
22). The results of both small scale (2 ft square) and full scale (4 ft square) tests of 4, 6 and 8-in. thick brick
masonry specimens are summarized in Table 6 and typical shear stress-strain curies are displayed in Fig. 23. All
specimens were constructed with type S mortar. The shearing stress V'm is determined by the equation


where: F = diagonal compressive force or load, in pounds

t = thickness of wall specimen, in inches

l = length of a side of a square specimen, in inches


Wallette Shear Test Setup
FIG. 21

24c http://www.gobrick.com/BIA/technotes/t24c.htm
22 of 26 9/13/2009 1:14 PM

Full Scale Shear Setup
FIG. 22


Typical Shear Stress-Strain Curves for 6 and 8 Inch Walls
FIG. 23
24c http://www.gobrick.com/BIA/technotes/t24c.htm
23 of 26 9/13/2009 1:14 PM


12 ft square.
24 ft square.

Effects of Normal Loads. The Brick Institute of America has also investigated the effects of compressive loads
normal to the bed joints on the shearing strength of plain (non-reinforced) brick masonry walls. Table 7 summarizes
the results of these tests with the normal compressive load on the wall varying from 0 (unloaded) to 375 psi. The
specimens were built with one type of brick utilizing type S mortar and inspected workmanship. All specimens were
two wythes, 8 in. in thickness and bonded with metal ties.


1Built with type S mortar, inspected workmanship and metal-tied.
fb = 11,100 psi

24c http://www.gobrick.com/BIA/technotes/t24c.htm
24 of 26 9/13/2009 1:14 PM
IRA = 20.7 g/min/30 in2

Allowable Stresses. The allowable shear stresses for non-reinforced masonry provided in the 1969 SCPI Standard,
Building Code Requirements for Engineered Brick Masonry, are shown in Table 8, and for reinforced masonry in
Table 9.

The Standard also provides a basis for the design of biaxially loaded shear walls. Section 4.7.12.1 states:
"In non-reinforced shear walls, the virtual eccentricity (el) about the principal axis which is normal to the length (l) of
the shear wall shall not exceed an amount which will produce tension. In non-reinforced shear walls subject to
bending about both principal axes, (etl + elt) shall not exceed (tl / 3) where et = virtual eccentricity about the principal
axis which is normal to the thickness (t) of the shear wall. Where the virtual eccentricity exceeds the values given in
this section, shear walls shall be designed in accordance with Section 4.7.9 or 4.7.11". (Reinforced or partially
reinforced walls.)

Provision is also made in the standard for increasing the shear capacity of the wall by taking into consideration
compressive loads on the wall. Section 4.7.12.3 states:
"The allowable shearing stresses in non-reinforced and reinforced shear walls shall be taken as the allowable
stresses given in Tables 3 and 4 (Tables 8 and 9 of this Technical Notes) ,respectively, plus one fifth of the average
compressive stress due to dead load at the level being analyzed. In no case, however, shall the allowable shear
stresses exceed the maximum values given in Tables 3 and 4.
"In computing the shear resistance of the wall, only the web shall be considered."



24c http://www.gobrick.com/BIA/technotes/t24c.htm
25 of 26 9/13/2009 1:14 PM


REFERENCES

1.Technical Notes on Brick Construction, BIA, (a monthly series).
2.Gross, J. G.; Dikkers, R. D.; and Grogan, J. C.; Recommended Practice for Engineered Brick Masonry,
SCPI, November 1969.
3.Allen M. H. and Watstein, D.; Compressive, Transverse and Shear Strength Tests of Six and Eight-lnch
Single-Wythe Walls Built with Solid and Heavy-Duty Hollow Clay Masonry Units, Research Report No. 16,
SCPI, September 1969.
4.Blume, J. A. and Prolux, J.; Shear in Grouted Brick Masonry Wall Elements, Western States Clay
Products Association, August 1968.
24c http://www.gobrick.com/BIA/technotes/t24c.htm
26 of 26 9/13/2009 1:14 PM

Technical Notes 24F - The Contemporary Bearing Wall - Construction
[Nov./Dec. 1974] (Reissued Sept. 1988)
INTRODUCTION
The Contemporary Bearing Wall concept as conceived and being applied today is based upon rational engineering
design. This concept requires floors and walls to work together as a system, each giving support to the other. A
building of high strength, in which the structure provides finish, closure, partition, sound control and fire resistance, is
thereby provided. In order to achieve this end, it is necessary that proper attention be given to design details and
construction procedures. It is of utmost importance that constructors follow the plans and specifications of the
designers.
Attention to detail and requirements for high quality materials and workmanship have not deterred the rapid
acceptance and application of the Contemporary Bearing Wall concept. Numerous cases can be cited where bearing
wall buildings have been built faster than scheduled or anticipated, and, in many cases, these buildings have been
built at less than the estimated cost of alternate designs utilizing other materials and structural systems.
SPECIAL REQUIREMENTS
The design of these buildings may be based upon the Building Code Requirements for Engineered Brick Masonry,
SCPI (BIA), August 1969, which contains various additional construction requirements not required by most other
modern building codes. The following sections relating to both materials and workmanship are specifically called to
the attention of the reader:
1.1 SCOPE
1.1.1 General This standard provides minimum requirements for the design and construction of brick masonry of solid masonry units, both plain
(non-reinforced) and reinforced. It does not include requirements for construction using hollow masonry units nor requirements for fire protection.
1.1.2 Analysis The design of brick masonry shall be based on a general structural analysis and the requirements of this standard.
1.1.3 Special Structures For arches, garden walls, retaining walls, tanks, reservoirs and chimneys, the provisions of this standard shall govern so
far as they are applicable.
1.2 PERMITS AND DRAWINGS
1.2.1 Copies of structural drawings and typical details showing the sizes and position of all structural members, steel reinforcement, design
strengths, and live loads used in the design shall be filed with the building department before a permit to construct such work shall be issued.
Calculations pertaining to the design shall be filed with the drawings when required by the building official.
1.3 INSPECTION
1.3.1 With Inspection When the design of brick masonry is based on the allowable stresses and other values given in Tables 1, 2, 3, 4 and 5 for
"With Inspection'', the construction shall be inspected by an engineer or architect, preferably the one responsible for the design, or by a competent
representative responsible to him. Such inspection shall be of a nature as to determine, in general, that the construction and workmanship are in
accordance with the contract drawings and specifications.
1.3.2 Without Inspection When there is no engineering or architectural inspection as specified in Section 1.3.1, the allowable stresses and other
values given in Tables 1, 2, 3, 4 and 5 for "Without Inspection" shall be used.
2.2.1 Brick
2.2.1.1 Brick and Solid Clay or Shale Masonry Units Standard Specification for Building Brick (Solid Masonry Units Made from Clay or Shale),
ASTM C 62, or Standard Specification for Facing Brick (Solid Masonry Units Made from Clay or Shale), ASTM C 216.
2.2.1.2 Grades and Types Brick subject to the action of weather or soil, but not subject to frost action when permeated with water, shall be of grade
MW or grade SW, and where subject to temperature below freezing while in contact with soil shall be grade SW. Brick used in loadbearing or shear
walls shall comply with the dimension and distortion tolerances specified for type FBS of ASTM C 216. Where such brick do not comply with these
tolerance requirements, the compressive strength of brick masonry shall be determined by prism tests. (See Section 4.2.2.1.)
http://www.gobrick.com/BIA/technotes/t24f.htm
1 of 11 9/13/2009 1:15 PM
2.2.1.3 Used Brick Used or salvaged brick shall not be permitted under the provisions of this standard.
2.2.2 Mortar and Grout
2.2.2.1 Non-Reinforced Brick Masonry Mortar for use in non-reinforced brick masonry shall conform to Standard Specification for Mortar for Unit
Masonry, ASTM C 270, types M, S or N. except that it shall consist of a mixture of portland cement (type I, II or III), hydrated lime (type S) and
aggregate where values given in Tables 1, 2 and 3 are used.
2.2.2.2 Reinforced Brick Masonry Mortar and grout for use in reinforced brick masonry shall conform to Standard Specification for Mortar and
Grout for Reinforced Masonry, ASTM C 476, except that mortar shall consist of a mixture of portland cement (type I, II or III), hydrated Iime (type S )
and aggregate where values given in Tables 2 and 4 are used.
2.2.2.3 Air-entraining admixtures or hydrated lime containing air-entraining admixtures shall not be used in mortar.
2.2.2.4 Calcium chloride or admixtures containing calcium chloride shall not be used in mortar or grout in which reinforcement, metal ties or anchors
are embedded.
2.2.2.5 Other mortars not specified in Sections 2.2.2.1 and 2.2.2.2 may be used when approved by the building of official, provided strengths for such
masonry construction are established by tests made in accordance with Section 4.2.2.1 and Standard Methods of Conducting Strength Tests of
Panels for Building Construction, ASTM E 72.
5.2.1 Mortar Joints All brick shall be laid with full head and bed joints and all interior joints that are designed to receive mortar shall be filled. (See
Section 5.8.3. ) The average thickness of head and bed joints shall not exceed 1/2 inch.
5.2.3 Tolerances for Brick Masonry Construction Based on Actual Dimensions
5.2.3.1 Variation from the Plumb
(1) In the lines and surfaces of columns, walls and arrises: in 10 feet-1/4 inch; in any story or 20 feet maximum-3/8 S inch; in 40 feet or more-1/2
inch.(2) For external corners, expansion joints and other conspicuous lines: in any story or 20 feet maximum-1/4 inch; in 40 feet or more-1/2 inch.
5.2.3.2 Variation from the Level or the Grades Indicated on the Drawings
(1) For exposed lintels, sills, parapets, horizontal grooves and other conspicuous lines: In any bay or 20 feet maximum-1/2 inch; in 40 feet or
more-3/4 inch.
5.2.3.3 Variation of the Linear Building Lines from Established Position in Plan and Related Portion of Columns, Walls and Partitions
(1) In any bay or 20 feet maximum-1/2 inch; in 40 feet or more-3/4 inch.
5.2.3.4 Variation in Cross-Sectional Dimensions of Columns and in the Thickness of Walls
(1) Minus 1/4 inch; plus 1/2 inch.
5.9 CHASES AND RECESSES
5.9.1 Chases and recesses shall be considered in the structural design and detailed on the building plans. Chases not shown on the plans shall be
permitted only when approved in writing by the structural engineer and the building official.
EXPERIENCE AND EXAMPLES
Pennley Park. This complex of eight buildings was one of the first major U.S. projects utilizing rationally designed
brick bearing walls as a major structural system, (Fig. 1). The cost was approximately $4,500,000 or $13.89 per sq
ft of floor area including architectural and engineering fees, soil analysis and site work. The project was scheduled
for 21 months of construction; it started May 1, 1964 and 12 months later was 98 per cent completed. The bearing
wall structural system resulted in approximately a 10 percent savings of the cost over a structural steel frame
alternate. Included in this savings of approximately $420,000 was $69,000 for fireproofing of the steel frame.
Tenants occupied some of the buildings seven months after the start of construction. The construction
superintendent of the project, Mr. A. M. DiFerio, attributed this low cost and speed of construction to a number of
reasons, including the simplicity of construction, the use of fewer cranes and other heavy construction equipment,
and a simple spread footing, which was used in lieu of caissons.
http://www.gobrick.com/BIA/technotes/t24f.htm
2 of 11 9/13/2009 1:15 PM
Pennley Park North, Pittsburgh, Pennsylvania
Tasso G. Katselas, Architect; R.M. Gensert, Structural Engineer
FIG. 1
Penn Plaza. This project, quite similar to the above, consists of six buildings, (see Fig. 2). The masonry contractor
for this project, Charles L. Cost, stated that his men "enclosed a floor every two and a half days and the masons
worked fast and efficiently. They were very enthusiastic about this construction. We went on the job with very little
advance notice, and with more planning time we could have finished the masonry work in seventy-four working days,
rather than ninety-four working days actually required." The cost per square foot of the Penn Plaza project was
$13.15.
Penn Plaza, Pittsburgh, Pennsylvania
Tasso G. Katselas, Architect; R.M. Gensert, Structural Engineer
FIG. 2
Oakcrest Towers. Oakcrest Towers is a series of 14 apartment buildings, eight stories in height, being constructed
on a 50-acre site outside of Washington, D. C., (Fig. 3). Each building contains a total of 161,334 sq ft. In good
weather, the contractors report construction of one story per week. Mr. Stanley Reed, Vice-President of L. F.
Jennings Inc., the masonry contractor for Oakcrest Towers, said that, in their experience, bearing wall buildings
were built faster than structural frame buildings and, typically, the time required to complete a bearing wall building
and have it ready for occupancy was approximately equal to the time required to erect the structural frame and
enclose the frame with masonry walls for a project of similar scope. Mr. Reed reported that savings on the Oakcrest
buildings resulted from lower initial cost for bearing wall than for structural frame construction, and elimination of
http://www.gobrick.com/BIA/technotes/t24f.htm
3 of 11 9/13/2009 1:15 PM
finishing and painting of interior exposed brick walls.
Oakcrest Towers, Prince Georges County, Maryland
Bucher-Meyers & Associates, Architects; Keller & Marchigiani, Structural Engineers
FIG. 3
Some bearing wall thicknesses in the first buildings were in excess of 20 in. in thickness. Buildings built after the
SCPI (BIA) standard was introduced had thicknesses reduced to 8 and 6 in. of brick masonry for exterior and
corridor walls, respectively.
Park Lane Towers. Park Lane Towers is a group of 20-story structures located in Denver, Colorado, (Fig. 4). Each
building is 206 ft high containing 38 one-bedroom, 73 two-bedroom, 4 three-bedroom units and one penthouse.
Fireplaces are featured in the two and three-bedroom units beginning at the 15th floor.
Ground breaking for Park Lane Towers No. 1 began in May of 1969. The masonry work was started in the fall of
1969 and progressed through the winter. By March 1, the masonry construction was at the 16th story. By May of
1970, the first building was topped out. BIA recommendations for cold weather construction were followed.
Total time sequence per floor averaged six to eight days, depending upon weather and other non-controllable
factors. An average of three and one-half days was required to construct all walls per story. Twin tee floor slab
erection averaged one day per floor.
The unit cost of Park Lane Towers, based on an area on 156,280 sq ft. excluding the area of the basement, was $
14.71 per sq ft. The total cost of the building was $2,300,000, including drapes, carpets, kitchens, parking garage,
basement and landscaping (excluding land and professional fees).
Park Lane Towers, Denver, Colorado
Joseph T. Wilson, AIA, Architect; Sallada & Hanson, Structural Engineers
FIG. 4
Woodlake Towers. Woodlake Towers is a group of ten-story, T-shaped buildings, each containing 215 units, which
consist of efficiencies, one, two and three-bedroom units, (Fig. 5).
http://www.gobrick.com/BIA/technotes/t24f.htm
4 of 11 9/13/2009 1:15 PM
Bearing walls consist of 8-in., single wythe, solid brick supporting a precast concrete floor system. The first building
was completed in 1970 and contains 268,450 sq ft at a cost of $2,700,000 or $ 10.06 per sq ft (excluding land,
interim financing, professional fees, builder's fee and entrance lobby finishing). The cost figure of $2,700,000
includes all general construction, electrical and mechanical work, parking lot paving, landscaping, furnished kitchens
and corridor carpeting.
Woodlake Towers, Fairfax County, Virginia
Collins & Kronstadt, Leahy, Hogan, Collins, Architects; Keller & Marchigiani, Structural Engineers
FIG 5
Twin Tower Apartments. Twin Tower Apartments consist of two identical structures of 11 stories in height for
elderly residents in Jacksonville, Florida, consisting of a total of 120 efficiency and 80 one-bedroom units, (Fig. 6).
The 10-in. bearing walls were grouted cavity walls (4-2-4) with varying amounts of grout and reinforcement,
depending on loading and location. A 6-in. cast-in-place concrete slab provided the floor system. The masons
rotated between buildings so the walls could be built on one building while the floor was formed and poured on the
other building.
Twin Tower Apartments, Jacksonville, Florida
Sheetz & Bradfield, Architects; Bennett & Pless, Structural Engineers
FIG. 6
Episcopal House of Reading. It is 15 stories tall and contains 141 units for middle income, elderly residents in
Reading, Pennsylvania, (Fig. 7). All the bearing walls are 8 in. thick, single wythe, non-reinforced, solid brick
masonry with the exception of some walls on the first floor level which are 12 in. in thickness.
http://www.gobrick.com/BIA/technotes/t24f.htm
5 of 11 9/13/2009 1:15 PM
Episcopal House of Reading, Reading, Pennsylvania
Muhlenberg & Greene, Architects; Long & Tann, Structural Engineers
FIG. 7
The floor system consists of two layers of concrete. The first is 2 1/4-in. precast concrete containing positive steel
and the second is 5 3/4-in. poured concrete with polystyrene foam sections to create internal voids.
The basic construction process consisted of building brick masonry walls and placing concrete floors in a "leapfrog"
fashion. While a crew of 8 to 12 masons, using corner poles, erected walls on one side of the building, the 2 1/4 in.
concrete floor slabs were placed by tower crane on the other side and were positioned and shored as required.
Construction continued throughout the winter. All materials were stored off the ground and covered for protection
from the elements. A mortar mixing station was constructed adjacent to the structure, complete with propane gas
equipment to heat sand and water. A custom-made sand measuring device permitted rapid and accurate
proportioning by volume of portland cement-lime and sand mortar. The quantity of sand held by this device was 9 cu
ft. Figure 8 shows this simple but workable setup for mixing mortar.
Measuring Sand Automatically.
FIG. 8
TECHNIQUES
Sequence and Scheduling. Most Contemporary Bearing Wall buildings can be planned so that the floor erection
crew will place the floor in about the same length of time as is required for the masons to build the walls. When the
crews are balanced in this manner, it is usually possible to schedule the construction so that neither the floor crew
nor the wall crew will be in the way of the other. An example is shown in Fig. 9 where a floor system of steel joists
http://www.gobrick.com/BIA/technotes/t24f.htm
6 of 11 9/13/2009 1:15 PM
with concrete topping is used. The plan of the Oakcrest project is tee-shaped with a short leg. The masons are
completing the walls on the leg of the building and the area immediately next to it. In the background, concrete is
being poured; in the center, steel decking is being placed; and in the foreground, steel joists are being set and
spaced. In this manner, the contractors built a story per week in good weather.
Floor and Wall Construction of Oakcrest Towers.
FIG. 9
Foundations. Studies have indicated that, in many cases, foundation costs are reduced by delivering the loads into
the soil in a series of lines or paths which might utilize spread footings, rather than in a series of points, which may
require piles or caissons. A case in point is the Oakcrest project in which spread footings for bearing walls resulted
in a savings in cost and time, (Fig. 10).
Corridor Bearing Walls at Oakcrest Towers.
FIG. 10
Formwork. In the construction of Contemporary Bearing Wall structures with joists or plank floor systems, there is
little or no need for formwork, thus permitting the other trades to work immediately below the level of the wall and
floor crews. In the case of a cast-in-place concrete building, the formwork is often such that other trades cannot
conveniently work for several stories below the level where concrete is being placed, (Fig. 11).
http://www.gobrick.com/BIA/technotes/t24f.htm
7 of 11 9/13/2009 1:15 PM
Formwork for Cast-in-Place Concrete Frame.
FIG. 11
Coordination. The space immediately below a floor system of plank or joist is available; therefore, mechanical and
other trades can closely follow the wall and floor crews, so as to contribute to the overall speed of construction,
(Fig. 12).
Working Space is Provided for Other Trades at Oakcrest Project.
FIG. 12
Scaffolding. The scaffolding requirements are simplified in bearing wall buildings, usually requiring only one lift per
story, (Fig. 13). Scaffolding used for one and two-story construction is often applicable to high-rise loadbearing
construction when the masons work overhand from the inside. All of the examples cited above were built in this
manner.
Scaffolding is Simple for Bearing Wall Construction.
FIG. 13
"SCR masonry process" (Reg. U.S. Pat. Off., SCPI (BIA)). The "SCR masonry process'' is a development of the
Structural Clay Products Research Foundation (now a part of BIA). It consists of a continuously adjustable scaffold
and corner poles which carry the lines. This provides a means whereby skilled masons can increase productivity and
enhance the quality of their work, resulting in better masonry at lower cost. For the construction of a series of
three-story barracks, at Ft. Gordon, Gal, bearing walls were selected under "Contractor's Option" in lieu of a
reinforced concrete frame with masonry curtain walls. The masonry contractor on the project, Phiffer and Goodwin,
stated that, four weeks after installing the "SCR masonry process", their records indicated a 37 percent increase in
overall production of the 75 masons working the project, (Fig. 14).
http://www.gobrick.com/BIA/technotes/t24f.htm
8 of 11 9/13/2009 1:15 PM
Use of "SCR masonry process" (Reg. U.S. Pat. Off., SCPI (BIA)), For Gordon, Georgia.
FIG. 14
Prefabricated Elements. Prefabricated stairs, door and window combinations and other elements which can easily
be set in place prior to the masonry work will eliminate cutting and fitting after the masonry is complete. In many
cases, the prefabricated elements can become guides for the masonry. In the case of the stairs, illustrated in Fig.
15, the landings become bench marks for the floors and the stairs themselves provide continuous access during
construction.
Prefabricated Metal Stairs Used in Bearing Wall Construction.
FIG. 15
Keeping Walls Clean. Frequently it will be desirable to place concrete for various members supported on brick
walls. In such cases, care should be exercised to prevent the dropping of concrete on brickwork to be left exposed.
This can be prevented by covering the walls with polyethylene sheets under formwork to make them watertight, (Fig.
16).
Polyethylene Sheets Under Corridor Formwork in Construction of Oakcrest Towers.
FIG. 16
http://www.gobrick.com/BIA/technotes/t24f.htm
9 of 11 9/13/2009 1:15 PM
Building in Accessories. When cavity walls are used, conduit and other accessories are often built in place as the
walls are constructed. Also, in multi-wythe walls, this is easily accomplished as the walls are built. In many cases,
through-the-wall units are designed so that they can be slipped over or conveniently built around conduit and other
such elements, (Figs. 17 and 18).
Building in Conduit and Sleeves at Park Lane Towers.
FIG. 17
Conduit in 4-in. Wall at Fort Gordon, Georgia.
FIG. 18
Lintels. Frequently, in the construction of brick masonry buildings, it is convenient to build horizontal elements on
shoring, with reinforcing steel to span openings, (Fig. 19). Reinforced brick lintels have several advantages, including
built-in fire resistance and elimination of structural steel, which requires maintenance and is more expensive in many
cases. For further information on reinforced brick lintels, see Technical Notes 17H.
Construction of Reinforced Brick Lintel on Shoring.
FIG. 19
http://www.gobrick.com/BIA/technotes/t24f.htm
10 of 11 9/13/2009 1:15 PM
Cold Weather. Complete enclosure of bearing wall structures is not always the most economical way of providing
protection for winter construction. In severe climates contractors may enclose the particular walls being worked
upon.
Figure 20 shows Jayhawker Towers, a dormitory at the University of Kansas in Lawrence. Kansas, where a
prefabricated wood-frame assembly covered with a polyethylene film was lowered over the work area to protect
walls under construction during cold weather.
Temporary Enclosure at Jayhawker Towers, Lawrance, Kansas.
FIG. 20
Where the weather is not severe enough to enclose the construction, heating of materials and proper covering of
walls may be all that is necessary. In all cold weather construction, the masonry should be constructed so strength
will develop and the mortar will lose sufficient water to prevent expansion upon freezing.
For further discussion of winter construction, see Technical Notes 1 Revised, "Cold Weather Masonry Construction-
Introduction", December 1967.
http://www.gobrick.com/BIA/technotes/t24f.htm
11 of 11 9/13/2009 1:15 PM

Technical Notes 24G - The Contemporary Bearing Wall - Detailing
[Dec. 1968] (Reissued Feb. 1987)
INTRODUCTION

The selection of a wall type and appropriate connection details is one of the most important decisions to be made in
the design of a bearing wall building. In most cases, the primary consideration will be a system which satisfies the
structural requirements of the building. Other considerations are also of importance, including the properties and
performance of the walls and floors that will result in an economical, maintenance-free and easy-to-construct
building.
Bearing walls offer the designer the opportunity to develop a complete building system in which the floors and walls
not only carry the vertical and lateral loads, but also provide separation, thermal and acoustical control, and
fire-resistive and low maintenance construction.

MATERIALS

Materials used in constructing the walls for bearing wall buildings will have considerable influence on the satisfactory
performance of the structure. In most cases, the compressive stresses will be relatively high, requiring medium to
high strength masonry units and mortar. Materials selected should comply with the requirements contained in the
standard, Building Code Requirements for Engineered Brick Masonry, BIA, 1969.
Brick. A large percentage of the brick produced in the United States have unit compressive strengths in excess of
5000 psi, with some having compressive strengths as high as 25,000 psi. Units are readily available throughout the
U.S. and Canada which have strengths of 10,000 psi and above. Esthetics, durability and economy should also be
considered in the selection of a particular type of brick unit for a given bearing wall building. The two sizes of brick
most widely used are often referred to as Standard (actual dimensions - 3-3/4 by 2-1/4 by 8 in.) and Standard
Modular (actual dimensions for 3/8-in. joint 3-5/8 by 2-1/4 by 7-5/8 in.). There are many different sizes of brick made
to suit local conditions. In some areas of the U.S., 3-in. bed depth brick are widely used. The sizes vary with the
manufacturer and locality. The size that is most widely made and distributed is called King size, with actual bed
dimensions of 3 by 9-5/8 in. and height of 2-5/8 in. (laying 4 courses to 12 in.) or 2-3/4 in. (laying 5 courses to 16
in.). Walls built with larger brick units usually provide greater economy than walls of smaller units. Table 1 lists a
number of modular brick sizes. These are illustrated in Fig. 1. The local availability and strength of specific units
should be ascertained prior to the design.

24g http://www.gobrick.com/BIA/technotes/t24g.htm
1 of 7 9/13/2009 1:16 PM
1
Also called Norman Economy, General and King Norman.
2
Reg. U.S. Pat. Off., SCPI.



Typical Modular Brick
24g http://www.gobrick.com/BIA/technotes/t24g.htm
2 of 7 9/13/2009 1:16 PM
FIG. 1


Mortar. There is no one mortar that is best for all purposes. Masonry mortar should be selected for the particular
use. ASTM Specifications C 270, "Mortar for Unit Masonry", are recommended as a basis for mortar specifications.
It is further recommended that the Proportion Specification be used rather than the Property Specification and that
mortar be constituted of portland cement (type I, II or III), type S hydrated lime (non-air-entrained), or lime putty and
sand.

Building Code Requirements for Engineered Brick Masonry recognizes three mortar types: Types M, S and N. In
general, type S mortar provides high bond and compressive strength with long durability and good workability. Type
N is a mortar with good strength properties, durability and excellent workability. In most cases, either type S or N
mortar will provide the desirable properties for loadbearing buildings. When high compressive stresses control the
design, type M mortar may be required. Type S mortar is proportioned 1C:1/2 L:4 - 1/2 S, type N mortar is
1C:1L:6S and type M is 1C: 1/4 L:3S. See Technical Notes 8 Revised, Portland Cement-Lime Mortars for Brick
Masonry".
Metal Ties and Joint Reinforcement. In years past, structural bonding of masonry walls was generally
accomplished by overlapping masonry units. In so doing, various visual patterns were created; two traditional ones
being English and Flemish bond.
In the construction of the thinner masonry walls of today, when these patterns are not required by esthetic
considerations, it is recommended that metal-tie bonding be used for multi-wythe solid walls in lieu of masonry
bonding. Laboratory tests at the Structural Clay Products Institute and independent laboratories indicate that
metal-tied walls are comparable to masonry-bonded walls in resisting compressive and transverse loads, as well as
water transmission, when the collar joint is full of mortar. In addition, metal-tied walls generally cost slightly less than
masonry-bonded walls. They also provide greater freedom in construction and pattern selection.
Metal ties used in bonding masonry walls should be equal to those required for cavity walls; i.e., 3/16-in. diameter
steel or metal ties of equivalent strength spaced so that at least one tie is provided for every 4-1/2 sq ft of wall area.
Distance between ties should not be over 24 in. vertically and 36 in. horizontally. Metal ties should be of corrosion-
resistant metal or protected with a corrosion-resistant metal. Ties may be shaped to form a Z, rectangle or
continuous tie of either the ladder or truss type. In cavity wall construction, the Z and rectangular ties, which are
often used, have depressions or drips in the middle of the ties. The drip is not necessary and reduces the strength of
the tie; however, most of the ties on the market have sufficient strength with or without the drips. There are some
advantages to the continuous ladder or truss type joint reinforcement, particularly where wythes are made of units
with different physical properties or in wall areas subject to stress concentration.
If stack bond masonry is used as a bearing wall, continuous joint reinforcement should be provided, spaced not more
than 16 in. on center vertically, one 9-gage or larger longitudinal wire for each 6 in. of wall thickness or portion
thereof.

CONSTRUCTION DETAILS

Even though the masonry details must be handled somewhat differently in bearing wall buildings than in structural
frame buildings, the same major considerations are prevalent:
1. Structural requirements of the buildings must be met. Anchorage must be provided which will transmit stress
where desired. Equally important are anchorage details which will not transmit unwanted forces.
24g http://www.gobrick.com/BIA/technotes/t24g.htm
3 of 7 9/13/2009 1:16 PM
2. Since the elements of the building are in a constant state of motion, such movement must be accommodated or
distress will occur.
3. The opportunity for uncontrolled entrance or condensation of water must be avoided. The problem of keeping
water out of the bearing wall building may be solved by using cavity or grouted walls. Where solid walls are used
with or without furring, pargeting, waterproofing and systems of flashing and drainage can be provided. See
Technical Notes 7A Revised, Water Resistance of Brick Masonry-Materials", and 7B Revised, "Water Resistance of
Brick Masonry-Construction and Workmanship".

Structural Requirements. Because of their structural simplicity, it is possible to design loadbearing clay masonry
buildings quite easily, using accepted theories. In some buildings, the floor plan will be such that the longitudinal
exterior and corridor walls may be loadbearing. In others, it may be more advantageous to have transverse bearing
walls. In both cases, horizontal forces are usually resisted by shear wall action. Since the transverse strength of
non-reinforced masonry walls is relatively low, structural positioning of building parts are so arranged as to minimize
this action and to exploit the compressive and shear resistance of the walls.

By utilizing the floors and roof as horizontal diaphragms and the shear walls as vertical diaphragms, the horizontal
forces are carried down into the foundation. Shearing stresses in bearing wall buildings will seldom control the wall
type and thickness. Although flexural stresses in shear walls may control the design under certain conditions, it is the
bearing stress that will generally govern. For additional information pertaining to structural design, see Technical
Notes 24 Series, "The Contemporary Bearing Wall".
Types of walls and floors used with bearing wall construction will depend on a number of factors, including
economics, climate, building type, height and arrangement of walls, and the structural requirements. In most cases,
clay masonry walls from 4 to 12 in. in thickness will satisfy the structural requirements.
In developing construction details for bearing wall buildings, it is important to understand the potential condition of
stress at junctures of horizontal and vertical elements. At connections, as a result of vertical and horizontal loads,
compression, tension and shear may occur. Seldom is it desirable to transmit bending moments at junctures of floors
and walls unless the walls are of reinforced masonry. It is relatively easy to develop details which will transmit
tension and compression, as well as shear. Often, however, it is desired to transmit tension and compression alone,
or shear alone. At times, it is desirable to take loads horizontally (produced by wind) but not vertically (due to
gravity). For these reasons, it is important for the structural designer to take an active part in developing the
construction details for a particular project.
Where concentrated loads are imposed upon masonry walls, it is recommended that bearing pads be used to
distribute the stress and to permit slight movement which may occur. Suitable materials for this use include 55-lb
roofing felt, pads of Neoprene, tempered Masonite or vinyl floor tile. It is advisable to design the connection so as to
position the resultant of the bearing stress in the middle two-thirds of the wall. If the vertical load develops an
eccentricity which falls outside of the middle two-thirds of the wall, the possibility of excessive tensile bending
stresses developing in the wall must be investigated. See detailed design requirements contained in the Building
Code Requirements for Engineered Brick Masonry.
Small amounts of reinforcing steel may be grouted in brick walls to provide extra strength. The designer may find it
helpful to use steel in piers where loads are high. Bond beams are often desirable to distribute loads at floor levels.
For information on the design and construction of reinforced brick masonry (RBM), see Technical Notes 17 series,
"Reinforced Brick Masonry".
Rigid anchorage of concrete floors to the masonry walls is seldom desired or needed. The friction will often meet the
structural requirements. Coefficients of friction between various materials in a dry state will vary. Standard
references indicate a range of values given in Table 2. In this table, masonry may be considered to be clay, stone or
concrete.

24g http://www.gobrick.com/BIA/technotes/t24g.htm
4 of 7 9/13/2009 1:16 PM


Differential Movement. All elements and materials which go into the makeup of the building are in a constant state
of motion. All building materials move with changes in temperature, some move with changes in moisture content,
some have plastic flow due to stress, and all have elastic deformation due to imposed loads. Provisions must be
made to permit the various materials and elements to move so as not to distress any part. See Technical Notes, 18
series, "Differential Movement".

The stress developed in restrained elements due to a change in temperature is equal to the modulus of elasticity
multiplied by the coefficient of expansion and by the change in temperature. Table 3 lists coefficients of lineal thermal
expansion for many commonly used building materials.

24g http://www.gobrick.com/BIA/technotes/t24g.htm
5 of 7 9/13/2009 1:16 PM

Wood, masonry and concrete may expand with changes in moisture content. For concrete and wood products, these
movements are reversible. For clay products masonry, in which the initial moisture expansion is not reversible at
atmospheric temperatures and pressures, a moisture expansion design coefficient of 0.0002 is recommended for
clay masonry. A shrinkage coefficient of 0.0005 is recommended for concrete.
Elastic deformation due to stress in the elastic range is approximately linear and is equal to the stress divided by the
modulus of elasticity of the material.
When some materials are continuously stressed, there is a gradual yielding in the direction of the stress application.
This is plastic flow, which is influenced not only by stress, but also by time and the physical properties of the
material. A design value of 0.000001 per unit of length per psi is recommended for concrete. For example, a 100-ft
long member, stressed to 1000 psi, would be expected to have a reduction in the length of 1.2 in.
In brick masonry, the units themselves are not subject to flow, although the mortar joints are. The joints, however,
24g http://www.gobrick.com/BIA/technotes/t24g.htm
6 of 7 9/13/2009 1:16 PM
seldom comprise more than 15 or 20 per cent of the volume in compression. Limited research on creep for brick
masonry suggests a design value of 0.0000002 per unit of length per psi.
Another major cause of movement in buildings is settlement and the action of unstable soil. Frequently, bearing wall
buildings can be utilized to an advantage because the loads are delivered in lines rather than points, thus keeping the
soil bearing stress low. Also, the walls can be designed as thin, deep beams which, when working with the footing,
will be able to span weak spots or depressions in the subsoil.
The spacing of expansion joints in bearing walls need not be as close as for non-bearing walls because bearing walls
are usually of greater strength and mass. Also, they are under higher stress, which in many cases will tend to
restrain the walls, thus reducing movement. Where expansion joints are required, it is important to be careful to
maintain the structural integrity of the wall.
Floor systems in bearing wall buildings will necessarily be influenced by structural requirements and by arrangements
of bearing and shear walls. Joist systems and precast plank systems are satisfactory for most buildings, since they
eliminate form work. Occasionally, two-way slabs, where the vertical load can be distributed in two directions, will
offer a better floor system.
Control of Moisture and Temperature. In many buildings, cavity walls will meet the structural requirements, as
well as provide walls which are resistant to water penetration, because of their internal drainage channels. Similar
drainage channels can be built into walls which are furred on the inside, if through-the-wall flashing is used to collect
any water which might penetrate the wall and divert it back to the outside through weep holes. These air spaces
also provide convenient spaces for thermal insulation. Two types of pouring-type wall insulation have been
investigated and found to be suitable for cavity walls. These are water-repellent vermiculite loose-fill insulation and
water-repellent perlite loose-fill insulation. See Technical Notes 21 Series. Cavity walls can often be used as interior
bearing walls to accommodate mechanical equipment which may run horizontally and vertically in the walls.
On exterior walls, care should be taken to avoid thermal bridges on which condensation may occur. In most
climates, condensation will not occur on uninsulated masonry walls; however, through-the-wall metal elements may
collect condensation on the inside. See Technical Notes 7C, "Moisture Control in Brick and Tile Walls -
Condensation", and 7D, "Moisture Control in Brick and Tile Walls-Condensation Analysis".
In bearing wall buildings, some of the detailing problems are more critical, while others are minimized. There are few
absolute answers to these problems. The best solution will stem from a rational analysis of the situation. Good
details require ingenuity and imagination. There are no standard details which should be used without question. Each
juncture of floor and wall should be considered as a separate problem and an appropriate detail developed for the
situation.
24g http://www.gobrick.com/BIA/technotes/t24g.htm
7 of 7 9/13/2009 1:16 PM

Technical Notes 26 - Single Wythe Bearing Walls
Sept. 1994
Abstract: Brick masonry bearing wall systems have been used for years for their strength, durability
and other inherent values. Once widely used in single family residential construction, this application
is experiencing a resurgence in interest. New designs possible with a single wythe of brick are
discussed in this Technical Notes. Selection of materials and recommended details for one and two
story designs are addressed.
Key Words: bearing wall, brick, reinforced brick masonry, single wythe wall.
INTRODUCTION
The rising cost of wood framing members has created a renewed interest in alternative building
systems for residential housing. The use of light-gage steel framing is one alternative. Another is the
use of single wythe brick bearing walls. The use of brick masonry as the load-carrying element of a
structure provides several benefits over other alternate systems. Using brick as both the building's
exterior skin and its structure capitalizes on brick masonry's strength and other inherent values. Brick
gives a home permanence and beauty. Brick homes have lower maintenance costs and often lower
insurance rates because of their fire resistant characteristics. Because of their thermal mass
properties, brick homes are more energy efficient than comparably insulated vinyl- or wood-sided
homes. For these reasons, brick homes have a higher resale value.
In a single wythe brick bearing wall system, the brick masonry serves as both the structural system
and the exterior facing. A wood, steel or masonry backing system is not necessary. The interior living
space of a brick bearing wall home may be the same as that of a framed home. Floor and roof
elements and interior partitions are constructed with the same materials as used in frame homes.
Home plans may include one or two story structures, expansive master bedrooms and other popular
amenities. In brick bearing wall homes, attractive features such as brick masonry fireplaces and
special brick details can be readily incorporated to set the house apart.
The design and construction of single wythe brick bearing wall systems are discussed in this
Technical Notes . Typical details for residential applications are provided. Although this Technical
Notes illustrates residential construction, single wythe brick bearing walls are equally appropriate in
commercial construction, and many of the design and material considerations discussed in this
Technical Notes are the same.
DESIGN CONSIDERATIONS
Single wythe brick loadbearing walls include the same design considerations as other types of wall
systems. Model building codes dictate the minimum loads to be resisted by the structural system, the
minimum thermal performance requirements and the necessary fire resistance of the wall system.
Since a brick bearing wall system forms the building envelope, the designer must also consider
resistance to moisture penetration and detailing of interior finishes. These concerns are addressed in
the sections that follow.
Structural Considerations
The design of any structural system begins with the determination of the design loads. Design loads
include vertical loads from the weight of the building materials and occupants and lateral (horizontal)
loads from wind, soil and seismic forces. In most residential structures, the controlling forces are the
expected vertical loads plus wind loads. In certain regions of the United States, predominantly the
t26 http://www.gobrick.com/BIA/technotes/t26.htm
1 of 16 9/13/2009 1:17 PM
west coast, seismic loads may control the structural design.
The model building codes and the associated structural loads will dictate the size of the building's
structural members. Model building codes have two methods of design: empirical and rational. In the
case of loadbearing masonry, all model building codes specify the minimum wall thicknesses and
maximum wall height or number of stories for empirical designs. Generally, these limitations are not
applicable to buildings which have been rationally designed. However, the distinction in model building
code limitations for engineered (rational) versus empirically designed masonry structures is not
always clear. Furthermore, even a rational design will include some prescriptive detailing
requirements. The masonry standard used for the structural design of brick bearing walls in this
Technical Notes is the Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS
402) and its companion Specifications for Masonry Structures (ACI 530.1/ASCE 6/TMS 602) [4],
referred to as the MSJC Code and Specifications, respectively. Other masonry design criteria, such
as Chapter 21 of the Uniform Building Code, could also be used.
The most widely used residential building code is the CABO One- and Two-Family Dwelling Code
[5]. The CABO Code is an empirical code which specifies an 8 in. (200 mm) minimum nominal wall
thickness for loadbearing masonry structures over one story in height. For single-story homes, a
nominal 6 in. (150 mm) wall is permitted, so long as the wall height does not exceed 9 ft (2.7 m) and
the gable height does not exceed 15 ft (4.6 m). In cases where these limits are applied, the minimum
wall thickness requirements will influence the type and size of brick unit used. When rationally
designed, these limits on wall size can be exceeded. Other model building codes have similar
prescriptive restrictions, but permit rational design, not subject to these limits, when using approved
masonry standards.
Loadbearing brick masonry houses may be built with walls less than 6 in. (150 mm) in nominal
thickness when rationally designed. These walls may require vertical steel reinforcing bars and
horizontal reinforcing bars or wires. Typically, vertical steel reinforcing bars are used to resist lateral
loads and horizontally reinforced bond beams are used to attach floor and roof members. Additional
horizontal reinforcement is required over large openings and in areas of high seismicity. This
Technical Notes provides details for location of vertical reinforcement and horizontal bond beams
when used in brick bearing wall construction.
The designer of a brick bearing wall must specify the properties of the materials necessary to meet
the structural requirements of the design. The compressive strength of the brick masonry assembly,
f'm, must be specified. Mortar and grout type or properties should be identified. Type, size and grade
of reinforcement, if used, should also be specified. See Technical Notes 3 Series for further
information on the structural design of brick bearing walls using the MSJC Code.
Energy Considerations
The model building codes contain requirements to ensure acceptable thermal performance of the
building envelope. Minimum levels of insulation are required, and in some cases, air leakage is
addressed. Type and installation of insulation in a single wythe brick bearing wall system differ from
those in other residential wall systems. In wood frame residential structures, batt insulation is
typically placed between the wood studs. For brick bearing wall homes, rigid board insulation is often
placed on the interior face of the brick wythe. Rigid board insulation has the advantages of being
easy to install and providing high insulation values. Installation of the insulation board on the interior of
the brick wythe is coordinated with the interior finish materials and with the flashing and drainage
system used to control water penetration. Options for attaching rigid insulation include square, "Z" or
other shaped furring strips, mechanical fasteners and adhesives. Alternately, insulation may be
placed in the cells of hollow brick units. However, this application is limited to large hollow units most
commonly used in commercial brick bearing wall buildings. Furthermore, such insulation is generally
not as effective as a continuous layer of insulation placed on the inside face of the single wythe wall,
due to the discontinuity of the insulation at the webs of the units.
Air leakage through the building envelope may also be a concern. Although the brick wall provides an
effective air barrier, there will be some leakage through weep holes and at the top of the brickwork.
Building paper or sheet membrane materials, called "house wraps", are commonly installed over
t26 http://www.gobrick.com/BIA/technotes/t26.htm
2 of 16 9/13/2009 1:17 PM
exterior sheathing materials in wood frame construction to prevent air leakage. However, these
materials are not appropriate for direct application on brick bearing walls. Alternate approaches to
further limit air leakage are the use of either foil-faced rigid board insulation or so-called "air-tight
drywall". These approaches rely on the air penetration resistance of the paper or other films on the
insulation or gypsum board. To achieve an impenetrable air barrier, the joints between the sheets of
insulation or gypsum board should be sealed or taped. Joints between dissimilar materials, and joints
around door and window frames, should also be sealed.
Water Penetration Resistance
Water penetration is one of the chief concerns in the performance of any exterior wall system.
Resistance to water penetration of the brick masonry wythe is of utmost importance in single wythe
construction. Full mortar joints and good extent of bond between units and mortar help to reduce
water penetration. The head joints in hollow brick masonry should be laid full, not face-shell mortar
bedded only. Technical Notes 7 Series provides further information on water penetration resistance
of brick masonry wall systems.
A single brick masonry wythe may not prevent water penetration entirely. Therefore, a drainage
cavity with flashing and weep holes should be provided. If a drainage cavity is not used, a bituminous,
damp-proof coating should be applied to the inside face of the brick bearing wall prior to installation
of the insulation and finishes. Material compatibility of the coating with adjacent materials should be
considered. In regions of the country which are subject to large amounts of rainfall or severe
wind-driven rain, the use of a clear water repellent coating on a wall built with good workmanship and
proper details may be appropriate with this wall type.
Location of Interior Work and Finishes
The installation of plumbing, heating and electrical systems in a loadbearing brick home will vary
slightly from their placement in conventional frame construction. There is no cavity between studs for
the placement of piping or conduit, and piping or conduit should not be placed within the brick wythe.
Instead, the plumbing, heating and electrical piping or conduit can be installed between furring strips
on the brick bearing wall, in the floor or ceiling or in interior frame walls. The type of foundation will
influence the location of interior systems. In slab-on-grade construction, it is easiest to route the
mechanical systems through the ceiling space. With basement or crawl space foundations, it is
possible to locate mechanical systems between the first floor joists.
The interior finish materials used in brick bearing wall homes are the same materials as those used in
frame construction. However, the installation of interior finishes varies. In brick veneer construction,
interior gypsum board is typically nailed or screwed to the studs. In brick bearing wall construction,
gypsum board or other interior finishes may be nailed or screwed to furring strips or light-gage metal
studs, anchored to the masonry with special clips or ties, nailed to special nailing inserts in the
masonry or adhesively attached over the brick masonry wythe and insulation. Some common
methods are discussed in Construction Details.
Cabinets and other built-in items can be attached directly to the brick masonry walls or to furring
strips attached to the masonry. Wood 2 by 4's, spaced to support the cabinets or other built-in items,
can be attached to the brick masonry using anchor bolts placed in the masonry wall or by expansion
anchors, adhesive anchors or other brick anchors and fasteners installed in the completed masonry.
Technical Notes 44 and 44A provide a complete description of suitable brick anchors and fasteners.
Insulation may be placed between the 2 by 4's, and cabinets then placed over insulation and
attached to the furring using conventional techniques. Cabinets may also be supported by stud
framing built inside the brick bearing wall, in a manner similar to brick veneer construction.
MASONRY MATERIAL SELECTION
Selection of masonry materials for a brick bearing wall system should consider structural, energy and
other performance requirements as discussed in this Technical Notes. Further information on
material selection for adequate strength and compliance with the MSJC Code and Specifications can
be found in Technical Notes 3 Series.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
3 of 16 9/13/2009 1:17 PM
Brick
Brick used in single wythe bearing wall structures may be either solid or hollow. Solid units should
meet the requirements of ASTM C 216 Specification for Facing Brick. Hollow units should meet the
requirements of ASTM C 652 Specification for Hollow Brick. Structural and model building code
requirements, aesthetics, availability and cost will determine the minimum unit compressive strength,
type and sizes of units used. Brick unit compressive strengths range from 1,700 psi to 36,000 psi (12
to 250 MPa), with a mean value of over 5,200 psi (36 MPa) for molded brick and over 11,300 psi (77
MPa) for extruded brick [1]. Brick unit strength is directly related to compressive strength of the brick
masonry assembly. Technical Notes 3A should be reviewed for material properties of brick masonry.
There are numerous brick sizes manufactured today. Solid units are commonly manufactured in
nominal widths of 3, 4 and 6 in. (75, 100 and 150 mm). Hollow units, which are less than 75 percent
solid, are manufactured in nominal widths of 4 in. (100 mm), 5 in. (125 mm), 6 in. (150 mm) and 8 in.
(200 mm). See Technical Notes 10B Revised for a more complete listing of sizes currently
manufactured. Nominal 5 and 6 in. (125 and 150 mm) wide hollow brick are the most common units
used to build reinforced brick bearing wall homes.
Hollow brick are prevalent in reinforced brick bearing walls because they have cells which can
accommodate vertical reinforcement and grout. The minimum size of cells in hollow units intended to
be reinforced is dictated by the MSJC Code and listed in Table 1. Larger bars, horizontal reinforcing
bars and coarse grout require larger cell sizes. When reinforcement is required, solid brick may be
cut to form grout pockets or incorporate pilasters to accommodate vertical reinforcement. For further
information on grouting requirements see the MSJC Code and Specifications and Technical Notes
3B.
Uniform spacing of vertical reinforcement is important in the design and construction of reinforced
loadbearing masonry walls. Cell size and unit length should be coordinated to provide cells which
align vertically for ease of grouting and uniform spacing of reinforcing bars. Most hollow brick
designed to accommodate reinforcing bars have lengths equal to twice their width, so that cells align
vertically when the masonry is laid in half running bond. Spacing of reinforcement should be selected
with these dimensions in mind.

1
Metric dimensions are based on metric reinforcing bar sizes as specified in ASTM A 615M.
2
Area of reinforcing bar should not exceed 6 percent of cell void area.
Mortar
The strength and water penetration resistance of a brick bearing wall is dependent upon the mortar
t26 http://www.gobrick.com/BIA/technotes/t26.htm
4 of 16 9/13/2009 1:17 PM
selected. Portland cement-lime mortars with an air content less than 12 percent are recommended
for their superior bond strength and resistance to water penetration. In unreinforced loadbearing
masonry, the MSJC Code allowable flexural tensile stresses are reduced approximately 50 percent
for assemblies made with masonry cement mortars or portland cement-lime mortars with air content
over 12 percent. In addition, some codes prohibit the use of masonry cement mortars and all Type N
mortars in Seismic Performance Categories D and E (formerly Seismic Zones 3 and 4). Mortar
should meet the proportion requirements of ASTM C 270 Specification for Mortar for Unit Masonry.
Type S, M or N mortar may be used in loadbearing brick masonry, although Type S is recommended
for use in reinforced brick bearing walls.
Grout
Grout is used in reinforced brick masonry to bond steel reinforcement to the surrounding brick
masonry. Grout for masonry may be made from either fine or coarse aggregate, although fine grout
is typically used for ease of grouting smaller cells. Aggregate type influences the size of grout space
needed. (See Table 1.) Grout should meet the proportion requirements of ASTM C 476 Specification
for Grout for Unit Masonry, and use of a shrinkage compensating admixture is recommended. The
water/cement ratio of grout is not typically specified, but the water content should be sufficient to
provide a mixture which has a slump of 8 to 11 in. (200 to 275 mm). Grout should be fluid enough to
fill voids, but not separate into its constituents.
Reinforcement
Vertical steel reinforcement is often used in brick bearing walls to provide resistance to lateral loads.
Size and spacing of reinforcement required are a function of design loads, unit size, compressive
strength of the masonry assemblage and cell spacing. The MSJC Code limits the maximum size of
reinforcement used in masonry to a No. 11 reinforcing bar. As a rule-of-thumb, the maximum bar
size should not exceed the nominal thickness of the wall in inches to ensure proper development of
the reinforcement. For example, a maximum reinforcing bar size of No. 6 is recommended for
nominal 6 in. (150mm) walls.
Steel reinforcing bars should conform to ASTM Specification A 615, A 615M, A 616, A 617 or A 706,
depending upon the type of bar used. Joint reinforcement, if used, should comply with ASTM A 82
and be hot-dipped galvanized or made from stainless steel to reduce the possibility of corrosion.
Construction Details
There are several features of a brick bearing wall system which differ from brick veneer wall
systems. The construction of loadbearing brick masonry may incorporate reinforcement and must
provide support and attachment of floors and the roof. Openings for windows and doors may be
spanned by self-supporting reinforced or unreinforced brick masonry. Water penetration resistance
and thermal resistance are generally provided by methods other than the traditional drainage cavity
and insulation between wood studs. Possible details for construction of these features in a brick
bearing wall system follow. The structural details and methods of attaching the insulation and interior
finishes shown vary from figure to figure to illustrate the variety of options available. No single
method is preferred in all cases, nor are all possible options shown.
Bearing Wall Reinforcement
A loadbearing brick wall often contains vertical steel reinforcement uniformly spaced along the length
of the wall and horizontal reinforcement in bond beams. Vertical reinforcement may also be
necessary around openings and at building corners. Figure 1 illustrates one means of incorporating
vertical reinforcing bars in a wall built of solid units. In this case, horizontal joint reinforcement may
be required in bed joints and formwork is necessary to contain the grout in the grout pocket until it
has cured. The brick units will require special cutting to form the grout pocket within the wall
thickness.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
5 of 16 9/13/2009 1:17 PM
Solid Brick With Grout Pocket
Fig. 1
Pilaster Built with Solid Brick
Fig. 2
Reinforced Hollow Brick
Fig. 3
Alternately, brick bearing walls built with solid units may incorporate pilasters, as shown in Fig. 2, to
provide the necessary confinement of vertical reinforcement. The advantage of using pilasters is that
no forms are required. They may, however, occupy a significant amount of floor space if located on
the interior side of the wall.
Masonry walls constructed with hollow units may be vertically reinforced as shown in Fig. 3. By
providing the necessary reinforcement within the cells of the unit, hollow brick bearing walls can
optimize the wall section. The design and detailing of reinforcement should follow the provisions of
the MSJC Code and Specifications. In some cases, such as at terminations or splices of vertical
reinforcing bars, special attention may be necessary to accommodate multiple or hooked reinforcing
bars within the confines of the cells of hollow brick. Technical Notes 3B provides examples of
common reinforced pilasters and reinforced hollow brick walls.
Horizontally reinforced bond beams are used to anchor bolts for attaching ledgers and plates and to
t26 http://www.gobrick.com/BIA/technotes/t26.htm
6 of 16 9/13/2009 1:17 PM
span wall openings. Bond beams are formed by removing part of the cross webs of hollow brick or
by using special U-shaped units. Necessary anchor bolts and reinforcement are placed, and the
bond beam is grouted solid. The depth of the bond beam required will depend on the design loads
for the structure, the material properties of the masonry and the amount of reinforcement used.
Connections
Foundation. Brick bearing walls may be supported on poured concrete, concrete masonry or brick
masonry foundation walls. If construction incorporates a slab on grade, the foundation wall may be
built as shown in Fig. 4.
Slab-On-Grade Foundation
Fig. 4
When a crawl space or basement is present, the floor joist system may be supported directly on the
foundation wall (Fig. 5),
Basement or Crawl Space Foundation
Fig. 5
on corbeled brickwork (Fig. 6) or on a ledger joist bolted onto a bond beam (Fig. 7). The details of
support will vary depending upon the size of foundation wall and the width of the brick bearing wall
above the foundation. A minimum bearing of 3 in. (75 mm) should be provided for floor joists which
bear on the foundation wall. Waterproofing should be provided for below-grade masonry.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
7 of 16 9/13/2009 1:17 PM
Basement or Crawl Space Foundation
Corbeled Support
Fig. 6
Basement or Crawl Sapce Foundation Bond Beam
Support
Fig. 7
Floors. Floors above the first story may be anchored to the brick bearing walls or be supported on
corbeled brickwork. In multi-story construction, anchor bolts cast into a continuous, reinforced,
grouted brick bond beam at the floor support level are often used. A continuous wood ledger is
bolted into place, and the floor joists are attached to the ledger with joist hangers as shown in Fig.
8. Flashing should extend a minimum of 8 in. (200 mm) above the ledger and at least 3 in. (75 mm)
below.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
8 of 16 9/13/2009 1:17 PM
Slab-On-Grade Foundation
Fig. 8
Roof. The roof should be supported directly on top of the bearing wall to minimize eccentric loading.
The roof must be anchored on top of the brick bearing wall to resist uplift forces on the roof. A wood
plate is attached to the top of the wall using anchor bolts embedded in a bond beam or masonry
below. A reinforced concrete bond beam or a reinforced and grouted brick bond beam may be
used, as illustrated in Fig. 9. In this case, the anchor bolts should extend a minimum of 12 in. (300
mm) into the grouted cells in the wall below and terminate with a standard hook. One alternative, for
use in unreinforced bearing walls, is to thread anchor bolts through the core holes of the solid units
and attach the bolts to a steel plate embedded in the masonry, as shown in Fig. 10.
Slab-On-Grade Foundation
Fig. 9
Window and Door Openings
Masonry over openings for windows and doors may be supported by loose steel lintels, reinforced
brick masonry lintels or brick masonry arches. The design of steel lintels is covered by Technical
Notes 31B. When steel lintels are used, flashing and weep holes should be provided over the lintel
as shown in Fig. 11. Alternately, loadbearing brick masonry can be self-supporting over many wall
opening. Horizontally reinforced brick masonry lintels are one option. Design of reinforced brick
lintels should be in accordance with the MSJC Code or other approved masonry standard.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
9 of 16 9/13/2009 1:17 PM
Roof Connection
Fig. 10
Window Detail With Steel Lintel
Fig. 11
Brick Bond Beam Lintel
Fig. 12b
t26 http://www.gobrick.com/BIA/technotes/t26.htm
10 of 16 9/13/2009 1:17 PM
Soldier Course Lintel
Fig. 12a
Reinforcement may be incorporated into voids in a soldier course of brick or in a bond beam, as
illustrated in Figs. 12a and 12b, respectively. Figure 13 provides required steel reinforcement for
solidly grouted bond beam lintels base on the MSJC Code. Required reinforcement is determined by
the total uniform dead and live loads on the lintel and the span of the opening. Figure 13 is applicable
to hollow or solid units of 5 to 6 in. (125 to 150 mm) in nominal width. Lintels made of hollow units
are assumed to be grouted over their entire height, ht. Generally, for loads greater than 400 lb/ft (6
kN/m), the maximum span length is limited by shear strength of the lintel. Spans may be increased if
shear reinforcement is provided in the lintel. Reinforced brick lintels should be shored for a minimum
of seven day after construction. Openings can also be spanned by reinforced or unreinforced brick
masonry arches. Consult the Technical Notes 31 Series for guidance on the structural design of
brick masonry arches.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
11 of 16 9/13/2009 1:17 PM
Reinforcement for Brick Bond Beam Lintels
Fig. 13
Flashing and Weep Holes
For best performance, a drainage cavity with flashing and weep holes should be incorporated in the
single wythe bearing wall system. The most common method to form a drainage cavity behind the
brick wythe is to space the rigid board insulation a minimum of 1/2 in. (13 mm) from the inside face of
the brick wythe using furring strips or other spacers. The interior finish is then attached directly over
the insulation. Any materials in direct contact with the brick wythe should be rot and corrosion
resistant for long-term durability. Alternately, rigid board insulation can be adhered to the brick
wythe, and light-gage (non-bearing) metal stud framing can be attached to floor and ceiling joists with
no direct contact with the brick wythe. This forms a cavity between the insulation and interior finish
which can be flashed similar to that in brick veneer wall systems. A minimum 1/2 in. (13 mm) cavity
is adequate for this wall system because the insulation and finishes are installed after the masonry is
completed. Any voids in the mortar should be filled, and any mortar protrusions should be removed
to provide a clear, open cavity. Flashing should be placed at the base of the cavity and at all
t26 http://www.gobrick.com/BIA/technotes/t26.htm
12 of 16 9/13/2009 1:17 PM
interruptions in the wall, such as over window and door openings. The effect of flashing placement
should be considered in the structural design. Splices in flashing must be sealed and discontinuous
flashing must have end dams.
Flashing should be turned up a minimum of 8 in. (200 mm) and attached with adhesive to the inside
surface of the rigid board insulation or outside surface of the gypsum board, or nailed or stapled to
furring strips. The flashing should extend past or be cut flush with the exterior face of the brickwork.
If the brick bearing wall is reinforced, the through-wall flashing at the base of the wall will be
punctured by vertical reinforcing bars. Flashing must be sealed around all reinforcement with mastic
at these locations. Suitable flashing materials are discussed in Technical Notes 7A Revised.
Weep holes should be placed directly above all flashing locations. Weep holes should be located
above grade and spaced a maximum of 24 in. (600 mm) on center when using open head joints or
brick vents, or 16 in. (400 mm) on center when using wicks or plastic tubes. Open head joint weep
holes are preferred over rope wicks or tubes. Vents, copper screening or stainless steel wool can
be placed in open head joint weep holes to prevent insects and rain from entering. If the brick
bearing wall is treated with a clear water repellent, brick vents in head joints at the base and top of
each story are recommended.
Insulation and Finishes
The attachment of insulation and interior finishes may be accomplished in several ways. One method
is to attach treated wood or plastic furring strips to wall plugs inserted into mortar joints as shown in
Fig. 14, at the top, bottom and mid-height of the wall. Rigid board insulation may be installed
between furring strips or over them to form a cavity for drainage. Gypsum board or other interior
finish is then placed over the insulation. Another option involves the use of a special "Z" clip. The clip
is attached to the brick wythe at 16 or 24 in. (400 or 600 mm) o.c. horizontally or vertically. The leg
of the clip extends beyond the rigid board insulation, and special hat channels (so-called because of
their shape) are screwed onto the clip, as shown in Fig. 15. The interior finish is then screwed to the
hat channels to finish the wall. A third alternative is light-gage (non-bearing) metal stud framing. The
framing is used to form a drainage cavity and to apply insulation and/or finishes in a manner similar to
that in brick veneer wall systems. Light-gage metal framing is installed by attaching a track to the
floor and ceiling joists at the desired distance from the brick wythe. Rigid board insulation may be
placed between the metal framing and brick bearing wall, or insulation may be placed between
studs. The framing provides a level surface for applying the interior gypsum board or other interior
finish. The framing does not provide support for the brick bearing walls. This permits a substantial
reduction in the size of the metal framing members. One and one-half inch (40 mm) studs are often
used.
Wall Plug Nailing Insert
Fig. 14
t26 http://www.gobrick.com/BIA/technotes/t26.htm
13 of 16 9/13/2009 1:17 PM
"Z" Clip Installation
Fig. 15
Construction Considerations
Quality Control
The quality of construction of a brick bearing wall is important for several reasons. The masonry wall
is the primary structural system for the building and must meet the minimum strength necessary for
adequate performance. Without quality construction, the expected strength of the masonry may not
be achieved. Good workmanship is also important in resisting water penetration. In a single wythe
wall, the masonry is the primary barrier to water penetration, and good workmanship directly affects
the water penetration resistance of the masonry.
To ensure quality construction, the MSJC Code and Specifications contain several provisions
regarding materials testing, inspection of masonry and workmanship. Brick units and mortar may
have to be tested to verify compliance with applicable standards. Verification of assembly
compressive strength may be determined by preconstruction testing of brick masonry prisms
constructed from the same brick and mortar to be used on the project. Alternately, assembly
compressive strength may be verified by the conservative unit strength method which requires
knowledge of the brick unit compressive strength and mortar Type to estimate the masonry
assemblage's strength. Inspection of the masonry during construction is required by the MSJC
Code, and limits on the dimensional tolerances of masonry elements ensure expected structural
performance.
In reinforced brick masonry, maintaining clear grout spaces during construction and proper location of
reinforcement are important. The cells of hollow brick intended to receive reinforcing bars and grout
should be free of mortar protrusions and debris. One method of keeping the cells of hollow brick
clean is to insert sponges into the cells to be grouted at the beginning of construction. The sponges
are pulled upward by a handle, wire or string as construction progresses, leaving clean cells ready
for grouting, as shown in Fig. 16. Another method, if sponges are not used, is to provide cleanout
openings at the base of the wall at all grout locations. The cleanouts allow the mortar protrusions to
be scraped off after construction and the debris removed at the wall base. See Fig. 17. The
cleanouts are sealed prior to commencing with grouting. One method of closing cleanouts in hollow
brick is to bevel cut the section of the face shell to be removed. When replaced, the cut piece is held
in position by pressure from the grout. Cleanouts are seldom used in single wythe construction
except with large cell hollow brick or at pilaster locations. For further information on grouting brick
masonry walls, see Technical Notes 17C and the MSJC Code and Specifications.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
14 of 16 9/13/2009 1:17 PM
Sponges In Cells To Be Grouted
Fig. 16
Grout Cleanouts
Fig. 17
To resist water penetration, full head joints should be used with solid and hollow units, and full bed
joints are recommended with solid units. Proper installation of flashing and weep holes and other
moisture control measures will control water which does penetrate the brick masonry wythe.
Sequence of Work
In a loadbearing wall system, brick masonry is the primary structural element. The masonry work
may begin as soon as the foundation is complete and properly cured. Bearing walls should be
braced during construction until lateral support is provided by the floors and roof. Once one story
height is laid, construction of the floor or roof systems follow, possibly serving as a work platform for
the remaining masonry work. The fasteners necessary for attachment of the cabinets, insulation and
interior finishes should be incorporated during the construction of the masonry. Interior frame walls
may be built simultaneously with the exterior loadbearing brick walls once the floor has been
constructed, at the builder's discretion.
Brick bearing walls should attain sufficient strength before any loads are applied. The curing
conditions will affect the rate of strength gain of loadbearing masonry. If sufficient moisture is
maintained, the masonry walls should cure a minimum of three days before supporting floor or roof
t26 http://www.gobrick.com/BIA/technotes/t26.htm
15 of 16 9/13/2009 1:17 PM
loads. Reinforced brick masonry beams require curing periods of at least seven days. Poor curing
conditions, such as exposure to cold temperatures, may require longer curing times. Once the brick
bearing walls are cured, the floors and roof may be attached. The windows, doors, plumbing,
electrical and heating systems, insulation and interior finishes can be installed as soon as the
masonry is complete.
SUMMARY
This Technical Notes covers the design and detailing of single wythe bearing walls. Selection of
materials and methods of reinforcing and finishing brick bearing walls are discussed. The information
provided illustrates a few of the many options available for a single wythe brick bearing wall system.
Design and construction considerations presented are equally applicable to residential and
commercial construction. Any design should be completed in accordance with the MSJC Code and
Specifications and applicable model building and energy codes.
The information and suggestions contained in this Technical Notes are based on the available data
and the experience of the engineering staff of the Brick Institute of America. The information
contained herein must be used in conjunction with good technical judgement and a basic
understanding of the properties of brick masonry. Final decisions on the use of the information
contained in this Technical Notes are not within the purview of the Brick Institute of America and must
rest with the project architect, engineer and owner.
REFERENCES
1. "Brick Masonry Material Properties," Technical Notes on Brick Construction 3A, Brick Institute of
America, Reston, VA, December 1992.
2. "Brick Masonry Section Properties," Technical Notes on Brick Construction 3B, Brick Institute of
America, Reston, VA, May 1993.
3. "Building Code Requirements for Masonry Structures ACI 530/ASCE 5 and Specifications for
Masonry Structures ACI 530.1/ASCE 6," Technical Notes on Brick Construction 3, Brick Institute of
America, Reston, VA, February 1990.
4. Building Code Requirements for Masonry Structures and Commentary (ACI 530/ASCE 5/TMS
402) and Specifications for Masonry Structures and Commentary (ACI 530.1/ASCE 6/TMS 602),
American Concrete Institute, Detroit, MI, 1992.
5. One-and Two-Family Dwelling Code, Council of American Building Officials (CABO), Falls Church,
VA, 1992.
6. "Structural Design of Brick Masonry Arches," Technical Notes on Brick Construction 31A Revised,
Brick Institute of America, Reston, VA, July 1986.
7. "Structural Steel Lintels," Technical Notes on Brick Construction 31B Revised, Brick Institute of
America, Reston, VA, May 1987.
t26 http://www.gobrick.com/BIA/technotes/t26.htm
16 of 16 9/13/2009 1:17 PM

Technical Notes 27 - Brick Masonry Rain Screen Walls
August 1994
Abstract: Pressure equalization across the exterior wythe of brick veneer and cavity walls allows the rain screen
principle to minimize the infiltration of rain into exterior walls. This Technical Notes focuses on the design and wall
components that contribute to the pressure equalized rain screen wall. A compartmented air cavity behind the
exterior brick wythe, a rigid air barrier system and adequate venting area of the exterior cladding in relation to the
leakage area of the air barrier are necessary elements.
Key Words: air barrier, air retarder, brick veneer, cavity wall, drainage wall, exterior cladding, pressure
equalization, rain screen, wind loads.
INTRODUCTION
Rain penetration through walls can damage the building envelope. Corrosion of metal accessories in the exterior
cladding, efflorescence of the masonry and damage to interior finishes and staining are just a few examples of
problems related to moisture penetration. Water penetration affects the appearance and function of a variety of
brick masonry wall systems.
Over the years, many methods have been used to prevent moisture penetration of walls, some more successfully
than others. Masonry barrier walls rely on the massive wall materials to deter water penetration. Drainage type
walls, such as brick veneer and cavity walls, provide good moisture penetration resistance. It must be recognized
that the exterior wythe can not be made watertight. Provisions for internal drainage are necessary for these wall
systems to function as intended.
The next step to provide better moisture penetration resistance for exterior brick walls is the use of the rain screen
principle. This concept introduces air into the cavity of conventional drainage type walls to provide pressure
equalization so that the cavity works in resisting wind-driven moisture penetration.
Cladding researchers and investigators increasingly recognize air pressure as a major cause of water penetration
problems. They are looking more carefully to the rain screen principle as a deterrent to moisture penetration. This
Technical Notes discusses the design criteria of the rain screen wall, how to develop the pressure equalization
feature within the cavity space and additional construction detailing needed to tailor conventional brick veneer and
cavity wall systems to the pressure equalized rain screen principle. Other Technical Notes in this series will address
construction considerations and material selection.
HISTORY
The rain screen principle has been used intuitively for many years. One of the first references to it was made in 1946
by C. H. Johansson entitled, "The Influence of Moisture on the Heat Conductance for Brick". It was not until sixteen
years later that researchers began to understand how to apply the fundamental laws of physics to the development
of the rain screen principle for practical use.
In 1962, Birkeland of the Norwegian Building Institute wrote Curtain Walls in which he stated:
"The only practical solution to the problem of rain penetration is to design the exterior rainproof finishing so open
that no super-pressure can be created over the joints or seams of the finishing. This effect is achieved by
providing an air space behind the exterior finishing, but with connection to the outside air. The surges of air
pressure created by the gusts of wind will then be equalized on both sides of the finishing. "
Birkeland noted six main sources of moisture leakage through wall systems. The processes discussed were: 1)
wind-induced air pressure differences; 2) pressure assisted capillarity; 3) gravity; 4) kinetic energy; 5) air currents;
http://www.gobrick.com/BIA/technotes/t27.htm
1 of 14 9/13/2009 1:18 PM
and 6) updrafts. Conventional means such as internal wall flashing, proper design of openings and overhangs provide
moisture resistance to items 2 through 6. But item 1 was the most difficult to counteract. He concluded that there
was no practical method of obtaining total watertightness in wall systems composed of joints when a pressure
gradient exists across the exterior rain barrier. These observations formed the basis of the rain screen principle.
Prompted by Birkeland, researchers in Canada began an intensive study into wall leakage. In 1963, Canadian
Building Digest (CBD) 40, "Rain Penetration and Its Control" was published by the Canadian National Research
Council's Division of Building Research. This publication. which remains a prime reference source on the subject,
popularized the term rain screen principle. G. K. Garden, who authored CBD 40 on wind-induced moisture
penetration wrote:
"It is not conceivable that a building designer can prevent the exterior surface of a wall from getting wet nor that he
can guarantee that no openings will develop to permit passage of water. It has, however, been shown that
through-wall penetration of rain can be prevented by incorporating an air chamber into the joint or wall where the
air pressure is always equal to that on the outside. In essence, the outer layer (wythe) is then an open rain screen
that prevents wetting of the actual wall or air barrier of the building".
The critical features of the rain screen principle are:
1. An exterior barrier (rain screen) containing protected openings which permit the passage of air but not water.
2. A confined cavity behind the rain screen in which air pressure is essentially the same as the external air pressure.
3. Insulation fixed to the outer face of the interior wall system, if provided in design.
4. An interior barrier (wall) which substantially limits the passage of air and water vapor and is capable of
withstanding all required design loads (e.g. wind and earthquake forces).
Many field studies and laboratory tests have applied this principle to a variety of wall systems, including masonry.
The developments made over the years can be effectively applied to conventional brick masonry drainage wall
construction, with some modifications as discussed in this Technical Notes.
DEFINITION AND PRINCIPLES
Drainage wall types, such as anchored brick veneers and cavity walls, which provide a space for drainage of
moisture that has penetrated the exterior wythe, are often confused with rain screen walls. When causes of rain
leakage problems are debated, the question usually arises of whether the wall system utilizes the rain screen
principle. Certainly, there is a cavity between the exterior wythe and interior wall which provides drainage of moisture
which has entered the wall. The concept of drainage type walls has been around for decades. More information on
brick veneer and cavity wall systems can be found in Technical Notes 28 Series and 21, respectively. However, the
basic premise of the rain screen principle is to control all forces that can drive moisture through the wall system.
The term pressure equalized rain screen wall should be used. This emphasizes the difference from the more
common drainage type wall. The pressure equalization in the cavity behind the exterior wythe is the major difference
between a rain screen wall and a drainage wall. A pressure equalized rain screen wall provides the best means of
resisting water penetration. As such, it should be used on projects located in areas which receive high volumes of
wind-driven rain and when resistance to water penetration is of prime concern.
Pressure Equalized Rain Screen Walls
The difference in air pressures across the exterior cladding is a significant force which causes infiltration of air and
water on windward facades. Air and moisture can infiltrate through units, mortar joints, hairline cracks, poorly
bonded surfaces and other openings that exist or develop over the life of the structure.
A rain screen wall is composed of two layers of materials separated by a cavity as shown in Figure 1. The exterior
cladding as discussed in this Technical Notes is a brick masonry wythe. The interior wall or inner layer can be either
the backing of an anchored brick veneer wall or the inner wythe of a cavity wall. When wind loads are imposed on
the wall assembly, a pressure difference between the exterior wythe and the cavity space is created. This pressure
difference forces water on the surface of the exterior cladding to penetrate any openings through the wall. If the
http://www.gobrick.com/BIA/technotes/t27.htm
2 of 14 9/13/2009 1:18 PM
exterior cladding has sufficient openings to permit air to flow to the cavity behind the cladding, the pressure in the
cavity increases until it equals the pressure resulting from the wind load being applied. This is the phenomenon of
pressure equalization design. To affect this air pressure transfer, the inner layer of the wall assembly must be
airtight. This is achieved by applying an air retarder at some location on the backing or inner wythe. The air barrier
seal at this location should last longer because it is not exposed to the exterior elements. Since the interior wall will
be airtight, stack effect and mechanical ventilation generated inside the building are effectively controlled. Rain
penetration through the exterior cladding should be reduced as the pressure difference on the exterior cladding which
drives rain into the cavity is reduced. The resultant wind load will be imposed on the air barrier and interior wall.
Rain Screen Wall Principle
FIG. 1

Moisture Migration
Exterior claddings primarily restrict the passage of water and wind and also function as part of a thermal barrier.
The extent to which the exterior cladding can be relied upon to serve these functions is variable, and the exterior
cladding is not considered to be the sole air or moisture barrier in the wall system.
Rain screen walls using brick veneer and cavity wall systems should be designed as a two-stage barrier for moisture
penetration resistance. The first stage is the exterior brick wythe. The backing assembly or the inner wythe is the
second stage. The exterior brick wythe should be detailed and constructed to provide moisture resistance so that
the second stage is not continually tested. If excess water penetrates the exterior brick wythe, the backing system
may become a single stage in itself which can lead to failure of the rain screen principle as a whole. A typical brick
masonry rain screen wall is shown in Fig. 2.
http://www.gobrick.com/BIA/technotes/t27.htm
3 of 14 9/13/2009 1:18 PM
Brick Rain Screen Wall
FIG. 2
The exterior brick wythe is the first stage of the rain screen principle, and a majority of the rain water will run down
the face of the brickwork. Some moisture in contact with the exterior wythe is absorbed by capillary action. If wind
pressure is applied to the face of the exterior brick wythe, the moisture will be forced into the brickwork, particularly
at mortar joints or openings. The use of dissimilar materials, the presence of mortar joints and the variations of
workmanship make it difficult to ensure a fully waterproof exterior wythe. Some moisture will penetrate the brick
wythe and infiltrate into the cavity space. If the cavity space is at the same air pressure as the exterior as a result of
air flow through vent openings and weep holes, the only moisture which will reach the cavity space is due to gravity
flow and capillary action. For the rain screen principle to work effectively, water which penetrates the exterior brick
wythe travels down the interior side, is collected on flashing and transferred to the exterior through weep holes.
It is not advisable to support the exterior cladding on the floor assembly with the backing system. If detailed in this
manner, moisture which penetrates the exterior cladding and its flashing has direct access through the joint at the
base of the interior wall to the interior of the building. Moisture can also flow under the flashing and penetrate directly
into the building. To avoid the chances of moisture penetration at this location, the exterior cladding support should
be lower than the support of the interior wall.
The rain screen principle relies on the use of intentional openings in the exterior brick wythe to create pressure
equalization in the cavity. The cavity pressure should be close to the external pressure. That depends upon the air
leakage characteristics of the exterior brick wythe and that of the air retarder on the backing or the inner wythe.
Sufficient openings in the exterior wythe to balance air pressures with the exterior of the wall system create the
pressure equalized rain screen wall. The openings are created by the use of weep holes and vents. Vents near the
top of the wall help permit air circulation through the cavity which helps in drying out the wall system.
http://www.gobrick.com/BIA/technotes/t27.htm
4 of 14 9/13/2009 1:18 PM
For the second stage to work effectively, both pressurization of the cavity and the provision for an airtight barrier are
extremely important. The extent to which the cavity can be pressurized will reduce the amount of moisture carried
through the exterior wythe by wind. It also tends to decrease the tendency for moisture that penetrates due to
capillary action to bridge the cavity and contact the backing.
Moisture may actually be transported through the wall system by air passing through weep holes and vents.
Movement of air within the cavity can transfer moisture to the interior wall and distribute it along the wall area. Air
leakage can then draw this moisture into and through the backing or inner wythe. Laboratory tests have shown that
high air leakage through the backing or inner wythe can even cause moisture to climb up and extend the area of wall
wetness. The backing or inner wythe should not permit air leakage to occur, thus vents will not have to be oversized
which could permit excess rain penetration.
Vapor and Air Retarders
There has been much confusion in the building industry about the functions of vapor and air retarders. Vapor
retarders are intended to control transmission of water vapor through building materials. A vapor retarder always
serves as an air retarder. Air retarders limit the amount of air flow through wall assemblies. An air retarder may or
may not serve as a vapor retarder. It is difficult that either retarder performs only one function. For example,
polyethylene film is commonly used as a vapor retarder but will also act to resist the passage of air. Most types of
sheathing used as air retarders tend to permit the passage of water vapor. This can result in a common problem:
many wall systems essentially have a two-stage setup of retarders.
In actual construction, this means that moisture may become trapped between the air and vapor retarder
installations if both are provided at different locations in the wall assembly. The amount of moisture and the duration
of wetness of certain critical elements may render the wall design vulnerable to premature deterioration and
distress. Of concern is the potential for corrosion of metal accessories within the wall system, deterioration of
sheathing materials and decrease in insulation capacity.
Vapor retarders are normally placed on the warm side of insulation in the wall assembly. Air retarders on the other
hand have no distinct position. If interior wall board or other finishes are used as retarders, concern arises over
punctures for utility services and wall hangings. Further, the incomplete seal around the perimeter of the wall section,
may render the air retarder ineffective. The use of the air retarder on the exterior side of the interior wall has
additional liabilities. Inspection for proper installation may be difficult and corrective repair may be troublesome.
When air exfiltration does occur, the exterior applied air retarder which has some vapor retarder qualities, may result
in condensing water being trapped in the wall. Walls designed with rigid insulation on the exterior side of the air
retarder can be designed so that the second (partial vapor retarder) barrier is at a temperature above the dew point
so condensation problems can be eliminated.
Generally, the mass of water which can penetrate into a wall assembly by air leakage, even through a very small
opening, is several times larger than the amount of water which infiltrates by other means. When this airborne water
condenses, it is unlikely that it can be removed by vapor transmission or evaporation to the exterior. Drying out of
this water is more likely to occur by air moving through the wall system under different temperature and humidity
conditions.
The location of the air and vapor retarders will depend on the wall system in question and whether inspection is
needed during construction. Construction not requiring inspection should have the air and vapor retarders placed on
the interior side of the interior wall. Incorrect installation or faults can easily be repaired or maintained. Where
inspection is employed on a project, the air and vapor retarders can be installed within the wall system. However,
the long term performance is questionable because these installations are not easily accessible for repair.
When detailing for both air and vapor retarders in pressure equalized rain screen walls, the following should be
accounted for:
1. wall openings,
2. disruption of the retarders due to building services such as utilities,
3. concealed spaces such as areas above suspended ceilings, behind heating elements and at junctions of interior
walls which connect with outer portions of the building envelope,
http://www.gobrick.com/BIA/technotes/t27.htm
5 of 14 9/13/2009 1:18 PM
4. minimize and seal joints of interior finishes,
5. minimize and seal joints between interior finishes and interruptions such as interior partitions and wall openings.
Most difficulties with installing the retarders occur at wall openings. A variety of materials intersect in one area,
which can be complicated. Damage by subsequent trades may breach or puncture the air and vapor retarders.
Openings must permit field construction tolerances which must be accommodated by field-fit and sealing of the
retarders. Also, attention to details of the air and vapor retarders can help minimize direct heat loss and other
detrimental effects due to exfiltration of air movement within the wall system. Figure 3 depicts the possibilities of
exfiltrating air in wall construction. Air can circulate through spaces between studs and cells of masonry units and
exit at leakage paths to the exterior. Where the components of a building assembly can be completely sealed to
prevent air leakage and the interior finish material provides the vapor resistance needed, a separate vapor retarder
is not usually required.
Sources of Exfiltrating Air Movement
FIG. 3
To provide effective air and vapor retarders, it is necessary to seal joints in these materials so that continuity is
provided. It is also necessary to seal around the edges of wall openings such as windows, doors and access for
utility services. Caulking or taping of the joints should be specified. The joints in the retarders should be detailed and
dimensioned with consideration of the need to accommodate variations in joint dimensions, building deflections and
differential movement of building materials.
If the sealant materials are inaccessible after construction, they must have qualities which will provide satisfactory
performance over the life of the structure. Even when the sealant material can be repaired or replaced, the sealant
must be suitable for the construction materials and environmental conditions. Field experience, tests or
manufacturer's literature should be used as the basis for selecting sealant materials for compatibility with the air and
vapor retarders selected.
The successful performance of the joint seal over the life of the structure depends on the ability of the material to
adhere to the surfaces and to deform without tearing, delamination or peeling under repeated cycles of expansion
and contraction. Workmanship is also very important. Air bubbles in the sealant or air voids between the sealant and
adjacent materials must be avoided. Sealant manufacturer's information should be followed for proper installation.
Axial and Lateral Loads
http://www.gobrick.com/BIA/technotes/t27.htm
6 of 14 9/13/2009 1:18 PM
Load application to the exterior brick wythe and the interior wall and load transfer are based on the type of wall
design and its associated construction. The exterior wythe is treated differently in cavity wall design from that in
veneer wall design. The treatment of the exterior wythe of a rain screen wall with respect to axial and lateral loads is
no different from that of conventional construction.
The exterior wythe of a veneer wall should not be subjected to any axial load other than its own weight. Veneers
must be designed, detailed, and constructed so that no axial loads are imposed. The exterior wythe of a cavity wall
can carry vertical loads and still perform as the exterior cladding in the rain screen wall. The imposed load depends
on the anchorage of floor or roof components to the exterior wythe.
Building codes state that the brick wythe of an anchored veneer wall does not resist lateral loads and that the
backing must be designed to carry all lateral loads. In cavity wall design, both wythes share a portion of the lateral
loads imposed. In fact, all lateral loads will initially be resisted by the exterior wythe and then transferred through the
wall system to the building frame. For lateral load distribution to occur, the exterior wythe must be anchored to the
backing or interior wythe with metal ties in a sufficient number and spacing. Design must consider lateral load
distribution, tie stiffness and deflection of the wall system. See Technical Notes 21 Series and 28 Series for the
proper design and construction of cavity walls and brick veneer over wood or steel backings, respectively.
Thermal Insulation
All wall components affect the thermal resistance of the assembly and contribute to the overall R-value. For masonry
wall systems, the insulation provides most of the thermal resistance. The designer should choose the level of
insulation required as part of the total wall design, considering its location.
The type and location of the insulation has a significant impact on the design and installation of the air and vapor
retarders. The possible locations for thermal insulation include: 1) in the cavity; 2) in the interior wall; and 3) on either
face of the interior wall. The general types of insulation used in drainage type walls are rigid board insulation,
fiberglass (batt) insulation or loose fill.
It is important to eliminate any gap between the insulation and the floor or ceiling. With suspended ceilings or ceilings
attached to the bottom chord of joist construction, the insulation should be continued above the ceiling to the bottom
of structural slabs, not as detailed in Fig. 4. For the pressure equalization to occur, air and vapor retarders must also
be continued to the floor or roof above the suspended ceilings. If the air retarder is not continued, the insulation may
separate from the backing wall by air infiltration pressure. Proper abutment of the edges of the insulation must be
considered to hinder air circulation from the interior of the building.
http://www.gobrick.com/BIA/technotes/t27.htm
7 of 14 9/13/2009 1:18 PM
Leakage Above Suspended Ceilings
FIG. 4
Cavity Placement. Rigid board insulation should be used when insulation in the cavity is desired. In order to be
effective, it must be adhered tightly to the exterior side of the interior wall. If any air is permitted to circulate behind
the insulation, the pressure equalization of the cavity is weakened as is the insulation capacity. As a result, the
method of securing the insulation is important. Generally, adhesives or mechanical fasteners are used. Consideration
on selecting adhesives includes: 1) assuring clean surfaces under field conditions; 2) compatibility of the adhesive
with the insulation; 3) long-term effectiveness of the adhesive; and 4) movement of the wall system.
Use of dabs of adhesive is not recommended because it creates an air gap which allows free air movement. Use of
a full adhesive bed is recommended, where the full adhesive bed acts as a vapor barrier. If a full adhesive bed is not
desired, a grid set-up is recommended because it will compartmentize the air spaces behind the insulation board and
retain the full insulation capacity.
There are a variety of mechanical fasteners available to attach rigid board insulation to the backing wall. Use of
mechanical fasteners is suggested where uneven surfaces or protrusion of mortar require bending of the insulation to
avoid air-gaps between the insulation and the interior wall. Mechanical fasteners are more reliable in the long term
than adhesively attached clips. Adhesively attached fasteners are more convenient, but the same considerations for
selecting adhesives to attach the insulation must be evaluated. Mortar joints should be cut flush to remove fins or
protrusions and provide as smooth a surface as possible for a tight fit of the insulation to the interior wall.
The insulation must support its own weight and accommodate expected wall movements. Further, the insulation may
be subjected to wind pressures. These wind pressures must be resisted by the insulation to avoid separation of the
insulation from the interior wall. The choice of mechanical fasteners, adhesives or a combination of both should take
into consideration the life of the structure, conformance with manufacturer's specifications and possible failure of the
fastener or adhesive. Thus, in some cases a combination of both may be advisable to ensure that costly repairs are
not needed for the life of the structure.
Loose granular fill insulation must not be placed in the cavity. This placement violates the rain screen principle.
However, loose fill insulation, such as vermiculite and perlite, can be used in the unobstructed vertical cells of the
masonry backing to increase thermal resistance.
Stud Wall Placement. For stud wall backings, fiberglass (batt) insulation is placed in the stud cavities. The
insulation is usually slightly larger than the stud depth to provide a friction fit. Stud spacing should be controlled so
that friction will be effective in keeping the insulation in place. The fiberglass insulation must fill the entire stud cavity
so that air circulation is minimized. If not completely filled, convection can greatly reduce the thermal resistance of
the interior backing wall.
Interior of Wall Placement. Insulation installed on the interior of the interior wall is usually limited to masonry
backings. It provides easy installation and is available to inspection of placement. The insulation can be applied by
many means, but the need to support interior finishes usually requires the use of metal or wood furring.
Differential Movement
All building materials change dimension with changes in temperature. Some building materials change dimension with
moisture content. All materials will deform elastically when subjected to loads. Some materials with cement matrices
will deform plastically (creep) when loaded. Adequate allowance for deformations of materials and building
movements are critical to the successful performance of the pressure equalized rain screen wall. Problems can arise
if these naturally occurring movements are not recognized and accommodated for in initial design. The air and vapor
retarders must not be disrupted by building movements, whether it be material generated or the building as a whole.
Sealing of movement joints is required to prevent the passage of water and air without restricting differential
movement. The sealant acts as the primary resistance to the passage of water through joints in exterior elements. A
backing material and, perhaps, a filler may be needed for large movement joints.
Discussion of building movements is beyond the scope of this Technical Notes. The reader is directed to Technical
Notes 18 Series for information on accommodating material and building movements, such as the building frame,
http://www.gobrick.com/BIA/technotes/t27.htm
8 of 14 9/13/2009 1:18 PM
exterior cladding and interior wall systems. Recommendations for other materials should be reviewed.
PRESSURE EQUALIZED RAIN SCREEN DESIGN PARAMETERS
The pressure equalized rain screen wall, whether a brick veneer or a cavity wall system, will be subject to both axial
and lateral loads. In the design of these innovative wall systems, imposed loads must be taken into account for the
rain screen wall to perform as intended. Other environmental loads, such as moisture leakage, thermal and air
retarder performance, must also be considered. There are many parameters which affect the pressure equalized
rain screen principle. These parameters include: 1) rate of applied wind load; 2) magnitude of applied wind load; 3)
cavity volume; 4) stiffness of the interior wall and the exterior cladding; 5) compartmentation of the cavity; and 6)
leakage areas of the air retarder and the exterior cladding. Many of these factors are interrelated.
Wind Loads
An advantage of the pressure equalized rain screen wall, in theory, is that no wind load should be imposed on the
exterior cladding. However, wind is dynamic and variable so that the pressures applied to the wall are constantly
changing. An ideal rain screen wall would pressure equalize instantly. In fact there is a time lag between the imposed
wind load and pressure equalization in the cavity. As a result a pressure difference does occur across the exterior
cladding.
This pressure is normally positive, a driving force pushing air into the wall. However, the cavity pressure can exceed
the positive wind pressure under gusting wind conditions. This situation occurs after the cavity pressure increases to
match that of a high wind, and the wind suddenly decreases. Until pressure equalization occurs, the cavity pressure
will exceed the exterior wind pressure which creates a negative load on the cladding. This negative loading will tend
to force water out of openings in the exterior cladding, reducing the likelihood of further moisture penetration. Under
the rain screen principle, wind loads on the exterior brick wythe may actually be reduced due to pressure
equalization of the cavity space. However, the entire design wind load should be applied to the exterior cladding and
the backing.
Pressure Differences and Distribution
Pressure differences are encountered in buildings from two main sources. The first is commonly referred to as stack
effect, which is created by temperature differences between the exterior and interior of the building. The second is
the wind forces that are imposed on the building envelope. The net pressure difference across a wall system at the
top and sides may be a combination of both and is not the same for all parts of the building envelope.
The stack effect which occurs mainly during the heating season results from warmer inside air rising as a result of a
lower density than cooler outside air. This difference in density creates an outward positive pressure at the top of a
building, while exerting negative pressure at the wall base. Thus, air will tend to infiltrate at the lower levels of the
building and exfiltrate at the upper levels.
Wind causes air infiltration on the windward side of buildings and exfiltration on the leeward side and also on the
sides parallel to the wind direction. A flat roof will experience exfiltration because of negative uplift pressure caused
by wind. Since wind velocity increases with height, the difference in pressure across the building envelope increases
with height.
Pressure distribution on the windward facade varies from a maximum at the center and decreases towards the
corners of the building. Suction pressures on the leeward wall vary from a maximum at the corners of the building,
diminishing towards the center. The pressure on the side walls parallel to the wind direction is normally negative, but
may change rapidly to positive pressure as the wind changes directions. This is usually why wind pressures are
normally higher at the top and corners of the building envelope.
The rain screen wall minimizes the air pressure differences across the exterior brick wythe by transferring the
pressure to the cavity space, see Fig. 1. Under the imposed wind load (Pe), air flows into the cavity causing the
cavity pressure (Pc) to increase until Pc = Pe and Pc = 0. When the entire wind load is imposed on the interior wall,
pressure equalization will occur. The veneer backing or the interior cavity wall wythe is thus designed for the entire
wind load.
Cavity Volume
http://www.gobrick.com/BIA/technotes/t27.htm
9 of 14 9/13/2009 1:18 PM
When positive pressure is applied to the exterior cladding, movement of air into the cavity causes the pressure in the
cavity to increase to match the external pressure applied. The volume of air required to achieve pressure
equalization is dependent on the volume of the cavity. The rate at which pressure equalization occurs is dependent
on the rate at which air can enter the cavity. Thus, as the cavity volume increases, the vent openings in the exterior
brick wythe must be increased in order to permit more rapid pressure equalization. The driving force causing air to
enter the cavity is the pressure difference across the exterior cladding. As air enters the cavity, this pressure
difference decreases. The flow rate is proportional to the pressure difference, and as air flows into the cavity, the
flow rate decreases.
Both the exterior cladding and air retarder applied to the interior wall will deflect under applied loads. Stiffness of
these elements will influence the volume of the cavity. Since these deflections also vary as the pressure differences
vary, it becomes clear that this situation is very complex.
Leakage of the Exterior Cladding and the Air Retarder
The relative airtightness of the exterior cladding with respect to that of the air retarder applied to the interior wall is
paramount for the cavity to pressure equalize with the exterior wind pressure. If the two layers have similar air
leakage characteristics, each layer will transmit the same volume of air, and the pressure differences will not
change. If the interior wall is made more airtight, a greater pressure difference will occur across it than that across
the exterior cladding. In the ideal case, the air retarder would be completely airtight, and the pressure difference
across the exterior cladding would be negligible. However, this is generally not the case.
The effectiveness of the rain screen wall is decreased as greater amounts of air are permitted to penetrate through
the air retarder. The Architectural Aluminum Manufacturers Association specifies that, for laboratory tests, air
leakage through a curtain wall should not exceed 0.06 cfm/ft
2
(0.0003 m
3
/s/m
2
) for a pressure difference equivalent
to a 25 mph (11 m/s) wind. This value for air retarders is currently being considered in ASTM Committee E 6.
However, in Canada, the acceptable rate is one-third of that currently being suggested in ASTM. It seems prudent to
use the lower value for air leakage through the air retarder.
Compartmentation
The wind pressure flowing around a building creates a distribution of positive and negative pressures over the
building exterior cladding as shown in Fig. 5. If the cavity of the rain screen wall is continuous, horizontally or
vertically, lateral flow of air in the cavity will occur.
If air is permitted to flow laterally in the cavity, pressure equalization will not occur. Moisture penetration into the wall
assembly may not be reduced.
To prevent lateral airflow, the cavity must be compartmented. The size of the compartments should be based on the
pressure differences across the exterior cladding. The corners and tops of buildings experience the greatest
pressure differences; hence, the compartments located in these areas should be small. Where pressure differences
are small, such as the center of the exterior cladding, the compartments can be larger. Many researchers suggest
that these compartments have closures no more than 4 ft (1.2 m) apart at the sides and top of the building in a 20 ft
(6 m) wide perimeter zone, see Fig. 6. Compartment closures should be within this recommendation for the sides
and top of the building. Based on the design wind pressure of the wall area in question, compartment dimensions
outside the 20 ft (6 m) zone of the tops and corners of the building can be increased. For a design wind pressure
less than 15 psf (718 Pa), compartment closures should be provided for every 400 ft
2
(37 m
2
) of wall area. When
the design wind pressure is between 15 psf (718 Pa) and 25 psf (1200 Pa), the compartment closures should be
decreased to 250 ft
2
(23 m
2
) of area. When the design wind pressure is greater than 25 psf (1200 Pa), the
compartment closures should have a maximum area of 100 ft
2
(9 m
2
). At the minimum, the cavity must be closed at
all corners and at the roof to prevent air from the windward side of the building laterally flowing to the adjacent
sides, as shown in Fig. 5.
http://www.gobrick.com/BIA/technotes/t27.htm
10 of 14 9/13/2009 1:18 PM
Moisture Movement Caused by Wind
FIG. 5
http://www.gobrick.com/BIA/technotes/t27.htm
11 of 14 9/13/2009 1:18 PM
Compartmentation of Rain Screen Walls
FIG. 6
Research conducted by Canada Mortgage and Housing Corporation used wind tunnel testing to determine the
effects of compartmentation in the cavity. One significant result of this testing was that the compartment closures
experienced at least two times the applied wind load. This research concluded that compartment closures must be
designed for high wind pressures and must be completely sealed to prevent lateral airflow from one compartment to
the next.
Exterior Cladding Openings
To provide pressure equalization in the rain screen wall, there must be a series of openings to connect the cavity to
the exterior of the wall system. The openings should be positioned at the top and bottom of each compartment. All
openings at the top and bottom should be placed at the same height, respectively, to avoid airflow loops in the
cavity.
There are no definitive guidelines for the required amount of openings for each compartment. The area of openings
depends on the airtightness of other components of the cavity, e.g., the air retarder system and the cavity closures.
If completely sealed compartment closures are used, a 10:1 ratio for cladding air leakage to air retarder leakage is
recommended. Most often, the cavity closures will not form an airtight seal of the individual compartments. To
account for this the required opening area should be larger. Some studies suggest a ratio of 25 to 40 times more air
flow volume through the openings in the exterior brick wythe than air leakage through the interior wall. Therefore, the
tighter the compartment, the less the area of openings in the exterior cladding required for pressure equalization of
the cavity. To obtain the airtightness value of the interior wall construction, testing of a mock-up wall compartment
may be required since little information on the range of air tightness of various field-applied air retarder components
is available. After having evaluated the air tightness of the interior wall, the openings in the exterior cladding should
be established to fit the recommended ratio.
http://www.gobrick.com/BIA/technotes/t27.htm
12 of 14 9/13/2009 1:18 PM
The following recommendations should provide sufficient venting to achieve pressure equalization for the rain screen
wall. Vents are installed at the top and bottom of each compartment. Open head joints should be spaced at a
maximum of 24 in. (600 mm) o.c. horizontally in the exterior brick wythe. If clear, round openings are used, they
should be at least 3/8 in. (10 mm) inside diameter, and the spacing should be reduced to 16 in. (400 mm) o.c.
horizontally. Open head joints may be positioned at flashing locations in the exterior brick wythe and serve as weep
holes. A minimum of two vents at both the top and bottom should be provided for each individual compartment. The
suggested minimum cavity width is 2 in. (50 mm). If the cavity space width is greater than 2 in. (50 mm), open head
joints should be used as vents. Vents should not be positioned at corners. Water drainage provisions in the wall
assembly should be evaluated carefully to avoid placing vents at high flow areas such as sills and heads of openings
in the wall system. It is imperative that the cavity have no blockage due to mortar bridging or mortar droppings that
can collect at the bottom of the cavity closures in each compartment, thus blocking the vent openings.
SUMMARY
This Technical Notes describes the pressure equalized rain screen principle as it applies to conventional brick
veneer and cavity wall systems. This innovative wall design has been used extensively in Canada and Europe for
many years. Pressure equalization across the exterior wythe of drainage type walls is the main emphasis of the rain
screen principle. Design recommendations that cover the prominent aspects of the pressure equalized rain screen
principle to minimize rain penetration through the exterior walls are described.
The information and suggestions contained in this Technical Notes are based on available data and the experience
of the engineering staff of the Brick Institute of America. The information contained herein must be used in
conjunction with good technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the purview of the Brick
Institute of America and must rest with the project architect, engineer and owner.
REFERENCES
1. Anderson, J.M. and Gille, J.R., "Rain Screen Cladding -A Guide to Design Principles and Practice,"
Construction Industry Research and Information Association - Building and Structural Design Report,
Butterworths, London, England, 1988.
2. "A Study of The Rainscreen Concept Applied to Cladding Systems on Wood Frame Walls," Canada
Mortgage and Housing Corporation, Report No. 39108.0R1, Ottawa, Ontario, Canada, August 1990.
3. Drysdale, R.G. and Suter, G.T., "Exterior Wall Construction in High-Rise Buildings -Brick Veneer on
Concrete Masonry or Steel Stud Wall Systems," Canada Mortgage and Housing Corporation, Report NHA
5450, Ottawa, Ontario, Canada, 1991.
4. Fornoville, L., "Rain Resistant Masonry Construction," Third North American Masonry Conference,
Matthys, J.H. and Borchelt, J.G., Eds., University of Texas at Arlington, Arlington, Texas, 1985.
5. Piper, R.S. and Kenney, R.J., "Brick Veneer Walls -Proposed Details to Address Common Air and Water
Penetration Problems," Masonry: Design and Construction, Problems and Repair, ASTM STP 1180,
Melander, J.M. and Lauersdorf, L.R., Eds., American Society for Testing and Materials, Philadelphia, PA,
1993.
6. Rousseau, M.Z., "Facts and Fiction of Rain Screen Walls," National Research Council of Canada,
Construction Canada, Vol. 32, Number 2, March/April 1990.
7. Ruggiero, S.S. and Myers, J.C., "Design and Construction of Watertight Exterior Building Walls," Water in
Exterior Building Walls: Problems and Solutions, ASTM STP 1107, Schwartz, T.A., Ed., American Society
for Testing and Materials, Philadelphia, PA, 1991.
8. "Structural Requirements for Air Barriers," Canada Mortgage and Housing Corporation, Report No.
30133.0RI, Ottawa, Ontario, Canada, August 1991.
9. Williams, M.F. and Williams, B.L., "Water Intrusion in Barrier and Cavity/Rain Screen Walls," Water in
http://www.gobrick.com/BIA/technotes/t27.htm
13 of 14 9/13/2009 1:18 PM
Exterior Building Walls: Problems and Solutions, ASTM STP 1107, Schwartz, T.A., Ed., American Society
for Testing and Materials, Philadelphia, PA, 1991.
http://www.gobrick.com/BIA/technotes/t27.htm
14 of 14 9/13/2009 1:18 PM

Technical Notes 28 - Anchored Brick Veneer, Wood Frame Construction
Rev August 2002
Abstract: This Technical Notes deals with the prescriptive design of brick veneer with wood frame construction in
buildings limited to three stories in height in new construction. The properties of the brick veneer/wood stud
system are described, which lead to design considerations. Selection of materials, construction details, and
workmanship techniques are included. The minimum requirements given have proven successful for this type of
wall construction.
Key Words: brick, flashing, foundations, lintels, ties, veneer, weep holes, wood frame.

INTRODUCTION
Anchored brick veneer construction consists of a nominal 3 in. (75 mm) or 4 in. (100 mm) thick exterior brick wythe
anchored to a backing system with metal ties in such a way that a clear air space is provided between the veneer
and the backing system. The backing system may be wood frame, steel frame, concrete or masonry. By
definition, a veneer wall is a wall having a facing of masonry units, or other weather-resisting, noncombustible
materials, securely attached to the backing, but not so bonded as to intentionally exert common action under load.
The brick veneer is designed to carry loads due to its own weight, no other loads are to be resisted by the veneer.
For many years brick veneer construction was limited principally to wood frame houses. It is now being used on
low-rise commercial and institutional construction and is used frequently for high-rise buildings, especially with
concrete masonry or steel stud backing systems. This Technical Notes discusses the prescriptive design of brick
veneer on wood frame buildings three stories or less in height. Other Technical Notes in this series cover brick
veneer with different backing systems.
The minimum requirements given in this Technical Notes are based on successful past performance of brick
veneer anchored to wood frame systems. The proper design, detailing and construction of anchored brick veneer
walls ensure that these walls function as complete systems. It is important to understand that the failure of any
part of the system, whether in design or construction, can result in improper performance of the entire system.
Satisfactory performance of brick veneer wood frame systems is achieved with: (1) an adequate foundation, (2) a
sufficiently strong, rigid, well-braced backing system, (3) proper attachment of the veneer to the backing system,
(4) proper detailing, (5) the use of proper materials, and (6) good workmanship in construction.

PROPERTIES OF BRICK VENEER
Strength
Factors that affect the strength of brick veneer are the type of brick and mortar used, the span of the veneer and
the backing, the stiffness of the backing, and the tie system. Although the brick veneer is not designed to carry
lateral load, it does carry a proportionate share. In fact, due to the relatively low stiffnesses normally achieved in
wood frame construction, the brick veneer usually carries the majority of any lateral load.
Support
With wood framing the brick veneer must carry its own weight and transfer this weight to a noncombustible
foundation or preservative-treated wood foundation. The weight of brick veneer should not be supported by wood
framing or other types of wood construction, except as noted. Table 1 contains empirical height limitations for
anchored brick veneer supported on a noncombustible foundation. The height of anchored veneer supported on
preservative-treated wood foundations cannot exceed 18 ft (5.49m) above the support. These limits, which are
found in the Building Code Requirements for Masonry Structures (ACI 530 / ASCE 5 / TMS 402-02) and model
building codes, are imposed because of the differences in relative stiffnesses of the brick veneer and the wood
28 http://www.gobrick.com/BIA/technotes/t28.htm
1 of 12 9/13/2009 1:23 PM
frame. Further, differences in movement resulting from wood shrinkage and brick expansion are controlled by
these limits.
TABLE 1
Empirical Height Limitations from Foundation for Anchored Brick Veneer

Height at Plate, ft (m) Height at Gable, ft (m)
30 (9.14) 38 (11.58)



Fire Resistance
Brick veneer wall assemblies can attain fire ratings of up to 2 hr. Figure 1 shows a brick veneer wall assembly with
a 2 hr fire rating. The combustible wood stud must be protected from fire on each side by noncombustible
materials which meet the required fire rating. A 4 in. (100 mm) nominal brick wythe provides a 1 hr fire rating.

2 Hr Fire Rated Brick Veneer Wall Assembly
FIG. 1

Moisture Resistance
Brick veneer wall assemblies are classified as drainage type walls. Walls of this type provide good resistance to
rain penetration. It is essential to maintain the clear air space between the brick veneer and the backing to ensure
proper drainage. Flashing and weep holes work with the air space to provide moisture penetration resistance.
Refer to Technical Notes 7 Series for more information. Brick veneer with wood frame backing has historically
been built with a 1 in. (25 mm) minimum air space. The protection provided by roof overhangs and the relatively
low wall heights aid in reducing water penetration.
Resistance to Heat Transmission
Brick veneer wall assemblies provide resistance to the transmission of heat and capacity insulation. The overall
coefficient of heat transmission, U-value, of these walls can be easily calculated using the procedure given in
Technical Notes 4 or the ASHRAE Handbook, Fundamentals Volume. The mass of the brick veneer provides
capacity insulation. It effectively lowers and delays the peak heating and cooling loads. The overall U-value
obtained for the wall assembly can be adjusted by the capacity insulation correction factor (M factor) given in
28 http://www.gobrick.com/BIA/technotes/t28.htm
2 of 12 9/13/2009 1:23 PM
Technical Notes 4B, Fig. 1. This adjustment of the overall U-value will help the designer to more accurately predict
the performance of the building envelope. The actual performance of brick masonry buildings shows that this
adjustment is very conservative, but it is an improvement over the steady-state assumptions normally used in
calculating heat flow.
Acoustical Properties
Brick veneer wall assemblies reduce sound transmission by several means. The mass of the veneer reduces
sound transmission by absorbing the energy of the sound vibrations. The discontinuity between the brick veneer
and the wood backing prevents vibrations of the exterior brick wythe from directly vibrating the rest of the wall
assembly, thereby retarding sound transmission to the interior. Further, a high percentage of the sound is reflected
by the brick wythe.
Although there are no specific data available on the sound transmission characteristics of brick veneer wall
assemblies, the brick veneer wall system shown in Fig. 1 has an estimated Sound Transmission Class (STC) in
excess of 45. See Technical Notes 5A for more information on the STC.

DESIGN AND DETAILS
Foundations for Brick Veneer
Brick veneer with wood frame backing must transfer the weight of the veneer through the veneer to the foundation.
Typical foundation details for brick veneer are shown in Fig. 2. It is recommended that the foundation or foundation
wall supporting the brick veneer be at least equal to the total thickness of the brick veneer wall assembly. Many
building codes permit a nominal 8 in. (200 mm) foundation wall under single-family dwellings constructed of brick
veneer, provided the top of the foundation wall is corbeled as shown in Fig. 2(c). The total projection of the corbel
should not exceed 2 in. (50 mm) with individual courses projecting beyond the course below not more than
one-third the thickness of the unit nor one-half the height of the unit. The top course of the corbel should not be
higher than the bottom of the floor joist and shall be a full header course.

Typical Foundation Detail
FIG. 2a

28 http://www.gobrick.com/BIA/technotes/t28.htm
3 of 12 9/13/2009 1:23 PM
Typical Foundation Detail
FIG. 2b

Typical Foundation Detail
FIG. 2c

28 http://www.gobrick.com/BIA/technotes/t28.htm
4 of 12 9/13/2009 1:23 PM
Typical Foundation Detail
FIG. 2d

Foundations must extend beneath the frost line as required by the local building code. Design of the foundation
should consider differential settlement and the effect of concentrated loads such as those from columns or
fireplaces. Appropriate drainage must be provided in order to maintain soil bearing capacity and prevent washout.
Brick walls which enclose crawl spaces must have openings to provide adequate ventilation. Openings should be
located to achieve cross ventilation.
Ties
Ties typically used with wood framing are shown in Fig. 3. There should be one tie for every 2 2/3 sq. ft. (0.25
m2 ) of wall area with a maximum spacing of 24 in. (600 mm) o.c. in either direction. The nail attaching a
corrugated tie must be located within 1/2 in. (13 mm) of the bend in the tie. The best location of the nail is at the
bend in the corrugated tie, and the bend should be 90.
Wire ties must be embedded at least 1 in. (38 mm) into the bed joint from the air space and must have at least
5/8 in. (16 mm) cover of mortar to the exposed face. Corrugated ties must penetrate to at least half the veneer
thickness or 1 in. (38 mm) and have at least 5/8 in. (16 mm) cover. Ties should be placed so that the portion
within the bed joint is completely surrounded by the mortar.

28 http://www.gobrick.com/BIA/technotes/t28.htm
5 of 12 9/13/2009 1:23 PM
Unit Ties
FIG. 3

Flashing and Weep Holes
Flashing and weep holes should be located above and as near to grade as possible at the bottom of the wall,
above all openings, and beneath sills. Weep holes must be located in the head joints immediately above all
flashing. Clear, open weep holes should be spaced no more than 24 in. (600 mm) o.c. Weep holes formed with
wick materials or with tubes should be spaced at a maximum of 16 in. (400 mm) o.c. If the veneer continues below
the flashing at the base of the wall, the space between the veneer and the backing should be grouted to the height
of the flashing. Flashing should be securely fastened to the backing system and extend through the face of the
brick veneer. The flashing should be turned up at least 8 in. (200 mm). Typical flashing details are shown in Figs.
2, 4 and 5. Flashing should be carefully installed to prevent punctures or tears. Where several pieces of flashing
are required to flash a section of the veneer, the ends of the flashing should be lapped a minimum of 6 in. (150
mm) and the joints properly sealed. Where the flashing is not continuous, such as over and under openings in the
wall, the ends of the flashing should be turned up into the head joint at least 2 in. (50 mm) to form a dam.
Sheathing
Wood frame construction requires exterior sheathing over the studs. Sheathing may be plywood, OSB,
gypsum sheathing, exterior insulation board, or similar materials. Building paper or housewraps are
placed over the sheathing to limit moisture penetration and air infiltration. See Figs. 1 and 2.

28 http://www.gobrick.com/BIA/technotes/t28.htm
6 of 12 9/13/2009 1:23 PM
Lintel Details
FIG. 4

Lintels, Sills and Jambs
Brick veneer backed by wood frame must always be supported by lintels over openings unless the masonry is
self-supporting. Lintel design information may be found in Technical Notes 17B and 31B. Loose steel, stone or
precast lintels should bear at least 4 in. (100 mm) at each jamb. All lintels should have space at the end of the
lintel to allow for expansion. The clear span for 1/4 in. (6.3 mm) thick steel angles varies between 5 ft (1.5 m) and
a maximum of 8 ft (2.4 m), depending on the size of the angle selected. Steel lintels with spans greater than 8 ft
(2.4 m) may require lateral bracing for stability. The maximum clear span may be restricted by the fire protection
requirements of some building codes. Concrete, cast stone and stone lintels must be appropriately sized to carry
the weight of the veneer.
Reinforced brick lintels are also a viable option. Some of the advantages of reinforced brick lintels are: more
efficient use of materials; built-in fireproofing; elimination of differential movement which may occur with steel
lintels and brick veneer; and no required painting or other maintenance. Typical residential construction details for
a lintel, sill and jamb using wood stud backing are shown in Figs. 4 and 5.
28 http://www.gobrick.com/BIA/technotes/t28.htm
7 of 12 9/13/2009 1:23 PM
Jamb and Sill Details
FIG. 5

Eave Details
A typical residential eave detail is shown in Fig. 6. This detail is suggested for the area at the top of the veneer.
The air space between the top of the brick veneer and wood framing is necessary to accommodate movement.
Larger overhangs and gutters are helpful to keep water from contacting the wall below.

28 http://www.gobrick.com/BIA/technotes/t28.htm
8 of 12 9/13/2009 1:23 PM
Eave Detail
FIG. 6

Movement Provisions
Design provisions for movement which include bond breaks, expansion joints, and joint reinforcement are not
usually required in residential and low-rise brick veneer construction. However, they may be required in specific
situations and the designer should analyze the project to determine such need.
Bond Breaks. Significant differential foundation settlement and horizontal movement may cause cracking in walls
rigidly attached to the foundation. Bond breaks will help to relieve the stresses caused by these movements
between the wall and the supporting foundation. Flashing at the base of the wall between the veneer and the
foundation will provide sufficient break in the bond.
Expansion Joints. Expansion joints to allow for horizontal movement may be required in brick veneer when there
are long walls, walls with returns or large openings. The placement of expansion joints and the materials used
should be in accordance with the information given in Technical Notes 18 Series.
Horizontal Joint Reinforcement
Masonry materials subject to shrinkage stresses, such as concrete masonry, require horizontal joint reinforcement
for control of cracking from such movement. Brick is not subject to shrinkage, therefore horizontal joint
reinforcement is never required in brick masonry for this purpose. It may be beneficial to use limited amounts of
horizontal joint reinforcement in brick veneer for added strength at the corners of openings and at locations where
running bond in the masonry is not maintained.
Horizontal joint reinforcement should be used to add integrity to veneer constructed in locations with intermediate
and higher seismic activity or when the units are laid in stack bond. It may be either single or double wire joint
reinforcement. The wire should engage the veneer ties as shown in Fig. 3(e) in seismically active areas. When
using horizontal joint reinforcement, it must be discontinuous at all movement joints.
Sealant Joints
Exterior joints at the perimeter of exterior door and window frames to be filled with sealant should be formed by
the adjacent materials or be a reservoir type joint. The joint should be no less than 1/4 in. (6.3 mm) nor more than
1/2 in. (12.7 mm) wide and 1/4 in. (6.3 mm) deep. If wider joints are required, the sealant depth should be
one-half of the joint width. A compressible backer rod or sealant bond break tape must be used. Fillet joints are
not recommended, but if used, should be at least 1/2 in. (12.7 mm) across the diagonal. Fig. 7 shows typical
sealant joints. These joints should be solidly filled with an elastic sealant forced into place with a pressure gun. All
joints should be properly prepared before placing sealants. Appropriate primers should be applied as necessary.
Expansion joints must be clear of all material for the thickness of the veneer wythe and closed with a backer rod
and sealant.
28 http://www.gobrick.com/BIA/technotes/t28.htm
9 of 12 9/13/2009 1:23 PM

Sealant Joints
FIG. 7

SELECTION OF MATERIALS
Brick
Brick should conform to ASTM C 62, C 216 or C 652 for Building Brick, Facing Brick and Hollow Brick,
respectively. Grade SW is required where high and uniform resistance to damage caused by cyclic freezing is
desired and where the brick may be frozen when saturated with water. Grade MW may be used where moderate
resistance to damage caused by cyclic freezing is permissible or where the brick may be damp, but not saturated,
when freezing occurs.
The brick selected should have an average initial rate of absorption (suction) of not more than 30 grams per 30
in.2 (1.5 kg/m2) per minute at the time of laying. Units having average initial rates of absorption exceeding this
value may be wetted immediately before they are laid. Alternately, the units may be wetted thoroughly 3 to 24
hours prior to their use so as to allow moisture to become distributed throughout the unit. With either method the
units should be surface dry when laid.
The use of salvaged brick is not recommended. In general, masonry constructed with salvaged brick contains
some weaker and less durable units than masonry constructed with new brick. Salvaged brick and the reasons
against its use are discussed in detail in Technical Notes 15.
Mortars
Mortar materials should comply with the requirements of ASTM C 270 Standard Specification for Mortar for Unit
Masonry. Three types of cementitious materials are permitted: portland cement-lime, mortar cement and masonry
cement. Portland cement-lime and masonry cement mortars made with non-air-entrained materials have greater
strength than those made with air-entrained materials and masonry cement. Proprietary mortar mixes, such as
masonry cements and mortar cements, are widely used because of their convenience and good workability. These
cements usually contain portland cement, ground limestone and additives which provide workability, water
retentivity and air entrainment. See Technical Notes 8 for information on portland cement-lime mortars, mortar
cement mortars and masonry cement mortars.
Type N mortar is suitable for most brick veneer although Type S or Type M may be used. Type S mortar is
recommended where a high degree of flexural resistance is required and may be required in areas of high seismic
28 http://www.gobrick.com/BIA/technotes/t28.htm
10 of 12 9/13/2009 1:23 PM
activity. Type M is recommended where the brick veneer is in contact with earth. For further information on the
selection of mortar see Technical Notes 8B.
Ties
Brick veneer with wood frame backing is supported on the foundation with lateral support provided by the ties and
backing system. The ties must be capable of resisting tension and compression resulting from forces
perpendicular to the plane of the wall. More information on wall ties is found in Technical Notes 44B.
Corrugated steel ties, at least 22 gage, 7/8 in. (22 mm) wide, 6 in. (150 mm) long, as shown in Fig. 3(d) have
historically been used to attach brick veneer to wood frame backing. However, corrugated metal ties are more
susceptible to corrosion than wire ties. Adjustable ties provide better load transfer and permit differential
movement in taller structures. Wire for such ties is either wire gage W1.7, 9 gage, or wire gage W2.8, 3/16 in.
(4.8 mm) diameter. Wire ties should be fabricated from wire conforming to ASTM A 82 Specification for Steel
Wire, Plain, for Concrete Reinforcement. Plate portions of adjustable ties are normally 14 gage in thickness. Steel
used to fabricate plate portions and corrugated ties should conform to ASTM A 366 Standard Specification for
Steel, Carbon, Cold-Rolled Sheet, Commercial Quality.
All tie components must be corrosion resistant. Zinc coating on steel must be at least 1.5 oz per square foot (458
g/m2). This corresponds to ASTM A 153 Standard Specification for Zinc Coating (Hot-Dip) on Iron and Steel
Hardware, Class B-2.
Ties are usually fastened to the wood frame with corrosion-resistant nails that penetrate the sheathing and are
driven a minimum of 1 1/2 in. (38 mm) into the studs.
Flashing and Weep Holes
There are many types of flashing available which are suitable for use in brick veneer walls. Sheet metals, plastics,
laminates or combinations of these have been used successfully. Plastic flashing should be at least 30 mil thick.
Asphalt impregnated felt (building paper) or an air-infiltration barrier is not acceptable for use as flashing. These
materials serve other purposes in the wall assembly. Building paper is applied as a moisture barrier to the
sheathing. Air-infiltration barriers function as their name implies and may also serve as a moisture barrier.
Selection of flashing is often determined by cost; however, it is recommended that only superior materials be
used, as replacement in the event of failure is exceedingly expensive.
Weep holes can be made in several ways. Some of the most common ways are leaving head joints open, using
removable oiled ropes or rods, using plastic or metal tubes, or using rope wicks. There are also plastic or metal
vents which are installed in lieu of mortar in a head joint. Clear openings without obstructions produce the best
weep holes. For further discussion on flashing and weep holes see Technical Notes 7A.
Sheathing Materials
Minimum thickness requirements for exterior gypsum sheathing, plywood, OSB and similar wood sheathing
materials should are stated in the model building codes. Building paper used to cover exterior sheathing should be
No. 15 asphalt saturated felt conforming to the requirements of ASTM D 226, Type I.
Horizontal Joint Reinforcement
Horizontal joint reinforcement should meet the requirements of ASTM A 951. It should have a corrosion-resistant
coating which conforms to ASTM A 153, Class B-2.
Lintel Materials
Lintels may be reinforced brick masonry, reinforced concrete, stone or steel angles. Reinforcement for reinforced
brick masonry lintels should be steel bars manufactured in accordance with ASTM A 615, A 616 or A 617, Grades
40, 50, or 60 and should be at least No. 3 bar size. Joint reinforcement can also be used in reinforced brick
masonry lintels.
Steel for lintels should conform to ASTM A 36 Standard Specification for Structural Steel. Steel angle lintels should
be at least 1/4 in. (6.3 mm) thick with a horizontal leg of at least 3 1/2 in. (89 mm) for use with nominal 4 in. (100
mm) wide brick veneer, and 3 in. (75 mm) for use with nominal 3 in. (75 mm) wide brick veneer. Steel lintels
should be painted before installation.
Sealants
There are numerous types of sealants available that are suitable for use with brick veneer. The material selected
should be flexible and durable. Superior sealants may have a higher initial cost, but their high flexibility and
28 http://www.gobrick.com/BIA/technotes/t28.htm
11 of 12 9/13/2009 1:23 PM
increased durability result in savings of maintenance costs due to the reduced frequency of reapplication. Good
grades of polysulfide, butyl or silicone rubber sealants are recommended. Oil-based caulking compounds are not
recommended since most lack the desired flexibility and durability, see Technical Notes 7A. Regardless of the
type of sealant chosen, proper primers and backer rods must be selected. Follow the recommendations of the
sealant manufacturer.

CONSTRUCTION
Protection of Materials
Masonry. Prior to and during construction, all materials should be stored off of the ground to prevent
contamination by mud, dust or other materials likely to cause stains or defects. The masonry materials should also
be covered for protection against the elements.
To limit water absorption, it is recommended that all brick masonry be protected by covering at the end of each
workday and for shutdown periods. The cover should be a strong, weather-resistant membrane securely attached
to and overhanging the brickwork by at least 24 in. (600 mm). Partially completed masonry exposed to rain may
become so saturated with water that it may require months after the completion of the building to dry out. This
saturation may cause prolonged efflorescence. See Technical Notes 23 Series for more information.
Flashing. Flashing materials should be stored in places where they will not be punctured or damaged. Plastic and
asphalt coated flashing materials should not be stored in areas exposed to sunlight. Ultraviolet rays from the sun
break down these materials, causing them to become brittle with time. Plastic flashing exposed to the weather at
the site for months before installation should not be used. During installation, flashing must be pliable so that no
cracks occur at corners or bends.
Workmanship
Good workmanship is as essential in constructing brick veneer as it is in all types of brick masonry construction. All
joints intended to receive mortar, including head joints with hollow brick, should be completely filled. Joints or
spaces not intended to receive mortar should be kept clean and free of droppings. Courses of brick laid on
foundations or lintels must have at least two-thirds of the brick width on the support.
The joints should be tooled with a jointer as soon as the mortar has become thumbprint hard. The types of joints
recommended for exterior use with brick veneer are concave, "V" and grapevine. These joints firmly compact the
mortar against the edges of the adjoining brick. Other joints are not recommended because they do not provide
the necessary resistance to moisture penetration. See Technical Notes 7B Revised for further information.
It is essential when constructing brick veneer, to keep the 1 in. (25.4 mm) minimum air space between the veneer
and the backing clean and free of all mortar droppings, so that the wall assembly will perform as a drainage wall.
If mortar blocks the air space, it may provide a bridge for water to travel to the interior. In addition, all flashing,
weep holes, ties and other accessories must be properly installed and kept clean.

SUMMARY
This Technical Notes is concerned primarily with the prescriptive design and conventional application of anchored
brick veneer in new wood frame buildings limited to three stories in height. Other Technical Notes in this series
consider brick veneer applied to existing structures, brick veneer with different backing materials for mid-and
high-rise structures, and adhered thin brick veneer.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Industry Association. The information and recommendations
contained herein must be used in conjunction with good technical judgment and a basic understanding of the
properties of brick masonry. Final decisions on the use of the information contained in this Technical Notes are not
within the purview of the Brick Industry Association, and must rest with the project architect, engineer and owner.

REFERENCES
For more detailed information on materials, design and construction procedures, the individual Technical Notes
referred to herein should be consulted.
28 http://www.gobrick.com/BIA/technotes/t28.htm
12 of 12 9/13/2009 1:23 PM
2008 Brick Industry Association, Reston, Virginia Page 1 of 10
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
28A
April
2008
Adding Brick Veneer to Existing Construction
Abstract: This Technical Note presents information on adding anchored brick veneer and thin brick adhered veneer to
existing construction. Considerations and recommendations for design, detailing, material selection and construction specific
to retrofitting existing walls with brick veneer are presented. Other Technical Notes are referenced for general brick veneer
construction information not specific to the retrofit of existing construction.
Key Words: adhered veneer, anchored veneer, anchors, brick veneer, design, existing construction, flashing, retrofit,
reveneering, thin brick, walls, weeps.
Preparation:
Inspect and repair existing siding as necessary to act as
a water resistive barrier
Where existing siding cannot be readily repaired, wrap
existing siding with a new water resistive barrier
Where neither option above is desired, remove existing
siding and inspect and repair existing water resistive
barrier and sheathing as necessary, installing a new
water resistive barrier where none currently exists
Anchored Brick Veneer:
Provide drainage wall details in accordance with
Technical Note 7, materials in accordance with Technical
Note 7A and workmanship in accordance with Technical
Note 7B
Provide a minimum nominal 1 in. (25.4 mm) air space
behind brick veneer
Support:
- Bear veneer on existing or extended footing or angle
fastened to existing foundation wall
- For angle support, provide corrosion resistant angle;
fasten angle to existing foundation wall with anchors
sized and placed in accordance with Table 2
- Where no footing or foundation wall exists, consult a
design professional
Anchorage:
- Select type and length of anchor based on existing wall
type
- Place anchors to penetrate existing sheathing and
securely fasten into existing structural members
Veneer Construction:
- Lap flashing at top of openings and base of veneer with
existing exterior finish
- Lap flashing at bottom of openings with existing sill
- Provide open head joint weeps immediately above
flashing at 24 in. (610 mm) o.c.
Edge Details:
- At perimeter of veneer openings, install molding and
sealant to close air space
- At top of veneer, provide at least in. (19.1 mm) clear-
ance to bottom of existing soffit, sealed at exterior edge
- Use proper construction techniques
Thin Brick Adhered Veneer:
Provide barrier wall details in accordance with Technical
Note 28C
Attach to existing wall with mortar or adhesive, or
proprietary clip anchoring system provided by thin brick
system manufacturer
INTRODUCTION
Brick veneer provides superior performance with many properties desired by designers, contractors and property
owners, such as attractive appearance, high fire resistance, high resistance to water penetration, low thermal trans-
mission rate, low maintenance and increased resale value. For an existing home or building clad in other materials,
adding a brick veneer exterior can result in a significant improvement, such as that shown in Photo 1 and Photo 2.
SUMMARY OF RECOMMENDATIONS:
Photo 1
Before
Photo 2
After
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 2 of 10
Air Space
Existing
Framing
Existing
Siding
New Brick
Veneer
New Flashing
and Weeps
Existing Continuous Foundation
Figure 1
New Brick Veneer on Existing Construction
Existing
Footing
New Brick
Veneer
Full Collar
Joint
Existing
Basement
Wall
Figure 2
Brick Veneer on Existing Foundation
New
Footing
New Brick
Veneer
Full Collar
Joint
New Bond
Break
Material
Existing
Footing
Existing
Basement
Wall
Figure 3
Brick Veneer on Footing Extension
Anchored brick veneer is a non-loadbearing
component, supported by a foundation and attached
to the structure by anchors fastened to studs or
embedded within masonry. Behind the veneer, an
air space and water-resistive barrier direct water
downward to flashing and weeps, providing an effective
drainage wall system. The brick veneer bears on an
existing or extended continuous footing, as shown in
Figure 1, Figure 2 and Figure 3, or angles attached to
an existing foundation wall, as shown in Figure 4.
Thin brick adhered veneer is attached to and supported
by the backing wall or framing with lath and plaster,
as shown in Figure 5, adhesive or a proprietary clip
system. For substrate walls with framing, a water
resistive barrier separates the veneer from the wall
framing. Mortar behind and surrounding the brick units
create a barrier wall system to manage water.
Existing Foundation Wall
New Steel
Angle
Wedge Anchor
Flashing and Weeps
Corrugated Anchor
New Brick Veneer
Existing Siding
Figure 4
Brick Veneer on Steel Support
Existing Siding
Removed
Existing Sheathing
Repair or Replace
As Necessary
New Felt Paper
Wire Lath
Cementitious Stucco
Thin Brick Veneer
Set into Stucco
Using Thick Set Method
Figure 5
Thin Brick Adhered Veneer on Existing Wood Studs
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 3 of 10
Building Plans Without Brick Veneer
Brick veneer may be added to drawings and plans where it currently is not included. For anchored brick veneer,
the existing design can be modified. For an existing design with a continuous footing, the details and explanation
presented herein are directly applicable. Where an existing design includes pier foundations without continuous
footings, such as with most modular homes or in coastal construction, brick veneer may likely require the addition
of a new footing and foundation wall. Guidance for various types of foundation wall construction can be found in
Technical Note 26, Technical Note 17L and in the Technical Note 21 Series. Footing design should be coordinated
with a contractor and/or design professional.
For thin brick adhered veneer, the thin brick simply may be substituted for other siding materials. Due to the lighter
weight of thin brick, further modification to the existing design is not necessary.
Existing Construction
Generally, the existing exterior walls of a building, whether wood stud, steel stud or masonry, are sufficient as
backing for anchored brick veneer or thin brick adhered veneer. Where exterior framing may allow excessive
deflections, such as steel stud backing or premanufactured home walls, stiffen the backing as necessary to meet
current code requirements for deflection limits.
In addition to structural preparations, the existing walls may need modification to provide a water resistive barrier.
Where existing siding materials cannot provide this barrier, either install a new barrier outside the existing siding or
remove the siding. If the siding is removed, either repair the barrier behind the siding as necessary or install a new
water resistive barrier. Existing sheathing should not be used alone as the water resistive barrier.
This Technical Note is one in a series focusing on brick veneer. It addresses adding brick veneer to an existing
building and focuses on anchored brick veneer. For more detailed information on thin brick adhered veneer, refer
to Technical Note 28C. Other Technical Notes in the series are devoted to various brick veneer systems in new
construction and provide additional useful information.
PROPERTIES OF BRICK VENEER CONSTRUCTION
The following properties are true of virtually all types of brick veneer. With thin brick adhered veneer, some
properties, such as fire resistance, may be somewhat less than those of full thickness units.
Aesthetics
Brick is a traditional sign of quality construction that will increase the value of an existing home or commercial struc-
ture. Brick is available in a large variety of sizes, colors, textures and coatings. Multiple bond patterns, colored mor-
tars and details are available to create unique looks. For further information on sizes and patterns, refer to Technical
Note 10 and Technical Note 30.
Thermal Performance
Adding brick veneer to an existing wall can increase thermal performance. Brickwork has a high thermal mass,
storing and releasing heat slowly. As a result, walls with brick veneer require less insulation than walls without brick.
Closed-cell rigid board insulation can be added over the existing wall behind the veneer. For further information
regarding the thermal resistance of brick assemblies, refer to the Technical Note 4 Series.
Fire Resistance
Brickwork is non-combustible and excellent fire resistance. A nominal 3 or 4 in. (76 or 102 mm) brick wythe has
a 1-hour fire resistance rating, significantly reducing the chance of fire spread [Ref. 4]. For additional information,
refer to the Technical Note 16 Series.
Acoustics
Brick veneer walls reflect a large portion of sound waves. The mass of the brickwork absorbs another portion of
sound energy. For anchored brick veneer, the air space separates the brickwork from the existing wall, further
reducing sound transmission. For additional information on sound transmission, refer to Technical Note 5A.
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 4 of 10
Moisture Resistance
Anchored brick veneer construction is separated from the existing wall by an air space, creating a drainage wall.
If wind-driven rain penetrates the brickwork, the air space drains water down the back face of the brickwork to
flashing and weeps. A water resistive barrier acts as a second line of defense against water penetration. Thin brick
adhered veneer uses the thickness of the brick and mortar backing as a barrier wall to resist water penetration. For
additional information regarding water penetration resistance, refer to the Technical Note 7 Series.
DESIGN AND DETAILING
Proper design and detailing of brick veneer added to a completed design or existing construction are essential
to ensure proper performance. Design and detailing considerations include supporting the weight of the veneer,
attachment of the veneer to the existing structure, drainage and movement provisions, and details around
openings.
Supporting Brick Veneer
The height limitations for anchored brick veneer above foundations are based on the history of successful
performance and depend on both the support offered by the backing and the anticipated seismic or other
lateral loads. Table 1 is based on the 2006 International Residential Code (IRC), Tables 703.7(1) and 703.7(2)
[Ref. 2]. Existing siding or sheathing may need to be upgraded to meet the minimum sheathing thickness and
coverage requirements of the code tables. The same height limits apply to nonresidential construction, except
that in Seismic Design Categories E and F, the veneer is required to be supported at each story. Refer to the
IRC, the International Building Code, and the MSJC Code for additional requirements for residential (IRC) and
nonresidential (IBC and MSJC) construction in Seismic Design Categories C through F [Refs. 1, 2, 5].
When adding anchored brick veneer, its weight may be supported directly on either existing or new concrete
foundations. Alternatively, where existing concrete or masonry foundation walls provide sufficient strength, the
veneer may be supported by steel angles anchored to the existing foundation walls.
TABLE 1
Height Limitations for Anchored Brick Veneer
Seismic Design
Category
Type of Backing
Empirical Height Limitations
Max, Stories Height at Plate, ft (m) Height at Gable, ft (m)
A, B or C
1
Wood stud 3 30 (9.14) 38 (11.58)
Steel stud 2 30 (9.14) 38 (11.58)
Concrete or masonry No specific limit
D
0
2
Wood stud 3
4
30 (9.14)
4
38 (11.58)
Steel stud Not permitted
Concrete or masonry No specific limit
D
1
2
Wood stud 3 20 (6.10)
5
28 (8.53)
5
Steel stud Not permitted
Concrete or masonry No specific limit
D
2
3
Wood stud 2 20 (6.10)
5
28 (8.53)
5
Steel stud Not permitted
Concrete or masonry No specific limit
1. Veneer to be maximum 5 in. (127 mm) nominal thickness and maximum 50 psf (245 kg/m) installed weight.
2. Veneer to be maximum 4 in. (102 mm) nominal thickness and maximum 40 psf (196 kg/m) installed weight.
3. Veneer to be maximum 3 in. (76 mm) nominal thickness and maximum 30 psf (147 kg/m) installed weight.
4. Maximum height of one- and two- story veneer limited to 20 ft (6.10 m) unless bottom 10 ft (3.05 m) is backed by concrete or masonry.
5. Maximum height may be increased to 30 ft (6.10 m) and 38 ft (11.58 m) in the gable if the bottom 10 ft (3.05 m) is backed by concrete or
masonry.
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 5 of 10
Concrete Footings
Where possible, support new brickwork with an existing concrete footing, as shown in Figure 2. If an existing
footing is not wide enough to support the brick wythe (minimum bearing of two thirds of the thickness of the
wythe), the existing footing can be extended by placing new concrete at the same depth as the existing footing, as
depicted in Figure 3. A bond break is recommended between the existing and new footing to allow for expected
slight differential movements. Provide reinforcement in any new concrete footing as required by the building code.
Coordinate any modification to the existing footing with the local building official.
Steel Angles
An alternate method of supporting the brick veneer is by attaching a continuous corrosion-resistant steel angle to
the existing foundation or basement wall. The preferred location of the horizontal leg of the angle is at or slightly
above grade. If the angle is to be placed below grade and above the frost line, the space beneath the angle can
be backfilled with freely draining granular material and a compressible pad to limit frost heave and to protect the
angle from displacement. It is recommended that the angle be attached to existing basement or foundation walls
constructed of concrete or masonry. Anchorage to concrete or masonry walls can be achieved either by anchors,
as depicted in Figure 4, or by using through-bolts. When through-bolting, seal the annular space around the shaft
of the bolt to prevent water penetration. When using mechanical anchors secured in poured concrete or filled
concrete masonry units, use maximum anchor spacing as defined in Table 2. Note that the angle size used will
depend on the proposed distance from face of brick back to face of foundation wall. The horizontal leg of the angle
must be long enough to allow a minimum of two thirds of the brick thickness to bear on the angle. Attach angles
through holes centered at two-thirds the height of the angle using non-corrosive shims as necessary.
This method of supporting veneer with retrofit steel angles may not be suitable for some applications and may
require review by a qualified design professional. Loads applied to the angle and foundation wall should be
carefully considered, as well as the strength of the foundation wall itself. Special consideration should be given
to the eccentricities of the applied loads, especially when acting in conjunction with existing soil backfill loads. In
general, this method of support should be confined to one-story structures where the total height to the plate does
not exceed approximately 12 ft (4.3 m).
Attachment
The brick veneer must be securely attached to the existing construction throughout its height. When using
adjustable two-piece W 1.7 (MW11) wire or 22-gage corrugated anchors, provide one anchor for each 2.67 sq
ft (0.25 m) of wall area. For other anchor types, provide one anchor for each 3.5 sq ft (0.33 m) of wall area.
Depending on the applicable code (IRC or IBC), the maximum spacing of anchors cannot exceed 24 in. (610 mm)
horizontally or vertically for residential construction and cannot exceed 32 in. (813 mm) horizontally or 18 in. (457
mm) vertically for nonresidential construction. This spacing applies above and below grade.
TABLE 2
Support Angle Anchor Bolt Spacing, in. (mm)
1
Brick Height, ft. (m)
Drop-In Anchors Wedge Anchors or Through-Bolts
2
in. (12.7 mm)
diameter
in. (15.9 mm)
diameter
in. (12.7 mm)
diameter
in. (15.9 mm)
diameter
4 (1.2) 42 (1067) 48 (1219) 48 (1219) 48 (1219)
6 (1.8) 28 (711) 48 (1219) 42 (1067) 48 (1219)
8 (2.4) 20 (508) 36 (914) 32 (813) 45 (1143)
10 (3.0) 18 (457) 28 (711) 24 (610) 36 (914)
12 (3.7) 14 (356) 24 (610) 21 (533) 30 (762)
Min. embedment, in. (mm) 2 (51) 2 (60) 2 (57) 2 (70)
Min. edge distance, in. (mm) 2 (64) 3 (79) 2 (64) 3 (79)
1. Assumes steel equal leg angles, in. (9.5 mm) thick, maximum leg dimension of 5 in. (127 mm), used in 10 ft. (3.0 m) sections with in.
(12.7 mm) gap between sections.
2. Minimum embeddment not applicable to through-bolting.
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 6 of 10
Supporting Lateral Loads
Lateral loads push or pull a wall surface due to
forces from wind and earthquakes. The addition
of brick veneer, when properly attached, does not
affect the ability of the existing construction to resist
wind pressures. The wind loads on the brickwork are
considered to be transferred through the anchors into
the existing walls, just as if the veneer had not been
added. Seismic loads, which are based on wall stiffness
and weight, among other factors, can be significantly
affected by the addition of the veneer. For all buildings
in areas defined by the building code as Seismic
Design Categories A and B, as well as detached one-
and two-family homes in Seismic Design Category C,
there are no requirements associated with the addition
of the veneer. The height limitations in Table 1 were
established based on the seismic resistance of wood
stud framing. In all other cases, a local building official
and/or design professional should be consulted to
determine if modification of the existing exterior walls
is needed to accommodate the seismic loads induced
by the new veneer.
Flashing and Weeps
Flashing details, similar to those depicted in Figure 4,
Figure 6 and Figure 7, are essential to brick veneer
construction. In order to divert the moisture out of
the air space through the weeps, install continuous
flashing at the bottom of the air space and above
grade. Where the veneer continues below grade,
completely fill the space between the veneer and the
existing construction below the flashing with mortar
or grout. Install flashing at the heads and sills of all
openings and wherever the air space is interrupted.
Turn the back of the flashing up a minimum of 8 in.
(203 mm) such that the top edge of the flashing is
covered by or sealed to the existing siding. The front
of the flashing should extend to the face of the brick
veneer. Where the flashing is not continuous, such
as at heads and sills, the ends should be turned up
approximately 1 in. (25.4 mm) to form an end dam.
Locate weeps in the head joints immediately above
all flashing. The maximum recommended spacing of
open head joint weeps is 24 in. (610 mm) on center.
When wick materials are used in the weeps, the
maximum recommended spacing of weeps is 16 in.
(406 mm) on center. Additional discussion of flashing
and weeps may be found in the Technical Note 7
Series.
Framing Around Openings
Typically, openings in an added veneer are
constructed similarly to those in new construction, with
existing window trim and details left in place behind
Figure 6
Base Detail
Full Collar Joint
Flashing and Weeps
Corrugated Anchor
New Brick Veneer
Existing Siding
Wire Anchor
Figure 7
Window Head and Sill Details
Corrugated Metal Anchor
Steel Angle Lintel
Weep
Existing Siding
Flashing
Sealant
New Molding
(a) Window Head/Lintel
Nominal 1 in.
(25.4 mm) Required
Air Space
Existing Sill
Sealant
New Brick Sill
Weep
New Brick Veneer
(b) Window Sill
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 7 of 10
the new veneer. At the window ledge, the new brick sill is made sufficiently deep to span the new air space and
tuck under the existing sill, as shown in Figure 6. Maintain a minimum in. (6.4 mm) gap below the existing sill
to allow for potential brick expansion. At the top and sides of the opening, install a new molding deep enough to
cover the air space, as depicted in Figure 7 and Figure 8.
Steel angles commonly used as lintels support veneer over door and window openings. Reinforced brick masonry
or precast concrete are also alternatives. Steel angle lintels must bear a minimum of 4 in. (102 mm) onto adjacent
masonry. For further information on the design, detailing and material selection of various types of lintels, refer to
Technical Notes 17H and 31B.
Top of the Veneer
A typical detail for the top of the brick veneer at an existing eave is shown in Figure 9. Maintain a minimum in.
(19.1 mm) clear space between the top of the last course of brick and the bottom of the soffit. Cover this space
with a new molding strip and sealant or otherwise protected from moisture. If the eave is insufficient to fully cover
the top of the veneer, extend it to protect the top of the brick.
Movement Provisions
Differential movements due to temperature, moisture, shrinkage, and creep are ordinarily insufficient in small
brick veneer buildings to require that movement joints or other provisions be installed. For large structures, such
as commercial buildings and large single-family houses, the design should include considerations of potential
differential movements and proper details to accommodate them. This is especially true for new veneer around
existing second story windows. Design and details for differential movement may include expansion joints, flexible
anchorage, joint reinforcement, bond breaks, and sealants. These items and their applications are discussed in the
Technical Note 18 Series, Technical Note 28 and the Technical Note 21 Series.
MATERIALS
The proper selection of quality materials is essential to the satisfactory performance of a brick veneer wall
assembly. No amount of design, detailing or construction can compensate for the improper selection of materials.
Brick
Use nominal 3 or 4 in. (76 or 102 mm) thick brick, conforming to the requirements of ASTM C216, Standard
Specification for Facing Brick (Solid Masonry Units Made from Clay or Shale), or ASTM C652, Standard
Specification for Hollow Brick (Hollow Masonry Units Made from Clay or Shale) [Ref. 3]. Grade SW brick is
recommended for all veneer applications because the brick wythe is isolated from the remainder of the wall by the
air space, thus exposing it to the maximum temperature extremes.
Figure 8
Window Jamb Detail
New Brick Veneer
New Brick Sill
Existing Siding
New Molding
Sealant
Figure 9
Eave Detail
Corrugated Metal Anchor
New Brick Veneer
Existing Siding
New Molding
or Sealant
Air Gap, Min. 3/4 in.
(19.1 mm) Recommended
Existing
Soffit
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 8 of 10
Generally, salvaged brick are not recommended since their original properties cannot be determined and they
may not provide the strength and durability required for satisfactory performance. For further information on
salvaged brick, refer to Technical Note 15.
Mortar
Selection of an appropriate mortar helps to ensure durable brickwork that meets performance expectations.
Mortar Type and mortar material selection should consider multiple aspects of a project, including design, brick
or masonry materials, exposure and required level of workmanship. Improper mortar selection may lead to lower
performance of the finished project. Mortars are classified by ASTM C270, Standard Specification for Mortar for
Unit Masonry, into four Types: M, S, N and O. These four Types of mortar may be made with portland cement,
masonry cement, mortar cement or blended cements, some of which are combined with hydrated lime. Type N
mortar is recommended for most brick veneer. Type M portland cement-lime mortar is recommended for brick
veneer below grade, where the brickwork is in contact with earth. Mortars for brick masonry are discussed in
Technical Note 8 Series.
Veneer Anchors
The type of anchor system used with brick veneer
depends on the construction of the existing wall.
Corrugated metal anchors are permitted to be used
with wood frame backing. Metal wire anchors are
required for other backing systems. Several types of
anchors that may be used in brick veneer applied to
existing construction are shown in Figure 10.
Corrugated Metal Anchors. Install corrugated
anchors that are at least 22 gage, in. (22.2 mm)
wide, 6 in. (152 mm) long and corrosion-resistant.
Corrugated metal anchors should comply with ASTM
A36, Specification for Carbon Structural Steel, or
ASTM A 1008/A 1008M, Specification for Steel, Sheet,
Cold-Rolled, Carbon, Structural, High-Strength Low-
Alloy, and High-Strength Low-Alloy with Improved
Formability.
Metal Wire Anchors. Use wire anchors that are at least 9 gage and corrosion-resistant. Metal wire ties should
comply with ASTM A82, Specification for Steel Wire, Plain, for Concrete Reinforcement.
Corrosion Resistance. Corrosion resistance is usually provided by a zinc coating, or by using stainless steel. To
ensure adequate resistance to corrosion, coatings or materials should conform to the following standards:
Zinc Coatings -
ASTM A123 or A153 Class B (for sheet metal anchors) or 1.50 oz/ft (458 g/m) (for wire anchors)
Stainless Steel -
ASTM A240 (for sheet metal anchors)
ASTM A480 (for stainless steel sheet metal)
ASTM A580 (for wire anchors)
Anchor Fasteners
The existing wall construction will influence the type of fastener used to attach anchors. Some systems have
manufacturer-specific attachment hardware for each type of backing. Where no manufacturer requirements are
given, the following guidelines apply.
Wood Frame. Use corrosion-resistant screws or nails to attach corrugated metal anchors to wood frame
construction. Use minimum No. 10 screws and 8d nails long enough to penetrate at least 1 in. (32 mm) into the
wood studs after passing through existing siding and sheathing.
Base & Vee Wire Corrugated Metal
Eye & Pintle Wire Wire & Screw
Figure 10
Veneer Anchors
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 9 of 10
Metal. Use corrosion-resistant, self-tapping metal screws, No. 10 minimum, to attach the wire receiver or strap to
metal construction. The screws should penetrate at least in. (12.7 mm) into the metal.
Concrete or Masonry. There are several methods of attaching the metal wire anchors to existing concrete
or masonry walls. Attach anchors with minimum in. (6.4 mm) lag bolts and expansion shields, minimum 8d
masonry nails, or minimum
3
16 in. (4.8 mm) masonry screws. Use corrosion-resistant fasteners and anchors
sufficiently embedded to provide the necessary capacity to resist lateral loading.
Other Wall Types. For pole buildings, pre-manufactured metal buildings, or other forms of construction where
structural members are widely spaced and lateral load resistance is provided mainly by the exterior skin,
coordinate the anchorage type and layout with a design professional.
Steel Angles
Steel for angles supporting new brick veneer at the foundation wall should conform to ASTM A36 and be treated or
coated for corrosion resistance. Bolts or other fasteners should also be corrosion-resistant. Steel angles for lintels
should be a minimum in. (6.4 mm) thick with at least 3 in. (76 mm) legs made of steel conforming to ASTM A36.
For information on steel lintels for brick masonry, refer to Technical Note 31B.
Flashing and Weeps
Flashing materials for use with brick veneer may be plastic or rubber membranes, sheet metals or a combination
of these materials. Selection of superior flashing materials is recommended since replacement in the event of
failure will be costly and difficult, if not impossible. Asphalt-impregnated felt paper is not recommended as a
flashing material. Open head joint weeps are recommended. For a more information on flashing and weeps, refer
to the Technical Note 7 Series.
CONSTRUCTION
Supports
Footings. Supporting brick veneer on an augmented or existing footing requires excavation. The excavation must
be sufficiently wide for the mason to work and sufficiently stepped, braced or shored to avoid collapse. Remove
loose soil and debris from the existing footing with a brush prior to placement of masonry.
Angles. When constructing brick veneer on continuous corrosion-resistant steel angles, lay the first course of brick
in a mortar setting bed. This provides a means to compensate for any variations and misalignment of the steel
angles.
Installing Additional Insulation
Applying brick veneer over existing construction offers an opportunity to better insulate the existing exterior walls.
The insulation materials used should comply with the criteria discussed in Technical Note 21A.
Rigid insulation may be installed directly over the existing finish prior to constructing the new brick veneer. Maintain
a minimum 1 in. (25.4 mm) air space between the brick veneer and the rigid insulation. If the existing wood or
metal framing contain little or no insulation, the existing siding of the wall may be removed to install insulation
within the wall. Replace materials removed from the existing wall with appropriate new materials to provide a
water-resistant barrier. When insulation is added, wire tie veneer anchors are required instead of corrugated
anchors due to the width of the cavity.
Workmanship
Good workmanship is necessary to achieve satisfactory performance of brick veneer. For information on
workmanship, refer to Technical Note 7B.
Mortar Joints. Completely fill all bed and head joints with mortar. Keep clean and free of mortar and mortar
droppings any locations not intended to receive mortar, such as air spaces, weeps and expansion joints. Tool
mortar joints to enhance the water resistance of the wall by consolidating the mortar. Joints should be properly
tooled when the mortar is thumbprint hard with a jointer tool slightly larger than the joint. Concave, V or
grapevine joints are recommended for the most water-resistant brickwork.
www.gobrick.com | Brick Industry Association | 28A | Adding Brick Veneer to Existing Construction | Page 10 of 10
Flashing and Weeps. Securely attach flashing to the existing wall with its top edge overlapped by the existing
siding. Extend the flashing to the face of the brick veneer and install weeps immediately above all flashing.
Anchor Placement. Embed anchors a minimum of 1 in. (38 mm) into the bed joints and completely surround
with mortar.
Sealants. Provide sealant joints at the perimeter of exterior door and window frames not less than in. (6.4 mm)
nor more than in. (12.7 mm) wide. Remove old sealant, dirt, debris, loose paint or coatings to a minimum depth
of in. (19.1 mm). Prime joints before placing sealant. Apply the sealant with a pressure gun.
Protection
As with any brick masonry construction, protect materials from weather before and during construction. Store brick
and mortar materials above the ground and under cover. Store flashing, anchors and other components indoors or
in a shed or trailer, or otherwise protected from weather. During construction, protect partially completed walls by
securely attaching a strong, weather-resistant membrane to the existing structure and allowing it to overhang the
brickwork by at least 2 ft (0.61 m). This will help keep excessive moisture out of the wall and materials, decreasing
the possibility of efflorescence and other deleterious effects.
MAINTENANCE
Most brickwork is virtually maintenance-free. If properly designed, detailed and constructed, minimal brickwork
maintenance is required. However, brick veneer added to existing wall systems should be inspected periodically
to ascertain performance and identify any potential problems. Inspections are recommended on a seasonal basis
or on an annual basis at a minimum. Such inspections should address sealant joints, plumbness of the wall,
cracking, etc., to identify repairs and corrections before severe issues develop. For additional information regarding
maintenance, refer to Technical Note 46.
SUMMARY
This Technical Note provides the basic information required to properly select materials, design, detail, and
construct brick veneer over existing construction. Added veneer relies on proper support at its base and
anchors to the existing framework of the building. Intact existing siding materials can remain in place, with water
penetration resistance provided by a new air space, water resistive barrier and flashing details. By following these
recommendations, existing buildings can attain the aesthetic, thermal, acoustic and fire resistive benefits of brick
veneer.
The information and suggestions contained in this Technical Note are based on the available
data and the experience of engineering staff and members of the Brick Industry Association.
This information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of the
information discussed in this Technical Note are not within the purview of the Brick Industry
Association, and must rest with the project architect, engineer and owner.
REFERENCES
1. 2006 International Building Code, International Code Council, Inc., Country Club Hills, IL, 2006.
2. 2006 International Residential Code, International Code Council, Inc., Country Club Hills, IL, 2006.
3. Annual Book of Standards, Vol. 04.05, ASTM International, West Conshohocken, PA, 2007
C216-07a Standard Specification for Facing Brick (Solid Masonry Units Made from Clay or Shale)
C652-07 Standard Specification for Hollow Brick (Hollow Masonry Units Made from Clay or Shale)
4. Borchelt, J.G., and Swink, J., Fire Resistance Tests of Brick Veneer/Wood Frame Walls, 14th
International Brick/Block Masonry Conference, Sydney, Australia, 2008.
5. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
Brick Veneer/Steel Stud Walls
Abstract: This Technical Note addresses the considerations and recommendations for the design, detailing, material selection
and construction of brick veneer/steel stud walls. This information pertains to behavior of the veneer and steel studs, differential
movement, anchors, air space, detailing, selection of materials and construction techniques.
Key Words: anchors, brick veneer, design, elastic properties, flashing, masonry, permeability, stability, steel studs, stiffness,
walls, weeps.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
28B
December
2005
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Veneer Height:
Maximum veneer height permitted to be supported on
foundation is 30 ft (9.14 m) to top of wall or 38 ft (11.58 m)
to top of gable unless wall is rationally designed
Support veneer above this height by shelf angle or other
means for each story
Air Space:
2 in. (51 mm) minimum air space recommended; 1 in.
(25.4 mm) minimum air space required
4
1
/2 in. (114 mm) maximum distance required between
back of brick veneer and steel framing unless anchors are
rationally designed
Flashing:
Do not stop flashing behind the face of the brickwork
Place flashing at all points where air space is interrupted
Extend flashing vertically up the backing to 8 in (203 mm)
minimum height
Lap flashing to 4 in. (102 mm) minimum height under
water-resistant barrier or behind sheathing above grade
Install base flashing minimum 6 in. (152 mm) above grade
Turn up flashing ends into head joint a minimum of 1 in.
(25.4 mm) to form end dam
Weeps:
Open head joint weeps spaced at no more than 24 in.
(610 mm) o.c. recommended
Most building codes permit weeps no less than
3
/16 in. (4.8
mm) diameter and spaced no more than 33 in. (838 mm)
o.c.
Wick and tube weep spacing recommended at no more
than 16 in. (406 mm) o.c.
Anchors:
Corrugated anchors not permitted with steel stud backing
Minimum W1.7 (9 gage, MW11) adjustable wire anchors,
hot-dipped galvanized, two-piece per ASTM A 153 Class
B-2
Minimum one anchor per 2
2
/3 ft
2
(0.25 m
2
) of wall area
(Note: some building codes may require one anchor per
2 ft
2
(0.19 m
2
))
Vertical spacing: maximum 18 in. (457 mm) o.c.
Horizontal spacing: maximum 32 in. (813 mm) o.c.
Securely attach anchors to the steel studs through the
sheathing, not to the sheathing alone
For high wind and seismic areas, see Anchors section for
anchor placement requirements
Shelf Angles and Lintels:
Shelf angles located above the height limit (see Veneer
Height) may support no more than one story of brick
Size horizontal leg of all shelf angles and lintels to provide
a minimum bearing of
2
/3 the thickness of the brick wythe
Sheathing:
Exterior grade gypsum sheathing or OSB or glass fiber
mat-faced sheathing or cement board; minimum
1
/2 in.
(12.7 mm) thick
Exterior grade plywood; minimum
3
/8 in. (9.5 mm) thick
Closed-cell rigid insulation meeting ASTM C 578 or C
1289; minimum
1
/2 in. (12.7 mm) thick
Water-Resistant Barrier:
Water-resistant barriers include No. 15 asphalt felt, build-
ing paper, qualifying high-density polyethylene or polypro-
pylene plastics (housewraps)
Install water-resistant barrier over sheathing
Seal water-resistant sheathing per manufacturer to per-
form as water-resistant barrier
Ship lap water-resistant barrier pieces minimum 6 in. (152
mm)
Steel Studs:
Galvanized steel studs with minimum G90 coating
Restrict allowable out-of-plane deflection of steel studs to
L/600 using service level loads
Minimum 0.043 in. (18 gage; 1.09 mm) studs for exterior
walls
Do not field weld steel studs
Screws:
Minimum No. 10 self-tapping corrosion-resistant screws
with a minimum nominal shank diameter of 0.190 in. (4.8
mm)
Corrosion resistance provided by polymer coating, zinc
plating or stainless steel
Mortar:
Comply with ASTM C 270
Type N recommended; Type S alternate
Expansion Joints:
Provide vertical and horizontal expansion joints through
brick veneer
Design and construct expansion joints complying with rec-
ommendations of Technical Notes 18 and 18A
Condensation Analysis:
Determine if potential for condensation exists in the wall
Make necessary changes to the wall design
Page 1 of 15
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 2 of 15
Moisture Resistance
Brick veneer construction incorporates a drainage
cavity to deter water penetration into the building. This
continuous, clean air space creates a physical separa-
tion between the brick wythe and the inner steel stud
wall. When wind-driven rain penetrates the veneer
wythe, the air space allows the water to drain down
the back face of the brickwork. This water is then col-
lected by flashing and channeled out of the veneer
wythe through weeps. When properly designed and
constructed, a brick veneer/steel stud system is a
water penetration resistant wall assembly. For addi-
tional information regarding water penetration resis-
tance, see the Technical Notes 7 Series.
Thermal Performance
Brick veneer systems incorporating an air space
can greatly reduce the amount of heat transmission
through the system. This air space provides a thermal
separation between the brick wythe and other system
components, increasing the resistance of the entire
wall system to heat loss or gain. Further, brickwork
has a high thermal mass giving it the ability to store
and slowly release heat over time. This is taken into
account in current energy codes by allowing a lower
R-value for walls with masonry. Batt insulation is typi-
cally placed between studs to increase the thermal
resistance of the wall. In addition, closed-cell rigid
board insulation can be placed inside an enlarged
air space for additional thermal resistance. With the
board insulation located outside of the steel stud wall,
there is increased resistance to heat transmission
and reduced thermal bridging. For further information
regarding the thermal resistance of brick assemblies,
refer to the Technical Notes 4 Series.
Fire Resistance
Brick masonry has superior fire resistance. Building
codes may require that exterior walls have a fire
resistance rating based on fire separation distance,
size of building and occupancy classification. Exterior
walls may require protection from one or both sides,
depending on whether the fire separation distance is
more or less than 5 ft (1.52 m), respectively. A nominal
4 in. (102 mm) brick wythe has a 1 hour fire resistance
rating and can provide this protection for the exterior
surface of the wall. For fire resistance from inside
the building, the steel stud must be protected on the
interior side. Fire-rated gypsum board is typically used
for this purpose and can be layered to provide the
required rating. For additional information, see the
Technical Notes 16 Series.
Acoustics
Brick veneer walls with cavities are well suited as
sound insulators. Three mechanisms reduce the
sound transmitted through the wall. The hard surface
of the brickwork reflects a large portion of sound
waves. The mass of the brickwork absorbs another
portion of sound energy. The remaining sound energy
which makes its way through the brick wythe must
continue through the air space and the sheathed
studs. This air space separates the brick from the
steel studs causing a dampening effect. With only
anchors bridging the air space, a further reduction in
sound wave propagation is realized due to discontinu-
ous construction. Finally, the energy must vibrate the
sheathing and stud to reach the inside of the building.
Additional information on sound transmission can be
found in Technical Note 5A.
Aesthetics
Brick is available in a large variety of colors, textures,
glazes and coatings. In addition, many sizes are
manufactured and special shapes can be created to
achieve a broad range of units. Add to this the ability
to achieve multiple bond patterns, the use of colored
mortars and interesting masonry detailing, and the
creative possibilities are nearly endless. For further
information on sizes and patterns, refer to Technical
Notes 10B and 30.
Ease of Construction
The steel studs and exterior sheathing of a brick
veneer/steel stud wall can be constructed prior to lay-
ing the brick veneer wythe. This allows the building to
be closed-in and placed under-roof quickly. Thus, inte-
rior work can begin with brick masonry construction
following at a convenient time. Further, other trades
can be scheduled to work and not interfere with the
mason. Care should be taken to ensure that the stud
system is not compromised prior to installation of the
brickwork.
Design Weight
The weight of a brick veneer/steel stud wall is less
than a wall constructed of brick and concrete masonry
units. Thus perimeter framing member sizes and seis-
mic forces used in the design may be reduced.
PROPERTIES OF BRICK VENEER/STEEL STUD WALLS
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 3 of 15
INTRODUCTION
The brick veneer/steel stud wall system offers several
advantages over other claddings. The system demon-
strates superior performance in many of the specific
areas of concern for designers, contractors and property
owners such as attractive appearance, high resistance
to water penetration, low thermal transmission rate,
ease of construction and low maintenance. Introduced
in the 1960's, the brick veneer/steel stud wall system
has evolved into a successful construction method used
in a wide variety of commercial, industrial and institu-
tional structures which include such building types as
churches, hospitals and office buildings. These build-
ings usually have structural frames of steel or reinforced
concrete. Unlike residential construction, they generally
are not designed with overhangs, eaves or gutters to
protect the veneer and frequently incorporate parapets.
They also are usually taller than residential structures.
Consequently, many commercial brick veneer/steel stud
wall systems have greater exposure to their environment than their residential counterparts. For this reason, it
is important to closely observe proper design, detailing and construction practices to ensure that expected and
required levels of performance are met.
The brick veneer/steel stud wall system is considered an anchored veneer wall. An anchored veneer is a brick
wythe secured to and supported laterally by the backing through anchors and supported vertically by the founda-
tion or other structural elements. The veneer transfers out-of-plane load directly to the backing and is not consid-
ered to add load-resisting capacity to the wall system. Anchored brick veneer with steel stud backing consists of a
nominal 3 or 4 in. (76 to 102 mm) thick exterior brick wythe mechanically attached to a steel stud backing system
with corrosion-resistant metal anchors so as to create a prescribed air space between the veneer and the backing
system as shown in Figure 1.
This Technical Note is one in a series dealing with brick veneer. This Technical Note addresses brick veneer with
steel stud backing in commercial construction. Others in the series discuss other types of brick veneer wall sys-
tems.
STRUCTURAL DESIGN CONSIDERATIONS
Brick veneer/steel stud walls must resist loads as prescribed by the governing building code(s). For exterior,
nonbearing walls, these loads are typically due to wind and seismic events. Although a veneer is defined as a
nonstructural facing, brick veneer does resist loads. Certainly the weight of the brick is supported by the brick-
work itself. But brickwork also contributes to the resistance of out-of-plane loads generated by wind and seismic
events. In addition, in-plane forces caused by the weight of the brickwork are also resisted internally, including in-
plane loads generated by seismic events. Returns and offsets in the veneer wythe can also act as flanges (in the
absence of expansion joints at these locations) and cause in-plane loads on the wall. Steel studs can be designed
to be nonloadbearing or loadbearing. Both nonloadbearing and loadbearing studs provide backing for the brick
wythe by resisting any out-of-plane loads such as those from wind or seismic events. Loadbearing studs also
serve as part of the structural system of a building by supporting a portion of the gravity load or acting as shear
walls while nonloadbearing studs only support their own weight.
Minimum standards for brick veneer/steel stud walls are established in the model building codes adopted by most
local jurisdictions. Some of these codes reference the ACI 530/ASCE 5/TMS 402 Building Code Requirements
for Masonry Structures, also known as the Masonry Standards J oint Committee (MSJ C) Code. [Ref. 2] Within
this Code, there is an entire chapter devoted to masonry veneers which outlines prescriptive, as well as alternate
design requirements, for anchored masonry veneers. To determine code provisions for a building in a specific
area, the local building code jurisdiction should be consulted.
Figure 1
Brick Veneer/Steel Stud Wall
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 4 of 15
System Behavior
Together, brick veneer and steel studs resist out-of-plane
loads by each taking a portion of the load relative to
its flexural stiffness, span length, and the ability of the
anchors to transfer the load. The flexural stiffness of the
brick is substantially greater than that of the steel stud
backing. In addition, the brick veneer typically spans a
greater distance than the steel stud system as shown in
Figure 2. Consequentially, the brick initially carries most of
the load, acting similar to a one-way beam. Deflections in
the brick wythe are transmitted to the steel studs through
the anchors. Anchors nearest the top and bottom of the
steel stud transfer more load than those located near the
center of the span. Frictional forces at the support for the
veneer resist a portion of the load.
As the masonry continues to deflect, flexural tensile
stresses developed in the veneer may cause a break in
bond near the point of highest moment, typically near the
center of the brickwork. This occurs when the modulus of
rupture of the brickwork has been exceeded. The veneer
will subsequently act as two separate segments with no
stress transfer across the crack. Thus, each segment will
act as a one-way beam spanning between the anchor
nearest the crack and the respective anchor at the top or
bottom of the veneer.
Steel Studs Deflection Criteria
Steel studs must be designed to provide adequate out-of-
plane support for all loads imposed on the wall system.
This is done by establishing a maximum deflection limit on
the stud while maintaining steel stress values in the stud
within permissible limits. This deflection is calculated assuming the entire out-of-plane load is resisted by the studs
alone, neglecting contribution of the brick veneer. While a number of design tables are based on a stud deflection
of stud span length divided by 360 (L/360), using this criterion may permit more deflection than the veneer is able
to tolerate. Therefore, to obtain sufficient backing stiffness, the allowable out-of-plane deflection of the studs due
to service level loads should be restricted to L/600. Such deflection criterion will allow a maximum crack width of
about 0.015 inches (0.38 mm) in the brick veneer wythe for typical floor-to-floor dimensions.
Steel Stud Design Recommendations
Studs surrounding all openings in the veneer should be designed with loads based on the tributary area of the
opening (windows and framing must be tied only to the metal studs). Further criteria for loadbearing studs include
providing adequate bearing capacity for the gravity loads. The flanges of the steel studs must be laterally braced
to resist compression in bending. This can be accomplished by fastening sheathing or board materials, such as
water-resistant gypsum sheathing, plywood or cement board, to each side of the stud. In general, rigid board
insulation does not contribute to adequate bracing. Alternatively, if sheathing or board materials are only installed
on the exterior side of the stud, bracing can be provided by attaching engineered steel straps or channels affixed
either horizontally or diagonally to the studs. However, it is suggested to provide sheathing on both sides, in addi-
tion to any engineered bracing, to help support the water-resistant barrier and any interior finishes. The design of
the bracing should follow appropriate codes and technical literature. [Ref. 8]
Seismic and Wind Requirements
The following requirements apply to all anchored veneer, not exclusively those with a backing of steel studs. For
further information, refer to Technical Note 3.
Seismic Requirements. As with all building materials, when the possibility and potential intensity of seismic activ-
ity increases, certain brick veneer seismic provisions are invoked. These requirements are specified in the model
Figure 2
Initial Moment and Anchor Load Distribution
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 5 of 15
building codes. Some of these codes reference the MSJ C Code. Refer to the appropriate building code to deter-
mine specific seismic provisions.
The veneer chapter of the MSJ C Code addresses seismic provisions for anchored masonry veneers. Beginning in
Seismic Design Category (SDC) C, these requirements become increasingly more stringent as the SDC increas-
es. The veneer is first required to be isolated from the structure. It is then required to have a reduced spacing of
anchors. Finally, joint reinforcement that is mechanically attached to the anchors is required. [Ref. 2]
Wind Requirements. In locations where the basic wind speed exceeds 110 mph (177 km/hr) but does not exceed
130 mph (209 km/hr) and the building's mean roof height is not greater than 60 ft (18.3 m), the permissible wall
area per anchor is required to be reduced and the spacing between anchors at perimeter openings must be
reduced. In locations where the basic wind speed exceeds 130 mph (209 km/hr), the veneer is permitted to be
rationally designed. [Ref. 2]
DETAILING
Foundations
Although some building codes permit the support of brick
veneer on wood foundations, it is recommended that the weight
(gravity load) of the veneer be supported on concrete or mason-
ry foundations or other noncombustible structural supports, such
as attached steel angles. The brick wythe may extend below
grade if it is properly detailed and constructed to minimize water
penetration. A typical foundation detail is shown in Figure 3.
Locating base flashing and weeps a minimum of 6 in. (152 mm)
above grade will allow the drainage system to function properly.
Base flashing should extend through the full wythe of the veneer
to the exterior to preclude any moisture from migrating by capil-
lary action up through the brickwork.
Brickwork below the base flashing should be detailed as a bar-
rier wall system by completely filling the cavity or air space with
grout or mortar to minimize water penetration. Anchors should
be located within the grout-filled cavity according to the same
spacing as in the brick veneer above grade. Steel studs should
be located a minimum of 6 in. (152 mm) above grade and
should not be used below grade on exterior walls under any
circumstances.
If soil immediately adjacent to the brickwork is not free-drain-
ing the brick wythe exterior should be waterproofed below
grade. Self-adhesive waterproofing membranes with protection
board to prevent damage during backfill operations can prevent
water from penetrating the brick. Drainboards with integral filter
fabric and waterproofing membrane can also be installed to
drain water to the foundation drain tile system. A French drain
between the soil and the wall, consisting of a gravel fill with a
fabric filter surround and drain tile below, sloped a minimum of
1
/8 in./ft (10 mm/m) can also provide some drainage. Finished
grade should provide positive drainage by sloping away from the
wall.
Water-Resistant Barrier
Water-resistant barriers are membranes which prevent liquid water from passing through them. These are dif-
ferent from vapor retarders, intended to prevent water vapor diffusion, and air barriers, intended to prevent air
flow through the wall system. Such a membrane should be located between the air space and the sheathing or
between the rigid insulation and the sheathing. A water-resistant barrier should keep out any water which finds
its way across the air space via anchors, mortar bridging or splashing. Individual pieces of water-resistant barrier
should be installed with their edges and ends lapped at least 6 in. (152 mm). While a separate membrane is pre-
Figure 3
Wall Section at Foundation
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 6 of 15
ferred, sheathing or rigid insulation with an inherent resistance to moisture penetration may also serve as a water-
resistant barrier when all edges and joints are completely taped or sealed.
A water-resistant barrier is required and can be provided by No. 15 asphalt felt, building paper, qualifying high-
density polyethylene or polypropylene plastics (housewraps) and qualifying water-resistant sheathings when
properly sealed. No. 15 asphalt felt should comply with Type I of ASTM D 226, Specification for Asphalt-Saturated
Organic Felt Used in Roofing and Waterproofing. Asphalt felt should not be left exposed to ultraviolet (UV) light for
an extended period of time, otherwise, it will lose the asphalt saturation and water resisting characteristics.
Some plastic membranes (housewraps) may have qualities similar to those of a water-resistant barrier, but ascer-
taining the effectiveness of a particular plastic as a water-resistant barrier can be difficult. While felts tend to seal
themselves when penetrated by fasteners, plastics may not. In addition, some plastic membranes also act as
vapor retarders and hence can potentially trap water vapor inside the stud wall where it can condense if the tem-
perature gradient in the wall drops below the dew point. The length of time a plastic membrane will be exposed
to sunlight should also be considered. Most show serious degradation with 3 to 12 months exposure to UV rays.
Thus, all plastic membranes should not be regarded as equivalent and caution should be used when using them
as a water-resistant barrier.
Care should be taken to reduce the likelihood of tearing the membrane or breaking the barrier. Such tears or
breaks must be corrected prior to installation of brickwork. Water-resistant sheathings with integral membranes
must be completely sealed with compatible tape or sealant to perform as water-resistant barriers. This means
components providing this seal must maintain their integrity and performance when subjected to moisture and
other environmental conditions over the life of the wall. These sheathing systems should also allow for the trans-
mission of vapor unless the effect of a vapor retarder is considered at this location in the wall.
Sheathing
An exterior grade sheathing or insulation material should be installed on the exterior side of the stud. Edges and
joints of sheathing or insulation board that also serve as the water-resistant barrier should be thoroughly sealed
with compatible tape or sealant to ensure against moisture intrusion over the life of the wall. Such joint treatment
will also reduce air infiltration. Careful detailing at the top of walls, at transition to other opaque materials and at
window openings should provide a watertight condition. If sheathing is used to laterally brace the studs, it should
be rigid enough to provide the required stiffness.
Exterior sheathing on steel studs should be suitably fastened with corrosion-resistant screws. The sheathing
should be one of the following: exterior grade gypsum sheathing or glass fiber mat-faced sheathing or cement
board, not less than
1
/2 in. (12.7 mm) in thickness; closed-cell insulating rigid foam not less than
1
/2 in. (12.7 mm)
thick conforming to ASTM C 578 or ASTM C 1289 or oriented strand board (OSB) not less than
1
/2 in. (12.7 mm) in
thickness; or exterior grade plywood not less than
3
/8 in. (9.5 mm) in thickness.
Screws
Corrosion-resistant screws with a minimum nominal shank diameter of 0.190 in. (4.8 mm) are required to attach
anchors to steel studs. A minimum #10 self-tapping screw is recommended. Screws used to attach exterior
sheathing and anchors can be either carbon steel or stainless steel. Carbon steel screws should have a non-cor-
rosive coating of zinc, polymer or composite zinc-polymer. Zinc-plated screws should be either mechanical-zinc
plated according to either 1) ASTM B 695, Specification for Coatings of Zinc Mechanically Deposited on Iron and
Steel or 2) electro-zinc plated in accordance with ASTM B 633, Specification for Electrodeposited Coatings of Zinc
on Iron and Steel. Polymer-coated screws do not have the self-healing properties of zinc, however they can offer
acceptable, long-term protection. A composite zinc-polymer coating offers superior protection to either coating
alone. Stainless steel screws may be acceptable even though a galvanic potential exists between stainless steel
and carbon steel. This is possible because of an area-relationship principle where the surface area of the steel
stud is much larger than that of the screw which results in a decreased corrosion potential. Copper-coated screws
are not recommended since they can react galvanically with steel studs having zinc coatings.
Screws incorporating an integral EPDM or neoprene sealing washer under the screw head may also assist in
water penetration resistance. Due to the area-relationship principle mentioned above, when stainless steel screws
are used with carbon steel anchors, sealing washers are highly recommended.
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 7 of 15
Steel Studs
The top connection of nonbearing studs must be detailed to prevent inadvertent vertical load transfer to nonload-
bearing studs. No rigid connection should be allowed between the top track and the studs. This allows for the
structural member above the track to deflect without transferring loads to the studs. Field welding of studs should
not be permitted. Shop welding may be permitted on steel studs with a minimum nominal thickness of 0.068 in.
(1.7 mm) (14 gage) studs. To increase quality assurance, welders and welding procedures should be qualified
as specified in AWS D1.3 by the American Welding Society. A corrosion inhibiting coating should be applied to all
welded areas after welding is completed.
Steel studs should have a minimum nominal thickness of 0.043 in. (1.1 mm) (18 gage) to provide sufficient thick-
ness to engage the threads of the screw. Studs should have a protective coating conforming to one of the follow-
ing ASTM standards: 1) ASTM A 653/653M, Specification for Steel Sheet, Zinc-Coated (Galvanized) or Zinc-Iron
Alloy-Coated (Galvannealed) by the Hot-Dip Process with a minimum G90/Z275 coating designation or 2) ASTM A
875/875M, Specification for Steel Sheet, Zinc-5% Aluminum Alloy-Coated by the Hot-Dip Process with a minimum
GF90/ZGF275 coating designation. For further information on selecting corrosion inhibiting coating weight, refer to
GalvInfoNote #19 [Ref. 4]
Air Space
The air space or drainage cavity provides a means to drain water which penetrates the brick veneer. The air
space between the back of the brickwork and the sheathing or rigid board insulation is recommended to be a mini-
mum of 2 in. (51 mm) and required to be a minimum of 1 in. (25.4 mm) in order to minimize the possibility of mor-
tar bridging the air space. A 4
1
/2 in. (114 mm) maximum distance is required between the back of the brick wythe
and the steel framing unless the anchors are rationally designed. If this distance is exceeded, additional or stron-
ger anchors may be required. Insulation must be attached to the backing by mechanical or adhesive means to
keep it from blocking the air space. When a high probability of mortar falling into the air space exists, such as for
tall brick veneer without shelf angles, drainage materials that catch mortar droppings may be specified at the base
to prevent mortar blocking the weeps. However, the use of drainage materials should not preclude good workman-
ship and an effort to keep the air space clean of excess mortar droppings.
Flashing
Flashing collects water at the bottom of the air space and directs it toward weeps which channel it to the exterior
face of the wall. Flashing must be placed at all locations where the air space is interrupted. These include above
and below all window and wall openings, above all shelf angles, at the base of the wall and under the coping at
parapets. Flashing should extend vertically up the backing a minimum of 8 in. (203 mm). If drainage materials that
catch mortar are placed at the bottom of the air space, flashing at the base of the wall may need to extend further
up the backing. This ensures that the flashing extends above the height of the drainage material and helps deter
water that migrates across mortar on the drainage material from entering the backing. The water-resistant barrier
on the backing should lap the top of the flashing a minimum of 4 in. (102 mm). Individual flashing pieces should
be lapped at least 6 in. (152 mm) and sealed to avoid water running under adjacent flashing pieces. Where flash-
ing is discontinuous, such as over and under openings in the wall, the ends should be turned up at least 1 in.
(25.4 mm) into the next head joint to form an end dam to channel water out of the wall. When possible, flashing
should extend beyond the face of the brickwork to form a drip. When using a flashing that deteriorates with UV
exposure, a metal or stainless steel drip edge can accomplish this. It is imperative that flashing be extended at
least to the face of the brickwork.
Flashing material should be waterproof and durable. It should be sufficiently tough and flexible so as to resist
puncture and cracking. In addition, flashings subject to deterioration from UV light should not be overly exposed
to sunlight. Flashing should not deteriorate when in contact with metal parts, mortar, sealants or water. Flashing
should also be compatible with adjacent adhesives and sealants. Flashings used in a wall system with water-
resistant sheathing acting as the water-resistant barrier should be self-adhesive or be mechanically attached with
a pressure bar and sealant. It is suggested that only superior flashing materials be selected, since replacement in
the event of failure is extremely expensive.
Weeps and Vents
Weeps should be placed immediately above the wall flashing to permit water to exit the wall. Open head joint
weeps are recommended with a spacing of no more than 24 in. (610 mm) on center. Wick and tube weeps are
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 8 of 15
recommended to be spaced no more than 16 in. (406 mm) on center. Most building codes require weep openings
to have a minimum diameter of
3
/16 in. (4.8 mm) and allow weeps to be spaced up to 33 in. (838 mm) on center.
Wicks should be at least 16 in. (406 mm) long and extend through the brick into the air space and along the back
of the brick. Non-corrosive metal, mesh or plastic screens can be installed in open head joint weeps if desired.
Vents (open head joints) may be placed at the top of the drainage air space to help reduce moisture buildup in the
air space by promoting ventilation. Vents should be spaced at the same horizontal spacing as weeps and should
be centered between weeps. Insect access to vents may be controlled through the use of covers or screens.
Anchors
Care must be taken to anchor the masonry veneer to the backing in a manner that will permit each to move freely,
in-plane, relative to the other. Anchors that connect the veneer to the backing must provide out-of-plane sup-
port, resisting tension and compression, but allowing shear. This permits in-plane differential movement between
the frame and the veneer without causing cracking or distress. Such anchors are shown in Figure 4. Corrugated
anchors are not permitted when brick veneer is anchored to steel stud backing.
Anchors should provide the capacity to transfer loads applied to a maximum of 2
2
/3 ft
2
(0.25 m
2
) of wall area. Each
anchor should be spaced a maximum of 18 in. (457 mm) on center vertically and a maximum of 32 in. (813 mm)
on center horizontally. They must be securely attached through the sheathing to the steel studs, not to the sheath-
ing alone. Around the perimeter of openings, additional anchors should be installed at a maximum of 3 ft (914
mm) on center within 12 in. (305 mm) of the opening.
All anchors must be embedded at least 1
1
/2 in. (38 mm) into the brick veneer with a minimum mortar cover of
5
/8
in. (15.9 mm) to the outside face of the wall. Anchors in Seismic Design Categories E and F must be mechanically
fastened to horizontal reinforcement in the brick veneer as depicted in Figure 5.
Anchors transfer load between the brick veneer and either the studs or the structural frame of the building. The
load that is transferred through a particular veneer anchor depends on many factors. Such factors include: anchor
stiffness; air space dimensions; the backing element the anchor is fastened to (the building frame or the steel
stud); where the anchor is fastened relative to the backing element's span; where the anchor is located relative to
the brick veneer's span; whether any cracks have occurred in the veneer; stud stiffness; and embedment.
For walls having shelf angles at each floor level, with either no windows or punched window openings in which
brick veneer supports the lintel, anchors carrying the highest load will be located near the bottom and top of the
floor span that are attached directly to the building frame. In one test performed for this configuration, the anchor
connected closest to the shelf angle supporting the veneer carried just over 30% of the total out-of-plane load of
the vertical strip on the story it served. [Ref. 1]
With brick veneer supported on a shelf angle above a window band, the anchors at the floor level will typically
carry the highest load. Again, anchors which are fastened directly to the building frame will carry more load than
Figure 5
Seismic Anchor Assemblies
Figure 4
Adjustable Anchor Assemblies
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 9 of 15
those attached to the studs.
Anchors are required to be made of carbon steel or stainless steel. Carbon steel anchors are required to conform
to ASTM A 82. Stainless steel anchors are required to conform to ASTM A 580. Anchors made of carbon steel are
required to be hot-dipped galvanized in accordance with Class B-2 of ASTM A 153/153M, Specification for Zinc
Coating (Hot-Dip) on Iron and Steel Hardware.
Two-piece adjustable anchors with a minimum wire size of W1.7 (MW11) are required. Eye and pintle adjustable
anchors are required to have a minimum wire size of W2.8 (MW 18) with a diameter of
3
/16 in. (4.8 mm). Wire
anchors are available in a variety of standard lengths from 3 to 5 in. (76 to 127 mm) and diameters from 0.15 to
0.25 in. (3.7 to 6.4 mm). In addition, anchors should have a maximum horizontal out-of-plane mechanical play of
1
/16 in. (1.6 mm) and should be detailed to prevent disengagement.
Anchors with formed drips in the wire should not be used since they have reduced load capacity. Corrugated
anchors are not permitted with steel stud backing. They may not fully engage the stud upon initial loading and do
not have sufficient compressive capacity for the given air space.
Anchors incorporating an EPDM sealing membrane between the sheathing or insulation and the wall base of the
anchor should be considered for superior water resistance. Prongs at each end of an adjustable anchor base, as
shown in Figure 4, may also be considered with non-rigid sheathing to provide a mechanical connection between
the anchor and the stud. These prongs provide positive, independent anchorage in the event of long-term deterio-
ration of sheathing or insulation and prevent compression of the insulation or sheathing. When using a prong-leg
base, a modified asphalt pad with self-adhesive is recommended. This pad is installed under the anchor base and
will seal openings created by the prongs and screws in the sheathing or insulation.
Lintels and Shelf Angles
Lintels provide support of brickwork over masonry openings by bearing on the brickwork on each side of the
opening. They are not attached to the building structure. Shelf angles provide support for the brickwork above by
attaching to the building structure. Shelf angles are at times referred to as relieving angles.
Steel for lintels and shelf angles should conform to ASTM A 36/A 36M, Specification for Carbon Structural Steel.
Steel angles should be a minimum of
1
/4 in. (6.4 mm) thick. All angles should be primed and painted as a minimum
to inhibit corrosion. Galvanized and stainless steel angles should be considered in harsh environments such as
coastal areas.
Lintel and shelf angle deflection between support points should not exceed the lesser of L/600 or 0.3 in. (7.6 mm)
and the total rotation of the toe of the angle should be less than
1
/16 in. (1.6 mm). The horizontal leg of all angles
should be sized to support a minimum of
2
/3 the thickness of the brick wythe.
Lintels should be installed over all masonry openings unless the brick is self-supporting. Lintels can be loose steel
angles, stone, precast concrete or reinforced masonry. They should bear a minimum of 4 in. (102 mm) on brick
on each side of the opening and should be sized to carry the brick veneer above them. For further information on
lintels, refer to Technical Note 31B.
Vertical expansion joints should not cross a lintel with-
out making provisions for potential movement. When an
expansion joint crosses a lintel, the full weight of the brick-
work above the lintel must be carried by the lintel.
Shelf angles should consist of steel angles sized and
installed to carry the brickwork above. Structures with a
maximum veneer height of 30 ft (9.14 m) from founda-
tion to top of wall and 38 ft (11.58 m) from foundation to
top of gable can have their entire brick veneer supported
directly on a foundation wall, footing or noncombustible
support without shelf angles. Unless rationally designed,
brick veneer above this height is required to be supported
by shelf angles for each story. Shelf angles are typically
located near the floor line or at the window head. Shelf
angles attached to rigid concrete or steel elements should
Figure 6
Shelf Angle with Concrete Frame
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 10 of 15
have full height shims to reduce rotation
as shown in Figure 6. Any shelf angle
attached to miscellaneous steel elements
must have bracing to prevent out-of-plane
movement of the wall as depicted in Figure
7.
Shelf angles should not be installed as
one continuous member. Space should
be provided at intervals to permit thermal
expansion and contraction of the steel
angle to occur without causing distress to
the masonry. Lipped brick may be used
above or below a shelf angle to maintain
the same joint width at the angle as other
joints in the brickwork.
Shelf angles should be supported by mis-
cellaneous structural steel elements and
not by steel studs. Field welding of shelf
angles to studs should never be permitted
since the thin wall of the steel stud increas-
es the potential for burn-through. Further,
a stud which supports a shelf angle may
require additional reinforcing and may be
more prone to corrosive action from expo-
sure to the moist air space.
Head, Jamb and Sill
Details
Openings in brick veneer walls should be
carefully detailed to prevent water from
entering the brick veneer/steel stud wall
system. Provision should be made for
movement between the brick veneer and
the frame or backing. Window frames, door
frames and opening sleeves must be attached to the backing, not the brick veneer. Window and door flashing
must be integrated with the water-resistive barrier to provide a continuous barrier to moisture intrusion as shown in
Figures 7 and 8. Sills should be sloped to the outside for drainage. Refer to Technical Note 36 for further informa-
tion.
Sealant Joints
Sealant joints prevent water penetration at expansion joints and perimeters of openings. These joints are typically
a compressible, foam backer rod recessed and covered by a sealant. Sealant joints should be free of mortar for
the entire thickness of the brick veneer and closed with the backer rod and sealant. If desired, a compressible
material may be included behind the backer rod.
The perimeter of all exterior window frames, door frames and sleeves should be closed with a sealant joint as
shown in Figure 8. This joint should be between
1
/4 and
1
/2 in. (6.4 and 12.7 mm) wide and
1
/4 in. (6.4 mm) deep.
Fillet joints are not recommended, but if used, should be at least
1
/2 in. (12.7 mm) across the diagonal.
Sealants should be selected for their durability, extensibility, compressibility and their compatibility with other
materials. A sealant should be able to maintain these qualities under the temperature extremes of the climate in
which the building is located. Sealant materials should be selected to comply with ASTM C 920, Specification for
Elastomeric J oint Sealants. Specific sealants recommended for brick include polysulfide, solvent release acrylic,
silicone and urethane sealants. A sealant primer may be required before applying some sealants on certain brick
to preclude staining. Since acetoxic silicone will attack cement in mortar, it should not be applied to masonry.
Oil-based caulks should not be used since they may stain the adjacent brickwork. Refer to Figures 6 and 11. For
Figure 7
Brick Veneer/Steel Stud Bracing System
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 11 of 15
further information on sealants, refer to ASTM C 1193,
Guide for Use of J oint Sealants.
Backer rods should be placed behind all sealant joints.
They should be of closed-cell plastic foam or sponge
rubber. Backer rods should be capable of resisting per-
manent deformation before and during sealant applica-
tion, non-absorbent to liquid water and gas, and should
not emit gas which may cause bubbling of the sealant. A
bond breaking tape may be required with some types of
backer rods.
Parapet Walls
Parapets are exposed on three sides and consequently
are potentially more vulnerable to water penetration and
condensation. Parapet walls should be avoided unless
required. If a parapet is required, it should be properly
designed, detailed and constructed. Steel studs are not
recommended as backing for parapet walls because of
potential moisture and movement issues. Reinforced
masonry provides the best means of constructing para-
pets above brick veneer/steel stud walls as shown in
Figure 9. A gravel stop detail as shown in Figure 10 can
be used instead of a parapet wall.
OTHER CONSIDERATIONS
Condensation
Experience has shown that most water or moisture
found between steel studs in a brick veneer/steel stud
wall can be attributed to condensation. Condensation
occurs at the point in the wall where the temperature
gradient exceeds the dew point. If this point is within
the air space, then the condensation will find its way out
of the wall via the drainage system. However, if it is on
the inside of the steel stud wall with batt insulation, then
it may dampen or eventually saturate the surrounding
materials and may lead to mold and/or corrosion prob-
lems.
Consequently, it is recommended that a condensation
analysis be conducted to determine if the potential for
condensation exists in a wall. If results indicate that
it may occur within the sheathing or steel stud wall,
then the wall design should be changed. Rigid board
insulation may be placed on the outside of the exterior
sheathing to increase the thermal resistance of the wall
or an air barrier or vapor retarder may be installed to
decrease air and vapor movement through the wall.
Installing some or all of the required insulation in the air
space also helps reduce or eliminate thermal bridging
through the steel studs. See Technical Notes 7C and 7D
for further information.
Air Barriers and Vapor Retarders
Where analysis indicates a probability of condensation,
an air barrier or vapor retarder should be provided. Air
a) Window Head Detail
b) Window J amb Detail
Figure 8
Window Details
c) Window Sill Detail
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 12 of 15
barriers are membranes made of polyethylene, polypropylene or polyolefin. They are intended to prevent air
leakage through the building envelope, hence reducing the associated energy losses and moisture movement.
Most allow the transmission of vapor, while some also act as vapor retarders. Manufacturers provide data based
on different standards, including 1) ASTM D 726, Test Methods for Resistance of Nonporous Paper to Passage of
Air and 2) ASTM E 283, Test Method for Determining the Rate of Air Leakage Through Exterior Windows, Curtain
Walls, and Doors Under Specified Pressure. For this reason, caution should be exercised when evaluating and
specifying air barriers.
Vapor retarders minimize moisture movement due to water vapor diffusion and are made of materials similar to
air barriers. While some air barriers will also inhibit vapor transfer, all vapor retarders can be air barriers if they are
installed and thoroughly sealed with no tears or holes. Some manufacturers cite test data based on ASTM E 96,
Test Methods for Water Vapor Transmission of Materials. However, this test does not account for fastener penetra-
tions, electrical outlets, or joints in the retarder. Materials which qualify as vapor retarders should have a perm rat-
ing of 1 or less.
Mortar Type
Mortar plays an important role in the flexural strength of a brick veneer wythe. Out-of-plane strength tests of full-
scale walls indicate that the bond between mortar and brick units is the most important single factor affecting wall
strength when resisting horizontal joint cracking. Mortar should conform to ASTM C 270, Specification for Mortar
for Unit Masonry. A designer should select the mortar with the lowest compressive strength that is compatible
with the project requirements. The compatibility between a particular brick and mortar should be examined when
determining mortar type. Flexural bond strength of a particular brick/mortar combination can be determined using
ASTM C 1357, Test Methods for Evaluating Masonry Bond Strength. Type N mortar is suitable for most veneer
brickwork, except in areas below grade, where Type S mortar should be used. Type S mortar is recommended
Figure 9
Masonry Parapet Wall
Figure 10
Gravel Stop
Counter Flashing
Dovetail Anchor
Metal Coping
Air Space, Min.
2 in. (51 mm)
Recommended
Sealant
Steel Reinforcement
Through Wall Flashing
J oint Reinforcement
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 13 of 15
where a higher degree of flexural resistance is required. Admixtures and additives for workability are not recom-
mended since they can potentially weaken the mortar. Admixtures containing chlorides should never be used
since they could greatly increase the probability of efflorescence and corrosion. For more information, refer to the
Technical Notes 8 Series.
Movement Provisions
Brickwork will expand and contract as will all building components. Brick is subject to permanent expansion as a
result of freezing and moisture absorption. Mortar will shrink as it cures. Changes in temperature will cause brick
to expand and contract. Moisture expansion can continue for years while thermal movement and mortar contrac-
tion will occur periodically, contingent upon temperature and moisture content. As a result, brickwork will continu-
ally change in size during its life.
To accommodate this movement, brick veneer should be designed in discrete sections which are allowed to move
independently of each other. This is accomplished through the use of expansion joints and bond breaks detailed
into the veneer. An expansion joint consists of a vertical or horizontal opening through the brick wythe that is
closed with a sealant joint and elastic materials. These joints separate each section of brickwork and isolate it from
other sections. Expansion joints must be designed and constructed to permit anticipated movements. Further,
expansion joints must be located and constructed so as not to impair the integrity of the wall.
The spacing and placement of vertical and horizontal expansion joints must be done on a case-by-case basis.
Each wall must be examined to determine its potential for movement based on its length, openings, offsets, corner
conditions, wall intersections, means of support, changes in wall heights and parapets. These features influence
how the brickwork reacts to movement in a wall. Any portion of wall not able to resist induced stress should be
isolated by an expansion joint. For more information, refer to the Technical Notes 18 Series.
Vertical Expansion Joints. A vertical expansion joint con-
sists of an opening through the brick wythe closed with a
backer rod and sealant. A compressible pad may be used
in the joint to ensure no mortar is placed in the expansion
joint. Such pads can be made of premolded foam or neo-
prene as shown in Figure 11. Vertical expansion joints should
extend from the foundation to the top of the brickwork without
deviating from vertical. When this is not possible, they can
be terminated at horizontal expansion joints. Generally, the
spacing of vertical expansion joints should not exceed 30 ft
(9.14 m) in walls without openings. Vertical expansion joints
are also recommended where site walls adjoin buildings and
at the corners of large openings. Building corners should
have a vertical expansion joint located within 10 ft (3.05 m) of
the corner. When a vertical expansion joint is located within
10 ft (3.05 m) of the corner, the vertical expansion joint on the other wall forming that corner should be placed at
the typical spacing between expansion joints. For example, if the spacing between vertical expansion joints on a
straight wall is 25 ft (7.62 m), then the spacing of expansion joints around a corner could be 10 ft (3.05 m) on one
side of the corner and 15 ft (4.57 m) on the other side. Plan offsets and setbacks of a wall should also include a
vertical expansion joint on inside corners.
Horizontal Expansion Joints. A horizontal expansion joint cannot function unless there is some means of sup-
porting the brickwork above it. Usually this is accomplished by a shelf angle. Shelf angles should have a horizontal
expansion joint below them. These joints are located between the bottom of the shelf angle and the brickwork
below. They consist of a sealant joint and either an opening or compressible pad behind them. Refer to Figure 6.
Bond Breaks. When a different material, such as concrete masonry, cast stone or precast concrete, is incorpo-
rated into a brick wall, differential movement between the two materials is likely to occur. In such cases, a bond
break may be used to separate it from the surrounding brickwork. This break allows for movement between the
two materials and diminishes horizontal or vertical cracking. A bond break is achieved by installing a layer of No.
15 asphalt felt or flashing between the other material and the mortar joint surrounding it. For further information on
bond breaks, see Technical Note 18A.
Figure 11
Vertical Expansion Joints
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 14 of 15
Horizontal Joint Reinforcement
Although not usually required for brick veneer construction, horizontal joint reinforcement can be incorporated into
brick veneer walls to alleviate cracking from high internal stress or to have the brick serve as a reinforced lintel.
Horizontal joint reinforcement is necessary for veneer laid in stack bond, in Seismic Design Categories E and F,
and possibly in joints adjacent to different materials. It may be either single or double wire joint reinforcement and
must have at least
5
/8 in. (15.9 mm) mortar cover. Horizontal joint reinforcement can also be used above and below
the corners of masonry openings for added strength.
Brick
Brick are usually selected on the basis of their appearance which includes color, texture and size. To assure
quality, brick units should conform to one of the following: 1) ASTM C 216, Specification for Facing Brick 2)
ASTM C 652, Specification for Hollow Brick 3) ASTM C 1405, Specification for Glazed Brick or 4) ASTM C 126,
Specification for Ceramic Glazed Structural Clay Facing Tile, Facing Brick and Solid Masonry Units. The use of
salvaged brick is not recommended since such brick may not bond properly with mortar and may be less durable.
For further information on brick specifications and salvaged brick, see the Technical Notes 9 Series and 15,
respectively.
CONSTRUCTION
Construction requirements are found in model building codes and in referenced specifications. Specification for
Masonry Structures, ACI 530.1-05/ASCE 6-05/TMS 602-05 [Ref. 11], is invoked by the MSJ C Code. [Ref. 2]
Project manuals prepared for specific buildings also contain construction requirements.
All materials at the job site should be stored off the ground and under adequate cover to prevent deterioration and
contamination. Cement and lime should be kept dry. Foreign material must be kept out of sand. Brick should not
be placed directly on the ground to preclude any potential staining from the earth.
A box or other measuring tool should be used for measuring sand when mortar is mixed at the job site. Only
full bags of cement and lime should be added to the mixer unless accurate volumetric measurements are used.
Retempering of mortar by adding water is permitted as necessary to maintain consistency. Caution should be
exercised when retempering white or colored mortar to avoid color changes. Water content and stiffness of mortar
during tooling also affect color. All mortar should be used within 2
1
/2 hours of mixing. See Technical Note 8B for fur-
ther information on controls for mixing mortar.
Brick which have an initial rate of absorption (suction) of more than 30 g/min30 in
2
(30 g/min194 cm
2
) should
be wetted and permitted to surface dry prior to laying when using mortar cement or masonry cement. This will
increase the bond between the mortar and the brick by slowing the absorption of water from the mortar. For addi-
tional information, refer to Technical Note 7B.
Hot or cold water protection may be necessary if temperatures are above 90 F (32.2 C) or below 40 F (4.4 C).
When temperatures are above 90 F (32.2 C) and wind exceeds 8 mph (12.9 km/hr), mortar should be used
within 2 hours of mixing and finished brickwork may need to be fog sprayed with water. For construction with tem-
peratures below 40 F (4.4 C), brick units should be at least 20 F (-6.7 C) or above when placed with mortar.
Mortar must not be frozen and should have a temperature between 40 F (4.4 C) and 120 F (48.9 C) when
placed. Additional wind breaks and enclosures may be necessary within certain lower temperature ranges. The
MSJ C Specification [Ref. 11] contains requirements for hot and cold water construction. For further information,
see Technical Note 1 in addition to the Hot and Cold Water Masonry Construction Manual. [Ref. 5]
Care should be taken to completely fill all bed and head joints with mortar. Conversely, any location not intended to
receive mortar, such as air spaces, weeps and expansion joints, should be kept clean and free of mortar and mor-
tar droppings. Mortar joints should be properly tooled to enhance the water resistance of the wall by consolidating
the mortar. J oints should be tooled when thumbprint hard with a jointer tool slightly larger than the joint. Concave,
"V" or grapevine mortar joints are the most water resistant since they do not provide a ledge for water to remain
on the brickwork.
Protection
Protection of unfinished walls is extremely important. The entry of rain or snow into brickwork in progress may
increase the potential for efflorescence and distress in the finished wall. Wind screens and enclosures may also be
necessary in hot or cold water.
www.gobrick.com | Brick Industry Association | TN 28B | Brick Veneer/Steel Stud Walls | Page 15 of 15
MAINTENANCE
Most brickwork is virtually maintenance free. If properly designed, detailed and constructed, brickwork main-
tenance should be minimal. However, brick veneer/steel stud wall systems should be inspected periodically to
ascertain performance and identify any potential problems. Ideally inspections should be performed on a seasonal
basis and on an annual basis as a minimum. Such inspections should address sealant joints, plumbness of the
wall, cracking, etc. In this way, repairs and corrections can be initiated prior to the occurrence of severe problems.
For additional information regarding maintenance, see the Technical Note 46.
SUMMARY
The brick veneer/steel stud wall system is a viable construction option when proper attention is given to design
and detailing, material specification, construction and maintenance procedures.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment and a
basic understanding of the properties of brick masonry. Final decisions on the use of the information
contained in this Technical Note are not within the purview of the Brick Industry Association and
must rest with the project architect, engineer and owner.
REFERENCES
1. Arumala, J .O. and R.H. Brown, Performance Evaluation of Brick Veneer with Steel Stud Backup, College
of Engineering, Clemson University, Clemson, SC, 1982.
2. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
3. Exterior Wall Construction in High-Rise Buildings, Canada Mortgage and Housing Corporation, Ottawa,
ON, 1991.
4. GalvInfoNote #19, Selection Coating Thickness (Weight or Mass) for Galvanized Steel Sheet Products,
Rev 2.2, International Lead Zinc Research Organization, Research Triangle Park, NC, September, 2003.
5. Hot and Cold Water Masonry Construction, Masonry Industry Council, Schaumburg, IL, 1999.
6. KPFF Consulting Engineers and Computech Engineering Services, Report on Behavior and Design of
Anchored Brick Veneer/Metal Stud Systems, Seattle, WA, September, 1989.
7. McGinley, W.M., Warwaruk, J ., Longworth, J . and Hatzinikolas, M., Masonry Veneer Wall Systems,
Structural Engineering Report #156, Department of Civil Engineering, University of Alberta, Edmonton, AB,
J anuary, 1988.
8. North American Specification for the Design of Cold-Formed Steel Structural Members, American Iron and
Steel Institute, Washington, DC, 2001.
9. Rutila, D. A., "Innovations in Brick Veneer", The Construction Specifier, Vol. 51, No. 11, October, 1998.
10. "Self Drilling Fasteners-The Basics", The Construction Specifier, Vol. 42, No. 9, August 1989.
11. Specification for Masonry Structures (ACI 530.1-05/ASCE 6-05/TMS 602-05), The Masonry Society,
Boulder, CO, 2005.
12. Tuluca, A., "Thermal Bridges in Buildings, The Construction Specifier, Vol. 49, No. 11, October, 1996.

Technical Notes 28C - Thin Brick Veneer - Introduction
[Jan. 1986] (Reissued Jan. 2001)
Abstract: This Technical Notes is an introduction to thin brick veneer, and covers the brick units, several application
procedures and its advantages and disadvantages. It is not the purpose of this Technical Notes to cover all aspects
of the use of thin brick veneer, nor to make specific recommendations for installation. Future issues on the topic of
thin brick veneer will address, in detail, design, detailing and construction requirements.
Key Words: adhered veneer, brick, panels, prefabrication, thick set, thin brick, thin set.
INTRODUCTION
This Technical Notes addresses thin fired clay units, often referred to as thin brick, as interior or exterior wall
coverings. Thin brick veneer is a relatively new product which is seeing increasing popularity in commercial,
residential and do-it-yourself markets. The kinds of thin brick units discussed are formed from shale and/or clay, and
are kiln-fired. These thin brick units are much like facing brick (ASTM C 216), except they are approximately 1/2 to 1
in. (12 to 25 mm) thick. The face sizes are normally the same as conventional brick and therefore, when in place,
give the appearance of a conventional brick masonry wall. ASTM C 1088 Thin Veneer Brick Units made from Clay or
Shale covers two grades for exposure conditions to weather which are defined as Exterior and Interior. The three
types of thin veneer brick are based on appearance and are defined as TBS, TBX and TBA. Minimum compressive
strengths are not required in C 1088 as there is no way to test thin brick in compression.
There were early uses of thin brick. In the early 1950's, the Structural Clay Products Research Foundation (now the
Brick Institute of America) began the development of "SCR Re-Nu-Veneer", a 3/4 in. (19 mm) thick fired clay unit
which had Norman size nominal face dimensions (2 - 2/3 in. by 12 in. [68 mm by 305 mm]). The decision to begin
development of this product was due to marketing research which recognized remodeling and reveneering areas as
substantial markets for a thin clay veneer wall covering. In addition to developing the thin units, the Foundation
developed special clips to attach the units to an existing wall, mortar for grouting the joints and a power-driven
grouting gun. Locations were chosen to test the product, manufacturers were licensed to produce the units and
applicators were licensed to install the "Re-Nu-Veneer". After approximately 4 years of effort, work was
discontinued on the project.
Today, thin brick are being installed using a variety of procedures. In Japan and in the United States, thin brick have
been placed into forms and cast integrally with concrete, thus providing a very attractive architectural precast
concrete panel. Another procedure involves bonding thin brick to a 16 in. by 48 in. (406 mm by 1220 mm) substrate,
resulting in small, lightweight, easily installed modular panels. Ceramic tile installation techniques are often used to
install the brick units, either at the jobsite or on prefabricated panels, and homeowners are renovating with do-it-
yourself thin brick products.
This Technical Notes addresses thin brick units, several methods for installing thin brick, as well as some of the
advantages and disadvantages of thin brick veneer.
THIN BRICK UNITS
Thin brick are available in various sizes, colors and textures. The most commonly found face size is standard
modular with nominal dimensions of 2-2/3 in. by 8 in. (68 mm by 203 mm). The actual face dimensions vary slightly
among manufacturers, but are typically 3/8 in. to 1/2 in. (10 mm to 13 mm) less than the nominal dimensions. The
economy size unit is 50% longer and higher, but this difference goes virtually unnoticed since the aspect ratio (length
to height) is the same for both the standard and the economy modular units. The economy modular face size, 4 in.
by 12 in. (102 mm by 305 mm), is popular for use in large buildings because productivity is increased, and the unit's
size decreases the number of visible mortar joints, thus giving large walls a more pleasing appearance by reducing
http://www.gobrick.com/BIA/technotes/t28c.htm
1 of 9 9/13/2009 1:26 PM
the visual scale of the wall. Other sizes, such as Norwegian, 3-in. (76 mm), non-modular, oversize, etc., may be
available. Table 1 contains face sizes of several modular brick units; however, thin brick may not be available in each
size. It is advisable to check with individual manufacturers or distributors regarding sizes available in a particular
area. Figure 1 illustrates the various types of thin brick units.
TABLE 1
Nominal Modular Face Sizes For Brick
Thin Brick Units
FIG. 1
As with all other fired clay or shale products, color depends on the chemical composition of the raw material, the
intensity of firing and controls used in the firing. The color ranges for thin brick units are as unlimited as those for
other fired clay brick. The texture of thin brick units depends on the method of manufacture and the surface
treatment used prior to firing. Also, some manufacturers provide glazed thin brick units.
The physical properties, such as modulus of rupture and compressive strength of the thin units, depend on the raw
materials and methods of molding or forming the units. Fire resistance properties of thin brick construction have not
been evaluated. But, because the units are made from fire clay or shale, it is likely that the overall performance of a
wall system would be improved as compared to non-masonry sidings. Thermal resistivity of thin brick is probably not
significantly different than that of solid face brick; however, because of the unit's thickness, the overall resistance
would be very little. Likewise, because of the thickness, the units would contribute little mass for thermal storage.
http://www.gobrick.com/BIA/technotes/t28c.htm
2 of 9 9/13/2009 1:26 PM
Primarily, thin brick functions as an architectural wall covering that has the maintenance-free benefits of conventional
brick masonry. Secondarily, thin brick will provide some protection to the material over which it is applied. In
comparison to conventional brick masonry, thin brick will have less fire resistance, sound resistance, structural
strength, thermal mass or insulation properties.
METHODS OF THIN BRICK INSTALLATION
Adhered Veneer
Adhered veneer relies on a bonding agent between the thin brick units and the backup substrate. Adhered veneer
construction may be classified as either thin bed set or thick bed set.
Thin Set. The thin bed set procedure typically utilizes an epoxy or organic adhesive, and is normally used on interior
surfaces only. For areas subject to dampness, only clear and dry masonry surfaces or concrete surfaces should be
used for backup. For dry locations, the backing material (substrate) may be wood, wallboard, masonry, etc. A
cross-section depicting a wood frame wall upon which thin brick veneer (thin set procedure) is installed is shown in
Fig. 2.
Thin Set Interior Finish Over Wallboard
FIG. 2
Thick Set. The thick bed set procedure is used on interior and exterior surfaces. The backing material may be
masonry, concrete, steel or wood stud framing. The thick bed setting procedure over concrete masonry is illustrated
in Fig. 3. The wire lath shown in Fig. 3 may be eliminated if the masonry wall is heavily scarified (sand-blasted).
(Williams, Griffith, Jr., "New Bricklike Tile Veneer", Building Standards, July-August, 1982). For applications over
steel studs, procedures are similar to those used for concrete or masonry backup; however, wallboard and building
felt must be installed over the studs before the lath and mortar bed are placed. Thick bed setting of thin brick over
steel studs is shown in Fig. 4.
http://www.gobrick.com/BIA/technotes/t28c.htm
3 of 9 9/13/2009 1:26 PM
Thick Set Method Over Masonry
FIG. 3
Thick Set Method Over Steel Stud Framing
FIG. 4
Prefabrication
Prefabrication, utilizing thin brick veneer units, has been accomplished using the "casting" method. This process
involves the combination of thin brick, grout and/or concrete cast into a prefabricated panel (similar to architectural
pre-cast concrete). This process requires the use of forms, a method of placing the units, and a system for grouting.
The usual practice is to place the units face down into a form (or waffle mold), and place a very fluid grout over the
back surface of the units. The grout flows into the space between the units, thus forming the appearance of mortar
http://www.gobrick.com/BIA/technotes/t28c.htm
4 of 9 9/13/2009 1:26 PM
joints (see Fig. 5). Concrete and reinforcement are placed over the grout to provide structural support. The
installation of a completed panel is shown in Fig. 6.
Grouting Over Back Surface of Thin Brick Panel
FIG. 5
Erection of Panel
FIG. 6
The use of steel studs and the thick bed setting procedure is another method of prefabrication with thin brick (see
Figs. 4, 7 and 8). The use of thin brick for prefabrication of this type results in panels which are lighter than many of
the conventional prefabricated panel systems.
Positioning Prefabricated Panel
FIG. 7
http://www.gobrick.com/BIA/technotes/t28c.htm
5 of 9 9/13/2009 1:26 PM
Electric Winch Used To Lower Panels Over Edge of Slab
FIG. 8
There are several advantages of prefabrication over laid-in-place masonry. By using panelized construction, the need
for on-site scaffolding is eliminated, which can be a significant cost savings in masonry construction. If an off-site
plant is used, the work and storage areas for materials at the jobsite are reduced, resulting in a less congested
jobsite. If proper scheduling of delivery is maintained, the panels can be erected as they are delivered, eliminating
any need for panel storage at the site. One of the distinct advantages of the factory set-up is that it permits
year-round work and multi-shift workdays. The use of prefabricated masonry may eliminate the need for, or actually
provide, the means of winterizing the structure.
The use of panelization makes possible the fabrication of complex wall shapes. These shapes can be accomplished
with ease. Complicated shapes with returns, soffits, arches, etc., are accomplished by using jigs, forms and
templates. Repetitive usage of these shapes can lower costs appreciably.
Prefabrication allows for the use of stringent quality control. Mortar batching systems can be tightly controlled, and
curing conditions, temperature and humidity can also be controlled. Panelization on some projects may save
construction time. It is possible for the masonry panels to be built before ground-breaking for the project, thus
keeping far enough ahead of the in-place construction work to permit panel erection when needed.
As with any construction method, prefabrication has inherent advantages as well as disadvantages. The use of
prefabricated brick masonry is limited to use with certain types of construction. The designer should be aware of the
limitations of prefabricated masonry. The size of brick masonry panels is limited primarily by transportation and
erection requirements. Architectural plan layout may, in some cases, preclude the use of prefabricated brick
masonry. Another disadvantage of prefabricated brick masonry, as in other panel systems, is the absence of
adjustment capabilities during the construction process. In-place masonry construction permits the craftsman to build
masonry to fit the other elements of the structure by adjusting joint thicknesses over a large area so that they are
not noticeable. This is not possible with prefabricated elements. The use of prefabricated elements sometimes
requires that other trades build to accuracies beyond the standard construction tolerances of those trades.
Modular Panels
A relatively new method of thin brick application is becoming popular in the United States and Canada. Modular
panels are produced by several different companies and each system differs slightly. Basically, thin brick units are
adhered to modular panels in the factory, or at the jobsite. The modular panels have dimensions of approximately 16
in. by 48 in. (406 mm by 1220 mm), as shown in Fig. 9. The backing materials to which the brick units are adhered
may consist of polystyrene, polyurethane, cementitious board, asphalt-impregnated fiber board, plywood, aluminum,
or a combination of these materials, depending on the manufacturer.
http://www.gobrick.com/BIA/technotes/t28c.htm
6 of 9 9/13/2009 1:26 PM
Thin Brick Panel
FIG. 9
The panels weigh approximately 35 lb (16 Kg), which is light enough for one person to handle easily. Installation
techniques vary only slightly among the different manufacturers. At the time of this writing (1985) most, but not all,
systems require that the head and bed joints between the thin brick units be grouted after the panels are secured to
the supporting wall.
The application of the modular panels is illustrated in Fig. 10. Since the construction materials and application
methods vary among manufacturers, the user must select the panel with installation techniques and materials which
best fit the job requirements. Measures must be taken to prevent water penetration and subsequent corrosion,
especially for multi-story buildings which are subjected to severe weather conditions.
http://www.gobrick.com/BIA/technotes/t28c.htm
7 of 9 9/13/2009 1:26 PM
Modular Panels Over Frame Construction
FIG. 10
BUILDING CODE ACCEPTANCE
The model building codes (Building Officials and Code Administrators, International, 4051 West Flossmoor Road,
Country Club Hills, Illinois; Southern Building Code Congress International, Inc., 900 Montclair Road, Birmingham,
Alabama; International Conference of Building Officials, 5360 South Workman Mill Road, Whittier, California) do not
specifically address the usage of thin brick veneer in all of the methods of installation mentioned in this Technical
Notes. Thick set and thin set adhered veneer have been used for many years with thin brick units, ceramic tile and
architectural terra cotta; therefore, these methods are addressed in the model codes. The other methods, such as
prefabrication or modular panels, may have to be approved on a case-by-case basis, through research compliance
reports from the various model code agencies, or through code changes.
ADVANTAGES AND DISADVANTAGES OF THIN BRICK VENEER
Some of the advantages of thin brick veneer are:
1. Interior thin brick veneer finishes can be applied by homeowners or other moderately skilled craftsmen.
2. Thin brick veneer is more durable and longer lasting than aluminum, wood or vinyl sidings.
3. Prefabrication with thin brick veneer is easily and economically done.
4. Better sound and fire resistance properties may be obtained using thin brick veneer than with some non-masonry
sidings.
5. Thin brick units are more durable than imitation brick units made from gypsum, cement or plastics.
6. Walls built with thin brick units are lighter in weight than conventional masonry veneer.
7. Cleanup costs often incurred in conventional brick veneer construction may be reduced.
http://www.gobrick.com/BIA/technotes/t28c.htm
8 of 9 9/13/2009 1:26 PM
8. Year-round installation is possible.
9. May be used where structural support for conventional brick veneer is not available.
Some of the disadvantages of thin brick veneer are:
1. The durability and overall quality of thin brick veneer systems may not be equivalent to conventional brick veneer.
2. Thin brick veneer does not provide the structural properties of conventional brick veneer.
3. Sound and fire resistance properties are less than those of conventional brick masonry veneer.
4. Thin brick veneer does not provide the thermal mass of conventional brick veneer.
SUMMARY
This Technical Notes has discussed a relatively new product in the masonry industry - thin brick veneer. Walls faced
with thin brick veneer may look like conventional brick masonry walls, yet weigh considerably less. Thin brick veneer
is popular with homeowners for redecorating or renovating because the homeowners can obtain an attractive finish
and may do the work themselves.
Thin units are also used in commercial construction, applied one unit at a time, or applied in large prefabricated
panels. Small, lightweight, interlocking modular panels are available and are installed as a siding. Thin brick veneer
can provide the same architectural effects as conventional brick masonry, but does not have the same structural,
thermal or fire resistance qualities.
The information contained in this Technical Notes is based on the available data and the experience of the technical
staff of the Brick Institute of America. Final decisions on the use of information, details and materials as discussed in
this Technical Notes are not within the purview of the Brick Institute of America and must rest with the project
designer, owner, or both.
http://www.gobrick.com/BIA/technotes/t28c.htm
9 of 9 9/13/2009 1:26 PM

Technical Notes 29 - Brick in Landscape Architecture - Pedestrian Applications
July 1994
Abstract: This Technical Notes describes brick paving systems used in landscape design. Landscape architecture
and its relationship to brick masonry is covered. Master planning and environmental aspects of landscape
architecture are briefly discussed. Applications covered include patios, walks, steps and ramps. Materials and
methods of construction of flexible and rigid paving applications, citing the most critical requirements, are outlined.
Key Words: landscape architecture, patios, pavements, ramps, steps, terraces.
INTRODUCTION
Landscape architecture is the planning and design of elements relating to the land, including trees, plants, paving,
streets and sometimes structures. The design must take into account all of these elements and their relation to each
other. Brick as a landscape material is an important design element. Brick paving applications can be used to create
a pathway through the landscape, delineating pedestrian elements from natural elements. Since brick is made from
the earth and is small in scale, it fits into many landscaping plans.
This Technical Notes covers the topic of brick as it relates to landscape architecture. It also covers environmental
issues concerning the use of brick and brick paving systems in landscaping. Paving applications addressed include
patios, walks, steps and ramps. Design, installation and material selections are discussed. Other installation
practices that must be considered, but are not included in this Technical Notes, are edging, expansion joints and
membranes. Other Technical Notes in this series cover garden walls and other miscellaneous landscape
applications. Paving systems and related issues are discussed in more detail in the Technical Notes 14 Series [5,6].
LANDSCAPE ARCHITECTURE
The art of landscape architecture is more than the placement of trees and shrubs. Often it involves the development
and planning of large areas within cities and suburban areas. This is typically organized through the development of a
master plan. Alternately, landscaping may be on a much smaller scale, as in the design of a small garden. The
landscape architect must always consider certain issues, including aesthetics, harmony, continuity/unity,
accessibility, economy and other design parameters. Material and system selections are usually based on these
issues.
Materials can be broadly classified as either landscape materials or hardscape materials. Landscape materials
include trees, plants, grasses, soil and gravel. Hardscape materials include brick, stone, concrete and other hard
materials. A comprehensive landscape plan usually combines both landscape and hardscape features.
Master Planning
The landscape architect plays a much larger role with all land development issues today than in the past. Buildings
and their relationship and integration into the site have become increasingly important design issues. This may apply
to entire subdivisions and cities as well. A master plan is usually developed to incorporate all elements into a
comprehensive land development design. Master plans will dictate where open spaces should be located and
locations of buildings, pavements and walks. Brick can play an important part in the development of master plans
since it can be used as a common thread throughout an entire project. This includes walls, pavements, fountains,
planters, fences, steps and other miscellaneous landscape uses. Continuity throughout the project can be achieved
by using brick in many of these applications.
Environmental Issues
http://www.gobrick.com/BIA/technotes/t29.htm
1 of 10 9/13/2009 1:27 PM
There is now more pressure than ever to consider the environmental effects of a particular landscape plan. A
movement, often termed "sustainable development", considers the environmental impact of land development before,
during and after design. Environmental issues, such as storm water runoff and the lack of water, are becoming more
important as landscape architects look more closely at potentially threatening issues.
Sustainable Development. Sustainable development can be defined as development that meets the needs of the
present without compromising the future. In the past, sites were often dramatically altered without adequate
consideration of the environmental impact. As with all designs, compromises must be made to achieve the design
requirements. Sustainable development takes into account the effects of materials used in the landscaping plan on
the environment. The embodied energy and the effects of the manufacturing of the material on the environment are
closely considered. Brick is a material made from clay and shale, some of the earth's most abundant materials. The
energy used to make brick, which is termed its embodied energy, is less than that of concrete, steel and many other
materials [3]. Since brick is inert, it does not pose any long-term environmental threats.
Water Issues. Environmental concerns have been raised regarding both storm water runoff and the lack of water in
some areas. When many parts of the landscape are being covered by impermeable surfaces, storm water runoff
becomes a larger problem. The amount of water that drains off of a shopping center parking lot, for example, can be
quite large causing flooding or erosion. Thus, the size of storm sewers and catch basins must be increased in size
accordingly, putting more stress on the infrastructure. Conversely, some areas of the country are so arid, they
cannot support plant life.
Porous Pavements - Most hardscaping materials, such as concrete or asphalt, will not allow water back into the
ground. This is also true of rigid (mortared) brick pavements and some flexible (mortarless) brick pavements over an
impermeable base. These systems can have a negative effect in urban areas which include trees as a part of the
urban landscape. Trees can die due to lack of water and nutrients when surrounded by impervious hardscapes. In an
effort to provide water for trees, grates have been used, but their small size can inhibit proper tree growth. Soil and
mulch have also been used around trees, but usually become compacted and allow rain to evaporate away too
quickly. When water infiltration into the ground is desired, a pavement which allows water to percolate back into the
ground should be used [4]. One alternative that can help water infiltration and reduce storm water runoff is the use of
porous pavements.
A porous pavement allows water to filter through it, percolate back into the ground and replenish the ground water.
In most cases, mortarless brick paving over an aggregate base can be constructed to allow water to percolate into
the ground. However, to allow more percolation to occur, the pavement must have joints between the pavers at least
1/4 in. (6 mm) wide. The joints allow water to enter easily and permeate to the base. The base should be an
open-graded aggregate, such as free- draining gravel or sand, to allow percolation. Although porous pavements
allow storm water runoff to be directed back into the ground water system, it may go against usual pavement design
practice. In a flexible brick pavement, it is desirable to have a dense base to resist loads from traffic above and from
frost heave from below. Using an open-graded base may not provide the stable base that is needed.
To use an open-graded base under brick paving, some simple recommendations must be followed. The open-graded
base must be compacted appropriately. Guidelines exist for the proper construction of open-graded bases [6]. A
membrane, such as a geotextile or filter fabric membrane, must be placed between the sand setting bed and the
open-graded base to avoid settling of the sand into the voids of the base. A geotextile may also be required
between the base and the soil or subgrade to prevent soil from pumping up into the base. The size of the joints can
be problematic when large amounts of water constantly run across the pavement. Jointing sand may wash out in
areas when the joints are larger than 1/4 in. (6 mm). Another issue to consider is that interlock of the pavers will not
be achieved when the joints are larger than 1/4 in. (6 mm). Interlock of the pavement occurs when the pavers are
compacted into the sand setting bed and the entire pavement - i.e. pavers, setting bed, and base - lock together and
act to withstand the loads as a single element. A flexible brick paving system can be designed for improved
percolation, but interlock of the pavement cannot be expected when sand-filled joints are larger than 1/4 in. (6 mm).
Xeriscapes - Xeriscapes are defined as water-efficient landscapes which not only require less water to grow
vegetation, but have a reduced need for mowing, fertilizing and pesticide application. They may be used in certain
areas of the country, such as parts of Southern California, Arizona and Nevada, where water supplies are low or
unreliable. Instead of introducing planting areas that require large amounts of water, it may be prudent to use brick in
the place of plants. In this manner, reliance on the local water system, expensive watering systems and plant
maintenance can be reduced. Wildfires in arid regions of the country are another concern which may require the use
of non-combustible materials adjacent to homes. Brush and shrubs can act as fuel sources for wildfires. Brick paving
http://www.gobrick.com/BIA/technotes/t29.htm
2 of 10 9/13/2009 1:27 PM
adjacent to the house can act as a fire break. To offset the use of all of the hardscape materials, patterns are laid in
the pavement to give the impression of plantings. Obviously, this must fit in with the entire landscaping plan.
Aesthetics
One of the most important features that a landscape architect faces is that of appearance. The look of any design
can evoke strong feelings, good or bad. So it is important that the aesthetics of the project be examined closely. As
in all architecture, form, color and pattern are the vehicles for achieving a certain aesthetic appeal. Brick paving
utilizes its variety of size, shape, color and pattern to conform to the chosen theme. Brick pavers are produced in a
variety of sizes. The most common sizes are shown in Table 1.
1Check with manufacturer for availability of chamfers.
Alternative sizes and shapes of pavers can also be manufactured or cut from standard units into the desired shape.
For example, radial brick are often used to create curves or circles in the pavement, as shown in Figure 1.
Special Brick Shapes
http://www.gobrick.com/BIA/technotes/t29.htm
3 of 10 9/13/2009 1:27 PM
FIG. 1
The color of brick pavers range from buffs to dark browns, pinks to deep reds. The pavers can be a uniform color or
there can be a range of colors. The color of brick will not fade over time. Different colors can be arranged within a
pavement to achieve a truly dramatic look.
Almost any pattern is possible with brick. The pattern can be simple diagonals or more complicated cross or weave
patterns. Different colored units can be used to create a flow pattern for pedestrian traffic. It may suggest a special
theme used throughout the landscape plan. The more traditional brick paving patterns are shown in Fig. 2. Brick can
also be cut to achieve a pattern; although in some cases, specially shaped pavers may be used.
Brick Paving Patterns
FIG. 2

BRICK PAVING SYSTEMS
Brick paving can be classified by two basic systems; flexible and rigid. Flexible brick pavements usually consist of
mortarless brick paving over a sand setting bed and an aggregate base. Rigid brick pavements consist of mortared
brick paving over a concrete slab. Mortarless brick paving can be used over any base. Mortared brick paving must
be supported by an adequate concrete slab or the mortar joints or pavers may crack if the base is not sufficiently
rigid. Examples of flexible and rigid brick pavements are shown in Figs. 3 and 4, respectively.
http://www.gobrick.com/BIA/technotes/t29.htm
4 of 10 9/13/2009 1:27 PM
Flexible Brick Paving
FIG. 3
Mortarted Brick Paving
FIG. 4
Although a flexible brick paving system is generally recommended, there are certain applications where rigid brick
paving is desired. An example is brick steps, which requires the edges to be mortared together to keep the brick in
place. The major advantages of using a flexible pavement include easier repairs to utilities beneath the pavement
and usually lower installation costs.
The design of brick paving systems can be rather complex, depending on the size of the project. Technical Notes 14
Series discusses many of the design and construction parameters in more depth than in this Technical Notes. Only
critical or unique information is contained here.
Patios and Walks
Some of the most widely used features of landscape design to which brick is adapted are patios and walks. Patios
may be outdoor extensions of the indoor living space and supplement the activities of the occupant. Patios are often
adjacent to living, family or dining rooms. Patios may be built as terraces, which are raised levels of earth supported
on one or more sides by a wall or bank. A terrace is used to extend living space along a hillside.
Walks are often effectively used to provide an interesting and inviting entrance path to a garden or home. Walks may
be used to define pedestrian travel routes and can provide geometric patterns in formal garden layouts. They can
also serve as a path through a natural or garden setting.
Bases. The proper design and construction of the base is often the most critical element for long-term performance
of the paving assembly. Insufficient base thickness or improperly compacted bases will lead to undulations (rutting)
or cracking of the pavement. Appropriate base thickness depends on the type of loading and weathering it will
receive. Most residential patios and terraces will only receive pedestrian traffic; therefore, the thickness may depend
more on its resistance to frost heave. The minimum recommended base thickness is 4 in. (200 mm) for concrete,
asphalt and aggregate bases. Thicker bases and the use of a subbase may be required in areas with poor soil
conditions or soils that are constantly saturated. In these cases, the base should be increased in thickness based on
local requirements.
Drainage. Drainage is another key design feature which affects long-term performance. Poor drainage will allow
water to stand on the pavement and saturate the brick pavement. Problems resulting from poor drainage include
deterioration of the paving, moss and algae growth and slippery pavements. Therefore, it is important to slope the
pavement to keep water from collecting. Primary drainage of all pavements should occur on the surface. Drainage
should occur away from buildings or other walls. For brick pavements, a slope of 1/8 to 1/4 in. per foot (1 to 2 mm
per 100 mm) is recommended. Lesser slopes will allow water to accumulate. Steps and ramps must also be sloped
to avoid standing water. Treads of steps should slope 1/8 to 1/4 in. per foot (1 to 2 mm per 100 mm). Cross-slopes
of the pavement help drain water off of the pavement, but the slope should not exceed 3 percent. Flexible brick
pavements may allow some water to percolate down into the ground. In this case, subsurface drains may be
necessary to remove water from the system.
http://www.gobrick.com/BIA/technotes/t29.htm
5 of 10 9/13/2009 1:27 PM
Steps and Ramps
Steps and ramps are used to connect different levels for easy access. Steps have traditionally been used to allow
movement up and down steep slopes or within structures. The size and configuration of steps is governed by a
combination of physical human dimensions and aesthetics. Ramps are used on gentle slopes and are used to
provide access for the physically impaired. Since ramps have low slopes, they will require more space than steps.
Brick has been used successfully in all configurations of steps and ramps mainly because it is a small element which
permits numerous configurations. Model building codes often dictate certain criteria for steps and ramps such as
riser-tread relationships and minimum slip resistance. Other design issues that should be considered include
structural support of the steps or ramps and other safety issues. Steps should be a minimum width of 60 in. (1.5 m)
for public spaces, or 42 in. (1.1 m) for private residences. There should be at least two, preferably three or more
steps, in a stepped walkway, since single steps can be overlooked and lead to trips.
Step Riser - Tread Relationships. Riser-tread relationships have been studied for many years. Most steps are
constrained to fit into a set elevation at the top and the bottom of the steps. For these steps, riser-tread
relationships have been developed for safe and efficient use. In other areas, such as plazas, the tread dimension
may not be constrained, which allows freedom of design. In these areas, riser and tread dimensions will be dictated
by human dimensions and appearance.
Most local building codes mandate the minimum and maximum riser dimension and minimum tread dimension. The
7-11 rule is used most frequently; that is, maximum riser height is 7 in. (180 mm), while the minimum tread depth is
11 in. (280 mm). The riser must also be greater than 4 in. (100 mm) in height. Due to normal walking and gait,
optimum riser-tread dimensions do exist. One formula for determining this relationship has been recommended for
use [1]. It provides a general guideline for riser and tread dimensions.
T x R = 77.5 (T x R = 500, for SI units)
where:
T = tread width, in. (cm)
R = riser height, in. (cm)
In this equation, the riser is restricted between 4 in. (10 cm) and 7 in. (18 cm). Since brick is a small element, there
are a variety of bonding patterns for the tread and riser. Figure 5 shows several examples of bonding arrangements
with modular brick.
Step Configurations
FIG. 5
Landings may be included in steps, especially when the cumulative height of steps is great. Building codes set the
maximum height between landings at 12 ft (3.7 m); however, it is usually more desirable to limit the maximum height
to 5 ft (1.5 m) between landings. The length of the landing should be long enough to allow easy cadence, which is
about 5 ft (1.5 m) or a multiple of 5 ft (1.5 m).
Edge Details. For safety reasons, several issues relating to the edge of the steps should be considered. In most
http://www.gobrick.com/BIA/technotes/t29.htm
6 of 10 9/13/2009 1:27 PM
pedestrian applications, it may be beneficial to highlight the edge of the step by varying the color of the tread and
riser brick, changing the bond pattern of the tread or by extending the tread slightly over the riser. These distinctions
of the edge allow easy visual indication that a change is coming, which helps to avoid tripping. Extending the tread
over the riser creates a shadow line highlighting the stairs and also allows the treads to be slightly deeper than if
they were squared off. However, the maximum projection should be 1 1/2 in. (40 mm) to avoid catching the foot
while stepping up. If a rounded tread is used, the leading edge should be a maximum 1/2 in. (13 mm) radius.
Ramps. The model building codes and other accessibility codes [2] usually limit the slope of ramps to no steeper
than 1:12 in most applications. Greater slopes are allowed only if the total rise is less than 6 in. (150 mm). Slopes
greater than those allowed by codes make it difficult for persons in wheelchairs to negotiate the ramp. The width of
a ramp should be at least 3 ft (0.9 m) wide for one-way traffic and 5 ft (1.5 m) wide for two way traffic. In addition,
landings should be provided every 30 ft (9 m) horizontally.
Support and Bonding. Brick steps and ramps are usually supported by a concrete base, but any material capable
of supporting the brick properly could be used, if designed properly. Deflections or settlement of the support must be
minimized to avoid cracking in the brickwork. Figure 6 shows a concrete support system for a step and ramp. Brick
should be adequately bonded to the support or restrained around its perimeter to avoid loosening of units. Mortar is
usually used to bond the brick to the concrete. This paving system is very effective when proper materials and
installation are used. Dowels or ties into the mortar joints are not necessary since the mortar provides adequate
bond. Newer types of adhesives are now being used to bond the brick directly to the concrete. These adhesives
must be durable to withstand the severity of its environment. Adhesives can only be used when the concrete surface
is fairly even and free of contaminants. Caulks and sealants are not appropriate for this purpose.
Stair and Ramp Sections
http://www.gobrick.com/BIA/technotes/t29.htm
7 of 10 9/13/2009 1:27 PM
FIG. 6
Adequate footings should be designed for the step or ramp support. The depth of the footings should extend below
the frost line. Since the paving assembly is supported on its own footing, an isolation joint should be used between
the pavement and building and between the pavement and ramps or steps.
Safety. In addition to the required physical dimensions of the element, the slip resistance of the surface should also
be considered. The static coefficient of friction is usually used to determine if a surface is considered slippery. There
is no consensus minimum value for coefficient of friction; however, some codes are promoting minimum static
coefficient of friction values of 0.6 for pavements and 0.8 for ramps [2]. Limiting the static coefficient of friction to
these values is believed to make the surfaces safe and passable by both able-bodied pedestrians and the physically
impaired. It is usually excessive water- ponding or the contamination of the pavement by other substances which
cause most slips and falls. Brick usually has an adequate slip resistance, with higher coefficient of friction values
achieved when a rough textured brick is used.
SELECTION OF MATERIALS
Pavements can be subjected to severe weather and abrasion; therefore, the materials used to construct them must
be of superior quality. Most of the materials in a pavement must conform to ASTM standards. Following are
recommendations for the selection of paving materials. Additional information on material selection can be found in
Technical Notes 14 Series
Brick Pavers
Pavers must be able to withstand the weather and the abrasion of pedestrian traffic. Pavers should conform to the
requirements of ASTM C 902 Specification for Pedestrian and Light Traffic Paving Brick. Units conforming to ASTM
C 1272 Specification for Heavy Vehicular Paving Brick may be used, but are usually not necessary for most
landscape applications, unless heavy vehicular traffic is expected. Heavy vehicular traffic is composed of high
volumes of heavy vehicles on a pavement. Two of the more critical requirements of ASTM C 902, durability and
abrasion, are discussed below. Other requirements, such as dimensional tolerances, chippage and warpage should
also be considered.
Durability. The resistance of pavers to weathering is determined by the Class of the paver. The Class of the paver
is based on the durability of the unit and is determined by compressive strength, cold water absorption and
saturation coefficient of the unit. Class SX pavers are intended for use where the paver may be frozen while
saturated with water. In exterior applications where freezing is not present, pavers should conform to Class MX or
SX. Class NX pavers are acceptable for interior use where they are protected from freezing when wet. Alternate
means of assessing durability of brick pavers are addressed in ASTM C 902 and Technical Notes 14A Revised.
Abrasion Resistance. Pavers must be able to resist the abrasive action of traffic. ASTM C 902 includes three
abrasion classifications of pavers; Types I, II and III. Type I pavers are appropriate for areas receiving extensive
abrasion, such as commercial driveways and entrances. Type II pavers are intended for exterior walkways and
floors in restaurants and stores. Type III pavers are used for residential floors and patios. The paver Type is
determined by its abrasion index, which is calculated by dividing the cold water absorption by the compressive
strength and multiplying by 100. The resistance to abrasion can also be determined by a laboratory test, as outlined
in ASTM C 902.
Setting Bed Materials
The setting bed, placed between the base and the brick pavers, functions as a leveling course for slight irregularities
in the base and units. Setting bed materials include sand, mortar, asphalt and building felt.
Sand. Sand used as a setting bed should be a washed, well-graded angular sand with a maximum particle sized of
3/16 in. (4.8 mm). Sand should conform to ASTM C 33 Specification for Concrete Aggregates, usually referred to as
concrete sand. Mason's sand can also be used as the setting bed material when the thickness of the setting bed is
less than 1 in. (25 mm). Mason's sand should conform to ASTM C 144 Specification for Aggregates for Masonry
Mortar. The thickness of the sand setting bed should be between 1/2 in. and 2 in. (13 mm and 50 mm).
Mortar. Mortar setting beds are used in mortared brick paving applications. Mortar should conform to ASTM C 270
http://www.gobrick.com/BIA/technotes/t29.htm
8 of 10 9/13/2009 1:27 PM
Specification for Mortar for Unit Masonry or ANSI A118.4 Specification for Latex-Portland Cement Mortar, when a
latex additive is used. For exterior mortared brick pavements, Type M mortar is preferred. Type M portland
cement-lime mortar consists of 1 part portland cement, 1/4 part hydrated lime and 3 3/4 parts sand. Type S mortar
can be used alternately. In severe freeze/thaw environments, mortars with better freeze/thaw resistance should be
used. Two mortar properties that greatly influence freeze/thaw resistance are air content and water/cement ratio. An
air content for paving mortars between approximately 10 percent and 15 percent is optimal. Increasing air content
too much will reduce bond between the brick paver and mortar. To address the water/cement ratio, the mortar
should be mixed with just enough water to make it workable. The thickness of the mortar setting bed should be
between 3/8 in. and 1 in. (10 mm to 25 mm).
Asphalt. Asphalt setting beds typically consist of approximately 7 percent asphalt and 93 percent sand. The asphalt
setting bed is used over a concrete or asphalt base. The thickness of the asphalt should be approximately 3/4 in. (20
mm).
Building Felt. Brick may be placed directly on a new or existing asphalt or concrete base. In these applications,
building felt may serve as a cushion between the pavers and the base, which can accommodate small dimensional
variations of the base and pavers. Two layers of No. 15 building felt or one layer of No. 30 building felt is
appropriate.
Base Materials
Base materials consist of crushed aggregate, gravel, sand, asphalt and concrete. Asphalt and concrete bases often
require an aggregate subbase. Steps and ramps usually are built on concrete bases, whereas patios and terraces
may be supported on any of the base materials listed.
Aggregate Bases. Aggregate bases include crushed stone, gravel and sand. Heavier loading or areas subjected to
frost heave may require crushed stone. Open-graded aggregate (gravel) is often used in areas of poor drainage,
areas subjected to frost heave or when porous pavements are designed. The proper aggregate size depends on the
depth of the layer and the size of the compaction equipment. Maximum aggregate size is usually 3/4 in. (20 mm)
diameter.
In residential pedestrian applications, sand bases can be used when the subgrade compaction is ensured, when
bearing on undisturbed earth and in areas where frost heave is not a consideration. Sand used as a base material
should be a concrete sand conforming to ASTM C 33 and be clean and free of deleterious materials.
Concrete Bases. New or existing concrete bases may be used to support brick paving. New concrete should be
installed following recommended concrete practices. Where mortar is used to bond brick pavers to the concrete, the
concrete should have a rough textured finish. Caution should be used if brick is placed over an existing concrete slab.
The existing concrete slab must be sound and any major cracks filled adequately with concrete or mortar.
Asphalt Bases. New or existing asphalt bases may used to support mortarless brick paving. Proper asphalt
materials are generally determined by paving contractors or the asphalt plant and are beyond the scope of this
Technical Notes. Asphalt bases should not be used to support mortared brick paving.
CONSTRUCTION
One of the most important factors in long-term pavement performance is proper installation. Critical elements include
proper base compaction, proper edge restraints and full mortar joints, if used. There are numerous ways to install
brick paving, and techniques tend to vary by region. The recommendations in this Technical Notes are based on
experience and provide a minimum level of workmanship necessary for satisfactory performance. More information
on construction of brick pavements may be found in Technical Notes 14 Series.
Base Preparation
Proper compaction of the subgrade (soil) and base is one of the most critical factors in pavement installation.
Pavements rely on the strength of the base to adequately resist loads. Poorly compacted aggregate bases usually
lead to undulations (rutting) or cracking of the pavement. Most patios, steps and ramps are built adjacent to a
building or residence. These are often areas over backfill rather than undisturbed earth, making proper compaction
even more critical. Compaction of the subgrade should be done with the largest equipment possible so that proper
compaction is achieved. Once the subgrade is compacted, the base may be built on top. The base material should
http://www.gobrick.com/BIA/technotes/t29.htm
9 of 10 9/13/2009 1:27 PM
be spread and compacted in layers or lifts. The thickness of each layer must be consistent with the size of
compaction equipment, but never exceed 4 in. (100 mm). Vibratory rollers may be necessary, although for most
residential applications plate compactors provide enough force. Typical compaction criteria is 95 percent maximum
density.
Mortar Installation
When brick are installed with mortar, standard bricklaying or tile setting procedures should be followed. The
preferred method of mortar placement is with a trowel. The concrete base should be clean and slightly dampened,
but surface dry prior to placing the mortar. Brick pavers should be buttered on the ends and shoved into the mortar
setting bed. All joints intended to receive mortar should be solidly filled. The joints should be tooled with a metal
jointer. A shallow concave joint profile is preferred for proper compaction of the joint and to keep water from ponding
on the mortar joints.
CONCLUSION
This Technical Notes describes elements considered in landscape design including master planning and
environmental aspects of landscape architecture. Design of pavements, steps and ramps is discussed. Materials
and construction of these elements and the most critical requirements for each are outlined.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Institute of America. The information contained herein must be used
in conjunction with good technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the purview of the Brick
Institute of America and must rest with the project architect, engineer and owner.
REFERENCES
1. Abdou, O. and Burdette, J., "Masonry Stairs: Design and Material Performance," Proceedings Sixth North
American Masonry Conference, Philadelphia, PA, June 1993, pp. 739-754.
2. "Americans with Disabilities Act: Accessibility Guidelines for Buildings and Facilities," U.S. Architectural and
Transportation Barriers Compliance Board, Washington, D.C., August 1992.
3. "AIA Environmental Resources Guide," American Institute of Architects, Washington, D.C., 1993.
4. Arnold Associates, "Urban Trees and Paving," Princeton, NJ, August 1982.
5. "Brick Floors and Pavements, Design and Detailing," Technical Notes on Brick Construction 14 Revised,
Brick Institute of America, Reston, VA, September 1992.
6. "Brick Floors and Pavements, Materials and Installation," Technical Notes on Brick Construction 14A
Revised, Brick Institute of America, Reston, VA, January 1993.
7. "Flexible Pavement Guide for Roads and Streets," National Stone Association, Washington, D.C., January
1985.
8. Harris, C.W. and Dines, N.T., Time-Saver Standards for Landscape Architects, McGraw-Hill Book Co.,
New York, NY, 1988.
http://www.gobrick.com/BIA/technotes/t29.htm
10 of 10 9/13/2009 1:27 PM

Technical Notes 29A - Brick in Landscape Architecture - Garden Walls
Rev [Nov. 1968] (Reissued Jan. 1999)
INTRODUCTION

The plan of a garden usually involves "leading" the sojourner through a series of spatial relationships. This can be
done formally, informally or subtly, depending upon the purpose and skill of the designer. Among the tools used for
this purpose are garden walls of brick. They may invite, enhance, lead, restrict, compel, separate, combine, protect,
screen or prohibit; all to the purpose of the artist and his skill.
The variations and possibilities with brick garden walls approach infinity. Still among the most popular of garden wall
structures is the serpentine wall. The serpentine wall is believed by some to predate recorded history. Early
examples of its use in America may be found in Colonial Williamsburg, and at the University of Virginia in
Charlottesville (Fig. 1) where over 160 years ago Thomas Jefferson designed and built serpentine walls.


Brick Serpentine Wall - University of Virginia
FIG. 1

MATERIALS AND WORKMANSHIP

29a http://www.gobrick.com/BIA/technotes/t29a.htm
1 of 8 9/13/2009 1:29 PM
Obviously, most garden walls will be subjected to the extremes of exposure to the elements. Rain, snow, freezing
weather, heat, direct sun and combinations of these will severely test the quality of the materials, the workmanship
and the design of garden walls. Consequently, the selection of materials and the workmanship are of paramount
importance in constructing successful and enduring garden walls of brick.
Brick. Brick for garden walls should meet the requirements for grade SW of the ASTM Standard Specifications for
Facing Brick C 216 (where exposed to view) or ASTM Standard Specifications for Building Brick C 62 (where not
exposed, such as below grade), unless the brick are known to have performed satisfactorily under similar conditions
of exposure.
Used or salvaged brick should not be used for garden walls unless they meet the SW requirements. Used brick
generally will not meet these requirements, since they often are non-uniform in degree of burning (see Technical
Notes 15, "Salvaged Brick").
Mortar. The recommended mortar for reinforced and non-reinforced brick garden wall construction is composed of:
1 part portland cement, 1/2 part hydrated lime and 4 - 1/2 parts sand by volume. This mortar conforms to ASTM
Standard Specifications for Mortar for Unit Masonry, C 270, Type S. It is very durable and develops good bond with
clay masonry units (see Technical Notes 8 Revised, "Portland Cement-Lime Mortars for Brick Masonry").
Workmanship. Workmanship on brick masonry garden walls should be that characterized by the complete filling of
all head, bed and collar joints. It is sometimes erroneously supposed that workmanship of lesser quality may be
tolerated in garden structures, since they are not weather barriers or enclosure walls, as in a building. Less than
excellent workmanship should not be permitted. Because of the extreme exposure of garden walls, any defect may
result in deterioration and perhaps ultimately in failure of the wall (see Technical Notes 7B Revised, "Water
Resistance of Brick Masonry-Design and Detailing").

DESIGN

The most meticulous design and details may be thwarted by the selection of inappropriate materials, or by poor
workmanship. The converse is also true. The use of the best possible materials and craftsmanship will not in
themselves insure a successful and permanent garden wall. Much depends on the design and attention to certain
critical details, such as foundations, copings and flashing.
Foundations. Since garden walls generally do not carry any vertical loads other than their own weight, the footings
and foundations sometimes do not receive proper attention. Foundations for garden walls should be placed in
undisturbed earth at a depth below the frost line. They may be of brick or of concrete and reinforced if necessary,
particularly in areas with active soils.
Copings. The appearance and character of a garden wall may be altered materially, depending on the type of cap
or coping used on the top of the wall. Many materials are suitable for coping, including: natural stone, cast stone,
slate, terra cotta, metals and brick. Figure 2 indicates a few suggestions for capping garden walls with various
materials.


29a http://www.gobrick.com/BIA/technotes/t29a.htm
2 of 8 9/13/2009 1:29 PM

Typical Copings
FIG. 2a



Typical Copings
FIG. 2b

29a http://www.gobrick.com/BIA/technotes/t29a.htm
3 of 8 9/13/2009 1:29 PM


Typical Copings
FIG. 2c

Whichever is used, there are certain general recommendations that should be followed. The coping should project
beyond the face of the brick a minimum of 1/2 in. on both sides. This projection should contain a positive drip to keep
water from flowing down the face of the wall. In most cases copings should be anchored to the brick wall with metal
anchors or bolts. If the coping is of material other than brick, its thermal and moisture expansion characteristics
should be compared to those of the brick masonry, and provisions made for the resulting differential movements
(see Technical Notes 18, "Differential Movement").
Flashing. In general, through-wall flashing is recommended immediately under the coping of garden walls. However,
this may be tempered with good judgment, depending on several factors, such as: the type of coping used, many
joints or relatively few; the climatic conditions of the area, high or low precipitation; and number of freezing and
thawing cycles (see Technical Notes 7A Revised, "Water Resistance of Brick Masonry - Materials").

WALL TYPES

There are many types of brick garden walls with new variations constantly appearing. Perforated walls of brick are
often used to complement contemporary architecture. Perforated walls can be designed in conjunction with any of
the typical types which are discussed below. If perforated walls are designed, the indicated dimensional proportions
and reinforcing steel requirements indicated under the specific wall types should be followed.
Straight Walls. This form of garden wall relies upon texture and color of the brick masonry for character. The
straight garden wall must be designed with sufficient thickness to provide lateral stability. Assuming the finished
grade on both sides of the wall is the same, so that no stresses result from earth pressure, it must be designed to
resist wind and impact loads. For 10 - psf wind pressure, it is recommended that the height above grade (h) not
29a http://www.gobrick.com/BIA/technotes/t29a.htm
4 of 8 9/13/2009 1:29 PM
exceed three-fourths of the wall thickness (t) squared where h and t are in inches. This recommendation
is based on the assumption that the overturning moment from the wind does not exceed three-fourths of the righting
moment from the wall dead load, and does not depend upon bond between the foundation and the wall. If bond is
assured or steel dowels are used to tie the wall to the foundation, or if greater lateral loads exist, straight walls can
be designed in accordance with the Recommended Building Code Requirements for Engineered Brick Masonry,
BIA, August 1969; or "Building Code Requirements for Reinforced Masonry", ANSI A41.2-1960 (R1970).

Pier and Panel Walls. The pier and panel wall is composed of a series of relatively thin panels, 4 in. in thickness,
which are braced intermittently by masonry piers (Fig. 3). This wall is relatively easy to build and is economical due
to the reduced thickness of the panels. It is also easily adapted to varying conditions of terrain.


Pier and Panel Garden Wall - Montgomery Village, Maryland
FIG. 3

Details for construction of pier and panel walls are shown in Fig. 4. Three typical design possibilities are indicated.
Table 1 provides the horizontal reinforcing steel requirements for panel walls and Table 2 gives the vertical
reinforcing steel for the piers. The required foundation diameter and depth of embedment are given in Table 3.


29a http://www.gobrick.com/BIA/technotes/t29a.htm
5 of 8 9/13/2009 1:29 PM

Pier and Panel Garden Wall
FIG. 4

Note:A= 2-No. 2 bars
B = 2 - 3/16 - in. diam. wires
C = 2 - 9 gage wires

29a http://www.gobrick.com/BIA/technotes/t29a.htm
6 of 8 9/13/2009 1:29 PM
1 Within heavy lines 12 by 16-in. pier required. All other values obtained with 12 by 12-in. pier (see Fig. 4).

1Within heavy lines 23-in. diam. foundation required. All other values obtained with 18-in. diam. foundation (see Fig. 4).

The reinforced brick masonry piers and panels are designed with an allowable flexural compressive stress for brick
masonry of 725 psi and an allowable tensile stress for reinforcing steel of 20,000 psi for No. 2 and larger bars (min.
yield strength 40,000 psi) and 24,000 psi for 3/16 -in. diam. and 9 gage wire (min. yield strength 60,000 psi). The
ratio of modulus of elasticity of the steel to that of the masonry (n) is assumed to be 13. The design is in compliance
with the Recommended Building Code Requirements for Engineered Brick Masonry, BIA, August 1969, and
"Building Code Requirements for Reinforced Masonry", ANSI A41.2-1960 (R1970). One of these standards should
be used to determine the required brick compressive strength. The pier foundation diameter and required
embedment below grade are based upon an average allowable soil pressure of 3000 psf. If poorer soil exists at the
site, a further investigation of the foundation is necessary.
The reinforcing steel requirements vary with the wind load, the wall height and the wall span. The two vertical pier
bars are placed 2 in. apart in the 12 by 12 - in. piers and 6 in. apart in the 12 by 16 - in. piers, one bar toward each
face. These bars are to be continuous from the bottom of the foundation to the top of the brick masonry pier. If
spliced, for ease of construction, they must be lapped at least 30 bar diameters. The panel wall reinforcement
consists of two bars or wires spaced not less than 2 in. apart in the mortar joint at a vertical spacing not to exceed
the dimensions given in Table 1. Individual bars or wires, or prefabricated joint reinforcement may be used. It is also
required that the panel wall reinforcement be continuous throughout the length of the wall. Where spliced, it should
be lapped 16 in.
Construction of these panel walls is simple, since foundations are required only under the piers. These may be
constructed by drilling or digging holes with the diameter given in Table 3.
The foundations should be placed in undisturbed earth and extend below the frost line, but not less than the required
embedment given in Table 3. The masons build the brick panel walls between the piers with a temporary 2 by 4 - in.
wood form under the first course of brickwork. The 2 by 4 - in. forms are removed after the wall has cured at least
seven days. The same mortar used to lay the brick may be used to grout the vertical reinforcing steel in the pier, if
sufficient water is added so that it flows readily and completely surrounds the steel.
Serpentine Walls. This ingenious technique of garden wall construction has been used for several hundred years.
The serpentine shape provides lateral strength to the wall so that it normally can be built only 4 in. in thickness
without additional lateral support. Since the serpentine wall depends on its shape for lateral strength, it is important
that the degree of curvature be sufficient. The following general rule is based upon the performance of many
successful serpentine walls. The radius of curvature of a 4 - in. wall should be no more than twice the height of the
wall above finished grade, and the depth of curvature should be no less than 1/2 of the height. Figure 5 provides
details of a typical serpentine garden wall.
The general rule stated below is limited in application to minor garden walls.

29a http://www.gobrick.com/BIA/technotes/t29a.htm
7 of 8 9/13/2009 1:29 PM

A Typical Serpentine Garden Wall
FIG. 5


29a http://www.gobrick.com/BIA/technotes/t29a.htm
8 of 8 9/13/2009 1:29 PM

Technical Notes 29B - Brick in Landscape Architecture - Misc. Applications
[Apr. 1967] (Reissued May 1988)
INTRODUCTION
In this issue of Technical Notes are suggestions for the use of brick in landscape architecture that take advantage of
the unique practicality and permanent beauty of the material. The color and texture of brick will complement the
masses and lines of contemporary architecture. And for traditional architecture, brick lends the same charm that has
endured for more than a century and a half on the grounds surrounding the splendid mansions of Colonial America.
The success of landscape architecture depends upon the intelligent spacing and inter-relationship of the various
elements, understanding the limits and possibilities of structural and plant materials, discreetly evaluating the
weakness and strength of various colors and textures, and choosing materials which are sympathetic with the site,
with each other and with the people who are to see and use them.
Brick is an ideal material for use in landscape architecture for it is made of natural earth material, clay or shale,
burned to permanent hardness. In the manufacturing process, brick take on colors which we know as earth colors -
reds, browns, buffs and yellows - which are entirely harmonious with nature.
MATERIALS AND WORKMANSHIP
Most garden and landscape structures will be subjected to the extremes of exposure to the elements. Therefore,
proper selection of materials and high quality workmanship cannot be emphasized too strongly.
Brick. Brick for garden structures should meet the requirements for grade SW of ASTM Standards for Facing Brick,
C 216 (where exposed to view) or ASTM Standard Specifications for Building Brick, C 62 (where not exposed such
as below grade). Used or salvaged brick should not be used for garden structures unless they are tested and meet
the grade SW requirements. Most used brick do not meet these requirements (see Technical Notes 15, "Salvaged
Brick").
Mortar. Types M and S. conforming to ASTM Specifications for Mortar for Unit Masonry, C 270, are recommended
for reinforced and non-reinforced brick garden structures.
Workmanship. All head, bed and collar joints should be completely filled with mortar. Less than excellent
workmanship should not be permitted, since a defect may result in deterioration due to the extreme exposure of
garden structures.
DESIGN
The best possible materials and workmanship will not in themselves assure successful and permanent garden
structures. Careful consideration must also be given to construction details. The following Technical Notes outline
details which may be helpful: 29 Rev "Brick in Landscape Architecture, Terraces and Walks", and 29A Rev Brick in
Landscape Architecture, "Garden Walls", and 7A Rev, "Water Resistance of Brick Masonry - Materials, Part II", and
7B Rev., "Water Resistance of Brick Masonry, "Construction and Workmanship", Part III.
Steps. Brick steps, such as those shown in Fig. 1, offer a practical solution to problems which develop in
landscaping a slope. The relatively small sizes of brick permit flexibility of design, such as adjustments of tread and
riser dimensions, and construction of curves. All treads should pitch outward slightly (1/4 in.) for drainage.
http://www.gobrick.com/BIA/technotes/t29b.htm
1 of 10 9/13/2009 1:31 PM
FIG. 1a
FIG. 1b
Brick Screens. Pierced brick screens offer beauty as well as privacy, without loss of light or air (see Fig. 2). A brick
screen can provide a handsome separation between the children's play area and the adults' terrace, and cooling
breezes are not thwarted by any of the numerous patterns available to the designer. Unlike hedges that require
trimming or wood fences with their need for repainting, brick walls are maintenance-free.
http://www.gobrick.com/BIA/technotes/t29b.htm
2 of 10 9/13/2009 1:31 PM
FIG. 2
Outdoor Fireplaces. An outdoor fireplace can be a garden structure of beauty when brick are used to execute an
imaginative but highly practical design, such as shown in Fig. 3. Fireplaces should be planned to face the prevailing
breezes. This orientation not only allows the smoke to blow away, but also provides the best draft. Fire brick should
be selected for the firebox of outdoor fireplaces.
FIG. 3
Planting boxes. Planters may be constructed indoors and outdoors, in a wide variety of designs. They protect
decorative plants from animals and facilitate waist-high gardening as shown in Fig. 4.
http://www.gobrick.com/BIA/technotes/t29b.htm
3 of 10 9/13/2009 1:31 PM
FIG. 4
In constructing brick planting boxes, adequate drainage must be provided. This may be accomplished by weep
holes, as indicated in Fig. 5. When the planting box is designed with a closed bottom, as indicated in Fig. 6, a drain
should be provided. Also, careful attention should be given to the waterproofing of the inside to prevent
efflorescence and staining on the outside face of the brick. In addition, walls must be checked to insure their
resistance to lateral earth pressures.
FIG. 5
http://www.gobrick.com/BIA/technotes/t29b.htm
4 of 10 9/13/2009 1:31 PM
FIG. 6
Edging. Brick edging, as indicated in Fig. 7, may be used to define the lawn area and keep it trim and neat for all
seasons. In addition, the brick strip provides a surface for the lawnmower and thereby eliminates hand trimming.
FIG. 7
Concealment Structures. Figure 8 shows the condenser of a central air-conditioning system that has been
concealed by a low perforated brick screen. The screen must have adequate openings to allow free circulation of air
required for efficient operation. Also, access for servicing should be provided.
http://www.gobrick.com/BIA/technotes/t29b.htm
5 of 10 9/13/2009 1:31 PM
FIG. 8
Trash cans are unsightly, but a low brick enclosure, as indicated in Fig. 9, will banish from sight such undesirable
items in landscape architecture. In addition to being handsome, the structure is maintenance-free.
FIG. 9
Tree Protection. Tree roots require air, water and minerals to survive. When a grade level is changed and the soil
depth over the roots is either increased or decreased, the roots have difficulty obtaining a normal amount of air,
water and minerals. Therefore, to insure the life of a tree, it is necessary to protect it from a grade change.
The properly designed brick retaining wall, reinforced or non-reinforced, is very effective in withstanding lateral
http://www.gobrick.com/BIA/technotes/t29b.htm
6 of 10 9/13/2009 1:31 PM
pressure from earth when changes in grade are necessary.
Visual evidence of respect for nature is shown in the photographs in Fig. 10. Mature trees were preserved, utilizing a
brick retaining wall and a handsome brick tree well.
FIG. 10
Raising the Grade. Minor fills, 6 in. or less in depth, will not harm most species of trees, if the fill is good top soil
that is high in organic matter and loamy in texture. For major grade changes, air and adequate water must be
supplied to the roots of the tree. This may be accomplished by constructing a brick retaining wall around the trunk of
the tree and placing a layer of gravel and a system of drain tile on original grade over the roots of the tree. Figure 11
indicates two plans for placing the drain tile. The tile should slope in the direction indicated. It is important that the tile
extend through the brick retaining wall so that water will not collect around the trunk of the tree. In addition to the tile
placed on original grade, it may be necessary to place a series of bell tile vertically over the roots, and connected to
the tile system for additional air and water circulation. This detail is shown in Fig. 12. The system must be designed
to fit the contour of the land so that water drains away from the tree trunk.
http://www.gobrick.com/BIA/technotes/t29b.htm
7 of 10 9/13/2009 1:31 PM
FIG. 11

http://www.gobrick.com/BIA/technotes/t29b.htm
8 of 10 9/13/2009 1:31 PM
FIG. 12
If economy is essential, or the questionable value of a tree will not permit the expense involved in the construction of
a complete areation system, a variation may be adopted. For example, if a tree was originally on a well drained
slope, sufficient drainage may be obtained through the fill by using coarse gravel around a series of bell tile placed
vertically over the roots, in which case the horizontal tile drains may be omitted.
Lowering the Grade. Protecting a tree from a lowered grade is usually less complicated than protecting it from a
raised grade. Lowering the grade can be equally harmful to a tree unless proper attention is given to cutting the
roots, pruning branches, stimulating root growth and watering. Generally, protection is achieved by terracing the
grade. If space is available, the tree may be unharmed if it remains on a gently sloping mound. Another way to
protect a tree from a lowered grade is to build a brick retaining wall between it and the lower grade, as shown in
Fig. 13.
http://www.gobrick.com/BIA/technotes/t29b.htm
9 of 10 9/13/2009 1:31 PM
FIG. 13

http://www.gobrick.com/BIA/technotes/t29b.htm
10 of 10 9/13/2009 1:31 PM

Technical Notes 30 - Bonds and Patterns in Brickwork
[March 1999]
BOND
The word bond, when used in reference to masonry, may have three meanings:
Structural Bond: The method by which individual masonry units are interlocked or tied together to cause the entire
assembly to act as a single structural unit.
Pattern Bond: The pattern formed by the masonry units and the mortar joints on the face of a wall. The pattern may
result from the type of structural bond used or may be purely a decorative one unrelated to the structural bonding.
Mortar Bond: The adhesion of mortar to the masonry units or to reinforcing steel.
STRUCTURAL BONDS
Structural bonding of masonry walls may be accomplished in three ways: (1) by the overlapping (interlocking) of the
masonry units, (2) by the use of metal ties embedded in connecting joints, and (3) by the adhesion of grout to
adjacent wythes of masonry.
The overlapped bond is based on variations of two traditional methods of bonding. The first is known as English
bond and consists of alternating courses of headers and stretchers (Fig. 1). The second is Flemish bond and
consists of alternating headers and stretchers in every course, so arranged that the headers and stretchers in every
other course appear in vertical lines (Fig. 1).
English and Flemish Bonds
FIG. 1a
http://www.gobrick.com/BIA/technotes/t30.htm
1 of 6 9/13/2009 1:32 PM
English and Flemish Bonds
FIG. 1b
The stretchers, laid with the length of the wall, develop longitudinal bonding strength; while the headers, laid across
the width of the wall, bond the wall transversely.
Modern building codes require that masonry-bonded brick walls be bonded so that not less than 4 per cent of the
wall surface is composed of headers, with the distance between adjacent headers not exceeding 24 inches,
vertically or horizontally.
Structural bonding of masonry walls with metal ties is used in both solid wall and cavity wall construction (Fig. 2).
Metal-Tied Masonry Walls
FIG. 2
http://www.gobrick.com/BIA/technotes/t30.htm
2 of 6 9/13/2009 1:32 PM
Most building codes permit the use of rigid steel bonding ties in solid walls.
At least one metal tie should be used for each 4 1/2 sq ft of wall surface. Ties in alternate courses should be
staggered. The distance between adjacent ties should not exceed 24 in. vertically nor 36 in. horizontally. Additional
bonding ties, spaced not more than 3 ft apart around the perimeter and within 12 in. of the opening, should be
provided at all openings.
If ties less than 3/16 in. in diameter are used, tie spacing should be reduced so that the tie area per square foot of
wall is not less than specified above.
Structural bonding of solid and reinforced brick masonry walls is sometimes accomplished by grout which is poured
into the cavity or collar joint between wythes of masonry.
The method of bonding will depend on the use requirements, wall type and other factors, However, the metal tie
method is generally recommended for exterior walls. Some of the advantages of this method are greater resistance
to rain penetration and ease of construction. Metal ties also allow slight differential movements of the facing and
backing which may relieve stresses and prevent cracking.
PATTERN BONDS
Frequently, structural bonds, such as English or Flemish, or variations of these, may be used to create patterns in
the face of the wall. However, in the strict sense of the term, pattern refers to the change or varied arrangement of
the brick texture or color used in the face. Therefore, it may be possible to secure many patterns using the same
structural bond. Patterns also may be produced by the method of handling the mortar joint or by projecting or
recessing certain brick from the plane of the wall, thus creating a distinctive wall texture that is not solely dependent
upon the texture of the individual brick.
There are five basic structural bonds commonly used today which create typical patterns. These are: Running bond,
common or American bond, Flemish bond, English bond and block or stack bond, as illustrated in Fig. 3. Through the
use of these bonds and variations of the color and texture of the brick, and of the joint types and color, an almost
unlimited number of patterns can be developed.
Traditional Pattern Bonds
FIG. 3
Running bond. The simplest of the basic pattern bonds, the running bond, consists of all stretchers. Since there are
no headers in this bond, metal ties are usually used. Running bond is used largely in cavity wall construction and
veneered walls of brick, and often in facing tile walls where the bonding may be accomplished by extra width
stretcher tile.
Common or American Bond. Common or American bond is a variation of running bond with a course of full length
headers at regular intervals. These headers provide structural bonding, as well as pattern. Header courses usually
appear at every fifth, sixth or seventh course.
In laying out any bond pattern, it is important that the corners be started correctly. For common bond, a three-
quarter brick should start each way from the corner at the header course.
http://www.gobrick.com/BIA/technotes/t30.htm
3 of 6 9/13/2009 1:32 PM
Common bond may be varied by using a Flemish header course.
Flemish Bond. Each course of brick consists of alternate stretchers and headers, with the headers in alternate
courses centered over the stretchers in the intervening courses. Where the headers are not used for the structural
bonding, they may be obtained by using half brick called "clipped" or "snap" headers.
Flemish bond may be varied by increasing the number of stretchers between headers in each course. If there are
three stretchers alternating with a header, it is known as "garden wall" bond. If there are two stretchers between
headers, it is designated as "double stretcher garden wall" bond. Garden wall bond may also be laid with four or
even five stretchers between the headers.
English Bond. English bond is composed of alternate courses of headers and stretchers. The headers are centered
on the stretchers and joints between stretchers in all courses are aligned vertically. Snap headers are used in
courses which are not structural bonding courses.
English Cross or Dutch Bond. English cross or Dutch bond is a variation of English bond which differs only in that
vertical joints between the stretchers in alternate courses do not align vertically. These joints center on the stretchers
themselves in the courses above and below.
There are two methods used in starting the corners in Flemish and English bonds. Figure 3 shows the so-called
"Dutch corner" in which a three-quarter brick closure is used, and the English corner in which a 2-in. or quarter brick
closure, called a "queen closure", is used. The 2-in. closure should always be placed 4 in. in from the corner, never
at the corner.
Block or Stack Bond. Block or stack bond is purely a pattern bond. There is no overlapping of units since all
vertical joints are aligned. Usually this pattern is bonded to the backing with rigid steel ties, but when 8 - in. bonder
units are available, they may be used. In large wall areas and in load bearing construction, it is advisable to reinforce
the wall with steel reinforcement placed in the horizontal mortar joints. In stack bond it is imperative that prematched
or dimensionally accurate masonry units be used if the vertical alignment of the head joints is to be maintained.
Figures 4 and 5 illustrate patterns that may be obtained by varying brick color. Figure 4 is a double stretcher garden
wall bond with the pattern units in diagonal lines. Figure 5 shows a garden wall bond with the pattern units set in
dovetail fashion.
Double Stretcher Garden Wall Bond with Units in Diagonal Lines
FIG. 4
http://www.gobrick.com/BIA/technotes/t30.htm
4 of 6 9/13/2009 1:32 PM
Garden Wall Bond with Units in Dovetail Fashion
FIG. 5
Wall Texture. Recently many contemporary modifications of the traditional bonds have been used by projecting and
recessing units, also by omitting units to form perforated walls or screens. Figure 6 illustrates contemporary uses of
masonry which imaginatively extend the traditional patterns.
Contemporary Bonds
FIG. 6
MORTAR JOINTS
As previously indicated, the treatment of mortar joints in the face of the wall affects the pattern and wall texture.
The mortar serves four functions:
1. It bonds the units together and seals the spaces between.
2. It compensates for dimensional variations in the units.
3. It bonds to and, therefore, causes reinforcing steel to act as an integral part of the wall.
4. It provides a decorative effect on the wall surface by creating shadow or color lines.
http://www.gobrick.com/BIA/technotes/t30.htm
5 of 6 9/13/2009 1:32 PM
Mortar joint finishes fall into two classes: troweled and tooled joints. In the troweled joint, the excess mortar is
simply cut off (struck) with a trowel and finished with the trowel. For the tooled joint, a special tool, other than the
trowel, is used to compress and shape the mortar in the joint.
Figure 7 shows a cross section of typical mortar joints used in good brickwork.
Mortar Joints
FIG. 7
Concave Joint ( 1 ) and V-Shaped Joint ( 2 ) . These joints are normally kept quite small and are formed by the
use of a steel jointing tool. These joints are very effective in resisting rain penetration and are recommended for use
in areas subjected to heavy rains and high winds.
Weathered Joint (3). This joint requires care as it must be worked from below. However, it is the best of the
troweled joints as it is compacted and sheds water readily.
Struck Joint (4). This is a common joint in ordinary brickwork. As American mechanics often work from the inside of
the wall, this is an easy joint to strike with a trowel. Some compaction occurs, but the small ledge does not shed
water readily, resulting in a less watertight joint than joints (1), (2) or (3).
Rough Cut or Flush Joint (5). This is the simplest joint for the mason, since it is made by holding the edge of the
trowel flat against the brick and cutting in any direction. This produces an uncompacted joint with a small hairline
crack where the mortar is pulled away from the brick by the cutting action. This joint is not always watertight.
Raked Joint (6). Made by removing the surface of the mortar, while it is still soft. While the joint may be compacted,
it is difficult to make weather-tight and is not recommended where heavy rain, high wind or freezing is likely to occur.
This joint produces marked shadows and tends to darken the overall appearance of the wall.
Colored mortars may be successfully used to enhance the patterns in masonry. Two methods are commonly used:
(1) the entire mortar joint may be colored or (2) where a tooled joint is used, tuck pointing is the best method. In this
technique, the entire wall is completed with a 1-in. deep raked joint and the colored mortar is carefully filled in later.
http://www.gobrick.com/BIA/technotes/t30.htm
6 of 6 9/13/2009 1:32 PM

Technical Notes 31 - Brick Masonry Arches
January 1995
Abstract: The masonry arch is one of the oldest structural elements. Brick masonry arches have been used for
hundreds of years. This Technical Notes is an introduction to brick masonry arches. Many of the different types of
brick masonry arches are discussed and a glossary of arch terms is provided. Material selection, proper
construction methods, detailing and arch construction recommendations are discussed to ensure proper structural
support, durability and weather resistance of the brick masonry arch.
Key Words: arch, brick, reinforced, unreinforced.
INTRODUCTION
In the latter part of the 19th century, an arch was discovered in the ruins of Babylonia. Archeologists estimate that
the arch was constructed about the year 1400 B.C. Built of well-baked, cigar-shaped brick and laid with clay mortar,
this arch is probably the oldest known to man. The Chinese, Egyptians and others also made use of the arch before
the Christian era. Later, more elaborate arches, vaults and domes with complicated forms and intersections were
constructed by Roman builders during the Middle Ages.
The brick arch is the consummate example of form following function. Its aesthetic appeal lies in the variety of forms
which can be used to express unity, balance, proportion, scale and character. Its structural advantage results from
the fact that under uniform load, the invoiced stresses are principally compressive. Because brick masonry has
greater resistance to compression than tension, the masonry arch is frequently the most efficient structural element
to span openings.
This Technical Notes addresses the detailing and construction of brick masonry arches. The common types of brick
masonry arches are presented, along with proper arch terminology. Methods of selecting the type and configuration
of brick masonry arches most appropriate for the application are discussed. Proper material selection and
construction methods are recommended. Other Technical Notes in this series discuss the structural design of brick
masonry arches and lintels.
ARCH TYPES AND TERMINOLOGY
Many arch forms have been developed during the centuries of use, ranging from the jack arch through the circular,
elliptical and parabolic to the Gothic arch. Figure 1 depicts examples of structural masonry arches used in
contemporary construction. An arch is normally classified by the curve of its intrados and by its function, shape or
architectural style. Figure 2 illustrates some of the many different brick masonry arch types. Jack, segmental,
semicircular and multicentered arches are the most common types used for building arches. For very long spans and
for bridges, semicircular arches are often used because of their structural efficiency.
http://www.gobrick.com/BIA/technotes/t31.htm
1 of 21 9/13/2009 1:35 PM
Structural Brick Arches
FIG. 1
Arch Types: Jack
FIG. 2a
http://www.gobrick.com/BIA/technotes/t31.htm
2 of 21 9/13/2009 1:35 PM
Arch Types: Segmental
FIG. 2b
Arch Types: Semicircular
FIG. 2c
Arch Types: Bullseye
FIG. 2d
http://www.gobrick.com/BIA/technotes/t31.htm
3 of 21 9/13/2009 1:35 PM
Arch Types: Horseshoe
FIG. 2e
Arch Types: Multicentered
FIG. 2f
Arch Types: Venetian
FIG. 2g
http://www.gobrick.com/BIA/technotes/t31.htm
4 of 21 9/13/2009 1:35 PM
Arch Types: Tudor
FIG. 2h
Arch Types: Triangular
FIG. 2i
http://www.gobrick.com/BIA/technotes/t31.htm
5 of 21 9/13/2009 1:35 PM
Arch Types: Gothic
FIG. 2j
Mainly due to their variety of components and elements, arches have developed their own set of terminology.
Following is a glossary of arch terminology. Figure 3 illustrates many of the terms defined in this glossary. Technical
Notes in this series will use this terminology.
Abutment: The masonry or combination of masonry and other structural members which support one end of the
arch at the skewback.
Arch: A form of construction in which masonry units span an opening by transferring vertical loads laterally to
adjacent voussoirs and, thus, to the abutments. Some common arch types are as follows:
Blind -An arch whose opening is filled with masonry.
Bullseye -An arch whose intrados is a full circle. Also known as a Circular arch.
Elliptical -An arch with two centers and continually changing radii.
Fixed -An arch whose skewback is fixed in position and inclination. Masonry arches are fixed arches by
nature of their construction.
Gauged -An arch formed with tapered voussoirs and thin mortar joints.
Gothic -An arch with relatively large rise-to-span ratio, whose sides consist of arcs of circles, the centers of
which are at the level of the spring line. Also referred to as a Drop, Equilateral or Lancet arch, depending
upon whether the spacings of the centers are respectively less than, equal to or more than the clear span.
Horseshoe -An arch whose intrados is greater than a semicircle and less than a full circle. Also known as an
Arabic or Moorish arch.
Jack -A flat arch with zero or little rise.
Multicentered -An arch whose curve consists of several arcs of circles which are normally tangent at their
intersections.
Relieving -An arch built over a lintel, jack arch or smaller arch to divert loads, thus relieving the lower arch or
lintel from excessive loading. Also known as a Discharging or Safety arch.
Segmental -An arch whose intrados is circular but less than a semicircle.
http://www.gobrick.com/BIA/technotes/t31.htm
6 of 21 9/13/2009 1:35 PM
Semicircular -An arch whose intrados is a semicircle (half circle).
Slanted -A flat arch which is constructed with a keystone whose sides are sloped at the same angle as the
skewback and uniform width brick and mortar joints.
Triangular -An arch formed by two straight, inclined sides.
Tudor -A pointed, four-centered arch of medium rise-to-span ratio whose four centers are all beneath the
extrados of the arch.
Venetian -An arch formed by a combination of jack arch at the ends and semicircular arch at the middle. Also
known as a Queen Anne arch.
Camber: The relatively small rise of a jack arch.
Centering: Temporary shoring used to support an arch until the arch becomes self-supporting.
Crown: The apex of the arch's extrados. In symmetrical arches, the crown is at the midspan.
Depth: The dimension of the arch at the skewback which is perpendicular to the arch axis, except that the depth of
a jack arch is taken to be the vertical dimension of the arch at the springing.
Extrados: The curve which bounds the upper edge of the arch.
Intrados: The curve which bounds the lower edge of the arch. The distinction between soffit and intrados is that the
intrados is a line, while the soffit is a surface.
Keystone: The voussoir located at the crown of the arch. Also called the key.
Label Course: A ring of projecting brickwork that forms the extrados of the arch.
Rise: The maximum height of the arch soffit above the level of its spring line.
Skewback: The surface on which the arch joins the supporting abutment.
Skewback Angle: The angle made by the skewback from horizontal.
Soffit: The surface of an arch or vault at the intrados.
Span: The horizontal clear dimension between abutments.
Spandrel: The masonry contained between a horizontal line drawn through the crown and a vertical line drawn
through the upper most point of the skewback.
Springing: The point where the skewback intersects the intrados.
Springer: The first voussoir from a skewback.
Spring Line: A horizontal line which intersects the springing.
Voussoir: One masonry unit of an arch.
http://www.gobrick.com/BIA/technotes/t31.htm
7 of 21 9/13/2009 1:35 PM
Arch Terms
FIG. 3
STRUCTURAL FUNCTION OF ARCHES
The brick masonry arch has been used to span openings of considerable length in many different applications.
Structural efficiency is attributed to the curvature of the arch, which transfers vertical loads laterally along the arch to
the abutments at each end. The transfer of vertical forces gives rise to both horizontal and vertical reactions at the
abutments. The curvature of the arch and the restraint of the arch by the abutments cause a combination of flexural
stress and axial compression. The arch depth, rise and configuration can be manipulated to keep stresses primarily
compressive. Brick masonry is very strong in compression, so brick masonry arches can support considerable load.
Historically, arches have been constructed with unreinforced masonry. Most brick masonry arches continue to be
built with unreinforced masonry. The structural design of unreinforced brick masonry arches is discussed in
Technical Notes 31A. Very long span arches and arches with a small rise may require steel reinforcement to resist
tensile stresses. Also, reduction in abutment size and arch thickness for economy may require incorporation of
reinforcement for adequate load resistance. Refer to the Technical Notes 17 Series for more information on
reinforced brick masonry. Elaborate and intricate arches are sometimes prefabricated to avoid the complexity of
on-site shoring. Most prefabricated brick masonry arches are reinforced. Prefabricated arches are built off site and
transported to the job or built at the site. Cranes are often used to lift the arch into place in the wall. Such
fabrication, handling and transportation should be considered in the structural design of the arch. Refer to Technical
Notes 40 for a discussion of prefabricated brick masonry.
If an unreinforced or reinforced brick masonry arch is not structurally adequate, the arch will require support.
Typically, this support is provided by a steel angle. This is the most common means of supporting brick masonry
arches in modern construction. The steel angle is bent to the curvature of the intrados of the arch. Curved sections
of steel angle are welded to horizontal steel angles to form a continuous support. The angle either bears on the
brickwork abutments or is attached to a structural member behind the wall. One example is shown in Fig. 4. When
an arch is supported by a steel angle, the angle is designed to support the entire weight of brick masonry loading the
arch, and the structural resistance of the arch is neglected. Consult Technical Notes 31B Revised for a discussion of
the structural design of steel angle lintels.
http://www.gobrick.com/BIA/technotes/t31.htm
8 of 21 9/13/2009 1:35 PM
Arch Supported by Curved Steel Angle
FIG. 4
WEATHER RESISTANCE
Water penetration resistance is a primary concern in most applications of the building arch. In the past, the mass of
a multi-wythe brick masonry arch was sufficient to resist water penetration. Today, thinner wall sections are used to
minimize material use for economy and efficiency. Still, the arch must provide an effective weather resistant facade.
Some arch applications do not require provisions for water penetration and insulation. For example, arch arcades
and arches supported by porch columns typically do not conceal a direct path for water migration to the interior of
the building they serve and may not require insulation. If this is the case, provisions for weather resistance need not
be included in the arch design and detailing.
Preventing water entry at an arch in an exterior building wall is just as important as at any other wall opening. Water
penetration resistance can be provided by using a barrier wall system or a drainage wall system. Refer to Technical
Notes 7 Revised for definitions and discussion of barrier and drainage wall systems. A drainage wall system, such
as a brick veneer or cavity wall, is the most common brick masonry wall system used today. For either wall system,
the arch should be flashed, with weep holes provided above all flashing locations.
Flashing and Weep Holes
Installation of flashing and weep holes around an arch can be difficult. Installation of flashing is easiest with jack
arches because they are flat or nearly flat. Flashing should be installed below the arch and above the window
framing or steel angle lintel. Flashing should extend a minimum of 4 in. (100 mm) past the wall opening at either end
and should be turned up to form end dams. This is often termed tray flashing. Weep holes should be provided at
both ends of the flashing and should be placed at a maximum spacing of 24 in. (600 mm) on centers along the arch
span, or 16 in. (400 mm) if rope wicks are used. An example of flashing a jack arch in this manner is shown in Fig.
http://www.gobrick.com/BIA/technotes/t31.htm
9 of 21 9/13/2009 1:35 PM
5a. Attachment of the flashing to the backing and formation of end dams should follow standard procedures. If the
arch is constructed with reinforced brick masonry, flashing and weep holes can be placed in the first masonry course
above the arch.
Flashing Arches
FIG. 5a
http://www.gobrick.com/BIA/technotes/t31.htm
10 of 21 9/13/2009 1:35 PM
Flashing Arches
FIG. 5b
http://www.gobrick.com/BIA/technotes/t31.htm
11 of 21 9/13/2009 1:35 PM
Flashing Arches
FIG. 5c
Installation of flashing with other arch types, such as segmental and semicircular arches, can be more difficult. This
is because most rigid flashing materials are hard to bend around an arch with tight curvature. If the arch span is less
than about 3 ft (0.9 m), one section of tray flashing can be placed in the first horizontal mortar joint above the
keystone, as illustrated in Fig. 5b. For arch spans greater than 3 ft (0.9 m), flashing can be bent along the curve of
the arch with overlapping sections, as illustrated in Fig. 4. Alternately, a combination of stepped and tray flashing can
be used, as shown in Fig. 5c. To form a step, the end nearest the arch should be turned up to form an end dam,
while the opposite end is laid flat. A minimum of No. 15 building paper or equivalent moisture resistant protection
should be installed on the exterior face of the backing over the full height of the arch and abutments. The building
paper or equivalent should overlap the arch flashing.
The design of a structural masonry arch should include consideration of the effect of flashing on the strength of the
arch. Flashing acts as a bond break. If flashing is installed above the arch, the loading on the arch will likely be
increased, and the structural resistance of the arch will be reduced. Installation of flashing at the abutments will
affect their structural resistance and should also be considered. Consult Technical Notes 31A for a more extensive
discussion of arch loads and structural resistance of brick masonry arches.
DETAILING CONSIDERATIONS
The brick masonry arch should serve its structural purpose and also provide an attractive architectural element to
complement its surrounding structure. Careful consideration should be given to the options available for the arch,
soffit and skewback. Proper configuration of the abutments and location of expansion joints should be considered for
any arch design.
Arch
Arches can be configured in a variety of arch depths, brick sizes and shapes and bonding patterns. The arch is
http://www.gobrick.com/BIA/technotes/t31.htm
12 of 21 9/13/2009 1:35 PM
normally composed of an odd number of units for aesthetic purposes. Some of the more common arch
configurations are illustrated in Fig. 6. Arch voussoirs are typically laid in radial orientation and are most often of
similar size and color to the surrounding brickwork. However, the arch can be formed with brick which are thinner or
wider than the surrounding brickwork and of a different color for variation. Another variation is to project or recess
rings of multiple-ring arches to provide shadow lines or a label course.
Typical Arch Configurations
FIG. 6
Brick masonry arches are constructed with two different types of units. The first is tapered or wedge-shaped brick.
These brick are tapered in the appropriate manner to obtain mortar joints of uniform thickness along the arch depth.
The second is uncut, rectangular brick. When rectangular brick are used, the mortar joints are tapered to obtain the
desired arch curvature. In some cases, a combination of these is used. For example, a slanted arch is formed with a
tapered keystone and rectangular brick. This arch is similar to a jack arch, but can be more economical because it
requires only one special-shaped brick.
Selection of tapered or rectangular brick can be determined by the arch type, arch dimensions and by the
appearance desired. Some arch types require more unique shapes and sizes of brick if uniform mortar joint
thickness is desired. For example, the brick in a traditional jack arch or elliptical arch are all different sizes and
shapes from the abutment to the keystone. Conversely, the voussoirs of a semicircular arch are all the same size
and shape. Arch types with many different brick shapes and sizes should be special ordered from the brick
manufacturer rather than cut in the field.
The arch span should also be considered when selecting the arch brick. For short arch spans, use of tapered brick
is recommended to avoid excessively wide mortar joints at the extrados. Larger span arches require less taper of
the voussoirs and, consequently, can be formed with rectangular brick and tapered mortar joints. The thickness of
mortar joints between arch brick should be a maximum of 3/4 in. (19 mm) and a minimum of 1/8 in. (3 mm). When
using mortar joints thinner than 1/4 in. (6 mm), consideration should be given to the use of very uniform brick that
meet the dimensional tolerance limits of ASTM C 216, Type FBX, or the use of gauged brickwork. Refer to Table 1
for determination of the minimum segmental and semicircular arch radii permitted for rectangular brick and tapered
mortar joints. Typically, the use of tapered brick and uniform thickness mortar joints will be more aesthetically
appealing.
http://www.gobrick.com/BIA/technotes/t31.htm
13 of 21 9/13/2009 1:35 PM
1
Based on 1/4 in. (6 mm) mortar joint width at the intrados and 1/2 in. (13 mm) mortar joint width at the
extrados. If the mortar joint thickness at the extrados is 3/4 in. (19 mm), divide minimum radius value by 2.
2
1 in.=25.4 mm; 1 ft=0.3m
Depth. The arch depth will depend upon the size and orientation of the brick used to form the arch. Typically, the
arch depth is a multiple of the brick's width. For structural arches, a minimum arch depth is determined from the
structural requirements. If the arch is supported by a lintel, any arch depth may be used.
The depth of the arch should also be detailed based on the scale of the arch in relation to the scale of the building
and surrounding brickwork. To provide proper visual balance and scale, the arch depth should increase with
increasing arch span. Because aesthetics of an arch are subjective, there are no hard rules for this. However, the
following rules-of-thumb will help provide an arch with proper scale. For segmental and semicircular arches, the arch
depth should equal or exceed 1 in. (25 mm) for every foot (300 mm) of arch span or 4 in. (100 mm), whichever is
greater. For jack arches, the arch depth should equal or exceed 4 in. (100 mm) plus 1 in. (25 mm) for every foot
(300 mm) of arch span or 8 in. (200 mm), whichever is greater. For example, the minimum arch depth for an 8 ft
(2.4 m) span should be 8 in. (200 mm) for segmental arches and 12 in. (300 mm) for jack arches.
The depth of jack arches will also be a function of the coursing of the surrounding brick masonry. The springing and
the extrados of the jack arch should coincide with horizontal mortar joints in the surrounding brick masonry. Typically,
the depth of a jack arch will equal the height of 3, 4 or 5 courses of the surrounding brickwork, depending upon the
course height.
Keystone. The keystone may be a single brick, multiple brick, stone, precast concrete or terra cotta. Avoid using a
keystone which is much taller than the adjacent voussoirs. A rule-of-thumb is that the keystone should not extend
above adjacent arch brick by more than one third the arch depth. When a keystone is used that is larger than
adjacent arch brick or formed with different material, one option is to use springers that match the keystone.
The use of a large keystone has its basis in both purpose and visual effect. With most arch types, the likely location
of the first crack when the arch fails is at the mortar joint nearest to the midspan of the arch. Use of a large
keystone at this point moves the first mortar joint further from the midspan and increases the resistance to cracking
at this point. Aesthetically, a large keystone adds variation of scale and can introduce other masonry materials in the
facade for additional color and texture.
If the keystone is formed with more than one masonry unit, avoid placing the smaller unit at the bottom. Such units
are more likely to slip when the arch settles under load. Also, it is preferred to have the arch crown (the top of the
keystone) coincident with a horizontal mortar joint in the surrounding brickwork to give the arch a neater appearance.
Soffit
http://www.gobrick.com/BIA/technotes/t31.htm
14 of 21 9/13/2009 1:35 PM
A brick masonry soffit is one attractive feature of a structural brick masonry arch. Many bonding patterns and
arrangements can be used to form the arch soffit. Deep soffits are common on building arcades or arched
entranceways. In this case, it is common to form a U-shaped wall section, as illustrated in Fig. 7. The arches on
either wall face should be bonded to the brick masonry forming the soffit. Bonding pattern or metal ties should be
used to tie the brick masonry forming the soffit together structurally and to tie the arches on either wall face to the
soffit. If metal ties are used to bond the masonry, corrosion resistant box or Z metal wire ties should be placed
along the arch span at a maximum spacing of 24 in. (600 mm) on center.
Structural resistance of the arch should be evaluated at sections through the soffit, the exterior wall face and the
interior wall face. Deeper soffits may require an increase in arch depth. If the arch is structural, connection of the
brick masonry forming the soffit to interior framing members with wall ties or connectors may not be required.
Structural Arch Soffit Option
FIG. 7
Skewback
For flat arches and arch types that have horizontal skewbacks, such as jack and semicircular arches, respectively,
the most desirable spring line location is coincident with a bed joint in the abutment. For other arch types, it is
preferred to have the spring line pass about midway through a brick course in the abutment, as illustrated in Fig. 8,
to avoid a thick mortar joint at the springing. The brick in the abutment at the springing should be cut or be a special
cant-shaped brick. This allows vertical alignment with the brick beneath, producing more accurate alignment of the
arch.
When two arches are adjacent, such as with a two-bay garage or building arcades, intersection of the arches may
occur at the skewback. Attention should be given to proper bonding of the arches for both visual appeal and
structural bonding. Creation of a vertical line between arches should be avoided. Rather, special shape brick should
http://www.gobrick.com/BIA/technotes/t31.htm
15 of 21 9/13/2009 1:35 PM
be used to mesh the two arches properly. One example is illustrated in Fig. 9.
Skewback Options
FIG. 8
Option For Intersecting Arches
FIG. 9

Abutments
An arch abutment can be a column, wall or combination of wall and shelf angle. Failure of an abutment occurs from
excessive lateral movement of the abutment or exceeding the flexural, compressive or shear strength of the
abutment. Lateral movement of the abutment is due to the horizontal thrust of the arch. Thrust develops in all arches
and the thrust force is greater for flatter arches. The thrust should be resisted so that lateral movement of the
abutment does not cause failure in the arch. If the abutment is formed by a combination of brickwork and a
non-masonry structural member, rigidity of the non-masonry structural member and rigidity of the ties are very
important. Adjustable ties or single or double wire ties are recommended. Corrugated ties should not be used in this
application because they do not provide adequate axial stiffness. Consult Technical Notes 31A for further discussion
of abutment and tie stiffness requirements.
Lateral Bracing
In addition to gravity loads, out-of-plane loads should be considered when designing a masonry arch. The arch
should have adequate resistance to out-of-plane loads or lateral bracing should be provided. In veneer construction,
http://www.gobrick.com/BIA/technotes/t31.htm
16 of 21 9/13/2009 1:35 PM
lateral bracing is provided by the backing through the use of wall ties. Arches which are not laterally braced may
require increased masonry thickness or reinforcement to carry loads perpendicular to the arch plane in addition to
vertical loads.
Expansion Joints
Thermal and moisture movements of brick masonry are controlled by the use of expansion joints. Expansion joints
avoid cracking of the brickwork and also reduce the size of wall sections. Reduction of wall size has a very important
effect upon the performance of structural brick masonry arches. The state of stress in a structural brick arch and the
surrounding masonry is very sensitive to the relative movements of the abutments. If an inadequate number of
expansion joints are provided, the differential movement of abutments can cause cracking and downward
displacement of brick in the masonry arch and surrounding masonry. Proper size and spacing of expansion joints is
discussed in Technical Notes 18A Revised.
If the arch is structural, care should be taken not to affect the integrity of the arch by detailing expansion joints too
close to the arch and its abutments. Vertical expansion joints should not be placed in the masonry directly above a
structural arch. This region of masonry is in compression, so an expansion joint will cause displacement when
centering is removed and possible collapse of the arch and surrounding brickwork. In addition, vertical expansion
joints should not be placed in close proximity to the springing. The expansion joint will reduce the effective width of
the abutment and its ability to resist horizontal thrust from the arch. If the arch is non-structural, placement of
expansion joints may be at the arch crown and also at a sufficient distance away from the springing to avoid sliding.
While permitted, placement of an expansion joint at the arch crown is not preferred because it disrupts ones
traditional view of the arch as a structural element. Refer to Fig. 10 for suggested expansion joint locations for
structural and non-structural arches.
Expansion Joints Near Arches
FIG. 10a
http://www.gobrick.com/BIA/technotes/t31.htm
17 of 21 9/13/2009 1:35 PM
Expansion Joints Near Arches
FIG. 10b
Detailing of expansion joints can be difficult with very long span arches or runs of multiple arches along an arcade.
Structural analysis of the arch should consider the location of expansion joints. For the particular case of multiple
arches closely spaced, vertical expansion joints should be detailed at a sufficient distance away from the end arches
so that horizontal arch thrusts are adequately resisted by the abutments to avoid overturning of the abutments. For
long arcades, expansion joints should also be placed along the centerline of abutments between arches when
necessary. In this case, horizontal thrusts from adjacent arches will not be counteracting, so the effective abutment
length should be halved and overturning of each half of the abutment should be checked. Refer to Technical Notes
31A for further discussion of abutment design for adequate stiffness.
MATERIAL SELECTION
To provide a weather resistant barrier and maintain its structural resistance, the arch must be constructed with
durable materials. The strength of an arch depends upon the compressive strength and the flexural tensile strength
of the masonry. Selection of brick and mortar should consider these properties.
Brick
Solid or hollow clay brick may be used to form the arch and the surrounding brickwork. Solid brick should comply
with the requirements of ASTM C 216 Specification for Facing Brick. Hollow brick should comply with the
requirements of ASTM C 652 Specification for Hollow Brick. Refer to Technical Notes 9 Series for a discussion of
brick selection and classification. The compressive strength of masonry is related to the compressive strength of the
brick, the mortar type and the grout strength. For structural arches, brick should be selected with consideration of
the required compressive strength of masonry. Typically, compressive strength of the brick masonry will not limit the
design of the arch.
http://www.gobrick.com/BIA/technotes/t31.htm
18 of 21 9/13/2009 1:35 PM
Tapered voussoirs can be cut from rectangular units at the job site or special ordered from the brick manufacturer.
Before specifying manufactured special arch shapes, the designer should determine the availability of special shapes
for the arch type and brick color and texture desired. Many brick manufacturers produce tapered arch brick for the
more common arch types as part of their regular stock of special shapes. Be sure to contact the manufacturer as
early as possible if special shapes are needed. In many instances, production of the special shapes may require a
color matching process and adequate lead time for the manufacturer.
Mortar
Mortar used to construct brick masonry arches should meet the requirements of ASTM C 270 Standard
Specification for Masonry Mortar. Consult Technical Notes 8 Series for a discussion of mortar types and kinds for
brick masonry. For structural arches, the flexural tensile strength of the masonry should be considered when
selecting the mortar. The flexural tensile strength of the masonry will affect the load resistance of the arch and the
abutments.
CONSTRUCTION AND WORKMANSHIP
The proper performance of a brick masonry arch depends upon proper methods of construction and attention to
workmanship. Layout of the arch prior to construction will help avoid poor spacing of voussoirs, which results in
thicker mortar joints and unsymmetrical arches. Some arch applications, such as barrel vaults and domes, can be
entirely self-supporting, even during construction. However, most applications of the masonry arch used today
require proper shoring and bracing.
Centering
Both structural and non-structural arches should be properly supported throughout construction. Brick masonry
arches are constructed with the aid of temporary shoring, termed centering, or permanent supports, such as a
structural steel angle.
Centering is used to carry the weight of a brick masonry arch and the loads being supported by the arch until the
arch itself has gained sufficient strength. The term "centering" is used because the shoring is marked for proper
positioning of the brick forming the arch. Centering is typically provided by wood construction. An example of
centering for an arch is shown in Fig. 11. Careful construction of the centering will ensure a more pleasing arch
appearance and avoid layout problems, such as an uneven number of brick to either side of the keystone.
http://www.gobrick.com/BIA/technotes/t31.htm
19 of 21 9/13/2009 1:35 PM
Centering
FIG. 11
Immediately after placement of the keystone, very slight downward displacement of the centering, termed easing,
can be performed to cause the arch voussoirs to press against one another and compress the mortar joints between
them. Easing helps to avoid separation cracks in the arch. In no case should centering be removed until it is certain
that the masonry is capable of carrying all imposed loads. Premature removal of the centering may result in collapse
of the arch.
Centering should remain in place for at least seven days after construction of the arch. Longer curing periods may
be required when the arch is constructed in cold weather conditions and when required for structural reasons. The
arch loading and the structural resistance of the arch will depend upon the amount of brickwork surrounding the arch,
particularly the brick masonry within spandrel areas. Appropriate time of removal of centering for a structural arch
should be determined with consideration of the assumptions made in the structural analysis of the arch. It may be
necessary to wait until the brickwork above the arch has also cured before removing the centering.
Workmanship
All mortar joints should be completely filled, especially in a structural member such as an arch. If hollow brick are
used to form the arch, it is very important that all face shells and end webs are completely filled with mortar. Brick
http://www.gobrick.com/BIA/technotes/t31.htm
20 of 21 9/13/2009 1:35 PM
masonry arches are sometimes constructed with the units laid in a soldier orientation. It may be difficult to lay units in
a soldier position and also obtain completely filled mortar joints. This is especially true for an arch with tapered
mortar joints. In such cases, the use of two or more rings of arch brick laid in rowlock orientation can help ensure full
mortar joints.
SUMMARY
This Technical Notes is an introduction to brick masonry arches. A glossary of arch terms has been provided. Many
different types of brick masonry arches are described and illustrated. Proper detailing of brick masonry arches for
appearance, structural support and weather resistance is discussed. Material selection and proper construction
practices are explained. Other Technical Notes in this Series discuss the structural design of arches.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Institute of America. The information contained herein must be used
in conjunction with good technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the purview of the Brick
Institute of America and must rest with the project architect, engineer and owner.
REFERENCES
1. Brickwork Arch Detailing, Ibstock Building Products, Butterworth & Co. (Publishers) Ltd., London, England,
1989, 114 pp.
2. Lynch, G., Gauged Brickwork, A Technical Handbook, Gower Publishing Company, Aldershot, Hants,
England, 1990, 115 pp.
3. Trimble, B.E., and Borchelt, J.G., "Jack Arches in Masonry Construction," The Construction Specifier,
Construction Specifications Institute, Alexandria, VA, January 1991, pp. 62-65.
http://www.gobrick.com/BIA/technotes/t31.htm
21 of 21 9/13/2009 1:35 PM

Technical Notes 31A - Structural Design of Brick Masonry Arches
[Oct. 1967] (Reissued July 1986)
INTRODUCTION
The railway bridge at Maidenhead, England, constructed in 1838, is a brick arch with a span of 128 ft and a rise of
24.3 ft. This arch was designed by engineer Marc Brunel, who is also credited with being the first to use reinforced
brick masonry. A similar brick arch railway bridge was constructed on North Avenue, Baltimore, Maryland, in 1895. It
has a span of 130 ft and a rise of 26 ft. These outstanding examples are cited only to illustrate the structural
capabilities of the brick arch - capabilities on which designers may rely when architectural or structural
considerations suggest their use in modern design.
This issue of Technical Notes covers the structural design of major and minor brick masonry arches.
Minor arches are those whose spans do not exceed 6 ft and with maximum rise-to-span ratios of 0.15. Coefficients
are given from which the horizontal thrust of such arches may be determined. Equations are presented for obtaining
compressive stresses developed in the masonry and for determining stability against sliding.
Derivation of thrust coefficients and equations are based on the hypothesis of least crown thrust, as described in
Technical Notes 31, and the following assumptions have been made:
1. The thrust at the crown is horizontal and passes through the upper 1/3 point of the arch.
2. The reaction passes through the lower 1/3 point of the arch at the skewback.
3. The equilibrium polygon lies completely within the middle 1/3 of the arch.
Figure 1 illustrates jack and segmental arches.


FIG. 1a
31a http://www.gobrick.com/BIA/technotes/t31a.htm
1 of 19 9/13/2009 1:35 PM




FIG. 1b



Major arches are those with spans in excess of 6 ft or rise-to-span ratios greater than 0.15. In this issue of
Technical Notes an example is given of major arch design based on the equations for redundant moments and
forces presented in the publication, "Frames and Arches". ("Frames and Arches," by Valerian Leontovich, McGraw
Hill Book Co., 1959.) The method of analysis presented in this book is substantially shorter than others in current
use.

MINOR ARCH LOADING

The loads falling upon a minor arch may consist of live loads and dead loads from floors, roofs, walls and other
structural members. These are applied as point loads or as uniform loads fully or partially distributed. A method of
determining imposed loads on a member spanning small openings is described in Technical Notes 17H. A brief
resume of that explanation is given here.
The dead load of a wall above an arch may be assumed to be the weight of wall contained within a triangle
immediately above the opening. The sides of this triangle are at 45-deg angles to the base. Therefore, its height is
1/2 of the span. Such triangular loading may be assumed to be equivalent to a uniformly distributed load of 1/3 times
the triangular load.
Superimposed uniform loads above this triangle may be carried by arching action of the masonry wall itself. Uniform
live and dead loads occurring below the apex of the triangle are applied directly upon the arch for design purposes.
Heavy concentrated loads should not be allowed to bear directly on minor arches. This is especially true of jack
31a http://www.gobrick.com/BIA/technotes/t31a.htm
2 of 19 9/13/2009 1:35 PM
arches. Minor concentrated loads bearing on, or nearly directly on, the arch may safely be assumed to be equivalent
to a uniformly distributed load equal to twice the concentrated load.
Figure 2 shows the use of minor arches in contemporary architecture.


High School in Columbus, Indiana
Harry Weese & Associates, Architects
FIG. 2

MINOR ARCH DESIGN

There are three methods of failure of unreinforced masonry arches: (a) by rotation of one section of the arch about
the edge of a joint; (b) by the sliding of one section of the arch on another or on the skewback; (c) by crushing of the
masonry
(a) Rotation. The assumption for the design of minor arches, that the equilibrium polygon lies entirely within the
middle third of the arch section, precludes the rotation of one section of the arch about the edge of a joint or the
development of tensile stresses in either the intrados or extrados.
(b) Sliding. The coefficient of friction between the units composing a brick or tile masonry arch is at least 0.60,
without considering the additional resistance to sliding resulting from bond between mortar and the masonry units.
This corresponds to an angle of friction of approximately 31 deg. If that angle, which the line of resistance of the
arch makes with the normal to the joint between arch sections, is less than the angle of friction, the arch is stable
against sliding. This angle can be determined graphically, as illustrated in Technical Notes 31, or may be determined
mathematically bar the following formula:

where:b = angle between line of resistance and normal to joint,
31a http://www.gobrick.com/BIA/technotes/t31a.htm
3 of 19 9/13/2009 1:35 PM

W = total equivalent uniform load on arch,
H = crown thrust and
g = angle of joint with vertical.

For minor segmental arches, the angle between the line of resistance and the normal to the joint is greatest at the
skewback. This will also be true for jack arches if the joints are radial about a center at the intersection of the planes
of the skewbacks. However, if the joints are not radial about this center, each joint should be investigated for
resistance to sliding. This can be done most easily by constructing an equilibrium polygon, assuming that the crown
thrust is applied at the top of the middle third and the reaction at the skewback is applied at the bottom of the middle
third of the section.
For segmental arches with radial joints, the angle (g) between the skewback and the vertical is


or in terms of the radius of curvature

For jack arches in which the skewback equals 1/2 in. per ft of span for each 4 in. of arch depth, the angle (g) that
the skewback makes with the vertical is


In equations 2, 3 and 4:
S = span,
r = rise,
R = radius of curvature.

(c) Crushing.
(1) Segmental Arch. Figure 3 is a graphic representation of thrust coefficients (H/W) for segmental arches subjected
31a http://www.gobrick.com/BIA/technotes/t31a.htm
4 of 19 9/13/2009 1:35 PM
to uniform load over the entire span. Once the thrust coefficient is determined for a particular arch, the horizontal
thrust (H) may be determined as the product of the thrust coefficient and the total load (W). To determine the proper
thrust coefficient, one must first determine the characteristics of the arch, S/r and S/d:
where:
S = the clear span,
r = the rise of the soffit and
d = the depth of the arch.
In these ratios and in the ratios and equations that follow, all terms of length must be expressed in the same units;
for example, in computing S/r and S/d, if S is in feet, r and d must be in feet also.


Thrust Coefficients for Segmental Arches
FIG. 3



If the applied load is triangular or concentrated, the above method may still be used, but the horizontal thrust
coefficient must be increased by 1/3 for triangular loading and doubled for concentrated loads.
31a http://www.gobrick.com/BIA/technotes/t31a.htm
5 of 19 9/13/2009 1:35 PM
Once the horizontal thrust has been determined, the maximum compressive stress in the masonry is determined by
the following formula:


In this equation:
fm = maximum compressive stress in the arch in pounds per square inch,

H = horizontal thrust in pounds,
b = breadth of the arch in inches and
d = depth of the arch in inches.

This value is twice an axial compressive stress on the arch, due to a load H. because the horizontal thrust is located
at the third point of the arch depth.
(2) Jack Arch. The common rule for jack arches is to provide a skewback (K, measured horizontally) of 1/2 in. per ft
of span for each 4 in. of arch depth. Jack arches are commonly constructed in depths of 8 and 12 in. with a camber
of 1/8 in. per ft of span.
For jack arches, applying the same assumptions as previously outlined, the horizontal thrust at the spring line may be
determined by the following formulae:
For uniform loading over full span,


For triangular loading over full span,

Maximum compressive stress (fm) in the jack arch may be determined from the following formulae:

31a http://www.gobrick.com/BIA/technotes/t31a.htm
6 of 19 9/13/2009 1:35 PM

The maximum compressive stress in a jack arch may be computed directly from the following formulae:
For uniform loading over full span,


For triangular loading,


Formulae 8, 9 and 10 include a factor which allows for non-axial loading. In formulae 6 through 10, inclusive:

H = horizontal thrust in pounds,
W = total load in pounds,
S = clear span in inches,
d = depth of arch in inches and
b = breadth of arch in inches.

THRUST RESISTANCE

Resistance to horizontal thrust, developed by the arch, is provided by the adjacent mass of masonry. In areas where
limited masonry is available, i.e. corners, openings, etc., it may be necessary to check the resistance of the wall to
the horizontal thrusts. Figure 4 illustrates how such resistance may be calculated.

31a http://www.gobrick.com/BIA/technotes/t31a.htm
7 of 19 9/13/2009 1:35 PM

FIG. 4



It is assumed that the thrust of the arch attempts to move a volume of masonry enclosed by the boundary lines
ABCD. For calculating purposes the area CDEF is equivalent in resistance. It can be seen that the thrust is acting
against two planes of resistance, CF and DE. The resistance to arch thrust is determined by the following formula:


By using the principle given in formula (11), the minimum distance from a corner or opening at which an arch may be
located is easily determined. This can be done by writing formula (11) to solve for x, substituting actual arch thrust
for resisting thrust:


In these formulae:
H1 = resisting thrust in pounds,
vm = allowable shearing stress in the masonry wall in pounds per square inch,

n = the number of resisting shear planes,
31a http://www.gobrick.com/BIA/technotes/t31a.htm
8 of 19 9/13/2009 1:35 PM
x = the distance from the center of the skew-back to the end of the wall in inches
and
t = wall thickness in inches.

The tendency for arch thrust to overturn a section of masonry, rather than slide it or rack it, must also be
investigated. In general, such overturning forces are found to govern only at arches near the top of a wall, since that
portion of masonry which tends to overturn must first become separated from the body of the wall.

ALLOWABLE STRESSES

Recommended allowable compressive stresses for use in the design of brick arches are given in Table 1.
Recommended allowable shearing stresses in unreinforced walls for use in the design of abutments are given in
Table 2. These are based on the requirements of Recommended Building Code Requirements for Engineered
Brick Masonry, SCPI, May 1966


1Based on Recommended Building Code Requirements for Engineered Brick Masonry,
SCPI, MA 1966
2Linear interpolation is permissible


31a http://www.gobrick.com/BIA/technotes/t31a.htm
9 of 19 9/13/2009 1:35 PM
MAJOR ARCH LOADING

The principal forces acting upon arches in buildings are the result of vertical dead and live loads and wind loads.
Many masonry arches are integral with surrounding masonry. In such instances, loads transmitted to the arch
through the masonry are indeterminate, due to arching action of adjacent masonry.
It often assumed that the entire weight of masonry, above the soffit, presses vertically upon the arch. This certainly
is not accurate, since even with dry masonry a part of the wall will be self-supporting. However, this assumption is
certainly on the safe side. The passive resistance of the adjacent masonry materially affects the stability of an arch.
The designer must rely on empirical formulae, based on the performance of existing structures, to determine the
loads on an arch. The dead load of masonry wall supported by an integral arch depends upon the arch rise and span
and the wall height above the arch. It may be considered to be either uniform (rectangular) or variable
(complementary parabolic) in distribution, or a combination thereof.
"Frames and Arches" gives solutions for arches with rise-to-span ratios (f/L) ranging from 0.0 to 0.6. The following
recommended assumptions for loading of such arches are believed to be safe:
For low rise arches, f/L = 0.2 or less, a uniform load may be assumed. This load will be the weight of wall above the
crown of the arch up to a maximum height of L/4.
For higher rise arches a dead load consisting of uniform plus complementary parabolic loading may be assumed.
The maximum ordinate of the parabolic loading will be equal to a weight of wall whose height is the rise of the arch.
The minimum ordinate of the parabolic loading will be zero. The uniform loading will be the weight of the wall above
the crown of the arch up to a maximum height of L
2
/100.

Uniform floor and roof loads are applied as a uniform load on the arch. Small concentrated loads may be treated as
uniform loads of twice the magnitude. Large concentrated loads may be treated as point loads on the arch.
Several major arches are shown in Fig. 5.

31a http://www.gobrick.com/BIA/technotes/t31a.htm
10 of 19 9/13/2009 1:35 PM

Grundtvig Church, Copenhagen, Denmark
Designed in 1913 by P.V. Jensen Klint - It was completed in 1940
FIG. 5

MAJOR ARCH DESIGN
General. "Frames and Arches" provides straightforward equations by which redundant moments and forces in
arched members may be determined. The reader is referred to a discussion of this book which appears in Technical
Notes 31.

Without repeating the aforementioned discussion here, let it suffice to say that, for relatively high-rise (f/L = 0.2)
constant-section arches, Method A of Section 22 yields the proper solutions. The recommendations for use of this
31a http://www.gobrick.com/BIA/technotes/t31a.htm
11 of 19 9/13/2009 1:35 PM
section are:
1. Establish principal dimensions of the arch.
2. On this basis and depending upon the established shape and f/L ratio of the arch, obtain the corresponding k
value of the arch (see Table 3).
3. Obtain the elastic parameters (a, b, g and d), load constants and general constants.

4. Perform the algebraic operations with the given equations.
Equations. The equations are based upon a horizontal and vertical grid coordinate system with origin at the
intersection of the arch axis and left skewback. Distances x and y are coordinates of the arch axis. The general
equation for the parabolic arch axis is:


Each set of equations depends upon the loading conditions. Among the solutions included with those in Section 22,
Method A are the following:

For vertical complementary parabolic loading:

When x L / 2:

31a http://www.gobrick.com/BIA/technotes/t31a.htm
12 of 19 9/13/2009 1:35 PM

For vertical uniform load over the entire arch:

M and Q are zero at any section of the arch.

When x L / 2:


"Frames and Arches" also contains equations for other loading conditions; e.g. concentrated loads.


Note: Adapted from Table 12, "Frames and Arches."

Notation. In these equations, the subscripts 1 and 2 denote the left and right supports respectively. The subscript x
denotes values at any horizontal distance, x, from the origin. f is the angle, at any point, whose tangent is the slope
of the arch axis at that point. (See Table 4.)

M = moment
N = axial force
31a http://www.gobrick.com/BIA/technotes/t31a.htm
13 of 19 9/13/2009 1:35 PM
Q = shearing force
f = rise of the arch
W = total load under consideration
H = horizontal thrust
V = vertical reaction
L = span of the arch

S and T are load constants (see Table 5).
J, F and K are constants, determined by:
J=1 + (d / g)
F=q - gJ2
K=S q / g)
a, b, g and d are parameters (see Table 6).

q=2 (a + b)


Note: From Table 10, "Frames and Arches."

31a http://www.gobrick.com/BIA/technotes/t31a.htm
14 of 19 9/13/2009 1:35 PM
Note: From Table 18, "Frames and Arches." Intermediate values may be obtained by interpolation.
.
Note: From Table 13, "Frames and Arches." Intermediate values may be obtained by interpolation.

ILLUSTRATIVE EXAMPLE
Problem. Using the equations given in the book, "Frames and Arches," design a parabolic brick masonry arch to
meet the following requirements. The arch is integral with a loadbearing, brick-and-brick cavity wall. Wall weight is
80 psf. The arch dimensions are: span, 20 ft; rise, 12 ft; depth, 16 in.; thickness, 12 in.; total wall height, 8 ft. The
uniform roof load, bearing on the wall above the arch, is 1200 lb per ft. The arch is of solid brick (4000 psi) and type
N mortar; allowable compressive stress in the arch is 300 psi.

Solution. The arch is a constant-section, high-rise, symmetrical, parabolic, hingeless arch; therefore, the equations
previously given in this issue of Technical Notes are applicable. Each different loading condition must be analyzed
separately. Similar loads, e.g. all uniform loads, may be added and treated as a single load. Moments, shears and
thrusts resulting from each loading condition are combined to give total values. For symmetrically loaded symmetrical
arches, only 1/2 of the arch need be analyzed.
The loads carried by the arch are:
31a http://www.gobrick.com/BIA/technotes/t31a.htm
15 of 19 9/13/2009 1:35 PM
Uniform loads
Wall dead load
(80) (202/100)=320 lb per ft
Roof dead + live load=1200
Arch dead load in excess
of wall weight (approx.)=260
Total uniform load = 1780 lb per It
W = (1780) (20) = 35,600 lb

Complementary parabolic loading
Maximum ordinate
p = (80) (12)=960 lb per ft
Minimum ordinate=0
W = pL / 3 = 960 (20)/3=6400 lb

Numbers before the following paragraphs refer to the outline of the recommended sequence.

1. The principal arch dimensions are:
d=16 in.
t=12 in.
L=20 ft
f=12 ft
f / L=0.6
2. From Table 3, k=2.80



3a. For parabolic loading:
31a http://www.gobrick.com/BIA/technotes/t31a.htm
16 of 19 9/13/2009 1:35 PM
From Table 5,
S=0.5943T=0.4286



From Table 6,
a=6.88g=8.046
b=2.72d=2.834



From the given relationships,
J=1.3522F=4.4884
q=19.20K=1.4182



3b. For vertical uniform loads:
From Table 5,
S=1.0061T=0.6800

From Table 6 (note that a, b, g and d are the same for any given arch dimensions),
a=6.88g=8.046
b=2.72d=2.834



From the given relationships:
31a http://www.gobrick.com/BIA/technotes/t31a.htm
17 of 19 9/13/2009 1:35 PM
J=1.3522F=4.4884
q =19.20K=2.4008

4. The necessary substitution may now be made to evaluate the design moments and forces. In this example,
moments, shears and axial thrusts are determined at increments of 0.1L (each 2 ft of span). Tabular computations
are suggested for ease in evaluating these equations. The results of such tabular computations are shown in Table
7.


Note: The above moments are in foot-pounds. Shears and axial thrusts are in pounds.

The stresses in the arch may be determined from the following equations:

In the above equations, fm denotes maximum and minimum fiber stresses and Vm denotes shearing stresses. All
quantities in the equations must be in units of inches and pounds. Table 8 shows stresses in the arch.
Plus signs indicate compression and minus signs tension, for values of fm. These signs have only directional
significance for values of Vm. No tensile stresses should be permitted in unreinforced masonry arches under static
loading conditions. The reader is referred to Technical Notes 31 for a discussion of mortars in arch construction.
See Table 1 for allowable compressive stresses in brick masonry.

31a http://www.gobrick.com/BIA/technotes/t31a.htm
18 of 19 9/13/2009 1:35 PM

Note: The above stresses are in pounds per square inch.

Interpretation of Results. Table 8 indicates that the compressive stresses are, in all instances, less than the
allowable 300 psi. No tensile stresses exist. The shearing stresses are insignificant. The arch is adequate. The
moments and shears are caused by other than uniform loads. For this arch, and perhaps for most arches, the
predominant load is uniformly distributed. As a result the moments and shears are relatively small and the arch is
predominantly in compression.
31a http://www.gobrick.com/BIA/technotes/t31a.htm
19 of 19 9/13/2009 1:35 PM
INTRODUCTION
A lintel is a structural member placed over an opening
in a wall. In the case of a brick masonry wall, lintels may
consist of reinforced brick masonry, brick masonry arches,
precast concrete or structural steel shapes. Regardless
of the material chosen for the lintel, its prime function is to
support the loads above the opening, and it must be
designed properly. To eliminate the possibility of structur-
al cracks in the wall above these openings, the structural
design of the lintels should not involve the use of "rule-of-
thumb" methods, or the arbitrary selection of structural
sections without careful analysis of the loads to be carried
and calculation of the stresses developed. Many of the
cracks which appear over openings in masonry walls are
due to excessive deflection of the lintels resulting from
improper or inadequate design.
This Technical Notes presents the considerations to
be addressed if structural steel lintels are to be used. It
also provides a procedure for the structural design of
these lintels. For information concerning reinforced brick
masonry lintels, see Technical Notes 17H and for brick
masonry arches, see Technical Notes 31, 31A and 31C
Revised.
CONSIDERATIONS
General
When structural steel lintels are used, there are sev-
eral considerations which must be addressed in order to
have a successful design. These include loading, type of
lintel, structural design, material selection and mainte-
nance, moisture control around the opening, provisions to
avoid movement problems and installation of the lintel in
the wall.
Types
There are several different types of structural steel lin-
tels used in masonry. They vary from single angle lintels
in cavity or veneer walls, to steel beams with plates in
solid walls, to shelf angles in brick veneer panel walls.
Most building codes permit steel angle lintels to be used
for openings up to 8 ft 0 in. (2.4 m). Openings larger than
this are usually required to have fire protected lintels.
Loose Angle Lintels. Loose angle lintels are used in
brick veneer and cavity wall constructions where the lintel
is laid in the wall and spans the opening. This type of lin-
tel has no lateral support. Figure 1a shows this condition.
Combination Lintels. In solid masonry walls, single
loose angle lintels are usually not capable of doing the
job. Therefore, combination lintels are required. These
combination lintels can take many forms, from a clustering
of steel angles, such as shown in Figs. 1b and 1c, to a
combination of steel beam and plates, as shown in Figs.
1d and 1e.
Angle Lintels - In solid masonry walls, it is usually sat-
isfactory to use multiple steel angles as a lintel. These
angles are usually placed back to back, as shown in Figs.
1b and 1c.
Steel Beam/Plate Lintels - In solid walls with large super-
imposed loads, or in walls where the openings are greater
than 8 ft 0 in. (2.4 m), it may be necessary to use lintels com-
posed of steel beams with attached or suspended plates, as
shown in Figs. 1d and 1e. This permits the beam to be fully
encased in masonry, and fire-protected.
Shelf Angles. In panel walls systems, the exterior
wythe of brickwork may be supported by shelf angles
rigidly attached to the structural frame. These shelf
angles, in some cases, also act as lintels over openings in
the masonry. This condition is shown in Fig.1f.
Reissued*
May
1987
31B
REVISED
Technical Notes
on Brick Construction
Brick Industry Association 11490 Commerce Park Drive, Reston, Virginia 20191
STRUCTURAL STEEL LINTELS
Abstract: The design of structural steel lintels for use with brick masonry is too critical an element to be left to
rule-of-thumb" designs. Too little concern for loads, stresses and serviceability can lead to problems.
Information is provided so that structural steel lintels for use in brick masonry walls may be satisfactorily
designed.
Key Words: beams (supports); brick; buildings; deflection; design; lintels; loads (forces); masonry; struc-
tural steel; walls.
*Originally published in Nov/Dec 1981, this Technical Notes has been reviewed and reissued.
Design
The proper design of the structural steel lintel is very
important, regardless of the type used. The design must
meet the structural requirements and the serviceability
requirements in order to perform successfully. Design
loads, stresses and deflections will be covered in a later
section of this Technical Notes.
Materials
The proper specification of materials for steel lintels is
important for both structural and serviceability require-
ments. If materials are not properly selected and main-
tained, problems can occur.
Selection. The steel for lintels, as a minimum, should
comply with ASTM A 36. Steel angle lintels should be at
least 1/4 in. (6 mm) thick with a horizontal leg of at least 3
1/2 in. (90 mm) for use with nominal 4 in. (100 mm) thick
brick, and 3 in. (75 mm) for use with nominal 3 in. (75
mm) thick brick.
Maintenance. For harsh climates and exposures,
consideration should be given to the use of galvanized
steel lintels. If this is not done, then the steel lintels will
require periodic maintenance to avoid corrosion.
Moisture Control
Proper consideration must always be given to mois-
ture control wherever there are openings in masonry
walls. There must always be a mechanism to channel the
flow of water, present in the wall, to the outside.
Flashing and Weepholes. Even where galvanized
or stainless steel angles are used for lintels in cavity and
veneer walls, continuous flashing should be installed over
the angle. It should be placed between the steel and the
exterior masonry facing material to collect and divert
moisture to the outside through weepholes. Regardless
of whether flashing is used, weepholes should be provid-
ed in the facing at the level of the lintel to permit the
escape of any accumulated moisture. See Technical
Notes 7A for further information on flashing and weep-
holes.
Movement Provisions
Because of the diversity of movement characteristics
of different materials, it is necessary to provide for differ-
ential movement of the materials. This is especially true
at locations where a number of different materials come
together. Technical Notes 18 Series provides additional
information on differential movement.
Expansion Joints. Expansion joints in brick masonry
are very important in preventing unnecessary and unwant-
ed cracking. There are two types of expansion joints
which will need to be carefully detailed when lintels are
involved: vertical and horizontal.
Vertical - Vertical expansion joints are provided to per-
mit the horizontal movement of the brick masonry. Where
these expansion joints are interrupted by lintels, the
expansion joint should go around the end of the lintel and
then continue down the wall.
Fig. 1
Types of Structural Steel Lintels
2
Horizontal - In multi-story walls where the lintels are a
continuation of shelf angles supporting masonry panels,
horizontal expansion joints to accommodate vertical
movement must be provided. Often a simple soft joint
below the shelf angle is all that is needed. See Technical
Notes 18A, 21 Rev, and 28B Rev for typical details.
Installation
The installation of steel lintels in masonry walls is a
conventional construction operation, familiar to most
members of the building team. The walls are built to the
height of the opening, the lintel is placed over the open-
ing, and the masonry work is continued. One item of spe-
cial construction that must be noted is temporary shoring.
Temporary Shoring. If the steel lintel is being
designed assuming in-plane arching of the masonry
above, then the lintel must be shored until the masonry
has attained sufficient strength to carry its own weight.
This shoring period should not be less than 24 hr. This
minimum time period should be increased to three days
when there are imposed loads to be supported. If the
masonry is being built in cold weather construction condi-
tions, the length of cure should be increased. If the lintel
is designed for the full uniform load of the masonry and
other superimposed loads ignoring any inherent arching
action, then no shoring is required.
STRUCTURAL DESIGN
General
The structural design of steel lintels is relatively sim-
ple. The computations are the same as for steel beams
in a building frame, but because of the low elasticity of the
masonry, and the magnitude and eccentricity of the load-
ing, the design should not be taken lightly. A proper
design must consider the loads, stresses, and serviceabil-
ity of the system. If these are not properly taken into
account, problems of cracking and spelling could occur.
Loads
The determination of imposed loads is an important
factor. Fig. 2 shows an example of a lintel design situa-
tion. On the left is an elevation showing an opening in a
wall with planks and a beam bearing on the wall. On the
right is a graphic illustration of the distribution of the
superimposed loads.
Uniform Loads. The triangular wall area (ABC) in
Fig. 2b above the opening has sides at 45-deg angles to
the base. Arching action of a masonry wall will carry the
dead weight of the wall and the superimposed loads out-
side this triangle, provided that the wall above Point B
(the top of the triangle) is sufficient to provide resistance
to arching thrusts. For most lintels of ordinary wall thick-
ness, loads and spans, a depth of 8 to 16 in. (200 mm to
400 mm) above the apex is sufficient. If stack bonded
masonry is used, horizontal joint reinforcement must be
provided to ensure the arching action.
Providing arching action occurs, the dead weight of
the masonry wall, carried by the lintel, may be safely
assumed as the weight of masonry enclosed within the tri-
angular area (ABC). To the dead load of the wall must be
added the uniform live and dead loads of the floor bearing
on the wall above the opening and below the apex of the
45-deg triangle. Again, providing arching occurs, such
loads above the apex may be neglected. In Fig. 2b, D is
greater than L/2, so the floor load may be ignored, but, in
order to use this assumed loading, temporary shoring
must be provided until the masonry has cured sufficiently
to assure the arching action.
If arching action is not assumed and temporary short-
ing is not to be used, the steel lintel must be designed for
the full weight of the masonry and other superimposed
live and dead loads above the opening. There could be
quite a substantial difference in the final lintel sizes
required in each case.
Fig. 2
Lintel Load Determination
3
Concentrated Loads. Concentrated loads from
beams, girders, or trusses, framing into the wall above the
opening, must also be taken into consideration. Such
loads may be distributed over a wall length equal to the
base of the trapezoid and whose summit is at the point of
load application and whose sides make an angle of 60
deg with the horizontal. In Fig. 2b, the portion of the con-
centrated load carried by the lintel would be distributed
over the length, EC, and would be considered as a par-
tially distributed uniform load. Arching action of the
masonry is not assumed when designing for concentrated
loads. Again, if stack bonded masonry is used, horizontal
joint reinforcement must be provided to assure this distrib-
ution.
Stresses
After the loads have been determined, the next step
in the design of the lintel is the design for stresses.
Which stresses need to be checked will depend upon the
type and detailing of the lintel.
Flexure. In a simply supported member loaded
through its shear center, the maximum bending moment
due to the triangular wall area (ABC) above the opening
can be determined by:
Mmax = WL
6
where:
Mmax = maximum moment (ft---lb)
W = total load on lintel (lb)
L = span of lintel, center to center of end bearing (ft)
As an alternative, the designer may wish to calculate an
equivalent uniform load by taking 2/3 of the maximum
height of the triangle times the unit weight of the masonry
as the uniform load across the entire lintel. If this is done,
the maximum bending moment equation becomes:
Mmax = wL
2
8
where:
w = equivalent uniformly distributed load per unit
of length (lb per ft).
To this bending moment should be added the bending
moment caused by the concentrated loading, if any.
Where such loads are located far enough above the lintel
to be distributed as shown in Fig. 2b, the bending moment
formula for a partially distributed uniform load may be
used. Such formulae may be found in the " Manual of
Steel Construction," by the American Institute of Steel
Construction (AISC). Otherwise, concentrated load bend-
ing moments should be used.
The next step is the selection of the required section.
The angle, or other structural steel shape, should be
selected by first determining the required section modu-
lus. This becomes:
S = 12Mmax
Fb
where:
S = section modulus (in
3
)
Fb = allowable stress in bending of steel (psi)
The allowable stress, Fb, for ASTM A 36 structural steel is
22,000 psi (150 MPa) for members laterally supported.
Solid brick masonry walls under most conditions provide
sufficient lateral stiffness to permit the use of the full
22,000 psi (150 MPa). This is especially true when floors
or roofs frame into the wall immediately above the lintel.
The design for non-laterally supported lintels should be in
accordance with the AISC Specification for the Design,
Fabrication and Erection of Structural Steel for Buildings.
Using the design property tables in the AISC Manual,
a section having an elastic section modulus equal to, or
slightly greater than, the required section modulus is
selected. Whenever possible, within the limitations of
minimum thickness of steel and the length of outstanding
leg required the lightest section having the required sec-
tion modulus should be chosen.
Combined Flexure and Torsion. In some cases, the
design for flexure will need to be modified to include the
effects of torsion. This is the case in cavity and veneer
walls where the load on the angle is not through the shear
center.
In some situations, such as veneers, panel or curtain
walls, the lintel may be supporting only the triangular por-
tion of masonry directly over the opening. If this is the
case, then the torsional stresses will usually be negligible
compared to the flexural stresses, and can be safely
ignored.
If, on the other hand, there are imposed uniform loads
within the triangle or imposed concentrated loads above
the lintel, then a detailed, combined stress analysis will be
necessary. The design of a lintel subjected to combined
flexure and torsion should be in accordance with the AISC
Specification for the Design, Fabrication and Erection of
Structural Steel for Buildings.
Shear. Shear is a maximum at the end supports, and
for steel lintels it is seldom critical. However, the computa-
tion of the unit shear is a simple calculation and should
not be neglected. The allowable unit shear value for
ASTM A 36 structural steel is 14,500 psi (100 MPa). To
calculate the shear:
vmax = Rmax
AS
where:
Vmax = the actual maximum unit shear (psi)
Rmax = maximum reaction (lb)
As = area of steel section resisting shear (sq. in.)
4
Bearing. In order to determine the overall length of a
steel lintel, the required bearing area must be determined.
The stress in the masonry supporting each end of the lin-
tel should not exceed the allowable unit stress for the type
of masonry used. For allowable bearing stresses, see
"Building Code Requirements for Engineered Brick
Masonry," BIA; "American Standard Building Code
Requirements for Masonry," ANSI A41.1-1953 (R 1970);
or the local building code. The reaction at each end of
the lintel will be one-half the total uniform load on the lin-
tel, plus a proportion of any concentrated load or partially
distributed uniform load. The required area may be found
by:
Ab = Rmax
fm
where:
Ab = required bearing area (sq in.)
fm = allowable compressive stress in masonry (psi)
In addition, any stresses due to rotation from bending
or torsion of the angle at its bearing must be taken into
account.
Since in selecting the steel section, the width of the
section was determined, that width divided into the
required bearing area, Ab, will determine the length of
bearing required, F and F1, in Fig. 2b. This length should
not be less than 3 in. (75 mm).
If the openings are close together, the piers between
these openings must be investigated to determine
whether the reactions from the lintels plus the dead and
live loads acting on the pier exceed the allowable unit
compressive stress of the masonry. This condition will not
normally occur where the loads are light, such as in most
one and two-story structures.
Serviceability
In addition to the stress analysis for the lintel, a ser-
viceability analysis is also important. Different types of
lintels have different problems of deflection and rotation,
and each must be analyzed separately to assure its prop-
er performance.
Deflection Limitations. After the lintel has been
designed for stresses, it should be checked for deflection.
Lintels supporting masonry should be designed so that
their deflection does not exceed 1/600 of the clear span
nor more than 0.3 in (8 mm) under the combined superim-
posed live and dead loads.
For uniform loading, the deflection can be found by:
t = 5wL
4
(1728)
384 EI
where:
t = total maximum deflection (in.)
E = modulus of elasticity of steel (psi)
I = moment of inertia of section (in.
4
)
For loadings other than uniform, such as concentrated
loads and partially distributed loads, deflection formulae
may be found in the AISC Manual.
Torsional Limitations. In cases where torsion is pre-
sent, the rotation of the lintel can be as important as its
deflection. The rotation of the lintel should be limited to
1/16 in. (1.5 mm) maximum under the combined superim-
posed live and dead loads. As mentioned before, all addi-
tional bearing stresses due to angle rotation must be
taken into account in the design for bearing.
Design Aids
In order to facilitate the design of steel angle lintels,
several design aids are included. These design aids are
not all-inclusive, but should give the designer some help
in designing lintels for typical applications. Conditions
beyond the scope of these tables should be thoroughly
investigated.
Table 1 contains tabulated load values to assist the
designer in the selection of the proper size angle lintel,
governed either by moment or deflection under uniform
load. Shear does not govern in any of the listed cases.
The deflection limitation in Table 1 is 1/600 of the span, or
0.3 in. (8 mm), whichever is less. Lateral support is
assumed in all cases.
Table 2 lists the allowable bearing stresses taken from
ANSI A41.1-1953 (R 1970). In all cases, allowable bear-
ing stresses set by local jurisdictions in their building
codes will govern.
Table 3 lists end reactions and required length in
bearing, which may control for steel angle lintels.
SUMMARY
This Technical Notes is concerned primarily with the
design of structural steel lintels for use in brick masonry
walls. It presents the considerations which must be
addressed for the proper application of this type of
masonry support system. Other Technical Notes address
the subjects of reinforced brick masonry lintels and brick
masonry arches.
The information and suggestions contained in this
Technical Notes are based on the available data and the
experience of the technical staff of the Brick Institute of
America. The information and recommendations con-
tained herein, if followed with the use of good technical
judgment, will avoid many of the problems discussed.
Final decisions on the use of details and materials as dis-
cussed are not within the purview of the Brick Institute of
America, and must rest with the project designer, owner,
or both.
5
TABLE 1
Allowable Uniform Superimposed Load (lb per ft) for ASTM A 36 Structural Steel Angle Lintels
1,2,3,4,5,6
Horizontal
Leg (in)
2 1/2
3 1/2
Angle Size
(in x in x in)
2 x 2 1/2 x 1/4
2 1/2 x 2 1/2 x 1/4
5/16
3/8
3 x 2 1/2 x 1/4
3 1/2 x 2 1/2 x 1/4
5/16
3/8
2 1/2 x 3 1/2 x 1/4
3 x 3 1/2 x 1/4
3 1/2 x 3 1/2 x 1/4
5/16
3/8
4 x 3 1/2 x 1/4
5/16
5 x 3 1/2 x 5/16
3/8
6 x 3 1/2 x 3/8
Weight per
ft (lb)
3.6
4.1
5.0
5.9
4.5
4.9
6.1
7.2
4.9
5.4
5.8
7.2
8.5
6.2
7.7
8.7
10.4
11.7
Span in Feet (Center to Center of
Required Bearing
Resisting
Moment
(ft-lb)
458
715
880
1045
1027
1393
1705
1998
752
1082
1448
1797
2108
1888
2310
3557
4198
5940
Elastic
Section
Modulus
(in
3
)
0.25
0.39
0.48
0.57
0.56
0.76
0.93
1.09
0.41
0.59
0.79
0.98
1.15
1.03
1.26
1.94
2.29
3.24
Moment of
Inertia
(in
4
)
0.372
0.703
0.849
0.984
1.17
1.80
2.19
2.56
0.777
1.30
2.01
2.45
2.87
2.91
3.56
6.60
7.78
12.90
3 4 5 6 7 8
352 146 73
631 279 141 80
777 336 170 96
923 390 197 112
908 467 237 135 83
1233 692 366 210 130 86
1509 846 446 255 158 104
1769 992 521 298 185 122
664 308 155 88
956 518 263 150 92
1281 718 409 234 145 95
1590 891 498 285 177 116
1865 1046 583 334 207 136
1672 938 594 341 212 140
2046 1147 726 417 260 172
3153 1770 1130 779 487 324
3721 2089 1333 918 574 381
5268 2958 1889 1308 958 638
1
Allowable loads to the left of the heavy line are governed by moment, and to the right by deflection.
2
Fb = 22,000 psi (150 MPa)
3
Maximum deflection limited to L/600
4
Lateral support is assumed in all cases.
5
For angles laterally unsupported, allowable load must be reduced.
6
For angles subjected to torsion, make special investigation.
TABLE 2
Allowable Compressive Stresses (psi) in Masonry
1
Type of Wall
Solid walls of brick or solid units
of clay when average compressive
strength of unit is as follows:
8000 plus psi
4500 to 8000 psi
2500 to 4500 psi
1500 to 2500 psi
Grouted solid masonry of
brick and other solid units of clay
4500 plus psi
2500 to 4500 psi
1500 to 2500 psi
Masonry of hollow units
Type of Mortar
M
400
250
175
125
350
275
225
85
S
350
225
160
115
275
215
175
75
N
300
200
140
100
200
155
125
70
O
200
150
110
75
-
-
-
-
1 Adapted from American Standard Building Code Requirements for Masonry, National Bureau of Standards, ANSI A41. 1-1953 (R 1970).
6
REFERENCES
1. AISC, Manual of Steel Construction, American
Institute of Steel Construction, Inc., New York, New
York, Eighth Edition, 1980.
2. AISC, Specification for the Design, Fabrication
and Erection of Structural Steel for Buildings,
American Institute of Steel Construction, Inc., New
York, New York, 1978.
3. ANSI, American Standard Building Code
Requirements for Masonry, ANSI A41.1-1953 (R
1970), American National Standards Institute, New
York, New York.
4. BIA, Building Code Requirements for Engineered
Brick Masonry, Brick Institute of America, McLean,
Virginia, 1969.
TABLE 3
End Reaction
1
and Required Length of Bearing
2
for Structural Angle Lintels
2 1/2 Leg Horizontal 31/2 Leg Horizontal
fm
psi
400
350
300
275
250
225
215
200
175
160
155
150
140
125
115
110
100
85
75
70
fm
psi
400
350
300
275
250
225
215
200
175
160
155
150
140
125
115
110
100
85
75
70
Length of Bearing Length of Bearing
3 4 5 6
3000 4000 5000 6000
2625 3500 4375 5250
2250 3000 3750 4500
2063 2750 3438 4125
1875 2500 3125 3750
1688 2250 2813 3375
1613 2150 2688 3225
1500 2000 2500 3000
1313 1750 2188 2625
1200 1600 2000 2400
1163 1550 1938 2325
1125 1500 1875 2250
1050 1400 1750 2100
938 1250 1563 1875
863 1150 1438 1725
825 1100 1375 1650
750 1000 1250 1500
638 850 1063 1275
563 750 938 1125
525 700 875 1050
3 4 5 6
4200 5600 7000 8400
3675 4900 6125 7350
3150 4200 5250 6300
2888 3850 4813 5775
2625 3500 4375 5250
2363 3150 3938 4725
2258 3010 3763 4515
2100 2800 3500 4200
1838 2450 3063 3675
1680 2240 2800 3360
1628 2170 2713 3255
1575 2100 2625 3150
1470 1960 2450 2940
1313 1750 2188 2625
1208 1610 2013 2415
1155 1540 1925 2310
1050 1400 1750 2100
893 1190 1488 1785
788 1050 1313 1575
735 980 1225 1470
1
End Reaction in lbs.
2
Length of Bearing in inches.
7

Technical Notes 31C - Structural Design of Semicircular Brick Masonry Arches
Rev [Feb. 1971] (Reissued August 1986)
INTRODUCTION
The semicircular arch is among the most popular arch forms used by architects today. Its smooth, continuous curve
makes it easily adaptable to many architectural styles and applications.
This issue of Technical Notes presents recommended procedures and tables for the structural design of
non-reinforced semicircular and segmental arches. Technical Notes 31 and 31A contain further information about
general arch forms and their design.

DESIGN ASSUMPTIONS

Since the brick masonry arch is usually an integral part of a wall and not free-standing, basic design assumptions are
made which assist in making the analysis.
The spring line is assumed to be located on a horizontal line one fourth the span length above the horizontal axis. The
arches are assumed to be fully restrained at the spring line and the portion of semicircular arch above this line is
analyzed in a manner similar to that for a parabolic arch, using the formulas in Section 10, Method A from Frames
and Arches, by Valerian Leontovich, M.S., McGraw-Hill 1959.


Holiday Inn, McLean, Virginia

31c http://www.gobrick.com/BIA/technotes/t31c.htm
1 of 10 9/13/2009 1:37 PM
NOMENCLATURE
d = arch ring depth, in inches,
f = rise of arch, in feet,
fm = allowable compressive stress, in psi,
H = horizontal thrust, in pounds,
HDL = horizontal thrust, in pounds, caused by a uniform dead load,
L = span length, in feet,
n = number of shear planes (see Technical Notes 31A.)
P = allowable concentrated load, in pounds,
P' = maximum allowable concentrated load in pounds under combined loading provisions,
P* = additional concentrated load capacity caused by the uniform dead load.
t = wall thickness, in inches
vm = allowable shear stress in brick masonry, in psi,
W = allowable uniform load, in pounds per foot,
x = length of wall required, in inches, to resist horizontal thrust

DEVELOPMENT OF TABLES

In the determination of the arch's capacity for uniform loads, the limiting factor was found to be the compressive
strength of the brick masonry. Additional stresses due to the circular loading of the masonry above the intrados are
also taken into consideration.
In determining the capacity for concentrated loads, the limiting factors were found to be bending at the center line of
span, shear at the spring line (vm = 40 psi), and maximum compressive stress (fm). Tensile stresses were not
permitted to develop at mid span.

Since axial forces develop in the arch ring from the concentrated and uniform loads, interaction formulas were
developed for each loading condition. These formulas combine the axial stresses with the bending stresses.

ALLOWABLE LOADING

In all formulas used in this publication, d and t are measured in inches, and L is measured in feet. The following
31c http://www.gobrick.com/BIA/technotes/t31c.htm
2 of 10 9/13/2009 1:37 PM
loading conditions were considered for analyzing a semicircular arch.

Uniform Load. Tables 1, 2, 3 and 4 give the allowable uniform loads occurring over the entire span length for a 1-in.
thick arch ring. Figure 1 illustrates the following requirements and limitations:


FIG. 1


1. Uniform load occurring between lines 1 and 3 (0.90 L and 0.70 L) are those provided for in the load tables.
2. Uniform loads occurring above line 1 may be ignored, at the discretion of the designer, provided arching action
occurs in the brick masonry above the arch ring. (See discussion in Technical Notes 31 and 31A.)
3. There must be a minimum height of masonry (line 3) equal to 0.70 L above the horizontal axis. No superimposed
loads are permitted below this line.
4. The maximum design height of masonry is 0.25 L above the crown for walls higher than line 2.
5. In all cases, the horizontal thrust (H) must be checked as shown in Technical Notes 31A, Fig. 4. For a given arch,
the horizontal thrust is directly proportional to the uniform load.
6. The portion of wall that resists the horizontal thrust is assumed to be non-yielding to any lateral movement.

Concentrated Load. Table 5 gives the allowable concentrated loads occurring at the center line of span for a 1-in.
thick arch ring. Figure 2 illustrates the following requirements and limitations:


31c http://www.gobrick.com/BIA/technotes/t31c.htm
3 of 10 9/13/2009 1:37 PM

FIG. 2


1. Concentrated loads occurring between lines 2 and 3 (1.20 L and 0.75 L) are those provided for in the load table.
2. Concentrated loads occurring between lines 1 and 2 may be divided by the span length (L) and considered as
equivalent uniform loads.
3. Concentrated loads occurring above line 1 (1.50 L) may be ignored, at the discretion of the designer, provided
arching action occurs in the brick masonry above the arch ring. (See discussion in Technical Notes 31 and 31A.)
4. In all cases, condition 4 for uniform loads must be used with the resulting thrusts added to those of the
concentrated loads.
5. There must be a minimum height of masonry (line 3) equal to 0.75 L above the horizontal axis. No superimposed
loads are permitted below this line.
6. In all cases, the horizontal thrust H must be checked as shown in Technical Notes 31A, Fig. 4. For a given arch,
the horizontal thrust is directly proportional to the concentrated load.
7. The portion of wall that resists the horizontal thrust is assumed to be non-yielding to any lateral movement.
31c http://www.gobrick.com/BIA/technotes/t31c.htm
4 of 10 9/13/2009 1:37 PM
31c http://www.gobrick.com/BIA/technotes/t31c.htm
5 of 10 9/13/2009 1:37 PM

*Values may be linearly interpolated except where horizontal lines occur. At these lines, the allowable load is ((0.241)(L + 0.083d^3
+ 0.134)(L + 0.083d)^2 * d)/(1.34(L + 0.083d) - 0.0778d)
or the value above the line whichever is smaller The horizontal: thrust is 0.778P + 0.134(L + 0.083d)
or value above the line whichever is smaller.

Combined Loading. When the uniform loads are combined with concentrated loads, the concentrated load capacity
of the arch ring increases. This additional capacity is due to the compressive stress from the uniform load equalizing
the tensile bending stress at mid span due to the concentrated load (M/S = P/A). This additional capacity may be
expressed by the following formula:


The values of P' and H' in Table 6 are the allowable capacities governed by compression or shear. They should be
used only as a check when combined loadings are used.
In all cases, the actual load must be less than P', and less than the allowable load, P, plus the additional capacity P*.
The total horizontal thrust must be checked and should be less than the maximum allowable for a uniform load.

31c http://www.gobrick.com/BIA/technotes/t31c.htm
6 of 10 9/13/2009 1:37 PM

SEGMENTAL ARCHES

Any segmental arch with f / L' > 0.29 but < 0.50 can be considered as an equivalent semicircular arch as shown in
Fig. 3. Twice the radius is the equivalent L for use with the tables.


FIG. 3


ILLUSTRATIVE EXAMPLE
Design an arch to meet the requirements as shown in Fig. 4. The arch is semicircular; the horizontal axis is 6 ft
above the base; the span, L, is 10 ft; the arch ring depth, d, is 12 in. (11 1/2 in. actual); and the nominal wall
thickness, t, is 8 in. (7 1/2 in. actual). A beam reaction of 5000 lb is located at the center line of the span and 17 ft
above the base. The uniform load consists of 1000 lb per ft dead load and 500 lb per ft live load occurring 14 ft
above the base. Assume fm = 400 psi and the brick masonry weighs 10 psf per 1-in. thickness.

31c http://www.gobrick.com/BIA/technotes/t31c.htm
7 of 10 9/13/2009 1:37 PM

FIG. 4



Uniform Load


All the following calculations will be with 1 in. of wall thickness; actual t = 7.5 in., fm = 400 psi, d = 11.5 in. and L= 10
ft.

Uniform Load

31c http://www.gobrick.com/BIA/technotes/t31c.htm
8 of 10 9/13/2009 1:37 PM

Use Table 2, since 0.7L < 0.8L < 0.9L
From Table 2, W = 724 lb per ft and
H = 2993 lb


Concentrated Load


Use Table 5 since 0.75L < 1.1 L < 1.2L
From Table 5, P = 88 lb and H = 84 lb


However, since there is combined loading, advantage can be taken of the increased capacity due to the uniform
load.


667 < 688 O.K.

31c http://www.gobrick.com/BIA/technotes/t31c.htm
9 of 10 9/13/2009 1:37 PM
From Table 6, P'= 1060 lb
667 < 1060 O.K.


Horizontal Thrust
H(total) =970 + 636 = 1606 < 2993 O.K.
At this point the wall shear caused by the horizontal thrust at the spring line should be checked. Assume Vm = 40 psi
and n = 2


The overturning moment of the support due to horizontal thrust should be checked next (see Technical Notes 31A).
In this example, the horizontal thrust is 1606 (7.5) (8.5) = 102,000 ft-lb.
The resistance to overturning is a function of the overall axial load, wall shape, and reinforcement, if any. This is a
separate analysis that should be performed after considering the total loading conditions on the entire structure.

CONCLUSION

This issue of Technical Notes has presented a simplified but conservative approach to a complex structural design
problem. To provide an analysis for all possible assumptions and loading conditions is beyond the scope of this
publication. Most loading conditions encountered will be similar to those in Fig. 1 and Fig. 2. To load an arch
unsymmetrically defeats its use as a natural load-carrying structure and induces bending stresses that may cause
failure.
If arches are to be loaded unsymmetrically or do not comply with the assumptions and limitations given in this
Technical Notes, consideration should be given to reinforced brick masonry. (See Technical Notes 17A Revised,
"Reinforced Brick Masonry - Flexural Design", and 17M, "Reinforced Brick Masonry Girders - Examples".) If
conditions exist other than those covered in the tables, special analysis should be made by the designer.
The Brick Institute of America can not assume responsibility for the results obtained when using this Technical Notes
Issue. It is beyond the scope of the Institute to anticipate every design situation that may arise. However, so long as
the design criteria agree with the assumptions and limitations, satisfactory results can be obtained which will save
countless hours of calculation time.


31c http://www.gobrick.com/BIA/technotes/t31c.htm
10 of 10 9/13/2009 1:37 PM

Technical Notes 36 - Brick Masonry Details, Sills, and Soffits
Rev [July/Aug. 1981] (Reissued Jan. 1988)
Abstract: Detailing of brick masonry is both an art and a science. Recommendations are provided for the
development of successful details using brick masonry and other materials. Detailing of sills and soffits is specifically
addressed. Performance, esthetic value and economics are the principal considerations in the development of
successful details.
Key Words: brick, connections, construction, design, detailing, economics, esthetic value, function,
performance, prefabrication, sills, soffits, structural stability.
INTRODUCTION
Successful detailing of brick masonry is both an art and a science. Proper details should result in a structure which is
pleasing to the eye, but more importantly, performs well over its lifetime. Good detailing is not accidental, it requires
proper planning. This planning may involve close cooperation between the architectural, engineering and construction
disciplines in the early stages of the design process.
There are three items which should be considered in the development of a successful detail. These are: 1.
Performance considerations; 2. Esthetic value considerations; and 3. Economic considerations. The last two of these
items may be traded off against each other. But, the first is mandatory and if it is not the primary concern, the detail
may, and probably will, be doomed to failure. This failure can manifest itself in several ways: cracking, structural
failure, moisture penetration to the interior, or efflorescence, to mention a few.
It is possible to have a successful detail while compromising either the esthetic value or the economic
considerations. But, it is impossible to have a successful detail if the performance considerations are compromised.
A successful detail can be developed with excellent esthetic value while completely ignoring the economic
considerations or vice versa, but to ignore the performance considerations is to invite trouble.
APPROACH TO DETAILING
General
Proper planning in the development of brick details is essential to the successful execution of that detail in the field.
The designer must be familiar not only with the properties of the various materials involved, but also how they go
together in the construction process and how they will perform, both individually and together in service. The most
esthetically pleasing detail is of no benefit if it can't be built, or does not perform its intended function.
The designer should always keep in mind that different materials react to temperature and moisture changes in
different ways. While in some cases these differences may be minor, in others they may be significant. If they are
not properly addressed, the result can be facade failures, such as leaking, bowing, cracking, etc. For a discussion of
differential movement, see Technical Notes 18 Series.
Performance Considerations
Performance is all-important if the detail is to be successful. There are three items which must be considered in the
development of a detail which will provide satisfactory performance. They are: 1. Functional considerations; 2.
Structural stability; and 3. Construction considerations. In the development of the detail, it is imperative that all of
these items be given proper consideration.
Functional Considerations. One of the first steps in the development of a successful detail is to determine the
function of the element. The designer must determine the purpose of the element, and how the element will affect
the overall performance of the building. Typical questions which should be addressed are: 1. Is the element to serve
http://www.gobrick.com/BIA/technotes/t36.htm
1 of 11 9/13/2009 1:38 PM
as a weather-tight enclosure? 2. Will stresses, axial, flexural or shear, be developed in the member? 3. Should it
channel and direct the flow of moisture? 4. Is it to seal the top of a vertical element? 5. Is its purpose merely for
esthetic value? Only after the designer has determined the required functions of the element can he begin to
consider the other factors which will dictate the final design.
Structural Stability. The designer must develop a detail which ensures that all applied loads can be adequately
resisted by the element or that they are transferred to other elements of the structure which can resist them. These
applied loads may be axial, transverse, shear or in the case of prefabricated elements, loads due to transportation
and erection. One area of concern is the manner and adequacy of the connection of the element to the structure. It
is imperative that these connections be structurally sound, to ensure structural stability of the element.
Construction Considerations. The designer should take great care to ensure that the details can be easily
executed in the field. This requires that the designer be knowledgeable in current construction practices. While some
innovation may be necessary and beneficial, the detail should not require radical deviation from conventional
construction practices. Typically, the more simple and straightforward the detail is, the easier it is to construct and
thus, the better its performance. In some instances, the construction can be simplified by prefabrication of the
element. Care should be taken by the designer to ensure, to the greatest extent possible, that the detail does not
require several crafts to be working in the same location at the same time.
Esthetic Value Considerations
The designer must also determine how best to fulfill the functional requirements and yet provide the desired esthetic
value. This involves decisions on materials, colors and textures, and other esthetic considerations. The configuration
of the element is also an important esthetic consideration. The designer may decide to project or recess parts of the
element to provide shadow lines or to use a different bond pattern to call attention to the detail. The esthetic value of
the detail is limited only by its function, its ease of construction, the designer's imagination and possibly its economic
feasibility.
Economic Considerations
A detail, to be successful, should have the capability of being constructed economically. Economics involves both
materials and labor. A successful detail requires that both the quantity and quality of materials be closely controlled.
The use of excess materials to achieve the function of the detail should be avoided.
Details which require very specialized skills by the crafts involved should be avoided. If very specialized skills are
required, there is usually a reduction in productivity of the craftsmen and an increase in cost.
SILLS
General
The prime function of a sill is to channel water away from the building. The sill may consist of a single unit or multiple
units; it may be built in place or prefabricated; and it may be constructed of various materials.
Esthetic Value
The desired esthetic effect may be achieved through the use of special shaped units, either manufactured or cut to
the desired shape. A word of caution concerning manufactured special shapes-while most manufacturers are
capable of making special shapes to match the color and texture of the units selected for the project, there will be an
added cost for each special-shaped unit. The added cost for the special shapes is dependent upon the complexity of
the configuration of the shape and the number of units of each special shape required. Some manufacturers carry
certain special shapes in stock. It may be advantageous to slightly alter the detail so that these stock special shapes
may be used in lieu of one with a slightly different configuration.
The appearance of the sill and the overall esthetic appeal of the structure may also be achieved by the use of a
contrasting color or texture or by use of materials other than brick for the sill. Esthetics may also be affected by the
use of a different bond pattern than that used in the adjacent wall. See Figs. 1 and 2.
http://www.gobrick.com/BIA/technotes/t36.htm
2 of 11 9/13/2009 1:38 PM
General Studies/Classroom/Instructional Resources Center -
State University Agricultural and Technical College - Delhi, New York
FIG. 1
Westgate Corporation Building - McLean, Virginia
FIG. 2
Materials
Sills for use in brick masonry construction are typically brick, concrete, stone or metal. The selection of material is
primarily dependent upon the required esthetic effect. But it is also important to note that metal, concrete and stone
sills normally require fewer joints than do brick sills, and therefore provide fewer potential avenues for water
penetration. Once the decision of which material to use is made, then decisions concerning the quality of that
material can be made. Whichever material is selected, it should be of high quality. A discussion of brick and mortar
properties is found in Technical Notes 7B Revised.
Flashings for use in sills can be of a number of materials, such as copper, lead or plastics, see Technical Notes 7A
Revised for additional information. Aluminum and asphaltic-impregnated felt are not recommended for use as
flashing materials. Aluminum is not recommended since alkalies in the cement of the mortar may attack it and cause
corrosion. Asphaltic-impregnated felt is not recommended because it is easily punctured during construction. For the
same reason, plastic films of less than 20 mil thickness should also be avoided. Once the flashing has been
punctured, it ceases to fulfill its function, thus in place flashing should be inspected for punctures and tears, and
appropriately repaired prior to laying brick masonry on the flashing. Also, some plastics are subject to continued
degradation after having been exposed to sunlight for an extended period of time.
Details
General. Since the primary function of sills is to divert water away from the building, the top surface should slope
downward and away from the building. In the case of brick sills, see Figures 3 and 4, the slope should be at least 15
http://www.gobrick.com/BIA/technotes/t36.htm
3 of 11 9/13/2009 1:38 PM
deg from horizontal. This may vary somewhat according to the sill configuration of the window unit, particularly in the
case of wood windows. The sill should extend a minimum of 1 in. (25 mm) beyond the face of the wall at its closest
point to the wall, see Fig. 3. In some instances, it may be necessary that the brick units at the ends of the sills be
uncored units so that no cores are exposed to view.
Sill in Frame/Brick Veneer Construction
FIG. 3
Sill in Cavity Wall Construction
FIG. 4
When concrete or stone sills are used, they should be sloped away from the building, and also sloped from the ends
toward the center, see Figs. 5 and 6. The slope away from the building should be at least 15 deg from horizontal,
the slope from the ends should be 1/8 in. (3 mm) to 12 in. (300 mm) toward the center of the sill. For sills longer
than 4 ft ( 1.2 m), the slope should extend for at least a distance of 2 ft (600 mm) from the ends, see Fig. 6.
http://www.gobrick.com/BIA/technotes/t36.htm
4 of 11 9/13/2009 1:38 PM
Concrete or Stone Sill
FIG. 5
Concrete or Stone Sill
FIG. 6
http://www.gobrick.com/BIA/technotes/t36.htm
5 of 11 9/13/2009 1:38 PM
Flashing and Weepholes. In general, when a collar joint, cavity or air space is interrupted, such as at sills, at the
base of the walls, at lintels over openings and at shelf angle supports, flashing should be provided in the wall. The
function of flashing is to serve as a collector for any moisture penetrating the wall or the sill. It is important that the
flashing extend through the brick to the exterior face of the wall at the lower end of the flashing and be turned down
at least to in. (6 mm) to form a drip. The flashing at the sill should extend beyond the ends of the sill to the first head
joint outside of the jamb of the opening, and should be turned up and outward for a distance of at least 1 in. (25 mm)
at each end, see Fig. 5. If the ends are not turned up and out, the moisture collected on the flashing will have a path
into the adjacent wall and there is no way to predict where it may go. The purpose of turning the flashing up and out
is to assure that the moisture stays on the flashing until it drains from the wall. See Technical Notes 7A Revised for
a discussion of materials to be used as flashing.
Once moisture penetrating the wall or sill has been collected on the flashing, it must be removed from the wall. This
is the function of weepholes. Weepholes may be installed in several ways, see Technical Notes 21C. Weepholes
should be placed on top of the flashing, not one course up. If wick-type materials are employed, or if hidden flashing
is used, the weepholes should have a maximum horizontal spacing of 16 in. (400 mm). If open weepholes with no
wicks are used, the horizontal spacing may be increased to 24 in. (600 mm) maximum.
Drips. Every sill should be provided with a drip. The function of the drip is to prevent water from returning to the
exterior face of the wall. The drip of a properly sloped brick sill is the lower corner of the brickwork. A drip in a
concrete or stone sill is usually formed, or cut into the bottom face of the sill, as shown in Figs. 5 and 6. The drip on
a concrete or stone sill can be cut in several shapes, Vee-shape, rectangular, semi-circular, or a combination of
these. The shape of the drip is not important, but its presence and location are important. The inner lip of the drip
should be located a minimum of 1 in. (25 mm) from the exterior face of the wall, , as shown in Fig. 5.
Connections. In brick masonry sills of short length, 4 ft. (1.2 m) or less, no special anchorage is necessary.
However, sills of brick, concrete, metal and stone having long runs should be anchored to the masonry below or
behind the sill, see Figs. 3 and 5. This will require penetration of the flashing below or behind the sill. Care must be
taken to ensure that these penetrations are adequately sealed so that the flashing functions as intended.
Attachment of the sill to the window will vary with window type and manufacturer. It is most important that the joint
where the sill and window make contact be sealed with a high-quality sealant, see Technical Notes 28 Revised and
28B Revised.
Expansion Joints. When expansion joints are necessary, it may be desirable to install them in vertical alignment
with window jamb lines. If this is done, the expansion joint should also be installed through the sill. This will enable
the expansion joint to perform as intended.
If the sill extends beyond the jamb of the opening and an expansion joint is required at the jamb, then the expansion
joint should be continuous around the entire sill extension, as should the flashing, see Fig. 6.
Construction
In the past, sills for use in brick masonry construction have generally been built in place, using conventional
construction practices. A trend during recent years has been to use prefabricated sills, particularly when combined
with a spandrel and soffit, see Technical Notes 40 Series. This type of construction will be further discussed in the
Soffits portion of this Technical Notes.
When prefabricated brick, pre-cast concrete, or stone sills are used, they should have section lengths as long as is
practical. The lengths will be determined by ease of handling and erection and the sills' ability to resist erection
stresses. The length of sill sections should be limited to a length that can easily be handled by equipment already on
the jobsite. The joints between long sill sections should be constructed using a soft joint. It may also be necessary, in
very long runs of sill, to provide expansion joints at the ends where the sill abuts the jamb.
SOFFITS
General
Detailing of soffits for brick masonry requires special considerations. The primary function of a brick masonry soffit is
to enclose the building while providing an esthetically pleasing appearance. There are two primary considerations in
http://www.gobrick.com/BIA/technotes/t36.htm
6 of 11 9/13/2009 1:38 PM
addition to esthetic value in the detailing of soffits: the structural stability of the system and whether it can be easily
and economically constructed using conventional methods.
Though prefabrication has not been widely used for total projects, it has been successfully used in many specialized
applications and is considered a conventional construction method. Prefabrication has been widely used in the
construction of soffits and may provide the most economical approach on certain projects. Construction of soffits in
place often requires expensive forming and shoring. However, if there is only a small area of soffits involved on a
given project, this may be the most efficient method.
Materials
Soffits generally are reinforced and grouted in some manner, whether built in place or prefabricated. Several
projects have been constructed using reinforced and grouted hollow units, conforming to ASTM C 652. See
Technical Notes 17 for information on reinforcement and grout. Properties for brick and mortar are discussed in
Technical Notes 7B Revised.
There are several high bond mortar additives available which may allow the designer to eliminate the reinforcement
and grout. However, it should be noted that the high bond mortars do not work well with all brick units. The
instructions of the additive manufacturer must be strictly followed, and a pre-design testing program should be
carried out, see Technical Notes 39A.
Design
There are several questions which must be answered when designing soffits. Some deal with esthetic value, some
with structural stability and some with construction. The primary esthetic concern is configuration. Should the soffit
be horizontal, or sloped, should it be integral with the spandrel or separate? The primary design concern is
structural. How should it be detailed to assure structural soundness, under all loading conditions, including any loads
imparted during erection? This may require detailing and construction practices unfamiliar to the designer and
contractor since this is not like a wall and demands careful consideration. One of the earliest decisions to be made
about the construction of the soffit is whether it will be best to construct it in place, or to prefabricate it. The
configuration, structural and economic considerations may dictate the method of construction to be used. See Figs.
7 and 8.
BIA Headquarters Building - McLean, Virginia
FIG. 7
http://www.gobrick.com/BIA/technotes/t36.htm
7 of 11 9/13/2009 1:38 PM
Evans Library, Texas A&M University - College Station, Texas
FIG. 8
Details
General. Each soffit entails its own unique detailing problems. These may include: configuration, support available
from the surrounding structure, space restriction on built in place soffits and construction sequencing. The manner in
which these problems are solved will determine how successfully the soffit will perform.
Flashing and Weepholes. Normally, soffits do not require flashing or weepholes. However, in some applications,
both may be required, see Fig. 9. In other applications, only weepholes may be required, since the inclusion of
flashing in some cases may impair the structural stability of the soffit. It can only be stressed that the detailer should
always keep in mind the primary function of flashing and weepholes in determining whether they are needed in any
particular application. Their primary functions are:
Flashing - collect and divert to the weepholes any moisture which might penetrate the element.
Weepholes - convey all collected and diverted water to the exterior.
Built-in-Place Brick Soffit
FIG. 9
Connections. Whether the soffit is prefabricated or built in place, its connection to the structure is the most
demanding detail for the designer to develop. Previously developed details may be totally inappropriate in the
present situation. Connection details are critical in providing structural stability to the soffit. In detailing connections, it
is important to keep one principal always in mind. That principle is: Keep It Simple. The simplest connection details
are in most cases the most successful.
http://www.gobrick.com/BIA/technotes/t36.htm
8 of 11 9/13/2009 1:38 PM
Expansion Joints. The installation of expansion joints, in most cases, should be avoided in soffits; however, it may
be necessary to provide expansion joints when soffits are to be installed over large areas. The installation of
expansion joints may cause problems in providing structural stability of the element and require additional
connections to the structure. If it is necessary that expansion joints be installed in soffits, it is important to remember
that the function is expansion control. This is provided by resilient joints which can be compressed to provide for the
movement of brick masonry, especially during hot weather, due to thermal expansion of the brick masonry and return
to its original shape when the temperature is cooler. Reinforced and grouted brick masonry does not usually require
expansion joints.
Construction
Structural and economic considerations normally determine the construction methods to be used. While the detailer
does not normally specify the manner in which the detail is to be executed during construction, the method of support
and economic aspects determined by the detail will affect the method of construction chosen.
The method of supporting the soffit, both its permanent support and support during construction, has a direct bearing
on the method of construction selected. The economics of constructing the element can be affected by configuration,
structural support and materials selection. Economics in turn may well be the final determining factor in the selection
of the construction methods employed.
When a soffit is constructed in place, it sometimes requires a complicated system of centering and falsework which
must be left in place for a number of days. Normal practice is to provide spacer strips on the forms which locate
each unit within the form and provide a joint on the exposed face suitable for tuckpointing once the form is removed,
see Fig. 10. These strips should be the width of the joint and a minimum of 1/2 in. (13 mm) in height. After the units
have been placed on the form, the upper side is grouted and ties are placed in the joints for anchorage to the
structure. After several days of curing, the forms are stripped and the joints can then be tuckpointed. The number of
days required for curing is dependent upon conditions at the site during the curing period and the materials used.
See Technical Notes 7 for tuckpointing recommendations.
Built-in-Place Brick Soffit Forming
FIG. 10
In some cases, the use of built in place soffits may be precluded. Then, prefabrication may be the most logical and
economical approach, see Technical Notes 40 Series. This method of construction has been used very satisfactorily
on many projects. On most of the projects where it has been used, the soffit is built integral with a spandrel cover
and a sloped sill, see Figs 11 and 12.
http://www.gobrick.com/BIA/technotes/t36.htm
9 of 11 9/13/2009 1:38 PM
Prefabricated Brick Sill, Spandrel and Soffit
FIG. 11
Prefabricated Brick Sill, Spandrel and Soffit
FIG. 12

http://www.gobrick.com/BIA/technotes/t36.htm
10 of 11 9/13/2009 1:38 PM
SUMMARY
The designer, when developing details for sills and soffits, should keep in mind the function of the element being
detailed, the esthetic value he wishes to achieve, the structural stability of the element, and the economics of
construction. It is essential to provide details which allow the elements to perform their primary functions as well as
possible. In order to do this, the designer must select the proper materials, locate them in the proper place, and
provide sufficient information so that the element can be properly constructed. Several decisions and assumptions
must be made by the designer because each project and each element on the project must be satisfactorily
addressed.
This Technical Notes addresses the major considerations necessary to successfully detail sills and soffits of brick
masonry. In some cases, other considerations may be necessary due to unusual or unique conditions. It is beyond
the scope of this Technical Notes to address all conditions and combinations of conditions which may occur,
therefore the designer or owner, or both, must make the final decision on the details, the materials selected and the
construction procedures used. The recommendations made in this Technical Notes are merely
that-recommendations. The final configuration of the detail must in the long run be based on the designer's
application of some or all of the principles set forth here.
http://www.gobrick.com/BIA/technotes/t36.htm
11 of 11 9/13/2009 1:38 PM

Technical Notes 36A - Brick Masonry Details, Caps and Copings, Corbels and Racking
Rev [Sept./Oct. 1981] (Reissued February 2001)
Abstract: Recommendations are provided for the development of successful details using brick masonry. Detailing
of caps, copings, corbels and racking is specifically addressed. Performance, esthetic value and economics are the
principal considerations in the develops meant of successful details.
Key Words: brick, caps, connections, construction, copings, corbels, design, detailing, economics, esthetic
values, function, performance, racking, structural stability.
INTRODUCTION
This Technical Notes is the second in a series that discusses brick masonry details. This Technical Notes will
address the detailing of caps, copings, corbels and racking. Technical Notes 36 Revised addresses the detailing of
sills and soffits.
The recommended approach to detailing is covered in Technical Notes 36 Revised. While that Technical Notes is
primarily for sills and soffits, it does provide the general approach applicable to all detailing. The following items
should be considered in the development of a successful detail: 1. Functional considerations; 2. Esthetic value; 3.
Construction considerations; 4. Economic considerations.
DEFINITIONS
Caps and Copings
The definitions for cap and coping are entirely dependent upon which dictionary or glossary is used as a reference.
In addition, there are other terms which are used interchangeably with them, such as water table, canting strip, and
offset. For the purpose of this Technical Notes, the word "coping" applies to the covering at the top of a wall, and
the term "cap" refers to a covering within the height of the wall, normally where there is a change in wall thickness.
The other terms cited will not be used.
Corbels and Racking
A corbel is defined as a shelf or ledge formed by projecting successive courses of masonry out from the face of the
wall. Racking is defined as masonry in which successive courses are stepped back from the face of the wall.
CAPS AND COPINGS
General
The primary function of caps and copings is to channel water away from the building. The cap or coping may be a
single unit or multiple units. They may be of several different materials. The tops may slope in one direction or both
directions. Additionally, where caps are discontinuous, a minimum slope from the ends of 1/8 in. (3 mm) in 12 in.
(300 mm) should be provided, as shown in Figures 4 and 6 in Technical Notes 36 Revised.
The esthetic value the designer wishes to achieve may come from the configuration of the element, its color, or its
texture. Caps and copings normally do not serve any structural function, and do not present any major problems in
their construction.
Materials
Caps and copings can be constructed of several materials: brick, pre-cast or cast-in-place concrete, stone, terra
cotta, or metal. It should be pointed out that because of their location in the structure, caps and copings are exposed
http://www.gobrick.com/BIA/technotes/t36a.htm
1 of 8 9/13/2009 1:39 PM
to climatic extremes. This severe exposure must be of prime concern to the designer. Because caps and copings
are subjected to extreme exposure, brick masonry may not be the best choice of materials. This is because caps
and copings of brick require more joints than do those made of other materials. This provides more avenues for
possible water penetration into the wall. If brick is the material selected, great care must be taken to provide for the
movement to which the element will be subjected and also to make sure all joints are properly filled with mortar.
Concrete, stone and metal caps and copings can be installed in relatively long pieces, thus requiring less joints than
do those made from brick.
Concrete, stone and terra cotta all have thermal expansion properties similar to those of brick masonry and normally
present no extreme problems with differential movement when applied as caps and copings, if properly detailed.
Metal has very different thermal expansion properties than brick masonry. Depending upon the metal used, its
thermal expansion coefficient may be 3 to 4 times that of brick masonry. The designer should be aware of this and
provide for this differential movement in the development of the details. Consideration must also be given to the
drying shrinkage of the element if cast-in-place concrete is the material selected.
If brick is the material chosen for the coping, it may be desirable in some applications to use a special shape to get
a positive slope in two directions. In most applications, the slope should be only in one direction, with drainage onto
the roof and not down the wall face. In such case, the coping can be built using regular shapes.
Design
The prime consideration in the design of caps and copings is the performance of the element in service. The
designer must take into consideration the movement of the element, differential movement between the element and
the wall, joint configuration and material, connection of the element to the wall, and type and location of flashings.
The esthetic value of the detail should be evaluated. As with details of other elements, selection of material, color,
texture and configuration will effect the esthetic value of the detail. The designer has a wide range from which to
choose, but he must keep in mind that the performance should not be compromised to achieve esthetic value.
The economic considerations are seldom a major consideration in the development of details for caps and copings.
The material selected may have a minor effect on the economics of the detail. It affects the economics not only by
its own costs, but also by the economics of installation. The economic considerations should not have a deleterious
effect on the performance of the details in service.
Details
General. The function of caps and copings is to prevent the entry of water into the wall where the wall becomes
partially or totally discontinuous vertically. Caps should have the top surface sloping downward, away from the face
of the wall above. Copings may slope in one or both directions. In all cases, the slope should be a minimum of 15
degrees from horizontal.
The caps should overhang the wall face on the exposed side. Copings should overhang the wall on both sides. The
overhang should be of sufficient dimension so that the inner lip of the drip is at least 1 in. (25 mm) from the face of
the wall. Since the function of caps and copings is to prevent moisture penetration, the fewer the number of joints,
the more assurance that the detail will perform its function.
Flashing and Weepholes. Flashings for caps and copings generally serve a different function from flashings used
elsewhere in the structure. Flashing used with caps and copings has as its prime function the prevention of the entry
of moisture into the wall. The collection and diversion of the water from the wall becomes a secondary, although
important function.
In order to properly anchor caps and copings to the wall, it may become necessary to penetrate the flashing with the
anchor, see Figs. 1, 3, and 4. To prevent moisture from entering the wall, at these points, it is absolutely necessary
that the penetrations be adequately sealed, or the flashing will fail to function as intended.
http://www.gobrick.com/BIA/technotes/t36a.htm
2 of 8 9/13/2009 1:39 PM
Precast Concrete or Stone Coping on Cavity Wall Parapet
FIG. 1
Coping for Cavity Wall Parapet
FIG. 2
http://www.gobrick.com/BIA/technotes/t36a.htm
3 of 8 9/13/2009 1:39 PM
Coping for Solid Masonry Parapet
FIG. 3
Rowlock Coping on Solid Masonry Parapet
FIG. 4
Flashings should be extended beyond the face of the wall and bent downward 1/4 in. (6 mm) to form a drip, as
shown in Figs. 2, 5, and 6. Metal copings may also serve as flashings. It should be recognized that exterior flashings
not contained within the wall serve the same functions as do interior flashing. Information on flashing materials is
provided in Technical Notes 7A Revised.
While the flashing for caps and copings may have a different prime function from normal usage, it is still necessary to
provide weepholes immediately above the flashings to convey the water collected on the flashing out of the wall,
unless exterior flashing is used. Weepholes should be spaced at a maximum of 24 in. (600 mm) o.c., unless wicks or
hidden flashing are used. Then the spacing should be reduced to 16 in. (400 mm) o.c. maximum.
Drips. Regardless of the material selected for caps or copings, drips should be provided. When brick caps and
copings are used, the drip is the lowest point on the element, as shown in Figs. 4 and 7. When metal caps and
copings are used, the drips can be formed by bending the material outward from the face of the wall, see Figs. 2, 5
and 6. With heavy gauge metals, stone concrete or terra cotta caps and copings, the drip is either cut or formed in
the bottom of the projection beyond the face of the wall, as shown in Figs. 1, 3, and 7. This drip can be in several
configurations, and still perform. The important thing is that a drip be provided and that the inner lip be at least 1 in.
(25 mm) from the face of the wall as shown in Figs. 1, 3, and 7.
http://www.gobrick.com/BIA/technotes/t36a.htm
4 of 8 9/13/2009 1:39 PM
Masonry Bearing Wall Coping
FIG. 5
Masonry Cavity Wall Coping
FIG. 6
Brick and Precast Concrete or Stone Caps
FIG. 7
Connections. Elements other than caps and copings require careful consideration of their connection to the
structure for the structural stability of the element. In the case of caps and copings, the structural stability becomes
secondary to the climatic considerations, such as moisture and temperature. Connections which are usually provided
http://www.gobrick.com/BIA/technotes/t36a.htm
5 of 8 9/13/2009 1:39 PM
for structural purposes are generally rigid. Because of the diversity of materials used for caps and copings in
conjunction with brick masonry walls, the connection in some cases should be of a flexible nature. Brick masonry,
concrete, stone and terra cotta, respond to climatic conditions in much the same manner, and rigid connections can
be used with little consideration of differential movement. Because of the dissimilarity of metal and brick masonry in
their reaction to climatic conditions, the connections require some flexibility.
Light gauge metal copings as shown in Figs. 5 and 6 should be nailed to the wall, and horizontal slots should be
provided at nailing locations to prevent buckling of the coping due to thermal expansion. Metal caps and copings
require an extension down the face of the wall, 4 in. (100 mm) min., and a sealant between the metal and the wall to
prevent wind uplift and water penetration. Care should be taken to seal each penetration of the metal cap or coping
where it is exposed to the exterior environment.
Expansion Joints. It is necessary to provide expansion joints in long walls to provide for movement of the wall due
to thermal and moisture expansion. This is particularly true in parapet walls and other masonry walls which are
exposed to the exterior climatic conditions on both sides. Expansion joints are discussed in Technical Notes 18
Series.
When expansion joints are required in the wall, the expansion joints should also be provided through any caps or
copings in the same locations. It may be necessary to provide additional joints in metal copings. Metal copings
should be so detailed and constructed that they function independently of the movement of the wall below. Expansion
joints should be of a compressible material, but should also be extensible. One method of providing expansion joints
is to leave the mortar from the head joints in a vertical line and insert a synthetic backer rod to the desired depth and
fill the remainder of the joint with a high-quality sealant.
Construction. Caps and copings require no special construction skills. If brick masonry is used as the cap or coping
material, great care should be taken to ensure that all head and bed joints are completely filled. If cast-in-place
concrete is used, some provision must be made to allow for the initial drying shrinkage of the concrete.
If precast concrete or stone are used for caps or copings, non-compressible shims should be placed on the top of
the wall at the exterior face of the wall. The shims are used because the weight of this type of cap or coping would
compress the plastic mortar and a smaller joint would result. Then the mortar for the bed joint is spread and the cap
or coping installed. The shims which should have a thickness equal to the bed joints should be left in place until the
mortar has set. Once the mortar has set, the shims should be removed and the joint tuckpointed.
CORBELS AND RACKING
General
Corbeling of brick masonry may be done to achieve the desired esthetics, or to provide structural support. There are
empirical requirements provided by most codes and standards for unreinforced corbels, as shown in Fig. 8. If these
requirements are to be exceeded, then the element will require a rational design as a reinforced element.
http://www.gobrick.com/BIA/technotes/t36a.htm
6 of 8 9/13/2009 1:39 PM
Limitations on Corbeling
FIG. 8
Corbels. The empirical approach requires that the total horizontal projection not exceed one-half the thickness of a
solid wall, or one-half the thickness of the veneer of a veneered wall. It is also required that the projection of a single
course not exceed one-half of the unit height or one-third of the unit bed depth, whichever is less. From these
limitations, the minimum slope of the corbeling can be established (angle measured from the horizontal to the face of
the corbeled surface is 63 deg 26 min. see Fig. 8). The required slope could be increased by the requirements that
the unit projection not exceed one-third of the bed depth if they are more restrictive. It should be pointed out that the
eccentricity induced into the wall by the corbeling must be considered in the wall design. If these limitations are
exceeded, the wall should be reinforced to resist the stresses developed by the corbeling.
Fig. 9 illustrates graphically the pattern of stresses within two corbels of different configurations under identical
loading conditions. The corbel on the left is 45 degrees from horizontal, which is not in accordance with building code
requirements. The corbeled wall on the right has an angle of corbel 60 degrees from horizontal and is very close to
the building code requirement of 63 degree 26 min discussed above. The 60 degree corbel shows a stress pattern
with axial and shear stresses with the only concentration of stresses directly below the applied load, P. The shear
stresses are well distributed within the wall section. The 45 degree corbel, on the other hand, has bending stresses
in addition to the axial and shear stresses, and the pattern of the stresses has been drastically altered. In addition to
the concentration of compressive stresses immediately beneath the load, P. there is another concentration of
compressive stress at the toe of the corbel. The bending stresses require that a corbel of this configuration be
rationally designed and reinforced. Those corbels having an angle from horizontal of 60 degree or greater do not
require reinforcement unless they exceed the other requirements given above.
http://www.gobrick.com/BIA/technotes/t36a.htm
7 of 8 9/13/2009 1:39 PM
Corbeling Stress Distribution
FIG. 9
Racking. When racking back to achieve the desired dimensions, care must be exercised to insure that, since there
is no limitation on the distance each unit may be racked, the cores of the units are not exposed. Preferred
construction consists of a setting bed over the racked face with the uncored brick or paving brick set to provide a
weather-resistant surface. Mortar washes may also be used. They may not, however, be as durable. When using a
mortar wash, it should not bridge over the rack, but should fill each step individually.
SUMMARY
The designer, when developing details for caps, copings, corbels and racking should keep in mind the function of the
element being detailed, the esthetic value he wishes to achieve, the structural stability of the element, and the
economics of construction. It is essential to provide details which allow the elements to perform their primary
functions as well as possible. In order to do this, the designer must select the proper materials, locate them in the
proper place and provide sufficient information so that the element can be properly constructed. Several decisions
and assumptions must be made by the designer because each project and each element on the project must be
satisfactorily addressed.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the technical staff of the Brick Industry Association. The information and recommendations contained
herein if followed with the use of good technical judgment, will avoid many of the problems discussed here. Final
decisions on the use of details and materials as discussed are not within the purview of the Brick Industry
Association, and must rest with the project designer, owner, or both.
http://www.gobrick.com/BIA/technotes/t36a.htm
8 of 8 9/13/2009 1:39 PM
Technical Notes 39 - Testing for Engineered Brick Masonry- Brick and Mortar
November 2001
Abstract: Testing of brick, mortar and grout is often required prior to and during construction of engineered brick
masonry projects. The tests involve a combination of laboratory and field procedures which are described in various
ASTM standards. The extent of testing is a decision made by the engineering or architectural firm responsible for the
masonry design, and may consist of only a few laboratory tests to determine the properties of the brick units, or may
involve extensive laboratory and field sampling and testing. This Technical Notes describes the testing of materials;
other issues in this series describe testing of brick masonry assemblages.
Key Words: brick, engineered brick masonry, grout, mortar, quality control, testing.

INTRODUCTION
The use of engineered brick masonry in the construction of loadbearing structures requires that the standard methods for
determining the physical properties of both the materials and the masonry assemblages be strictly followed. The
standards and specifications for engineered brick masonry are based, for the most part, on the results of American
Society for Testing and Materials (ASTM) methods of testing.
It is not the intent of this Technical Notes to supersede the various applicable ASTM standards, but to supplement them.
The ASTM standards have been carefully developed by balanced technical committees composed of people experienced
and knowledgeable in their chosen fields. Therefore, if the prescribed methods of tests are not adhered to, inaccurate
and inconsistent test data and erroneous conclusions can result. This can be quite serious when the design of a masonry
bearing wall structure is based on such tests, or when such tests are used as quality controls during construction.
This Technical Notes covers testing of masonry materials for obtaining information needed to determine design
properties for engineered brick masonry. Additional testing required for assessment of material compliance to various
ASTM specifications is not included. In addition, field testing of brick, mortar and grout for quality control is discussed.
This Technical Notes is the first in a series on testing. Other Technical Notes in this series discuss the construction,
preparation and testing of masonry assemblages (in the laboratory); and the sampling, preparation and handling of
jobsite test specimens for the purpose of quality control of the construction.

ENGINEERED BRICK MASONRY STANDARDS
There are several standards used in the United States for the design of brick masonry structures, all of which contain
some requirements for testing of masonry materials or assemblages. Likewise, other standards and building codes
require testing in order to establish various design parameters.
In addition to predesign and preconstruction testing, testing for the purpose of quality control is often implemented.
Building Code Requirements for Masonry Structures (ACI 530 / ASCE 5 / TMS 402-99) and Specification for Masonry
Structures (ACI 530.1/ ASCE 6 / TMS 602-99) [3], known as the MSJC Code and Specifications, contain several quality
assurance requirements. For example, the MSJC Code and Specification require that the initial rate of absorption (IRA)
of brick at the time of laying not exceed 1 gram per sq in. per min. ASTM C 62, ASTM C 216 and ASTM C 652 also
recommend that the limit on IRA be 30 g/30 in.
2
/min. The determination of this property may be made in the laboratory
on oven-dry brick, or at the construction site as a field test. The tests outlined within this Technical Notes are those
which are most commonly performed to satisfy the requirements of the MSJC Code and Specification.

TESTING STANDARDS
The ASTM standards which are most frequently utilized when testing brick masonry materials should be readily available
to all laboratory personnel, and to individuals involved in field testing. The applicable standards are as follows:
TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
1 of 8 9/13/2009 1:40 PM
Clay Masonry Units -ASTM C 67, Standard Test Methods of Sampling and Testing Brick and Structural Clay Tile
Mortar -ASTM C 270, Standard Specification for Mortar for Unit Masonry
-ASTM C 91, Standard Specification for Masonry Cement
-ASTM C 109, Standard Test Method for Compressive Strength of Hydraulic Cement Mortars
(Using 2-in. or 50-mm Cube Specimens)
-ASTM C 780, Standard Test Method for Preconstruction and Construction Evaluation of Mortars
for Plain and Reinforced Unit Masonry
Grout -ASTM C 476, Standard Specification for Grout for Masonry
-ASTM C 1019, Standard Test Method for Sampling and Testing Grout

For the most part, these standards provide clear and concise explanations of the procedures for sampling and testing
masonry materials; however, for the novice, some areas may present some confusion. The following sections will explain
some of the procedures required by the various ASTM standards.

BRICK TESTING FOR ENGINEERED BRICK MASONRY
The strength of brick varies considerably, depending on raw material, method of manufacture and degree of firing. The
range in compressive strength is on the order of 2000 psi to in excess of 20,000 psi. The MSJC Code does not dictate
minimum compressive strength requirements for brick, but since the allowable stresses and elastic moduli of masonry are
a function of compressive strength of brick, testing to determine compressive strength is required.
For the determination of unit compressive strength, f'
b
, the procedures given in ASTM C 67 [1] should be followed.
The initial rate of absorption (IRA) is another important property. If the IRA of brick exceeds an acceptable upper limit,
problems with excessive shrinkage of mortar and grout, and poor bond, are apt to occur. The procedures for
determining the IRA, in the laboratory and in the field, are contained in ASTM C 67.
Compressive Strength
Specimen Size. ASTM C 67 requires that the specimen be full height and width, and approximately one-half of a brick in
length, plus or minus 1 in. (25 mm). For example, an 8-in. (200 mm) long brick may be tested using a piece of brick with
a length between 3 and 5 in. (75 and 125 mm). However, if the testing machine being used is not capable of providing
sufficient force to crush the approximate half-brick, a piece of brick having a length of one-quarter of the original full brick
length may be used, so long as the total cross-sectional area is not less than 14 in.
2
(90 cm
2
), see Figure 1.

Compressive Strength Specimens
FIG. 1

TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
2 of 8 9/13/2009 1:40 PM
Although ASTM C 67 does not specifically state the method in which the samples are to be obtained, it has been
common practice to use pieces of brick which are left over from modulus of rupture tests. If modulus of rupture tests are
not being performed, then sawing the units to the desired size is acceptable. A minimum of five specimens is required.
The compressive strength test specimens should be oven-dried. The amount of moisture in the brick can affect its
compressive strength - the higher the moisture content, the lower the apparent strength. Therefore, by drying the
specimens before testing, one variable that can affect the results is eliminated. If they are wet-cut with a masonry saw,
the drying should follow the cutting. If a wet capping material, such as high-strength gypsum, is used, it is generally
agreed that the small amount of moisture absorbed by the specimens will not make additional oven drying necessary.
The 24-hr curing period in laboratory air will suffice.
Capping Specimens. The importance of careful capping procedures cannot be over-emphasized. Brick units, by their
inherent nature, are not perfectly formed and their bearing surfaces may not be parallel and free from surface
irregularities. The purpose of capping the bearing surfaces is to assure reasonably parallel and smooth opposite bearing
surfaces; thus reducing the likelihood of uneven bearing and stress concentrations, and the resulting premature failure of
the test specimen.
Laboratory technicians responsible for capping compressive test specimens should be thoroughly familiar with the
capping procedures prescribed in ASTM C 67. Poor caps, resulting from careless capping techniques, can result in
erratic test results and a lowering of the apparent compressive strengths of the specimens.
Placing Specimens in Testing Machine. The requirement in ASTM C 67 that the specimen be centered under the
spherical upper bearing block within 1/16 in. (1.6 mm) is not a capricious one. The introduction of an eccentric load, if the
specimen is not carefully centered, can result in a lower apparent compressive strength for the test specimen. It should
be understood, however, that this requirement assumes that the specimen is symmetrical about both horizontal axes or
its center of gravity. For symmetrical specimens, the center of gravity will be the geometrical center of the unit. Such is
not the case with unsymmetrical test specimens. Therefore, the centers of gravity of unsymmetrical specimens should
be determined and marked, and it is those marks that should be aligned with the center of the upper bearing block.
To determine the center of gravity for an unsymmetrical test specimen, a small steel rod, 1/8 in. to 1/4 in. (3 to 8 mm) in
diameter, may be used. The location of the center of gravity is determined by finding the balance point of the brick
specimen. To place the specimen over the rod in the exact position such that it balances perfectly is difficult, but a very
good estimate of this location is not hard to achieve.
Speed of Testing. The speed of testing specified in ASTM C 67 should be adhered to, primarily for the purpose of
obtaining consistent results. Past experience on the effect of the rate of loading on the compressive strength of
specimens has shown that, as the rate increases, there can be significant increases in the apparent compressive
strengths of the specimens. The requirements of ASTM C 67, while not particularly specific, do provide a moderate rate
of loading which, if followed, will produce consistent results that will represent more accurately the true compressive
strengths of the specimens.
ASTM C 67 specifies that the specimen should be loaded to one-half of the expected maximum load, and then the rate
should be adjusted such that the test is completed in not less than one minute and not more than two minutes. For this
reason, it is a good idea to do one or two preliminary tests to get an estimate of the maximum strength. Figure 2
illustrates the time vs. loading criteria of ASTM C 67.

Compressive Test Loading Rate
FIG. 2

Determination of Minimum Net Area (Percent Voids). There are two reasons for determining the void area of brick:
TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
3 of 8 9/13/2009 1:40 PM
the first reason is to obtain the percentage of voids (percentage coring) in order to assess whether the brick will be
classified as solid or hollow brick; the second reason is to obtain the average net cross-sectional area for determination
of net area compressive strength of the units. ASTM C 216 and C 62 for solid brick, and ASTM C 652 for hollow brick
require calculation of gross area compressive strength.
If the net area compressive strength is required, the section "Measurement of Void Area in Cored Units" in ASTM C 67
should be followed. To perform these measurements, a sample of ten brick is specified by ASTM C 67. Following the
procedure in this section, the cores are filled with sand. The sand is then placed in a graduated cylinder to determine the
volume. Using the equation given in ASTM C 67 for percent void area, the void area can be determined. The net area
can be determined by subtracting the void area from the gross area.
Calculation and Report. The compressive strength is determined by dividing the maximum compressive load by the
gross cross-sectional area of the specimen. If the net area compressive strength is required, the net area, as
determined in the previous section, must be used to obtain the desired results. Since five specimens are used, the
arithmetic average should be determined.
Initial Rate of Absorption
The initial rate of absorption (IRA) is an important property of brick because it affects mortar and grout bond. Brick IRA
and mortar retentivity should be considered when selecting brick and mortar type. If the initial rate of absorption is over
1 gram per minute per in
2
, brick will absorb moisture from the mortar or grout at a rapid rate, and may impair the
strength and extent of the bond. Thus, determining the IRA is important.
In the laboratory, the IRA is measured using brick which are oven-dried to equilibrium. The IRA of a dry brick is apt to be
higher than one which contains some moisture. The field test for initial rate of absorption is performed on brick in their
field condition, i.e., no attempt is made to dry the units. The laboratory test will give an idea of the order of magnitude of
the IRA and the field test can be used to determine if additional wetting is necessary.
Laboratory Procedure. As previously mentioned, the laboratory procedure is performed on oven-dried specimens. Five
full-size specimens are required. The technician performing the test should be aware that the larger the tray size, the
less effect the absorption has on the water level. ASTM C 67 requires a tray with a cross-sectional area of at least 300
in.
2
(1935.5 cm
2
). For a brick with an IRA of 40 g/min/30 in.
2
, the water level would drop less than 1/100 in., which is
hardly measurable. Nevertheless, ASTM C 67 provides recommendations on maintaining the water level. Figure 3
illustrates the tray with a brick positioned for testing. The method is relatively straightforward and easy to perform. The
results are reported in grams of water gained per 30 sq in. when the brick are immersed in 1/8 in. (3 mm) of water for 1
min. The calculation of IRA is as follows:
IRA = 30 W / LB (Eq. 5)
where: W = actual gain in weight of specimen in grams,
L = length of specimen, in in. and
B = width of specimen, in in.

What some laboratory technicians fail to realize, however, is that the above equation is for specimens that are not cored.
If the test specimens are cored brick, or are non-prismatic, the net area must be substituted for LB in Eq. 5.

Determination of IRA
FIG. 3

Field Procedures. Brick units on the jobsite may have a different rate of absorption than that of the same units tested
for IRA in the laboratory. The IRA may be lower due to moisture which brick absorb after leaving the manufacturing
TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
4 of 8 9/13/2009 1:40 PM
plant. Two tests are available for field determination of brick absorption. One is an ASTM procedure, described in
ASTM C 67, which measures quantitatively the absorption rate. The other is an approximate, but effective, test which is
not covered by an ASTM standard, and yields a qualitative indication of the bricks' absorption rate and necessity for
wetting prior to use.
ASTM Field Method for IRA - This method is described in detail in ASTM C 67, and is accomplished through volumetric
means rather than by weight measurements. Using this method, the brick are placed in a pan of water for 1 min.
removed and the quantity of water remaining in the pan is measured using a pycnometer (Fig. 4). The pycnometer is
used to measure the initial quantity of water to be placed in the pan. The difference in the original amount of water and
the quantity remaining after placement of the brick into the pan for 1 min is the amount absorbed by the brick. It is very
important to use the correct size pan and to wet and drain the pan prior to testing.

Pycnometer
FIG. 4

Test for Wetting Brick - The following test is useful for determining the necessity of wetting brick prior to use:
A circle, approximately 1 in. (25 mm) in diameter, is drawn on the bed surface of the brick, using a wax pencil and a
twenty-five-cent coin as a guide. Twenty drops of water are placed into the circle using an eyedropper. If, after 90
seconds, all of the water has been absorbed, wetting the brick prior to placement is recommended.

MORTAR TESTING FOR ENGINEERED BRICK MASONRY
Technical Notes 8 Series discusses the various types of mortar, properties and mix designs. Also, ASTM C 270,
Specification for Mortar for Unit Masonry, gives both prescriptive and performance requirements for mortar. Another
standard specification for mortar, BIA M1-88, provides recommendations on selection proportions and test requirements
of portland cement-lime mortars. This section will outline the various mortar tests which are important when designing
and building engineered brick masonry elements.
Laboratory Testing of Mortar
Laboratory testing of mortar is performed in accordance with ASTM C 270 [2]and other standards referenced in ASTM C
270. The tests are performed on mortar samples which are prepared in the laboratory. ASTM C 270 is not a
specification to determine mortar strength and properties through field testing. The amount of testing required by ASTM
C 270 depends on the method in which the mortar is specified, i.e., proportion or property specification. If the mortar is
specified by the proportion specifications, there are no testing requirements for mortar. For mortars specified by the
property specifications, water retention, compressive strength and air content tests must be performed.
The following sections describe the methods of tests for mortar which are specified by the property specifications.
Water Retention. ASTM C 270 refers to the procedures of ASTM C 91 for water retention determination, except that
the laboratory-mixed mortar shall be of the same materials and proportions to be used in the construction. Since the
water content of mortar used on the jobsite varies somewhat, and is not a specified quantity, the laboratory technician
should proportion the cementitious materials and sand in accordance with the job specification and add sufficient water to
bring the flow up to 110 +/- 5%.
TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
5 of 8 9/13/2009 1:40 PM
To perform the water retention tests, the technician should review ASTM C 91 on Water Retention, ASTM C 305 on
Mechanical Mixing, and ASTM C 109 on Performing Flow Tests. The flow test apparatus must meet the specifications of
ASTM C 230. The chart in Fig. 5 indicates the ASTM standards relative to water retention testing of mortar specified by
the property specifications.

Related ASTM Standards for Property Specifications
FIG. 5

Compressive Strength. Compressive strength testing of laboratory-prepared mortar is required under the ASTM C 270
property specifications. To determine compressive strength, samples are to have the same proportions as in the actual
construction. As with the water retention test, the amount of water to be used is not clearly stated; therefore, it is
recommended that sufficient water be used to bring the flow to 110 +/- 5%. As shown in Fig. 5, other associated ASTM
standards which must be used are ASTM C 109, C 305 and C 230.
The technician should become familiar with the procedures of ASTM C 109 for specimen molding and load application
since these procedures must be followed closely in order to obtain reliable results.
Air Content. Air content determination is the third and last property which must be assessed for mortars specified under
the property specifications. The air content is determined using a weight-volume relationship to determine the absolute
volume of solids and water. ASTM C 91 and ASTM C 185 are used to determine air content, except that the equation
for percent air content is given in ASTM C 270. The equation for air-free mortar density is:

(Eq. 2)

and the volume of air in percent is
TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
6 of 8 9/13/2009 1:40 PM
(Eq. 3)
where: W1 = weight of portland cement, g,
W
2
= weight of hydrated lime, g,
W
3
= weight of masonry cement (if used), g,
W
4
= weight of sand, g,
V
w
= volume of water used, mL,
P
1
= unit weight of air-free portland cement, g/cm
3
,
P
2
= unit weight of air-free hydrated lime, g/cm
3
,
P
3
= unit weight of air-free masonry cement (if used), g/cm
3
,
P
4
= unit weight of air-free sand, g/cm
3
,
W
m
= weight of 400 mL of mortar, g.

The air-free unit weights of the various materials in Eq. 2 are equal to the specific gravity of the material times the unit
weight of water (which is unity); thus, the unit weight is numerically equal to the specific gravity. The specific gravity for
the various materials should be obtained from the manufacturers or determined by testing. Table 1 gives the
approximate specific gravities for several mortar materials

In performing the air-content tests, it is very important to weigh and measure the quantities accurately, since errors in
weights and volumes would have significant impact upon the calculated air content.
Field Testing of Mortar
For purposes of quality control, field testing of mortar is sometimes required. Field testing should not be confused with
laboratory testing, or be performed using the standards and procedures for laboratory testing of mortar. The
appropriate standard for this type of testing is ASTM C 780 "Standard Test Method for Preconstruction and Construction
Evaluation of Mortars for Plain and Reinforced Unit Masonry". The main purposes of field testing is to ensure that mortar
is proportioned properly by the mixer operator, and to obtain an indication of variability or change in constituent materials,
quality and performance.
There are several tests which are covered in ASTM C 780, not all of which are required. Eight tests are outlined in the
Annexes of ASTM C 780 which are: A1) Consistency by Cone Penetration Test Method, A2) Consistency Retention of
Mortars for Unit Masonry, A3) Initial Consistency and Consistency Retention or Board Life of Masonry Mortars Using a
Modified Concrete Penetrometer, A4) Mortar Aggregate Ratio Test Method, A5) Water Content Test Method, A6)
Mortar Air Content Test Method, A7) Compressive Strength of Molded Masonry Mortar Cylinders and Cubes and A8)
Splitting Tensile Strength of Molded Masonry Mortar Cylinders.
The testing agency and the specifier should be aware that the compressive strength of mortar, as determined by field
testing, does not have to meet the minimum compressive strength requirements of ASTM C 270.
The specifier must decide which of the eight tests is to be performed, then preconstruction testing of the materials can be
performed in order to establish requirements for construction site-sampled mortar.
A complete discussion of the test procedures of ASTM C 780 is not within the scope of this Technical Notes; therefore,
the technician in charge of performing the tests should become thoroughly knowledgeable with ASTM C 780 and its
TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
7 of 8 9/13/2009 1:40 PM
referenced documents.

GROUT TESTING FOR ENGINEERED BRICK MASONRY
The specification for grout for engineered brick masonry, ASTM C 476, does not require any laboratory testing.
Experience with grout mixed in accordance with the provisions of ASTM C 476 has been extremely favorable, and grout,
therefore, does not require extensive testing if mixed with the materials and in the proportions stipulated by the standard.
There is a relatively new standard for both field and laboratory sampling and compressive testing of grout used in
masonry construction, entitled ASTM C 1019, "Standard Test Method for Sampling and Testing Grout".
Sampling and Testing Grout for Engineered Brick Masonry
According to ASTM C 1019, the use of the standard may be to select grout proportions by comparing test values or as a
quality control test for uniformity of grout preparation during construction. The standard specification for grout, ASTM C
476, does not contain provisions for mixing grout to property specifications; therefore, the use of ASTM C 1019, at this
time, for grout mix design is not advised. For purposes of quality assurance, the grout testing standard may be useful.
The specimens are prepared by using masonry units as forms (Fig. 6). The masonry units are those which are to be
used in the project under construction or to be constructed. The laboratory technician may find it strange to use the brick
units as forms, but the reason is to simulate the conditions of the grout after placement into the brick masonry element.
Grout is placed with a high water/cement ratio, slump of 10 to 11 in. (250 mm to 275 mm), in order to facilitate
consolidation and eliminate voids. Due to the absorptive nature of the masonry, the water content of the grout is reduced
after placement.
The methods of sampling and testing, as described in ASTM C 1019, are easily accomplished; therefore, additional
description and explanation will not be given in this Technical Notes.

Grout Mold Using 2 in. (152.4 mm)
High Standard Size Brick
FIG. 6

SUMMARY
This Technical Notes has discussed testing of brick, mortar and grout used in engineered brick masonry. Most
laboratory and field tests are covered by ASTM standards. Testing agencies using these tests should be fully aware of
the procedures and limitations, so that improper application and erroneous results are avoided.

The information and suggestions contained in this Technical Notes are based on the available data and the experience of
the engineering staff of the Brick Industry Association. The information contained herein must be used in conjunction with
good technical judgment and a basic understanding of the properties of brick masonry. Final decisions on the use of the
information contained in this Technical Notes are not within the purview of the Brick Industry Association and must rest
with the project architect, engineer and owner.

TN 39 Revised - TESTING FOR EGNINEERED BRICK MASONRY-B... http://www.gobrick.com/BIA/technotes/t39.htm
8 of 8 9/13/2009 1:40 PM

Technical Notes 39A - Testing for Engineered Brick Masonry - Determination of Allowable Design Stresses
[July/Aug. 1981] (Reissued December 1987)
INTRODUCTION

Prior to the development of a rational design procedure for brick masonry, it was sufficient to know that brick
masonry units and mortar used were in compliance with the standards outlined in Technical Notes 39 Revised,
"Testing for Engineered Brick Masonry - Brick, Mortar and Grout." These quality control tests provided assurance
that the same quality of materials were being used throughout the building project. This did not give any assurance
or knowledge as to the actual performance of the masonry in the wall.
With the development of a rational design method, it became important that the architect and/or engineer have
knowledge of the expected performance of the brick and mortar, not as individual parts of the wall, but as the total
wall system. With this need in mind, this Technical Notes outlines several ASTM Standard Methods of Tests for
Masonry Assemblages which will give the architect and/or engineer the ability to predict in-the-wall performance of
masonry and determine allowable design stresses.
It is essential in all of the tests described in this Technical Notes that the units, mortar and construction of the
assemblage be nearly identical with the materials and methods to be used in the actual construction process. Only in
this way can the actual performance of the masonry be accurately predicted.
This Technical Notes will cover ASTM standards for the determination of all necessary design stresses for brick
masonry as specified in the design standard, Building Code Requirements for Engineered Brick Masonry, BIA,
August 1969, and the model building codes in present-day usage. It will also stipulate the revisions necessary to
determine the same properties for hollow brick units. Subsequent issues of Technical Notes will discuss
miscellaneous tests for masonry not to be used for design stress determinations. These tests will be used primarily
for quality control, material comparability and in-the-wall performance predictions for properties other than strength.

STANDARD METHODS OF TESTS

The ASTM test standards with which this Technical Notes is concerned are contained in the Annual Book of ASTM
Standards. The methods of tests described in the ASTM standards and listed below should be strictly adhered to;
otherwise, the test performed is no longer a standard test and erroneous or misleading results may be obtained.
The applicable ASTM standards for masonry are as follows:
Compressive Strength and Modulus of Elasticity:
Test Methods for Compressive Strength of Masonry Prisms, ASTM Designation E 447.
Diagonal Tension (Shear) and Modulus of Rigidity:
Test Method for Diagonal Tension (Shears in Masonry Assemblages, ASTM Designation E 519.
Method for Conducting Strength Tests of Panels for Building Construction, ASTM Designation E 72.
39a http://www.gobrick.com/BIA/technotes/t39a.htm
1 of 18 9/13/2009 2:04 PM
Flexural Tensile Strength:
Method for Conducting Strength Tests of Panels for Building Construction, ASTM Designation E 72.
In addition to the above listed standards, the brick and mortar used should conform to the standards listed in
Technical Notes 39 Revised.

PURPOSE AND APPLICATION OF TESTS
General. If a rational design approach for masonry is employed, it is essential to establish allowable design
stresses early in the design process. Present design standard requirements provide two methods to establish these
values. Under these requirements, the ultimate compressive strength (f'm) may be determined by (a) prism test, or
(b) an approximation based upon brick strength and mortar properties. The prism test method is the preferred
method as it provides the designer with more exact information; whereas the approximation method, of necessity,
provides more conservative values.

Allowable design stresses may be determined under the design standard, once an ultimate compressive strength of
masonry has been determined. However, in some cases, it may be desirable, or necessary, to establish tensile or
shear strength to closer tolerances than is obtained by design standard values usually given as a function of ultimate
compressive strength or of brick strengths and mortar types. Such conditions may occur when masonry is to be
used in prefabricated panels, is to be subjected to unusual loading conditions or when high, early strengths are
desirable.
All masonry specimens for establishing design stresses should be built using "inspected workmanship;'' that is to
say, all head, bed and collar joints should be completely filled (see Technical Notes 7B Revised for proper
procedures). Once strength tests have been performed and masonry properties established, it is necessary to
decide if, indeed, inspected workmanship can be achieved at the jobsite. If inspected, workmanship will be achieved,
the ultimate stress, f'm, need not be reduced. If, however, uninspected workmanship is expected, the ultimate
strength for "uninspected workmanship" must be used. This value is based upon inspected workmanship in which the
ultimate strength is reduced by 33 1/3, percent in the design standard.

Test methods for determining strengths and other properties of masonry necessary to establish allowable design
stresses are outlined below.

DETERMINATION OF f'm AND Em
The ultimate compressive strength (f'm) of masonry may be approximated if the brick to be used have been tested in
accordance with ASTM C 67 (see Technical Notes 39 Revised) and mortar type has been established. These
values of f'm are given in tabular format in the design standard for both types of workmanship. Values from this table
for a known brick strength, mortar type and workmanship classification may be used directly in determination of f'm,
and thus the allowable design stresses.

The ultimate compression strength of masonry is best determined by testing of compressive prisms in accordance
with ASTM Standard Test Methods for Compressive Strength of Masonry Prisms, E 447. There are two methods of
performing this test. Method A, which is used primarily for strength comparisons of different brick and mortars, could
be used in the selection process to determine what unit or mortar to use. For determination of ultimate compressive
strength of a specific brick and mortar for a specific project, Method B of E 447 should be used. This Technical
Notes will concern itself only with Test Method B.
39a http://www.gobrick.com/BIA/technotes/t39a.htm
2 of 18 9/13/2009 2:04 PM
The test specimens for Method B shall be built to conform as nearly as possible with the actual wall they represent.
They should have the same thickness as the wall represented; that is, if the wall is to be a solid wall of two wythes
with filled collar joint, the specimens should be the same. Use the same joint dimensions and bonding pattern as the
wall. The length of the specimen should be equal to, or greater than, the thickness. It has been general practice to
construct specimen lengths equal to 1 1/2 times the thickness. The specimen height should be at least twice the
thickness of the specimen or a minimum of 15 in. (3.81 cm). The height generally should not exceed five times the
thickness. See Fig. 1 for prisms with h/t from two to five. The height of specimens may be controlled by the testing
facilities available. Not all laboratories have testing machines of dimensions which will permit specimens of height-
to-thickness of five to be placed in the machine for test. Prior to construction of specimens, a check of facilities
available should be made and the specimens constructed with the greatest height-to-thickness ratio which the
available machine can accommodate. Correction factors for different height-to-thickness ratios and reasons for them
are discussed elsewhere in this Technical Notes. A minimum of three specimens representing each wall type should
be constructed and tested. Less than three tests will not give a representative sample. Five specimens of each type
of wall are desirable and will give the designer a more accurate ultimate compressive strength value on which he can
base allowable stresses.


Compressive Prisms - Slenderness Ratios Two Through Five
FIG. 1

All compressive strength specimens on which design stresses are to be based shall be tested at 28 days. The use
of 7-day tests for quality control during construction will be discussed in a subsequent Technical Notes. Seven-day
tests should not be used for determination of design stresses. However, if for some reason only 7-day test results
are available, an approximation of the 28-day strength may be made. The estimated 28-day strength can be
obtained by dividing the 7-day strength by 0.90.
When prisms with height-to-thickness ratios of less than five are used for design determinations, a reduction factor
must be used to determine the ultimate compressive strength of the masonry. Research experience indicates that
the mode of failure of masonry walls under compressive loading is by vertical tensile splitting. Therefore, to
accurately predict wall strength, the prism failures should be similar. Laboratory studies also show that masonry
specimens having slenderness ratios (h/t) of five or greater consistently fail in compression by the mode of vertical
tensile splitting and shorter prisms do not. Therefore, a slenderness ratio of five was selected as unity, and lesser
slenderness ratio results must be corrected by the factors shown in Table 1. Research has shown a definite
relationship between ultimate compressive strengths of prisms ranging from a slenderness ratio of two up to five.
Table 1 is based upon this relationship.

39a http://www.gobrick.com/BIA/technotes/t39a.htm
3 of 18 9/13/2009 2:04 PM

a
Height to thickness.
b
Interpolate to obtain intermediate values.


Solid Brick Compressive Prism - Tensile Splitting Failure
FIG. 2



39a http://www.gobrick.com/BIA/technotes/t39a.htm
4 of 18 9/13/2009 2:04 PM

Hollow Brick Compressive Prism - Tensile Splitting Failure
FIG. 3



The h/t of the specimens shall be determined by dividing the actual measured height of the specimen by the actual
measured thickness of the specimen. These specimen dimensions shall be determined in accordance with paragraph
6.2 of ASTM E 447.
The cross-sectional area shall also be determined based upon actual dimensions of the specimen in accordance with
paragraph 6.2 of ASTM E 447. The cross-sectional area to be used for determination of ultimate compressive
strength shall be the specimen thickness times the specimen length. For masonry units in accordance with ASTM C
62 and C 216, the gross cross-sectional area shall be used (t x l). If units are hollow brick (ASTM C 652), the net
cross-sectional area must be used for determination of ultimate compressive strength. The net cross-sectional area
shall be determined as follows: the actual gross cross-sectional area (t x l), using measured dimensions, less the
area of voids in the total cross section as measured or determined as outlined in Technical Notes 39 Revised.
If the coefficient of variation (v) of the test results on the specimens exceeds 10 percent, the ultimate compressive
strength to be used must be modified. This should not be confused with the 12 percent coefficient of variation
requirement for the test samples of individual units as covered in Technical Notes 39 Revised. If less than 10
percent, the average of the specimen tests should be used for (f'm) ultimate compressive strength. When the
coefficient of variation exceeds 10 percent, modify the average compressive strength of the specimens by the
following equation to obtain f'm:

39a http://www.gobrick.com/BIA/technotes/t39a.htm
5 of 18 9/13/2009 2:04 PM

where:
f'm = ultimate design compressive strength, psi (Mpa)

v = coefficient of variation of the specimen samples tested, percent
_
X = average compressive strength of all specimens, psi (Mpa)

The test report should include the average compressive strength, the standard deviation and the coefficient of
variation. If this information is not included, they may be calculated as follows:


where:

X = compressive strength of individual specimen, psi (Mpa)
Xt = total of all individual specimen compressive strengths, psi (Mpa)

n = number of specimens
s = standard deviation, psi (Mpa)
v = coefficient of variation, percent
In many instances, it is desirable or necessary to know the modulus of elasticity, Em, of the masonry being used.
The modulus of elasticity of the masonry can be determined by instrumentation of the specimens to be tested for the
determination of ultimate compressive strength. General practice for obtaining the strain of masonry in compression
requires the installation of strain gages on compressive prisms. These strain gages, having equal gage lengths, are
installed on each end of the prism along the neutral axis of the section (see Fig. 4). It is necessary in the case of
multiple wythe wall constructions and/or multiple wythes of dissimilar materials to determine the neutral axis prior to
loading as the load should also be applied at the neutral axis. The gage lengths should be as long as practicable.
39a http://www.gobrick.com/BIA/technotes/t39a.htm
6 of 18 9/13/2009 2:04 PM
Dial strain gages during test should be read at predetermined load levels up to approximately 75 to 80 percent of the
anticipated ultimate load and then removed to prevent damage to the gages at specimen failure. The strain in the
masonry is determined by averaging the strain gage readings and dividing by the gage lengths as given by the
formula:


where:

g = strain, average over entire section, in./in. (mm/mm)
D V1 = dial reading gage No. 1, in. (mm)
D V2 = dial reading gage No. 2, in. (mm)

g = vertical gage length in. (mm)




Compressive Prism Instrumentation for Modulus of Elasticity
FIG. 4

39a http://www.gobrick.com/BIA/technotes/t39a.htm
7 of 18 9/13/2009 2:04 PM


Once the strains at the various load levels, determined by formula (5) are obtained and stresses are calculated, a
stress-strain curve for the specimen should be plotted (see Fig. 5). There are several methods of determining the
modulus of elasticity from the stress-strain curve. The most common for masonry are the initial tangent modulus and
secant modulus methods. The modulus of elasticity is the slope of the tangent or the secant of the curve. The secant
modulus is most commonly used for masonry and is easier to determine. The two points selected on the stress-
strain curve are generally at 0 psi (Mpa) and 250 psi (1.72 Mpa) stress levels and the modulus is calculated as
follows:


where:
Em = secant modulus of elasticity, psi (Mpa)
fm0 = 0 psi (Mpa) stress
fm250 = 250 psi (1.72 Mpa) stress
g0 = strain at 0 psi stress, in./in. (mm/mm)
g250 = strain at 250 psi stress, in./in. (mm/mm)
In addition to the determination of the modulus of elasticity by actual tests, the modulus may be based upon f'm of
the compressive prism tests stated as a function of f'm. (See Tables 3 and 4 of the design standard. )

39a http://www.gobrick.com/BIA/technotes/t39a.htm
8 of 18 9/13/2009 2:04 PM

Idealized Stress Strain Curve
FIG. 5


DETERMINATION OF f'V AND EV (Vm AND G)

There are two methods of test provided in ASTM standards for the determination of the shear strength of masonry.
Shear or diagonal tensile strength is of considerable concern to structural designers, especially in geographical
areas where seismic design is required. Until recently, ASTM standards provided only one method of test for
determining shear strength. The method of test is described in ASTM E 72, Method for Conducting Strength Tests of
Panels for Building Construction. This method of test, referred to as the racking load test in the standard, has been
supplemented for masonry by ASTM E 519, Standard Test Method for Diagonal Tension (Shear) in Masonry
Assemblages. The E 72 racking load test provides for testing materials and constructions of all types, while E 519
applies only to masonry.
It has long been recognized that the method of test provided for in E 72 introduces compressive stresses into the
test specimen at the tie down which cannot be measured. See Figs. 6 and 7 for the testing apparatus used for this
39a http://www.gobrick.com/BIA/technotes/t39a.htm
9 of 18 9/13/2009 2:04 PM
test and method of failure. The tie down is required to prevent rotation of the specimen when load is applied. In
addition to the uncertainty of the tie-down stresses, this method of test requires a specimen 8 ft by 8 ft (2.438 m x
2.438 m) in size. This method of test generally is available only in large laboratories active in masonry research. On
the other hand, E 519 provides a method of test which is easier to perform and provides very reliable data. The
smaller specimens, 4 ft by 4 ft (1.219 m x 1.219 m), plus more simplified equipment place this method of test within
the capabilities of many private testing facilities. See Figs. 8 and 9 for test setup and loading shoes required for this
test.
The specimens for both E 72 and E 519 should be constructed using the brick, mortar, bonding pattern and wall
thickness that will be utilized in the construction. These specimens should be constructed using "inspected
workmanship" as previously described. Specimens for both tests should be cured for 28 days prior to testing.


Racking Test Frame and Specimen
FIG. 6




Racking Test Frame and Specimen After Testing
39a http://www.gobrick.com/BIA/technotes/t39a.htm
10 of 18 9/13/2009 2:04 PM
FIG. 7




Diagonal Tension Test Instrumentation for Modulus of Rigidity
FIG. 8




Diagonal Tension Loading Shoe
FIG. 9

39a http://www.gobrick.com/BIA/technotes/t39a.htm
11 of 18 9/13/2009 2:04 PM

The E 72 method of test calls for three 8-ft by 8-ft (2.438 m x 2.438 m) specimens. The panel can be instrumented
as shown in the standard and the horizontal deflection plotted against the load applied in graph form as described in
the standard. It has been common practice within the masonry industry to slightly modify this test. In lieu of the
instrumentation shown in the standard, a series of strain gages are placed to measure horizontal displacement of the
panel under load. These gages are placed along the vertical face of the panel where tie downs and load devices do
not occur. The horizontal displacement or strain is then taken at various load levels. The strain is the calculated
average of all dial readings at a particular load. The shear stress is calculated by dividing the horizontally applied
load by the panel width times the panel thickness. From these data stress-strain curves may be plotted. The
instrumentation should be removed at approximately 75 to 80 percent of the calculated load and the specimen tested
to failure. Data pertinent to the determination of allowable design stresses are (f'v) ultimate shear stress, a plotted
stress-strain curve and the modulus of rigidity at predetermined stress levels, usually 20 percent and 50 percent of
ultimate shear stress.

The E 519 method of test also specifies three specimens. Instrumentation of the specimens is provided along the
vertical and horizontal diagonals, as shown in Fig. 8. The vertical diagonal instrumentation measures the shortening
along that diagonal. The horizontal instrumentation measures the lengthening along that diagonal.
The calculations for shear stress for specimens constructed of solid units shall be based on gross area, while the
shear stress for hollow unit specimens shall be based on net area. The shear stress shall be calculated as follows:


where:
S
s
= shear stress on gross or net area, psi (Mpa)

P = applied load, lb (N)
A = average of the gross or net areas of the two contiguous upper sides of the
specimen, sq in. (mm
2
)

Formula (8) shall be used when specimens are built of solid units and formula (8a) shall be used for specimens of
hollow units.


where:
39a http://www.gobrick.com/BIA/technotes/t39a.htm
12 of 18 9/13/2009 2:04 PM
(t x l)1 . . . (t x l)2 =thickness and length or gross area
of the two upper contiguous sides of the specimen, sq
in. (mm
2
)
(t x l) = thickness and length or gross area of upper
side of specimen built of hollow units, sq in. (mm
2
)
Av = area of voids of the upper side of specimen built
of hollow units, sq in. (mm
2
)

The shear strain shall be calculated as follows:


where:

g = shearing strain, in./in. (mm/mm)
DV = vertical shortening, in. (mm)
DH = horizontal lengthening, in. (mm)

g = vertical gage length, in. (mm)
DH must be based on the same gage length as DV.

The modulus of rigidity shall be calculated as follows:


where:
Ev = G = modulus of rigidity, psi (Mpa)

The modulus of rigidity is calculated for predetermined stress levels, usually at approximately 20 percent and 50
39a http://www.gobrick.com/BIA/technotes/t39a.htm
13 of 18 9/13/2009 2:04 PM
percent of ultimate load.
The allowable shear stress should be determined by dividing the ultimate shear strength of the specimens by a
safety factor selected by the designer, when E 72 or E 519 are used to determine allowable design values. The
safety factor used should be based upon the designer's experience, type of workmanship expected, type of loading
the masonry will be subjected to, or as recommended in the design standard or Recommended Practice for
Engineered Brick Masonry.

DETERMINATION OF f't

At present, only one method of test is available in ASTM standards for determining the ultimate and design flexural
tensile strengths for masonry. This method of test is covered in ASTM E 72. Recently ASTM adopted E 518,
Standard Test Methods for Flexural Bond Strength of Masonry. This test, however, is to be used only as a
compatibility test for brick and mortar or as a quality control test and should not be used for determination of flexural
or transverse design stresses. ASTM E 518 will be more fully discussed in a subsequent issue of Technical Notes
39 series.
ASTM E 72 provides four methods for testing the large scale panels. The specimen may be tested in either a
horizontal (Fig. 10) or a vertical position (Fig. 11). In addition to these methods, the orientation of the masonry panel
itself within the loading frame will have great effect on the results obtained. If the span is normal to the bed joints,
simulating a wall supported by floor and roof framing in normal construction, the ultimate strengths obtained will be
considerably less than those with spans parallel to the bed joints. The panel oriented with a span parallel to bed
joints simulates a wall which in normal masonry construction is laterally supported by columns or pilasters.


Transverse Test - Horizontal Uniform Loading (Air Bag)
FIG. 10

39a http://www.gobrick.com/BIA/technotes/t39a.htm
14 of 18 9/13/2009 2:04 PM

Transverse Test - Vertical Uniform Loading (Air Bag)
FIG. 11



The specimens for this method of test should be at least 4 ft by 8 ft (1.219 m x 2.438 m) and the same thickness as
the proposed project walls. Three specimens are required for this method of test. The specimens should be built
using the type of brick, mortar and bonding pattern proposed for the construction project. The specimens should be
built using inspected workmanship as described earlier.
The test procedure is as follows: The specimen once placed in the test frame, which usually has a span of 6 in.
(152.5 mm) less than the specimen size, is instrumented only to measure the center of span deflection. The
concentrated load method, whether the specimen is vertical or horizontal, applies two equal loads at a distance of
one quarter of the span length from each support.
The uniform loading is applied in either position, using an air bag. See Figs. 10 and 11. The loads should be applied
in increments with deflection readings taken and recorded at each increment. Instrumentation should be removed at
approximately 75 to 80 percent of anticipated ultimate load to prevent damage to the instrument at failure.
The report of the test should provide a stress-deflection curve and ultimate transverse strength. The transverse
stress at ultimate may be calculated as follows:


where:
f't = ultimate transverse stress, psi (Mpa)
39a http://www.gobrick.com/BIA/technotes/t39a.htm
15 of 18 9/13/2009 2:04 PM

M = bending moment for 1-ft (305 mm) wide strip in.-lb (N-m)
S = section modulus of the specimen for 1-ft (305 mm) wide strip, in.3 (mm3)

The moment for uniformly loaded specimens is calculated as follows:

where:w = uniform load, psf (Mpa)

l = span length, ft (mm)
The moment for the concentrated loading may be calculated as follows:


where:
P = concentrated loads at the quarter points for 1-ft (305 mm) wide strip, lb (N)
The section modulus for the specimen must take into account whether the masonry units are solid (up to 25 percent
cored) or hollow (26 to 40 percent cored). Calculations for the section modulus of solid units would be as follows:


where:
b = width of 1-ft (305 mm) wide strip, in. (mm)
d = thickness of specimen, in. (mm)
The calculation for the section modulus of hollow units becomes somewhat more complicated and is as follows:

39a http://www.gobrick.com/BIA/technotes/t39a.htm
16 of 18 9/13/2009 2:04 PM

where:
b1 . . . bn = width of cores, in. (mm), see Fig. 12
d1 . . . dn = depth of cores, in. (mm), see Fig. 12

The illustration for this calculation method, Fig. 12, is based on a 3-core unit with nominal length of 12 in. (305 mm).
However, the formula can be adapted to fit other sizes of units and coring patterns. All dimensions shall be actual
dimensions.
Once the ultimate transverse strength has been determined by the test of three specimens, the designer should
select a safety factor to apply to arrive at an allowable design stress. This safety factor should be based upon the
designer's experience, type of workmanship expected, the in situ loadings expected or as recommended by the
design standard.


Typical Hollow Brick Cross Section
FIG. 12


CONCLUSION

39a http://www.gobrick.com/BIA/technotes/t39a.htm
17 of 18 9/13/2009 2:04 PM
The methods of tests described in this Technical Notes provide the design professional with methods to determine
allowable design stresses which may be used in the rational design of brick masonry. The values derived from tests
will remain valid only so long as the brick properties, mortar properties and workmanship remain relatively close to
that specified. The next issue of this Technical Notes series will detail the quality control tests available to insure the
designer that he is obtaining masonry properties of sufficient quality to achieve the performance desired.
39a http://www.gobrick.com/BIA/technotes/t39a.htm
18 of 18 9/13/2009 2:04 PM

Technical Notes 39B - Testing for Engineered Brick Masonry - Quality Control
March 1988
Abstract: Testing prior to and during the construction of engineered brick masonry may be required to provide a
means of quality assurance. Testing may cover materials, to determine compliance with the project requirements,
assemblies, to determine the properties of the masonry as constructed or to establish the properties of masonry in
existing structures. The extent of testing required must be determined by the engineering or architectural firm
responsible for the project design and will depend upon the complexity and importance of the project. This Technical
Notes describes quality assurance procedures applicable to brick masonry assemblies; other issues in this series
address testing of component materials and testing to establish allowable design stresses.
Key Words: brick, bond strength, diagonal tension, engineered brick masonry, grout, masonry testing, mortar,
prism testing, quality assurance, shear, testing.
INTRODUCTION
Engineered brick masonry design is a rational design procedure based on material properties and fundamental
engineering analysis principles. This type of approach, as opposed to an empirical approach, permits the designer to
retain the aesthetic qualities of brick masonry and make efficient use of brick masonry's structural properties.
Since engineered brick masonry design is dependent on material properties, minimum material strength requirements
are determined in the preliminary design phase. Materials (brick, mortar and grout) are selected and may be tested
to determine allowable design stresses for the combination of materials selected (see Technical Notes 39A). Quality
assurance testing is then performed during construction to evaluate the properties of constructed masonry. The
results of these tests are then used to determine if the constructed masonry is acceptable.
This Technical Notes discusses standards developed by the American Society for Testing and Materials (ASTM)
that may be used for quality assurance testing of engineered brick masonry. This Technical Notes is not intended to
replace applicable ASTM standards, but to supplement them.
The purpose of this Technical Notes is to serve as an aid in selecting, applying and interpreting tests. The engineer,
architect or other responsible person must use judgment in selecting and applying these test methods, but it is hoped
that this Technical Notes will aid in that process.
PURPOSE OF QUALITY ASSURANCE TESTING
Several standards are used in the United States for the design of brick masonry structures. These standards are
referenced in most building codes and contain some type of requirement for testing of materials and assemblages to
evaluate material properties, design parameters or as a means of quality assurance. Quality assurance testing is
specifically performed to determine that the materials, construction and workmanship meet the project
specifications.
The BIA Standard (Building Code Requirements for Engineered Brick Masonry, Brick Institute of America, McLean,
Virginia, August 1969), for example, requires inspection and testing in order for the designer to make use of higher
allowable design stresses. Allowable stress values under the BIA Standard are divided into two categories: "With
Inspection" and "Without Inspection". If no inspection is provided, the design allowables for "Without Inspection" are
used and represent a thirty-three percent reduction in magnitude, as compared to the values permitted for "With
Inspection". Therefore, it is advantageous to implement quality assurance measures in some cases to permit higher
allowable stress values.
The type of inspection required in the BIA Standard typically consists of an inspector (the engineer, architect or other
39b http://www.gobrick.com/BIA/technotes/t39b.htm
1 of 13 9/13/2009 2:05 PM
responsible party) and some type of testing. The tests outlined in this Technical Notes are those that are most
commonly performed to satisfy the requirements of the BIA Standard.
TESTING METHODS
The ASTM standards commonly used for quality assurance testing of brick masonry materials and assemblies are
contained in the Annual Book of ASTM Standards. Current copies of applicable standards should be readily available
to laboratory personnel, individuals involved in field sampling and testing, and individuals involved in interpreting test
results. The applicable ASTM standards are:
Component Materials
Clay Masonry Units
ASTM C 67, Standard Method of Sampling and Testing Brick and Structural Clay Tile.
Mortar
ASTM C 270, Standard Specification for Mortar for Unit Masonry.
ASTM C 109, Standard Test Method for Compressive Strength of Hydraulic Cement Mortars (using 2-in. or 50-mm
Cube Specimens).
ASTM C 780, Standard Method for Preconstruction and Construction Evaluation of Mortars for Plain and Reinforced
Unit Masonry.
Grout
ASTM C 476, Standard Specification for Grout for Masonry.
ASTM C 1019, Standard Method for Sampling and Testing Grout.
Assemblies
Masonry Compressive
ASTM E 447, Standard Test Methods Strength for Compressive Strength of Masonry Prisms.
Flexural Bond Strength
ASTM E 518, Standard Test Methods for Flexural Bond Strength of Masonry.
ASTM C 1072, Standard Test Method for Measurement of Flexural Bond Strength.
In addition to the preceding standards, other standards, while not generally used for quality assurance testing, may
be performed in conjunction with compressive and/or flexural bond strength tests to establish a relationship between
test methods for quality assurance purposes. These methods listed below are discussed in detail in Technical Notes
39A.
Flexural Tensile Strength
ASTM E 72, Standard Methods of Conducting Strength Test of Panels for Building Construction.
Shear Strength Diagonal Tension (Shear)
ASTM E 519, Standard Test Method for Diagonal Tension (Shear) in Masonry Assemblages.
Typically, ASTM standards provide clear and concise explanations of the procedures involved in sampling and
testing; however, for the novice, some areas may be confusing. Technical Notes 39 Revised presents a complete
39b http://www.gobrick.com/BIA/technotes/t39b.htm
2 of 13 9/13/2009 2:05 PM
discussion of the preceding standards for testing component materials.
The standards are listed here for the sake of completeness only. The remaining standards relating to the testing of
masonry assemblages are the subject of this Technical Notes.
LABORATORY SELECTION
A laboratory selected to perform masonry testing should be properly staffed and be experienced in masonry testing.
The equipment available at a laboratory will directly affect the types of tests that can be performed, and the
specimens that can be tested. As a minimum, a laboratory will require a curing room with controlled temperature and
humidity, and a compression testing machine with a minimum capacity of 300,000 lb and a 15-in. stroke to perform
prism tests. Other test methods described in this Technical Notes require more specialized equipment that may not
be available at some laboratories.
EVALUATION OF MASONRY STRENGTH
Compressive Strength
General. Masonry assembly compressive strength should be determined by prism tests in accordance with ASTM E
447, Method B (see Figure 1).
Prism Test Schematic
FIG. 1
Specimens. A minimum of three prisms should be constructed, using the same materials and workmanship as used
in the project. The mortar bedding, joint thickness, joint tooling, bonding arrangement and grouting pattern should be
the same as that in the project. No structural reinforcement should be included; however, metal wall ties may be
included if used in the project. Prisms should not be grouted unless all hollow cells and spaces in the actual
construction are to be grouted.
The prism thickness should be the same as that of the actual construction. The prism length should be equal to or
greater than the prism thickness. The height of the prism should be at least twice the prism thickness or a minimum
of 15 in. (375 mm).
39b http://www.gobrick.com/BIA/technotes/t39b.htm
3 of 13 9/13/2009 2:05 PM
Handling and Curing. Prisms should be constructed on the jobsite in an area where they will not be disturbed or
damaged. Prisms should be subjected to atmospheric conditions similar to those of the masonry they represent for a
period of 48 hr prior to being prepared for transportation to the testing laboratory. Prisms should be secured and
transported in such a manner so as not to damage them.
After prisms are delivered to the laboratory, they should be cured in laboratory air, free of drafts, at 75 deg F +/- 15
deg F (24 deg C +/- 8 deg C), with a relative humidity between 30 and 70% for a period of 26 additional days.
Capping. Proper capping of prisms cannot be over-emphasized. Brick units are not perfectly formed and their
bearing surfaces may not be parallel and free from surface irregularities. The purpose of capping the bearing
surfaces is to assure reasonably parallel and smooth bearing planes. This reduces the likelihood of uneven bearing
and stress concentrations that can result in premature prism failure. The capping material itself should have a
compressive strength in excess of that expected of the prism to insure that the capping material does not fail before
the prism.
Laboratory personnel responsible for capping prisms should be knowledgeable of the capping procedures
prescribed in ASTM C 67 and C 140. Poor capping techniques and inappropriate capping materials can result in
erratic test results and lower apparent prism compressive strengths.
Testing. Prisms should be centered under the spherical upper bearing block of the testing machine so that the
resulting load will be applied through the center of gravity of each specimen. This is extremely important since the
introduction of an eccentric load, if the specimen is not properly centered, can result in lower apparent prism
compressive strength.
The speed of testing specified in ASTM E 447 should be followed to obtain consistent results. Past experience on
the effect of the loading rate on compressive strength has shown that, as the loading rate increases, there may be a
significant increase in apparent compressive strength. The prescribed loading rate provides a moderate rate of
loading that produces more consistent results and more accurately represents the true prism compressive strength.
Calculation and Report. The ultimate compressive strength of a prism is calculated by dividing the maximum
compressive load by the cross-sectional area of the prism. For prisms constructed with solid units (ASTM C 216 or
ASTM C 62), or units grouted solid, the gross cross-sectional area is used to calculate compressive strength. For
prisms constructed with ungrouted hollow units (ASTM C 652), the net cross-sectional area (determined by the
procedure described in ASTM C 67) is used in the calculation. When brick masonry prisms with height-to-thickness
ratios (h/t) of less than 5 are tested, the ultimate compressive strength, as calculated above, must be multiplied by
the factors given in Table 1 to correct for slenderness effects.
The report should contain the prism dimensions, prism age, description of materials, maximum compressive load for
each prism, cross-sectional area of each prism, compressive strength of each prism, average compressive strength
of the specimens, standard deviation and coefficient of variation.

a
These values are different from those now given in the August 1969 BIA Building Code Requirements for
Engineered Brick Masonry. They are based on subsequent research and more nearly reflect the masonry
behavior in prisms with h/t less than 5.
b
Height to thickness (h/t).
c
Interpolate to obtain intermediate values.

39b http://www.gobrick.com/BIA/technotes/t39b.htm
4 of 13 9/13/2009 2:05 PM
Recommendations and Evaluation. When prism tests are used as a means of quality assurance for the BIA
Standard, not less than 3 prisms should be constructed for each 5000 sq ft of wall area or each story height,
whichever is more frequent. Additional test prisms may be constructed at the discretion of the engineer or architect.
Often it is desirable to establish strength relationships for prisms cured less than 28 days to prisms cured for 28
days. This may be established by testing one set of 3 prisms (5 prisms preferred), constructed for each curing
period. The prisms should be cured at the site for 24 hr and transported to the laboratory and stored with an
ambient temperature and humidity as prescribed in ASTM E 447 for the remainder of the curing period. One set of
prisms should be tested at 28 days and the other set tested at the desired age level, typically, 3 or 7 days. From
this data, strength relationships between shorter curing periods and 28-day curing periods may be developed.
It is desirable to establish the relationship between early prism strengths to 28-day strengths by testing. However, if
this relationship is not or cannot be established, an approximate method may be used to predict the 28-day prism
compressive strengths.
The work represented by the quality assurance specimens may be deemed acceptable if the average 28-day
compressive strengths or the projected average 28-day compressive strengths are not less than the specified
design compressive strength.
Diagonal Tension (Shear) Strength
General. Under certain circumstance, it is sometimes necessary to directly establish design shear stresses more
accurately than values established as a function of compressive strength (see allowable shear stresses in the BIA
Standard). When this is the case, the methods outlined in ASTM E 519 or ASTM E 72 are used to establish design
values (see Technical Notes 39A). These test methods require large masonry specimens and cannot be used
practically as quality assurance tests. However, these tests may be performed with companion compressive test
specimens to develop a relationship between shear and compressive test results. This permits testing of smaller
compressive specimens as a means of quality assurance.
Companion Specimens. Compressive test prisms should be constructed and tested as outlined in Technical Notes
39A. A minimum of 3 prisms (preferably 5 prisms) should be constructed and tested. The data collected from the
compressive prism tests and diagonal tension tests can be used to establish a relationship between prism
compressive strength and design shear strength. When this relationship is established, tests can then be conducted
by ASTM E 447, in lieu of ASTM E 72 and E 519.
FLEXURAL BOND STRENGTH
General. ASTM E 518 or C 1072 may be used as quality assurance tests to measure the flexural bond strength
between masonry units and mortar. These tests are not intended for use in establishing design stresses. Design
stresses should be established through ASTM E 72, as described in Technical Notes 39A.
Companion Specimens. A relationship between the flexural bond strength obtained by ASTM E 518 or ASTM C
1072 and the transverse strength of ASTM E 72 may be developed by testing companion specimens. This requires
that ASTM E 72 transverse load tests be performed and that companion specimens, as prescribed in ASTM E 518
or ASTM C 1072 be constructed and tested using the same units, mortar and workmanship as the E 72 tests. The
data from these tests can then be used to establish the relationship between the transverse flexural strength and the
flexural bond strength.
Once this relationship has been established, ASTM E 72 tests need not be conducted. Quality assurance tests can
then be made by ASTM E 518 or ASTM C 1072.
ASTM E 518 Test. E 518 provides two methods for performing tests on flexural beams. Method A uses
concentrated loads at 1/3 points of the span (see Fig. 2). Method B uses a uniform loading over the entire span (see
Fig. 3) applied by an air bag.
39b http://www.gobrick.com/BIA/technotes/t39b.htm
5 of 13 9/13/2009 2:05 PM
E518 Method A Test
FIG. 2
E518 Method B Test
FIG. 3
Specimens - Prisms should be built at the jobsite with the same materials and workmanship as the actual
construction.
Prisms constructed in the field for quality assurance testing should be protected from damage, but exposed to the
same atmospheric conditions as the constructed masonry. These prisms should be stored at the jobsite until the
testing date.
Testing - While ASTM E 518 does not specify the orientation of the specimens, specimens for both Method A and
Method B (see Figs. 4 and 5) should be placed with the tooled joints downward; that is, loads should be applied to
the unfinished face. This provides a more standardized test and allows a more accurate comparison of results.
39b http://www.gobrick.com/BIA/technotes/t39b.htm
6 of 13 9/13/2009 2:05 PM
E518 Method A Setup
FIG. 4

E518 Method B Setup
FIG. 5
If Method A is used and failure of any specimen occurs outside of the middle third of the specimen, the test results
for that specimen should be discarded.
Calculation and Report - After testing is completed, the gross area modulus of rupture (tensile bond strength) can
be calculated using one of the following formulae:
Method A test: specimen made of solid masonry units.

where:
39b http://www.gobrick.com/BIA/technotes/t39b.htm
7 of 13 9/13/2009 2:05 PM

R = gross area modulus of rupture, psi (Mpa)
P = maximum machine-applied load, lb (N)
Ps = weight of specimen, lb (N)
l = span, in. (mm)
b = average width of specimen, in. (mm)
d = average depth of specimen, in. (mm)
Method B test: specimen made of solid masonry units.

Method A test: specimen made of hollow masonry units.

where:
S = section modulus of actual net bedded area, in.
3
(mm
3
)
Method B test: specimen made of hollow masonry units.

For calculation of the section modulus based on the net bedded areas of hollow units, the following formulae may be
used:
Fully bedded hollow units; (see Fig. 6).
39b http://www.gobrick.com/BIA/technotes/t39b.htm
8 of 13 9/13/2009 2:05 PM
Cross Section-Full Bedded Hollow Unit
FIG. 6

where:
b1 = width of cores, in. (mm)
d1 = depth of cores, in. (mm)
Face shell bedded hollow units: (see Fig. 7)
39b http://www.gobrick.com/BIA/technotes/t39b.htm
9 of 13 9/13/2009 2:05 PM
Cross Section-Face Shell Bedded Hollow Unit
FIG. 7

ASTM C 1072 Test. ASTM C 1072, commonly known as the "bond wrench test", permits individual mortar joints to
be tested for flexural bond strength by applying an eccentric load to a single joint in a prism (see Fig. 8). This
method has several advantages over the ASTM E 518 test method in that: 1) More data is collected from each
prism. 2) The data gathered is more representative since each joint in a specimen is tested instead of the weakest
joint in the specimen. 3) It may be used to test specimens extracted from existing structures. 4) Joints remaining
after testing by the ASTM E 518 method may be tested and the results of the two methods compared.
Specimens- Prisms should be constructed at the jobsite with the same materials and workmanship used in the
actual construction. Prisms should be constructed in a location where they will not be disturbed or damaged, but be
subjected to atmospheric conditions similar to those of the actual masonry.
Prisms should be a minimum of 2 units in height, with a minimum width of 4 in. (200 mm). It is recommended that the
prisms be a full unit in width. Joints should be 3/8 in. 1/16 in. (9.4 mm 1.6 mm) in thickness. One face of each
prism should be tooled to match the tooling of the project. Prisms should be stored at the jobsite until the testing
date. As a minimum, 5 joints should be tested.
Testing- Prisms should be placed in the support frame so that the tooled joints face the clamping bolts in the loading
arm and are subjected to flexural tension (see Fig. 8). Prisms should be positioned vertically such that a single brick
projects above the lower clamping bracket. A soft bearing material a minimum of 1/2 in. (13 mm) in thickness should
be placed between the bottom of the prism and the adjustable prism support base. The loading arm clamping bolts
should be tightened using a torque of not more than 20 lb-in. (2.3 N-m). Loading should be applied at a uniform rate
such that the total load is applied in not less than 1 min. nor more than 3 min.
39b http://www.gobrick.com/BIA/technotes/t39b.htm
10 of 13 9/13/2009 2:05 PM
Bond Wrench Test Apparatus
FIG. 8
Calculation and Report- After testing is completed, the flexural bond strength can be calculated as:
Specimens made of solid masonry units



where:
Fg = gross area flexural tensile strength, psi (MPa),
P = maximum machine applied load, lb (N),
P1 = weight of loading arm, lb (N) (See Appendix X1 in ASTM C 1072)
L = distance from center of prism to loading point, in. (mm),
L1 = distance from center of prism to centroid of loading arm, in. (mm)
(See Appendix X1 in ASTM C 1072)
b = average width of the cross-section of failure surface, in. (mm),
d = average thickness of cross-section of failure surface, in. (mm)
39b http://www.gobrick.com/BIA/technotes/t39b.htm
11 of 13 9/13/2009 2:05 PM
Specimens made of hollow masonry units



where:
Fn = net area flexural tensile strength, psi (MPa),
S = section modulus of actual net bedded area, in.
3
(mm
3
),
A = net bedded area, in.
2
(mm
2
).

For calculation of section modulus, see Eqs. 5 and 6 and Figs. 6 and 7. The net bedded area may be calculated as:

Fully bedded hollow units; (see Fig. 6)

Face shell bedded hollow units; (see Fig. 7)

EVALUATION OF TEST RESULTS
General
In most cases, strength levels are established by testing or by the selection of design values. Evaluation then
becomes a simple matter of comparing the results of the quality assurance tests with the desired strength levels.
Unsatisfactory Test Results
Examination of Procedures. Several alternatives are available when test results fall below the required level. If
backup specimens are not available for testing, then close examination should be made of the method of prism
construction, the handling of the specimens during transportation and storage and of the laboratory facilities and test
procedures. Actual stresses should be checked to determine if the lower strength will provide structural stability.
After the above observations and calculations are completed, some judgments should then be made. They are:
1. Did mortar proportions or properties change?
2. Did brick properties change?
3. Were there unusual curing conditions?
4.Were specimens damaged during transit or storage?
39b http://www.gobrick.com/BIA/technotes/t39b.htm
12 of 13 9/13/2009 2:05 PM
5. Were specimens properly constructed?
6. Were test procedures properly followed?
7. Were calculations correctly performed?
As a result of these questions, the possible cause of low test results may be determined.
Alternate Test Procedure. If no immediate solution is evident and reduced strengths result in safety factors below
an acceptable level, prisms may be cut from the area in question and tested as described previously.
After specimens are cut from the wall, they should be transported to the lab for testing. If specimens are cut by a
water-cooled saw, they should be allowed to dry prior to testing.
Exercise of Judgment. If the test results are still low, then a judgment is required. If the field-cut specimen tests
result in strengths that lower the factor of safety below an acceptable level, then removal of the masonry in question
must be considered. Obviously, this is the last resort after all other possibilities have been closely examined.
SUMMARY
This Technical Notes has discussed quality assurance testing based on procedures developed by ASTM. Testing
agencies using these ASTM test methods should be fully aware of their procedures and limitations, so that improper
application and erroneous results are avoided.
While the testing procedures outlined in this Technical Notes are primarily for engineered brick masonry under the
BIA Standard, they may also be used for non-structural masonry testing and quality assurance testing under other
standards. Excessive testing can add unnecessary cost to the project. It is important that care be exercised to avoid
excessive testing.
It should also be pointed out that, while strengths are important properties of masonry, they are not the only
desirable properties. Strengths should not become so important that other desirable properties of masonry are
sacrificed. It is still important that masonry be resistant to water penetration, provide sound control and be properly
detailed. Quality assurance testing is just one tool to provide assurance that all of the desirable properties of brick
masonry are obtained.
Testing procedures described in this Technical Notes may involve the use of hazardous materials, operations and/or
equipment. This Technical Notes does not purport to address all of the safety practices associated with the use of
these test methods. It is the responsibility of the user of this Technical Notes to establish appropriate safety and
health practices and determine the applicability or regulatory limits prior to the use of the test methods described.
The information contained in this Technical Notes is based on the available data and experience of the technical staff
of the Brick Institute of America. The information should be recognized as recommendations which, if followed with
judgment, should prove beneficial to the performance of masonry construction.
Final decisions on the use of information, details and materials as discussed in this Technical Notes are not within
the purview of the Brick Institute of America and must rest with the product designer, owner or both.
39b http://www.gobrick.com/BIA/technotes/t39b.htm
13 of 13 9/13/2009 2:05 PM
RE

Technical Notes 40 - Prefabricated Brick Masonry - Introduction
Revised August 2001
INTRODUCTION
The desire of the construction industry to minimize on-site labor and reduce construction time has resulted in the
prefabrication of building components. Methods for the prefabrication of masonry have been developed by several
segments of the brick industry: mason contractors, brick manufacturers, equipment manufacturers and others
closely associated with the industry.
This Technical Notes covers the history, advantages, disadvantages, considerations for prefabrication, fabrication
methods, materials, specifications and present applications of prefabricated brick masonry.
There are several recent developments which make prefabrication of brick masonry possible. The most important is
the development and acceptance of a rational design method for brick masonry. Other factors, such as research
with new and improved brick units and mortars, have aided the rapid progress in the prefabrication process.
This Technical Notes deals only with prefabricated brick masonry using full size brick units. Prefabricated elements
of thin brick facing units, in conjunction with concrete, fiberboard or other backings, are discussed in Technical Notes
28C.
HISTORY AND DEVELOPMENT
Individual uses of prefabricated brick masonry have occurred for more than 100 years. Brick piers were laid on
boards for use below sea level in Galveston, Texas prior to 1900. The prefabrication of brick masonry involving
equipment rather than bricklayers had its early development during the 1950s in France, Switzerland and Denmark.
At the same time, the Structural Clay Products Research Foundation, which was once a part of the Brick Industry
Associations Engineering & Research Division, developed a prefabricated brick masonry system. This system,
known as the SCR building panel was used in the construction of several structures in the Chicago area.
Most of the early methods of panelization were attempts to mechanize the bricklaying process to produce standard
panels, using unskilled labor. Later trends, especially in the United States, have been toward the retention of skilled
labor using conventional masonry construction practices and devising various means to increase mason productivity.
(See FIG. 1)
There are several different methods or systems of prefabrication being used in the U.S. today. Some systems
employ proprietary mechanized equipment, while others are not patented but are merely methods of prefabrication
developed by individual manufacturers or mason contractors.
ADVANTAGES OF PREFABRICATION
There are several advantages of prefabrication over conventional masonry construction. By using panelized
construction, the need for on-site scaffolding is virtually eliminated. If an off-site plant is used, the work and storage
area for masonry materials at the job site are kept to a minimum. When proper scheduling of delivery is maintained,
the panels can be erected as they are delivered, eliminating any need for panel storage at the site.
The use of panelization also makes possible the fabrication of complex shapes. These shapes can be accomplished
without the need for expensive falsework and shoring necessary for laid in-place masonry. Complicated shapes with
returns, soffits, arches, etc., are accomplished by using jigs and forms. Repetitive usage of these shapes can lower
costs appreciably; the more re-uses, the lower the per panel cost. The designer is able to obtain more complex
shapes and different bonding patterns when including brick masonry panels. (See FIG. 2) Fabricating a panel in
http://www.gobrick.com/BIA/technotes/t40.htm
1 of 11 9/13/2009 2:07 PM
running bond and rotating it results in soldier courses in running bond which are expensive and difficult to execute if
conventionally laid. Sloped sills and soffits, along with other shapes and patterns, are easily achieved in this manner.
One of the distinct advantages of prefabrication is the possibility of year round work and multi-shift workdays.
Off-site construction permits the labor force to work under conditions not affected by weather. The use of
prefabricated masonry may eliminate the need for cold weather construction practices. Prefabrication requires
stringent quality control. Building panels on a standard form in a single location makes this easier to attain. Mortar
batching systems can be tightly controlled by automation or sophisticated equipment. In a factory, the curing
conditions are more consistent, since they are less affected by weather changes.
Panelization on some projects may save construction time. In most cases it is possible to fabricate masonry panels
prior to ground breaking, thus keeping far enough ahead of the in-place construction work to permit panel placement
when needed. In the case of bearing wall structures, construction time could be shortened since panels have
completely cured when erected. This allows the construction crew to immediately start erection of the next floor level
and thus expedite construction.
In some cases, the structure of the building, or the perimeter beams, may be downsized due to the ability of the
prefabricated masonry panel to span column to column. This distributes the wall load directly to the columns thus
lowering the floor load at the slab edge.
The savings in construction and time can also provide economy to the building owner. Earlier completion allows
earlier occupancy. In the case of rental or commercial properties, this allows the owner to have income production
start sooner.
DISADVANTAGES OF PREFABRICATION
As with any construction method, prefabrication has inherent disadvantages as well as advantages. Prefabrication of
masonry to date has not achieved the economy of construction originally desired on flat wall areas. Typically,
prefabricated masonry costs are the same or higher than most conventionally laid-in-place masonry on a square foot
(square meter) cost basis.
The size of brick masonry panels is limited primarily by transportation and erection limitations. Architectural design
may, in some cases, need modification to use prefabricated brick masonry.
Another disadvantage of prefabricated brick masonry, as in other panel systems, is the limited adjustment
capabilities during the construction process. In-place masonry construction allows the brick mason to build masonry
to fit the other elements of the structure by adjusting mortar joint thickness over a large area so that it is not
noticeable. With prefabricated elements, the adjustments must take place in the connections and the joints between
panels. The use of prefabricated elements sometimes requires other crafts or trades to adopt more stringent
construction tolerances for their work beyond the standard construction practices for their respected trades.
CONSIDERATIONS OF PREFABRICATION
The designer must evaluate each project to determine the feasibility and adaptability of prefabrication to that project.
Basic questions that must be considered prior to a decision should include, but not necessarily be limited to, the
following for each individual project:
1. Is the building layout, plan and elevation suitable to prefabrication?
2. Is there a location on-site suitable for panel fabrication and storage?
3. Is it desirable to use off-site prefabrication due to the limited size of the building site?
4. What is the completion schedule and what time of year is construction to take place?
5. Are structural design solutions unrealistic when prefabrication is used?
6. Can a reasonable level of quality control in all trades be achieved if prefabricated brick masonry is utilized?
7. Is prefabrication, based on answers to questions 1 through 6, an economical answer?
Contacting a fabricator during the design development or construction document phase of a project may be
advisable to determine whether particular elements can be constructed using prefabricated masonry. A fabricator
can also advise how different panels can be supported from the building structure.
FABRICATION METHODS
http://www.gobrick.com/BIA/technotes/t40.htm
2 of 11 9/13/2009 2:07 PM
There are two basic manufacturing methods being used in brick masonry prefabrication: hand-laying and casting.
The equipment used in prefabrication as practiced today varies widely. It ranges from simple conventional masonry
hand tools to highly sophisticated automated machinery.
Hand-laying
The hand-laying method of prefabrication is similar to conventional laid in-place masonry. That is, the brick are
placed in mortar by a mason, except it is accomplished in an area removed from the final location of the masonry
element. The bricklaying may be done using conventional bricklaying tools, automated equipment or both. Automated
equipment typically has several components:
1. Devices (such as forms or jigs) for establishing the shape of the panel and locating courses.
2. Scaffolding to keep the bricklayer at a comfortable position (See FIG. 3).
3. Material delivery to the bricklayer.
4. Mortar application.
The hand-laying method will usually employ the conventional masons tools: trowel, jointing tool, etc. In addition, the
hand-laying method may also utilize corner poles, jigs, and templates for special shapes. The hand-laying method
will usually employ some type of adjustable scaffolding. Adjustable scaffolding can greatly increase mason
productivity and reduce fabrication costs. Mechanized and pneumatic mortar spreaders may also be used in the
hand-laying method to distribute mortar to the bed joints.
This method is particularly adapted to a mason contractor serving as the prefabricator since the contractors regular
labor force can be employed. The fabrication operation can be performed either at an off-site plant or an on-site
temporary production facility. Casting
The casting method of fabrication involves the combining of masonry units, mortar and grout into a prefabricated
element using unskilled labor. The casting method is performed with the element either in a horizontal or a
non-vertical position. In general, the casting method lends itself to automated equipment that requires a form or an
alignment device, some method of placing units and reinforcement, and a method for introducing mortar or grout.
The usual practice is to place the units, either by hand or machine, and fill the form with a grout at atmospheric
pressure or under moderate pressure. This method of prefabrication usually takes place in an off-site plant.
There are specialized tools used in the casting method. Jigs and forms provide for the alignment of the brick and the
spacing for the joints. The casting method may require that the face of the unit be protected from contamination by
the mortar or grout. This is usually done by applying a contact surface to the exterior face of the brickwork.
Pressure at the contact surface of the brick face is created by either an inflated form face or by applying a load to
the brick and forcing it into a soft material. Some prefabrication has employed the casting method and automated
unit placing machinery. This equipment places the unit with proper joint width by machine in lieu of hand placement of
units in jigs or forms. Pressurized grouting systems have also been widely used in the casting method of
prefabrication.
MATERIALS
Masonry Units
Both solid brick and hollow brick have been used in prefabrication. Solid brick masonry units are those which have
coring of less than 25 per cent of the bedding area. Hollow brick units are cored in excess of 25 per cent but no
more than 60 per cent of the bedding area. The hollow units are suitable for and used in applications where
reinforcement is required. Reinforcement is often required not only for in-place structural reasons but for loads and
stresses included during panel handling.
Unit face size is based on economy and appearance, just as in conventional masonry. Dimensional tolerances of the
size and face of brick may be more stringent if the casting method is used to form the prefabricated masonry.
Specially-shaped units are often employed in fabricated masonry panels. Special units made for returns other than at
right angles allow continuity in bond. Channel-shaped units accommodate the placement of reinforcement. Single and
multiple wythe panels of the units outlined above have been used in prefabricated work, but single wythe panels are
most commonly used.
http://www.gobrick.com/BIA/technotes/t40.htm
3 of 11 9/13/2009 2:07 PM
Mortar and Grout
Prefabrication may use either conventional mortar or mortar with additives to increase bond strength. If the panel is
not reinforced, care must be taken to determine compatibility of the brick and mortar. Testing should be performed
with the mortar and brick selected for construction to determine if the combination will indeed produce the required
flexural strengths for the project.
Most masonry panels utilize steel reinforcement to resist tensile and shear stresses due to handling and in-service
loads. Fine grout is used to surround the reinforcement and create a homogeneous element. (See Technical Notes
17A for material requirements of reinforced brick masonry.)
SPECIFICATION FOR PREFABRICATED PANELS
The American Society for Testing and Materials (ASTM) has developed a standard specification for prefabricated
masonry panels that includes many items that should be considered when using prefabricated brickwork. The ASTM
C 901 Specification for Prefabricated Masonry Panels contain the following sections: Materials and Manufacturing;
Structural Design; Dimensions and Permissible Variations; Workmanship, Finish and Appearance; Quality Control;
Identification and Marking; Shop Drawings; and Handling, Storage and Transportation. Each section is discussed
below.
Materials and Manufacture
The appropriate ASTM material standards for brick and structural clay tile for use in prefabricated masonry are
referenced in this section of ASTM C 901. Brick must meet the requirements of one of the following standards:
ASTM C 62 for building brick, ASTM C 216 for facing brick, ASTM C 652 for hollow brick, or ASTM C 126 for
ceramic glazed units. Standards for brick that are not included in ASTM C 901, but which can be used to construct
prefabricated brick masonry panels are ASTM C 1088 for thin brick veneer and ASTM C 1405 for single fired,
glazed brick.
Mortar and grout must meet the requirements specified in ASTM C 270 and ASTM C 476, respectively.
All metal embedded in masonry panels, except structural reinforcement, must be coated with a corrosion-resistant
material or be made of stainless steel Type 304 or 316. This includes all ties, fittings, anchors and lifting inserts. The
corresponding specifications for zinc coatings, copper-coated wire, stainless steel, and all types of reinforcement
are referenced in this section of ASTM C 901.
Structural Design
Structural design of prefabricated masonry panels must be performed in accordance with the local building code and
ASTM C 901. Where there is no local building code, a national model code should be used. Panels must be
designed for all loads and restraining conditions from fabrication to installation and in-service performance. Wind,
seismic, and other dynamic loads must be considered as mandated by the building code.
Differential movement between dissimilar materials within a panel and between panels and their supports must also
be considered.
Lifting devices and their connections must have an ultimate capacity of four times the dead weight of the appropriate
portion of the panel. Inclination of the lifting forces must also be considered.
Dimensions and Permissible Variations
Panel sizes are based on multiples of the nominal sizes of the individual masonry units. The nominal thickness of
panels shall be the sum of the nominal thickness of the masonry plus the nominal thickness of any cavities. Actual
panel thickness must be determined for adequate strength, fire resistance, and other design criteria as required for
the type of structure and occupancy.
Specified dimensions of the panel can vary from the nominal size by the thickness of one mortar joint or ? in. (13
mm) maximum. Custom dimensions are permitted and should be shown on the drawings or specified, however
modular dimensioning is recommended. (See Technical Notes 10A for more information on this subject.) Dimensional
tolerances for panel size, thickness, and out-of-square are stated in this section of ASTM C 901.
http://www.gobrick.com/BIA/technotes/t40.htm
4 of 11 9/13/2009 2:07 PM
Workmanship, Finish and Appearance
A sample panel should be used to establish acceptable workmanship and appearance for facing panels. Individual
units and joints should be properly aligned. The location of anchors, inserts, and lifting and connection devices should
not vary more than 3/8 in. (10 mm) from the specified location. Warpage is limited to a maximum of 1/8 in. (3 mm)
for each 6 ft (1.8 m) of panel height or width.
Quality Control Brick.
The brick unit compressive strength and the initial rate of absorption of brick must be determined. A sample of at
least ten units for each 50,000 units used in panel fabrication must be tested in accordance with ASTM C 67.
Mortar and Grout.
The proportions of mortar and grout must be determined as given in ASTM C 270 or C 476, respectively. Bond
enhancing admixtures must be mixed in accordance with the manufacturers specifications. Once the proportions are
determined, the compressive strength of a sample of 12 specimens should be determined at intervals of 1, 3, 7, and
28 days to determine the relationship between early-age strengths and the 28-day strength for both mortar and
grout. During production, at least one batch of mortar and grout should be sampled each day to determine 1, 3, or 7
day strengths.
Completed Panels
Masonry assemblies must also be tested. One sample of three compression specimens must be tested for every
5000 ft2 (465 m2) of panel production or every story height. Test one sample of three flexural specimens for each
days work. Specimens should be constructed and tested in accordance with ASTM Test Method C 1314 for
compressive strength and ASTM Test Method E 518, horizontal beams with third-point loading, for flexure. Also,
flexural bond strength may be evaluated by ASTM Test Method C 1072 or ASTM Test Method C 1357 in lieu of the
method specified in ASTM C 901.
Identification and Marking.
Each prefabricated member must be marked to indicate its location on the structure, its top surface, and the date of
fabrication. These marks shall correspond to those on the placing drawings.
Shop Drawings
Shop drawings consist of fabrication drawings and placing drawings. Fabrication drawings show details and
locations of reinforcement, inserts, anchors, bearing seats, lifting inserts, coursing, size and shape of openings, and
panel size and shape. Placing drawings show panel identification, location, reference dimensions, panel dimensions,
dimension of joints between panels, and connection details.
Handling, Storage and Transportation
Care must be taken not to overstress, warp or otherwise damage the panels during manufacturing, curing, storage,
and transportation. Damaged panels must be replaced, unless authorized by the architect or engineer.
INSTALLATION OF PANELS
Most panels are trucked to the job site and lifted into place by cranes. (See FIG. 4) Lifting devices are built into the
panels for this purpose. These panels are usually attached to the structure by welding or bolting. (See FIG. 5)
Connections between the panels and other structural elements of the building provide transfer of both vertical and
horizontal loads. Typically per panel, there are only two connections located near the bottom that transfer the
weight. These connections also transfer horizontal loads. The remaining connections transfer only horizontal loads.
PREFABRICATION EXAMPLES
The use of prefabricated brick masonry in construction has become quite widespread. Prefabricated brick panels
have some very dramatic and aesthetically pleasing applications throughout the United States. Most of the projects
built in the United States use single wythe, reinforced brick panels as non-loadbearing curtain wall panels. However,
http://www.gobrick.com/BIA/technotes/t40.htm
5 of 11 9/13/2009 2:07 PM
panelized brick construction is not limited to curtain wall applications and some loadbearing panels have been
constructed.
Figures 6 through 10 show several recent construction projects using prefabricated brick masonry panels. The
panels for these projects have been built utilizing the full spectrum of methods previously outlined.
CONCLUSION
Prefabricated brick masonry panels are an excellent solution to a multitude of problems commonly found on jobsites
such as material storage concerns, tight construction schedules, and quality assurance issues. Also, there are a
wide variety of possibilities that can be achieved with prefabricated panels that would either be cost prohibitive or
impossible to construct with brick otherwise.
Architects, engineers, and specifiers are urged to work with the manufacturer of the panels to ensure that the
desired effects and final goals can be realistically met as addressed previously in Considerations for Prefabrication;
they can also reference ASTM C 901 for industry standards and material requirements.
Prefabrication of brick masonry is a rapidly developing field and future innovations and needs could greatly affect its
value as a design solution.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Industry Association. The information contained herein must be used
in conjunction with good technical judgment and a basic understanding of the properties of brick masonry. Final
decisions on the use of the information contained in this Technical Notes are not within the purview of the Brick
Industry Association and must rest with the project architect, engineer and owner.
REFERENCES
1. Standard C901-00a Specification For Prefabricated Masonry Panels, American Society for Testing and Materials,
Vol. 04.05, 2001.
2. Reinforced Brick Masonry - Materials and Construction, Technical Notes on Brick Construction 17A, Reston, VA,
August 1997.
3. Thin Brick Veneer, Technical Notes on Brick Construction 28C, Brick Industry Association, Reston, VA, February
1990.
FIG.1
http://www.gobrick.com/BIA/technotes/t40.htm
6 of 11 9/13/2009 2:07 PM
The The Wendel Wyatt Building, Portland, Oregon ZGF Architects, Portland, Oregon (Courtesy of L.C.
Pardue, Inc.)
FIG. 2
Complex Shapes With Different Bonding Patterns (Courtesy of Vet-O-Vitz)
FIG. 3
Panel Prefabrication Plant, Automated Scaffolding (Courtesy of Vet-O-Vitz)
http://www.gobrick.com/BIA/technotes/t40.htm
7 of 11 9/13/2009 2:07 PM
FIG. 4
Placing Panel (Courtesy of Vet-O-Vitz)
FIG. 5
Connection Detail
http://www.gobrick.com/BIA/technotes/t40.htm
8 of 11 9/13/2009 2:07 PM
FIG. 6
Turn Pike Metroplex, East Brunswick, New Jersey Gatarz Venezia Architects, East Brunswick, New Jersey
(Courtesy of Vet-O-Vitz)
FIG. 7
The Center for Molecular Studies University of Cincinnati, Cincinnati, OH Frank O. Gehry & Associates,
Santa Monica, CA (Courtesy of Vet-O-Vitz)
http://www.gobrick.com/BIA/technotes/t40.htm
9 of 11 9/13/2009 2:07 PM
FIG. 8
Sand Lake Plaza, Dayton, OH Hixson Architects and Engineering, Cincinnati, OH (Courtesy of Vet-O-Vitz)
FIG. 9
The Port of Portland Headquarters Portland, Oregon ZGF Architects, Portland, Oregon (Courtesy of L.C.
Pardue, Inc.)
http://www.gobrick.com/BIA/technotes/t40.htm
10 of 11 9/13/2009 2:07 PM
FIG. 10
Safeco Field, Seattle, Washington NBBJ Architects, Seattle, Washington (Courtesy of L.C. Pardue, Inc.)
http://www.gobrick.com/BIA/technotes/t40.htm
11 of 11 9/13/2009 2:07 PM
2008 Brick Industry Association, Reston, Virginia Page 1 of 9
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
41
January
2008
Hollow Brick Masonry
Abstract: This Technical Note presents information about the use of hollow brick in both structural and anchored veneer
applications. Basic properties of hollow brick units are presented, including applicable ASTM standards. Issues specific to
hollow brick masonry are discussed, including design details, structural performance and construction methods.
Key Words: compressive strength, hollow brick, reinforced masonry, veneer.
INTRODUCTION
Originally, brick was formed by placing moist clay in a mold by hand. As modern industrial methods were
implemented in the brick manufacturing process, the majority of production was changed from a molded process
to an extrusion process. Extrusion more easily accommodates the inclusion of holes in a brick unit, which in
turn can make the manufacture and use of brick more cost-effective and material-efficient. Traditionally, the size
and number of holes in a brick unit have varied based on manufacturer capabilities, type of clay being extruded,
type of firing process, and intended use of the product. As part of the evolution of brick unit manufacture and
classification, these various hole patterns were categorized into two basic designations: solid brick and hollow
brick. Solid brick are defined as having holes (or voids) not greater than 25 percent of the units bed area.
Hollow brick are defined as having greater than 25 percent and at most 60 percent void areas. Hollow brick are
further classified into those with a void area not greater than 40 percent and those with greater than 40 percent
voids.
In todays construction, the majority of hollow brick produced are used in two basic applications. The first is in
reinforced or unreinforced single-wythe structural walls. Hollow brick units provide both the structural component
and the brick finish without the need for additional materials. Hollow brick for this type of use generally range in
size from 4 to 8 in. (102 to 203 mm) in nominal thickness with void areas in the 35 to 60 percent range. Typical
single-wythe applications of hollow brick include commercial, retail and residential buildings; hotels; schools;
noise barrier walls; and retaining walls. The second application of hollow brick is as veneer units. These brick are
generally 3 to 4 in. (76 to 102 mm) in nominal thickness with void areas typically between approximately 26 and
35 percent.
Material Selection
Use hollow brick units conforming to the requirements
of ASTM C652
Specify Grade SW brick for most exterior exposures
and Grade MW for interior projects and exterior
exposures protected from freezing
Specify Type HBS brick for typical projects, Type HBX
for projects where higher precision units are desired,
and Type HBA for projects where unique variations in
dimension or appearance are desired
Structural (Non-Veneer) Construction
When placing mortar, use face-shell mortar bedding
Place reinforcement in accordance with applicable
building codes
In single-wythe walls, add a drainage space to the
interior and through-wall flashing where additional water
penetration resistance is necessary
Veneer Construction
When placing mortar, use full mortar bedding
Use typical brick veneer details
Other Details
For corbelling, use solid units or solidly filled units with
approval of building official
For fireboxes and chimney construction, use solid units
or fully grouted units
SUMMARY OF RECOMMENDATIONS:
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 2 of 9
Figure 1 presents the three typical configurations
of hollow brick units. Actual coring patterns vary by
manufacturer and may depend on raw materials,
extrusion equipment, firing methods or other factors.
Note that as used in Figure 1 and throughout this
Technical Note, a core is a void having a cross-
sectional area of 1.5 in. (968 mm) or smaller, and a
cell is a void larger than a core.
Many designers are familiar with the design,
construction and performance of masonry built
with solid units. Hollow brick masonry is similar in
many instances. This Technical Note describes the
classifications, properties and uses of hollow brick.
Further information regarding single-wythe walls is
presented in Technical Note 26. Information on brick
veneer construction can be found within the Technical
Note 28 Series.
PROPERTIES OF HOLLOW
BRICK MASONRY
Strength
The structural design of hollow brick masonry is
governed by model building codes and ACI 530/
ASCE 5/TMS 402 Building Code Requirements for
Masonry Structures, also known as the Masonry
Standards Joint Committee (MSJC) Code [Ref. 4].
Hollow brick masonry can be designed by empirical
requirements or by rational design procedures.
Depending on materials and mortar bedding,
prescriptive stresses can be different for hollow brick
masonry than for solid brick masonry. The following
sections highlight some of the specific requirements
for hollow brick units.
Compressive Strength of Units. Compressive strength of hollow brick can be reported on either a gross or net
cross-sectional area basis, depending on how the value is to be used. The gross area compressive strength is used
to determine compliance with ASTM C652, Standard Specification for Hollow Brick (Hollow Masonry Units Made
From Clay or Shale) [Ref. 2] for purposes of durability and empirical design requirements. The net area compressive
strength is needed for structural computations in structural applications using rational design of masonry.
An internal BIA survey conducted in 1994 showed that the range of compressive strength of 6 to 8 in. (152 to
203 mm) thick hollow brick based on gross cross-sectional area is between 2190 psi (15.1 MPa) and 12,795 psi
(88.2 MPa), with an average compressive strength equal to 6740 psi (46.5 MPa). More recent testing indicates
hollow brick of 3- to 4-in. (76- to 102-mm) nominal thickness have similar compressive strengths as solid units of
the same size [Ref. 5].
Brick units generally have higher compressive strengths than other loadbearing masonry materials. This makes
hollow brick particularly well-suited for reinforced masonry applications where the increased strength of the unit
can allow thinner wall sections to handle the same loading.
Compressive Strength of Masonry. The compressive strength of hollow brick masonry depends on unit strength,
mortar type, mortar bedding area, grouting and thicknesses of face shells and webs. The design strength of the
masonry can be determined by testing sample prisms (prism test method). During construction, strength can
be verified using prism testing or from tabulated values based on brick strength and mortar type (unit strength
method). Ungrouted prisms exhibit failure in compression by a splitting of the unit through the cross webs due to
Solid Face Shell
Cell or Core
Cell
Webs
End Shell or Web
Figure 1
Hollow Brick Configurations
(a) Solid Shell Hollow Brick
(b) Double Shell Hollow Brick
(c) Cored Shell Hollow Brick
Cored Face Shell
Cell
Webs
End Shell or Web
Double Face Shell
Cell or Core
Cell
Webs
End Shell or Web
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 3 of 9
the lateral expansion of the mortar. Filling the cells of hollow brick with grout will generally increase the masonrys
capacity; however, the result is a decrease in the net area compressive strength due to the increased area of the
grouted section. The strength of grouted hollow prisms is significantly affected by both the tensile strength of the
unit and by the compressive strength of the mortar [Ref. 8].
The compressive strength of hollow brick masonry is based on the minimum net cross-sectional area. This is
normally the net mortar bedded area (face-shell bedding) and is used in structural calculations. When using prism
testing to determine or verify compressive strength, the MSJC Code requires that the prisms be built with units fully
bedded in mortar (i.e., all face shells and webs fully mortared).
Values obtained from prism tests must be corrected based on the height-to-thickness (h/t) ratio of the prism. The h/t
ratio provides a uniform basis for the determination of compressive strength. Codes stipulate the correction factors
to use for masonry prisms. An h/t of 2 has been adopted as the base level.
Research shows typical values of ungrouted hollow brick masonry compressive strength based on net area ranging
from 3470 psi (23.9 MPa) to 6620 psi (45.6 MPa). Figure 2 shows typical values of the net area compressive
strength of grouted and ungrouted hollow brick masonry prisms from the research [Refs. 3, 8].
Flexural Strength. The flexural tensile strength of hollow brick masonry is influenced by mortar and unit
configurations and the use of reinforcing steel. Previous research has indicated that face-shell bedded hollow brick
masonry exhibits a lower flexural tensile strength than solid brick masonry laid with the same mortar, fully bedded.
This is likely due to the relative thickness of the face-shell bedded mortar joints and the drying of the mortar before
the hollow unit is laid (face-shell bedding has more surface area exposed to air relative to its volume than does a
full mortar bed). More recent research has indicated that the percentage of voids, ranging from approximately 22 to
35 percent, in nominal 4-in. (102-mm) brick has no significant effect on the flexural strength of the resulting fully
bedded masonry prism [Ref. 5].
Fire Resistance
The excellent fire-resistant qualities of brick masonry are well known. However, there have been relatively few
full-scale fire tests of hollow brick masonry walls. This is due in part to the acknowledgment that brick masonry is
inherently fire-resistant. Fired clay products provide superior fire resistance. Fire resistance ratings for hollow brick
Unit Net Area Compressive Strength, psi (MPa)
8,000 10,000 12,000 14,000 16,000 18,000
(55) (69) (83) (97) (110) (124)
P
r
i
s
m

N
e
t

A
r
e
a

C
o
m
p
.

S
t
r
e
n
g
t
h
,

p
s
i

(
M
P
a
)
7,000
6,000
5,000
4,000
3,000
(48)
(41)
(34)
(28)
(21)
2,000
(14)
Mortar Type: M S N
Grouted:
Ungrouted:
Figure 2
Hollow Brick Prism Compressive Strengths
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 4 of 9
masonry assemblies can be taken from results of actual testing or can be calculated using industry standards.
Results from actual wall tests performed in accordance with ASTM E119 are listed in Table 1 [see Ref. 6 and
Technical Note 16].
Wall Assembly
2
Fire Resistance
Rating
3
, Hours
3-in. (76-mm) hollow brick veneer, 1-in. (25.4-mm) air space, building felt, OSB sheathing, wood studs,
-in. (12.7-mm) gypsum wallboard (fire exposure on brick side)
1
4-in. (102-mm) hollow brick veneer, 1-in. (25.4 mm) air space, building felt, OSB sheathing, wood studs, -in. (12.7-
mm) gypsum wallboard (fire exposure on brick side)
1
4-in. (102-mm) hollow brick masonry, solid grouted 1
5-in. (127-mm) hollow brick masonry 1
6-in. (152-mm) hollow brick masonry 1
8-in. (203-mm) hollow brick masonry, units at least 71% solid, combustible members framed in 1
5-in. (127-mm) hollow brick masonry, solid grouted 2
6-in. (152-mm) hollow brick masonry, solid grouted 3
8-in. (203-mm) hollow brick masonry, units at least 71% solid, with noncombustible members or no members framed in 3
10-in. (254-mm) hollow brick masonry 3
8-in. (203-mm) hollow brick masonry, units at least 60% solid, with noncombustible members or no members framed
in, cells filled with loose fill insulation
4
8-in. (203-mm) hollow brick masonry, solid grouted 4
TABLE 1
Fire Resistance Ratings of Hollow Brick Masonry
1
1. Adapted from Refs. 1 and 6.
2. Nominal thicknesses given for masonry.
3. When a -in. (15.9-mm) layer of plaster is added to the surface of the masonry, the fire resistance rating may be increased by one hour.
An alternative way of determining the fire resistance of a wall assembly is by calculating the equivalent
thickness of the brick. This approach has been approved by the model building codes to determine fire
resistance ratings of walls not physically tested by ASTM E119. The fire resistance rating of hollow brick
masonry is determined by its equivalent solid thickness. The equivalent thickness is calculated by subtracting
the volume of core or cell spaces from the total gross volume of a brick unit and dividing by the exposed face
area of the unit. The resulting thicknesses can be compared with the requirements given within the building
code. As an example, the 2006 International Building Code requires that for a one-hour rating, an equivalent
thickness of 2.3 in. (58 mm) of hollow brick be provided [Ref. 1]. For a 28 percent void area, this would equate
to an actual brick thickness of 3.2 in. (81 mm). By contrast, the requirement for solid brick is an equivalent
thickness of 2.7 in. (69 mm), meaning that a unit with 22 percent coring would need to be 3.5 in. (88 mm) in
actual thickness.
For additional information on fire resistance ratings and calculations, refer to Technical Note 16.
Water Penetration Resistance
The water penetration resistance of hollow brick masonry depends upon the materials, wall construction and
workmanship used. The best water penetration resistance is provided by drainage wall systems, such as those
incorporating brick veneer. For hollow brick veneer applications, water penetration resistance is provided by
proper detailing, including clean air spaces, through-wall flashing and weeps. Testing has shown fully mortar-
bedded hollow brick veneer to have water penetration resistance similar to that of solid brick veneer [Ref. 5]. For
all brick veneer, water penetration resistance depends on proper design and detailing, as presented in Technical
Note 7 and the Technical Note 28 Series. Flashing should be provided at the wall base, below and above all wall
openings, at roof/wall intersections, and at the tops of parapet walls. Flashing and weeps will collect water that
enters the wall and direct it back to the exterior.
Many hollow brick used as single-wythe walls are designed to act as a variation of a barrier wall, relying on the
thickness and mass of the materials to act as a barrier to water penetration. The single-wythe wall design is not
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 5 of 9
inherently as resistant to water penetration as are drainage wall systems or multi-wythe barrier wall systems
and may not be appropriate for some severe exposures. With careful detailing and good construction practices,
however, they can perform well. For example, vertically reinforced and grouted brickwork often provides good
water penetration resistance. With single-wythe masonry, it is especially important to use a mortar joint profile
that sheds, rather than collects, water. Concave and V joints are recommended over raked joints, for example.
Appropriate details and methods to increase the water penetration resistance of single-wythe hollow brick masonry
walls can be found in the Technical Note 7 Series and Technical Note 26.
In cases where water penetration resistance is critical, a drainage space should be provided on the interior of the
wall assembly. The interior may be furred out and insulation and gypsum board attached. Flashing and weeps
are used to drain the space. Another precaution may be the use of a water-resistant membrane placed on the
inside face of the wall. Waterproof membranes or polyethylene sheets have been used to resist water that has
penetrated the hollow brick wall. Any puncture in the membrane must be properly sealed.
Sound Resistance
Because sound insulation increases with increasing wall weight, brick masonry provides very good sound
penetration resistance. The sound transmission class (STC) rating is used to determine the sound insulation of
walls. Hollow brick veneer generally has an STC of approximately 40 to 45, slightly less than the STC of 45 for
solid brick veneer of the same thickness. The STC for through-the-wall brick units is typically calculated as a linear
function of weight. A grouted 8-in. (203-mm) brick wall generally has an STC of approximately 50 to 55. In the
absence of test results for a particular wall, calculated values of STC ratings can be determined from the following
equation:
STC = 19.6W
0.23
(Imperial) or 913.6W
0.23
(SI)
The STC rating is a function of the weight of the wall, W, expressed in pounds per cubic foot (kg/m
3
). This
equation is a best-fit curve based on the average of historical test data [Ref. 7].
DESIGN AND DETAILING
Mortar Bedding
Requirements for brickwork constructed of hollow brick vary depending on the intended use of the brickwork.
For larger hollow brick units used in structural (non-veneer) walls, mortar should be applied to the full thickness
of the face shell (face-shell bedding). For smaller hollow brick units that are used in veneer applications, mortar
should be applied to the full width of the brick veneer (full bedding) to maintain proper anchor embedment and
cover.
Reinforcement
Although reinforcement is not always used in hollow brick masonry, the large cells allow the units to be easily
reinforced and grouted. The reinforcing must be embedded in grout, not mortar. Reinforcing is most often
positioned in the center of the wall but may be placed to one side to maximize the distance from the compression
face. Reinforcement is grouted into hollow brick walls to increase the flexural strength, to provide ductility and
to carry tensile forces. The flexural strength of a reinforced hollow brick wall depends primarily on the amount of
vertical reinforcement because the compressive strength is rarely the limiting factor. The reinforcement resists the
flexural tension and the brickwork resists the flexural compression. Building codes may dictate a minimum amount
of reinforcement for improved ductility in seismic regions. In reinforced masonry design, any tension resistance
provided by the masonry is neglected.
Corbelling and Other Design Details
Certain construction details require the use of solid masonry units while others require either solid units or solidly
filled hollow units. The use of hollow brick is restricted by the model building codes as follows:
Corbelling. For corbelling, use solid units or solidly filled hollow units with approval of building official.
Masonry Piers. The height of masonry piers constructed of unfilled hollow units is limited to four times their least
dimension.
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 6 of 9
Parapet Walls. An unreinforced hollow unit masonry parapet must be no less than 8 in. (203 mm) thick, and its
height must not exceed three times its thickness.
Fireboxes and Chimneys. Fireboxes and chimneys constructed of hollow units are required to be grouted solid.
HOLLOW BRICK SPECIFICATION AND SIZES
Hollow brick should be specified to comply with ASTM C652, Standard Specification for Hollow Brick (Hollow
Masonry Units Made from Clay or Shale [Ref. 2]. When initially issued in 1970, ASTM C652 covered units with
void areas up to and including 40 percent in any plane parallel to the surface containing the voids. In 1987, the
standard was modified to allow void areas up to and including 60 percent of the units gross area.
Grade
Two Grades exist in ASTM C652: Grades SW and MW. As with solid units (governed by ASTM C216, Standard
Specification for Facing Brick [Ref. 2]), the Grade establishes requirements to ensure adequate freeze/thaw
resistance. Grade SW units provide high and uniform resistance to frost action while saturated with water. Grade
MW units are intended for applications that are unlikely to be saturated with water when exposed to freezing
temperatures. When the Grade is not specified, ASTM C652 stipulates that the requirements for Grade SW govern.
The physical property requirements are shown in Table 2. Two alternates exist in the specification to demonstrate
durability without meeting the requirements of Table 2. Brick with a cold water absorption less than 8 percent are
exempt from the saturation coefficient requirements. Brick passing a 50-cycle freezing and thawing test are exempt
from the boiling water absorption and saturation coefficient requirements. Brick meeting Table 2 requirements or
the cold water absorption alternate are not required to be subjected to the freeze/thaw test.
TABLE 2
ASTM C652 Physical Property Requirements for Hollow Brick
Designation
Compressive Strength, Gross Area,
min.
1
, psi (MPa)
Five-hour Boiling Water Absorption,
max., %
Saturation Coefficient,
max.
Average of 5 Individual Average of 5 Individual Average of 5 Individual
Grade SW 3000 (20.7) 2500 (17.2) 17.0 22.0 0.78 0.80
Grade MW 2500 (17.2) 2200 (15.2) 22.0 25.0 0.88 0.90
1. Unit in stretcher position with load applied perpendicular to bed surface.
Type
Four Types of hollow brick are defined by ASTM C652: Types HBS, HBX, HBA and HBB. Each of these Types
relates to the appearance requirements for the brick. Dimensional variation, chippage, warpage and other
imperfections are qualifying conditions of Type. The most common, Type HBS, is considered to be standard and
is specified for most applications. When the Type is not specified, ASTM C652 stipulates that the requirements
for Type HBS govern. Type HBX brick are specified where a higher degree of precision is required. Type HBA
brick are unique units that are specified for nonuniformity in size or texture. Where a particular color, texture or
uniformity is not required, Type HBB brick can be specified (these applications are typically unexposed locations).
Class
The extent of void area of hollow brick is separated into two Classes: H40V and H60V. Brick with void areas
greater than 25 percent but not greater than 40 percent of the units gross cross-sectional area in any plane
parallel to the surface containing the voids are classified as Class H40V. Brick with void areas greater than 40
percent but not greater than 60 percent of the gross cross-sectional area are classified as Class H60V. When the
Class is not specified, ASTM C652 stipulates that the requirements for Class H40V govern.
Hollow Spaces (Voids)
Hollow spaces may be cores, cells, deep frogs or combinations of these. In ASTM C652, a core is defined as
a void having an area equal to or less than 1 in. (968 mm), while cells are voids larger than a core. A deep
frog is an indentation in the bed surface of the brick that is deeper than in. (9.5 mm). The thickness of face
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 7 of 9
shells and webs are limited by ASTM C652. Figure 1 and Table 3 define the nomenclature associated with hollow
brick units and the minimum required thickness of face shells and cross webs.
TABLE 3
ASTM C652 Hollow Brick Cross-Sectional Requirements
Type of Void
Minimum Distance from
Void to Exposed Edge,
1, 2
in. (mm)
Minimum Distance from
Void to Unexposed
Edge,
3
in. (mm)
Minimum Web
Thickness (Between
Void and Core), in. (mm)
Minimum Web
Thickness (Between
Void and Cell), in. (mm)
Core (15.9) (12.7) (6.4) (9.5)
Cell (19.1) (12.7) (9.5) (12.7)
Additional Requirements for Class H60V Units
Nominal Width of Units,
in. (mm)
Minimum Solid Face-Shell
Thickness, in. (mm)
Minimum Cored or Double
Face-Shell Thickness, in. (mm)
Minimum End-Shell or End
Web Thickness, in. (mm)
3 and 4 (76 and 102) (19.1) N/A (19.1)
6 (152) 1 (25.4) 1 (38) 1 (25.4)
8 (203) 1 (32) 1 (38) 1 (25.4)
10 (254) 1 (35) 1 (41) 1 (29)
12 (305) 1 (38) 2 (51) 1 (29)
1. Cored-shell hollow brick with cores greater than 1 in. (650 mm) in cored shells shall be not less than in. (12.7 mm) from any edge. Cores
not greater than 1 in. (650 mm) in shells cored not more than 35 percent shall be not less than in. (9.5 mm) from any edge.
2. Double-shell hollow brick with inner and outer shells not less than in. (12.7 mm) are permitted to have cells not greater than in.
(15.9 mm) in width nor 5 in. (127 mm) in length between the inner and outer shell.
3. Permitted where recess in unexposed edge is in. (12.7 mm) or greater.
TABLE 4
Typical Nominal Hollow Brick Sizes
Unit Designations
1
Thickness, in. (mm) Height, in. (mm) Length, in. (mm)
Queen 3 (76) 2 or 3
1
5 (68 or 81) 8 (203)
King 3 (76) 2 or 3
1
5 (68 or 81) 10 (254)
Modular, Engineer Modular 4 (102) 2 or 3
1
5 (68 or 81) 8 (203)
Engineer Norman, Utility 4 (102) 3
1
5 or 4 (81 or 102) 12 (305)
Meridian, Double Meridian 4 (102) 2, 4 or 8 (68, 102 or 203) 16 (406)
4 (102) 8 (203) 8 (203)
5 (127) 3
1
5 (81) 10 (254)
6 (152) 4 (102) 12 (305)
6" Through-Wall Meridian 6 (152) 4 or 8 (102 or 203) 16 (406)
8 (203) 3
1
5 or 8 (81 or 203) 12 (305)
8" Through-Wall Meridian, Double
Through-Wall Meridian
8 (203) 4 or 8 (102 or 203) 16 (406)
12 (305) 4 (102) 16 (406)
1. Unit designations given are standardized nomenclature and encompass the vast majority of current brick production. Additional sizes may be
available from individual or regional manufacturers. Refer to Technical Note 10B for a complete list of standardized designations.
The dimensions of the unit and the configuration of its voids are critical for reinforced brick masonry. The cells
intended to receive reinforcement must align so that reinforcing bars can be properly placed. Most Class H60V
hollow brick contain two cells that are aligned when laid in running and stack bonds. Other bond patterns, such as
one-third bond and bonds at corners, may require different unit configurations to permit placement of reinforcement.
Size of cores will also influence grout type and grout placement methods. It is advisable to check with the brick
manufacturer to determine the coring patterns available.
Sizes and Shapes. Hollow brick are commonly available in a variety of sizes, as listed in Table 4. Hollow brick are
also made in a variety of special shapes. Special shapes include radial, bullnose, interior and exterior angled corner
units and others. Bond beam units are often used where horizontal reinforcing is required. They may be specially
made at the plant or cut on site. The brick manufacturer should be consulted for the availability of special shapes.
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 8 of 9
CONSTRUCTION REQUIREMENTS
Mortar Bedding
In structural (non-veneer) applications, hollow brick units are typically laid with face-shell bedding. Face-shell
bedding consists of mortar coverage on the inner and outer face shells of the unit. Cross webs or end webs of the
unit may require mortar bedding when grout must be confined within certain cells of partially grouted masonry or
on the first course of brickwork.
In veneer applications, hollow units should be laid in full mortar beds. Field experience has demonstrated that
a veneer constructed of hollow brick units with a nominal thickness of 3 to 4 in. (76 to 102 mm) and laid in a full
mortar bed has not significantly increased mortar usage compared to the same veneer constructed of solid brick
units. Care should be taken to avoid using excessively plastic mortar or placement methods that would force
excessive amounts of mortar into the cells or cores of the brick below. If these steps are taken, the rule of thumb to
use seven bags per thousand brick for estimating mortar usage is valid for most hollow brick veneer applications.
Anchors and Ties
In some loadbearing and all veneer applications, hollow brick masonry is connected to either another wythe of
masonry or to some other structural system. For face-shell bedded hollow brick masonry, the rectangular wire
tie or joint reinforcement used must be embedded across one face shell of the hollow masonry and at least
in. (12.7 mm) into the other face shell, as depicted in Figure 3. For hollow masonry veneer, where full mortar
bedding is required, anchors must be fastened to the backing and embedded into the mortar a minimum of
1 in. (38 mm), as depicted in Figure 4. Wire ties, joint reinforcement or sheet-metal ties are used for veneer
applications. When a backing of wood stud framing is used, corrugated sheet-metal anchors can be used to
anchor hollow masonry veneer. Both wall ties and veneer anchors must be recessed from the exposed exterior
face of the mortar by a minimum of in. (15.9 mm).
SUMMARY
Hollow brick are a natural evolution of clay brick manufacturing, providing durable brick units with less raw
material. For typical veneer applications, hollow brick can provide a water-resistant drainage wall system. For
single-wythe applications, the cells of larger hollow brick units can be reinforced to provide a structural solution.
In all applications, hollow brick provide the durability, aesthetics, fire resistance, thermal resistance and overall
performance characteristic of clay brick masonry.
Wire Tie or Joint
Reinforcement
Hollow Brick
Masonry Wythe
Additional
Masonry
Min. in.
(12.7 mm)
Embedment
1
2 /
Min. in.
(15.9 mm) Cover
5
8 /
Hollow Brick
Veneer
Backing
Veneer Anchor
(Style Varies)
Min. 1 in.
(38 mm)
Embedment
1
2 /
Min. in.
(15.9 mm) Cover
5
8 /
Figure 3
Wall Tie Placement in Hollow Brick Masonry
Figure 4
Anchor Placement in Hollow Brick Veneer
www.gobrick.com | Brick Industry Association | TN 41 | Hollow Brick Masonry | Page 9 of 9
The information and suggestions contained in this Technical Note are based on the available
data and the experience of the engineering staff and members of the Brick Industry
Association. The information contained herein must be used in conjunction with good technical
judgment and a basic understanding of the properties of brick masonry. Final decisions on the
use of the information discussed in this Technical Note are not within the purview of the Brick
Industry Association and must rest with the project architect, engineer and owner.
REFERENCES
1. 2006 International Building Code, International Code Council, Inc., Country Club Hills, IL, 2006.
2. Annual Book of Standards, Vol. 04.12, ASTM International, West Conshohocken, PA, 2007:
C216 Standard Specification for Facing Brick (Solid Masonry Units Made From Clay or Shale)
C652 Standard Specification for Hollow Brick (Hollow Masonry Units Made From Clay or Shale)
3. Brown, R.H., and Borchelt, J.G., Compression Tests of Hollow Brick Units and Prisms, Masonry:
Components to Assemblages, ASTM STP 1063, J.H. Matthys, ed., ASTM, Philadelphia, PA, 1990.
4. Building Code Requirements for Masonry Structures (ACI 530-05/ASCE 5-05/TMS 402-05), The Masonry
Society, Boulder, CO, 2005.
5. Sanders, J.P., and Brosnan, D.A., The Effect of Void Area on Brick Masonry Performance, Journal of
ASTM International, Volume 4, Issue 1, ASTM International, West Conshohocken, PA, 2007.
6. Southwest Research Institute, Fire Performance Evaluation of Load-Bearing Brick-Veneer Wall Assemblies
Tested in Accordance with ASTM E119-00a, Standard Test Methods for Fire Tests of Building Construction
and Materials, San Antonio, TX, 2007.
7. Standard Method for Determining the Sound Transmission Class Rating for Masonry Walls, (TMS 0302),
The Masonry Society, Boulder, CO, 2007.
8. Whitlock, A.R., and Brown, R.H., Compressive Strength of Grouted Hollow Brick Prisms, Masonry:
Materials, Properties, and Performance, ASTM STP 778, J.G. Borchelt, ed., ASTM, Philadelphia, PA,
1982.

Technical Notes 42 - Empirical Design of Brick Masonry
November 1991
Abstract: This Technical Notes provides requirements for the empirical design of masonry structures. These
requirements are based on past proven performance. The provisions are taken from ACI 530-92/ASCE 5-92,
"Building Code Requirements for Masonry Structures", Chapter 9. Subjects discussed pertaining to ACI 530/ASCE 5
are: lateral stability; allowable stresses; lateral support; thickness of masonry; bonding; anchorage and
miscellaneous requirements. Seismic considerations and material requirements are also included.

Key Words: brick, building codes, design standards, empirical design, masonry, stresses.

INTRODUCTION

Empirical design is a procedure for sizing and proportioning masonry elements to form an entire structure or parts of
a structure. Empirical design does not require a rational analysis. It is based on rules of thumb and formulas
developed over many years of experience. This design method has been used successfully for many decades.
Empirical design is generally used for buildings of a small scale nature. The basic premise is that masonry walls are
incorporated into two directions of the building along with floor and roof systems for lateral support.
Chapter 9 of ACI 530-92/ASCE 5-92 is devoted solely to empirical design procedures. The provisions of earlier
empirical standards have been modified to reflect contemporary construction materials and methods. Many
requirements remain the same as earlier standards but new restrictions have been added to reflect recent
developments.
The current model building codes contain requirements for empirical design of masonry. Until the development of ACI
530/ASCE 5, most of the model building code empirical design procedures were based on the ANSI A41.1 (R1970)
document which has been discontinued.
It is the purpose of this Technical Notes to review many of the pertinent design and construction requirements
included in Chapter 9 of ACI 530/ASCE 5. In this Technical Notes, sections of ACI 530/ASCE 5 referenced are
given in parenthesis.

SCOPE (9.1)

Chapter 9 of ACI 530/ASCE 5 covers empirical design criteria which can be used for masonry components and
masonry buildings in lieu of the design requirements in Chapters 5, 6, 7 and 8. Chapters 5 through 8 contain a
rational design for masonry based on the working stress method. The scope of Chapter 9 has three basic
restrictions that have not been incorporated in other previous empirical procedures: 1) buildings cannot be located in
Seismic Zones 3 and 4 as defined in ASCE 7-88, "Minimum Design Loads for Buildings and Other Structures"
42 http://www.gobrick.com/BIA/technotes/t42.htm
1 of 15 9/13/2009 2:08 PM
(formerly referred to as ANSI A58.1); 2) lateral load forces, i.e. wind loads, are restricted to a maximum of 25 psf
(1.2 kPa) as referenced in ASCE 7; 3) buildings relying on masonry walls for lateral load resistance cannot exceed
35 ft (10.7 mm) in height.
Chapter 9 permits empirical design of masonry elements not acting as a portion of the lateral force resisting system
even though the main lateral force resisting system is rationally designed by other chapters contained in ACI
530/ASCE 5. Further, Chapter 9 can be used to design masonry elements in frame structures.

DESIGN

Consideration of lateral stability and lateral support are of prime importance in empirical design. Compressive
stresses, thickness of masonry, bonding and anchorage requirements are incorporated in this design methodology.

Lateral Stability (9.3)
Shear walls are necessary when lateral support is provided by masonry construction. Shear walls must be provided
in two directions, parallel and perpendicular to the assumed direction of the lateral load resisted. The minimum
cumulative length of shear walls in any one direction must be at least 40 percent of the long dimension of the
building. Portions of walls with openings cannot be included when determining the cumulative length of shear walls.
BIA recommends that the cumulative length of shear walls include only wall lengths greater than or equal to one-half
the story height of the building. Bearing walls are permitted to serve as shear walls. Shear walls must be a minimum
nominal thickness of 8 in. (200 mm). Figure 1 provides an example calculation to determine the cumulative shear wall
length.

42 http://www.gobrick.com/BIA/technotes/t42.htm
2 of 15 9/13/2009 2:08 PM

MINIMUM CUMULATIVE LENGTH OF SHEAR WALLS = 0.4 X LONG DIMENSION
MINIMUM CUMULATIVE LENGTH = 0.4 X 36 FT = 14.4 FT
X-DIRECTION = 2 ( 6 + 6 + 6 + 6 ) = 48 FT > 14.4 FT OK
Y-DIRECTION = 2 ( 24 + 10 +10 ) = 88 FT > 14.4 FT OK

Cumulative Length of Shear Walls
FIG. 1

Shear wall spacing requirements are based on the type of floor or roof construction used in the building under
consideration. When using stiffer elements such as cast-in-place concrete floors, the shear wall spacings are
greater. Table 1 provides the maximum ratio of shear wall spacing to shear wall length based on the type of floor or
roof construction.

42 http://www.gobrick.com/BIA/technotes/t42.htm
3 of 15 9/13/2009 2:08 PM


Allowable Stresses (9.4)

Allowable compressive stresses permitted in Chapter 9 are given in Table 2. Compressive stress calculations are
based on gross area, not minimum net area as is the case in the rational analysis chapters of ACI 530/ASCE 5.
Gross area is based on the actual dimensions of the masonry unit under consideration. When multi-wythe walls are
used in construction, the allowable stress taken from Table 2 should be based on the weakest combination of the
unit and mortar used for each wythe.
The allowable stresses in Table 2 are considered as allowable average stresses, not maximum fiber stresses.
These allowable stresses only pertain to vertically applied loads reasonably centered on the wall. Any influence of an
eccentrically applied load is limited by the minimum wall thickness and maximum lateral support requirements.

42 http://www.gobrick.com/BIA/technotes/t42.htm
4 of 15 9/13/2009 2:08 PM

1
Linear interpolation for determining allowable stresses for masonry units having compressive strengths which are intermediate between those given in the
table is permitted
2
1 psi = 6.9 kPa
3
Where floor and roof loads are carried upon one wythe, the gross cross-sectional area is that of the wythe under load; if both wythes are loaded, the gross
cross-sectional area is that of the wall minus the area of the cavity between the wythes. Walls bonded with metal ties shall be considered as non-composite
walls unless collar joints are filled with mortar or grout.

Lateral Support (9.5)

42 http://www.gobrick.com/BIA/technotes/t42.htm
5 of 15 9/13/2009 2:08 PM
Chapter 9 contains arbitrary limits on the ratios of wall thickness to distance between lateral supports. These limits
provide controls on the flexural tension stresses within the wall and limit possible buckling under compressive
stresses. Maximum h/t or l/t ratios and minimum thickness used for determining distance between lateral supports
are consistent with past masonry standards. Definitions for height (h), length (l) and thickness (t) for use in the
allowable lateral support ratios are as follows: h = the vertical distance or height between lateral supports; I = the
horizontal distance or length between lateral supports: and t = the nominal thickness of the masonry wall under
consideration. ACI 530/ASCE 5 does not provide guidance for computing the thickness of masonry bonded hollow
walls or cavity walls bonded with metal ties. BIA suggests that the value for thickness be the sum of the nominal
thicknesses of the inner and outer wythes.
Masonry walls should be laterally supported in either the horizontal or vertical direction at intervals not exceeding
those given in Table 3. Lateral support should be provided by cross walls, pilasters, buttresses or structural frame
members when the limiting distance is taken horizontally. Floors, roofs or structural frame members should be used
when the limiting distance is taken vertically (see Fig. 2).



42 http://www.gobrick.com/BIA/technotes/t42.htm
6 of 15 9/13/2009 2:08 PM

Lateral Support Requirements
FIG. 2

Cantilever type walls also have a minimum lateral support criteria. The h/t ratio for cantilever walls should not exceed
6 for solid masonry walls nor 4 for hollow masonry walls.

Thickness of Masonry (9.6)
Empirical design requirements pertaining to the thickness of bearing walls and foundation walls are found in Section
9.6. Masonry walls must conform to thickness requirements as well as lateral support and allowable stress
requirements. Thicknesses given are nominal dimensions. These requirements are more conservative than empirical
design criteria in previous masonry standards.
Bearing Walls. The minimum thickness of masonry bearing walls more than one story in height must be 8 in. (200
mm). Bearing walls of one story buildings may be reduced to 6 in. (150 mm). The height to thickness limitation in
Table 3 requires a wall of 6 in. (150 mm) in thickness to have a maximum height of 10 ft (3.1 m)
Specific provisions are incorporated due to a change in wall thickness between floor levels or floor and roof levels. If
a change in wall thickness between floors or between floor and roof levels is desired, the greater wall thickness
42 http://www.gobrick.com/BIA/technotes/t42.htm
7 of 15 9/13/2009 2:08 PM
must extend to the lower support level. Wall thicknesses may be changed to meet fire, sound or thermal
requirements. Where walls of hollow masonry units or masonry bonded hollow walls are decreased in thickness, a
course or courses of solid masonry should be constructed between the thicker wall below and the thinner wall
above. Special units or construction are permitted to be used as long as the loads from face shells or wythes of
masonry above are transmitted to the wall system below.
Foundation Walls. Foundation walls have empirical thickness requirements which are shown in Table 4. Foundation
walls must be constructed of either Type M or S mortar. The height of unbalanced fill (height of finished ground
above the basement floor or inside ground level) and the height of the wall between lateral supports must not exceed
8 ft (2.4 m), and the equivalent fluid weight of unbalanced fill must not exceed 30 pcf (480.5 kg/m3). Most
well-drained sand and gravel backfills have an equivalent fluid weight of less than 30 pcf (480.5 kg/m3). When these
conditions are not met, foundation walls must be designed in accordance with Chapters 5 and 6 or 5 and 7 of ACI
530/ASCE 5.


1
1 in = 25.4 mm
2
1 ft = 0.3048

Parapets. Parapets are required to have a minimum thickness of at least 8 in. (200 mm). Their height cannot exceed
three timed their thickness.

Bonding (9.7)
Multi-wythe masonry walls may be bonded together by either masonry headers or metal wall ties. Limitations on the
area and spacing of masonry headers or metal ties for both solid and hollow units are contained in Chapter 9 of the
code.
Masonry headers are typically used when bonding barrier type walls (walls of solid units built without air spaces) or
hollow walls composed of solid masonry units. Metal ties can be used for barrier type walls (with grouted collar
joints) and drainage type walls (a clear air space between wythes of masonry).
42 http://www.gobrick.com/BIA/technotes/t42.htm
8 of 15 9/13/2009 2:08 PM
The necessary requirements for bonding multi-wythe walls with masonry headers is shown in Fig. 3. Masonry
headers of solid units must comprise not less than 4 percent of wall surface area and extend at least 3 in. (75 mm)
into each wythe. The distance between adjacent full-length headers should not exceed 24 in. (610 mm) horizontally
or vertically along the wall surface.


Multi-Wythe Bond With Masonry Headers
FIG. 3
Two options exist when bonding multi-wythe walls with metal ties, the use of unit metal ties and the use of
prefabricated horizontal joint reinforcement. When using unit metal ties, such as Z-ties or rectangular ties (box ties),
one tie must be provided for each 4 1/2 ft
2
(0.42 m
2
) of wall area. Ties should be at least 3/16 in. (4.76 mm) in
diameter and be corrosion resistant. The maximum vertical distance between ties should not exceed 24 in. (610
mm), and the maximum horizontal distance should not exceed 36 in. (914 mm). Z-ties may not be used with hollow
masonry units. Additional metal ties should be provided at all openings, spaced not more than 3 ft (0.91 m) apart
around the perimeter and within 12 in. (300 mm) of the opening. These provisions are similar to those for cavity wall
construction.
When bonding multi-wythe walls with horizontal joint reinforcement, there should be one crosswire metal tie for each
2 2/3 ft
2
(0.25 m
2
) of wall area. The vertical spacing should not exceed 16 in. (400 mm). Crosswires should not be
smaller than No. 9 gage wire (W 1.7) and be corrosion resistant.

Pattern Bond. Masonry walls can be laid in either running or stack bond. Running bond is defined by each wythe of
masonry head joints in successive courses being offset by at least one-quarter the unit length (see Fig. 4). It is
considered stack bond if the longitudinal bond is offset less than one-quarter the unit length, and horizontal joint
reinforcement or bond beams with a maximum spacing of 4 ft (1.2 m) vertically with a minimum area of steel equal to
0.0003 times the vertical cross-sectional area of the wall must be provided.

42 http://www.gobrick.com/BIA/technotes/t42.htm
9 of 15 9/13/2009 2:08 PM

Typical Foundation Details
FIG. 4

Anchorage (9.8)

Masonry elements must be anchored to various components of the building which provide lateral support when using
empirical design. Anchorage must occur at intersecting walls, at floors and roofs which adjoin masonry walls and
where masonry walls abut structural framing. Anchorage requirements for masonry walls contained in ACI
530/ASCE 5 are as follows:

"9.8.2 Intersecting walls - Masonry walls depending upon one another for lateral support shall be anchored
or bonded at locations where they meet or intersect by one of the following methods:
9.8.2.1 Fifty percent of the units at the intersection shall be laid in an overlapping masonry bonding pattern,
with alternate units having a bearing of not less than 3 in. (75 mm) on the unit below.
9.8.2.2 Walls should be anchored by steel connectors having a minimum section of 1/4 in. (6.4 mm) by 1
1/2 in. (38.1 mm) with ends bent up at least 2 in. (50 mm) or with cross pins to form anchorage. Such
anchors shall be at least 24 in. (600 mm) long and the maximum spacing shall be 4 ft (1.22 m).
9.8.2.3 Walls shall be anchored by joint reinforcement spaced at a maximum distance of 8 in. (200 mm).
Longitudinal rods of such reinforcement shall be at least 9 gage (W 1. 7) and shall extend at least 30 in.
(762 mm) in each direction at the intersection.
9.8.2.4 Interior non-loadbearing walls shall be anchored at their intersection, at vertical intervals of not more
than 16 in. (400 mm) with joint reinforcement or 1/4 in. (6.4 mm) mesh galvanized hardware cloth.
42 http://www.gobrick.com/BIA/technotes/t42.htm
10 of 15 9/13/2009 2:08 PM
9.8.2.5 Other metal ties, joint reinforcement or anchors, if used, shall be spaced to provide equivalent area
of anchorage to that required by this section.
9.8.3 Floor and roof anchorage - Floor and roof diaphragms providing lateral support to masonry shall be
connected to the masonry by one of the following methods:
9.8.3.1 Wood floor joists bearing on masonry walls shall be anchored to the wall at intervals not to exceed 6
ft (1.8 m) by metal strap anchors. Joists parallel to the wall shall be anchored with metal straps spaced not
more than 6 ft (1.8 m) on centers extending over or under and secured to at least 3 joists. Blocking shall be
provided between joists at each strap anchor.
9.8.3.2 Steel floor joists shall be anchored to masonry walls with 3/8 in. (9.5 mm) round bars, or their
equivalent, spaced not more than 6 ft (1.8 m) on center. Where joists are parallel to the wall, anchors shall
be located at joists cross bridging.
9.8.3.3 Roof structures shall be anchored to masonry walls with 1/2 in. (12.7 mm) bolts 6 ft (1.8 m) on
center or their equivalent. Bolts shall extend and be embedded at least 15 in. (381 mm) into the masonry, or
be hooked or welded to not less than 0.20 in
2
(129 mm2) of bond beam reinforcement placed not less than
6 in. (150 mm) from the top of the wall.

9.8.4 Walls adjoining structural framing - Where walls are dependent upon the structural frame for lateral
support they shall be anchored to the structural members with metal anchors or otherwise keyed to the
structural members. Metal anchors shall consist of 1/2 in. (12. 7 mm) bolts spaced at 4 ft (1.2 m) on center
embedded 4 in. (100 mm) into the masonry, or their equivalent area. "

Miscellaneous Requirements (9.9)
General limitations for masonry structures such as masonry over chases and recesses, lintels over openings,
noncombustible supports for masonry walls and corbeling have empirical requirements for proper design and
construction.
Chases and recesses in masonry walls are sometimes used for visual effects or to receive pipes, conduits or ducts.
When chases or recesses are wider than 12 in. (300 mm), the masonry above the chase must be supported by
noncombustible lintels, which could be steel angle lintels or reinforced brick masonry lintels.
The design of lintels must be in accordance with Section 5.6 which stipulates that the deflection of lintels due to
vertical loads should not exceed the span divided by 600 nor 0.3 in. (7.6 mm) when supporting unreinforced
masonry. Minimum bearing for lintels is 4 in. (100 mm) on each end of the masonry opening.
Masonry is not permitted to be supported by combustible construction, i.e. wood. Even though wood construction
may meet the deflection requirements for lintels, this restriction is a fire safety requirement.
Corbeling limitations are the same as those required by the model building codes used throughout the country. The
maximum corbeled projection beyond the plane of the wall should not be more than one-half of the wall thickness or
one-half the wythe thickness for hollow walls. The maximum projection of any single course of masonry should not
exceed one-half the unit height or one-third the unit thickness. Solid units are required for corbeled courses of
Masonry. Figure 5 illustrates these criteria for corbeling masonry.

42 http://www.gobrick.com/BIA/technotes/t42.htm
11 of 15 9/13/2009 2:08 PM

Corbeling Limitations
FIG. 5

Seismic Considerations for Empirically Designed Masonry

Appendix A of ACI 530/ASCE 5 contains special requirements for masonry in seismic zones as specified in ASCE 7
(formerly ANSI A58.1). The provisions of Chapter 9 on empirical design of masonry may be used in Seismic Zones
0, 1 and 2, and are modified by Appendix A. Empirical design cannot be used in buildings located in Seismic Zones 3
and 4.
For Seismic Zones 0 and 1, all provisions of Chapter 9 apply without modification. There are no restrictions on
materials or design methods since these areas of the country represent low seismic risk.
Masonry elements in Seismic Zone 2 must meet more stringent requirements. Connections are strengthened and
minimum vertical and horizontal reinforcement is required in order to provide more ductility in the structure. All
materials also are permitted to be used in the structure.
Seismic requirements for buildings or structures in Seismic Zone 2 as contained in Appendix A are as follows:

"A.3.5 Veneer and units not specifically intended for structural use shall not be designed to resist loads
other than their own weight or their own shear loads.
A.3.6 Masonry walls shall be anchored to all floors and roofs which provide lateral support for the walls.
Such anchorage shall provide direct connection capable of resisting horizontal forces required in Section
5.2 or a minimum of 200 lb (90.9 kg) per lineal foot (meter) of wall, whichever is greater. Walls shall be
designed to resist bending between anchors where anchor spacing exceeds 4 ft (1.2 m). Anchors in
masonry walls shall be embedded in reinforced bond beams or reinforced vertical cells.
A.3.7 Structural members framing into or supported by masonry columns shall be anchored thereto. Anchor
bolts located in the tops of columns shall be set entirely within the reinforcing cage composed of column
42 http://www.gobrick.com/BIA/technotes/t42.htm
12 of 15 9/13/2009 2:08 PM
bars and lateral ties. A minimum of two #4 lateral ties shall be provided in the top 5 in. (127 mm) of the
column. Welded or mechanical connections for reinforcing bars in tension shall develop 125 percent of the
yield strength of the bars in tension.
A.3.8 Vertical reinforcement of at least 0.20 in.
2
(129 mm
2
) in cross-sectional area shall be provided
continuously from support to support at each corner, at each side of each opening and at the ends of walls.
Horizontal reinforcement not less than 0.20 in.
2
(129 mm
2
) in cross section area shall be provided: (1) at the
bottom and top of wall openings and shall extend not less than 24 in. (610 mm) nor less than 40 bar
diameters past the opening, (2) continuously at structurally connected roof and floor levels and at the top of
walls, (3) at the bottom of the wall or in the top of the foundations when doweled to the wall, (4) at maximum
spacing of 10 ft (3.1 m) unless uniformly distributed joint reinforcement is provided. Reinforcement at the
top and bottom openings when used in determining the maximum spacing specified in Item No. (4) above
shall be continuous in the wall. "

Minimum reinforcement requirements are shown in Fig. 6.



Minimum Reinforcement Requirements for Seismic Zone 2
FIG. 6

"A.3.9 Where head joints in successive courses are horizontally offset less than one-quarter of the unit
length, the minimum horizontal reinforcement shall be 0.0007 times the gross cross-sectional area of the
wall. This reinforcement shall be satisfied with uniformly distributed joint reinforcement or with horizontal
reinforcement spaced not over 4 ft (1.2 m) and fully embedded in grout or mortar "

MATERIALS AND CONSTRUCTION
General
42 http://www.gobrick.com/BIA/technotes/t42.htm
13 of 15 9/13/2009 2:08 PM

The provisions of ACI 530.1-92/ASCE 6-92, "Specifications for Masonry Structures" have minimum material and
construction requirements for masonry structures designed in accordance with Chapter 9 empirical provisions.
Masonry units, mortar, grout, reinforcement and accessories are included. This document should be referenced in
the project specifications and can be modified as required for the particular project.

Masonry Units
The products section permits the use of clay brick, concrete masonry units and stone masonry in empirically
designed masonry structures. ASTM standards for clay or shale masonry covered by ACI 530.1/ASCE 6 are ASTM
C 34, C 56, C 126, C 212, C 216 and C 652. Grade or class of the units to be used in construction are determined
by exposure conditions and required durability. For further information on the manufacture, designation and selection
of clay masonry units, see Technical Notes 9 Series.

Mortar and Grout
Mortar is required to conform to ASTM C 270 Mortar for Unit Masonry. When job site pigments are used to color
mortar there are maximum percentages of color pigment by weight of the cement content which can be added. For
portland cement-lime mortars, the maximum content of the coloring pigment is limited to 10 percent for mineral oxide
pigments and 2 percent for carbon black. If masonry cements are used, the percentage by weight for color pigments
are halved.
Grout is required to conform to ASTM C 476 Grout for Unit Masonry. This is a proportion specification for either fine
or coarse grout used in construction.

Reinforcement and Accessories
ACI 530.1/ASCE 6 contains provisions for reinforcement and metal accessories. All reinforcement and metal
accessories are required to be corrosion resistant. Procedures described represent current construction practices
and are consistent with model building codes now in existence. Topics that are covered are ASTM standards for the
materials, inspection, and detailing and placement of reinforcement and accessories which include tolerances.
Corrosion Resistance. Conventional corrosion protection methods attempt to protect metals embedded in masonry
by isolating them with impervious coatings, by using metals that are corrosion resistant or by providing cathodic
protection. ACI 530.1/ASCE 6 provides requirements for corrosion protection for carbon steel by galvanized
coatings. The amount of galvanizing required increases with the severity in exposure of the masonry wall. Anchors,
ties and joint reinforcement must meet minimum corrosion protection requirements. Table 5 shows the minimum
corrosion protection requirements needed for metal accessories used in masonry walls.

42 http://www.gobrick.com/BIA/technotes/t42.htm
14 of 15 9/13/2009 2:08 PM


Construction

Construction requirements within ACI 530.1/ASCE 6 cover the conventional construction practices used in projects
that involve empirically designed masonry. The provisions are similar to those found in the model building codes. The
basic premise under the construction requirements is to ensure proper placement of materials. Mortar joint filling
depends on the type of unit used in construction. Solid units have full head and bed joints. Hollow units are laid with
face shell bedding. Requirements include tolerances for erection, collar joints and placement of embedded items
such as wall ties and reinforcement. Grout placement is also covered.

SUMMARY

This Technical Notes reviews empirical design procedures contained in ACI 530/ASCE 5. The discussion centers on
the requirements which are needed by engineers and architects to fully understand the empirical design of masonry
structures within the limits of Chapter 9.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Institute of America. The information contained in this publication
must be used with good technical judgment. Final decisions on the use of materials and suggestions contained herein
are not within the purview of the Brick Institute of America and must rest with the project architect, engineer and
owner.
42 http://www.gobrick.com/BIA/technotes/t42.htm
15 of 15 9/13/2009 2:08 PM

Technical Notes 43 - Passive Solar Heating with Brick Masonry - Part 1 Introduction
June 1981
Abstract: Brick masonry passive solar energy systems can be used to significantly reduce the use of fossil fuels for
heating and cooling buildings. The basic concepts and necessary considerations for the design of passive solar
heating systems are discussed. The basic concepts involve the incorporation of the passive solar heating system into
the architectural design of the intended use and operation of the building. Consideration of environmental factors is
also discussed.
Key Words: attached sunspaces, bricks, buildings, cavity wall systems, climatology, conservation, direct gain
systems, energy, masonry, passive solar heating systems, solar radiation, system operation, thermal storage
walls.
INTRODUCTION
Energy conservation and fuel consumption have become a major concern in recent years. Much of the nation's fuel is
used in the heating of buildings. The use of solar heating systems will help to reduce this consumption of
non-renewable energy resources. Solar energy is an immediately available renewable energy source. Most buildings
can easily be designed to benefit from solar heating.
Two types of solar energy systems may be used to heat buildings, active and passive. Active solar heating systems
are those which require mechanical equipment for operation. Pumps and other mechanical devices are required to
circulate liquids or gases through solar collectors, to storage media, and then to transfer the collected heat to the
occupied spaces of the building.
Passive solar heating systems do not require the use of mechanical equipment. The heat flow in passive solar
heating systems is by natural means: radiation, convection, and conductance. The thermal storage is in the structure
itself. Although passive solar heating systems do not require mechanical equipment for operation, this does not mean
that fans or blowers may not, or should not, be used to assist the natural flow of thermal energy. The passive
systems assisted by mechanical devices are referred to as ''hybrid" heating systems.
Passive solar systems utilize basic concepts incorporated into the architectural design of the building. They usually
consist of: buildings with rectangular floor plans, elongated on an East-West axis; a glazed South-facing wall; a
thermal storage media exposed to the solar radiation which penetrates the South-facing glazing; overhangs or other
shading devices which sufficiently shade the South-facing glazing from the summer sun; and windows on the East
and West walls, and preferably none on the North walls. Passive solar systems do not have a high initial cost or
long-term payback period, both of which are common with many active solar heating systems.
This Technical Notes introduces the general features and requirements for the development and application of
passive solar heating systems. Passive solar cooling systems are discussed in Technical Notes 43C. Due to the
variations in building type and environment which must be considered, it is not normally feasible for passive solar
systems to be the sole source of heat in most climatological areas. Construction details are provided in Technical
Notes 43G.
43 http://www.gobrick.com/BIA/technotes/t43.htm
1 of 15 9/13/2009 2:11 PM
Passive Solar Building with Thermal Storage Wall Under Construction
FIG. 1
Combined Thermal Storage Wall System and Attached Sunspace
FIG. 2
ENVIRONMENTAL DATA AND REQUIREMENTS
Many environmental factors must be considered to fully utilize the concepts of passive solar heating systems.
Environmental data is given in Tables 1 and 2 of this Technical Notes.
Temperature
Exterior design temperatures are important considerations in developing passive solar heating systems. The size of
the system will depend upon daily, monthly and annual temperature fluctuations. In mild, sunny climates, the required
glazing and thermal storage areas may be relatively small. In temperate, cloudy climates, the required glazing area
may be small, but the thermal storage requirements may be greater. In colder climates, the amount of glazing and
thermal storage is usually large.
The average monthly heating degree days are related to exterior temperature conditions. These values are
necessary to determine the total monthly thermal load of the building. Average monthly heating degree days and
exterior temperatures are given in Table 2 at the end of this Technical Notes.
Latitude
Latitude is important to determine the amount of solar radiation and the appropriate summertime shading provided
by overhangs and other devices. The further North the building is to be located, the less winter solar radiation it will
receive. This is because the sun is above the horizon for a shorter period of time and the solar radiation must
penetrate more of the atmosphere. Values of solar radiation at various latitudes are given in Table 1.
At higher latitudes, the sun appears lower in the sky. At these latitudes, where the position (altitude) of the sun in the
sky is low, larger overhangs are required to shade the South-facing wall from the summer sunlight. Figure 3 shows
how the altitude of the sun changes from winter to summer, demonstrating how the South-facing wall may be shaded
from summer solar radiation and still be exposed to winter solar radiation by using an overhang. The length of
projection required to shade a South-facing wall from the summer sun is given in Table 3.
43 http://www.gobrick.com/BIA/technotes/t43.htm
2 of 15 9/13/2009 2:11 PM
Sun Altitude-Winter and Summer

aReprinted with permission from the 1972 ASHRAE Handbook of Fundamentals Volume. ASHRAE HANDBOOK & Product
Directory.
bProjection greater than 20 ft required.
Solar Radiation Data
Solar radiation data is required to determine the amount of radiation transmitted through the South-facing glazing.
Actual average solar radiation data for various geographical locations is given in Table 2. The amount of solar
radiation is dependent on climate, elevation and latitude. Clear day solar radiation for various latitudes is given in
Table 1.
Orientation
Orientation is extremely important in the design of passive solar buildings. The best performance will usually result
when the passive solar system faces true South. True South may be obtained from isogonic (magnetic variation)
charts developed by the United States Department of Commerce, Coast and Geodetic Survey, or by consulting a
local land surveyor.
When the passive solar system faces true South, the system will be exposed to the maximum amount of winter solar
radiation. Deviations of more than 30
o
East or West of true South are not recommended, especially where maximum
performance is desired.
Site Topography
The topography of the site is of major concern. If the South-facing wall of the building is shaded by natural or
man-made elements, it will probably not be feasible to consider passive solar systems. An ideal siting for a passive
solar building is to be bermed into a South-facing slope. This provides a South wall exposed to the sun, and a North
wall protected from environmental changes by the earth berm. Berming the North wall of the building should be done
cautiously to avoid problems caused by ground water and earth pressure.
BUILDING TYPE AND USE
In addition to environmental considerations building type and use are very important in developing and applying
passive solar heating systems. Building type and use are flexible requirements which allow the designer to make
43 http://www.gobrick.com/BIA/technotes/t43.htm
3 of 15 9/13/2009 2:11 PM
appropriate adaptations to the structure to provide the desired energy performance.
Thermal Load Requirements
Thermal load requirements are important in the selection and sizing of passive solar heating systems. The effects of
building type and use on the thermal load are determined by the interior design temperature and the allowable
temperature fluctuation. A warehouse may not require the same interior design temperature as a residential
structure. Many commercial buildings are only occupied during daylight hours and do not have to maintain the higher
interior working hour temperatures overnight. In many applications, the passive solar heating systems may provide
similar performance as conventional heating systems with night-time setbacks.
Another aspect which affects the requirements of the building's use is human comfort. Passive solar systems provide
conditions which contribute to human comfort. The brick storage areas of the system are warm. When surrounded
by warm surfaces, the human body receives radiation from the warm surfaces. This permits the occupants to feel
comfortable at lower interior air temperatures because heat is radiated to the body rather than from the body.
Glazing and Lighting Quality
The amount of natural lighting required will affect the selection of the type of passive solar heating system. Fabrics
and even the glazing material itself may suffer from ultraviolet degradation when exposed to direct sunlight. In
applications such as studios, admitting large quantities of diffuse solar radiation provides appropriate lighting.
The amount of glazing for most conventional structures is typically determined by the need or desire to provide
contact with the exterior or to meet building code egress requirements. This is not usually a primary design
consideration for the passive solar heating system
Material Properties
Massive brick masonry is recommended for thermal storage because of its inherent ability to store heat. Typically,
brick exposed to direct sunlight should be of a dark color wherever it is to perform as a thermal storage media. The
American Society of Heating, Refrigerating and air-conditioning Engineers (ASHRAE) defines dark colors as dark
blue, red, brown and green. The properties of brick as related to passive solar applications are discussed in
Technical Notes 43D.
System Operation
Passive solar heating systems may be shaded from the summer sun by fixed, adjustable or removable shading
devices. Adjustable or removable overhangs or shading devices require operation, but permit the optimum use of the
winter sun and can completely eliminate any solar exposure on the South-facing glass in the summer.
The performance of passive solar systems may be greatly enhanced by the use of night insulation. The insulation
may be applied on the interior in the form of drapes or panels. Insulation may also serve as reflector panels or
shading devices. Reflector-insulating panels may be hinged at the base of the South-facing glazing so that, when
opened during the day, they reflect additional solar radiation through the glazing and when closed, provide night
insulation. Night insulation may be operated manually or automatically.
Building Design and Appearance
There is no reason for passive solar heating systems to have an extremely unconventional design or appearance.
The only required variations are: additional South-facing wall glazing, reduced glazing on the East and West walls,
and preferably no glazing on the North wall; sufficient overhang or some other shading device to prevent the South-
facing glazing from being exposed to the summer sun; and interior brick masonry. The interior brick masonry
exposed to direct sunlight is used as the thermal storage component of the passive solar energy system. Additional
interior brick masonry unexposed to direct sunlight is used to provide a thermal flywheel which reduces interior
temperature fluctuations.
Spatial Requirements
The spatial requirements may dictate the type of system used. The depth of penetration of solar radiation into the
43 http://www.gobrick.com/BIA/technotes/t43.htm
4 of 15 9/13/2009 2:11 PM
structure may affect the system type selected. Buildings should be arranged with a longitudinal East-West
orientation to maximize the solar exposure of the South-facing glazing. This minimizes the distance from the South
wall to the North wall, across which the thermal energy from the passive solar energy system has to be distributed.
Building energy performance may be increased by heating the North wall with solar radiation entering through the
South-facing glazing.
DIRECT GAIN SYSTEMS
The direct gain system is simple and often used. The system consists of South-facing glazing which allows winter
sunlight to enter the habitable spaces of the building. This thermal energy is stored in brick floors and walls. A
schematic of a direct gain system is shown in Fig. 4. The South-facing glazing may be windows (operable or fixed),
or glass doors. The brick masonry exposed to the solar radiation should generally be a dark color and 4 to 8 in.
thick. All walls or other components not exposed to solar radiation should have light-colored surfaces.
In the direct gain system, the South-facing glazing permits sunlight to strike the brick masonry construction. The brick
masonry, because of its color, mass and thermal properties, provides the thermal storage for the system. The brick
masonry absorbs the thermal energy from the sunlight striking its surface. The heat, which is stored during the
daylight hours, is released gradually. The heat that is reflected from the brick masonry provides heat to the habitable
space during the daylight hours. The light-colored surfaces reflect the heat radiated or reflected from the brick
masonry to the air and surroundings in the habitable space. If large amounts of heat are required during the daytime
hours and less during night-time hours, this may be accomplished by using lighter colors of brick masonry.
Direct gain systems provide rapid temperature increases in the habitable space and may have large temperature
fluctuations. This is because such systems often must be designed to prevent overheating. The systems may have
limited amounts of brick masonry exposed to the winter sunlight. This is especially true in the lower latitudes where
the winter sun has a higher altitude. This may be overcome by providing clerestories to obtain solar radiation on the
North wall, as shown in Fig. 4.
Increased Building Depth Using Direct Gain System with Clerestory
FIG. 4
Ultraviolet degradation is of the greatest concern when direct gain systems are utilized. Materials subject to
ultraviolet degradation should not be exposed to direct sunlight. This may become an inconvenience in the living
areas heated by direct gain. The walls and floors exposed to the sunlight and used for thermal storage should not be
covered. Wall hangings and carpet greatly decrease the performance of the system.
THERMAL STORAGE WALL SYSTEMS
The thermal storage wall system, often referred to as a Trombe Wall System, is schematically represented in Fig. 5.
The thermal storage wall may be vented, as shown in Fig. 5, and provide heat by radiation and convection, or it may
43 http://www.gobrick.com/BIA/technotes/t43.htm
5 of 15 9/13/2009 2:11 PM
be unvented and supply heat by radiation alone. A thermal storage wall system is shown on the left of Fig. 2. It
consists of glazing, usually spaced 2 to 4 in. on the exterior of a South-facing wall, constructed of brick masonry.
The massive brick wall, usually 10 to 18 in. thick, may be loadbearing, or non-loadbearing.
Vented Thermal Storage Wall System
FIG. 5
The winter sunlight penetrating the South glazing heats the brick, the heat slowly penetrates the brick wall and
warms the interior. Thermal storage walls may have sufficient heat storage to maintain comfortable temperatures in
buildings for periods up to three completely overcast days. The thermal storage wall systems have considerably less
temperature fluctuation than do direct gain systems, but usually do not achieve the same high initial interior
temperatures.
The massive brick thermal storage wall prevents ultraviolet degradation of materials contained in the living space
because solar radiation does not directly enter the habitable space. The performance may be substantially increased
by providing vents at the top and bottom of the brick wall to provide convection in addition to the heat radiated from
the interior face of the wall. Vented walls may be used to decrease the temperature fluctuations and increase the
maximum temperature achieved in the living space. Fig. 1 shows a vented thermal storage wall under construction.
When venting the storage wall system, vents with automatic or manual closures should be used so that the system
does not reverse at night, creating a heat loss.
If controlled vents are not installed on the vented thermal storage wall systems, night insulation is essential to
prevent heat losses at night. Night insulation may be required on unvented thermal storage walls and those with
controlled vents to increase the efficiency of the system.
COMBINED SYSTEMS
The best thermal performance and living conditions result by combining the thermal storage wall system and the
direct gain system. This combination permits some direct sunlight into the living spaces, achieves higher interior
temperatures than the thermal wall system alone, provides less temperature fluctuation than the direct gain system
alone and provides natural lighting. The combination essentially utilizes the best of the two systems.
ATTACHED SUNSPACES
Attached sunspaces are a combination of the components of the direct gain system and the thermal storage wall
system, as shown in Fig. 2 on the right, and in Fig. 6. The sunspace is a room, or space, which typically has both a
glass roof and a glass South-facing wall. The East and West walls may also be glass. The floor is similar to that of
the direct gain system. It consists of 4 to 8-in. thick brick masonry. The North wall is a 10 to 18-in. thick brick
thermal storage wall. The room is vented or ducted to other areas of the structure. With the assistance of fans and
blowers, the structure is heated by the extreme temperatures achieved in the sunspace. The sunspace usually has
severe temperature fluctuations and is often unbearably hot during daylight hours. They do require removable
shading devices to prevent solar gains in the summer. They will also require night insulation if they are to become
useable living space in the evening hours.
43 http://www.gobrick.com/BIA/technotes/t43.htm
6 of 15 9/13/2009 2:11 PM
Attached Sunspace
FIG. 6
43 http://www.gobrick.com/BIA/technotes/t43.htm
7 of 15 9/13/2009 2:11 PM
CAVITY WALL SYSTEM
The cavity wall system, shown in Fig. 7, is a modification of the double envelope system. The concept of the cavity
wall system is that the South-facing thermal storage wall heats up and creates a convective loop around the entire
building envelope. The warmed air space minimizes the temperature differential from the interior of the building
through the inner wythe of the cavity wall. There are no generally accepted design procedures for this type of
system presently available. Some experts in the passive solar design field feel that the increased thermal
43 http://www.gobrick.com/BIA/technotes/t43.htm
8 of 15 9/13/2009 2:11 PM
performance may be accounted for by the insulation in the interior and exterior shells of the double envelope system.
Others feel that there is no convective loop occurring, i.e., the air between the double envelope shells is stagnant.
Cavity Wall System
FIG.7
The use of a properly constructed, insulated brick cavity wall on the North side of the building could be used to
provide a moderate heat loss to drive the convective loop through the air space in the building envelope. This would
reduce the temperature of the air being circulated through the cavity, but the air should still reach high enough
temperatures as it passes through the air space of the thermal storage wall system to provide a net heat gain.
Since there is still considerable controversy regarding this type of system, and since accurate performance analysis
is not easily accomplished, these systems should only be designed and constructed with the appropriate awareness
of the expected and achievable performance level of the system.
METRIC CONVERSION
Because of the possible confusion inherent in showing dual unit systems in the calculations, the metric (SI) units are
not given in this Technical Notes. Table 13 in Technical Notes 4 provides metric (SI) conversion factors for the more
commonly used units.
SUMMARY
This Technical Notes has provided general information concerning passive solar heating systems. It has described
several passive solar heating systems, the basic principles of their operation and general design consideration. This
introduction to passive solar heating systems hopefully provides sufficient familiarization with concepts so that the
design of such systems will be understood. Passive solar cooling is discussed in Technical Notes 43C. The material
properties of brick masonry, as related to passive solar energy systems, is provided in Technical Notes 43D..
Details and construction information are provided in Technical Notes 43G.
This Technical Notes does not and is not intended to provide information for specific designs and applications, but
rather offers general information to assist in the consideration and use of brick masonry in passive solar heating
systems. The decision to use these concepts in the design specific applications is not within the purview of the Brick
Institute of America, and must rest with the owner or designer of any specific project.
Environmental Data for Passive Solar Systems
TABLE 2
a,b
43 http://www.gobrick.com/BIA/technotes/t43.htm
9 of 15 9/13/2009 2:11 PM
43 http://www.gobrick.com/BIA/technotes/t43.htm
10 of 15 9/13/2009 2:11 PM
43 http://www.gobrick.com/BIA/technotes/t43.htm
11 of 15 9/13/2009 2:11 PM
43 http://www.gobrick.com/BIA/technotes/t43.htm
12 of 15 9/13/2009 2:11 PM
43 http://www.gobrick.com/BIA/technotes/t43.htm
13 of 15 9/13/2009 2:11 PM
a
Reprinted from the U.S. Department of Commerce, National Oceanic and Atmospheric Administration, Environmental Data and Information Service, National
Climate Center, Asheville, North Carolina - "Input Data for Solar Systems," by V. V. Cinquemani. J. R. Owenby, Ir., and R. G. Baldwin.
b
Based on 1941 - 1970 Period. Zeros appearing for all values appearing in these columns signify that 1941 - 1970 period normals were not available.
43 http://www.gobrick.com/BIA/technotes/t43.htm
14 of 15 9/13/2009 2:11 PM

a
Reprinted with permission from the 1972 ASHRAE Handbook of Fundamentals Volume, ASHRAE HANDBOOK & Product Directory
Projection greater then 20 ft required.
43 http://www.gobrick.com/BIA/technotes/t43.htm
15 of 15 9/13/2009 2:11 PM

Technical Notes 43C - Passive Solar Cooling with Brick Masonry - Part 1 - Introduction
[March 1980] (Reissued Feb. 2001)
Abstract: Brick masonry passive solar energy systems can be used to significantly reduce the use of fossil fuels for heating and
cooling buildings. The concepts of passive solar cooling systems discussed here are simple modifications to passive solar
heating systems. For locations where humidity is high, or there is little exterior temperature fluctuation, or applications where
low interior design temperatures are required, passive solar cooling may not be viable. Several methods of pre-cooling and the
concept of dehumidifying air with these systems are introduced.
Key Words: attached sunspace, bricks, buildings, cavity wall systems, climatology, conservation, direct gain systems,
effective temperature, energy, masonry, passive solar cooling systems, passive solar heating systems, solar radiation, system
operation, temperature, thermal storage wall systems.
INTRODUCTION
The application of passive solar energy systems using brick masonry can help to significantly reduce the amounts of
fossil fuels and electric energy currently being used for heating and cooling buildings. Other Technical Notes in this
Series address passive solar heating systems with brick masonry. They discuss the general concepts, the
procedures for sizing the systems, and the performance calculations. This Technical Notes introduces the concept of
passive solar cooling systems using brick masonry.
PASSIVE SOLAR COOLING
The terminology "passive solar cooling" does not necessarily refer to the actual reduction of the interior air
temperature of the building. "Passive solar cooling" is a means of providing comfortable interior conditions by
properly using the natural flow of thermal energy to create air movement. These "cooling" systems provide comfort
by controlling the effective temperature of the interior of a building.
The effective temperature is a measure of the comfortable air conditions in a building dependent upon the actual
temperature of the air, the level of relative humidity, and the amount of air movement. By properly varying any one,
or any combination of these factors, more comfortable interior conditions can be achieved. The American Society of
Heating, Refrigerating and Air-Conditioning Engineers (ASHRAE) provides methods which may be used to determine
the amount of change or fluctuation necessary to achieve comfortable interior conditions. These methods are
described in ASHRAE 1997 Handbook of Fundamentals, and ASHRAE Standard 55-92, Thermal and
Environmental Conditions for Human Occupancy.
The actual determination of the effectiveness of passive solar cooling is complex and its performance is not yet
satisfactorily predicted with calculation procedures alone. The type of passive solar cooling system selected, and its
performance can be greatly affected by the site and the climatological conditions.
SYSTEMS AND OPERATION
The basic passive solar heating systems, utilizing brick masonry, are discussed in Technical Notes 43, These
systems are: thermal storage wall systems, direct gain systems, attached sunspaces and combinations of these.
These passive solar heating systems can be easily modified to provide interior comfort during the cooling season.
Obtaining all the necessary cooling with passive solar cooling systems usually is neither economically nor
thermally feasible for the entire cooling season. These simple modifications to passive solar heating systems can
be used to create more comfortable interior conditions for at least part of the cooling season in most climates.
The necessary modifications to passive solar heating systems to provide passive solar cooling are provisions for 1)
exhausting air from the interior, and 2) intaking exterior air. Schematics are shown in Figs. 1, 2, and 3 for the direct
http://www.gobrick.com/BIA/technotes/t43c.htm
1 of 8 9/13/2009 2:12 PM
gain system, attached sunspace, and thermal storage wall system, respectively. The principal modification is to
provide controlled openings for exhausting the internal heat gained by the passive solar heating system. The
controlled openings should be at the highest points of the structure, preferably in the roof/ceiling, or gable. Control of
the openings may be provided with operable vents, or registers. Similar openings can be placed at the low points of
the structure for intaking exterior air. The openings for intaking exterior air may be the windows or doors of the
structure.
The operation of each of these systems is very similar in the cooling mode: (1) sunlight strikes the south-facing
glazing, (2) solar energy is transmitted through the south-facing glazing to the brick masonry thermal storage media,
(3) the brick masonry absorbs and stores the heat, (4) radiant heat from the surface of the brick masonry rises, (5)
the heated air is exhausted through the controlled openings at the top of the structure, (6) as the heat is exhausted,
exterior air is drawn into the structure, and (7) the air movement created by exhausting and intaking air through the
structure creates the effect of cooling and provides more comfortable interior conditions.
Cooling With the Direct Gain System
FIG.1
Direct Gain System
The direct gain system, when applied as passive solar cooling, is the most economical, but probably the least
effective. The minimum 4-in. ( 100 mm) thick brick masonry floors and walls on the interior are exposed to direct
sunlight to absorb and store heat.
The interior brick masonry should be dark to absorb most of the heat and radiate and reflect only a small portion
during the day. The gradual release of radiant heat through the night draws the cool night air into the structure and
cools the structure.
The system is only advantageous when the nighttime temperatures consistently fall below the interior design
temperature and when internal solar heat gain can be adequately controlled to prevent overheating in the daytime.
A major problem with using a direct gain system is that the interior space used to store heat is also an integral part
of the habitable space of the building.
Attached Sunspace
Using the attached sunspace for passive solar cooling is probably more effective but less economical than direct gain
cooling. In the attached sunspace, the heat storage element is not usually part of the space that is to be cooled. The
system schematic is shown in Fig. 2.
Although the intent in many applications is to use the attached sunspace as a greenhouse, this is not advantageous in
most applications because the greenhouse will be vented to the interior and the humidity from watering plants may
result in uncomfortable interior conditions and condensation problems. The major disadvantage of this system is the
http://www.gobrick.com/BIA/technotes/t43c.htm
2 of 8 9/13/2009 2:12 PM
cost of the additional floor area which has limited use.
Cooling With the Attached Sunspace
FIG. 2

The system consists of minimum 4-in. ( 100 mm) thick brick masonry floors, south-facing glazing and preferably a 10
to 18-in. (250 to 450 mm) thick vented brick masonry thermal storage wall between the sunspace and the habitable
portion of the building.
In the cooling mode, the top vents of the brick masonry storage wall are closed and the bottom vents are open. The
air in the sunspace is heated by radiant heat from the brick masonry. The heated air rises through operable openings
in the roof of the sunspace, drawing air from the habitable spaces through the bottom vents of the brick masonry
thermal storage wall. The air drawn from the habitable space is replaced by exterior air drawn in through operable
windows or doors.
Thermal Storage Wall System
One of the most economical and effective passive solar cooling systems is the vented thermal storage wall, shown
schematically in Fig. 3. The greatest advantage of the thermal storage wall is that the heat used for the passive solar
cooling does not directly enter the interior spaces of the habitable portion of the building.
The system consists of exterior glazing 2 to 4 in. (50 to 100 mm) in front of a 10 to 18-in. (250 to 450 mm) thick
vented brick masonry wall used for storing heat. Operation is similar to that of the attached sunspace. The operable
openings for exhausting the heated air may be located at the top of the exterior glazing. The exhaust system may
also be operable vents at the top of the airspace from which the air may be exhausted through additional vents in the
roof. When using the latter exhaust system, additional vents from the habitable space through the roof/ceiling
component may be used to increase the heat flow from the interior, thereby drawing additional air into the building.
http://www.gobrick.com/BIA/technotes/t43c.htm
3 of 8 9/13/2009 2:12 PM
Cooling With the Vented Thermal Storage Wall System
FIG. 3


CONTROLS
These three basic passive solar heating systems, as modified for passive solar cooling, require controls to regulate
internal heat gain and the level of comfort, "cooling", achieved. These controls may be automatically or manually
operated vents, or registers. The controls should be such that the system can be totally "shut down" when it is not
being effective, i.e., when exterior conditions are such that a comfortable effective temperature cannot be maintained
inside the building. Shading devices are required as a means of controlling the amount of sunlight permitted to enter
the structure for operation of the passive solar cooling system. These may be automatic or manual devices, but are
necessary to prevent overheating that can occur during the cooling season when the interior temperature, i.e.,
effective temperature, will no longer be within the comfort range. The entire system must be completely shut down
before mechanical/refrigeration cooling systems are put into operation. Shading the south-facing glazing and closing
openings are required for efficient use of any conventional cooling system. Obviously, operation is a critical factor in
the performance of "passive solar cooling systems".
CAVITY WALL SYSTEM
A system which may be effectively used for cooling, with less consideration of the climatic conditions, is the cavity
wall system, schematically represented in Fig. 4. The north and south walls of the structure are uninsulated cavity
walls (see Technical Notes 21 Series). The south-facing wall, above grade, is an unvented thermal storage wall. The
airspace in the thermal storage wall system is open at the bottom to the cavity of the basement or foundation wall,
and at the top to the roof/ceiling. The cavity of the wall is open to ductwork extended in the north-south direction
through the basement floor, or crawl space. As shown in Fig. 4, this provides an air passageway within the building
envelope components. A ductwork system is provided from the base of the cavity to vents on the exterior. Exhaust
vents are provided in the roof or gable ends.
Cavity Wall System Cooling Mode
FIG. 4
http://www.gobrick.com/BIA/technotes/t43c.htm
4 of 8 9/13/2009 2:12 PM
With the cavity wall system, the south-facing glazing exposes the brick thermal storage wall to sunlight, which heats
up and causes air to rise through the cavity. As the air rises, it is vented to the exterior from the top of the cavity,
and exterior air is drawn into the cavity via the ductwork system and exterior vents. This provides a means of
keeping the entire building envelope cooled. Since the surfaces warmed during the daytime hours retain heat, this
continues through the evening hours, further cooling the building envelope. The cooled building envelope and interior
require a longer time period to be heated up to uncomfortable temperatures during the daytime hours of the next
day. The advantages of passive solar cooling with the cavity wall system are: (1) shutdown is not essential for
conventional cooling to work effectively (the cooling systems are isolated from each other), and (2) since the exterior
air for cooling is not brought into the habitable space, the effects of humidity (a major drawback in most passive
solar cooling systems, depending on climate), may be reduced.
The east and west walls do not have to be uninsulated cavity walls. By keeping all walls cavity walls, the east and
west walls may perform as a buffer zone between the north and south walls. This may increase the overall
performance of the system.
A schematic of the system in the heating mode is shown in Fig. 5. The vents are closed, which creates a convective
loop around the entire shell of the building; through the floor, wall and roof/ceiling components. This thermal
convective loop warms both the interior and exterior wythes of the building envelope. Since this operation warms the
interior wythe, there is little or no heat loss through those portions of the building envelope. This system, properly designed and
operated, may provide the most effective passive solar heating and cooling.
Cavity Wall System Heating Mode
FIG. 5

FACTORS AFFECTING PERFORMANCE
The effects of the environmental conditions and building use on passive solar heating systems are discussed in
Technical Notes 43. These must also be considered for passive solar cooling systems, however, the necessary
considerations of these factors vary for passive solar cooling systems. The major variations and additional effects
which must be addressed specifically are: temperature, humidity and shading.
Exterior Design Temperature
The exterior design temperature may be such that the effective temperature range cannot be achieved or maintained
within the structure. Since the effects of cooling are principally achieved by air movement, this may make the cooling
system ineffective. One option is to take maximum advantage of the daily temperature swing. When the nighttime
temperatures drop below the interior design temperature, the structure may be cooled during the night, delaying the
time to heat up the next day. Caution must be used when considering the daily temperature swing to guard against
http://www.gobrick.com/BIA/technotes/t43c.htm
5 of 8 9/13/2009 2:12 PM
overcooling in moderate climates.
Humidity
Humidity is an additional environmental factor not generally addressed in passive solar heating systems. In areas
where the effects of high humidity cannot be eliminated by air movement, these simple versions of passive solar
cooling systems may not be effective. Additional complex modifications to the basic passive solar cooling systems
may be necessary to dehumidify the air.
Shading Devices
Operable shading devices are usually required in passive solar cooling systems. The shading devices are used to
control the amount of solar radiation permitted to strike the system. This is necessary to prevent overheating,
especially when the system is marginal because the effective temperature cannot be attained by natural air flow. In
this case, the system should be completely shaded from the summer sunlight.
In instances where the system is providing cooling by night air intake, it may be advantageous to have the system
shaded from the morning and possibly early afternoon sunlight. Exposure to only the late afternoon sunlight may
result in sufficient performance to draw cool night air through the structure.
SPECIAL CONSIDERATIONS
The performance of the passive solar cooling system may be greatly increased by pre-cooling, or dehumidifying the
air before introducing it into the structure. Pre-cooling and dehumidifying the air are both fairly straightforward
concepts. Adapting the system for pre-cooling air is usually simple, but dehumidifying the air is much more
complicated.
Pre-Cooling Air
Air may be cooled before it is introduced into the structure by providing underground ductwork or piping, and venting
it to the surface as shown schematically in Figs. 6, 7 and 8. This is easily adaptable for direct gain, attached
sunspace and thermal storage wall cooling systems. The ductwork should be corrosion-resistant and installed for a
sufficient length and at the appropriate depth to pre-cool the air. The number of ducts and their length and depth
requirements are beyond the scope of this Technical Notes because they are a function of climate, soil type,
elevation of ground water and other related factors, all of which affect the amount of pre-cooling, both required and
attainable. General design information and calculation procedures may be obtained from the References 1, 2, 6 and
7 of this Technical Notes.
http://www.gobrick.com/BIA/technotes/t43c.htm
6 of 8 9/13/2009 2:12 PM
Pre-Cooling and Dehumidification of Exterior Air
for Cooling With the Direct Gain System
FIG. 6
Pre-Cooling and Dehumidification
of Exterior Air With Attached Sunspace
FIG. 7
Pre-Cooling and Dehumidification
of Exterior Air With Vented Thermal Storage Wall System
FIG. 8
DEHUMIDIFICATION
Air may be dehumidified by using the concept shown in Figs. 6 through 8, for pre-cooling. Dehumidifying the air with
http://www.gobrick.com/BIA/technotes/t43c.htm
7 of 8 9/13/2009 2:12 PM
the passive solar system alone can be very difficult. Again, it is a function of climate, soil type, level of ground water
and other related factors. The procedure for determining the amount of dehumidification involves fairly complex
calculations. These calculation procedures are similar to those in the ASHRAE 1997 Handbook of Fundamentals and
ASHRAE Standard 55-92. The temperature fluctuations necessary to saturate air and condensate water by the
natural flow of air further complicates the use of passive solar cooling systems for providing dehumidification.
SUMMARY
This Technical Notes provides general information concerning passive solar cooling systems. In addition to
describing modifications of the passive solar heating systems which may be used to supply successful passive solar
cooling, it introduces an innovative system - the cavity wall system - which may be quite effectively used for heating
and cooling buildings. The basic concepts of the passive solar cooling systems and the principles of their operation
are also discussed. The purpose of this Technical Notes is to provide general information on passive solar cooling
systems with brick masonry. It discusses type, operation, advantages and disadvantages of these systems. This
Technical Notes does not and is not intended to provide information for specific designs or applications, but rather
offers general information to assist in the consideration of the use of passive solar cooling systems of brick masonry.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the engineering staff of the Brick Industry Association. The information contained herein must be used
in conjunction with good technical judgement and a basic understanding of the properties of brick masonry. Final
decisions on the use of information contained in this Technical Notes are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. Proceedings of the Fourth Passive Solar Conference, October 3-5, 1979, Kansas City, Missouri, published by the publishing
office of the American Section of the International Solar Energy Society, Incorporated, McDowell Hall, University of Delaware,
Newark, Delaware, 1979.
2. Proceedings of the International Solar Energy Society, Silver Jubilee Congress, Atlanta, Georgia, May 1979, Pergamon
Press, Inc., Maxwell House, Fairview Park, Elmsford, New York 10523, 1979.
3. Technical Notes 43 Revised, "Passive Solar Heating with Brick Masonry, Part I - Introduction" Brick Industry Association,
Reston, Virginia, May/June 1981.
4. ASHRAE 1997 Handbook of Fundamentals, American Society of Heating, Refrigerating and Air-Conditioning Engineers,
Inc., Atlanta, GA, 1997.
5. ASHRAE Standard 55-92, Thermal Environmental Conditions for Human Comfort, American Society of Heating,
Refrigerating and Air-Conditioning Engineers, Inc., Atlanta, GA, 1992.
http://www.gobrick.com/BIA/technotes/t43c.htm
8 of 8 9/13/2009 2:12 PM

Technical Notes 43D - Brick Passive Solar Heating Systems - Part 4 - Material Properties
Reissued September 1988
Abstract: The inherent properties of brick masonry make it one of the most advantageous storage media materials
for passive solar energy systems. Brick masonry may be used to provide an aesthetic effect, structural capacity and
other design considerations in addition to thermal storage. Most of these inherent properties of brick masonry are
already well understood for conventional applications. However, in order to properly use brick masonry as a thermal
storage media for passive solar energy systems additional information may be needed by the designer. This
additional information has to do with the effective thermal storage of brick masonry.

Keywords: absorptivity, brick, density, emissivity, energy heat transfer, masonry, material properties, passive
solar energy systems, reflectance, solar radiation, specific heat, temperature, effective thermal storage, thermal
conductivity, thermal diffusivity.

INTRODUCTION

Brick masonry can be used most advantageously as the thermal storage media in direct gain systems, thermal
storage wall systems and attached sunspaces. The general concepts, empirical procedures for sizing systems and
performance calculations are discussed in Parts I through III of this Technical Notes Series. This Technical Notes
provides information and references regarding the material properties of the basic components of passive solar
energy systems. This information includes the properties of brick masonry when used for thermal storage, with
major emphasis on the effective thermal storage of brick masonry and a general discussion of the properties of
glazing materials when used as collectors.

BRICK MASONRY
General

Most of the design requirements and performance of brick masonry as a general building material are discussed in
other Technical Notes. The inherent properties of brick masonry offer many design advantages in addition to those
required for use as a thermal storage material.

Structural
Brick masonry has many applications as a structural element in buildings. Brick masonry is commonly used as
loadbearing elements in commercial and residential structures. Brick masonry when considered as a thermal storage
media for passive solar energy systems may also be considered as a structural element. Information on loadbearing
brick masonry is provided in Technical Notes 24 Series.
43d http://www.gobrick.com/BIA/technotes/t43d.htm
1 of 13 9/13/2009 2:13 PM
One of the reasons brick masonry is less frequently considered as a structural element in one and two-family
construction is because of the difficulties in insulating solid masonry walls to meet prescriptive energy code
requirements for the reduction of heat loss. This can be overcome by using loadbearing insulated cavity walls which
provide a durable facade, sufficient space for insulation and interior brick masonry which may be used as thermal
storage in direct gain systems. Cavity wall construction is addressed in Technical Notes 21 Series.
The structural design may require reinforced brick masonry. Reinforcement in brick masonry usually has little if any
effect on the thermal performance of the wall. This is because the reinforcement is usually horizontal and/or vertical
in the plane of the wall, and occurs at or near the center of the wall section resulting in very little increase of thermal
transmission through the wall. Information regarding reinforced brick masonry is provided in Technical Notes 17
Series.

Durability
Brick masonry is an extremely durable building material requiring little or no maintenance. It does not require
coatings or coverings which could reduce its thermal performance as a storage media. Coatings and coverings may
decrease the emissivity and thermal conductivity of the brick masonry. This is not desired when trying to optimize on
the available thermal storage and thermal energy retrieval. Since coatings and coverings are not required, brick
masonry may be exposed to enhance the aesthetics of the building. The use of coating applied to exterior brick
masonry is discussed in Technical Notes 6A.

Aesthetics
Brick masonry is normally used as an exterior facade, not only because of its durability but also because it provides
architectural freedom. Brick masonry offers many bond patterns, colors and textures. As elements of the building,
brick masonry provides options for architectural freedom that no other building material can offer. For instance, not
only the texture of the brick itself is available in many varieties, but brick allows variation in wall texture, also. The
texture of the wall may be varied by using projected or recessed brick or even sculptured brickwork. The typical
modular sizes of brick masonry are given in Technical Notes 10B and common bond patterns are given in Technical
Notes 30. Brick masonry used as paving is discussed in Technical Notes 14 Series. Information on the use of brick
masonry sills and soffits is provided in Technical Notes 36 Series. The use of brick masonry arches and reinforced
brick masonry lintels are discussed in Technical Notes 31 Series and 17H, respectively.

Fire Resistance
Depending upon the specific application of brick masonry in passive solar energy systems, brick masonry may be
designed and placed to offer fire protection. The fire resistance of brick masonry is discussed in Technical Notes 16
Series.

Sound Transmission Resistance
Brick masonry, because of its inherent properties offers considerable reduction in sound transmission. Thus,
depending on specific design applications, strategically placed thermal storage elements may be used to reduce
sound transmission from one area of the building to another or from the exterior to the interior of the building.
Information on the sound transmission classification of brick masonry is provided in Technical Notes 5A.

Effective Thermal Storage
43d http://www.gobrick.com/BIA/technotes/t43d.htm
2 of 13 9/13/2009 2:13 PM
The overall performance of the brick masonry as a passive solar energy system thermal storage component is
dependent on its absorptivity, emissivity, and ability to store heat. The ability of a material to store heat is usually
referred to as heat capacity which is a function of the specific heat and density of a material. In addition to the heat
capacity, the way the wave of thermal energy penetrates the material being used to store heat should also be
considered. The performance as a thermal storage media may be estimated using the value of the thermal diffusivity
of the material. Thermal diffusivity is not only a good value for assisting in the selection of materials but is also useful
in simplified heat flow calculations to determine the amount of heat penetrating a material and the number of hours it
takes for the heat transmission to occur. This information is useful for selecting the thickness of thermal storage. The
thermal diffusivity is a function of the specific heat, density and thermal conductance of a material.
Specific Heat. The specific heat, c, of material is the amount of heat required to increase the temperature of a unit
weight of material one degree. The specific heat, c, in Btu per pound per degree Fahrenheit, for brick may vary from
0.20 to 0.26. Typically this variation is due to the impurities in the clay used to manufacture the brick. The greater
the percentage of metallic oxides in the clay, usually the greater the specific heat. Building brick which usually have a
low percentage of metallic oxides by weight have low specific heats usually between 0.20 to 0.22 Btu/lb/
o
F, whereas
face brick which contain larger amounts of metallic oxides, typically up to 35%, have specific heats ranging between
0.22 to 0.26 Btu/lb/
o
F. A value of specific heat of face brick which may be used when the actual specific heat is not
known is 0.24 Btu/lb/
o
F.
For building brick or brick containing a low percentage of metallic oxides, a value of 0.22 Btu/lb/
o
F may be used.
Generally red, brown and blue brick contain high amounts of metallic compounds.
The value of the specific heat for brick may be assumed for brick masonry. The specific heat of grouted hollow brick
may be approximated by determining the percent of the brick masonry which is to be grouted, and averaging the
specific heat, accordingly. This may be done by adding the product of the specific heat of face brick times the
fraction of the brick which is solid, at least 0.60, and the specific heat of grout times the fraction of the brick which is
cored, less than or equal to 0.40. For grouted hollow walls, the specific heat for the masonry wall may be modified
for the grout by using Equation 1:
c
w
= [(t
b1
X c
b1
) + (t
b2
X c
b2
)+ (t
g
X c
g
)] / (t
b1
+ t
b2
+ t
g
) (1)
where: cw = Average specific heat of a grouted brick masonry wall, in Btu/lb/
o
F.
tb1 = Nominal thickness of the exterior wythe of brick masonry, in inches.
c
b1
= Specific heat of the brick in the exterior brick masonry wythe, in Btu/lb/
o
F.
t
b2
= Nominal thickness of the interior wythe of brick masonry, in inches.
c
b2
= Specific heat of the brick in the interior brick masonry wythe, in Btu/lb/
o
F.
tg = Nominal thickness of the grout, in inches.
cg = Specific heat of the grout, in Btu/lb/
o
F.

Consider a 14-in. thick brick thermal storage wall constructed of 4-in. face brick, a 4-in. grouted space and a 6-in.
grouted hollow brick wythe. Although this example is not representative of typical brick masonry thermal storage
components, it is offered to include the available combinations of brick masonry construction. The specific heat of
the wall assembly components may be determined by using Table 1 where:
tb1 = 4 in., cb1 = 0.24 Btu/lb/
o
F
t
b2
= 6 in., c
b2
= 0.22 Btu/lb/
o
F
tg = 4 in., cg = 0.20 Btu/lb/
o
F
43d http://www.gobrick.com/BIA/technotes/t43d.htm
3 of 13 9/13/2009 2:13 PM


a
These values are representative of the information available to the Brick Institute of America, and are typical for brick being manufactured today. These
values may vary by plus or minus 10 % depending on the specific brick being considered.
b
Hollow brick are assumed to be 60 % solid and the core space fully grouted.
c
Source is Reference 3.
d
The thermal conductivity of grouted hollow brick should be determined by dual path analysis. Typically grouted hollow brick, 60 % solid, fully grouted will have
a thermal conductivity of approximately 10 Btu/hr/
o
F/ft
2
, per inch.


The approximate average specific heat of the wall assembly is found by substituting these values into Equation 1:
c
w
= [(4 X 0.24) + (6 X 0.22) + (4 X 0.20)] / (4 + 6 + 4)
c
w
= 0.22 Btu/lb/
o
F
Density. The density, r, of brick varies with the type of clay, the additives used in manufacturing and with the extent
43d http://www.gobrick.com/BIA/technotes/t43d.htm
4 of 13 9/13/2009 2:13 PM
of firing. Generally the longer the firing and the higher the temperature, the more dense the brick. Typical densities
for various brick are provided in Table 1. These values are average densities. The density for grouted hollow brick
assumes 130 lb per cu ft density brick and 120 lb per cu ft density grout. Hollow brick may range from 75% to 60%
solid. This may significantly change the density, however the values provided in Table 1 are for grouted hollow brick
assuming 60% solid.

The density, as the specific heat, of brick masonry is slightly less than that of brick, however for simplified effective
thermal storage calculations these differences are usually insignificant. Typically, the maximum amount of mortar in
solid brick or grouted hollow brick walls constructed with full collar joints would be about 30% of the wall volume;
however, considering that mortar has a density of about 120 lb per cu ft. this results in a less than 3% reduction in
the density of the wall as compared to the density of the brick.
when considering the use of a grouted hollow wall, two wythes of masonry constructed with a grouted space in
between, the density of the wall should be approximated by calculation using Equation 2:

rw = [(tb1 X rb1) + (t
b2
X r
b2
) + (tg X rg)] / (tb1 + tb2 + tg) (2)
where: rw = Average density of the wall, in Ib/cu ft.

r
b1
= Density of exterior brick, in Ib/cu ft.
r
b2
= Density of interior brick, in Ib/cu ft.
rg = Density of grout, in Ib/cu ft.

For a 14-in. thick brick thermal storage wall, constructed of a wythe of 4-in. face brick, a 4-in. grouted space and a
6-in. grouted hollow brick wythe, the densities may be selected from Table 1 where:
t
b1
= 4 in., rb1 = 130 Ib/cu ft
t
b2
= 6 in., r
b2
= 26 Ib/cu ft
t
g
= 4 in., r
g
= 120 Ib/cu ft
By substituting these values into Equation 2, the average density of the wall, r
w
, is found to be:

rw = [(4 X 130) + (6 X 126) + (4 X 120)] / (4 + 6 + 4)
rw = 125 Ib/cu ft

Thermal Conductivity. The thermal conductivity, k, of brick is discussed in Technical Notes 4 Revised. Typical
values of the thermal conductivity of brick are provided in Table 1. The thermal conductivity of brick varies with
density. The denser the brick, generally the greater the thermal conductivity. The thermal conductivity of grouted
brick should be determined by the dual path procedure described in Technical Notes 4 Revised.
43d http://www.gobrick.com/BIA/technotes/t43d.htm
5 of 13 9/13/2009 2:13 PM
For hollow brick which is grouted, there are two paths for heat flow, one path is through the webs of the hollow brick
and the other path is through the face shells and grout. The average thermal resistivity of grouted hollow brick may
be determined by using Equation 3:
rh = [(rb X tw) / l ] + [[(rb X 2tf) + rg x (t - 2tf)] X [(I - tw) / (l X t)]](3)
where: rh = Average thermal resistivity of grouted hollow brick, in (
o
F ft
2
hr)/Btu. in.
rb = Thermal resistivity of brick, in (
o
F ft
2
hr)/Btu. in.
r
g
= Thermal resistivity of grout, in (
o
F ft
2
hr)/Btu. in.

t = Thickness of the brick, in inches.
tf = Thickness of the face shell, in inches.
tw = Total thickness of the webs, in inches.

/ = Length of the brick, in inches.

Considering the thermal resistivities and thickness of a 6 X 4 X 12 grouted hollow brick:
rb = 0.11 (
o
F ft
2
hr)/Btu. in.t = 6 in.
rg = 0.08 (
o
F ft
2
hr)/Btu in. tf = 1.25 in.
/ = 12 in.tw = 4 in.

and substituting these values into Equation 3, the average thermal resistivity of a 6 X 4 X 12 grouted hollow brick
would be:
rh = [(0.11 X 4) / 12] + [[(0.11 X 2.50) + 0.08 X (6 - 2.50)]
X [(12 - 4) / (12 x 6)]]
rh = 0.037 + [[0.275 + 0.280] X 0.11]
rh = 0.098 (
o
F ft
2
hr)/Btu in. or approximately 0.10 (
o
F ft
2
hr)/Btu. in.
The average thermal conductivity is the inverse of the average thermal resistivity. The average thermal conductivity,
kh, for a 6 X 4 X 12 grouted hollow brick would be:
kh = 1/rh(4)
kh = 1/0.10 (
o
F ft
2
hr)/Btu in.
kh = 10.00 Btu/hr/
o
F/ft
2
per inch of thickness
43d http://www.gobrick.com/BIA/technotes/t43d.htm
6 of 13 9/13/2009 2:13 PM

The average thermal resistivity of a storage media is the summation of the thermal resistivity times the thickness of
the materials divided by the total thickness. This is expressed in Equation 5:
r
w
=[(r
1
X t
1
) + (r
2
X t
2
) + . . . ] / (t
1
+ t
2
+ . . . )(5)
where: r
w
= The average thermal resistivity of the storage media, in (
o
F ft
2
hr)/Btu. in.
r = The thermal resistivity of each component, in (
o
F ft
2
hr)/Btu. in.

t = The thickness of each component, in inches.

Thus consider again the 14-in. grouted hollow wall constructed of a 4-in. wythe of face brick, a 4-in. grouted space,
and a 6-in. wythe of grouted hollow brick. The nominal thickness and respective average thermal resistivities of each
component would be:
t
1
= 4 in.,r
1
= 0.11 (
o
F ft
2
hr)/Btu in.
t
2
= 4 in.,r
2
= 0.08 (
o
F ft
2
hr)/Btu. in.
t
3
= 6 in.,r
3
= 0.10 (
o
F ft
2
hr)/Btu in.
Substituting these values into Equation 5, the average thermal resistivity of the wall, rw, is determined to be:
rw = [(0.11 x 4) + (0.08 x 4) + (0.10 x 6)] / (4 + 4 + 6)
rw = [1.36(
o
F ft
2
hr)/Btu in.] / 14 in.
r
w
= 0.097 (
o
F ft
2
hr)/Btu per inch of thickness

Thermal Conductance. Thermal conductivity and thermal resistivity refer to the value of heat loss for one inch of
thickness. The thermal conductance is the value of heat loss for a specified thickness. The average thermal
conductance for a one foot thickness of the storage media is used in the simplified equations of heat transfer for
determining effective thermal storage. The average thermal conductance for a one foot thickness of the brick
storage media may be determined using Equation 6:
Ca = 1 / (r
w
X 12 in./ft) (6)
where: Ca = The average thermal conductance of the storage media for
one foot of thickness. in (Btu/hr/
o
F/ft
2
) / ft.

The 4-in wythe of face brick, 4-in. grouted space and 6-in. wythe of grouted hollow brick is calculated to have a
thermal conductance per foot of thickness of:
Ca = 1 / [0.097 (
o
F ft
2
hr) / Btu in. X 12 in./ft]
Ca = 0.859 (Btu/hr/
o
F/ft
2
) / ft
43d http://www.gobrick.com/BIA/technotes/t43d.htm
7 of 13 9/13/2009 2:13 PM

Thermal Diffusivity. Typically the storage capacity of a material is represented by the amount of heat which can be
stored in the material, the heat capacity of the material. The heat capacity may be determined by using Equation 7:

b = c X r(7)
where: b = Heat capacity, in Btu/cu ft/
o
F.

Typical values for the heat capacity of various brick are provided in Table 1. Thermal diffusivity is a function of the
heat capacity and thermal conductance per foot of material thickness. The thermal diffusivity of a material may be
determined by using Equation 8:

d = Ca/ b or d = Ca / (c X r) (8)
where: d = Thermal diffusivity, in ft
2
/hr.

The values of thermal diffusivity for typical brick masonry are provided in Table 1. The value of thermal diffusivity
may be used to provide the designer with a better concept of heat storage in the passive solar energy system
thermal storage component. The use of the thermal diffusivity in simplified heat transfer equations may provide a
more rational approach for selecting the thickness of thermal storage walls.
The average value of thermal diffusivity may be determined by using Equation 9:

da = Ca / (cw X rw)(9)
where: da = Average thermal diffusivity, in ft
2
/hr.
Thus for the 14-in. thick wall assembly constructed of 4-in. solid brick wythe, a 4-in. grouted space, and a 6-in.
wythe of grouted hollow brick the average thermal diffusivity may be determined using Equation 9. Ca for this wall
was determined to be 0.859 (Btu/hr/
o
F/ft
2
)/ft, rw was determined to be 125 Ib/cu ft and cw = 0.22 Btu/Ib/
o
F. Thus,
the average thermal diffusivity would be:

da = 0.859 (Btu/hr/
o
F/ft
2
)/ft/(0.22 Btu/lb/
o
F X 125 Ib/cu ft)
da = 0.031 ft
2
/hr

The value of the average thermal diffusivity is useful in simplified heat transfer equations, however if precise values
are desired, each component in section should be analyzed individually.
Emissivity. The emissivity of a surface is its ability to radiate heat to the surroundings. This is the basis of heat
retrieval in passive solar energy systems as discussed here. The radiant heat from the surface of the brick masonry
43d http://www.gobrick.com/BIA/technotes/t43d.htm
8 of 13 9/13/2009 2:13 PM
is what causes the useable natural flows of thermal energy; i.e., surface to air conduction, convection between
surfaces and radiation to surfaces and the air, which heat the interior spaces of the building. Typical values of
emissivity for brick are provided in Table 1.
Exposed brick masonry allows the use of various bond patterns, projections and even sculptured brick work to
increase the aesthetic value of the thermal storage media. Although at first the idea of leaving the brick exposed
seems merely aesthetic, it does serve a function important to the thermal performance of the thermal storage media.
Brick masonry does not require nor should it have any coverings; i.e., gypsum wall board, paints, wall papers, or
carpeting which could decrease the emissivity of the surface. The addition of coatings and coverings not only may
reduce the emissivity of the thermal storage element but will usually decrease the thermal conductivity thus
decreasing the surface temperatures and the amount of surface radiant heat available.
If the value of emissivity, the surface temperature of the thermal storage element, and the surface temperature of
the interior materials being radiated to are known, the amount of radiant thermal energy emitted may be
approximated. The approximate amount of thermal radiation emitted per square foot of thermal storage surface area
may be calculated using Equation 10:
qr = [0.174 X e X [Tr4 - Tc4]] / 108 (10)
where: qr = Radiation, in Btu/hr/ft
2

e = Emissivity.
Tr = Temperature of the radiant surface measured from absolute zero,
degrees Fahrenheit plus 459.6.
Tc = Temperature of the receiving surface measured from absolute zero,
degrees Fahrenheit plus 459.6. For most applications in passive solar
energy systems, the value of Tc is usually interior design temperature in
o
F
plus 459.6, typically 72
o
F + 459.6 or 531.6.
Consider an average surface temperature of a radiating surface at 83
o
F and an interior design temperature of 72
o
F.
Measuring these temperatures from absolute zero would result in Tr = 83 + 459.6 or 542.6 and Tc = 72 + 459.6 or
531.6. The radiation from a dark brown brick wall, with e = 0.93, may be determined using Equation 10 to be:
qr = [0.174 X 0.93 X [(542.6)4 - (531.6)4]] / 108
qr = [0.168 X (8.668 X 1010)] / 108
qr = (1.103 X 109) / 108
qr = 11.03 Btu/hr/ft
2

Absorptivity. The solar absorptivity of a material is mostly dependent on color. The solar absorptivity is the ratio
between how much solar radiation is absorbed by a material to that absorbed by a standard black surface.
Typically, passive solar thermal storage components (or any finish applied to such components) should be as dark a
color as possible to provide sufficient energy absorption. However, trade-offs do exist between color, wall thickness
and the amount of surface area exposed to sunlight. Trade-offs also exist between darkness of color and how much
heat is desired and when the available heat is wanted. These trade-offs can only be adequately determined by
rigorous analysis and are not recommended for use with rule-of-thumb approaches. Surfaces (such as frame walls)
not being used for storage should be painted light colors in order to reflect as much energy as possible to the darker
storage material. Although black is the most desirable storage material color from a thermal point of view, it has
been determined that the darker natural brick colors (browns, blues and reds) will perform almost as effectively,
without deterioration problems which may result when using paint or other coverings. Typical values for solar
43d http://www.gobrick.com/BIA/technotes/t43d.htm
9 of 13 9/13/2009 2:13 PM
absorptivity of brick are given in Table 2. Brick with glossy glazed ceramic coatings should be avoided as they will
reflect too great a percentage of the solar radiation striking them. Several brick manufacturers can supply brick with
dull black ceramic glazed faces, which may increase the solar radiation absorbed.


Although it would seem at first glance that rough-textured brick, by providing more surface area for the collection of
energy, would be more effective than smooth brick as an energy storage media, but it has been determined that this
is not the case. It appears that brick texture does not have a major impact on the performance of passive solar
installations and that any desired texture can be used without significant loss or gain in effectiveness.

GLAZING MATERIALS
General

Information regarding the transmittance, reflectance, absorptance, thermal performance and durability of glazing
materials for passive solar energy systems should be obtained from the glazing manufacturer. Some general
suggestions to assist in the design and selection of glazing materials for passive solar energy systems applications
are discussed.

Transmittance
The glazing material used should have a high transmittance, low absorptance and low reflectance of solar radiation.
The high transmittance is important so that the maximum amount of solar energy may be transmitted to the interior
of the passive solar building where it can be absorbed and stored in the interior brick masonry. Generally, the higher
the transmittance the lower the absorptance and reflectance. When the glazing is desired to provide specific
daylighting requirements, it may be preferred to use a glazing material which diffuses the direct solar radiation.

Absorptance
Depending on the glazing material and the frame assembly, absorptance should be a factor in selecting the glazing
material or a proper framing assembly. The higher the absorptance, the lower the amount of energy transmitted to
the interior. Typically, the absorptance of most glazing materials is low. The amount of solar absorptance may be
important because of thermal stress that may occur within the glazing material itself or between the glazing material
43d http://www.gobrick.com/BIA/technotes/t43d.htm
10 of 13 9/13/2009 2:13 PM
and framing assembly. The framing assembly should be such that it allows for thermal expansion of the glazing
material being used to avoid structural failure of the collector component.

Reflectance
The reflectance of the glazing material should be kept to a minimum, so that the maximum amount of solar radiation
may be transmitted to the interior. In addition to the consideration of reflectance, as related to transmittance and
maximum solar performance, reflectance might also require consideration of exterior glare which may be annoying or
even hazardous because of impairing vision.

Thermal Conductivity
Typical values for overall coefficient of heat transmission of glass are given in Table 3. Table 3 also provides solar
transmittance correction factors for glass. The solar transmittance correction factors are based on double glass
having a solar transmittance value of 1. Information regarding plastics or other glazing materials must be obtained
from the material manufacturer.
The basic material properties for single, double and triple glass given in Table 3 may be used as a means of
selecting the appropriate glazing material by considering the trade-offs between the transmittance and U value.
Single glass has about a 21% increase in transmittance and 124% increase in heat loss, over double glass, and thus
should only be used in lieu of double glass when night insulation is used. However, triple glass has approximately an
18% reduction in transmittance and a 37% reduction in heat loss over double glass and may be used in lieu of
double glass with night insulation, to increase economic feasibility at only a slight reduction in overall performance.
When considering such trade-offs it is extremely important to consider the specific environmental factors at the site.
Triple glass in lieu of double glass may be excellent in areas of high solar radiation and cold exterior temperatures,
but not effective in areas of low solar radiation and moderate exterior temperatures. The effectiveness should be
determined by steady-state heat loss calculations and passive solar performance analysis.
aValues may vary with material, see manufacturers recommendations.

Durability

The durability of the glazing material must be such that it resists failure due to thermal stresses and that it does not
degrade or discolor when exposed to solar radiation for extended periods of time. Of course these considerations
will vary with economics. If a material degrades frequently but is inexpensive to replace, it may be more economical
than a more expensive, more durable glazing material. Additional considerations in selection of a glazing material
may be resistance to impact, and ease of replacement due to breakage.
No matter what glazing material is selected, the designer should be assured it will maintain an expected condition
that is not detrimental to the performance of the passive solar energy system for its intended period of use.
Materials that discolor or degrade rapidly and have significant reduction of solar radiation transmittance should not
43d http://www.gobrick.com/BIA/technotes/t43d.htm
11 of 13 9/13/2009 2:13 PM
be used.

METRIC CONVERSION

Because of the possible confusion inherent in showing dual unit systems in the calculations, the metric (Sl) units are
not given in the examples. Table 13 in Technical Notes 4 provides metric (Sl) conversion factors for the more
commonly used heat transmission units.

SUMMARY

This Technical Notes provides information on the component materials for passive solar energy system applications.
This offers a designer or owner sufficient information regarding the material properties of brick to assist in the design
and use of brick masonry in passive solar applications. Consideration of the properties of brick masonry could result
in a thermal storage media that is an aesthetic, durable, maintenance free, fire resistant structural component of a
building that provides sound transmission reduction. In addition, sufficient values of the material properties of brick
masonry are provided for passive solar energy system analysis techniques, manual or computer calculations. The
decision to use the information and concepts presented in this Technical Notes is not within the purview of the Brick
Institute of America, and must rest with the designer or owner of any specific project.

REFERENCES

1.ASHRAE Handbook and Product Directory, 1977, Fundamentals Volume, American Society of Heating,
Refrigerating and Air-Conditioning Engineers, Inc., 345 East 47th Street, New York City, New York 10017.
2
.Brick Masonry for Thermal Storage by Stephen S. Szoke, a paper presented at "Passive Solar Building
Construction Program," 21-22 November 1980, Madison, Wisconsin.
3.Brick Walls for Passive Solar Use by G. C. Robinson, C. C. Fain, Stephen M. Jansen and Paul Harshman,
February 1980, Clemson University, South Carolina.
4.Chemical Engineers' Handbook prepared by a staff of Specialists, John H. Perry, Ph.D., Editor, Third
Edition, 1950, McGraw-Hill Book Company, Inc., New York, Toronto, and London.
5.The Chemistry and Physics of Clays and Other Ceramic Materials, by Alfred B. Searle and Rex W.
Grimshaw, Third Edition, 1959, Ernest Benn Limited, London, England.
6.Heating and Ventilating's Engineering Handbook by Clifford Strock, First Edition, 1948, The Industrial
Press, New York City, New York.
7.Proceedings of the Solar Glazing 1979 Topical Conference 22-23 June 1979 Stockton State College,
Pomona, New Jersey, Sponsored by the Mid-Atlantic Solar Energy Association.
8.Properties of Engineering Materials by Glenn Murphy, C.E., Ph.D., Second Edition, 1952, International
Textbook Company, Scranton, Pennsylvania.
43d http://www.gobrick.com/BIA/technotes/t43d.htm
12 of 13 9/13/2009 2:13 PM
9.Smithsonian Physical Tables by William Elmer Forsythe, Ninth Edition, 1956, Smithsonian Institute,
Washington, D.C.
10.Thermal Environmental Engineering by James L. Threlkeld, Second Edition, 1970, Prentice-Hall Inc..
Englewood Cliffs. New Jersey.
43d http://www.gobrick.com/BIA/technotes/t43d.htm
13 of 13 9/13/2009 2:13 PM

Technical Notes 43G - Brick Passive Solar Heating Systems - Part 7 - Details and Construction
[Mar./Apr. 1981] (Reissued Sept. 1986)
Abstract: Details and construction of brick masonry for passive solar energy system applications vary only slightly
from conventional residential and commercial brick masonry construction. Typical construction details are provided
for direct gain and thermal storage wall systems. These details, with slight modifications, are also applicable for
attached sunspaces. Construction variations from conventional construction and considerations for compliance with
the major model building codes are also discussed.

Key Words: attached sunspaces, bricks, building code requirements, details, direct gain systems, energy,
masonry, passive solar energy systems, thermal storage wall systems.

INTRODUCTION

Brick masonry construction and recommended details for passive solar energy systems are similar to conventional
residential and commercial brick masonry construction and details. The general concepts of direct gain systems,
attached sunspaces and thermal storage wall systems are discussed in Technical Notes 43. Empirical sizing,
rational approaches for determining the thickness of brick masonry as a storage medium, material properties and
performance calculations are discussed in other Technical Notes in this series. In these passive solar applications,
brick masonry may be used as a storage medium and structural component of the building. Brick masonry also
offers the capability for esthetic designs, fire resistance and sound transmission reduction.
These recommended details are presented in an effort to show as many applications of brick masonry in passive
solar heating systems as possible and are not offered as typical combinations of details. The details may be slightly
varied and different combinations of the details may be used to satisfy the requirements of any specific passive solar
heated building design.

DIRECT GAIN
General

Details for brick masonry floors and walls used for thermal storage in direct gain systems are provided in Figs. 1
through 3. Each of these figures shows a typical connection detail for the ground floor, interim floors and roof.

Exterior Loadbearing Walls
Exterior loadbearing brick masonry walls may be constructed as insulated cavity walls to provide an interior brick
masonry wythe for thermal storage and an exterior brick masonry wythe for durability, as shown in Fig. 1. The brick
43g http://www.gobrick.com/BIA/technotes/t43g.htm
1 of 15 9/13/2009 2:14 PM
masonry should be continuous through all floor intersections so that the brick masonry bears on the foundation or
foundation wall, provides adequate support and complies with building code fire safety requirements. Where wood
joists frame into brick masonry wall construction, the wood joists should be fire cut.


Cavity Wall Construction
FIG. 1


The design should consider the local code requirements for minimum bearing. A thicker interior wythe may be
required for bearing or special provisions incorporated into the detail so that both the exterior and interior wythes
may be used for bearing. Bearing on both wythes should only be used when other alternatives are not practical,
43g http://www.gobrick.com/BIA/technotes/t43g.htm
2 of 15 9/13/2009 2:14 PM
since there may be difficulty in properly constructing and detailing such a connection without interfering with the
performance of the cavity wall. Additional information on the design, detailing, construction and insulating of cavity
wall construction is provided in Technical Notes 21 Series.
Providing clerestories with the appropriate pitched roof in conjunction with a cathedral type ceiling and exposed
beams or trusswork may allow even the North wall to be exposed to sunlight and used for thermal storage. This type
of detailing may require consideration of exposed trusses in the roof/ceiling component. The trusses or other means
of eliminating the thrust at the top of the cavity wall is necessary because the building codes do not allow lateral
thrust on cavity wall construction. When considering the use of trusses or other members to relieve a cavity wall of
this thrust, the spacing of the trusses or other members should be such that the interior of the wall is subjected to
only minimal shading if it is to be used as thermal storage for direct gain.
When considering the use of insulated cavity walls, the exterior wythe of brick masonry is thermally isolated from the
rest of the wall system. Thus, the exterior wythe of the cavity wall is usually subjected to greater temperature
fluctuations than the interior wythe used for thermal storage. For cavity wall construction, both the interior and
exterior wythes may require expansion joints for thermal movement.

Exterior Non-Loadbearing Walls
Cavity wall construction may also be used for exterior non-loadbearing walls. East or West-facing walls may be
positioned in the structure so that they are exposed to morning or afternoon sunlight for direct gain storage.
Typically, passive solar buildings require a large amount of additional interior mass which may be unexposed to
direct sunlight. This mass provides supplementary thermal storage, resulting in a thermal flywheel for reduced
interior temperature fluctuations. The interior wythe of the cavity wall may be considered when determining the
amount of additional mass.

Interior Loadbearing Walls
Typical details for interior loadbearing brick masonry walls are shown in Fig. 2. These details are similar to
conventional loadbearing construction. Wood floor joists bearing on the brick should be fire cut.

43g http://www.gobrick.com/BIA/technotes/t43g.htm
3 of 15 9/13/2009 2:14 PM

Interior Loadbearing
FIG. 2


A roof construction detail is provided, offering the option to use a skylight to expose the brick masonry loadbearing
wall to sunlight. The interior brick masonry wall may be exposed to direct sunlight through South-facing windows and
doors, or a clerestory may be used, depending on the distance from the South-facing wall.
The use of interior loadbearing brick masonry construction does not require any special consideration over and
above conventional construction. The only exception is that provisions for thermal expansion may be required.

43g http://www.gobrick.com/BIA/technotes/t43g.htm
4 of 15 9/13/2009 2:14 PM
Interior Non-Loadbearing Walls
Interior non-loadbearing brick wall construction is quite similar to conventional brick veneer construction. The brick
veneer should be constructed as shown in Fig. 3. The brick masonry should be continuous through all floor
intersections so that all the brick masonry bears on the foundation or foundation wall and complies with building code
fire safety requirements. Additional information on brick veneer construction is provided in Technical Notes 28
Series.



Interior Non-Loadbearing
FIG. 3

43g http://www.gobrick.com/BIA/technotes/t43g.htm
5 of 15 9/13/2009 2:14 PM

The requirements in Technical Notes 28 Series apply to interior brick veneer construction, except that the
requirements for the effect of weathering may be disregarded. The backup material; wood frame, metal stud, etc.,
should be constructed as in conventional construction.
If the interior brick veneer is constructed with wood frame or metal studs without sheathing between the backup and
the brick veneer, the 1-in. airspace between the brick and the backup, as recommended in Technical Notes 28
Series, may be eliminated. If the brick veneer is constructed with a framing system that requires sheathing on the
side to be veneered, it is recommended that a 1-in. airspace be maintained. This provides ''finger room" to facilitate
the laying of the brick. The use of the sheathing on the side of the backup material which is to be veneered may be
required to provide the appropriate structural rigidity of the backup system. This sheathing may also be used to
increase the fire resistance and sound transmission classification of the wall.

Interior Flooring
Typical details for brick flooring are provided in Figs. 1 through 6. The interim floor details show mortarless paving,
and the ground floor details show brick masonry set in a mortar bed. These details are interchangeable. The interim
floor detail shown in Fig. 3 is a typical detail for mortarless brick paving in a sand bed. The difference in thermal
performance of mortarless paving as compared to paving units set in a mortar bed is insignificant. There may be a
slight reduction in heat transfer from unit to unit, but this will typically have a negligible effect on overall thermal
performance of the floor system being used as direct gain thermal storage. Paving units are used as the flooring in
the thermal storage wall details, Figs. 4 through 6. These paving units in combination with glazing incorporated into
the thermal storage wall for daylighting and visual contact with the exterior may be used to form a direct gain
system. The brick flooring may also be used to achieve the additional interior mass required by many passive solar
heated buildings.

43g http://www.gobrick.com/BIA/technotes/t43g.htm
6 of 15 9/13/2009 2:14 PM

Solid Brick Thermal Storage Wall
FIG. 4


43g http://www.gobrick.com/BIA/technotes/t43g.htm
7 of 15 9/13/2009 2:14 PM

Grouted Hollow Thermal Storage Walls
FIG. 5


43g http://www.gobrick.com/BIA/technotes/t43g.htm
8 of 15 9/13/2009 2:14 PM

The Use of Hollow Brick in Thermal Storage Walls
FIG. 6


A soft joint should be installed around the perimeter of the brick paving, mortarless or set in a mortar bed, to provide
relief of the stresses due to thermal movement, deflection and differential movement between the brick flooring and
adjacent construction. Additional soft joints may be required for thermal expansion.
Supporting brick masonry paving on floor systems requires sufficient stiffness of the system to adequately support
the additional weight in such a manner as to satisfy the minimum deflection requirements of the brick paving. For
mortarless brick paving, the maximum deflection should be less than or equal to L/360. For brick paving set in a
mortar bed, the maximum deflection should be less than or equal to L/600. For wood floor systems supporting brick
flooring, the sizing and spacing of the floor joists should be adequate to support the additional weight, satisfy the
floor joist structural requirements and the deflection requirements of the brick flooring.
The floor connection details shown in Figs. 2 and 3 should be such that the top surface of the brick flooring is level
43g http://www.gobrick.com/BIA/technotes/t43g.htm
9 of 15 9/13/2009 2:14 PM
with other floor finishes. If this is not desirable, or possible, the appropriate riser distance between the surfaces of
the different floor finishes should be provided to comply with the governing building code.
The use of brick paving is discussed in detail in Technical Notes 14B. The brick flooring may also be constructed by
laying face brick in a rowlock position. Another option is to use reinforced brick masonry floors, as discussed in
Technical Notes 14B.


THERMAL STORAGE WALLS
General

Figures 4 through 6 show the thermal storage wall being used as a structural component of the building, supporting
various floor and roof systems. These combinations are not typical, but are offered to demonstrate the various
alternatives available.
The thickness required for thermal storage walls is usually sufficient for the wall to be used as a loadbearing
component of the building without any special considerations. However, it may be necessary to check the structural
adequacy of the wall.
The thermal storage wall may be several wythes of solid brick, as shown in Fig. 4; solid through-the-wall units; a
grouted cavity wall system, as shown in Fig. 5; or grouted hollow units or combinations of grouted hollow units and
solid units, as shown in Fig. 6.

Details

Details for solid brick thermal storage walls are shown in Fig. 4. Corbeling the thermal storage wall to provide
support for the exterior glazing is one way to eliminate the need for thick foundation walls. Brick masonry may be
laid as projected headers to provide a durable support for attaching the glazing assembly to the wall. This eliminates
the use of combustible materials exposed to high temperatures for extended periods of time. Projected headers may
provide a durable non-combustible, horizontal separation between individual floors for multi-story vented thermal
storage wall systems. This may be used to comply with the building code requirements regarding the fire-stopping of
plenums. Vertical separation to provide a means of closing the sides of the thermal storage wall air space may also
be achieved with projected headers.
Depending upon the structural loads imposed on the projected headers and to avoid exposing cores, corbeling may
be required, as shown in Fig. 4. The air space between the glazing and the thermal storage wall should be of a
thickness that satisfies the building code requirements for unreinforced corbeling. If these limitations cannot be met,
an alternate means of support for the glazing will be required.
Additional glazing is provided in each detail to show that the thermal storage wall need not be a solid barrier
eliminating any view of the exterior or daylighting. This glazing may be used as a direct gain collector with interior
brick masonry floors and walls as the thermal storage.
The air space between exterior glazing and the thermal storage wall may be interrupted at various intervals and the
thermal storage wall made discontinuous, as shown in Fig. 7. This may be used to incorporate direct gain and
thermal storage walls into a combined system.
Operable or stationary shading devices may be attached to the structural framing of the glazing assemblies. The
43g http://www.gobrick.com/BIA/technotes/t43g.htm
10 of 15 9/13/2009 2:14 PM
glazing assemblies should be sufficiently anchored to the brick masonry to accommodate these additional loads.
a Solar Savings Fraction as determined by using Method I of Technical Notes 43B
Vents
If the thermal storage wall is to be vented, each opening through the thermal storage wall should be approximately
64 sq in. The length of the opening should be about 4 times the height of the opening. The vents should occur as
sets, one at the top of the wall directly over one at the bottom of the wall, to facilitate air flow. The number of sets of
vents may be approximated by using Equation 1.
nv' = Fv [(lw X hW) / (lv X hv)] (1)
where: nv' = approximate number of sets of vents.
Fv = vent area factor from Table 1.
lw = length of the vented thermal storage wall, in ft.
hw = height of the vented thermal storage wall, in ft.
Iv = length of the vent opening, in inches, approximately 4 X hv
hv = height of the vent opening, in inches.


The actual number of sets of vents to be installed, nv, should be a whole number. Performance tends to decrease as
the percentage of vent area to wall area increases. The next lower whole number to nv' should typically be used as
the actual number of vents to be installed, if nv' is less than nv plus 0.70. If the value of nv' is greater than or equal to
nv plus 0.70, the next larger whole number would typically be used as the number of sets of vents to be installed.

Both the vertical and horizontal spacing of vents will also affect performance. Top vents should be located as close
43g http://www.gobrick.com/BIA/technotes/t43g.htm
11 of 15 9/13/2009 2:14 PM
to the ceiling as possible and the bottom vents as close to the floor as possible. The vertical distance between top
and bottom vents should be at least 6 ft for full story height vented thermal storage walls.
The horizontal spacing of vents, sv, may be determined by using Equation 2.
sv = lw / nv (2)

Example. A 25-ft long vented thermal storage wall system 8 ft high is expected to supply about 35% of a building's
heating load, SSF = 0.35. Vent openings are formed by omitting one and one-half courses of standard modular brick
vertically and two standard modular brick horizontally as shown in Fig. 7. Thus, the opening has a height of about 4
in. and a length of about 16 in. The dimensions of the vent opening satisfy the criteria of the length being
approximately 4 times the height and the area being approximately 64 sq in. If other size brick are used, the courses
and number of brick omitted to meet the area and height-to-length requirements of the vent opening will vary.



Locating Vents
FIG. 7
Using Equation 1, and Fv = 1.58 from Table 1 for a solar savings fraction of 0.35, the approximate number of vents
would be:
nv' = 1.58 (25 X 8) / (16 X 4)
nv' = 4.94
43g http://www.gobrick.com/BIA/technotes/t43g.htm
12 of 15 9/13/2009 2:14 PM
Thus, nv is 5 sets of vents to be installed.

The horizontal spacing of vents may be approximated by using Equation 2.
sv = 25/5 = 5 ft

The result is that the 25 ft long vented thermal storage wall should have 5 sets of top and bottom vents, each having
an opening of approximately 64 sq in., spaced horizontally at 5 ft o.c.

ATTACHED SUNSPACES

The typical details for direct gain thermal storage and thermal storage walls may be used for attached sunspaces
with only modifications to the glazing details. Depending on the type of attached sunspace, the thermal storage
components may be direct gain floors and walls or direct gain floors and vented or unvented thermal storage walls.

CONSTRUCTION

Solid brick masonry used as a thermal storage medium, as in all brick masonry construction, requires that all head,
bed and collar joints be solidly filled with mortar. Solid brick are units which are cored less than 25 percent of the
gross cross-sectional area parallel to the bedding plane. In typical running bond or stack bond construction, the brick
should be shoved into full bed joints. This results in sufficiently filled cores so that there is little or no effect on the
overall thermal performance of the wall. When soldier courses or projected headers are being considered, uncored
units may be preferred.
Hollow brick are brick units in which the coring is less than 40 percent and greater than 25 percent of the gross
cross-sectional area in the bedding plane. Hollow brick masonry used for thermal storage requires all head and
collar joints and bedding surfaces to be solidly filled with mortar and all cores fully grouted. Projected headers and
corbels may best be achieved by combining solid brick masonry with hollow brick masonry construction.
Grouted hollow walls are discussed in Technical Notes 17, 17C and 17D. When considering the use of grouted
hollow walls, constructed of two wythes of brick separated by a fully grouted space, the only control over thickness
will be requirements for adequate thermal storage. Thus, grouted hollow brick masonry walls may be advantageous
when the thickness desired is not easily achieved by using modular sizes of brick. As in all brick masonry
construction, the brick wythes should have all head and bed joints solidly filled with mortar.

MISCELLANEOUS CONSIDERATIONS
Thermal Expansion

For most applications of brick masonry as interior direct gain thermal storage, the temperatures within the brick
masonry will probably range from 72 to 96F. Thus, thermal expansion will not normally be a problem except where
long interior walls or floors are used. Interior brick masonry used for direct gain thermal storage occurring in lengths
43g http://www.gobrick.com/BIA/technotes/t43g.htm
13 of 15 9/13/2009 2:14 PM
longer than 100 ft or exposed to a higher maximum temperature should be analyzed for thermal expansion. The
thermal expansion of brick masonry is discussed in Technical Notes 18A.
Thermal storage walls may be subjected to larger temperature fluctuations than direct gain thermal storage
components. Usually, the difference between the maximum temperature and minimum temperature at the center of
the thermal storage wall is small and no provision for thermal expansion is necessary. Generally, thermal expansion
need only be considered for long or high thermal storage walls or for walls exposed to extreme temperature
fluctuations.
The maximum mean temperature of brick thermal storage walls may be determined by using the temperature
fluctuation equation in Technical Notes 43. The minimum mean temperature may be determined by using the
steady-state temperature gradient through the wall as discussed in Technical Notes 7C.

Flashing
Flashing brick masonry thermal storage components is usually not required because the brick masonry is on the
interior of the building. Cavity walls will require flashing as discussed and shown in Technical Notes 21B. Flashing
may be required for thermal storage wall systems, depending on the type of glazing assembly and how it is mounted
in front of the thermal storage wall.

Reinforced Brick Masonry
Reinforced brick masonry, as discussed in Technical Notes 17 Series, may be required depending on the structural
design loads. For thermal storage walls, this is easily accomplished by using reinforced grouted hollow brick or
reinforced hollow wall construction. Reinforced brick masonry may be designed so that the wall will be able to
sustain lateral thrust.
Typically in reinforced brick masonry construction, the reinforcement is both horizontal and vertical, placed as near to
the center of the wall as practical. This, in combination with the minimum required spacing to sufficiently reinforce a
brick masonry wall does not result in any significant decrease in the wall's thermal performance due to thermal
bridges.

Lintels and Sills
The thermal storage wall details provide several options for constructing lintels and sills. Additional information on
lintels is provided in Technical Notes 17H and 31B. Information regarding the construction of brick masonry arches is
provided in Technical Notes 31 Series. Brick masonry sill details are provided in Technical Notes 36 Series,
however, most sills for thermal storage walls do not require a sloped top surface or a drip since they are not
exposed to exterior weather.

Fireplaces
Interior fireplaces may be used to obtain additional mass to decrease the interior temperature fluctuations. Brick
masonry fireplaces may be incorporated in the thermal storage component in any of these passive solar heating
systems. A fireplace may be used for direct gain storage or may be constructed in a thermal storage wall. The
design and construction of fireplaces is discussed in Technical Notes 19 Series.

Glazing
43g http://www.gobrick.com/BIA/technotes/t43g.htm
14 of 15 9/13/2009 2:14 PM
It is desirable that the glazing component of these passive solar energy systems be operable to facilitate cleaning,
exhausting excess heat, providing a means of egress or a combination of these. The glazing may be sliding glass
doors, awning type windows, hinged glass doors or other options. Hinged doors installed vertically or horizontally
may greatly facilitate the cleaning of vented or unvented thermal storage wall collectors.
Depending upon building classification, building codes may require a 3-ft vertical separation between openings
located vertically one above the other. This is not typically a requirement for residential buildings, or any building
under 3 stories in height.

METRIC CONVERSION

Because of the possible confusion inherent in showing dual unit systems in the calculations, the metric (SI) units are
not given in this Technical Notes. Table 13 in Technical Notes 4 provides metric (SI) conversion factors friar the
more commonly used units.

SUMMARY

This Technical Notes provides information on the construction and detailing of brick masonry thermal storage
components for passive solar energy systems. The information, recommendations and details contained in this
Technical Notes are based on the available data and experience of the Institute's technical staff. They should be
recognized as suggestions and recommendations for the consideration of the designers and owners of buildings
when using brick in passive solar energy applications.
All of the possible variations cannot be covered in a single Technical Notes. However, it is believed that the
information is presented in a form such that specific details are interchangeable. The final decision for details to be
used is not within the purview of the Brick Institute of America, and must rest with the project designer, owner or
both.

REFERENCES

1. Passive Solar Design Handbook, Volume Two of Two Volumes: Passive Solar Design Analysis, January
1980, prepared by Los Alamos Scientific Laboratory, University of California, J. Douglas Balcomb, Dennis
Barley, Robert McFarland, Joseph Perry, Jr., William Wray and Scott Noll, prepared for the U.S. Department
of Energy, Office of Solar Applications, Passive and Hybrid Solar Buildings Program, Washington, D.C.
43g http://www.gobrick.com/BIA/technotes/t43g.htm
15 of 15 9/13/2009 2:14 PM

Technical Notes 44 - Anchor Bolts for Brick Masonry
April 1986
Abstract: Anchor bolts are used extensively in brick masonry to make structural attachments and connections. To
date, a limited amount of information has been available to aid designers in the selection and design of anchor bolts
in brick masonry. This Technical Notes addresses the types of anchor bolts available, detailing of anchor bolt
placement and suggested design procedures. A discussion of current and proposed codes and standards is also
presented.

Key Words: anchors, bolts, conventional anchors (bent bar, plate, sleeve, wedge), edge distance, headed bolts,
loads, proprietary anchors (adhesive, expansion), shear, tension, through bolts.

INTRODUCTION

Anchor bolts are used in masonry construction with few or no guidelines for the practicing designer to follow. This
Technical Notes offers basic information covering 1) the types of anchor bolts available for structural applications in
brick masonry, 2) typical details of proper anchor bolt installation, 3) suggested allowable anchor bolt design loads
and 4) the current and proposed codes and standards governing anchor bolts in brick masonry construction.
In new masonry construction, anchor bolts are commonly embedded in walls and columns to support beams, plates
and ledgers. In prefabricated panel construction, anchor bolts are used to facilitate connections to the structural
frame. Renovation and rehabilitation of existing masonry structures usually require that anchor bolts be used to
attach stair risers, elevator tracks and various frame assemblages for equipment installation. This is only a fraction
of the possible uses of anchor bolts in masonry construction and with the increase of new, innovative architectural
masonry designs, the uses of anchor bolts in masonry construction are likely to increase.
This Technical Notes is the first in a series on masonry anchors, fasteners and ties, and addresses anchor bolts for
brick masonry. Other Technical Notes in this series will address brick masonry fasteners and ties.

ANCHOR BOLT TYPES

Anchor bolts can be divided into two major groups: conventional (unpatented) anchors and proprietary (patented)
anchors. Conventional anchor bolts are usually embedded in the masonry during construction and require careful
attention to bolt location and grip length requirements to avoid problems with connection alignment and erection.
Proprietary anchors, however, are typically installed after completion of construction and therefore, permit a larger
degree of freedom in anchor placement. For this reason, proprietary anchors are becoming popular in masonry
construction and add new concerns in the area of anchor bolt design.

44 http://www.gobrick.com/BIA/technotes/t44.htm
1 of 24 9/13/2009 2:15 PM
Conventional Anchor Bolts
Conventional bolts are usually made to the specific project requirements by steel fabricators or they may be
purchased in standard sizes (diameters and lengths) from steel suppliers. The availability and cost of conventional
bolts are generally based on demand and fabrication requirements. The types of conventional anchor bolts most
often used are discussed below.
Headed Bolts. Square or hex-headed ASTM A 307 bolts are frequently used as anchor bolts due to their wide
availability and relatively low cost (see Figure 1). Higher strength bolts, such as ASTM A 325 bolts, are available and
can be used, but are more expensive. A washer placed against the bolt head is often used with the intention of
increasing the bearing area and thus increasing the anchor strength. However, the actual strength increase obtained
by adding a washer is small, if any, and under certain conditions (small edge distances), may actually decrease the
tensile strength.


Headed Bolts
FIG. 1

Bent Bar Anchors. Bent bar anchors, frequently used in masonry construction, are usually made in "J" or "L"
shapes (see Fig. 2). Even though the "J" and "L" shapes are the more popular, a variety of shapes (see Fig. 3) is
available since there currently is no standard governing the geometric properties of bent bar anchors. These anchors
are usually made from ASTM A 36 bar stock and are shop-threaded.

44 http://www.gobrick.com/BIA/technotes/t44.htm
2 of 24 9/13/2009 2:15 PM

"L" and "J" Bent Bar Anchors
FIG. 2

44 http://www.gobrick.com/BIA/technotes/t44.htm
3 of 24 9/13/2009 2:15 PM

Other Bent Bar Anchors
FIG. 3

Plate Anchors. Plate anchors are usually made by welding a square of circular steel plate perpendicular to the axis
of a steel bar that is threaded on the opposite end (see Fig. 4). There are no standards governing the dimensions
(length, width or diameter) of the plate. The American Institute of Steel Construction does limit the fillet weld size
based on the plate thickness (see Table 1). Both the plate and bar are usually made from ASTM A 36 steel.


44 http://www.gobrick.com/BIA/technotes/t44.htm
4 of 24 9/13/2009 2:15 PM
Plate Anchors
FIG. 4

Through Bolts. As the name implies, through bolts extend completely through the thickness of the masonry and are
composed of a threaded rod or bar with a bearing plate located on the surface opposite the attachment (see Fig. 5).
In the early 1900's, through bolts were used in loadbearing masonry structures to tie floor and wall systems
together. Often decorative cast bearing plates were used since through bolts were visible on the exterior masonry
surfaces (see Fig. 6). Today, through bolts are primarily used in industrial construction where aesthetics are not a
principal concern, or in retrofitting existing structures. Through bolt rods are usually made from ASTM A 307
threaded rod or threaded ASTM A 36 bar stock. Bearing plates are typically made from ASTM A 36 steel plate.


Through Bolt
FIG. 5



44 http://www.gobrick.com/BIA/technotes/t44.htm
5 of 24 9/13/2009 2:15 PM

Decorative Through Bolt Bearing Plate
FIG. 6


* American Institute of Steel Construction

Proprietary Anchor Bolts

Proprietary anchors are available through a number of manufacturers under numerous brand names. Although the
style and physical appearance of the anchors differ between manufacturers, the basic theories behind the anchors
are very similar. For this reason, proprietary anchors can be divided into two generic categories: expansion-type
anchors and adhesive or chemical-type anchors.
Expansion Anchors. Two different types of expansion anchors are generally recommended by their manufacturers
for use in brick masonry: the wedge anchor and the sleeve anchor (see Fig. 7). These anchors develop their
strength by means of expansion into the base material. Wedge anchors develop their hold by means of a wedge or
wedges that are forced into the base material when the bolt is tightened. The wedges create large point bearing
stresses within the hole; therefore, this anchor requires a solid base material to develop its full capacity. For this
reason, voids formed by brick cores and partially filled mortar joints in some brick masonry may make the
construction unsuitable for wedge anchor installation.

44 http://www.gobrick.com/BIA/technotes/t44.htm
6 of 24 9/13/2009 2:15 PM

Proprietary Expansion Anchors
FIG. 7



Sleeve anchors develop their strength by the expansion of a cylindrical metal sleeve or shield into the base material
as the bolt is tightened. The expansion of the sleeve along the length of the anchor provides a larger bearing surface
than the wedge anchor, and is less affected by irregularities and voids in the base material than is the wedge
anchor. For this reason, sleeve anchors are recommended by their manufacturers for use in brick masonry more
often than wedge anchors.
Drop-in and self-drilling anchors (see Fig. 8) are two other types of expansion anchors available, but are typically not
recommended by their manufacturers for use in masonry. The reason for this is due to the embedment and setting
characteristics of the two anchors. Both anchors are produced to allow shallow embedment depths and are
expanded or set by an impact setting tool. The combination of shallow embedment and high stresses imparted by
the expansion tend to cause cracking or splitting in masonry. Depending on the extent of cracking or splitting, the
anchor could experience a reduction in load-carrying capacity or undergo complete failure during installation.

44 http://www.gobrick.com/BIA/technotes/t44.htm
7 of 24 9/13/2009 2:15 PM

Other Proprietary Expansion Anchors
FIG. 8

There are several considerations that should be examined when contemplating the use of expansion-type anchors in
brick masonry. These are: 1) Expansion anchors should not be used to resist vibratory loads. Vibratory loads tend to
loosen expansion anchors. 2) Specific torques are required to set expansion anchors. Excessive torque can reduce
anchor strength or may lead to failure as excessive torque is applied. 3) Expansion anchors require solid, hard
embedment material to develop their maximum capacities. Some brick construction may not provide a good
embedment material due to voids formed by brick cores and partially filled mortar joints.
Adhesive Anchors. Two basic types of adhesive anchors are currently available. The major difference between the
two is that one anchor is manufactured as a pre-mixed, self-contained system, whereas the second type requires
measurement and mixing of the epoxy materials at the time of installation. The more popular self-contained types
use a double glass vial system (see Fig. 9) to contain the epoxy. The outer vial contains a resin and the inner vial
contains a hardener and aggregate. The glass vial is placed in a pre-drilled hole and a threaded rod or bar is driven
into the hole with a rotary hammer drill, breaking the vials and mixing the adhesive components. The other type of
adhesive anchor requires that the epoxy components be hand-measured and mixed before the epoxy is placed into a
pre-drilled hole. A threaded rod or bar is then set into the epoxy mixture, as shown in Fig. 10. Adhesive epoxies
usually vary slightly between manufacturers, but the steel rods or bars are typically ASTM A 307 or ASTM A 325
threaded rod, or ASTM A 36 shop-threaded bar.


Self-Contained Adhesive Anchor
FIG. 9


44 http://www.gobrick.com/BIA/technotes/t44.htm
8 of 24 9/13/2009 2:15 PM



Site-Mixed Adhesive Anchor
FIG. 10

There are special requirements and limitations. that should be considered when contemplating the use of adhesive
anchors in brick masonry. They are: 1) Specially designed mixing and/or setting equipment may be required. 2) Dust
and debris must be removed from the pre-drilled holes to insure proper bond between the adhesive and base
material. 3) The adhesive mixture tends to fill small voids and irregularities in the base material. 4) Large voids (due
to brick cores, intentional air spaces and partially filled joints) may cause reductions in anchor capacities. This is
especially true with the self-contained adhesive anchors since a limited volume of epoxy is available to fill the voids
and provide a bond to the anchor. 5) The adhesive bond strength is reduced at elevated temperatures and may also
be adversely affected by some chemicals (see Table 2 and Fig. 11).
44 http://www.gobrick.com/BIA/technotes/t44.htm
9 of 24 9/13/2009 2:15 PM
aThe manufacturer should always be consulted when adhesive anchors are to be used in areas where contact with chemicals is likely.


Effect of Temperature on Ultimate Tensile Capacity
FIG. 11

INSTALLATION DETAILS
Conventional Anchor Bolts

Typical embedment details for each type of conventional anchor used in grouted collar joint construction are shown in
Fig. 12. The conventional embedded anchors (headed bolts, bent bar and plate anchors) are usually placed at the
intersection of a head joint and bed joint. By using this location, the brick units adjacent to the anchor can be chipped
44 http://www.gobrick.com/BIA/technotes/t44.htm
10 of 24 9/13/2009 2:15 PM
or cut to accept the anchor without altering the joint thickness.




Conventional Anchors in Grouted Collar Joints
FIG. 12

Typical embedment details of conventional anchors in multi-wythe brick construction are shown in Fig. 13. A brick, or
portion of a brick, is left out of the inner wythe to form a cell for the embedded anchor (Fig. 14). After the anchor is
placed, the cell is filled with mortar or grout prior to placement of the next course.



44 http://www.gobrick.com/BIA/technotes/t44.htm
11 of 24 9/13/2009 2:15 PM

Conventional Anchors in Multi-Wythe Brick Masonry
FIG. 13




Plan View of Grout Cell in Multi-Wythe Brick Masonry
44 http://www.gobrick.com/BIA/technotes/t44.htm
12 of 24 9/13/2009 2:15 PM
FIG. 14

In hollow brick construction, the units are laid so that the cells are aligned and provide continuous channels for
reinforcing steel placement and for grouting. Depending on the design, every cell or intermittent cells may be
reinforced and grouted (see Technical Notes 41 Revised). The anchor embedment detail will depend on the
reinforcing pattern used in the construction. Figure 15 shows typical embedment details for conventional anchors
embedded between reinforcing cells. The anchor should be solidly surrounded vertically and horizontally by grout for
a minimum distance of twice the embedment depth (1b) (Figs. 14 and 15) for full tension cone development. The
tension cone theory is discussed in following sections. This may require that some cells be partially grouted. A wire
mesh screen can be placed in the bed joint across cells that are to be partially grouted to restrict the grout flow
beyond a certain point. Figure 16 shows typical embedment details for conventional anchors embedded in reinforced
cells. In this detail, the anchor may be tied with wire to the reinforcing to secure the anchor during the grouting
process. Again, the anchor should be solidly surrounded by grout to a minimum distance of twice the actual anchor
embedment depth, both vertically and horizontally.




Conventional Anchors in Reinforced Hollow Brick
FIG. 15



44 http://www.gobrick.com/BIA/technotes/t44.htm
13 of 24 9/13/2009 2:15 PM

Conventional Anchors in Partially Grouted Hollow Brick
FIG. 16

Two typical embedment details for conventionally embedded anchor bolts installed in composite brick and concrete
block construction are shown in Fig. 17. As shown, anchor bolts may be placed in the collar joint between the brick
and block wythes or placed into cells in the concrete block wythe and grouted into place. In details similar to Fig.
17(a), the anchor bolt type and diameter may be controlled by the width of the collar joint. Collar joints should be a
minimum of 1 in. (25 mm) wide when fine grout is used, or a minimum of 2 in. (50 mm) wide when coarse grout is
used (see Technical Notes 7A Revised). When the collar joint dimension is in the 1 in. (25 mm) range, it may
become difficult to position anchor bolts in the collar joint and maintain the recommended clear distance between the
masonry and the anchor (Fig. 17). The practice of using soaps to accommodate anchors larger than the collar joint
is not recommended because the reduction in the brick masonry thickness around the anchor could lead to strength
reductions. If the anchor dimensions required are larger than the collar joint, a detail similar to that shown in Fig.
17(b) should be considered.



44 http://www.gobrick.com/BIA/technotes/t44.htm
14 of 24 9/13/2009 2:15 PM

Conventional Anchors in Composite Brick/Block Masonry
FIG. 17

Through bolts are typically installed after construction and grouting by drilling through the completed masonry work.
When through bolts are to be installed after construction in reinforced brick masonry, care should be taken during
installation to avoid cutting or damaging reinforcement while drilling the through bolt holes. Reinforcing bar locations
can be identified by specially tooled joints or other marks made during construction.

Proprietary Anchors
Proprietary expansion and adhesive anchors typically require special installation procedures and equipment. The
manufacturer should be contacted to determine the appropriate anchor for a particular application, the correct
installation procedure and if any special installation equipment is required. Improper application and installation of
proprietary anchors may lead to less than satisfactory structural performance.
Typical proprietary anchor details are shown in Fig. 18. It is suggested that proprietary anchors be embedded in
head joints when facing or building brick are used. This reduces the possibility of placing anchors in brick cores that
occur within the thickness of the brick and adjacent to the bed joint surfaces. Anchors set in grouted hollow brick
should be placed in holes drilled in the bed joints so that they intersect grouted cells, or should be placed in holes
drilled through the faces of the units into the grouted cells. As with conventional anchors, proprietary anchors should
be solidly surrounded vertically and horizontally by grout for a minimum distance of twice their embedment depth.

44 http://www.gobrick.com/BIA/technotes/t44.htm
15 of 24 9/13/2009 2:15 PM

Typical Proprietary Anchor Details
FIG. 18

ANCHOR BOLT DESIGN

Anchor bolts are used as a means of tying structural elements together in construction and therefore, provide
continuity in the overall structure. In virtually all applications, anchor bolts are required to resist a combination of
tension and shear loads acting simultaneously due to combinations of imposed dead loads, live loads, wind loads,
seismic loads, thermal loads and impact loads. For this reason, and also to insure safety, anchor bolt details should
receive the same design considerations as would any other structural connection. However, due to a lack of
available research and design guides, anchor bolt designs are based largely on past experience with very little
engineering backup. This situation may lead to conservative, uneconomical designs at one extreme, or
nonconservative designs at the other.
Recently, however, research investigating the strength of conventional and proprietary anchors in masonry has been
completed. Reports have been issued that evaluate anchor performance and suggest equations to predict ultimate
anchor strengths. By combining the research findings with design practices currently used in concrete design,
equations for allowable tension, shear and combined tension/shear loads for plate anchors, headed bolts and bent
bar anchors are under consideration for adoption in the proposed "Building Code Requirements for Masonry
Structures" (ACI/ASCE 530). These equations are outlined below.

Tension
The tensile capacity of an anchor is governed either by the strength of the masonry or by the strength of the anchor
44 http://www.gobrick.com/BIA/technotes/t44.htm
16 of 24 9/13/2009 2:15 PM
material. For example, if the embedded depth of an anchor is small relative to its diameter, a tension cone failure of
the masonry is likely to occur. However, if the embedded depth of the anchor is large relative to its diameter, failure
of the anchor material is likely. For these reasons, the allowable tensile load is based on the smaller of the two loads
calculated for the masonry and anchor material. Thus, the allowable load in tension is the lesser of:


or

where:TA = Allowable tensile load, lb,
Ap = Projected area of the masonry tension cone, in.
2
, f'm = Masonry prism compression strength (In
composite construction, when the masonry cone intersects different materials, f'm should be based on the
weaker material), psi,
AB = Anchor gross cross-sectional area, in.
2
,
fy = Anchor steel yield strength, psi.
The value of Ap in Eq. 1 is the area of a circle formed by a failure surface (masonry cone) assumed to radiate at an
angle of 45
o
(see Fig. 19) from the anchor base. When an anchor is embedded close to a free edge, as shown in
Fig. 20, a full masonry cone cannot be developed and the area Ap must be reduced so as not to over-estimate the
masonry capacity. Thus, the area Ap, in Eq. 1 will be the lesser of:


or

where:Ap = Projected area of the masonry tension cone, in.2,

1b = Effective embedded anchor length, in.,
1be = Distance to a free edge, in.

44 http://www.gobrick.com/BIA/technotes/t44.htm
17 of 24 9/13/2009 2:15 PM

Full Masonry Tension Cone
FIG. 19

44 http://www.gobrick.com/BIA/technotes/t44.htm
18 of 24 9/13/2009 2:15 PM

Reduced Masonry Tension Cone
FIG. 20a




Reduced Masonry Tension Cone
FIG. 20b

The effective anchor embedded length (1b) is the length of embedment measured perpendicular from the surface of
the masonry to the plate or head for plate anchors or headed bolts. The effective embedded length of bent bar bolts
(1b) is the length of embedment measured perpendicular from the surface of the masonry to the bearing surface of
the bent end minus one bolt diameter. Where the projected areas of adjacent anchors overlap, Ap of each bolt is
reduced by one-half of the overlap area. Also, any portion of the projected cone falling across an opening in the
masonry (i.e., holes for pipes or conduits) should be deducted from the value of Ap calculated in Eqs. 3 or 4.

Shear
The allowable shear load is based on the same logic as the allowable tension load. That is, the anchor capacity is
governed by either the masonry strength or the anchor material strength. The distance between an anchor and a
free masonry edge has an effect on the masonry shear capacity. Calculations have shown that for edge distances
less than twelve times the anchor diameter, the masonry shear strength controls the anchor capacity. (Calculations
based on masonry with f'm = 1000 psi and anchor steel yield strength with fy = 60 ksi. Therefore, where the edge
distance equals or exceeds 12 anchor diameters. the allowable shear load is the lesser of:

44 http://www.gobrick.com/BIA/technotes/t44.htm
19 of 24 9/13/2009 2:15 PM

or

where:VA=Allowable shear load, lb.

When anchors are located less than 12 anchor diameters from a free edge, the allowable shear load is determined
by linear interpolation from a value of VA obtained in Eq. 5 at an edge distance of 12 anchor diameters to an
assumed value of zero at an edge distance of 1 in. (25 mm). This takes into consideration the reduction in the
masonry shear capacity due to the edge distance.

Combined Tension and Shear

Allowable combinations of tensile and shear loads are based on a linear interaction equation between the allowable
pure tension and pure shear loads calculated in Eqs. 1, 2, 5 and 6. Anchors subjected to combinations of tension
and shear are designed to satisfy the following equation:
T / TA + V / VA 1.0(Eq. 7)
where:T = Applied tensile load, lb.,

V = Applied shear load, lb.

Proprietary Anchor Bolts

The allowable load equations previously presented are intended for use with plate anchors, headed bolts and bent
bar anchors and have been proposed to the ACI/ASCE 530 Committee on Masonry Structures. However, when the
allowables from these equations are compared to test results for proprietary anchors, they appear to produce
acceptable safety factors.
Allowable Loads. Average factors of safety are 4.0 for tensile tests and 5.0 for shear tests on proprietary anchors.
The combined tension/shear interaction equation produced an average safety factor of 7.0 when compared to test
results on proprietary anchors. Therefore, based on comparison to test results, the allowable load equations
proposed in this Technical Notes are suggested for use in the design of proprietary anchors in brick masonry. The
embedment depth used to calculate the allowable load values should be equal to the embedded depth of the
proprietary anchor.
Edge Distance. Edge distance is of particular concern when expansion anchors are used in brick masonry, due to
44 http://www.gobrick.com/BIA/technotes/t44.htm
20 of 24 9/13/2009 2:15 PM
lateral expansion forces produced when the anchors are tightened. These forces are often large enough to cause
cracking or spelling of the brick when edge distances become small. To date, no research has been conducted in
this area. Therefore, due to the lack of information, it is suggested that a minimum edge distance of 12 in. (300 mm)
be maintained when expansion anchors are installed in brick masonry.

Through Bolts
There are no known published reports available addressing the strength characteristics of through bolts in brick
masonry. However, based on the conservatism in the allowables for bent bar anchors and proprietary anchors, the
allowable load equations should provide acceptable allowable load values for through bolts used in brick masonry.
The embedment depth used to calculate the allowable load values should be taken as equal to the actual thickness
of the masonry.

Current Codes and Standards
At the present time, one model code and one design standard contain provisions for anchor bolt design in brick
masonry. The BIA Standard, Building Code Requirements for Engineered Brick Masonry, and the Uniform Building
Code cover design allowables and embedment depths for anchors loaded in shear. There are no provisions for axial
tensile loads or combined tension/shear loads in these documents. Tables 3 and 4 show the allowable shear loads
and minimum embedment depths from the two documents. The values in Table 4(a) are based on rational analysis
and in Table 4(b) on empirical analysis. As can be seen, the tables are very similar and are generally more
conservative than the allowable shear loads obtained from Eqs. 5 and 6 for the same embedment depths (Table 5).


* From Building Code Requirements for Engineered Brick Masonry, Brick Institute of
America, August 1969.
1
In determining the stresses on brick masonry, the eccentricity due to loaded bolts and
anchors shall be considered.
2
Bolts and anchors shall be solidly embedded in mortar or grout.
3
No engineering or architectural inspection of construction and workmanship.
4
Construction and workmanship inspected by engineer, architect or competent
representative.

44 http://www.gobrick.com/BIA/technotes/t44.htm
21 of 24 9/13/2009 2:15 PM

* Reproduced from the Uniform Building Code, 1985 Edition, Copyright 1985 with permission of the publisher, The International
Conference of Building Officials."

44 http://www.gobrick.com/BIA/technotes/t44.htm
22 of 24 9/13/2009 2:15 PM


*American Concrete Institute/American Society Of Civil Engineers Committee 530 on Masonry
Structures.
1Assuming: f'm = 2,000 psi

ASTM A36 steel fy = 36 ksi
Edge Distance = 12 Bolt Diameters

SUMMARY

This Technical Notes is the first in a series on brick masonry anchors, fasteners and ties. It covers anchor bolt
types, detailing and allowable loads for anchor bolts in brick masonry. Other Technical Notes in this series will
address brick masonry fasteners and ties.

The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the technical staff of the Brick Institute of America. The information and recommendations contained
herein should be used along with good technical judgment and an understanding of the properties of brick masonry.
Final decisions on the use of the information discussed in this Technical Notes are not within the purview of the Brick
Institute of America and must rest with the project designer, owner or both.

REFERENCES

1.Manual of Steel Construction, 8th Edition, American Institute of Steel Construction, Inc., Chicago, Illinois,
1980.
44 http://www.gobrick.com/BIA/technotes/t44.htm
23 of 24 9/13/2009 2:15 PM
2.Whitlock, A.R. and Brown, R.H., Strength of Anchor Bolts in Masonry, NSF Award No. PRF-7806095,
"Cyclic Response of Masonry Anchor Bolts", August 1983.
3.Brown, R.H. and Dalrymple, G.A., Performance of Retrofit Embedments in Brick Masonry, NSF Award
No. CEE-8217638, "Static and Cyclic Behavior of Masonry Retrofit Embedments (Earthquake Engineering)",
Report No. 1, April 1985.
4.Hatzinikolas, M.; Lee, R.; Longworth, J. and Warwaruk, J., "Drilled-In Inserts in Masonry Construction",
Alberta Masonry Institute, Edmonton, Alberta, Canada, October 1983.
5.Building Code Requirements for Engineered Brick Masonry, Brick Institute of America, McLean, Virginia,
August 1969.
6.Uniform Building Code, International Conference of Building Officials, Whittier, California, 1985.
7.Technical Notes on Brick Construction 17 Revised, "Reinforced Brick Masonry, Part I of IV", Brick Institute
of America, McLean, Virginia, October 1981.
8.Technical Notes on Brick Construction 41 Revised, "Hollow Brick Masonry-Introduction", Brick Institute of
America, McLean, Virginia, 1983.
9.Specification for the Design and Construction of Load-Bearing Concrete Masonry, National Concrete
Masonry Association, McLean, Virginia, April 1971.
10.The BOCA Basic/National Building Code, 9th Edition, Building Officials and Code Administrators,
International, Country Club Hills, Illinois, 1984.
11.Standard Building Code, Southern Building Code Congress, International, Inc.. Birmingham, Alabama,
1985.
12.Technical Notes on Brick Construction 7A Revised, "Water Resistance of Brick Masonry-Materials, Part
II of III", Brick Institute of America, Reston, Virginia, 1985.
44 http://www.gobrick.com/BIA/technotes/t44.htm
24 of 24 9/13/2009 2:15 PM

Technical Notes 44A - Fasteners for Brick Masonry
May 1986 (Reissued Aug. 1997)
Abstract: Fasteners are used extensively in brick masonry construction to attach fixtures, equipment and other
objects. This Technical Notes discusses the different types of fasteners used in brick masonry construction, their
applications, appropriate fastener selection based on brick type, fixture weight, environmental exposure and
aesthetics.
Key Words: adhesives, bolts, brick, fasteners, fixtures, hardware, masonry, screws.
INTRODUCTION
This Technical Notes is the second in a series that addresses brick masonry anchor bolts, fasteners and ties. The
term "fastener", as used in this text, refers to devices for securing equipment, fixtures or other objects to brick
masonry. This Technical Notes discusses the different fastener types used to attach these items to brick masonry.
When other materials, fixtures, etc., are to be attached to brick masonry, the procedure is relatively simple and can
be executed either during or after construction. The designer or builder has a wide variety of fastening methods from
which to choose. The final selection will depend largely upon what is to be attached, when it will be attached and the
type of brick used in the construction.
TYPES OF FASTENERS
Fasteners can be divided into two general categories: those installed during the construction of the masonry, and
those that are installed after the completion of the masonry work.
Fasteners Installed During Construction
Nailing Blocks and Wall Plugs. Wooden nailing blocks and metal wall plugs are placed in mortar joints as the brick
are laid (see Figure 1). Wooden nailing blocks are not used today as frequently as they were in the past, but do
provide an acceptable means of attachment to brick masonry walls. If wooden blocks are used, they should be of
seasoned soft wood to prevent shrinkage and treated to inhibit deterioration. Wooden blocks should be placed only
in head joints.
http://www.gobrick.com/BIA/technotes/t44a.htm
1 of 10 9/13/2009 2:15 PM
Fasteners Installed During Construction
FIG. 1
Metal wall plugs are made of galvanized metal, and may contain wooden or fiber inserts. Metal plugs are preferred
over wooden blocks since problems with shrinkage and decay are not associated with metal plugs. Metal plugs may
be placed in either head joints or bed joints of masonry.
The primary consideration when using fasteners installed during construction is location. Their exact location is not a
serious problem when used to attach moldings, such as baseboards, chair rails, etc., but it may be difficult to
predetermine fastener locations for fixtures, cabinets, shelving, etc. For this reason, post-construction fasteners
have virtually replaced wooden blocks and metal nailing plugs for fastening to masonry.
Post-Construction Fasteners
Screw Shields and Plugs. Screw shields and plugs are produced in plastic, fiber, rubber, nylon and lead (see Fig.
2). Some are advertised by their manufacturers as vibration-resistant, chemical-resistant or water-resistant. These
fasteners are generally used for lightweight attachments and are typically installed in mortar joints or may be placed
directly into solid masonry units (see Fig. 3).
http://www.gobrick.com/BIA/technotes/t44a.htm
2 of 10 9/13/2009 2:15 PM
Shields and Plugs
FIG. 2
Shields and Plugs Installed
FIG. 3
http://www.gobrick.com/BIA/technotes/t44a.htm
3 of 10 9/13/2009 2:15 PM
Bolts and Screws. Several types of bolts and screws are available for use in both solid and hollow masonry. These
fasteners are generally used to attach medium to heavy-weight fixtures. Toggle bolts (made of steel or plastic),
hollow wall screws, small diameter sleeve anchors and screws are used to attach fixtures to walls constructed of
hollow units (see Fig. 4). These fasteners may be placed in holes drilled through bed joints or through the unit faces
into hollow cells (see Fig. 5). Small diameter sleeve anchors, wedge anchors, screws and lag bolt shields (see Fig.
6) are used to attach fixtures to solid masonry and are usually installed in mortar joints (see Fig. 7).
Fasteners for Hollow Masonry Units
FIG. 4
http://www.gobrick.com/BIA/technotes/t44a.htm
4 of 10 9/13/2009 2:15 PM
Fasteners Installed in Hollow Units
FIG. 5
Fasteners for Solid Masonry Units
FIG. 6
http://www.gobrick.com/BIA/technotes/t44a.htm
5 of 10 9/13/2009 2:15 PM
Fasteners Installed in Solid Masonry Units
FIG. 7
Nails. Case-hardened cut and spiral nails (masonry nails) are often used to attach furring strips to masonry walls
(see Fig. 8). If used, the nails should be hammered directly into the mortar joints and not into the brick units. Caution
should be exercised when nails are used in single-wythe walls with exposed exterior faces. The nails could open
small cracks in the mortar joints, allowing water to penetrate the wall (see Technical Notes 7F for problems
associated with water penetration).
http://www.gobrick.com/BIA/technotes/t44a.htm
6 of 10 9/13/2009 2:15 PM
Masonry Nails
FIG. 8
Powder-Driven Fasteners. Powder-driven fasteners are hardened steel pins that are driven into masonry by means
of a powder-actuated tool (Fig. 9). The power for the tool is provided by a powder charge typically ranging from .22
to .38 caliber with varying charges, depending on the material and required pin penetration. Powder-driven fasteners
are generally used on commercial or industrial projects where large volumes of fasteners are required. Several pin
styles and lengths are produced for different fastening requirements (see Fig. 10).
Powder-Driven Fastening Tool
FIG.9
Powder-Driven Pins
FIG. 10
http://www.gobrick.com/BIA/technotes/t44a.htm
7 of 10 9/13/2009 2:15 PM
Powder-driven fasteners require special installation equipment, safety equipment and inspection procedures. For this
reason, the manufacturer should be contacted to determine proper equipment and installation specifications.
Adhesives. A multitude of adhesives, such as epoxies, mastics and contact cements, is produced for various
bonding applications. Many of these produce high bond strengths, have short setting times and offer versatility in
bonding different materials. Adhesives may be used to attach furring, electrical boxes, wall paneling, etc. (see Fig.
11). The manufacturer's literature should be referred to when determining the suitability of an adhesive for a
particular application. Some adhesives may not bond properly to masonry, may not have the elasticity required to
accommodate movements of dissimilar materials and may be affected by exposure to weather, chemicals or
temperature extremes.
Adhesive Fastening
FIG. 11
FASTENER SELECTION
The selection of an appropriate fastener can usually be based on four considerations: 1) the type of brick used in the
construction, 2) the weight of the attachment, 3) the environmental exposure (i.e., interior or exterior) and 4)
aesthetics.
Construction and Attachment
The type of brick used in construction will determine the choice of a fastener as either a solid or hollow wall type
fastener; the weight of the fixture will determine the size of the fastener required. The fastener selection chart shown
in Table 1 can be used as a general guide in selecting a fastener type based on the brick type, installation location
and fixture weight.
http://www.gobrick.com/BIA/technotes/t44a.htm
8 of 10 9/13/2009 2:15 PM
Environment
Environmental factors may have a definite impact on the long-term service life of fasteners and should be considered
in their selection. Environmental factors do not, in general, influence the type of fastener selected, but should
influence the choice of fastener based on the material from which the fastener is made. Corrosion is a major
concern, especially when fasteners are exposed to the elements or when fasteners are used in areas where contact
with corrosive agents is likely.
Steel fasteners used for applications under normal exposure conditions should be galvanized (zinc-coated) to resist
corrosion. Lead, copper-coated or brass fasteners also provide adequate corrosion resistance for normal
exposures. In applications where fasteners are subject to severe exposure conditions or exposed to chemicals,
stainless steel fasteners should be used.
Aesthetics
In most applications, the fastener or fasteners installed will be hidden by the attachment (i.e., cabinets, baseboards,
electrical boxes or furring), and the physical appearance of the fastener (usually the head of a screw or bolt) will not
be of importance. However, when fasteners are used to attach privacy partitions, lighting fixtures or rails, the head
of the fastener is usually visible and required to match or accent the finish of the fixture. In these cases, finished
screws or bolts (i.e., chrome or brass-plated, solid brass or painted) can be purchased to match the fixtures. The
manufacturers should be contacted to determine the availability and range of finishes available in their products.
SUMMARY
This Technical Notes is the second in a series on brick masonry anchors, fasteners and ties. It addresses the types
of fasteners available for use in brick masonry construction. Other Technical Notes in this series address brick
masonry anchor bolts and wall ties.
http://www.gobrick.com/BIA/technotes/t44a.htm
9 of 10 9/13/2009 2:15 PM
The products described in this Technical Notes may involve the use of hazardous materials, operations and/or
equipment. This Technical Notes does not purport to address all of the safety practices associated with the use of
these products. It is the responsibility of the user of this Technical Notes to establish appropriate safety and health
practices and determine the applicability of regulatory limitations prior to the use of the products described.
The information and suggestions contained in this Technical Notes are based on available data and experience of
the technical staff of the Brick Institute of America. This information should be recognized as recommendations and
should be used with judgment. Final decisions on the use of the information discussed herein are not within the
purview of the Brick Institute of America, and must rest with the project owner, designer or both.
REFERENCES
1. Technical Notes on Brick Construction 7F, "Moisture Resistance of Brick Masonry - Maintenance", Brick
Institute of America, Reston, Virginia, February 1986.
2. Construction Sealants and Adhesives, 2nd Edition, J. R. Panek and J. P. Cook, John Wiley and Sons,
New York, 1984.
3. Architectural Graphic Standards. 7th Edition, C. G. Ramsey and H. R. Sleeper, John Wiley and Sons, New
York, 1981.

http://www.gobrick.com/BIA/technotes/t44a.htm
10 of 10 9/13/2009 2:15 PM


Tech Notes 44B - Wall Ties for Brick Masonry
[Revised May 2003]
Abstract: The use of metal ties in brick masonry dates back to loadbearing masonry walls in the 1850's. Historically, the size, spacing
and type of ties have been entirely empirical. Over time, ties of various sizes, configurations and adjustability have been developed for
loadbearing masonry, cavity walls and brick veneer construction. These ties are used to connect multiple wythes of masonry, often of
different materials; anchor masonry veneer to backing systems other than masonry; and connect composite masonry walls. This
Technical Notes addresses the selection, specification and installation of wall tie systems for use in brick masonry construction.
Information and recommendations are included which address tie configuration, detailing, specifications, structural performance and
corrosion resistance.
Key Words: anchors, brick, cavity walls, corrosion, design, differential movement, fasteners, grout, masonry, structural masonry,
ties, veneer, walls.

INTRODUCTION
This Technical Notes is the third in a series that addresses anchor bolts, fasteners and wall ties for brick masonry. This Technical
Notes discusses wall ties commonly used in brick construction, their function, selection, specification and installation. The term "wall
tie", as used in this Technical Notes, refers to wire or sheet metal devices used to connect two or more masonry wythes or used to
connect masonry veneers to a structural backing system. The later of these are more properly identified as veneer anchors.

GENERAL
The first use of wall ties in brick masonry construction can be traced to England in the mid-nineteenth century, where wrought iron ties
were used in brick masonry cavity walls. Use of wall ties in the United States grew after testing showed that metal-tied walls were
more resistant to water penetration than were masonry-bonded walls. Bonders, or "headers", used in masonry-bonded walls may
provide direct paths for possible water penetration. Testing also indicated that the compressive strength of metal-tied cavity walls and
solid walls, and the transverse strength of metal-tied solid walls were comparable to those of masonry-bonded walls.
The use of wall ties has continued to increase over the years due to a trend away from massive, multi-wythe masonry walls to
relatively thin masonry cavity walls, double-wythe walls and veneers. The use of backing systems other than masonry, i.e., steel,
concrete and wood, has rendered bonding with masonry headers impossible, leading to the development of a number of different
metal tie systems. Investigation into the performance of masonry-bonded walls in which the bonded wythes are of different materials
indicates frequent shear failures in the headers.
During this period of transition, little progress was made in the area of rational design of wall tie systems. Typically, the sizing and
spacing of wall ties has been based largely on empirical information and the designer's judgment. Questions concerning strength,
stiffness, corrosion and the effects of these on the long-term performance of wall ties, have been posed. Selection of a tie system to
function properly under these conditions is further complicated by the vast number of tie types available and the variety of materials
from which they are fabricated. Most tie systems perform well for their intended application. Some tie systems, however, are poorly
designed and do not provide adequate load transfer for brick masonry. The distinction is often subtle and requires an understanding
of the properties and characteristics of brick masonry. With the addition of Chapter 6 on Veneers to the Building Code Requirements
for Masonry Structures (ACI 530 / ASCE 5 / TMS 402-02), the empirical requirements for type, size and spacing of metal ties have
been reviewed and refined. This Code, known as the MSJC Code, contains requirements for most types of tie systems.
Function of Wall Ties
Typically, wall ties perform three primary functions between a wythe of brick and its backing or another wythe of masonry: 1) provide
a connection, 2) transfer lateral loads, 3) permit in-plane movement to accommodate differential movements and, in some cases,
restrain differential movement. In addition to these primary functions, metal ties (as joint reinforcement) may also be required to serve
as horizontal structural reinforcement or provide longitudinal continuity.
For a tie system to fulfill these functions, it must: 1) be securely attached to both masonry wythes or the brick veneer and its backing,
2) have sufficient stiffness to transfer lateral loads with minimal deformations, 3) have a minimum amount of mechanical play, 4) be
corrosion-resistant and 5) be easily installed to reduce installation errors and damage to the tie system. This listing is far from
complete; special project conditions, unusual details and special building code requirements must also be considered. Availability and
cost are always factors in product specifications. However, cost should not have a major influence on the selection of a wall tie
system since the cost of ties is typically a very small part of the total wall cost.
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
1 of 15 9/13/2009 2:16 PM

TYPES OF WALL TIES
General
There are a number of different wall tie systems available for brick masonry wall systems. These include unit ties, continuous
horizontal joint reinforcement, adjustable ties (unit and continuous) and re-anchoring systems. Placement requirements for ties are
shown in Figure 1.
FIG. 1
Unit Ties
Unit ties are rectangular boxties, "Z" ties and corrugated ties, as shown in Figure 2. Rectangular and "Z" ties are usually fabricated
from cold-drawn steel wire conforming to ASTM A 82. Rectangular and "Z" ties made of stainless steel conforming to ASTM A 580
are also available for use in more corrosive environments. Corrugated sheet steel ties are typically manufactured from steel sheet
conforming to ASTM A 1008 and are also available in stainless steel conforming to A 240.

Unit Ties
FIG. 2

Rectangular and "Z" ties are used to bond walls constructed of two or more masonry wythes. "Z" ties should only be used to bond
walls constructed with solid units (not less than 75% solid) or grouted units. Rectangular ties may be used with either solid or hollow
units. Such wire ties should not have a bend or drip to reduce water transfer. Such a bend in the tie reduces the capacity of the tie to
transfer lateral load.
Corrugated ties are typically used in low-rise, residential veneer over wood frame construction and are not recommended for
construction incorporating brick veneer over steel studs, masonry-backed cavity walls, multi-wythe walls or grouted masonry walls.
Typical installation details are shown in Fig. 3.
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
2 of 15 9/13/2009 2:16 PM
Unit Tie Placement Details
FIG. 3

Joint Reinforcement
Continuous horizontal joint reinforcement is typically made from #8, 9, 10, or 11 gage wire, or 3/16 in. (5 mm) diameter wire,
conforming to ASTM A 951, in lengths of 10 to 12 ft (3 to 4 m). The most common configurations are the ladder, truss, and tab types
(see Fig. 4).
Continuous Joint Reinforcement
FIG. 4

Structural testing performed in the early 1960's indicated that multi-wythe walls tied with joint reinforcement performed as well as
walls tied with unit ties or masonry bonders. Joint reinforcement may be used in multi-wythe solid walls, masonry cavity walls, brick
veneer with masonry backing, and grouted masonry walls (see Fig. 5). As with wire ties, the cross wires should be without drips.
Truss-type joint reinforcement is not recommended for use in cavity walls or brick veneer with masonry backing. Test results also
indicated that truss-type joint reinforcement, in such wall systems, did not contribute to any composite action in the vertical span, but
did develop a degree of composite action in the horizontal span .The configuration of the truss diagonals can restrain differential
movement between wythes and possibly result in bowing of the walls.
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
3 of 15 9/13/2009 2:16 PM
Joint Reinforcement Details
FIG. 5

Adjustable Ties
Adjustable tie systems were initially developed to accommodate the use of face brick whose bed joints did not align vertically with
interior masonry wythes. This concept has been extended to ties used to attach brick to other systems, resulting in the use of both
adjustable unit ties and adjustable ties with joint reinforcement.
The use of adjustable ties has increased rapidly for a number of reasons: 1) Adjustable ties permit the construction of interior
masonry wythes and other backings prior to the construction of exterior facing wythes, permitting the structure to be enclosed faster.
2) Adjustable ties are two-piece systems. One piece is installed as the backing is constructed and the other piece is installed as the
facing wythe is constructed, reducing the risk of damage to exposed ties that might occur when unit ties or standard joint
reinforcement are used. 3) Adjustable ties can accommodate construction tolerances common in multi-material construction. 4)
Adjustable ties can accommodate larger differential movements than standard unit ties or joint reinforcement.
The advantages offered by adjustable tie systems are not without possible problems: 1) Mislocation of adjustable ties placed prior to
construction of facing wythes, if extreme, can render the ties useless. 2) Adjustable ties may encourage less than perfect layout of
the wall system since a built-in adjustment allowance is available. 3) Large variations in construction tolerances may not allow full
engagement of ties installed before facing wythes are constructed. 4) Improperly positioned ties may result in large vertical tie
eccentricity. 5) The structural performance of some adjustable ties in regard to strength and stiffness is less than that of standard unit
ties or joint reinforcement.
Adjustable Unit Ties. Adjustable unit ties produced for use with masonry backing, concrete backing, steel frames and steel studs
are shown in Figs. 6 and 7. Slot-type ties (dovetail, channel slot, etc.) have been used for a number of years with concrete, steel
frame and steel stud backing systems, and are recognized as tie systems capable of accommodating differential movement, as
further discussed in Technical Notes 18 Series (see Fig. 6). Other types of adjustable unit tie systems are available for brick with
masonry backing and other backing systems. These ties are typically two-piece systems, consisting of a single or double eye and
pintle arrangement (see Fig. 7). The Building Code Requirements for Masonry Structures (ACI 530 / ASCE 5 / TMS 402-02)
requires that for veneer masonry all pintle anchors have at least two pintle legs of wire size W2.8 (3/16 in., MW18) each and have an
offset not exceeding 1 in. (31.8 mm). Typical installation details are shown in Fig. 8

TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
4 of 15 9/13/2009 2:16 PM
.
Adjustable Unit Ties for Steel, Concrete and Stud Backup
FIG. 6
Adjustable Unit Ties for Masonry Backup
FIG. 7

TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
5 of 15 9/13/2009 2:16 PM

Adjustable Unit Tie Details
FIG. 8

Adjustable Assemblies. Adjustable ladder and truss-type joint reinforcement assemblies are available for use in masonry backed-
cavity wall, veneer and grouted wall construction. This joint reinforcement typically consists of rectangular tab type extensions,
connected to standard joint reinforcement by means of an eye and pintle arrangement (see Fig. 9). Installation details are shown in
Fig. 10.
Adjustable Joint Reinforcement Assemblies
FIG. 9


TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
6 of 15 9/13/2009 2:16 PM
Adjustable Assembly Details
FIG. 10
Masonry Re-anchoring Systems.

Masonry re-anchoring systems are the most recent development in masonry tie systems. Three general types of systems are being
produced and typically consist of a mechanical expansion system, helical screw system and an epoxy adhesive system (see Fig.
11).These systems are primarily used to: provide ties in areas where ties were not installed during original construction, 2) replace
existing ties, 3) replace failed masonry bonding units, 4) upgrade older wall systems to current code levels, or 5) attach new veneers
over existing facades.

Masonry Re-Anchoring Systems
FIG. 11

As stated, re-anchoring systems are relatively new and many designers and contractors may not be fully familiar with their installation
or limitations. For this reason, consultation with the tie system manufacturer is essential to assure proper application, detailing,
installation, inspection, and performance.

TIE SELECTION
Strength and Deformation
The strength and deformation characteristics of tie systems are not generally analyzed nor investigated during the project design or
specification phase. Building codes and standards have typically required minimum tie size (diameter or gage) and maximum tie
spacing limits to control tie loading and deformation. Present tie size and spacing requirements have been derived from some testing
and from the past performance of traditional tie systems (rectangular ties, "Z'' ties and standard joint reinforcement). The growing use
of adjustable tie systems has caused some concerns in regard to tie strength and deformation. Most adjustable ties permit vertical
adjustment up to approximately one-half the height of a standard brick unit, some permit greater adjustments. Depending on the tie
configuration, the deflections of adjustable ties can become quite large as vertical adjustment eccentricities are increased. This
deflection is further increased if mechanical play is present in the tie system.
Analytical and experimental investigations of cavity wall and veneer wall systems have shown that tie loads and deformations are a
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
7 of 15 9/13/2009 2:16 PM
function of: 1) the relative stiffness between facing and backing materials, 2) tie spacing, 3) tie stiffness, 4) support conditions of the
facing and backing systems, 5) location of edges and openings, 6) cavity width and 7) applied loads.
Estimating tie loads based on tributary area can lead to large errors, depending on the geometry and properties of the wall system.
Fig. 12 shows tie loads and deflections calculated from a simplified model of a cavity wall system. As shown, adjustable tie
deflections become large as the adjustment eccentricity becomes large. These values were calculated assuming that no mechanical
play existed in the tie system. Mechanical play must be added to these values to determine the total deflection of the exterior wythe.
Typical adjustable ties have values of mechanical play ranging from approximately 0 to 0.3 in. (0 to 8 mm). Some adjustable ties may
have an extreme amount of mechanical play when not properly installed (see Fig. 13). The MSJC Code limits mechanical play to a
maximum of 1/16 inch (1.6 mm). If satisfactory performance is to be expected, total tie and backing deflections must be maintained
within the working range of the masonry facade under full design loads.

Calculated Tie Loads and Deflections
FIG. 12
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
8 of 15 9/13/2009 2:16 PM
Mechanical Play
FIG. 13

Recommendations. At present, analysis techniques that accurately model metal-tied wall systems are still in the developmental
stage and require further refinement and verification through testing. Until more accurate methods are available, the Brick Industry
Association feels that acceptable strength and deformation characteristics can be achieved by one or more of the following
measures: 1) Reduce or eliminate lateral mechanical play in adjustable tie systems. Limit the total mechanical play to 1/16 in. (1.6
mm), see Fig. 13. 2) Reduce or eliminate adjustment eccentricity in adjustable tie systems. This can be accomplished by installing ties
as facing wythes are constructed or by using starter courses or ledges when facing wythes are constructed over masonry backing. 3)
Eliminate possible disengagement of adjustable ties by providing positive vertical movement limitations. 4) Provide additional ties
within 8 in. (200 mm) of openings and discontinuities, i.e., windows, shelf angles, vertical expansion joints, etc. 5) Do not specify ties
with formed drips. Testing has shown that drips can reduce the ultimate buckling load by approximately 50 percent. 6) Space ties as
shown in Table 1, based on the tie system and wall system. 7) Specify stiff ties. This can be accomplished by specifying ties with
maximum deflections of less than 0.05 in. (1.2 mm) when tested at an axial load of 100 lb in tension or compression. When adjustable
ties are specified, the deflection limit should be satisfied at the eccentricity expected in the field. See Table 2 for minimum tie gage
and diameter recommendations. 8) Select an appropriate tie system for the wall system (see Table 3). Many of these
recommendations have been incorporated in the MSJC Code.

TABLE 1
Tie Spacing Requirements
1,2
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
9 of 15 9/13/2009 2:16 PM
1.
Masonry laid in running bond. Consult applicable building code for special bond patters such as stack bond.
2.
Based on the requirements in the 2002 MSJC Code.
3.
Maximum allowable distance between inside face of veneer and framing material, per MSJC Code, unless noted otherwise.

TABLE 2
Required Tie Sizes
1

TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
10 of 15 9/13/2009 2:16 PM
1
Based on the requirements in the 2002 MSJC Code.
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
11 of 15 9/13/2009 2:16 PM

1See Table 1 for spacing; Table 2 for sizes and gages; Table 5 for corrosion protection.

Corrosion
General. Awareness of possible corrosion problems in metal-tied masonry walls has increased due to corrosion damage found on
reinforcement in concrete highway pavements, bridge decks and some masonry structures. The potential for corrosion problems in
masonry has increased as construction and design philosophies have changed and as environmental conditions have changed over the
last decades. These changes include use of thinner masonry walls and masonry veneers that are more susceptible to water
penetration, increases in atmospheric pollutants, use of accelerators containing calcium chloride, increased use of insulated cavities
(resulting in the relocation of the dew point within the wall section) and combinations of different metals in brick veneer wall systems.
This list is not all-inclusive; corrosion potential can also be affected by the function of a structure, geographic location, compatibility of
construction materials, detailing and workmanship.
Corrosion Protection. In order to provide corrosion protection, environmental factors must be controlled or metals used in
construction must be protected. Conventional corrosion protection methods attempt to protect metals embedded in masonry by
isolating them with impervious coatings (barrier protection), by using metals that are corrosion-resistant, or by providing cathodic
protection in which one metal becomes sacrificial to protect another.
Galvanizing Galvanizing (zinc-coating) provides resistance to corrosion by two methods. First, the zinc coating acts as a barrier
shielding the underlying steel from corrosive action. Second, it acts as a sacrificial element that is consumed before the base steel is
attacked. This sacrificial nature protects the base steel at scratches and discontinuities in the zinc coating caused by fabrication,
handling or installation, until most of the adjacent zinc coating is consumed. Studies have shown that the protective value of zinc
coating is proportional to its thickness. Thus, for longer periods of protection, a thicker zinc coating is required. Also, when the
protective zinc coating is depleted, the corrosion of the base steel will progress as if no galvanizing were present.
Two methods of galvanizing are used to protect metal masonry ties: mill galvanizing and hot-dip galvanizing. Mill galvanizing takes
place after steel wire or sheets have been processed to their specified dimensions and prior to fabrication of the tie. During the mill
galvanizing process, zinc can be applied in a variety of thicknesses, as shown in Table 4. Hot-dip galvanizing is performed by dipping
completely fabricated assemblies into molten zinc until a specified amount of zinc is bonded to the base metal. Hot-dip galvanized
coatings are typically thicker than mill galvanized coatings and therefore, provide longer periods of protection.
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
12 of 15 9/13/2009 2:16 PM

TABLE 4
Coating Requirements
1
Class B Rolled, pressed or forged articles.
B-1: 3/16 in. (4.8 mm) and over in thickness and over 15 in. (381 mm) in length
B-2: Under 3/16 in. (4.8 mm) in thickness and over 15 in. (381 mm) in length.
B-3: Any thickness and 15 in. (381 mm) and under in length.

Stainless Steel - Stainless steel ties are often specified for use in very corrosive environments. Stainless steel ties are
specified under ASTM A 240 or A 580 and are generally made from one of the austenitic stainless steels. Stainless steel
resists corrosion well; however, if in contact with carbon steel, a galvanic cell can result and actually increase the potential for
corrosion. For this reason, combining stainless steel ties or screws with carbon steel or galvanized steel components is not
recommended.
Fusion-Bonded Epoxy Epoxy coating is the newest process used to provide corrosion protection for metal ties. The process
has been adapted from epoxy-coated reinforcement bars used successfully in concrete systems with severe environmental
exposures. Epoxy coating provides protection by acting as an impervious barrier. The epoxy coating is bonded to the base steel by a
heat-induced chemical reaction through which a chemical and mechanical bond is formed. The combination of the two types of
adhesion helps to prevent cracking of the coating due to handling, installation or stress reversals. The epoxy coating is not sacrificial
like zinc; therefore, nicks and voids in the coating can lead to corrosion of the base steel. Epoxy coatings for joint reinforcement
should meet the requirements of ASTM A 884, Class A, Type 1 - 7 mils. Epoxy coatings for wire ties and anchors are specified in
ASTM A 899, Class C - 20 mils. Sheet metal ties and anchors should be coated with 20 mils of epoxy per surface or per
manufacturer's specification.
Recommendations. The past performance of metal ties in regard to corrosion has generally been satisfactory. The American
Galvanizers Association has developed a Zinc Coating Life Predictor program that provides an estimate of service life for zinc coating
in an exposed environment. This does not specifically address performance of ties in masonry walls. Until research can produce
accurate methods of assessing corrosion potential and predicting adequate levels of protection, the Brick Industry Association
suggests minimum levels of corrosion protection for metal ties and hardware as indicated in Table 5. As with all other engineering
considerations minimum recommendations may not be adequate in every situation, and should not serve as substitutes for engineering
investigation or judgment. Decisions must be based on individual project conditions, performance requirements and safety.

TABLE 5
Recommended Minimum Corrosion Protection
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
13 of 15 9/13/2009 2:16 PM
SUMMARY
This Technical Notes is the third in a series addressing brick masonry anchor bolts, fasteners and wall ties. It is primarily concerned
with the types of wall ties commonly used in brick masonry construction, their function, selection, specification and installation. Other
Technical Notes in this series individually address anchor bolts and fasteners for brick masonry.
The information and suggestions contained in this Technical Notes are based on the available data and the experience of the
engineering staff of the Brick Industry Association. The information contained herein must be used in conjunction with good technical
judgment and a basic understanding of the properties of brick masonry. Final decisions on the use of the information contained in this
Technical Notes are not within the purview of the Brick Industry Association and must rest with the project architect, engineer and
owner.

REFERENCES
1. DeVekey, R.C., ''Corrosion of Steel Wall Ties: Recognition, Assessment and Appropriate Action, Building Research
Establishment Information Paper, IP 28/79, Building Research Establishment, Garston, Watford, England, October 1979.
2. Fishburn, C.C., ''Water Permeability of Walls Built of Masonry Units, Report BMS 82, National Bureau of Standards, Department
of Commerce, Washington, D.C., April 1942.
3. Allen, M. H., Research Report No. 10, Compressive and Transverse Strength Tests of Eight-Inch Brick Walls, Structural Clay
Products Research Foundation, Geneva, Illinois, October 1966.
4. Allen, M.H., Research Report No. 14, Compressive Strength of Eight-Inch Brick Walls with Different Percentages of Steel Ties
and Masonry Headers, Structural Clay Products Research Foundation, Geneva, Illinois, May 1969.
5. Bortz, S.A., Investigation of Continuous Metal Ties as a Replacement for Brick Ties in Masonry Walls, Summary Report ARF
6620, Armour Research Foundation, Chicago, Illinois, June 1960.
6. Investigation of Masonry Wall Ties, ARF Project B870-2 (Revised), Armour Research Foundation, Chicago, Illinois, December
1962.
7. ''Flexural Strength of Cavity Walls, ARF Project B870, Armour Research Foundation, Chicago, Illinois, March 1963.
8. Brown, R.H. and Elling, R.E., ''Lateral Load Distribution in Cavity Walls, Proceedings of the Fifth International Brick Masonry
Conference, Washington, D.C., October 1979.
9. Bell, G.R. and Gumpertz, W.H., ''Engineering Evaluation of Brick Veneer/Steel Stud Walls, Part 2 Structural Design, Structural
Behavior and Durability, Proceedings of the Third North American Masonry Conference, Arlington, Texas, June 1985.
10. Arumala, J.O. and Brown, R.H., ''Performance of Brick Veneer With Steel Stud Backup, Clemson University, Clemson, South
Carolina, April 1982.
11. ''Development of Adjustable Wall Ties, ARF Project No. B869, Armour Research Foundation, Chicago, Illinois, March 1963.
12. Catani, Mario J., ''Protection of Embedded Steel in Masonry, The Construction Specifier, January 1985.
13. Catani, Mario J. and Whitlock, A. Rhett, ''Coping With Wide Cavities, The Construction Specifier, August 1986.
14. Zhang, X. Gregory, Zinc Coating Life Predictor, The International Lead Zinc Research Organization, published online at
http://zclp.galvanizeit.org:8180/zclp/index.html, 2002.
TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
14 of 15 9/13/2009 2:16 PM



TN 44B - WALL TIES FOR BRICK MASONRY http://www.gobrick.com/BIA/technotes/t44b.htm
15 of 15 9/13/2009 2:16 PM

Technical Note 45 - Brick Masonry Noise Barrier Walls - Introduction
[Feb. 1991] (Reissued July 2001)
Abstract: Because our national highway system has grown significantly over the last few
decades, public awareness of traffic noise on neighborhood communities has increased.
Neighborhood associations and governmental bodies look for ways to reduce traffic noise without
adversely affecting the surrounding environment. A solution to this problem lies in brick masonry
noise barrier walls. Brick masonry noise barrier walls can easily blend into the environment and
give residential communities protection from unwanted highway noise.
Key Words: acoustics, brick, noise barrier walls.
INTRODUCTION
Continued growth of our national highway system combined with an increase in public awareness
of environmental issues has focused on a need to evaluate the impact of traffic noise associated
with highway systems on neighboring communities. When noise levels exceed acceptable limits,
community action generally alerts governmental bodies to the problem or potential problems.
Governmental bodies then investigate measures to prevent or alleviate noise problems.
The severity of the noise and the stage at which the problem is identified determine the measures
available to reduce the impact of highway noise. Measures to alleviate highway noise include
traffic controls and regulations, modification of the highway configuration, land-use planning and
zoning, and brick noise barrier walls.
When new highway systems are in the planning and design stages, a comprehensive analysis of
and consideration to noise abatement measures can be given. However, when existing highway
systems are renovated or if restrictions are placed on the routing of new highway systems or use
of adjacent land, the most practical solution to noise control may be the use of noise barrier walls
to isolate the highway noise sources from the surrounding communities.
Three major types of noise barriers are currently being used in the United States: earth berms,
walls and berm-wall combinations. Of these three, the noise barrier wall is typically the most
common means of achieving noise abatement and is the primary topic of this Technical Notes.
This Technical Notes, the first in a series, addresses acoustical, visual, structural, construction,
detailing and maintenance considerations of brick masonry noise barrier walls. The other
Technical Notes in this series addresses the structural design of brick masonry noise barrier walls.
ACOUSTICAL CONSIDERATIONS
To understand the function of a noise barrier wall or how the wall reduces the noise level perceived
by a receiver, it is necessary to discuss some of the fundamental principles involved in sound
propagation and noise reduction.
When there are no obstacles or barriers between highway noise sources and receivers, sound
travels in a direct path from the source to the receiver (Figure 1). When a noise barrier wall is
placed between the noise source and the receiver, the barrier disperses the sound along three
paths: a diffracted or bent path over the top of the wall, a reflected path away from the receiver
and a transmitted path through the wall (Fig. 2).
t45 http://www.gobrick.com/BIA/technotes/t45.htm
1 of 18 9/13/2009 2:17 PM

Direct Noise Path
FIG. 1

Noise Path With Barrier Wall
FIG. 2
Diffraction of sound over the top of the wall produces a shadow zone behind the barrier. The
boundary of this shadow zone is outlined by a straight line drawn from the noise source over the
top of the barrier wall (Fig. 3). All receivers located within the shadow zone will experience some
degree of sound attenuation. The amount of reduction or attenuation is directly related to the
diffraction angle -. As this angle increases, the barrier attenuation increases. Thus, barrier
attenuation is a function of the wall height and the distances between the source, barrier and
receiver. Two other factors also affect the amount of attenuation: the sound transmission
characteristics of the material from which the barrier is constructed and the length of the barrier.

Noise Barrier Shadow Zone
FIG. 3
The sound transmission characteristics of a material are related to its weight, stiffness and loss
factors. The sound transmission characteristics of materials can be assessed and compared by
means of transmission loss values. The sound transmission loss is related to the ratio of the
incident noise energy to the noise energy transmitted through the material. Typically, transmission
loss values can be expected to increase with increasing square foot surface weights of barrier
materials. Table 1 lists the transmission loss values at a frequency of 550 hertz (Hz) for materials
commonly used in noise barrier wall construction. 550 Hz is the accepted frequency used to
determine the transmission loss of highway noise barrier wall materials. As a general rule for
design, the transmission loss value should be a minimum of 10 decibels (dB) above the attenuation
resulting from the diffraction over the top of the barrier. The transmission loss values for brick
masonry are at the higher end of the range and sound transmission through a brick barrier will not
significantly affect the attenuation. However, when less massive materials are used, the
transmission loss values may not be adequate and the noise reduction provided by the barrier can
t45 http://www.gobrick.com/BIA/technotes/t45.htm
2 of 18 9/13/2009 2:17 PM
be severely affected.
1
Grouted cavity is 2 3/4 in.
The actual acoustical design of a barrier system to determine the length and height requirements
are beyond the scope of this Technical Notes. A detailed discussion of noise barrier acoustical
design procedures and considerations can be found in Reference 1.
VISUAL CONSIDERATIONS
General
Highway noise barriers tend to dominate the visual environment adjacent to roadways (Fig. 4).
They are often thousands of feet long and can be as high as 25 ft (7.6m) above the road surface.
When noise barrier walls higher than 16 ft (4.9m) are acoustically required, visual consideration of
surrounding features should be evaluated. Exceptionally high walls can have an unsightly impact
on the aesthetic features of the territory and can give the driver a claustrophobic feeling. For
safety reasons, the designer should reduce the visual impact of the noise barrier wall. The
motorist must pass the barrier with as little visual disruption as possible. The primary attention of
the driver should be on the road ahead and adjacent traffic conditions. This can be achieved by
doing one of several things.

Brick Noise Barrier Wall
FIG. 4
For relatively low walls, the line of the noise barrier should reflect similar lines of the surrounding
environment. For instance, in rolling terrain, a straight line will be out of place and attention will be
drawn to that line. However, in a flat terrain where the horizon is visible as a straight line, a
straight line in a noise barrier wall may not appear to be visually dominant. The introduction of
vertical lines, such as with pilasters, placed along relatively low walls is recommended to achieve
visual balance. Plantings such as columnar trees can emphasize vertical lines in a noise barrier
wall. Further, shrubbery can be used to soften the transition between ground and wall
intersection. Wherever possible, the wall should step back to open up the view for the motorist
(Fig. 5). However, this can only be practically achieved in rolling or hilly terrain. In an urban
environment where the horizon is composed of alternating heights of buildings, an appropriate wall
may vary in height as a reflection of the city's profile.
t45 http://www.gobrick.com/BIA/technotes/t45.htm
3 of 18 9/13/2009 2:17 PM

Effect of Wall Placement on Sight Lines
FIG. 5
Another way to reduce the visual impact on the environment is through changes in height and
location of the wall. A wall with offsets can break the monotony of a straight wall and create
pockets which can be used for plantings (Fig. 6). These transitions may further be used as areas
for change in texture, color or wall height. A serpentine wall can create the same visual interest as
a wall with offsets (Fig. 7). Moreover, due to their geometry, both of these walls have the added
advantage of being more resistant to seismic and wind forces than their straight counterparts.
t45 http://www.gobrick.com/BIA/technotes/t45.htm
4 of 18 9/13/2009 2:17 PM

Offset Wall
FIG. 6a

Offset Wall
FIG. 6b
t45 http://www.gobrick.com/BIA/technotes/t45.htm
5 of 18 9/13/2009 2:17 PM

Serpentine Wall
FIG. 7a

Serpentine Wall
FIG. 7b
Regardless of the shape, noise barrier walls should not begin or end abruptly. The best transition
of beginning and end is to tie the wall into a natural hillside or a man-made earth berm. If no
natural hills or berms are available, the wall termination should taper down and angle away from
t45 http://www.gobrick.com/BIA/technotes/t45.htm
6 of 18 9/13/2009 2:17 PM
the roadway. Not only is this visually pleasing, it is also functional. This transition can effectively
reduce the amount of noise traveling around the end of the wall as a result of approaching traffic.
Access through noise barrier walls may be needed in certain instances. Maintenance personnel
may require doors for equipment or service. Firefighters may require access to hydrants or water
sources on the opposite side of the barrier wall. The appropriate highway and emergency
agencies should be consulted regarding access locations and requirements. Openings through
noise barrier walls must not reduce the acoustic or structural performance of the noise barrier.
Larger openings are best located at offsets in the wall, or with piers or pilasters at the jamb of the
opening. This geometry provides an easier means of accommodating loads and reducing sound
penetration. Openings in straight wall sections change the load distribution and this influence must
be considered. Loose steel lintels or reinforced masonry beams should be used to span over the
openings.
Texture
A change of texture on noise barriers helps to create a pleasant variety for motorist, adjacent
residents and pedestrians. The requirements of each are different, however, and must be treated
separately. Since motorists usually drive at high rates of speed, they have little opportunity to
examine details. To be effective, textures along the highway need to be bold or coarse and visible
at a glance because the motorists' attention should not be diverted from the highway. However,
textures on the opposite of the highway should be more detailed. The residents and pedestrians
on this side view the barrier at much slower speeds and at closer distances. Bold textures can be
overbearing and monotonous to them and, therefore, should not be used.
Unlike other materials, masonry can be adapted to create the bold textures for the motorist and
the subtle, more detailed textures for those on the other side. Because of its versatility, the
possibilities for brick masonry are almost limitless. Bold textures can be created by offsetting
brick in random patterns which can cast varying textural shadows during the day. The use of
pilasters, special shape brick and copings can also create bold textural interest. Further, brick
sculpture can create detailed textures for residents (Fig. 8). Brick can be carved to portray a
desired logo, mural or composition.

Brick Sculpture
FIG. 8
Color
Color
The color of the wall plays an important role in blending the wall into the surrounding environment.
Since brick barrier walls are man-made structures placed in a natural environment, their color
should not attempt to match the color of trees, grass, or shrubbery because they are not related
to such natural features by form. Earthen colors, such as browns, grays, and rusts of varying
t45 http://www.gobrick.com/BIA/technotes/t45.htm
7 of 18 9/13/2009 2:17 PM
tones, when used on barrier walls help to blend the structures into their environment. Repetitious
polychromatic patterns are not recommended on the highway side of the barrier. These types of
patterns draw the motorists' attention away from the road ahead. However, they can be used on
the side of the wall opposite of the highway. Moreover, placing units of different color in
alternating bonding patterns can also easily create visual interest. Further, color interest and
variety may be achieved through the use of plants and trees. Foliage which changes color will
impart a pleasing seasonal variation.
STRUCTURAL CONSIDERATIONS
Structurally, brick masonry noise barrier walls can be designed in various ways. The most popular
designs though are the pier and panel, pilaster and panel, and the cantilever walls.
Pier and Panel Wall
The pier and panel wall is composed of a series of single-wythe panels, usually four inches in
thickness. These panels are braced periodically by piers (Fig. 9). This type of wall is relatively
easy to build and is economical due to the efficient use of materials. It is easily adapted to varying
terrain and is acoustically adequate for a highway noise barrier. The pier and panel wall can also
be built with returns of varying angles. However, the most easily constructed and economical
return is one which is perpendicular to an adjacent panel. The panels, usually built from 8 to 20 ft
long (2.4 to 6.1 m), are placed between piers of reinforced masonry, concrete, or steel. The
panels can either be prefabricated or built in place and can be as high as acoustically or
aesthetically necessary. However, any space left between the bottom of the wall and the ground
must be adequately backfilled to prevent noise penetration underneath the wall.
.
t45 http://www.gobrick.com/BIA/technotes/t45.htm
8 of 18 9/13/2009 2:17 PM

Pier and Panel Assembly
FIG. 9
The panels, supported on piles or clip angles attached to piers, essentially act as thin, simply
supported beams. The panel, which spans horizontally between the piers, will develop flexural
tensile stresses parallel to the bed joints due to out-of-plane wind and seismic loads (Fig. 10).
Horizontal joint reinforcement is required if the calculated flexural stresses exceed the allowable
stresses found in the local building code. If horizontal reinforcement is required, it must be
distributed the full height of the panel.
t45 http://www.gobrick.com/BIA/technotes/t45.htm
9 of 18 9/13/2009 2:17 PM

Out-of-Plane Deflection of Panel in Pier and Panel Wall
FIG. 10

The panel also develops in-plane flexural stresses due to its own dead weight and any incidental
vertical loading which may occur (Fig. 11). The in-plane bending will cause flexural tensile stresses
at the bottom of the panel. Although the building codes do not now define allowable flexural tensile
values for in-plane bending, the allowable flexural tensile stresses parallel to the bed joint for
out-of-plane bending can conservatively be used. Both the in-plane and out-of-plane flexural
tensile stresses must be calculated and added because the bottom of the panel is subjected to
both maximum in-plane and out-of-plane moments. If the sum of the calculated stresses exceeds
the out-of-plane allowable flexural tensile stress parallel to the bed joint the panel must be
reinforced. This reinforcement is usually placed in the bottom two or three courses of masonry.
In-of-Plane Deflection of Panel in Pier and Panel Wall
FIG. 11
The piers, on the other hand, act as vertical cantilevers and must be designed to resist all lateral
loads transferred from the panels. The piers are usually anchored to or embedded in reinforced
concrete piles, which vary in depth due to local soil conditions. The piles must be designed to
resist all shear and axial loads and the overturning moment caused by the panel due to
t45 http://www.gobrick.com/BIA/technotes/t45.htm
10 of 18 9/13/2009 2:17 PM
out-of-plane wind and seismic forces (Fig. 12). Vertical reinforcement may be required in a panel
if the out-of-plane deflection of the pier exceeds the maximum allowable deflection of the panel.
This maximum allowable deflection for an unreinforced panel is based on the allowable flexural
tensile stress perpendicular to the bed joints. If vertical reinforcement is required, then hollow
brick units can be used to facilitate the reinforcement and grouting process. However, it is
recommended that the piers be stiff enough so vertical reinforcement in the panels is not
necessary.
Out-of-Plane Deflection of Pier and Panel
FIG. 12
Due to the deflection requirements of the panel, the web length of the pier may be larger than the
width of the panel, especially for piers made of steel. The space between the pier and panel must
be filled with a non-compressible material, placed either uniformly or intermittently along the height
of the pier. This non-compressible material ensures proper load transfer from the panel to the
pier. However, if intermittent supports are used, a filler material must be placed between supports
to block noise transfer around the end of the panel. Further, a clear space the entire height of the
panel must be maintained between the end of the panel and the web of the pier. This space
allows for the in-plane expansion and contraction of the brick panel (Fig. 13).
Steel Pier/Panel Detail
FIG. 13
When reinforced concrete or masonry piers are used, the flanges should be analyzed to ensure
that the shear and bending forces imposed on them by the adjacent panel do not exceed allowable
stresses. If for aesthetic reasons, an exposed steel pier is not desirable, brick can be built around
the steel in the form of a pilaster (Figs. 13 through 15). Corrosion protection of the pier should be
considered when steel piers are used.
Finally, the panels can bear directly on the pile or a steel clip angle which is attached to the pier.
The bearing stress requirements of each material must be considered in the design.
t45 http://www.gobrick.com/BIA/technotes/t45.htm
11 of 18 9/13/2009 2:17 PM
Steel Pier and Panel Wall
FIG. 14
Steel Pier and Panel Wall with Brick Surround
FIG. 15
Pilaster and Panel Walls
The pilaster and panel and the pier and panel wall appear to be very similar. Both are composed
of single-wythe panels periodically braced by vertical elements and both are equally adaptable to
varying terrain and returns. However, there are some fundamental differences which must be
carefully analyzed. First, unlike the pier and panel wall, the panel in the pilaster and panel wall is
integrally bonded to the pilaster at most intersections (Fig. 16). This seemingly innocuous
difference actually has a marked effect on the structural characteristics of the wall. The end
condition of the panel in a pier and panel wall is considered simply supported while that in a
bonded pilaster and panel is considered fixed. Because of the fixed-end condition, the designer
must satisfy the negative moments which are generated at the pilaster (Fig. 17). Depending on
the geometry of the wall, horizontal reinforcing steel may be required in both the top and bottom
courses of brick due to vertical in-plane bending. If required, it must be fully developed and
adequately anchored in the pilaster. The horizontal out-of-plane deflection of the panel will also
generate negative moments at the pilaster (Fig. 18). Any horizontal reinforcement will help resist
negative moments due to out-of-plane bending. However, the reinforcement must also be fully
developed in the pilaster. The pilaster should be stiff enough so the allowable flexural tension
developed in the panels due to the out-of-plane deflection of the pilaster is not exceeded. The
pilaster must be rigidly attached to the pile below, and the pile must be designed to resist all shear
and axial loads and overturning moments.

t45 http://www.gobrick.com/BIA/technotes/t45.htm
12 of 18 9/13/2009 2:17 PM
Pilaster and Panel Assembly
FIG. 16

In-Plane Deflection of Panel in a Pilaster and Panel Wall
FIG. 17


Out-of-Plane Deflection of Panel in a Pilaster and Panel Wall
FIG. 18
t45 http://www.gobrick.com/BIA/technotes/t45.htm
13 of 18 9/13/2009 2:17 PM

Another difference between the pier and panel and pilaster and panel wall is the placement of
expansion joints. Since the pier and panel are not bonded together, the in-plane horizontal
movement can be accounted for at the end of each panel. However, this is not the case with the
pilaster and panel wall because they are integrally bonded together. A vertical break provided by
an expansion joint is necessary to permit horizontal expansion. The best location for an expansion
joint is at the pilaster and panel intersection. The expansion joints should not be placed more than
a maximum of 30 ft (9.1m) on center, and the pilaster must not restrict horizontal in-plane
movement due to expansion. Further, the connection between pilaster and panel must be able to
resist the out-of-plane loads imposed on it.
Finally, because the pilaster and the panel are bonded together, the pilaster and panel wall must
be built in place. Forms or centering must support the panel during construction and can only be
removed after the wall is adequately cured. However, a continuous footing running between the
piles could be used to support the dead weight of the panel.
Cantilever Walls
The cantilever wall acts, as its name implies, like a vertical cantilever supported on a continuous
footing. Unlike the panel walls, this type of wall is subjected primarily to out-of-plane bending (Fig.
19). The cantilever wall must be built of either reinforced grouted hollow or multi-wythe masonry
(Fig. 20). To function properly this wall must be supported on a continuous foundation, usually
made of reinforced concrete. The foundation must be designed to support the weight of the wall
and be able to resist rotation caused by out-of-plane loads imposed on the wall. The reinforced
masonry wall is anchored to the foundation by steel reinforcement placed in the cells of hollow
masonry or between wythes in a multi-wythe wall. The steel reinforcement should be designed to
resist the flexural tension developed in the wall and be fully developed in both the foundation and
grouted masonry.

Out of Plane Deflection of Cantilever Wall
Fig. 19
t45 http://www.gobrick.com/BIA/technotes/t45.htm
14 of 18 9/13/2009 2:17 PM
Cantilever Wall Cross Sections
FIG. 20
Expansion joints should be placed at a maximum of 30 ft (9.1m) on center and may be detailed in a
staggered fashion for multi-wythe construction (Fig. 21). This detail ensures that the sound from a
highway cannot pass directly through the wall if the sealants fail.

Staggered Expansion Joint
FIG. 21
Other Load Considerations
Foundations. Additional stresses can be introduced in brick masonry noise barrier walls by
differential settlement or rotation of the foundation system. Soil conditions should be evaluated to
keep both differential settlement and differential rotation to a minimum in all wall systems.
However, horizontal reinforcement can be used to resist in-plane loads resulting from differential
settlement in pilaster and panel and in cantilever walls. Further, more frequent spacing of vertical
expansion joints can reduce the effect of differential settlement in these walls.
Traffic Impact. The possibility of vehicles hitting a noise barrier wall must be considered. This is
of special concern if the wall is immediately adjacent to the shoulder. Concrete deflector barriers
t45 http://www.gobrick.com/BIA/technotes/t45.htm
15 of 18 9/13/2009 2:17 PM
are recommended in this instance, and any time such devices are used traffic impact loads on the
noise barrier walls need not be considered. If traffic can reach the noise barrier wall, then these
additional loads must be considered. Horizontal and vertical reinforcement may be necessary in
the brick noise barrier wall to add ductility and post-cracking integrity.
Due to the varying traffic and site conditions it is beyond the scope of this Technical Notes to
evaluate traffic impact effects. Local highway officials should be consulted to establish these
design parameters.
Seismic. If brick masonry noise barrier walls are built in Seismic Performance Categories C or D,
they must be reinforced with a minimum amount of both horizontal and vertical reinforcement.
These reinforcement requirements can be found in the Building Code Requirements for Masonry
Structures (ACI 530 / ASCE 5 / TMS 402-02) or the local building codes. Moreover, an analysis
should be made to ensure that sufficient reinforcement is present to resist the seismic forces.

CONSTRUCTION AND DETAILING CONSIDERATIONS
Good workmanship and detailing are key to the success of all masonry assemblages, including
noise barrier walls. Full head and bed joints and proper location and installation of reinforcement,
ties, flashing and expansion joints are required for proper performance. Any unfilled joint will result
in water penetration and will degrade the effectiveness of the noise barrier wall. Proper mixing
and consistency between batches of mortar and grout is necessary. All spaces to be grouted
must be completely filled, and grouting procedures found in the local building codes must be
followed. Generally, Type S mortar as specified by proportion in ASTM C 270 Mortar for Unit
Masonry is recommended for construction of noise barrier walls. Grout should conform to ASTM
C 476 Specification for Grout for Masonry.
Two critical details in a noise barrier wall are the location and placement of copings and flashing.
Copings should project beyond the faces of the wall a minimum of 1 in. (2.5 cm) on both sides.
Stone or masonry copings should have a minimum slope of 15 degrees from horizontal and contain
a positive drip to keep water from flowing down the face of the wall. It is important that the
copings be anchored to the brick wall with metal anchors or bolts, especially in high wind and
seismic areas. Natural stone, cast stone, terra cotta, metals, and brick are suitable for copings.
If metal copings are used, they should extend down each side of the wall a minimum of 4 in. (10
cm). A sealant should be placed between the metal coping and the wall to prevent wind uplift and
water penetration (Fig. 22). When stone or concrete copings are used, an elastic sealant should
be placed between the head joints of the coping pieces (Fig. 23). If brick is used as a coping (Fig.
24), it may be prudent to use units which have the same physical requirements as brick pavers.
ASTM C 902 Standard Specifications for Pedestrian and Light Traffic Paving Brick is the
specification for these units. However, brick units that have been used successfully as a coping in
the past, should be adequate.

Metal Coping
FIG. 22

t45 http://www.gobrick.com/BIA/technotes/t45.htm
16 of 18 9/13/2009 2:17 PM
Stone Coping
FIG. 23
Brick Coping
FIG. 24
Through-wall flashing is required directly under the coping. The flashing should extend beyond the
faces of the wall to form a drip. All penetrations through the flashing made by the anchors must be
adequately sealed with a compatible material.
Brick in noise barrier walls should not be in direct contact with the ground. Salt laden ground
water could be absorbed into the brick causing efflorescence or possible spalling in the lower
courses. In some instances it may be visually and functionally necessary to have the base of the
wall in contact with the ground. In these cases, gravel instead of earth should be placed in contact
with the wall. The gravel not only keeps ground water from being absorbed by the brick masonry
but also keeps the lower courses free from staining by rain splashed earth.

MAINTENANCE CONSIDERATIONS
Brick masonry walls maintain their aesthetic appeal and remain virtually maintenance free
throughout their life. The expansion joint sealant and any sealants used in conjunction with copings
are the only elements in the wall which will require intermittent inspection and maintenance.
In some areas the noise barrier wall may be subjected to graffiti. In such an instance, an
anti-graffiti coating should be considered. However, some coatings may reduce the durability of
clay brick. Also, to remain effective, these materials may have to be re-applied.
Further, sufficient rights-of-way should be established where possible to allow for accumulations of
snow on the leeward side of the wall. The location and alignment of noise barriers should be
analyzed in order to prevent or reduce problems of drifting snow across roadways.

SUMMARY
Because our national highway system has grown significantly over the last few decades, public
awareness of traffic noise on neighborhood communities has increased. Neighborhood
associations and governmental bodies look for ways to reduce traffic noise without adversely
t45 http://www.gobrick.com/BIA/technotes/t45.htm
17 of 18 9/13/2009 2:17 PM
affecting the surrounding environment. A solution to this problem lies in brick masonry noise
barrier walls. Brick masonry noise barrier walls can easily blend into the environment and give
residential communities the protection from highway noise.
The information and suggestions contained in this Technical Notes are based on the available data
and the experience of the engineering staff of the Brick Industry Association. The information
contained herein must be used in conjunction with good technical judgment and a basic
understanding of the properties of brick masonry. Final decisions on the use of the information
contained in this Technical Notes are not within the purview of the Brick Industry Association and
must rest with the project architect, engineer and owner.

REFERENCES
1. Noise Barrier Design Handbook, Report No. FHWA-RD-76-58, United States Department of
Transportation, Federal Highway Administration, 1976.
2. A Guide to Visual Quality in Noise Barrier Design, Implementation Package 77-12, United
States Department of Transportation, Federal Highway Administration, 1977.
3. Guide Specifications for Structural Design of Sound Barriers, American Association of State
Highway and Transportation Officials, 1989.
4. Technical Notes on Brick Construction 5A, Sound Insulation-Clay Masonry Walls, Brick
Institute of America, Aug. 1986.

t45 http://www.gobrick.com/BIA/technotes/t45.htm
18 of 18 9/13/2009 2:17 PM

Technical Notes 45A - Brick Masonry Noise Barrier Walls - Structural Design
April 1992
Abstract: Rationally designed brick masonry noise barrier walls provide an attractive wall form with reliable
structural function. This Technical Notes addresses the structural design of pier and panel, pilaster and panel, and
cantilever brick noise harrier walls. Suggested design methodology and design examples are provided. The
information presented in this Technical Notes can be applied with slight modifications to the many design schemes
and loading demands of noise barrier walls. The result is an attractive noise barrier wall with the durability and
versatility inherent in brick masonry structures.

Key Words: brick, cantilever, noise barrier, pier, pilaster, structural design, wall system.

INTRODUCTION

Technical Notes 45 presented an introduction to acoustical, visual, structural, and construction considerations for
brick masonry noise barrier walls. In continuation of the series, this Technical Notes addresses structural design
considerations in greater detail and provides design examples. Recommended procedures are presented on the
structural design of the wall system assuming acoustic and visual considerations are previously addressed. Design
recommendations for noise barrier wall footings and caissons have not been addressed in this Technical Notes.
A design approach is presented which follows criteria contained in the Building Code Requirements for Masonry
Structures (ACI 530/ASCE 5/TMS 402-92) and, where applicable, the Load and Resistance Factor Design Manual
of Steel Construction-First Edition (AISC LRFD). Refer to Technical Notes 3 series for a discussion of the
ACI/ASCE/TMS document.

NOTATION1
A
s
Area of steel, in.
2
bWidth of section, in.
cDistance from extreme compression fiber to the neutral axis of the cross section, in.
dDistance from extreme compression fiber to the centroid of tension reinforcement, in.
EmModulus of elasticity of masonry in compression, psi
E
s
Modulus of elasticity of steel, psi
fa Calculated compressive stress in masonry due to axial load only, psi
45a http://www.gobrick.com/BIA/technotes/t45a.htm
1 of 24 9/13/2009 2:17 PM
fcr Modulus of rupture, psi
f
b
Calculated compressive stress in masonry due to flexure only, psi
F
b
Allowable compressive stress in masonry due to flexure only, psi
fm Compressive stress in masonry, psi
Specified compressive strength of masonry, psi
fsCalculated tensile or compressive stress in reinforcement, psi
FsAllowable tensile or compressive stress in reinforcement, psi
ftCalculated tensile stress in masonry, psi
FtAllowable flexural tensile stress in masonry, psi
fvCalculated shear stress in masonry, psi
FvAllowable shear stress in masonry, psi
f
y
Specified yield stress for reinforcement, psi
h
Height of wall or panel, ft
IcrMoment of inertia of cracked masonry cross section, in.4
IeEffective moment of inertia, in.4
IgMoment of inertia of uncracked masonry cross section, in.4
IxMoment of inertia of steel pier about the strong axis, in.4
jRatio of distance between centroid of flexural compressive forces and centroid of tensile forces to depth
kRatio of distance between compression face and neutral axis to distance between compression face and centroid
of tensile forces
MDesign moment, ft-lb
M
n
Nominal moment strength, ft-lb
M
o
Overturning moment, ft-lb
Mpx Moment due to spanning between piers, ft-lb
Mpy Moment induced by Mpx due to plate effects, ft-lb
Mr Resisting moment, ft-lb
Mu Ultimate moment strength, ft-lb
Mw Moment due to spanning between caissons, ft-lb
45a http://www.gobrick.com/BIA/technotes/t45a.htm
2 of 24 9/13/2009 2:17 PM
nElastic moduli ratio, Es/Em
PDesign axial load, lb
pReinforcement ratio, As/bd
rRadius of gyration, in.
tThickness of wall or panel, in.
VDesign shear force, lb
VnNominal shear strength, lb
VpxShear due to spanning between piers, lb
V
u
Ultimate shear strength, lb
WLateral load, lb/ft
xWidth of footing, ft
yDepth of footing, ft
DDeflection, in.
fbResistance factor

1
Metric equivalents:
1 in.= 25.4 mm
1 ft = 0.3048 m
1 lb = 4.448 N
1 psi = 0.006895 N/mm
2

SELECTION OF A WALL SYSTEM

As discussed in Technical Notes 45, there are three typical brick masonry noise barrier wall systems: cantilever
walls, pier and panel walls, and pilaster and panel walls. Preliminary consideration of design parameters can help
select the wall system that is most appropriate and efficient without having to develop and compare three separate
designs.

Cantilever Noise Barrier Walls
A cantilever wall system is better suited for shorter noise barriers, i.e. walls that are 12 ft (3.7 m) or less in height.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
3 of 24 9/13/2009 2:17 PM
In most instances, taller cantilever walls are less desirable because strip footings become too massive. Cantilever
walls are more efficient for shorter heights because they are likely the easiest and most economical to construct,
and will require the least quality control and inspection. This is because construction techniques used are similar to
building wall construction familiar to mason contractors.

Pier and Panel Noise Barrier Walls
Pier and panel wall systems are best for quick site erection. Brick panels can be prefabricated on or off site, or laid
in place. Also, pier caissons are typically constructed faster and require less concrete than strip footings. Strip
footings under the panel are not required, as the panel can span from pier to pier. Material costs for pier and panel
wall systems will typically be the least of the three systems. Disadvantages of the pier and panel system include
increased construction supervision and inspection, tight construction tolerances for pier-to-panel connections, and
increased costs to install the panels.

Pilaster and Panel Noise Barrier Walls
Pilaster and panel walls, like pier and panel walls, typically utilize caissons for quick foundation construction. Wall
construction is done on site, as the panel is built integral with the pilaster. This requires panel support between
caissons during construction. Supervision and inspection are required to ensure proper construction. However,
construction tolerances are more liberal than those for pier and panel systems. Generally, pilaster and panel wall
systems permit longer pier spacing and taller wall height. A pilaster and panel wall assembly is structurally more
efficient than a pier and panel wall assembly, as a fixed condition may be developed at the pilaster-panel connection.

DESIGN ASSUMPTIONS

It has been widely accepted that masonry stress-strain behavior is similar to that of concrete. Thus, design
assumptions made for masonry under working stress and strength conditions are analogous to assumptions made in
concrete design. Figure 1 depicts the assumed stress-strain relationship for masonry in flexure under working loads.
In all cases, the principles of equilibrium and compatibility of strains of masonry materials are assumed to apply.
Assumptions made following a working stress design are as follows: 1) plane sections before bending remain plane
after bending, 2) moduli of elasticity of masonry and steel remain constant, 3) reinforcement is completely bonded to
masonry, and 4) in cracked masonry members, the tensile capacity of masonry is neglected.


45a http://www.gobrick.com/BIA/technotes/t45a.htm
4 of 24 9/13/2009 2:17 PM
Straight Line Stress Distribution
FIG. 1

In this Technical Notes, a number of additional assumptions will be made to facilitate design. It is assumed that the
brick masonry will be reinforced. Most noise barrier wall applications demand tall slender walls to meet acoustic
requirements and minimize material costs and land use. Reinforcing is required for brick masonry to meet these
criteria. Additional assumptions placed on both material properties and wall behavior are as follows.

Material Properties
Grout and concrete are assumed to have compressive strength equal to or greater than the masonry compressive
strength, and elastic moduli of the masonry and the grout are assumed to be equal. The method of transformation of
areas may be used in lieu of these assumptions.

Wall Behavior
Masonry walls are plate structures. Thus, a masonry wall loaded perpendicular to its plane will experience strain
along its length and its height. However, the traditional masonry wall design approach is to use the strip method. In
this method, a one foot wide section of wall is designed considering one span direction. Strains perpendicular to the
strip span direction are ignored. For cantilever walls, this method is nearly exact, as plate effects are negligible. Pier
and panel and pilaster and panel walls, however, exhibit wall behavior which can make plate effects significant. This
does not mean a rigorous plate analysis is necessary for these walls. Rather, a few simple observations and
assumptions can be made to simplify design.
For pier and panel and pilaster and panel wall systems, the panel is subject to three different deflection conditions: a
horizontal simple span between piers or pilasters subject to wind or seismic load, a horizontal simple span between
caissons subject to panel weight, and a vertical cantilever span subject to deflection of the pier or pilaster. If the
panel is free to deflect both in-plane and out-of-plane, the moment due to simple spanning between piers or
pilasters, Mpx, and the moment due to simple spanning between caissons, Mw, are combined by vector addition to
calculate the maximum design moment for the horizontal span of the panel. However, the panel must be flush with
the ground to avoid noise penetration under the wall. Thus, the panel may, in fact, be supported along its entire
length by the ground. This is significant, because the design moment in this case is solely Mpx, the moment about the
weak axis of the panel. This condition will require the most amount of horizontal reinforcement for the panel. In the
panel design examples that follow it is assumed that the ground supports the entire length of the panel.
Because of plate effects, Mpx will induce moment about the horizontal axis as well, denoted as Mpy. The strip
solution does not and cannot calculate Mpy, as plate effects are ignored in this method. However, plate analysis
shows that Mpy can be significant and that the ratio of Mpy to Mpx increases as the height to length ratio of the panel
increases. As an approximation, Mpy is calculated as one tenth the height to length ratio times Mpx. Mpy is a
maximum at the middle of the panel. However, moment due to vertical cantilever deflection of the wall is a maximum
at the bottom of the panel. Thus, the design moment about the horizontal axis is the greater of: 1) the moment due
to vertical cantilever deflection at the bottom of the panel, or 2) the sum of Mpy and the moment due to vertical
cantilever deflection at the middle of the panel.

Vertical cantilever deflection of the panel is a function of the rigidity of the piers or pilasters. If the piers or pilasters
are very rigid, cantilever deflection of the panel will be negligible. However, optimal flexural design may result in less
rigid piers or pilasters with considerable deflection, especially when steel piers are used. Induced tensile stresses in
the panel must be within allowable tensile stresses for unreinforced masonry if the panel cannot be reinforced in the
vertical direction. Thus, deflection criteria will often govern pier and pilaster design.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
5 of 24 9/13/2009 2:17 PM
Reinforced brick masonry pilaster and panel and pier and panel wall systems are typically very rigid, so deflections
in many cases will be small. However, the deflection of the pilaster must be calculated considering the ratio of
applied moment to cracking moment. Cracking moment is calculated using the gross moment of inertia of the pier or
pilaster.
In the pilaster and panel design example that follows, a two span continuous panel is assumed. Thus, the pilaster
panel interface is assumed to be a fixed connection for the middle pilaster, and a simple connection for the two
exterior supports. This allows for expansion joints at the simple supports to accommodate horizontal expansion of
the panels.
Lastly, compression steel in the pilaster is usually ignored in design. If consideration of the increased compressive
strength due to the compression steel is made, the steel must be properly confined within the pilaster with lateral or
spiral ties.

DESIGN PROCEDURE

It is important to establish a set design procedure to ensure an accurate and comprehensive noise barrier wall
design. The following nine steps are presented as a guide to the structural design of a brick masonry noise barrier
wall. Additional criteria may be warranted for a particular wall design scheme.
1) Determine required wall height based upon acoustical considerations.
2) Determine critical lateral and axial load combinations on wall elements. Loads should be determined according to
the recommendations of the local building code or as contained in the document Minimum Design Loads for
Buildings and Other Structures (ASCE 7). For the examples that follow, inertial wall force due to seismic base shear
is divided by wall surface area for comparison with wind loads.
3) To determine required reinforcement, assume j = 0.9:
A
s
req'd=M/F
s
jd
4) Calculate masonry compressive stresses and the steel tensile stress:
f
b
= 2M/jkbd
2
f
a
= P/bkd
f
s
= M/A
s
jd
5) Check the allowable compressive stress in masonry and the tensile stress in steel (Table 1). Axial compression
and buckling seldom govern design of noise barrier wall elements. However, axial compression must be included to
calculate the maximum flexural compression. Note that allowable stresses for wind or seismic load conditions may
be increased by one-third over those given in Table 1.

45a http://www.gobrick.com/BIA/technotes/t45a.htm
6 of 24 9/13/2009 2:17 PM
1
Allowable stresses for wind and seismic loading conditions may be increased by one-third.

6) Calculate shear stress:
f
v
= V/bjd
7) Check the allowable shear stress in masonry (Table 1). If exceeded, member must be reinforced for shear and
the shear stress checked.
8) Design the pier or pilaster, if applicable. If the pier or pilaster is made of reinforced masonry, design will follow
steps 3 through 7. If a steel pier is used, follow the design recommendations given in the Load and Resistance
Factor Design Manual of Steel Construction. Flexural tensile stresses developed in unreinforced masonry panels
due to pier or pilaster deflection may not exceed the allowable flexural tensile stresses given in Table 2.

45a http://www.gobrick.com/BIA/technotes/t45a.htm
7 of 24 9/13/2009 2:17 PM

1
Allowable stresses for wind and seismic loading conditions may be increased by one-third.
2
For partially grouted masonry allowable stresses shall be determined on the basis of linear
interpolation between hollow units which are fully grouted or ungrouted and hollow units
based on amount of grouting.
9) Calculate overturning and resisting moments, and sliding resistance (Fig.2). These are functions of the wall,
footing, and caisson dimensions, as well as the soil pressure resistance. The factor of safety is the ratio of resisting
moment, Mr, to the overturning moment, Mo. A factor of safety of 2 or greater is recommended.

45a http://www.gobrick.com/BIA/technotes/t45a.htm
8 of 24 9/13/2009 2:17 PM

Wall/Footing Forces
FIG. 2

Hollow Brick Cantilever Wall
FIG. 3

45a http://www.gobrick.com/BIA/technotes/t45a.htm
9 of 24 9/13/2009 2:17 PM
DESIGN EXAMPLES
Design Example #1: Hollow Brick Cantilever Noise Barrier Wall
Type S portland cement/lime mortar, f
'
m
= 3600 psi,
E
m
= 3.0 x 10
6
psi, n = 9.7
Wall dimensions shown in Fig. 3, running bond, face shell bedding
Grade 60 reinforcement,
Loads: wind = 20 psf, wall weight =73 psi
Step 1: Based on acoustical considerations, the wall height shall be 10 ft minimum.
Step 2: Critical load combinations result in the following design values (per ft of wall):
M = (0.5)(W)(h)
2
= (0.5)(20 psf) (10ft)
2
= 1000 ft-ob
V = (W)(h) = (20 psf)(10ft) = 200 lb
P = (wall weight)(h) = (73 psf)(10 ft) = 730 lb

Step 3: Calculate required reinforcement.



Try #5 bars @32 in. o.c., per foot of wall (Table 3).

Step 4: Calculate the masonry compressive stress and the steel tensile stress.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
10 of 24 9/13/2009 2:17 PM

1
Area of steel listed is for one wire.

Step 5: Check compressive stress in masonry and the tensile stress in steel.
Step 6: Calculate shear stress.
Step 7: Check shear stress.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
11 of 24 9/13/2009 2:17 PM
Step 8: Does not apply.
Step 9: Neglect the soil pressure resistance to overturning. Calculate overturning and resisting moments, and the
safety factor on overturning (Fig. 2). Note: Sliding must also be considered. Sliding is a function of the soil
conditions, and is beyond the scope of this Technical Notes.
Design Example #2: Steel Pier with 4 Inch Panel Noise Barrier Wall
Type S portland cement/lime mortar, f
'
m
= 3000 psi,
E
m
= 2.8 x 10
6
psi
Wall dimensions shown in Fig. 4
Running bond, full bedding, spacing of piers shall be 15ft
Ladder type joint reinforcement:
Pier I-beam: Grade 50 W Section,
Loads: wind = 15psf, panel weight = 40psf, seismic = 25psf


Noise Barrier Wall Design Example #2
45a http://www.gobrick.com/BIA/technotes/t45a.htm
12 of 24 9/13/2009 2:17 PM
Steel Pier With 4 Inch Panel
FIG. 4

Step 1: Based on acoustical considerations, the wall height shall be 16 ft minimum.

Step 2: Critical load combinations result in the following design values:
a) For the panel span between piers (per foot of wall):
M
px
= (1/8)(W)(L)
2
= (1/8)(25 psf)<15 ft) = 703 ft-lb
m
py
= (1/10)(h/L)(M
px
) = (1/10)(16 ft/15 ft) (703 ft-lb) = 75 ft-lb
V
px
= (0.5)(25 psf)(15 ft) = 188 lb
b) For the panel span between caissons, assume the panel is supported along its entire base by the ground.
c) For an interior steel pier:
V = (2)(V
px
)(h) = (2)(188 lb/ft)(16 ft) = 6016 lb
V
u
= 1.3V = 1.3 (6016 lb) = 7820 lb
Step 3: Calculate required reinforcement for the panel for the design moment Mpx. Assume the distance from the
extreme compression face to the joint reinforcement is 3 inches.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
13 of 24 9/13/2009 2:17 PM

1
Area of steel listed is for one wire.

Step 4: Calculate the masonry compressive stress and the steel tensile stress.
Step 5: Check compressive stress in masonry and the tensile stress in steel.
Step 6: Calculate shear stress.
Step 7: Check shear stress.
Step 8: Assume the pier design will be governed by deflection limitations of the panel, not the required flexural
45a http://www.gobrick.com/BIA/technotes/t45a.htm
14 of 24 9/13/2009 2:17 PM
strength of the pier. Calculate the maximum vertical cantilever deflection the panel may undergo without exceeding
the allowable tensile stress value for the panel found in Table 2. From Table 2:
Step 9: Sliding and overturning resistance of the caisson is a function of the lateral soil pressure and is beyond the
scope of this Technical Notes.

Design Example #3: Reinforced Brick Masonry Pier and 4 Inch Panel Noise Barrier Wall
45a http://www.gobrick.com/BIA/technotes/t45a.htm
15 of 24 9/13/2009 2:17 PM

Noise Barrier Wail Design Example #3
Reinforced Brick Pier and 4 Inch Panel
FIG. 5

Step 1: Based on acoustical considerations, the wall height shall be 10 ft minimum.

Step 2: Critical load combinations result in the following design values:
45a http://www.gobrick.com/BIA/technotes/t45a.htm
16 of 24 9/13/2009 2:17 PM
Step 3: Calculate the maximum permissible spacing of piers with the given joint reinforcement based on the design
moment, Mpx. Assume the distance from the extreme compression face to the joint reinforcement is 3 inches.
Step 4: Calculate the masonry compressive stress and the steel tensile stress.
Step 5: Check the compressive stress in masonry and the tensile stress in steel.
Step 6: Calculate shear stress.
Step 7: Check shear stress.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
17 of 24 9/13/2009 2:17 PM
Step 8: Design the reinforced brick masonry pier. Repeat steps 3 through 7 with the following design values:
Step 3a: Calculate required pier reinforcement.
Step 4a: Calculate the masonry compressive stress and the steel tensile stress.
Step 5a: Check compressive stress in masonry and the tensile stress in steel.
Step 6a: Calculate shear stress.
Step 7a: Check shear stress.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
18 of 24 9/13/2009 2:17 PM
Step 9: Sliding and overturning resistance of the caisson is a function of the lateral soil pressure and is beyond the
scope of this Technical Notes.

Design Example #4: Reinforced Brick Pilaster and 6 inch Panel Noise Barrier Wall
45a http://www.gobrick.com/BIA/technotes/t45a.htm
19 of 24 9/13/2009 2:17 PM
Noise Barrier Wall Design Example #4
Reinforced Brick Pilaster and 6 inch Panel
FIG. 6

Step 1: Based on acoustical considerations, the wall height shall be 6 ft minimum.
Step 2: Critical load combinations result in the following design values:
45a http://www.gobrick.com/BIA/technotes/t45a.htm
20 of 24 9/13/2009 2:17 PM
Step 3: Calculate required reinforcement for the panel for the design moment, Mpx. Assume the distance from the
extreme compression face to the joint reinforcement is 4.9 inches.
Step 4: Calculate the masonry compressive stress and the steel tensile stress.
Step 5: Check compressive stress in masonry and the tensile stress in steel.
Step 6: Calculate shear stress.
Step 7: Check shear stress.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
21 of 24 9/13/2009 2:17 PM
Step 8: Design the reinforced brick masonry pilaster. Repeat steps 3 through 7.
Step 3a: Calculate required pilaster reinforcement.
Step 4a: Calculate the masonry compressive stress and the steel tensile stress.
Step 5a: Check compressive stress in masonry and the tensile stress in steel.
Step 6a: Calculate shear stress.
Step 7a: Check shear stress.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
22 of 24 9/13/2009 2:17 PM
Step 9: Sliding and overturning resistance of the caisson is a function of the lateral soil pressure and is beyond the
scope of this Technical Notes.
SUMMARY
This Technical Notes discusses the structural design of brick masonry noise barrier walls. Design procedures are
given following working stress analysis provided in the ACI/ASCE/TMS standard. Example designs are presented for
recommended design of pier and panel, pilaster and panel and cantilever brick noise barrier walls.
The information and suggestions contained in this Technical Notes are based on the available data and the
experience of the technical staff of the Brick Institute of America. The information and recommendations contained in
this publication must be used in conjunction with good engineering judgment and a basic understanding of the
properties of brick masonry and related construction materials. Final decisions on the use of the information
contained in this Technical Notes are not within the purview of the Brick Institute of America and must rest with the
45a http://www.gobrick.com/BIA/technotes/t45a.htm
23 of 24 9/13/2009 2:17 PM
project designer and owner.

REFERENCES

1.Anand, S.C., Brick Masonry Noise Barrier Walls, A Refined Structural Analysis of Wall Panel for
Pier/Panel Case, Submitted to the Brick Institute of America, December 1991, 16 pp.
2.Building Code Requirements for Masonry Structures (ACI 530/ASCE 5/TMS 402-92) and Specifications
for Masonry Structures (ACI 530.1/ASCE 6/TMS 602-92), American Concrete Institute, Detroit, MI, 1992.
3.Guide Specifications for Structural Design of Sound Barriers, American Association of State Highway and
Transportation Officials, Washington, D.C., 1989.
4.Load and Resistance Factor Design Manual of Steel Construction-First Edition, American Institute of Steel
Construction, Chicago, IL, 1986.
5.Minimum Design Loads for Buildings and Other Structures (ASCE 7-88), American Society of Civil
Engineers, New York, NY, 1990.
6.Noise Barrier Design Handbook, (Report No. FHWA-RD-76-58), United States Department of
Transportation, Federal Highway Administration, Washington, D.C., 1976.
7.Technical Notes on Brick Construction 3, "Building Code Requirements for Masonry Structures ACI
530/ASCE 5 and Specifications for Masonry Structures ACI 530.1/ASCE 6," February 1990.
8.Technical Notes on Brick Construction 45, "Brick Masonry Noise Barrier Walls - Introduction," February
1991.
9.Uniform Building Code, 1991 Edition, International Conference of Building Officials, Whittier, California,
1991.
45a http://www.gobrick.com/BIA/technotes/t45a.htm
24 of 24 9/13/2009 2:17 PM
Maintenance of Brick Masonry
Abstract: Even though one of the major advantages of brick masonry construction is durability, periodic inspections and
maintenance can extend the life of brickwork in structures. This Technical Note discusses the benefits and elements of sug-
gested inspection programs and describes specific maintenance procedures including replacement of sealant joints, grouting
of mortar joint faces, repointing of mortar joints, removal of plant growth, repair of weeps, replacement of brick, installation of a
dampproof course, installation of flashing in existing walls and replacement of wall ties.
Key Words: anchors, cleaning, dampproof course, efflorescence, flashing, inspection, maintenance, moisture penetration,
mortar, repointing, sealant, ties, weeps.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction 46
December
2005
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Perform periodic inspections, preferably each season
Determine moisture source before attempting repairs to
correct moisture penetration
Remove and replace torn, deteriorated or inelastic seal-
ants
When repairing mortar joints, surface grout hairline cracks
and repoint damaged or deteriorating mortar joints
Repoint with prehydrated Type N, O or K mortar, mixed
drier than for conventional masonry work
Remove ivy and plant growth that contributes to moisture
penetration or deterioration of brickwork
Exercise care in opening existing or drilling new weeps, to
ensure that flashing is not damaged
Install a dampproof course if missing or required
Install remedial anchors and ties in accordance with man-
ufacturers recommendations
Inspect masonry and correct all deficiencies before appli-
cation of external coatings
Page 1 of 7
INTRODUCTION
This Technical Note discusses maintenance of brick masonry with an emphasis on preventing moisture penetra-
tion. All buildings are unique and may experience different problems. A given solution may not remedy similar prob-
lems on all buildings. It is therefore suggested that a repair method which will effectively suit the particular needs of
a building be selected when a problem occurs.
Generally, if brickwork is properly designed, detailed and constructed, it is very durable and requires little mainte-
nance. However, many of the other components incorporated in the brickwork such as caps, copings, sills, lintels
and sealant joints may require periodic inspection and repair. Neglecting maintenance of these components may
lead to deterioration of other elements in the wall.
Maintenance of buildings may be broken into two general categories: 1) general inspection to identify potential
problems with the performance of exterior walls; and 2) specific maintenance to correct problems which may
develop. This Technical Note addresses both general and specific maintenance procedures. A checklist is provided
for general inspections and specific repair techniques are described.
GENERAL INSPECTION
A thorough inspection and maintenance program may
help extend the life of a building. It is a good idea to
become familiar with the materials used in a building
and how they perform over a given time period. Table
1 lists various building materials and the estimated
time before repair may be needed, given normal expo-
sure. These times are based on brickwork in vertical
applications, constructed of proper materials and
workmanship and exposed to normal weathering con-
ditions in the United States. Sills, parapets, chimneys
and copings which experience more severe exposures
may require repairs at shorter intervals.
Periodic inspections should be performed to determine
Brick Walls 100+
Sealant J oints 5-20
Metal Coping/Flashing 20-75
Metal Anchors & Ties 15+
Mortar Walls 25+
Plastic Flashing 5-25
Fi ni shes
Paint Appearance 3-5
Water Repellents Dampproofing 5-10
Stucco Appearance 5-10
Materi al Use
Esti mated Ti me to Repai r
(Years)
TABLE 1
Estimated Time to Repair of Materials
the condition of the various materials used on a building. These inspections can be performed monthly, yearly,
biennially, or any time period deemed appropriate. Seasonal inspection periods are recommended so that the
behavior of building materials in various weather conditions can be noted. Inspection records, including conditions
and comments, should be kept to identify changes in materials, potential problems and needed repair. Table 2 is
a suggested checklist of conditions that may require maintenance or repair. It is not all-inclusive; however, it may
establish a guideline for use during inspections.
Conditions that may necessitate maintenance or repair actions include efflorescence, spalling, deteriorating mor-
tar joints, interior moisture damage and mold. Once one or more of these conditions becomes evident, the origin
of the problem should be determined and action taken to correct both the cause and visible effect of the condi-
tion. Table 3 lists various conditions affecting brickwork and their most probable sources. The items checked in
the table represent each source that should be considered when such conditions are observed in brick masonry.
SPECIFIC MAINTENANCE
After investigating all of the possible contributors the actual cause(s) of distress conditions may be determined
through the process of elimination. Often the source will be self-evident as with deteriorated and missing materi-
als; however, in instances such as improper flashing or differential movement the source may be hidden and
determined only through building diagnostics. In any case, it is suggested to first visually inspect for the self-evi-
dent source before performing a more extensive investigation as it may save time and money in detecting the
cause. Such a process should always be followed if the condition involves water penetration. Once the source is
determined, measures can be taken to effectively remedy the moisture penetration source and its effects on the
brickwork.
www.gobrick.com | Brick Industry Association | TN 46 | Maintenance of Brick Masonry | Page 2 of 7
NORTH SOUTH EAST WEST
Cracked Units
Loose Units
Spalled Units
Hairline Cracks in Mortar
Deteriorated Mortar J oints
Missing or Clogged Weeps
Plant Growth
Deteriorated/Torn Sealants
Out-of-Plumb
Efflorescence
Stains
Water Penetration
Damaged
Open Lap J oints
Missing
Stains
Inadequate Slope
Cracked Units
Hairline Cracks in Mortar
Loose J oints
Open J oints
Out-of-Plumb
Drips Needed
Deteriorated Mortar J oints
Cracks
Separation from Flooring
Inadequate Drainage
Water Penetration
Spalled Units
Deteriorated Mortar J oints
Cracks
Out-of-Plumb
Dampness
Inadequate Drainage
Roof Overhangs
Gutters/Leaders
Seal at Adjacent Materials
Grade/Drainage
M
a
s
o
n
r
y
C
a
p
s
/

C
o
p
i
n
g
s
/

S
i
l
l
s
ITEM OR CONDITION
BUILDING ELEVATION
LOCATION
A
b
o
v
e

G
r
a
d
e
B
e
l
o
w

G
r
a
d
e
O
t
h
e
r

E
l
e
m
e
n
t
s
R
e
t
a
i
n
i
n
g

W
a
l
l
s
F
o
u
n
d
a
t
i
o
n

W
a
l
l
s
F
l
a
s
h
i
n
g
/

C
o
u
n
t
e
r
-
f
l
a
s
h
i
n
g
TABLE 2
Brick Masonry Inspection Checklist
www.gobrick.com | Brick Industry Association | TN 46 | Maintenance of Brick Masonry | Page 3 of 7
Removing
Efflorescence
Generally, efflorescence is water-
soluble and easily removed by natural
weathering or by scrubbing with a
brush and water. Proprietary cleaners
formulated specifically for use on brick-
work are effective in removing stubborn
efflorescence (see Technical Note 20).
Use solutions specifically manufactured
to remove efflorescence from brick-
work. Improper acid cleaning proce-
dures such as insufficient prewetting,
rinsing and strong acid concentrations
may cause additional staining, etched
mortar joints and increase moisture
penetration in brickwork. Stains caused
by improper cleaning are not water-
soluble, but can be removed by propri-
etary cleaners.
All cleaning procedures should first
be tried at different concentrations in
an inconspicuous area to judge their
effectiveness and potential harm to the
brickwork. Additional recommendations and cleaning methods for brick masonry are presented in Technical Note
20. After cleaning, the mortar joints should be inspected. Repointing or grouting of the joints, as discussed later in
this Technical Note, may be necessary.
Sealant Replacement
Missing or deteriorated sealants in and between brickwork and other materials such as windows, door frames and
expansion joints may be a source of moisture penetration. The sealant joints in these areas should be inspected
closely to discover areas where the sealant is missing, or was installed but has deteriorated, torn or lost elastic-
ity. Deteriorated sealants should be carefully cut out and the opening cleaned of all existing sealant material. The
clean joint should then be properly primed and filled with a backer rod (bond breaker tape if the joint is too small to
accommodate a backer rod) and a full bead of high-quality, elastic sealant compatible with adjacent materials.
Mortar Joint Repair
Repair of cracked or deteriorating mortar joints is very effective in reducing the amount of water that enters exte-
rior masonry. Cracks in brickwork that are more than a few millimeters in width or that are suspected to have
been caused by settlement or other structural problems (for example, cracks that continue through multiple brick
units and mortar joints, or follow a stepped or diagonal pattern along mortar joint) are beyond the scope of this
Technical Note. These cracks often require professional investigation to determine the cause and appropriate
method of repair.
Grouting of Hairline Cracks. If the mortar joints develop small hairline cracks, surface grouting may be an
effective measure to fill them. The impact of surface grouting on brickwork aesthetics should be considered before
work begins as the appearance of the mortar joints will change somewhat. A recommended grout mixture is 1 part
portland cement,
1
/3 part hydrated lime and 1
1
/3 parts fine sand (passing a No. 30 sieve). The joints to be grouted
should be dampened. To ensure good bond, the brickwork must absorb all surface water. Clean water is added to
the dry ingredients to obtain a fluid consistency. The grout mixture should be applied to the joints with a stiff fiber
brush to force the grout into the cracks. Two coats are usually required to effectively reduce moisture penetration.
Tooling the joints after the grout application may help compact and force the grout into the cracks. The use of a
template or masking tape may be effective in keeping the brick faces clean.
Repointing Mortar Joints. Moisture may penetrate mortar which has softened, deteriorated or developed visible
TABLE 3
Possible Sources and Effects of Masonry Distress
Observed
Condi ti on
P
o
t
e
n
t
i
a
l

C
a
u
s
e

o
f













C
o
n
d
i
t
i
o
n
I
n
c
o
m
p
l
e
t
e
l
y

F
i
l
l
e
d

M
o
r
t
a
r

J
o
i
n
t
s

S
e
e

T
e
c
h
n
i
c
a
l

N
o
t
e

7
B
M
i
s
s
i
n
g
/
C
l
o
g
g
e
d

W
e
e
p
s
P
l
a
n
t

G
r
o
w
t
h
D
e
t
e
r
i
o
r
a
t
e
d
/
T
o
r
n

S
e
a
l
a
n
t
s
C
a
p
i
l
l
a
r
y

R
i
s
e
M
i
s
s
i
n
g
/
D
a
m
a
g
e
d

F
l
a
s
h
i
n
g














S
e
e

T
e
c
h
n
i
c
a
l

N
o
t
e
s

7

S
e
r
i
e
s
D
i
f
f
e
r
e
n
t
i
a
l

M
o
v
e
m
e
n
t















S
e
e

T
e
c
h
n
i
c
a
l

N
o
t
e
s

1
8

S
e
r
i
e
s
P
r
e
v
i
o
u
s

A
c
i
d

C
l
e
a
n
i
n
g















S
e
e

T
e
c
h
n
i
c
a
l

N
o
t
e

2
0
P
r
e
v
i
o
u
s

S
a
n
d
b
l
a
s
t
i
n
g
















S
e
e

T
e
c
h
n
i
c
a
l

N
o
t
e

2
0
. . .
. . . . . .
. . . . . . . .
. . . . . .
. . . . . .
. . . . .
. . . . . .
. . . . . .
Deteriorated Mortar
Spalled Units
Cracked Units
Moisture Related Stains
Corrosion of Backing
Materials
Mildew/Algae Growth
Damaged Interior Finshes
Efflorescence
See TN 23 Series
www.gobrick.com | Brick Industry Association | TN 46 | Maintenance of Brick Masonry | Page 4 of 7
cracks, as shown in Photo 1. When this is the case, repointing
(sometimes referred to as tuckpointing) may be necessary to
reduce moisture penetration. Repointing is the process of remov-
ing damaged or deteriorated mortar to a uniform depth and plac-
ing new mortar in the joint, as shown in Photo 2 and Figure 1.
Prior to undertaking a repointing project, the following should be
considered: 1) The potential for power tools to damage the brick
surrounding the mortar being cut out. 2) Repointing operations
should only be performed by qualified and experienced repoint-
ing craftsmen. An individual who is an excellent mason may
not be a good repointing craftsman. Skills should be tested and
evaluated prior to the selection of the contractor or craftsman.
3) When repointing for historic preservation purposes, refer to
Preservation Brief 2: Repointing Mortar Joints in Historic Masonry
Buildings. [Ref. 7]
The deteriorated mortar should be removed, by means of a
toothing chisel or a special pointers grinder, to a uniform depth
(refer to Figure 1b) that is twice the joint width or until sound mor-
tar is reached. Care must be taken not to damage the brick edges.
Remove all dust and debris from the joint by brushing, blowing
with air or rinsing with water.
Repointing mortar should be carefully selected and properly pro-
portioned. For best results, the original mortar constituents and
proportions should be duplicated. If this is not possible, select a
mortar that is similar or lower in compressive strength. Type N, O
and K mortar are generally recommended, as mortars with higher
cement contents may be too strong for proper performance.
Proper proportions for Type K mortars are 1 part portland cement,
4 parts hydrated lime and 11
1
/4 to 15 parts fine sand. Refer to
Technical Note 8 for material proportions of Type N and O mortar.
The repointing mortar should be prehydrated to reduce excessive
shrinkage. The proper prehydration process is as follows: All dry
ingredients should be thoroughly mixed. Only enough clean water
should be added to the dry mix to produce a damp consistency
which will retain its shape when formed into a ball. The mortar
should be mixed to this dampened condition 1 to 1
1
/2 hr before
adding water for placement.
The joints to be repointed should be dampened, but to ensure a
good bond, the brickwork must absorb all surface water before
repointing mortar is placed. Water should be added to the pre-
hydrated mortar to bring it to a workable consistency (somewhat
drier than conventional mortar). The mortar should be packed
tightly into the joints in thin layers (
1
/4 in. [6.4 mm] maximum), as
shown in Figure 1c. The joints should be tooled to match the origi-
nal profile after the last layer of mortar is thumbprint hard, as in
Figure 1d. As it may be difficult to determine which joints allow
moisture to penetrate, it is advisable to repoint all mortar joints in
the affected wall area.
If only portions of the wall area are repointed, the repointing mor-
tar should match the color of the existing mortar. Mortar materi-
als should be mixed and the color matched to existing mortar
that has been wetted. Several mix proportions can be made and
placed on extra brick. Selection is made after the mortar specimens
a) Deteriorated Mortar
J oint
b) Mortar Cut Back to
Uniform Depth
c) Pack Pointing Mortar
in Thin Layers
d) Tool J oint to Match
Original Profile
Figure 1
Repointing Mortar Joints
Photo 1
Mortar Joints in Need of Repointing
Photo 2
Repointing Mortar Joints
are dried and compared to dry existing mortar.
Plant Removal
Certain types of plant growth may contribute to moisture pen-
etration. For example, ivy shoots, sometimes referred to as
suckers, penetrate voids in mortar and may conduct mois-
ture into these voids. If this is the case, ivy removal may be
necessary.
To effectively remove ivy and similar plants, the vines should
be carefully cut away from the wall. The vines should never
be pulled from the wall as this could damage the brickwork.
After cutting, the shoots will remain. These suckers should be
left in the wall until they dry up and shrivel. This usually takes
2 to 3 weeks. Care should be taken not to allow the suckers
to rot as this could make them difficult to remove. Once the
shoots dry, the wall should be dampened and scrubbed with
a stiff fiber brush and water. Laundry detergent or weed killer
may be added to the water in small concentrations to aid in
the removal of the shoots. If these additives are used, the wall
must be thoroughly rinsed with clean water before and after
scrubbing.
To determine how the wall will appear once the ivy is
removed, it is suggested that a small portion of the ivy (5-10
ft
2
[0.5 to 1.0 m
2
]) be removed from an inconspicuous area
first. Repointing of the mortar joints may be necessary if the
mortar has cracked or deteriorated.
Opening Weeps
Weeps should be inspected to ensure that they are open and
appropriately spaced so that moisture within the walls is able
to escape to the exterior. If weeps are clogged, they can be
cleaned out by probing with a thin dowel or stiff wire. If the
weeps were not properly spaced, drilling new weeps may be
necessary. Technical Note 7 outlines suggested types and
spacing of weeps.
Since weeps are placed directly above flashing, care must be
exercised to not damage the flashing when probing or drilling.
The use of a stopper to limit the depth of penetration of the
probe or drill bit may be effective in reducing the possibility
of damaging the flashing where it turns up inside of the brick
wythe.
Replacement of Brick
Moisture may penetrate brick that are broken or heavily spalled. When this occurs, it may be necessary to replace
the affected units. The procedure shown in Figure 2 is suggested for removing and replacing brick.
The mortar that surrounds the affected units should be cut out carefully to avoid damaging adjacent brickwork, as
shown in Figure 2b. For ease of removal, the brick to be removed can be broken. Once the units are removed,
all of the surrounding mortar should be carefully chiseled out, and all dust and debris should be swept out with a
brush. If the units are located in the exterior wythe of a drainage wall, care must be exercised to prevent debris
from falling into the air space, which could block weeps and interfere with moisture drainage.
The brick surfaces in the wall should be dampened before new units are placed, but the masonry should absorb
all surface moisture to ensure a good bond. The appropriate surfaces of the surrounding brickwork and the
b) Remove Brick and Mortar
www.gobrick.com | Brick Industry Association | TN 46 | Maintenance of Brick Masonry | Page 5 of 7
a) Damaged Brick
c) Butter Replacement Brick and
Carefully Shove into Place
Figure 2
Replacement of Deteriorated Brick
www.gobrick.com | Brick Industry Association | TN 46 | Maintenance of Brick Masonry | Page 6 of 7
replacement brick should be buttered with mortar.
The replacement brick should be centered in the
opening and pressed into position, refer to Figure 2c.
The excess mortar should be removed with a trowel.
Pointing around the replacement brick will help to
ensure full head and bed joints. When the mortar
becomes thumbprint hard, the joints should be tooled
to match the original profile.
Mortar proportions are selected as discussed in the
section on Repointing. Matching the existing mortar
color is important to keep the replacement location from
being different in appearance. Similarly, replacement
brick must match the color, texture and size of the exist-
ing brick. Locating a matching brick may take consider-
able effort.
Installation of a Dampproof
Course
Moisture may migrate upward through brickwork by
capillary action. This condition appears as a rising water
line or tide mark on the wall and is referred to as ris-
ing damp.
Model building codes require the use of a dampproof-
ing material on below grade masonry walls and flash-
ing above grade. If these are omitted or improperly
installed, rising damp may occur. The insertion of a
dampproof course at a level above the ground, but
below the first floor, may stop the rising moisture. The
installation procedure can take one of two forms. One
form is the injection of a synthetic chemical that forms
a continuous dampproof barrier into an existing brick
course. Holes are drilled into the course of brick and the
synthetic material is injected. The other form of installa-
tion is the insertion of flashing through the brick wythe.
One or more brick courses are removed, flashing is
inserted, and the brick is replaced. Recommendations
for brick removal and replacement are discussed in the
following section.
Installation of Flashing
Flashing that has been omitted, damaged or improp-
erly installed may permit moisture to penetrate to the
building interior. If this is the case, a difficult procedure
of removing brick, installing flashing and replacing the
units may be required.
To install continuous flashing in existing walls, alternate
sections of masonry in 2 to 5 ft (610 mm to 1.52 m)
lengths should be removed. The flashing is installed in
these sections and the masonry replaced, refer to Photo 3. Alternately, temporary braces can be installed as lon-
ger sections of brickwork are removed, as shown in Photo 4. The flashing can then be placed in these sections.
The lengths of flashing should be lapped a minimum of 6 in. (152 mm) and be completely sealed to function prop-
erly. See Technical Note 7 for other flashing installation recommendations. The opening is then filled as discussed
under Replacement of Brick. The replaced masonry should be properly cured (5 to 7 days) before the intermedi-
ate masonry sections or supports are removed.
Photo 3
Flashing Installed in Alternating Sections
Photo 4
Flashing Installation Using Temporary Support
a) Steel Stud Backing /
Mechanical Expansion Anchor
b) Concrete Backing /
Epoxy Adhesive Anchor
c) Masonry Backing /
Grout Expansion Anchor
d) Wood Stud Backing /
Helical Screw Anchor
Figure 3
Masonry Re-Anchoring Systems
www.gobrick.com | Brick Industry Association | TN 46 | Maintenance of Brick Masonry | Page 7 of 7
Installation of Wall Ties and Anchors
In instances where masonry walls have been constructed without a sufficient number of connectors or the existing
connectors have failed, retrofit anchors may be used to attach the wythes or veneer and transfer lateral loads.
Installing anchors in such a wall improves its strength and reduces the potential for cracking. Installation of most
retrofit anchors involves drilling small holes in the masonry, usually in a mortar joint, through which the anchors
are attached to the substrate. Generally, mechanical expansion, helical screws, grout- or epoxy-adhesive systems,
shown in Figure 3, are used to make the connection. Because the installation methods and limitations of each
product are unique, consultation with the manufacturer is essential to assure proper application, detailing, installa-
tion, inspection, and performance.
Coatings and Water Repellents
The use of external coatings on brick masonry should be considered only after completing repair and replace-
ment of brick, mortar joints and other building elements, and careful consideration of the possible consequences.
Properly designed and constructed brickwork can be expected to satisfactorily resist water penetration without the
application of water repellents or external coatings. However, they may be used successfully to correct some defi-
ciencies. For example, some coatings are helpful in reducing the amount of water absorbed by barrier walls and
masonry subject to extreme exposures such as chimneys, parapets, copings and sills.
External coatings are most effective in reducing water penetration when their intended use corresponds with the
nature of the existing water penetration problem. Water repellents and coatings should not be considered equiva-
lent to essential, code-required details that resist water penetration. Use of coatings for reasons outside their
intended application rarely reduces water penetration and may lead to more serious problems.
Only water repellents that permit evaporation and the passage of water vapor, such as siloxanes and silanes,
should be used on exterior brickwork. Film-forming coating should not be applied to exterior brickwork. Technical
Notes 6 and 6A and manufacturers literature should be consulted before any coating is applied to brickwork.
SUMMARY
This Technical Note has presented maintenance procedures for brick masonry. Routine inspection of the build-
ing is suggested to determine the condition of the brickwork and related materials. If distress is noted, appropri-
ate maintenance tasks should be performed. If the problem is moisture related, the source of moisture should be
determined and corrected before other repairs are initiated.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
REFERENCES
1. Brick Brief, Ivy on Brickwork", Brick Industry Association, Reston, VA, J uly 2005.
2. Brick Brief, Repointing (Tuck-Pointing) Brick Masonry", Brick Industry Association, Reston, VA, J uly 2005.
3. Grimmer, Anne E., A Glossary of Historic Masonry Deterioration Problems and Preservation Treatments,
Department of the Interior, National Park Service, Preservation Assistance Division, Washington, D.C.,
1984.
4. Higgins, L.R., Maintenance Engineering Handbook, Fifth Edition, McGraw-Hill Publishing Company, New
York, NY, 1995.
5. Kreh, Richard T., Sr., Masonry Skills, Fifth Edition, Delmar Learning, Clifton Park, NY, 2003.
6. Old House Journal, No. 5, Vol. XIII, Brooklyn, NY, J une 1985.
7. Preservation Briefs No. 2, Repointing Mortar J oints in Historic Masonry Buildings, Heritage Preservation
Services, U.S. Department of the Interior, Washington, D.C., October 1998.
Condensation - Prevention and Control
Abstract: This Technical Note describes a variety of conditions that can cause condensation to occur in brick walls and ana-
lytical tools used to determine the likelihood of occurrence. Use of air barriers and vapor retarders to control condensation is
discussed.
Key Words: air barrier, condensation, dew point, humidity, permeance, relative humidity, saturated vapor pressure, saturation
temperature, vapor pressure, vapor resistance, vapor retarder.
SUMMARY OF RECOMMENDATIONS:
TECHNICAL NOTES on Brick Construction
47
June
2006
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
Consider the possibility and location of condensation in
building design
Reduce the possibility of condensation by:
- Providing adequate ventilation/dehumidification to reduce
humidity in the building
- Adding insulation and minimizing thermal bridges to
increase heat resistance of the wall
- Add an air space to increase the drying potential of the
wall
- Add an air barrier or vapor retarder at the appropriate loca-
tion when analysis indicate the probability for condensation
Page 1 of 6
INTRODUCTION
Moisture within masonry walls may enter during construction, in service as wind-driven rain, through poor detailing
or maintenance, or result from condensation of water vapor. Air movement within walls and differences in humidity
between inside and outside air can move water vapor to areas where condensation can occur, resulting in wetting
of wall elements.
When moisture and air movement are considered during the design phase, walls can be designed and construct-
ed to prevent detrimental condensation. Condensation tends to form on nonporous surfaces. Materials such as
brick may absorb water and not show evidence of water droplets formed by condensation. The presence of this
moisture within the brick masonry wythe can damage the brick when freezing and thawing occurs. In most cases,
the brick wythe dries out with no damage caused. However, other wall elements can be damaged by this mois-
ture, posing structural, health and aesthetic concerns.
CONDENSATION BASICS
Dew Point
Air is a mixture of gases and water vapor. At a given temperature and pressure, there is a limited amount of water
vapor that air can hold. When this limit is reached, air is considered saturated. At a constant pressure, the amount
of water vapor that air can hold increases with temperature. If saturated air at a temperature of 50 F (10 C) is
warmed to a temperature of 70 F (21 C), the mixture is no longer saturated and can absorb additional water
vapor. However, at a constant pressure, unsaturated air can be cooled to a temperature at which the air is satu-
rated. This temperature is called the dew point and, if the mixture is cooled below the dew point, water vapor will
condense, returning to liquid water. When water vapor moves through a wall system, condensation can result if
water vapor reaches the dew point. This liquid water can cause deterioration or greatly reduce the performance of
some materials. Dew is one of the most common examples of condensation.
Relative humidity is the common term used to describe the ratio of the amount of water vapor air contains to the
amount it can hold at saturation at a given temperature. For air with a fixed amount of water vapor, the relative
humidity increases as the temperature is lowered and decreases as the temperature rises. Air which is saturated
has a relative humidity of 100 percent.
In Figure 1, the difference in temperature between the air and the dew point is shown for relative humidities from
50 percent to 100 percent. The figure illustrates that for relative humidities above 80 percent, a drop in tempera-
ture of 6.8 F (3.7 C) or more will cause condensation. These high humidity conditions usually occur during the
www.gobrick.com | Brick Industry Association | TN 47 | Condensation - Prevention and Control | Page 2 of 6
summer. If the temperature difference between the
inside and outside surfaces of a wall exceed the tem-
perature drop in Figure 1, there is a possibility of con-
densation occurring on or within the wall.
Vapor Pressure
The concentration of water vapor may also be stated
by giving its pressure. Part of the atmospheric pres-
sure of air is maintained by water vapor and the
remainder of the pressure by the other constituents of
the atmosphere. Water vapor pressure is independent
of the other gases in the air. The vapor pressure when
air is saturated, as described above under Dew Point,
is defined as the saturated vapor pressure for that tem-
perature. Data on saturated vapor pressures are listed
in tables of the ASHRAE Handbook of Fundamentals
[Ref. 1] and Table 1 gives saturated vapor pressures
for various temperatures as included in the 2005 edi-
tion.
The ratio of the actual pressure of the water vapor to
the saturation pressure of the water vapor for the par-
ticular temperature is also termed relative humidity.
When quantified, the value is essentially the same as
that given previously.
Per Daltons Law of Partial Pressures, water vapor from a higher pressure zone will move to a lower pressure
zone. In other words, when a vapor-pressure differential exists, water vapor will move toward the lower pressure
independently of air. Differences in vapor pressure through different parts of the wall from inside to outside distrib-
ute themselves in proportion to the vapor resistance of the respective parts. When vapor passes through pores
of homogeneous walls that are warm on one side and cold on the other, it may reach its dew point and condense
into water within the wall; but, if the flow of vapor is impeded by a vapor-resistant material on the warm side of the
wall, the vapor cannot reach that point in the wall at which the temperature is low enough to cause condensation.
Permeance
The water vapor transmission coefficient for a material is expressed as permeance. Permeance is measured in
perms. (1 perm =1 grain of water passing through 1 ft
2
of wall in 1 hour under a vapor pressure differential of 1
inch of mercury). A perm-inch is the permeance of 1-inch thickness of a homogeneous material. The reciprocal of
permeance is called vapor resistance. It is analogous to thermal resistance in that differences in vapor pressure
through each part of the wall are proportional to the vapor resistance of these parts.
Temperature Gradient
Thermal resistance, or R-value, describes the steady-state resistance to heat flow. Walls with a higher R-value tend
to have better insulating ability. The temperature gradient or temperature differential across a wall is directly propor-
tional to the thermal resistance of individual elements. For example, if the total wall R-value is 8.0 hrft
2
F/Btu
(1.4 m
2
K/W) and the resistance of one element is 2.0 hrft
2
F/Btu (0.35 m
2
K/W), the temperature drop across this
element under steady-state conditions will be 2/8 of the air temperature difference across the wall.
INFLUENCERS ON CONDENSATION
Air Leakage and Air Barriers
Air leakage has the potential to carry significantly greater amounts of water vapor than does vapor diffusion, and
consequently tends to have a greater impact on the condensation potential. Air leakage through building enve-
lopes can occur through doors and windows, at the sill plate, through electrical and plumbing penetrations, and
through the walls themselves. The building envelope should be as airtight as possible. Not all cracks and open-
ings can be sealed, so the designer should assume that some water vapor flows into, or through, the wall assem-
bly and provide a means for it to exit.
Figure 1
Temperature Drop Curve for Relative Humidity
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
40 50 60 70 80 90 100
Relative Humidity in Percent
T
e
m
p
e
r
a
t
u
r
e

D
r
o
p

(
o
F
)

T
o

D
e
w

P
o
i
n
t
www.gobrick.com | Brick Industry Association | TN 47 | Condensation - Prevention and Control | Page 3 of 6
In some cases, operating the building's mechanical system to provide a slight
positive or negative pressure can help control air flow through the envelope.
For example, during the cooling season, a slight positive indoor pressure (rela-
tive to the outdoor air pressure) helps prevent humid outdoor air from being
drawn into the building.
Air barriers are membranes made of polyethylene, polypropylene or polyolefin
and are intended to prevent air leakage through the building envelope, hence
reducing the associated energy losses and moisture movement. Most air bar-
riers allow water vapor transmission, while some also act as vapor retarders.
Manufacturers may provide data based on different standards, including 1)
ASTM D 726, Test Method for Resistance of Nonporous Paper to Passage
of Air and 2) ASTM E 283, Test Method for Determining Rate of Air Leakage
Through Exterior Windows, Curtain Walls, and Doors Under Specified
Pressure Differences Across the Specimen. For this reason, caution should be
exercised when evaluating and specifying air barriers. In addition, convention-
al wall elements may act as air barriers. For example, when joints are caulked
or taped, either sheathing or gypsum board can serve as an air barrier.
Vapor Diffusion and Vapor Retarders
As stated previously, water vapor will naturally diffuse from an area of higher vapor pressure to one of lower vapor
pressure. When a vapor pressure differential occurs across a wall section, the water vapor diffuses through the
wall elements in proportion to each elements vapor permeance. A vapor retarder is a material with low vapor per-
meability, intended to interrupt water vapor diffusion.
Vapor retarders are often made of materials similar to air barriers. While some air barriers will also inhibit vapor
transfer, all vapor retarders can be air barriers if they are installed continuously and thoroughly sealed, with no
tears, gaps or holes. Some manufacturers cite test data based on ASTM E 96, Test Methods for Water Vapor
Transmission of Materials. However, this test does not account for fastener penetrations, electrical outlets, or joints
in the retarder. Materials that qualify as vapor retarders have a perm rating of one or less.
Thermal Bridging
Thermal bridging occurs when materials with a low thermal resistance penetrate materials with higher thermal
resistance. Heat flowing through a wall will take the path of least resistance, causing the heat to funnel through
areas of low resistance. An example of this phenomenon occurs in a steel stud wall where the R-value of the batt
insulation is on the order of 100 times that of the steel. Because the insulation is placed between the studs, the
steel then acts as a bridge through the insulation, allowing the heat to bypass the insulation. Thermal bridging, in
this example, can lead to local cold spots at the steel stud locations, resulting in condensation in those areas even
though analysis of the wall as a whole may not indicate a high potential for condensation. Continuous insulation or
insulated sheathing on the exterior of the studs helps break the thermal bridge and keeps the steel studs at a tem-
perature closer to that of the building interior, reducing the chances of local condensation. It is also important that
all walls are completely insulated without gaps that could result in thermal bridging.
REDUCING CONDENSATION POTENTIAL
The possibility of condensation can be reduced by:
- Lowering the humidity of the air if the high humidity is inside the building. This may be accomplished by
adequate ventilation if the high humidity is caused by conditions inside the building.
- Increasing the temperature of the surface on which the condensation occurs by increasing the interior
temperature, increasing the heat resistance of the wall or minimizing thermal bridging in the construction.
This is usually accomplished by adding insulation behind the interior finish.
- Promoting drying by increasing the movement of air over the surface on which condensation is likely to
occur.
- Adding an air space to increase the walls drying potential; this is a good redundant measure to help the
wall system recover from the presence of moisture.
Fahr enheit
Temper at ur e
Vapor Pr essur e,
in. Hg
-10 0.0220
0 0.0376
10 0.0629
15 0.0806
20 0.1027
24 0.1243
30 0.1645
40 0.2477
50 0.3624
60 0.5216
70 0.7392
75 0.8750
80 1.0323
90 1.4219
100 1.9333
Table 1
Saturated Vapor Pressures
www.gobrick.com | Brick Industry Association | TN 47 | Condensation - Prevention and Control | Page 4 of 6
- Installing an air barrier or vapor retarder in locations appropriate to the climate and building use where
analysis indicates a probability of condensation.
CONDENSATION ANALYSIS TOOLS
Traditionally, hand-calculation procedures, such as the steady-state dew point analysis, have been used to
estimate the potential for condensation within a wall. The dew point analysis looks at temperatures and vapor
pressures across a wall section at a set of interior and exterior climate conditions. Currently, however, computer
programs are widely available for condensation analysis. Contrary to the dew-point analysis, computer aided
analysis tools have the advantage of considering dynamic effects. These programs use up to a full year of specific
weather-related input for temperature ranges, humidity and rainfall. They also include material property databases
for common construction materials, as well as the ability to add custom material data. One such program, that is
publicly available via the internet, is WUFI [Ref. 3].
WUFI-ORNL/IBP Research and Education version for USA and Canada is publicly available at the following URL:
www.ornl.gov/ORNL/BTC/moisture [Ref. 3]. Proper application of WUFI-ORNL/IBP requires experience in the field
of hygrothermics and basic knowledge of numerical calculation methods.
WUFI-ORNL/IBP is a menu-driven program which allows realistic calculation of one-dimensional heat and mois-
ture transport in walls. It is based on the latest findings regarding vapor diffusion and liquid transport in building
materials and has been validated by detailed comparison with measurements obtained in the laboratory and in
outdoor tests.
WUFI provides the user with a wide range of output options for each user-defined monitoring position, including
temperature, water content in each construction layer, heat fluxes and profiles across the wall section of tempera-
ture, relative humidity and water content at discrete points in time.
4 in. (100 mm)
Brick
2 in. (50 mm)
Air Space
Water-Resistant
Barrier on
Extruded
Polystyrene
Insulation
6 in. (150 mm)
Batt Insulation
Between Studs
in. (13 mm)
Wallboard
Monitor Position 3
(See Figure 5)
4 in. (100 mm)
Brick
2 in. (50 mm)
Air Space
Water-Resistant
Barrier on
in. (12 mm)
Fiberglass
Sheathing
6 in. (150 mm)
Batt Insulation
Between Studs
in. (13 mm)
Wallboard
Monitor Position 3
at Vapor Retarder
(See Figure 4)
Figure 3
Example Wall Section - San Antonio
Figure 2
Example Wall Section - Minneapolis
www.gobrick.com | Brick Industry Association | TN 47 | Condensation - Prevention and Control | Page 5 of 6
Because WUFI does not include the effects of air movement on heat and moisture transfer, the program does not
adequately evaluate the drying capabilities of an air space in the wall system. The presence of an air space can
reduce the moisture content of adjacent wall components. In addition, since WUFI is a one-dimensional tool, the
potential effects of thermal bridges are not accounted for.
Example
Brick veneer steel stud wall systems, such as shown in Figures 2 and 3, were evaluated using WUFI in two cli-
mates: Minneapolis, Minnesota and San Antonio, Texas. Each was evaluated for a full year, and example results
are shown in Figures 4 and 5. These figures show relative humidity at specific positions within the wall over the
course of a year. In Minneapolis, the wall includes a polyethylene vapor retarder between the batt insulation and
the interior gypsum wallboard. In cold weather, this prevents interior moisture from entering the wall system and
condensing within the wall. In San Antonio, this interior vapor retarder is replaced with extruded polystyrene insu-
lation on the exterior of the steel studs. The joints are sealed to intercept exterior moisture and prevent it from
condensing within the wall as it moves closer to the cooler, dryer interior under summer cooling conditions. For the
Minneapolis wall section Monitoring Position 3 is at the the batt insulation/vapor retarder interface, and for the San
Antonio wall section Monitoring Position 3 is at the extruded polystyrene insulation/batt insulation interface.
It is important to note that condensation occurs when relative humidity reaches 100 percent. Figures 4 and 5 show
that the wall systems in Figures 2 and 3 do not reach the level where condensation will occur. Therefore, the
example wall sections are appropriate for the given locations.
SUMMARY
Proper design of brick masonry walls must consider the potential for condensation. This includes reducing the
influences of air leakage, vapor diffusion and thermal bridging. Prevention and control of condensation can be
achieved through proper design, which can be aided by contemporary analysis tools, as well as material selection
and placement. Incorporating these tools and methods into design will add to the durability of brick and longevity
of the building.
The information and suggestions contained in this Technical Note are based on the available data
and the combined experience of engineering staff and members of the Brick Industry Association.
The information contained herein must be used in conjunction with good technical judgment
and a basic understanding of the properties of brick masonry. Final decisions on the use of
the information contained in this Technical Note are not within the purview of the Brick Industry
Association and must rest with the project architect, engineer and owner.
0
20
40
60
80
0 60.8 121.6 182.4 243.2 304.0 364.8
Time [days]
R
e
l
a
t
i
v
e

H
u
m
i
d
i
t
y

[
%
]
Monitorpos. 3
Monitorpos. 4 (Interior Surface)
Figure 4
Relative Humidity Profile of Brick Veneer/Steel Stud
Wall in Minneapolis, Minnesota
0
25
50
75
100
0 60.8 121.6 182.4 243.2 304.0 364.8
Time [days]
R
e
l
a
t
i
v
e

H
u
m
i
d
i
t
y

[
%
]
Monitorpos. 3
Monitorpos. 4 (Interior Surface)
Figure 5
Relative Humidity Profile of Brick Veneer/Steel
Stud Wall in San Antonio, Texas
Monitorpos. 3
Monitorpos. 4 (Interior Surface)
Monitorpos. 3
Monitorpos. 4 (Interior Surface)
REFERENCES
1. Handbook of Fundamentals, American Society of Heating, Refrigeration and Air Conditioning Engineers,
Inc., Atlanta, GA, 2005 Edition.
2. Masonry Structures: Behavior and Design, Second Edition, The Masonry Society, Boulder, CO, 1999.
3. WUFI-ORNL/IBP, Oak Ridge National Laboratory, ORNLs Buildings Technology Center, Oak Ridge, TN,
2005.
www.gobrick.com | Brick Industry Association | TN 47 | Condensation - Prevention and Control | Page 6 of 6
2009 Brick Industry Association, Reston, Virginia Page 1 of 14
TECHNICAL NOTES on Brick Construction
1850 Centennial Park Drive, Reston, Virginia 20191 | www.gobrick.com | 703-620-0010
48
June
2009
Sustainability and Brick
Abstract: This Technical Note discusses sustainability and sustainable design and their relationship to brick manufacturing,
use and recycling. Sustainable practices in manufacturing are identified, as are ways to utilize brick as part of sustainable
design strategies. Ways that brickwork can be used to earn points in the LEED and other green building rating systems also
are identified.
Key Words: environmental impacts, green building, LEED, life cycle assessment (LCA), recycled content, resources,
sustainability, sustainable design, thermal mass, volatile organic compounds (VOCs).
Sustainable design balances environmental, economic
and societal goals. It is more than just a certification from
a rating system
Brick is made from abundant natural resources (clay
and shale) and is readily recycled for use in the
manufacturing process or other uses
Brick manufacturers address sustainability by locating
plants in close proximity to mines; incorporating waste
products and recycled materials into brick; reducing
energy use, water use and atmospheric emissions; and
utilizing landfill gas and other wastes for fuel
Brickwork used in sustainable designs can provide
structure, finish, acoustic comfort, thermal comfort, good
indoor air quality, fire resistance, impact resistance and
durability, all in one product
Brickwork can help meet requirements of many
certification rating systems in the areas of development
density, storm water management, heat island
effect, improved energy performance, building reuse,
construction waste management, materials reuse,
recycled content and regional materials
Brickwork can contribute to improved indoor air quality by
eliminating the need for paints and the resulting volatile
organic compounds (VOCs) and by eliminating a food
source for mold
Brickwork is durable, having a life expectancy of
hundreds of years. Brick buildings can be and are
reused, thereby distributing their environmental impact
over an extended life span
INTRODUCTION
Sustainable design is a term that has entered the vernacular of building design and construction. As more
buildings are designed and constructed using sustainable design principles, the need for information on building
products and their sustainable design features also grows. In assessing the sustainable attributes of building
products, consideration must be given to how the product is manufactured, used and disposed of. This Technical
Note provides information on brick usage, manufacturing and recycling as it relates to sustainability.
WHAT IS SUSTAINABILITY?
Sustainability is defined as meeting the needs of the present without compromising the ability of future
generations to meet their own needs [Ref. 2]. Sustainable buildings are designed in a way that uses available
resources efficiently and in a responsible manner, balancing environmental, societal and economic impacts to
meet the design intents of today while considering future effects. Sustainably designed buildings are energy-
efficient, water-efficient and resource-efficient. They address the well-being of the occupants by considering
thermal comfort, acoustics, indoor air quality and visual comfort in the design. They also consider the impact of
a buildings construction, operation and maintenance on the environment, and the environmental impact of the
buildings constituent materials. A sustainably designed building considers all of these aspects through the entire
life cycle of the building, including its operation and maintenance.
Often the tendency is to focus on one aspect of sustainable design, such as energy use or environmental impacts.
This approach leaves out other equally important elements necessary for true sustainability. Truly sustainable
design is best described as achieving the triple bottom line, that balance of environmental goals, societal goals
and economic goals.
A high-performance, sustainable building design should include accessibility, aesthetics, cost-effectiveness,
durability, functionality/operation, productivity of occupants, security and safety, and environmental performance.
SUMMARY OF RECOMMENDATIONS:
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 2 of 14
THE NEED FOR SUSTAINABLE BUILDING DESIGN
Each year building construction accounts for 25 percent of the virgin
wood used globally, and 40 percent of the raw stone, gravel and sand
used. Buildings also use 48 percent of the energy and 16 percent
of the water used annually worldwide (see Figure 1). In the United
States, about as much construction and demolition waste is produced
as municipal garbage [Ref. 14].
Architect Edward Mazria, who founded Architecture 2030, a nonprofit
organization dedicated to research and solutions in architecture to
address global climate change, determined that Combining the
annual energy required to operate residential, commercial, and
industrial buildings along with the embodied energy of industry-
produced building materials like carpet, tile, glass, and concrete,
exposes buildings as the largest energy consuming and greenhouse
gas emitting sector in the United States [Ref. 12].
This focus on energy use of buildings is one of many aspects of evaluating sustainable building design that has
multiple impacts. Reduction in building energy use has a positive impact on the environment because most energy
is generated by processes that negatively impact the environment. A reduction in building energy use has positive
economic impacts for building owners and occupants. Because any negative impact on the environment concerns
all of society, a reduction in energy use also has positive societal impacts. This limited example illustrates the
interrelated nature of sustainable design.
Another aspect to be considered in sustainable design is the indoor environment. Improvements in this area can
increase productivity and provide an economic benefitsalaries are the largest cost of occupied buildingsas well
as a societal health benefit. The environmental impacts of the building location and construction process, as well
as the environmental impacts of the manufacture and use of building materials, also must be considered.
There are as many ways to achieve this triple bottom line as there are design possibilities. Understanding
general sustainable design issues helps identify areas in which design has an impact.
SUSTAINABLE DESIGN WITH BRICK
While there is general agreement on many of the elements of sustainable building design, defining and measuring
it poses a challenge. The first step is to define sustainable design for buildings. As discussed earlier, to be
sustainable a design must consider more than just environmental impacts. A holistic approach is necessary. The
Whole Building Design Guide developed by the National Institute of Building Sciences discusses sustainable
design in terms of whole building design. In their words, Whole Building design in practice requires an integrated
team process in which the design team and all affected stakeholders work together throughout the project phases
and to evaluate the design for cost, quality-of-life, future flexibility, efficiency; overall environmental impact;
productivity, creativity; and how the occupants will be enlivened. The Whole Buildings process draws from the
knowledge pool of all the stakeholders across the life cycle of the project, from defining the need for a building,
through planning, design, construction, building occupancy, and operations [Ref. 15].
Sustainable Design Elements
Every sustainable building is unique, designed specifically for its site and the programming requirements of the user.
However, all high-performance, sustainable buildings should consider the following components of design [Ref. 14]:
The versatility and durability of brick facilitate its use as part of many elements of sustainable design.
Figure 1
U.S. Energy Consumption [Ref. 12]
Environmentally responsive site planning
Energy-efficient building shell
Thermal comfort
Energy analysis
Renewable energy
Water efficiency
Safety and security
Daylighting
Commissioning
Environmentally preferable materials and products
Durability
Efficient use of materials
High-performance HVAC
High-performance electric lighting
Life cycle cost analysis
Acoustic comfort
Superior indoor air quality
Visual comfort
Transportation
27%
Industry
25%
Buildings
48% Industry
25%
Buildings
48%
Transportation
27%
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 3 of 14
Environmentally Responsive Site Planning
Environmentally responsive site planning includes consideration of site selection, site disturbance, storm water
management and effect of the building on its surroundings. The use of brick masonry is an appropriate choice for
achieving several elements of environmentally responsive site planning.
Reuse and Renovation. The first step in site planning is selection of the building site. Reuse or renovation of an
existing building (see Photo 1) can result in significant reductions in environmental impacts as compared with new
construction. Because of aesthetic appeal and durability, brick masonry buildings often are chosen for reuse. In
many cases, load-bearing brick buildings are reused in their entirety. In other cases, the brick faade is retained
while a new structure is constructed.
Using Brick in Urban Areas. When locating new construction, it is desirable to select sites near existing
infrastructure. Utilizing brick masonry in urban development can help meet requirements for fire resistance and
limitations on site accessibility and can also accommodate irregularly shaped lots.
Maximizing Open Space. On any site it is desirable to maximize the amount of open space on the site, either by
limiting the building footprint or by minimizing the extent of site disturbance adjacent to the building. Because brick
masonry construction does not require large staging areas or large equipment for placement, the amount of site
disturbed can be kept to a minimum. In addition, brick paving in an open space can provide a pedestrian-friendly
surface.
Managing Storm Water Runoff. By managing storm water runoff, increasing on-site filtration and eliminating
contaminants, the disruption and pollution of natural water flows are limited. Flexible brick pavements can be
designed as permeable pavements to allow percolation of storm water through the pavements, thereby reducing
runoff, recharging groundwater aquifers and removing contaminants from surface water (see Photo 2).
Reducing the Heat Island Effect. Building projects have an effect on their surroundings, particularly in urban
areas. The use of light-colored materials can help reduce the heat island effect. Light-colored brick pavers can
be used on vegetated roofs to provide access paths or on non-roof pavements as part of a strategy to reduce
this effect.
Energy-Efficient Building Shell,
Thermal Comfort and Energy Analysis
Energy-Efficient Building Envelope. An energy-efficient building envelope is a key component in sustainable
building design. Incorporation of brick masonrys thermal mass provides numerous energy benefits, including the
reduction of peak heating and cooling loads, moderation of indoor temperature swings (improved thermal comfort),
and potential reduction in the size of the HVAC system. The benefits of thermal mass have been demonstrated
when brick is used as a veneer, and are even more pronounced when brick masonry also is exposed on the
interior of the building.
Photo 1
Reused (left) and New Brick Buildings
Photo 2
Permeable Clay Pavement
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 4 of 14
Energy Analysis. In order to thoroughly account for the thermal mass benefits of masonry, energy analysis using
simulation software is necessary. BLAST [Ref. 4] or EnergyPlus [Ref. 5] are the most suited to analysis of buildings
with masonry.
Brick Rain Screen Wall. A brick rain screen wall is another example of a high-performance brick wall. Moisture
penetration is one of the most common causes of problems in buildings. Rain screen walls minimize rain infiltration
by applying principles of drainage and pressure equalization. A brick masonry pressure-equalized rain screen wall
utilizes intentional openings in the brick masonry and compartmentation of the cavity to equalize the pressure in
the cavity behind the exterior brick and thus minimize rain penetration. Details of the design of pressure-equalized
rain screen brick masonry walls can be found in Technical Note 27.
Safety and Security
Brick masonry promotes occupant health and safety through fire-resistant construction and resistance to impacts
and wind-borne debris. In addition, the durability of brick masonry gives long-lasting results.
Life Cycle Cost Analysis
Due to durability of brickwork construction, a life cycle cost analysis can often demonstrate the long-term benefits
of building with brick.
Acoustic Comfort
Acoustic comfort is a key element in a superior indoor environment. Brick masonry walls provide superior
resistance to sound penetration with a sound transmission class (STC) of 45 or greater.
Renewable Energy
Incorporation of renewable energy sources into a building design can significantly reduce reliance on fossil fuels
used by the building during operation. Passive solar energy is a free resource, and brick masonry can be utilized
as part of several passive solar design strategies. Brick paving can be used in interior applications to store heat
and moderate temperature swings.
Environmentally Preferable Materials and Products
Life Cycle Assessment. Consideration of the environmental impact of building materials and products is an
important element in a sustainable design, though it is only one of several criteria to be considered for product
selection. Materials should be evaluated over their entire life cycle, from raw material extraction to end of useful
life. This life cycle assessment (LCA) of a building material or product must include accurate evaluation of product
service life.
Construction Waste. Building construction can generate significant amounts of waste. Because of the small,
modular nature of brick, on-site construction waste can be dramatically reduced through careful design and
detailing. In addition, scrap brick is easily crushed and recycled for new uses, thus avoiding the landfill. Packaging
from brick is minimal and easily recycled.
Salvaged Materials. Use of salvaged materials avoids the environmental impacts associated with new products.
Salvaged brick, especially sand-set units, can be reused when care is taken to determine material performance
characteristics.
Recycled Content. Many environmentally preferred product listings focus on materials that incorporate recycled
content. By utilizing recycled materials, the assumption is that the environmental impact is lowered. Recycled
materials can come from either postconsumer or postindustrial (preconsumer) sources. Brick masonry can contain
many recycled products. Brick units may incorporate recycled materials such as sawdust and metallic oxides.
Mortar and grout can include recycled materials, such as fly ash, and most steel reinforcement used in reinforced
brick masonry has a high recycled content.
Regional Sources. By selecting materials from regional sources, environmental impacts associated with the
transport of materials can be reduced. Most brick are manufactured from materials obtained from within a few
miles of the manufacturing plant. Of the 50 largest metropolitan statistical areas (MSAs) in the United States,
there are more than 25 plants on average within 500 miles of each, and there are at least two brick plants within
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 5 of 14
500 miles of 49. More than 70 percent of these
MSAs have at least one plant within 200 miles (see
Figure 2).
Durability and Designing for
Longer Life Expectancy
Designing a high-performance wall considers not only
the multiple functions a wall can perform, but also
the different life expectancies of those elements that
make up the wall. It is important when designing
systems with many components, such as curtain
walls, endeavor to make all components equally
durable to prevent premature failure of one part of the
system. When possible, provide for easy inspection,
maintenance, and/or replacement of less durable
components [Ref. 13].
This approach is especially important in achieving a
sustainable building design. Brickwork is extremely
durable, having a life expectancy of hundreds of years.
Repointing may be required only every 50 years or
more. Thus it is important to recognize this fact when
detailing those elements that interface directly with
brick masonry that have shorter life expectancies or
require more frequent maintenance. One example
of this is flashing. Some flashing materials, such
as stainless steel and copper sheet, have been documented to have a life expectancy of more than 100 years.
The detail shown in Figure 3 is designed with service life in mind. The metal cap and roof membrane will require
Metal Cap
Anchored to P.T.
Wood and Backing
Compressible
Material
Wood Anchored
to Backing Only
Metal Panel
System
Roofing
Membrane
Air Space
Water-Resistive
Barrier from
Wall Below
Figure 3
Detailing for Longer Life Expectancy
Figure 2
Brick Plant Proximity to 50 Largest Metropolitan Statistical Areas
500 mi. from
N
e
w
Y
o
r
k
,

N
Y
500 mi. from
Lo
s
A
n
g
e
le
s
,

C
A
5
0
0

m
i
.

f
r
o
m

S
e
a
t
tle
, W
A
5
0
0

m
i. fro
m
D
allas, TX
500 mi. from
C
h
ic
a
g
o
, IL
500 m
i. fro
m
M
ia
m
i,

F
L
5
0
0
m
i. from Denver, C
O
Metropolitan Statistical Area
Brick Plant Locations
Multiple Brick Plants, Same Location
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 6 of 14
periodic replacement. By utilizing separate components, they can be replaced without damage to or repointing of
the brick masonry wall. Such a detail is not required by the building code, but can extend the service life of a wall.
Efficient Use of Materials
Multiple Functions. How a building material is used also should be considered when examining the sustainability
of a material. Brick masonry walls are able to perform multiple functions that often require several components in
other wall systems. By designing walls that serve multiple functions, materials are used efficiently. This translates
into reduced environmental impacts for the building. A single brick wythe can do all of the following:
Serve as a load-bearing structural element.
Provide an interior or exterior finish without the need for paints or coatings.
Provide acoustic comfort with a sound transmission class (STC) rating of 45 or greater.
Regulate indoor temperatures as a result of thermal mass.
Provide fire resistance (a nominal 4-in. (100 mm) brick wall has a one-hour fire rating).
Provide impact resistance from wind-borne debris or projectiles.
Improve indoor air quality by eliminating the need for paint and coatings (no VOCs).
Provide a non-combustible material which does not emit toxic fumes in fires.
Provide an inorganic wall that is not a food source for mold.
Serve as a heat-storing element in a passive solar design.
Last for generations.
In addition, other innovations in brick masonry design can further decrease the raw materials used. The use of
prestressed brick walls capitalizes on the inherent compressive strength of brickwork, resulting in typically thinner,
taller walls.
Reduced Thickness. Brick manufactured to a reduced thickness use less raw material. Brick manufactured
to meet ASTM C1088, Specification for Thin Brick Veneer Units Made from Clay or Shale, have a maximum
thickness of 1 in. (44 mm). When adhered to a load-bearing wall substrate, these units can provide a brick finish
with minimal thickness. For more information on thin brick construction, refer to Technical Note 28C.
Anchored masonry veneer can be as little as 2 in. (67 mm) thick, according to the Building Code Requirements
for Masonry Structures, which is referenced by the International Building Code. Such brick are available from most
brick manufacturers and most often are used in residential applications.
Hollow Brick. Brick units meeting the requirements of ASTM C652, Specification for Hollow Brick, utilize less raw
material than other brick used in anchored veneers while performing the same function. In addition, these units can
be used in reinforced or prestressed brick masonry walls.
Superior Indoor Air Quality
Avoiding Volatile Organic Compounds. Because brick masonry can be used on the interior of a building,
serving as structure and finish material without the need for paints or coatings, brick can contribute to improved
indoor air by avoiding volatile organic compounds (VOCs). In addition, because the appearance of brick will last a
lifetime without costly paints or other maintenance, this benefit continues for the life of the building.
Resisting Mold. Mold is another area of concern for indoor air quality. Brick masonry is not a food source for
mold. As a result, it does not promote mold growth, even if wetted, and is easily cleaned if needed. Other interior
wall materials can be literally eaten up by mold if moisture problems occur.
In addition, interior brick paving can be used in lieu of carpeting, particularly in high-traffic areas, thereby reducing
indoor VOC content associated with carpet and adhesives and eliminating the need for regular replacement of
flooring.
BRICK MANUFACTURING AND SUSTAINABILITY
In order to understand how brick can contribute to sustainable building design, it is important to consider how brick
is made, as well as how it is used. Brick manufacturing is a highly efficient process incorporating many sustainable
practices as described below. Technical Note 9 discusses the manufacture of brick in detail.
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 7 of 14
Raw Material Use
Brick is made primarily from clay and shale, which are abundant natural resources [Ref. 6]. Most of the clays and
shales used in brick making are mined in open pits located near brick manufacturing facilities many of which
are less than a mile away. Most plants use material from the same pit extracted through multiple soil layers for a
minimum of 50 years, thus minimizing their impact to the surface area. Conveyors and other power equipment
typically are used to transport the clay from the mine to the plant. Brick manufacturing plants are located
throughout the country, putting them within a short distance of most urban areas.
Storm water runoff from clay pits is controlled by regulations from the Mining Safety and Health Act. Manufacturers
use techniques such as settling ponds, filtration through marshes and wetlands and catch basins. Dust is
controlled by spraying organic, biodegradable oils or water.
Once the clay is mined, it is ground to suitable particle size and then mixed with water. This mixture is then formed
into brick. Non-hazardous waste products are sometimes incorporated into the mixture. For example, petroleum-
contaminated soil or sludge can be used. Recycled waste from other industries, such as bottom ash and fly ash
from coal-fired generators, glass, stone dust, and ceramic tile may be incorporated. Reclaimed industrial metallic
oxides can be used as colorants in brick. Because fired brick are inert, brick can safely encapsulate many materials.
Additionally, nearly all of the mined, raw materials result in finished brick product. Nearly 95 percent of all the
mined clay and shale goes to the plant, and an average of 3 percent of the manufactured product ends up as
scrap most of which is returned to the manufacturing process or recycled for secondary uses such as structural
fill.
Sustainable Practices in Manufacturing
One of the results of todays manufacturing process is that brick manufacturing is more energy efficient now than it
was a few short decades ago. In the 1970s, a standard brick required 14,000 Btu of energy to mine, manufacture
and transport [Ref. 6]. Today, the average embodied energy for U.S.-manufactured brick is about 4300 Btu per
standard brick [Ref. 3]. The reason for this significant change is that the contemporary brick manufacturing process
incorporates many practices intended to conserve resources and promote sustainability.
For example, the majority of brick plants use renewable materials within the brick-making process. Lubricants
made from a waste by-product derived from processing organic materials can be used in the forming of brick.
Heat required for dryer chambers usually is supplied from the exhaust heat of kilns to maximize thermal efficiency.
Water used in brick production is recycled and reused. Improvements in automation result in even less energy
being used.
Additionally, more than 35 percent of brick plants have investigated alternative energy sources. Many
manufacturers are using waste products such as methane gas from landfills and sawdust. While natural gas is the
most frequently used fuel for firing brick, utilizing waste materials enables brick plants to reduce their consumption
of fossil fuels as well as provide a beneficial means of disposal for potential wastes. Brick manufacturers also are
taking steps to further improve their efficiency by using sawdust and petroleum coke as a burnout material in the
clay or shale mixture, producing lower-weight units with less raw material. In fact, more than 80 percent of brick
manufacturers have made recent improvements to reduce the energy used by their plants. These improvements
include providing more energy-efficient kilns, forming more energy-efficient brick and installing energy-efficient
lighting.
Also, the majority of brick today are packaged in self-contained, strapped cubes, which can be broken down
into individual strapped sections for ease of handling on the jobsite. Layers of brick are separated by wooden or
paper strips to reduce chippage and breakage. Such packaging does not use wooden pallets. Plastic straps and
wood dividers can be recycled, resulting in little or no waste. Approximately 15 percent of brick are transported to
the distributors yard or jobsite by rail and 85 percent by truck, often less than 500 miles. About half of U.S. brick
manufacturing facilities accept unused brick that was shipped to a distributor or end user.
Another result of todays brick manufacturing is that the embodied energy per pound required to make brick is
much less than for other materials. Most of the energy required for brick production is used in the firing of the kilns.
The 1998 AIA Environmental Resource Guide states that the energy used to mine, manufacture and transport
brick (known as embodied energy) is approximately 4000 Btu per pound. Thats less per pound than concrete,
glass, steel or aluminum [Ref. 6]. This number has decreased since then, with most manufacturers increasing the
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 8 of 14
energy efficiency of existing plants or constructing new plants that operate more efficiently. The current industry
average is approximately 1239 Btu per pound [Ref. 3].
With todays environmental concerns, the brick industry recognizes the need for compliance with state and federal
regulations for clean air and the environment. Air emissions are minimized with controls such as scrubbers
installed on kiln exhausts. Lime waste that accumulates in scrubbers often is recycled as a beneficial additive to
soil. Dust in brick plants is controlled through the use of filtering and containment systems, vacuums, additives and
water mists. Even vehicular emissions are being addressed, with brick manufacturers monitoring truck emissions;
recycling waste oil, antifreeze and hydraulic oil; and regulating truck speeds for improved fuel efficiency [Ref. 7].
As part of a commitment to sustainability and good community citizenship, mined areas are reclaimed for future
use by replacing overburden and topsoil so the resulting property can be used for a wide variety of functions,
including farmland, residential and commercial sites, and even wetlands. In this way valuable clay and shale
resources are obtained while still allowing for land to be reused for a different purpose later.
Environmental Information
Information on the environmental impacts of brick manufacturing can be provided in several ways. The simplest of
forms is an environmental audit report. An environmental audit examines and identifies the environmental impact
of the various manufacturing processes in a qualitative manner. Exact quantification of effects is not determined. A
more thorough examination of the manufacturing process is conducted in an environmental life cycle assessment
(LCA). In an LCA, quantitative data are collected on the effects of the manufacturing process in a prescribed
number of assessment areas, which are then attributed to corresponding environmental impacts such as global
warming potential, stratospheric ozone depletion and eutrophication (the gradual increase in the concentration of
plant nutrients in an aging aquatic ecosystem). The boundaries of a typical building product LCA as conducted by
a manufacturer consider all impacts during the manufacturing process, from resource extraction to final production,
or cradle to gate. Transport to the project site also may be included.
As sustainable building design and standardization of LCA grow, it is anticipated that more U.S. brick
manufacturers will undertake this process. It is important to recognize that environmental assessments can be
done for industries as a whole and also for individual manufacturing facilities. Environmental assessments for
individual manufacturing facilities reflect the environmental impacts of the different choices in fuel type, water
conservation, etc. for that particular plant or product.
BRICK RECYCLING
AND REUSE
Brick can be recycled in many ways. As discussed
previously, raw brick and fired brick are recycled in
the manufacturing process. Scrap brick and brick from
demolition can be crushed and recycled into new brick
or used as brick chips for landscaping (see Photo 3),
baseball diamonds and tennis courts. Recycled brick
also can be used as subbase material for pavements,
on quarry roads or even as aggregate for concrete.
Brick also can be reused. Individual brick units can be
salvaged and reused with proper precaution. Sand-set
brick pavers are the easiest to reuse because they
are easily removed and typically remain physically
unaltered. These units can be reused in a variety of
new ways: for flooring, in pavements and in other
landscape elements. Mortared units, pavers and facing units must be carefully cleaned of old mortar before reuse.
In addition, care must be taken when reusing older brick units, as they may not be as durable as those currently
manufactured. Reused brick should be tested to verify that they meet the requirements of the current specification
for their intended use. Units that will be laid in mortar must have adequate open pores to ensure proper bonding.
Perhaps the best example of bricks sustainability is the reuse of brick masonry buildings. Because of the durability
of brick masonry construction and the historic value often associated with brick buildings, reuse of brick masonry
Photo 3
Recycled Brick Chips
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 9 of 14
buildings is increasingly popular. By adapting existing structures to new uses, both resources and energy are
saved and environmental impacts are reduced. This adaptive reuse of brick masonry buildings is a testimony to
the longevity and durability of brick masonry.
OTHER MATERIALS IN BRICKWORK
Brick masonry includes not only brick, but also mortar, grout, metal anchors and ties, reinforcement, flashing, and
drainage systems. Not all projects will use all of these other materials, and generally they constitute a relatively
small percentage of brick masonry construction. However, it is important to understand how they fit in sustainable
design as well. This Technical Note serves as an introduction to sustainable issues surrounding these other
materials. Readers are encouraged to contact product manufacturers for further information on these materials.
Mortar and Grout
The majority of brick masonry is constructed using mortar. For a typical brick masonry wall built with modular size
brick laid in running bond, about 20 percent of the surface is mortar. Mortar is composed of cementitious materials,
sand and water. Cementitious materials for mortar include portland cement, blended cement, lime, masonry
cement and mortar cement, or a combination of these. Details on mortar properties and proportions are found in
Technical Note 8.
Grout, like mortar, is composed of cementitious materials and aggregate. Lime is used in much smaller quantities
in grout than in mortar, if used at all. Depending on project requirements, the aggregate may be coarse (gravel-
sized particles) or a combination of fine and coarse aggregate (sand and gravel).
Cement. Cements are high embodied energy materials. Increases in fuel efficiency and use of waste fuels have
reduced the environmental impact of cement production. Mortar and grout may incorporate fly ash, a waste
product from the burning of coal, as a partial replacement for cement. By reducing the amount of cement, the
embodied energy and environmental impact of the mortar or grout is reduced, and an industrial waste is recycled.
Hydrated Lime. Hydrated lime is also a high embodied energy material. Softer mortars that contain only lime or
high proportions of lime with a small amount of portland cement were common before the early part of the 20th
century, so they are now used for historic restoration of buildings from that era. Mortars with high lime content can
be more easily removed from brick when masonry is deconstructed and can be crushed and used as a beneficial
additive to soil.
Sand and Aggregates. Sand and other aggregates are naturally occurring materials. They are available in almost
all geographic locations and require little processing, and as such, they typically have a low embodied energy and
a small environmental impact. Sand is sometimes considered a renewable material, being constantly formed from
the erosion of rock.
Metal Anchors, Ties and Reinforcement
The majority of metal anchors, ties and reinforcement used in brick masonry construction are made from steel.
Stainless steel and other metals are sometimes used, but to a lesser extent. The quantity of metal elements found
in brick masonry is dependent on the type of construction. Brick veneer typically contains only metal anchors and
possibly metal flashing. Reinforced brick masonry often will contain reinforcing bars and wire reinforcement in
addition to ties and flashing.
Steel is made from iron ore and is a high-embodied energy material. However, steel is commonly recycled.
Increasingly, anchor and tie manufacturers are utilizing raw materials that contain recycled content. The Concrete
Reinforcing Steel Institute estimates that nearly 100 percent of feedstock for reinforcing bars is recycled scrap.
Wire joint reinforcement also may contain recycled steel. Product manufacturers should be contacted for specific
information.
ASSESSING SUSTAINABLE DESIGN
Many organizations have developed programs to measure various attributes of sustainable design. Some focus
on the entire picture of sustainable design. Others focus primarily on the environmental aspects, so-called green
building. Still others apply sustainable design metrics to a particular type of construction, such as schools. Though
these rating systems are extensive, they are still limited in scope.
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 10 of 14
Green Building Rating Systems
The Leadership in Energy and Environmental Design (LEED) Green Building Rating System, developed by
the United States Green Building Council (USGBC), is currently the most widely used green building rating
system in the United States. As its name implies, the focus of the LEED rating system is on the energy use and
environmental impact of a building. There are a number of different LEED rating systems focusing on different
types of building construction, such as commercial interiors, core and shell, homes and existing buildings. The
oldest and most widely used is the LEED for New Construction and Major Renovation Rating System (LEED-NC)
[Ref. 10], and all of the other rating systems follow the general categories of LEED-NC.
Green Globes is another national green building rating system promulgated by the Green Building Initiative. Based
originally on the British Research Establishment Environmental Assessment Method (BREEAM) rating system
developed in the United Kingdom, Green Globes was created in Canada. Introduced in the United States in 2004,
Green Globes is not as well known or as widely used as LEED-NC. Though many aspects of Green Globes are
similar to those found in LEED, Green Globes differs in that it is an online tool that allows for self-certification. As
a Web-based tool, Green Globes also provides interactive, educational links to information on sustainable design
strategies and systems.
LEED-NC and Green Globes focus on commercial construction. There are rating systems for residential
construction as well: LEED for Homes developed by USGBC, and the 2008 National Green Building Standard,
developed by the National Association of Home Builders (NAHB). These guidelines are tailored to new home
construction and contain many elements unique to that market. However, the general categories and goals of this
program are similar to those found in LEED-NC.
These and many other green building rating systems are available for use, including programs developed by states
and local jurisdictions. Though specifics of each vary, there is general agreement in categories and scope. The
matrix shown in Table 1 correlates sustainable design strategies with brick and the various rating system credits. It
is important to recognize that revisions to the rating systems are ongoing, and the credits indicated are based on
the following versions:
LEED-NC: version 3 (LEED 2009) (110 total points) [Ref. 11]
Green Globes: Green Building Assessment Protocol for Commercial Buildings (April 10, 2009, Draft)
(1000 total points) [Ref. 8]
LEED for Homes: January 2008 (136 total points) [Ref. 10]
2008 National Green Building Standard: January 2009 (approximately 2000 total points) [Ref. 1]
Specific criteria and requirements for each of the rating systems can be found in the publications referenced.
Limitations of Rating Systems
The focus of these and other green building rating systems is energy use and the environment. All contain
numerous requirements and credits intended to reduce building operational energy use, to promote the use of
building products with lower environmental impacts and to provide a healthy indoor environment. However, what
is often lacking in all of these rating systems is a means by which to promote and measure the avoidance of
negative impacts. Only one of these systems currently contains methods to measure the avoidance of construction
waste. All measure the diversion of waste from landfills, but only the NAHB effort recognizes materials that have
little or no onsite waste to begin with. Similarly, efficient use of materials is not well-recognized. Materials such as
brickwork that perform multiple functions avoid the use of other materials, such as paints, sound insulation, etc.
At present, efficiencies such as these are included only in the NAHB rating system. Only a whole-building, holistic
approach can capture the true intent of sustainable design.
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 11 of 14
LEED-NC
1
Green Globes
2
LEED for
Homes
3
2008 National
Green Building
Standard
4
Environmentally Responsive Sites
Reuse Existing Buildings. Brick masonry
buildings can be renovated and reused.
Materials &
Resources (MR)
1 3 points
Resources/
Materials
10.4.1.1,
10.4.2.1 and
10.4.3.1 up to
18 points
Site Design and
Development
(SD) 403.9
6 points, and
Resource
Efficiency (RE)
up to 12
points
Urban Development. Brick masonry is suitable
and highly adaptable to urban infill projects.
Sustainable
Sites (SS) 2
5 points
Site 7.1.1.4
2 points
or Site 7.1.2.2
3 points
Location &
Linkage (LL)
3.1 1 point or
LL 3.2 1 point
SD 401.1
4 points; and
Lot Design,
Preparation and
Development
(LD) 501.1
4 points
Location on Site. Site building to optimize solar
radiation (passive solar heating and cooling
possible).
Maintain open space (brick construction requires
minimal disruption of site).
SS 5.2 1 point
Innovation &
Design (ID) 1.5
1 point
SD 403.2
6 points
Storm Water Design. Reduce quantity and
improve quality of runoff with permeable brick
pavements.
SS 6 2 points Site 7.3.1.1
up to 10 points
SS 4.1
4 points
SD 403.5
up to 5 points
and LD 503.4
up to 5 points
Heat Island Effect. Light-colored pavers can help
reduce heat build-up.
SS 7.1 1 point Site 7.2.2.3
up to 10 points
SS 3 1 point LD 505.2
4 points
Energy Efficiency, Thermal Comfort and Energy Analysis
Improved Energy Performance. Thermal mass
of brick helps reduce heat transfer; pressure-
equalized brick rain screen walls.
Energy &
Atmosphere
(EA) 1 up to
19 points (about
2-4 points from
brick)
Energy 8.2.1.1,
8.2.1.2 and
8.2.1.3 up to
12 points
Energy &
Atmosphere
(EA) 1 up to
34 points (about
1-2 points from
brick)
Energy
Efficiency (EE)
703.1.3 up to
6 points and EE
704.3.1 up to
10 points
Thermal Comfort. Thermal mass of brick helps
reduce indoor temperature swings.
EA 1 up to 19
points (about
2-4 points from
brick)
Energy 8.2.1.1,
8.2.1.2 and
8.2.1.3 up to
12 points
EA 1 up to 34
points (about
1-2 points from
brick)
EE 703.1.3 up
to 6 points and
EE 704.3.1 up
to 10 points
Energy Analysis. Energy modeling reflects
benefits of thermal mass of brick.
EA 1 up to 19
points (about
2-4 points from
brick)
Energy 8.2.1.1,
8.2.1.2 and
8.2.1.3 up to
12 points
EA 1 up to 34
points (about
1-2 points from
brick)
EE 703.1.3 up
to 6 points and
EE 704.3.1 up
to 10 points
Environmentally Preferable Materials
Life Cycle Assessment. Innovation &
Design (ID) 1.2
1 point
Resources/
Materials
10.1.1.1
up to 33 points
RE 609.1
3 points per
product
1. LEED-NC version 3 (LEED 2009) Total possible points = 110
2. Green Globes: Green Building Assessment Protocol for Commercial Buildings (April 10, 2009, Draft) Total possible points = 1000
3. LEED for Homes (January 2008) Total possible points = 136
4. 2008 National Green Building Standard, January 2009 Total possible points = approximately 2000
TABLE 1
Potential Sustainable Design Contributions from Brickwork
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 12 of 14
LEED-NC
1
Green Globes
2
LEED for
Homes
3
2008 National
Green Building
Standard
4
Environmentally Preferable Materials
Avoidance of Construction Waste. Use modular
designs to avoid waste.
Materials &
Resources (MR)
1.2, 1.3
2 points
RE 601.3
3 points
Recycling of Construction Waste. Brick and
packaging are 100% recyclable.
MR 2
2 points
Resources/
Materials
10.5.1.1,
10.5.2.1 up to
7 points
MR 3.2
3 points
RE
605.2, 605.3
8 points
Salvaged Materials. Salvaged brick and pavers
can be reused.
MR 3
2 points
Resources/
Materials
10.3.1.1 up to
6 points
MR 2.2
point per
product
RE
603.2
3 points
Recycled Content. Brick may contain recycled
sawdust, sludge, metallic oxides. Mortar/grout may
use fly ash.
MR 4
2 points
Resources/
Materials
10.1.2.1,
10.2.2.1 up to
10 points
MR 2.2
point per
product
RE
604.1
up to 6 points
Regional Materials. Brick manufacturing plants
are located near raw materials and available
throughout the United States.
MR 5
2 points
Resources/
Materials
10.1.3.1,
10.1.4.1,
10.2.3.1,
10.2.4.1 up to
14 points
MR 2.2
point per
product
RE 608.1
2 points per
material
Materials That Do Not Require On-site Finishes.
No finishes are required of brickwork, can be used
inside as well.
RE 601.7
2 or 5 points for
each material
Materials Made with Renewable Energy. Several
brick manufacturers use landfill gas or sawdust to
fire their brick.
RE 606.3
2 for each
material
Durability and Design for Service Life
Durability. Brick has a useful life of more than 100
years.
Resources/
Materials
10.6.1.1,
10.6.2.1 up to
4 points
Termite Resistant Materials in Areas of Termite
Infestation. Insects do not eat brick.
SS 5
up to 1 point
RE 602.8
up to 6 points
Weather-resistant Barrier or Drainage Plane
Inside Siding or Veneer. Brick veneer introduced
the drainage wall.
Resources/
Materials
10.7.6.1
5 points
Flashing. Flashing is always present in well-
detailed brick buildings.
Resources/
Materials
10.7.2.1
5 points
RE 602.12
6 points
1. LEED-NC version 3 (LEED 2009) Total possible points = 110
2. Green Globes: Green Building Assessment Protocol for Commercial Buildings (April 10, 2009, Draft) Total possible points = 1000
3. LEED for Homes (January 2008) Total possible points = 136
4. 2008 National Green Building Standard, January 2009 Total possible points = approximately 2000
TABLE 1 (continued)
Potential Sustainable Design Contributions from Brickwork
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 13 of 14
LEED-NC
1
Green Globes
2
LEED for
Homes
3
2008 National
Green Building
Standard
4
Renewable Energy
Renewable Energy. Thermal mass of brick walls
and floors can be used in passive solar designs.
EA 1 up to 19
points (about
2-4 points from
brick)
Energy
8.2.1.1
4 points,
8.2.1.2
4 points,
8.2.1.3
4 points
Energy
Efficiency (EE)
703.1.3, 704.3.1
up to 16
points
Safety and Security
Fire-resistant Construction. Brick will not burn or
emit toxic fumes.
Impact-resistant Construction. Brick masonry
resists damage from wind-borne debris and other
impacts.
Efficient Use of Materials
Materials with Multiple Functions. Brick walls
can serve as structure and finish, provide acoustic
separation, and provide thermal mass.
RE 601.9
4 points
Materials Made with Less Content. Thinner brick
units use less material and weigh less; hollow brick
units use less material and can be reinforced.
RE 607.1
3 for each
material
Foundations that Require Less Material. Pier
and panel foundations of brick meet this practice.
RE 601.8
3 points
Structural Systems That Optimize Material Use.
Engineering design, rather than empirical design,
of brick walls provides better utilization of materials.
RE 601.2
3 points
Acoustic Comfort
Acoustic Comfort. Brick walls provide an STC of
45 or higher
ID Credit 1.1
1 point
Indoor
Environment
(IE) 12.6.1.2
2 points
Superior Indoor Air Quality
Avoid VOCs. Interior brick walls avoid paints.
Interior brick floors avoid carpets and adhesives.
ID Credit 1.3
1 point
IE 12.2.1.1
up to 10 points
ID 3.1 1 point Indoor
Environmental
Quality (EQ)
901.8 5 points
Masonry Fireplaces and Heaters. Fireplaces
with gasketed doors, outside combustion air and
a means of sealing the flue provide heat without
compromising indoor air or heat loss. Masonry
heaters of brick are energy efficient and clean
burning.
Indoor
Environmental
Quality (EQ)
2.2 2 points,
masonry
heaters only
EE 703.2 up
to 15 points and
EQ 901.2.1 4
points for fire-
places; 6 points
for heaters
Mold. With moisture-tolerant materials and
finishes, brick is not a food source for mold and
can be easily cleaned.
IE 12.2.2.1 up
to 5 points
Miscellaneous
Product manufacturer is ISO 14001 certified. RE 610.1
1 point per
percent up to 10
points
1. LEED-NC version 3 (LEED 2009) Total possible points = 110
2. Green Globes: Green Building Assessment Protocol for Commercial Buildings (April 10, 2009, Draft) Total possible points = 1000
3. LEED for Homes (January 2008) Total possible points = 136
4. 2008 National Green Building Standard, January 2009 Total possible points = approximately 2000
TABLE 1 (continued)
Potential Sustainable Design Contributions from Brickwork
www.gobrick.com | Brick Industry Association | TN 48 | Sustainability and Brick | Page 14 of 14
SUMMARY
This Technical Note provides information on sustainability and sustainable design as it relates to brick
manufacturing, use and recycling. The Brick Industry Association is committed to sustainable design and has
adopted the following environmental policy statement:
The brick industry recognizes that the stewardship of our planet lies in the hands of our generation.
Our goal is to continually seek out innovative, environmentally friendly opportunities in the
manufacturing process and for the end use of clay brick products. As demonstrated over time,
we are committed to manufacturing products that provide exceptional energy efficiency, durability,
recyclability, and low maintenance with minimal impact on the environment from which they originate.
We will ensure that our facilities meet or exceed state and federal environmental regulations, and
we will continue to partner with building professionals to help them in using our products to create
environmentally responsible living and working spaces for todays and future generations.
The information and suggestions contained in this Technical Note are based on the available
data and the combined experience of engineering staff and members of the Brick Industry
Association. The information contained herein must be used in conjunction with good technical
judgment and a basic understanding of the properties of brick masonry. Final decisions on the
use of the information contained in this Technical Note are not within the purview of the Brick
Industry Association and must rest with the project architect, engineer and owner.
REFERENCES
1. 2008 National Green Building Standard National Association of Home Builders, www.nahb.org,
Washington, DC, January 2009.
2. ASTM E2114-06a, Standard Terminology for Sustainability Relative to the Performance of Buildings, Vol.
4.12, ASTM International, West Conshohocken, PA, 2006.
3. Building for Environmental and Economic Sustainability, BEES 4.0, National Institute of Standards and
Technology, Gaithersburg, MD, www.nist.gov, May 2007.
4. Building Loads Analysis and System Thermodynamics (BLAST), University of Illinois, Building Systems
Laboratory, Urbana, IL, www.wbdg.org/tools/blast.php, June 2007.
5. EnergyPlus, U.S. Department of Energy, Washington, DC, www.eere.energy.gov/buildings/energyplus/,
June 2007.
6. Environmental Resource Guide, American Institute of Architects, John Wiley & Sons, New York, NY,
1998.
7. Environmental Stewardship & Employee Health and Safety Programs Awards, Brick Industry
Association, Reston, VA, April 1, 2007, press release.
8. Green Globes: Green Building Assessment Protocol for Commercial Buildings, Public Review Draft,"
Green Building Initiative, Portland, OR, www.thegbi.org, April 10, 2009.
9. High Performance School Buildings Resource and Strategy Guide, Second Edition, Sustainable
Buildings Industry Council, Washington, DC, www.sbicouncil.org, 2003.
10. LEED for Homes, U.S. Green Building Council, Washington, DC, www.usgbc.org, January 2008.
11. LEED Reference Guide for Green Building Design and Construction, U.S. Green Building Council,
Washington, DC, www.usgbc.org, 2009.
12. Mazria, Edward, AIA, 2030 Challenge, Architecture 2030, Santa Fe, NM, www.architecture2030.org,
April 2007.
13. Rand, J. Patrick, AIA, Designing Masonry for Durability and Sustainability, as presented to the University
Professors Masonry Workshop, Louisville, KY, March 2007.
14. Roodman, D.M., and N. Lenssen, A Building Revolution: How Ecology and Health Concerns are
Transforming Construction, Worldwatch Paper 124, Worldwatch Institute, Washington, DC, www.world-
watch.org, March 1995.
15. Whole Building Design Guide, National Institute of Building Sciences, www.wbdg.org, April 2007.

You might also like