You are on page 1of 188

ATC 115

Roadmap for the use of


high-strength reinforcement in
reinforced concrete design
95% Draft

Applied Technology Council


Funded by the
Charles Pankow Foundation

ATC-115
Roadmap for the Use of High-Strength
Reinforcement in Reinforced Concrete Design
95% Draft Report

by
APPLIED TECHNOLOGY COUNCIL
201 Redwood Shore Parkway, Suite 240
Redwood City, California 94065
www.ATCouncil.org
Funded by
CHARLES PANKOW FOUNDATION
P.O. Box 820631
Vancouver, Washington 98682

CHARLES PANKOW FOUNDATION REPRESENTATIVE


Mark J. Perniconi
ATC PROJECT MANAGER
Jon A. Heintz
PROJECT MANAGEMENT COMMITTEE
Dominic J. Kelly (Project Technical Director)
David Darwin
David C. Fields
Robert J. Frosch
Andres Lepage
Joseph C. Sanders
Andrew Whittaker

2014

PROJECT REVIEW PANEL


Wassim Ghannoum
S.K. Ghosh
Ramon Gilsanz
James O. Jirsa
Mike Mota
Thomas C. Schaeffer
Loring A. Wyllie, Jr.

Preface
In 2012, the Charles Pankow Foundation (CPF) began to investigate the
feasibility of incorporating reinforcing steel in excess of 60ksi into the ACI
318 Building Code. This investigation was prompted by interest on the part
of structural engineering practitioners, structural concrete constructors and
key academic researchers who felt that the use of higher strength reinforcing
bars could provide a significant benefit to the industry. The initial
investigative efforts of CPF first included several informal meetings with an
expert panel followed by the commissioning of several research projects that
studied the technical feasibility of using higher strength reinforcing bar as
well as developing a technical definition of the product. At the same time,
CPF engaged the producers of steel reinforcing bars to evaluate the technical
and financial feasibility of making high strength reinforcing bars
commercially available.
By mid-2013, initial feasibility studies commissioned by CPF confirmed the
technical feasibility of using high strength reinforcing bar in design. In
addition, reinforcing bar producers verified that higher strength reinforcing
bars could be manufactured and be made available through normal
distribution channels. Also in 2013, an unrelated study funded by NIST
(NIST GCR 14-917-30) and titled Use of High-Strength Reinforcement in
Earthquake-Resistant Concrete Structures (also called the ATC 98 Project)
was completed. This study confirmed the feasibility of using high strength
reinforcing bar in seismic applications. The next task was to determine the
supporting research necessary to effect an ACI 318 Building Code update to
incorporate the general use of reinforcing bar in excess of 60ksi. The last
comprehensive ACI 318 Code update related to rebar strength was done in
1971.
The vehicle chosen by CPF to define the steps necessary to effect ACI 318
Code changes was a roadmap that essentially reviewed every applicable
section of the ACI 318-14 Building Code and identified new research or
engineering studies necessary to support the required code changes and
updates. The entity chosen by CPF to manage this process is the Applied
Technology Council (ATC). CPF and ATC have a long history of successful
collaboration on research projects. The name of the project is The
Development of a Roadmap on Use of High-Strength Reinforcement in

ATC-115

Preface

iii

Reinforced Concrete Design, and the working title is the ATC 115 Project.
This research project may be found on the CPF website
(www.pankowfoundation.org) as Research Grant Agreement #05-13.
CPF wishes to acknowledge the hard work of Jon A. Heintz, Project
Manager for ATC. CPF is indebted to the leadership of Dominic J. Kelly,
Project Technical Director, and to the members of the Project Management
Committee, consisting of Andres Lepage, David C. Fields, David Darwin,
Joseph C. Sanders, Robert J. Frosch and Andrew Whittaker, for their
contributions in developing this report and the resulting recommendations.
The Project Review Panel, consisting of Wassim Ghannoum, James O. Jirsa,
Ramon Gilsanz, Mike Mota, Thomas C. Schaeffer and Loring A. Wyllie, Jr.,
provided technical review and commentary at key developmental milestones
during the project. CPF also wishes to acknowledge the contributions of
Conrad Paulson and Jack Moehle as special consultants to CPF and Robert
Risser, CEO of CRSI and its many producers and member who contributed
to this effort. Finally, CPF wishes to thank its Board of Directors, Richard
Kunnath, Ron Klemencic and Tim Murphy for their support and
contributions to this effort.
The completion of the ATC 115 Project is an important next step in the
process of adding high strength reinforcing bar into the ACI 318 Building
Code. The research effort defined in the ATC 115 Report will be one of the
largest research programs ever undertaken in the industry. To accomplish
this goal, it will take the support of the entire reinforced concrete industry.
The net result will be much more than just the addition of high strength
reinforcing bar into the ACI Building Code. The process of adding high
strength reinforcing bar into the ACI 318 Building Code will essentially
bring most of the technical specifications in the ACI 318 Building Code into
the 21st Century.
Mark J. Perniconi, P.E.
Executive Director
Charles Pankow Foundation

iv

Preface

ATC-115

Table of Contents
Preface........................................................................................................... iii
List of Figures............................................................................................... ix
List of Tables ................................................................................................ xi
1.

Introduction ...................................................................................... 1-1


1.1 Purpose and Scope of Work ..................................................... 1-1
1.2 Historical Perspective of High-Strength Reinforcement and
ACI 318 .................................................................................... 1-2
1.3 Prospects for Adoption of High-Strength Reinforcement into
ACI 318 .................................................................................... 1-3
1.4 Prospects for Adoption of High-Strength Reinforcement into
other U.S. Codes and Standards ............................................... 1-4
1.5 Key Issues with the Use of High-Strength Reinforcement....... 1-5
1.6 Report Organization and Content ............................................. 1-5

2.

Production, Fabrication, and Construction Issues ........................ 2-1


2.1 Reinforcement Production and Specification Issues ................ 2-1
2.1.1 Production of Deformed Steel Reinforcement ............ 2-1
2.1.2 Specification of High-Strength Reinforcement ........... 2-5
2.2 Fabrication Issues ..................................................................... 2-7
2.3 Constructability Issues and Construction Efficiencies ............. 2-9

3.

Current State-of-Knowledge on Key Design Issues ...................... 3-1


3.1 Strength and Ductility Issues .................................................... 3-2
3.1.1 Flexural and Axial Load Strength ............................... 3-2
3.1.1.1 Strain Limits................................................ 3-2
3.1.1.2 Beams .......................................................... 3-5
3.1.1.3 Columns ...................................................... 3-7
3.1.2 Shear Strength ........................................................... 3-10
3.1.2.1 Shear Strength of Members without Shear
Reinforcement ........................................... 3-10
3.1.2.2 Shear Strength of Members with Shear
Reinforcement ........................................... 3-13
3.1.2.3 Shear Friction ............................................ 3-15
3.1.2.4 Deep Beams .............................................. 3-16
3.1.2.5 Ordinary Structural Walls ......................... 3-17
3.1.3 Strut and Tie Modeling.............................................. 3-19
3.1.4 Other Strength Issues ................................................ 3-20
3.1.4.1 Concrete Strength...................................... 3-20
3.1.4.2 Tension Regions of Shells and Folded
Plates ......................................................... 3-21
3.1.4.3 Bonded Reinforcement Ratios for Members
with Unbonded Post-Tensioning ............... 3-22

ATC-115

Table of Contents

3.1.4.4

3.2

3.3

3.4

3.5

3.6

vi

High-Cycle Elastic Fatigue of


Reinforcement ........................................... 3-22
Serviceability .......................................................................... 3-23
3.2.1 Deflections ................................................................. 3-23
3.2.2 Drift ........................................................................... 3-27
3.2.3 Crack Control ............................................................ 3-27
Reinforcement Limits ............................................................. 3-29
3.3.1 Beams ........................................................................ 3-29
3.3.2 Slabs and Footings ..................................................... 3-30
3.3.3 Columns ..................................................................... 3-31
3.3.4 Walls .......................................................................... 3-32
Detailing and Other Design Considerations ........................... 3-33
3.4.1 Bend Test Requirements and Minimum Bend
Diameters ................................................................... 3-33
3.4.2 Transverse Reinforcement Spacing ........................... 3-35
3.4.2.1 Spacing of Transverse Reinforcement in
Members Resisting Gravity and Wind Loads
or in Ordinary Seismic-Force-Resisting
Systems...................................................... 3-35
3.4.2.2 Transverse Reinforcement Spacing for
Members of Intermediate and Special
Seismic-Force-Resisting Systems ............. 3-36
3.4.3 Head Attachment of Headed Deformed Bars ............ 3-38
3.4.4 Development and Splice Lengths .............................. 3-39
3.4.4.1 Straight Bar Development Lengths ........... 3-39
3.4.4.2 Hooked Bar Development Length ............. 3-41
3.4.4.3 Headed Deformed Bar Development
Length........................................................ 3-42
3.4.4.4 Mechanical Splice ..................................... 3-43
3.4.5 Bar Extensions in Slabs ............................................. 3-44
3.4.5.1 Bar Extensions in One-Way Slabs ............ 3-44
3.4.5.2 Bar Extensions in Two-Way Slabs ............ 3-44
3.4.6 Horizontal Support of Offset Column
Reinforcement ........................................................... 3-45
3.4.7 Cover for Fire Protection ........................................... 3-46
3.4.8 Beam-Column and Slab-Column Joints (Type 1)...... 3-48
3.4.8.1 Interior Beam-Column Joints .................... 3-48
3.4.8.2 Exterior Beam-Column Joints ................... 3-50
3.4.8.3 Exterior Slab-Column Joints ..................... 3-51
General Considerations for Analysis ...................................... 3-51
3.5.1 Flexural Stiffness ....................................................... 3-51
3.5.2 Moment Redistribution .............................................. 3-53
Seismic-Force-Resisting Systems ........................................... 3-54
3.6.1 Shape of Stress Strain Relationship ........................... 3-55
3.6.2 Intermediate Moment Frames .................................... 3-58
3.6.3 Special Moment Frames ............................................ 3-59
3.6.3.1 Beams ........................................................ 3-60
3.6.3.2 Columns .................................................... 3-62
3.6.3.3 Shear Demand on Beams and Columns .... 3-65
3.6.3.4 Strong Column-Weak Beam Behavior ...... 3-66
3.6.3.5 Beam-Column Joints ................................. 3-67
3.6.4 Flexure-Critical Special Structural Walls .................. 3-70

Table of Contents

ATC-115

3.6.5
3.6.6
4.

Shear-Critical walls ................................................... 3-73


Diaphragms ............................................................... 3-75

Research Studies............................................................................... 4-1


4.1 Objectives ................................................................................. 4-1
4.2 Overview .................................................................................. 4-1
4.3 Bar Production and Specification ............................................. 4-1
4.3.1 Mechanical Properties of Recent Heats of High
Strength Reinforcement ............................................... 4-2
4.3.2 Detailed Mechanical Property Tests of Grade 100 and
Grade 120 Reinforcement ........................................... 4-3
4.4 Strength of Members ................................................................ 4-4
4.4.1 Flexural Strength and Tensile Strain Limits ................ 4-4
4.4.2 Deflection Warning for Flexural Members Subjected
to Gravity Loads .......................................................... 4-5
4.4.3 Column Strength ......................................................... 4-7
4.4.4 Tension Regions of Shells and Folded Plates.............. 4-8
4.4.5 Shear Strength of Beams and Slabs without Shear
Reinforcement ............................................................. 4-9
4.4.5.1 One-Way Shear in Beams without Shear
Reinforcement ............................................. 4-9
4.4.5.2 Two-Way Shear in Slabs without Shear
Reinforcement ........................................... 4-10
4.4.6 Shear Strength of Beams with Shear
Reinforcement ........................................................... 4-11
4.4.7 Shear Friction ............................................................ 4-12
4.4.8 High-Cycle Elastic Fatigue of High-Strength
Reinforcing Bars ....................................................... 4-13
4.5 Serviceability .......................................................................... 4-14
4.5.1 Deflection of Flexural Members ............................... 4-14
4.5.2 Crack Control of Flexural Members ......................... 4-16
4.6 Reinforcing Limits ................................................................. 4-17
4.6.1 Minimum Reinforcement Ratio for Beams ............... 4-17
4.6.2 Minimum Reinforcement Ratio for Slabs and
Footings ..................................................................... 4-18
4.7 Detailing of Members ............................................................. 4-19
4.7.1 Development and Splice Lengths .............................. 4-19
4.7.2 Hooked Bar Development Length ............................. 4-20
4.7.3 Headed Bar Development Length ............................. 4-21
4.8 General Considerations for Analysis ...................................... 4-22
4.8.1 Flexural Stiffness....................................................... 4-22
4.8.2 Effective Stiffness for Column Slenderness .............. 4-23
4.8.3 Moment Redistribution.............................................. 4-23
4.9 Earthquake-Resistant Structures............................................. 4-24
4.9.1 Moment Curvature and Rotational Capacity ............. 4-24
4.9.2 Factor for Estimating Expected Flexural Strength .... 4-26
4.9.3 Cyclically Loaded Beams and Columns Initial
Tests and Analytical Studies ..................................... 4-27
4.9.4 Cyclically Loaded Beams, Columns, and Joints ....... 4-28
4.9.4.1 Cyclically Loaded Beams ......................... 4-28
4.9.4.2 Cyclically Loaded Columns ...................... 4-29
4.9.4.3 Cyclically Loaded Interior Joints .............. 4-30

ATC-115

Table of Contents

vii

4.9.4.4
4.9.4.5

Cyclically Loaded Exterior Joints ............. 4-31


Two-Way Shear in Slab-Column
Intermediate Moment Frames.................... 4-33
4.9.5 Performance of Moment Frames Systems ................. 4-33
4.9.6 Multi-Bay, Multi-Story Frames ................................. 4-35
4.9.7 Ordinary Flexure-Critical Walls ................................ 4-35
4.9.8 Special and Ordinary Shear-Critical Walls ................ 4-37
4.9.9 Special Flexure-Critical Walls Initial Tests ............ 4-38
4.9.10 Special Flexure-Critical Walls ................................... 4-39
4.9.11 Performance of Flexure-Critical Wall Systems ......... 4-41
4.10 Engineering Design Studies .................................................... 4-42
4.10.1 Trial Engineering Designs for Use of Grade 80
Reinforcement in Special Seismic Systems ............... 4-42
4.10.2 Trial Engineering Designs for Use of Grade 100
Reinforcement in General Applications .................... 4-44
4.10.3 Trial Engineering Designs for Use of Grade 100
Reinforcement in Special Seismic Systems ............... 4-45
5.

viii

Program Recommendations ............................................................ 5-1


5.1 Summary of Program ................................................................ 5-1
5.2 Estimated Budget Requirements............................................... 5-1
5.2.1 Budget Assumptions .................................................... 5-4
5.3 Priority and Schedule Recommendations ................................. 5-5
5.3.1 Priority Level 1 Revise ASTM A615 to Include
Grade 100 Reinforcement ............................................ 5-6
5.3.2 Priority Level 2 Modify ACI 318 to Allow the use
of ASTM A706 Grade 80 Reinforcement in Special
Seismic Systems .......................................................... 5-6
5.3.3 Priority Level 3 Modify ACI 318 to Allow the use
of ASTM A615 Grade 100 Reinforcement in General
Applications (Gravity, Wind, and Ordinary Seismic
Systems)....................................................................... 5-7
5.3.4 Priority Level 4 Develop a new ASTM specification
for Grade 100 Reinforcement for use in Special
Seismic Systems .......................................................... 5-9
5.3.5 Priority Level 5 Modify ACI 318 to Allow the use
of Grade 100 Reinforcement in Special Seismic
Systems ........................................................................ 5-9
5.4 Other Conclusions and Recommendations ............................... 5-9
5.4.1 Potential Code Changes that can be Implemented
without Additional Research ..................................... 5-10
5.4.2 Issues Not Requiring Further Action or Code
Change ....................................................................... 5-11
5.4.3 Other Potential Studies .............................................. 5-11
5.5 Implementation Recommendations ........................................ 5-12
5.5.1 Key Collaborators ...................................................... 5-12
5.5.2 Technical Oversight ................................................... 5-13
5.5.3 Technical Synthesis ................................................... 5-13
5.5.4 Adoption into Codes and Standards .......................... 5-14

Table of Contents

ATC-115

References .................................................................................................. A-1


Project Participants ...................................................................................B-1

ATC-115

Table of Contents

ix

List of Figures
Figure 3-1

Strength reduction factor based on strain........................ 3-2

Figure 3-2

Types of stress-strain curves with distinct shapes ............. 3-4

Figure 3-3

Influence of reinforcement ratio on shear strength .......... 3-11

Figure 3-4

Ratios of tested to predicted punching shear capacity


using ACI 318-05............................................................. 3-13

Figure 3-5

Developed bar stresses and concrete compressive


strengths for development and splice tests on bars
without confining transverse reinforcement .................... 3-40

Figure 3-6

Developed bar stresses and concrete compressive


strengths for development and splice tests on bars with
confining transverse reinforcement.................................. 3-40

Figure 3-7

Yield strength and tensile strength of reinforcing bars


after heating and cooling.................................................. 3-47

Figure 3-8

Yield strength and tensile strength of reinforcing bars at


high temperature .............................................................. 3-47

Figure 3-9

Permissible redistribution in accordance with ACI


318 ................................................................................... 3-53

Figure 3-10

Beam specimens used to study the effect of the tensilestrength-to-yield-strength ratio ........................................ 3-57

Figure 3-11

Load-deflection curves for beam specimens with: (a) yield


ratio of 75% with splice; and (b) yield ratio of 90% without
splice. ............................................................................... 3-57

Figure 3-12

Details for beam specimens with high-strength reinforcing


bars ................................................................................... 3-61

Figure 3-13

Measured shear versus drift ratio in beam tests: (a) Specimen


CC4-X, with Grade 60 reinforcing bars; and (b) Specimen
UC4-X, with Grade 97 reinforcing bars .......................... 3-61

Figure 3-14

Hysteretic response of two circular columns based on


Restrepo et al. (2006): Unit 1 including Grade 60
reinforcement, and Unit 2 including Grade 100
reinforcement ................................................................... 3-63

ATC-115

List of Figures

ix

Figure 3-15

Details for column specimens with high-strength


reinforcing ........................................................................ 3-64

Figure 3-16

Measured shear versus drift ratio in column tests:


(a) Specimen CC-3.3-20, with Grade 60 reinforcement;
and (b) Specimen UC-1.6-20, with Grade 120
reinforcement ................................................................... 3-64

List of Figures

ATC-115

List of Tables
Table 3-1

Minimum Percentage of Fracture Elongation for ASTM


A615, A706, and A1035 Reinforcing Bars ...................... 3-24

Table 3-2

Minimum Bend Diameter of Hooks in ACI 318 and Bend


Test Requirements for ASTM A615, A706, and A1035
Reinforcing Bars .............................................................. 3-24

Table 3-3

Stress-Strain Parameters for Reinforcing Steel Used to


Create Moment-Curvature Relationships in NIST GCR
14-917-30 ......................................................................... 3-56

Table 4-1

Summary of Studies Related to Bar Production and


Specification ...................................................................... 4-1

Table 4-2

Summary of Studies Related to Strength of Members ....... 4-4

Table 4-3

Summary of Studies Related to Serviceability ................ 4-14

Table 4-4

Summary of Studies Related to Reinforcing Limits ........ 4-17

Table 4-5

Summary of Studies Related to Detailing of Members ... 4-19

Table 4-6

Summary of Studies Related to General Considerations


for Analysis ...................................................................... 4-22

Table 4-7

Summary of Studies Related to Earthquake-Resistant


Structures ......................................................................... 4-25

Table 4-8

Summary of Studies Related to Trial Engineering


Designs ............................................................................ 4-42

Table 5-1

Summary of Proposed Research and Engineering


Studies................................................................................ 5-2

Table 5-2

Estimated Budget Requirements by Program Area ........... 5-3

Table 5-3

Recommended Priorities Based on Implementation


Objectives and Target Milestone Dates ............................. 5-5

Table 5-4

Priority Level 1 Revise ASTM A615 to Include Grade


100 Reinforcement ............................................................. 5-6

Table 5-5

Priority Level 2 Modify ACI 318 to Allow the use of


ASTM A706 Grade 80 Reinforcement in Special Seismic
Systems .............................................................................. 5-7

ATC-115

List of Tables

xi

Table 5-6

Priority Level 3 Modify ACI 318 to Allow the use of


ASTM A615 Grade 100 Reinforcement in General
Applications (Gravity, Wind, and Ordinary Seismic
Systems) ............................................................................. 5-8

Table 5-7

Priority Level 4 Develop a new ASTM specification for


Grade 100 Reinforcement for use in Special Seismic
Systems .............................................................................. 5-9

Table 5-8

Priority Level 5 Modify ACI 318 to Allow the use of


Grade 100 Reinforcement in Special Seismic Systems.... 5-10

Table 5-9

Potential Code Changes that can be Implemented without


Additional Research ......................................................... 5-11

Table 5-10

Issues Not Requiring Further Action or Code Change..... 5-11

Table 5-11

Other Potential Studies ..................................................... 5-12

xii

List of Tables

95% Draft

Chapter 1

Introduction

Concrete reinforcement with yield strength greater than 75 ksi is becoming


more available in the United States. In 2009, ASTM International included
Grade 80 in specification A615, which is the most commonly referenced
reinforcing bar specification. ASTM A1035 includes Grades 100 and 120,
which are available from a single producer. A bar produced in Germany,
SAS670, has a minimum yield strength of 97 ksi. ASTM A1035 Grade 100
and SAS670 have been approved by the New York City Building
Department as column reinforcement. In the near future, many producers in
the United States will be capable of reliably producing Grade 100 bars, and
within 10 years, will likely be able to produce Grade 120 bars.
1.1

Purpose and Scope of Work

High-strength reinforcement refers to reinforcement with grades higher than


Grade 60. The primary objective of the ATC-115 Project was to prepare a
detailed Roadmap that specifically identifies the technical support required,
whether it be the results of new research, engineering studies, or reevaluation of existing research findings, to effect updates of ACI 318-14,
Building Code Requirements for Structural Concrete and Commentary (ACI,
2014a), to allow the general use of steel reinforcement in excess of Grade 80
for gravity and wind load applications, and in excess of Grade 60 for special
seismic systems. The project specifically considered the use of Grade 80,
Grade 100, and Grade 120 high-strength reinforcement, and considered
applications in members designed for gravity loads, wind loads, or seismic
loads, as part of a gravity system, or as part of an ordinary, intermediate, or
special seismic-force-resisting system.
The resulting Roadmap outlines the steps needed to:

further investigate the use of high-strength reinforcement in general


reinforced concrete design and construction in building and non-building
structural applications; and

develop code-change proposals and encourage adoption of high-strength


reinforcement in the code development process.

Development of this Roadmap was based on the following activities:

ATC-115

1: Introduction

1-1

95% Draft

identification of production, fabrication, and design issues to be


considered;

identification of constructability challenges;

search for relevant research and other available information on highstrength reinforcement in the published literature;

determination of the current state-of-knowledge regarding issues and


challenges associated with high-strength reinforcement;

evaluation of the current-state-of knowledge and determination of when


existing information was sufficient or when additional research or study
was needed.

identification of additional experimental research and engineering studies


needed; and

estimation of the approximate budget, schedule, and prioritization for a


recommended program.

1.2

Historical Perspective of High-Strength


Reinforcement and ACI 318

American Concrete Institute (ACI) Committee 318 has a history of revising


provisions to include the use of higher strength reinforcement than was
readily available in the past. As of the 1950s and early 1960s, Intermediate
Grade (Grade 40) reinforcement and Hard Grade (Grade 50) reinforcement
had been available for about 50 years, and was commonly used. In 1959, the
American Society for Testing and Materials (ASTM) specifications A432
and A431 were published, which introduced Grade 60 and Grade 75
reinforcement, respectively. The 1963 version of ACI 318 allowed the use of
steel bars with a yield strength of 60 ksi. ASTM A615, Standard
Specification for Deformed and Plain Carbon-Steel Bars for Concrete
Reinforcement, which included Grades 40, 60 and 75, was introduced in
1968.
In the late 1950s and 1960s, the Portland Cement Association (PCA)
conducted a series of tests reported in eight parts that examined beams,
girders, and columns (Hognestad, 1961; Hognestad, 1962; Gaston and
Hognestad, 1962; Kaar and Mattock, 1963; Pfister and Mattock, 1963; Pfister
and Hognestad, 1964; Kaar and Hognestad, 1965; and Kaar, 1966). These
tests covered flexural strength, control of flexural cracking, compression
splices in columns, and fatigue. Reinforcement strengths ranged from 55 ksi
to 120 ksi. At about the same time, Thomas and Sozen (1965) published the
results of tests of beams reinforced with unstressed prestressing

1-2

1: Introduction

ATC-115

95% Draft

reinforcement with a yield strength of 230 ksi. These early tests were
considered in changes to the 1971 edition of ACI 318 when the upper limit
for yield strength was increased to 80 ksi, even though there were no ASTM
specifications for reinforcement with yield strengths more than 75 ksi at the
time.
In the 1971 edition of ACI 318, the maximum specified yield strength was
restricted to 60 ksi for reinforcement in special seismic systems, and this
limit is still in effect in ACI 318-14. The Structural Engineers Association of
California (SEAOC) developed a specification for reinforcement with more
restrictive tensile properties and chemistry controls, published as ASTM
A706, Standard Specification for Low-Alloy Steel Deformed and Plain Bars
for Concrete Reinforcement (ASTM, 1974). ASTM A706 was permitted in
the 1977 edition of ACI 318. It became required for reinforcement in special
seismic systems in 1983, but an exception was made for the use of A615
reinforcement if certain mechanical properties were met.
The 2009 versions of ASTM A615 and A706 specifications (ASTM, 2009a;
ASTM, 2009b) were the first to include requirements for Grade 80
reinforcement. ACI Committee 318 adopted these specifications without
restriction in the main body of ACI 318-11 (ACI, 2011) because the use of
Grade 80 reinforcement was already permitted. However, Grade 80
reinforcement is currently not permitted for use in special moment-resisting
frames and special structural walls due to the perceived lack of test data for
cyclically loaded members with Grade 80 reinforcement.
1.3

Prospects for Adoption of High-Strength


Reinforcement into ACI 318

Based on the current availability of high-strength reinforcement, its use in


Japan, research completed to date, and progress made towards codifying its
use, the prospects for adoption of high-strength reinforcement into ACI 318
are good. High-strength reinforcement is available from some producers in
the United States, and many other producers will soon be capable of rolling
it. Significant research was completed in Japan as part of the New RC
Project (Aoyama, 2001), which took place between 1988 and 1993.
Research has continued in Japan, and is being performed in Taiwan, Korea,
and the United States. Much of this research is identified and described in
NIST GCR 14-917-30, Use of High-Strength Reinforcement in EarthquakeResistant Concrete Structures (NIST, 2014).
High-strength reinforcement has been used in hundreds of high-rise buildings
in Japan, and has been used in some buildings in New York City.
Documents such as ACI ITG-6, Design Guide for the Use of ASTM
ATC-115

1: Introduction

1-3

95% Draft

A1035/A1035M Grade 100 Steel Bars for Structural Concrete (ACI, 2010a),
and NCHRP Report 679, Design of Concrete Structures Using High-Strength
Steel Reinforcement (Shahrooz et al., 2011), have made progress towards
identifying how some code provisions in ACI 318 and the AASHTO Bridge
Design Specifications should be changed.
Adopting the use of high-strength reinforcement into ACI 318 will require a
substantial effort because many sections of the code will require new or
revised provisions, and many questions and concerns will likely be raised by
the members of ACI Committee 318. Research and engineering studies
identified in this Roadmap are intended to develop the necessary code change
proposals, and are also intended to comprehensively address the questions
and concerns that are likely to be raised.
1.4

Prospects for Adoption of High-Strength


Reinforcement into other U.S. Codes and Standards

The nuclear industry is keen on using high-strength reinforcement in new


construction to reduce congestion, improve the quality of placed concrete,
and speed construction time. Prospects for the adoption of high-strength
reinforcement into ACI 349-13, Code Requirements for Nuclear SafetyRelated Concrete Structures and Commentary (ACI, 2014b), are high, and its
use has been debated at recent ACI committee meetings. The reduction in
minimum elongation associated with the use of higher strength reinforcement
is of lesser concern in the nuclear industry because component nonlinear
deformation demands are much smaller than those in buildings and bridges at
design level intensities of shaking.
The bridge industry is also interested in using high-strength reinforcement in
new construction. Goals for the use of high-strength reinforcement are
similar to those of the nuclear industry. The California Department of
Transportation is currently working with North Carolina State University to
provide strain information on ASTM A706 Grade 80 reinforcement. If
adequate ductility is shown, A706 Grade 80 bars could be used in capacityprotected components, such as bent caps, where rebar congestion is the
highest. For use in seismic members that are expected to yield (e.g.,
columns), however, rigorous testing of couplers and splices will be
necessary. Other State Departments of Transportation (DOTs), and the
Federal Highway Administration (FHWA), have also shown significant
interest in high-strength reinforcement for general reinforced concrete
design.

1-4

1: Introduction

ATC-115

95% Draft
1.5

Key Issues with the Use of High-Strength


Reinforcement

Several challenges, questions, and concerns must be resolved before highstrength reinforcement can be used extensively in the construction of
reinforced concrete structures. These stem from bar production and
fabrication challenges, design requirements, and impacts on constructability,
as described in Chapters 2 and 3 of this Roadmap.
1.6

Report Organization and Content

This Roadmap identifies key issues in production, fabrication, design, and


construction related to the use of high-strength reinforcement, identifies the
current state-of-knowledge, and outlines a recommended program of
experimental research and engineering studies to investigate the use of highstrength reinforcement and support its adoption in building codes and
standards for reinforced concrete design.
Chapter 2 identifies the current state-of-knowledge on issues related to
production and fabrication of high-strength reinforcement, and construction
using high-strength reinforcement.
Chapter 3 identifies the current state-of-knowledge on issues related to
design using high-strength reinforcement.
Chapter 4 outlines recommended experimental research and engineering
studies needed to investigate the use of high-strength reinforcement and
support the development of code changes.
Chapter 5 provides a summary of the overall program, including estimated
budget requirements, priority and schedule recommendations, and additional
recommendations related to implementation of the program.

ATC-115

1: Introduction

1-5

95% Draft

Chapter 2

Production, Fabrication, and


Construction Issues

Challenges associated with the production and fabrication of high-strength


steel for concrete reinforcement will be most affected by the required
mechanical properties and the cost of production. This chapter describes
production, fabrication, and construction issues associated with the use of
high-strength reinforcement.
2.1

Reinforcement Production and Specification Issues

In general, higher yield strength is accompanied by reductions in the tensileto-yield strength ratio, uniform elongation (elongation at tensile strength),
and length of yield plateau. In some cases, the yield plateau is completely
eliminated.
The processes used to produce high-strength reinforcement will likely
depend on the mechanical properties of the reinforcement needed for safe
and serviceable designs. These properties are addressed in sections of this
report that deal with design issues. The requirements for of mechanical
properties to be specified must also realistically account for the capabilities
of producers and the economics of producing the bar.
The requirements for mechanical properties of high-strength steel bars and
the means by which bars will be specified needs to be established as part of
the program described in this report. Issues related to inventory and storage,
identifications of bars on the site, production costs, and required properties
for various design conditions will affect the criteria used for specifying highstrength reinforcement.
2.1.1

Production of Deformed Steel Reinforcement

Reinforcement must be rolled in a very malleable state so that bar


deformations will form properly when it is rolled. Therefore, reinforcing
steel is red hot when being rolled. Higher strength reinforcement is
commonly achieved by micro-alloying. It can also be achieved by quenching
and self-tempering or by work hardening, although work hardening is not
commonly used to produce reinforcing bars in the United States.

ATC-115

2: Production, Fabrication, and Construction Issues

2-1

95% Draft

The means by which high-strength reinforcement will be produced may be


somewhat dependent on the properties that need to be achieved for the
reinforcement. Reinforcement used for resisting gravity loads (i.e., dead,
live, ice, snow, and rain) requires reliable yield strength with less emphasis
on ductility, whereas reinforcement used to resist earthquake forces requires
more emphasis on the elongation and ratio of tensile-to-yield strength.
Where quenching is used to produce bars, it involves the use of water to
rapidly cool the steel that has been heated to the austenite phase (the phase at
which solid steel recrystallizes), while self-tempering is achieved by the
gradual release of the heat that is trapped in the core of the quenched steel.
Quenching results in a hard metal structure while self-tempering softens the
steel and increases its toughness.
The quenching and self-tempering process for reinforcing steel involves
spraying bars with water just after rolling, producing a layer of Martensite
(hard phase), and then allowing the residual heat from the core of the bar to
re-heat the quenched outer layer. Upon cooling, the core of the bar contracts
and converts to Ferrite-Pearlite. This contraction develops compressive
stresses in the cooler outer portion of the bar while the core develops tensile
stresses. These stresses improve bendability and fatigue resistance of the bar.
When the bar is subjected to tension, the residual stresses in the bar result in
the core yielding before the outer portion leading to a stress-strain curve
without a sharp yield point.
Increasing the yield strength of reinforcing steel by means of micro-alloying
involves the addition of small quantities of certain alloying elements to the
molten steel to induce grain refinement. Vanadium (V) is the most
commonly used alloy due to its dependability and consistent results. It
refines the grain size and increases the strength without a negative impact on
weldability or notch toughness. Molybdenum (Mo), Niobium (Nb), Tantalum
(Ta), and Titanium (Ti), to name a few, are also used for micro-alloying.
The addition of Chromium (Cr) in higher than trace amounts is also effective
in increasing yield strength and in combination with copper improves
corrosion resistance. Generally, Grade 80 reinforcement produced in the
United States is micro-alloyed, as is also commonly done for large Grade 60
bars (i.e., No. 9 and larger). Less grain refinement is achieved by rolling
larger bars so more micro-alloying makes up the difference to achieve the
required properties.
Generally, micro-alloyed bars have a higher ratio of tensile-to-yield strength
than quenched and tempered bars, and a lower ratio of tensile-to-yield
strength than plain carbon steel bars.

2-2

2: Production, Fabrication, and Construction Issues

ATC-115

95% Draft

For the New Zealand market, both quenching and tempering and microalloying are used to produce ductile reinforcement with strength on the order
of 75 ksi specifically for use in structures resisting earthquake effects. The
use of these processes (or a combination of the two) has potential for
producing even higher strength reinforcement.
Work hardening involves the plastic deformation of a metal at or near room
temperature. Cold drawing plain and deformed wire reinforcement is a form
of work hardening. In some countries, the strength of bars is increased by
the process of cold twisting after hot rolling. Work hardening improves yield
strength but eliminates a yield plateau, and reduces both the achievable
elongation and the ratio of tensile-to-yield strength. Work hardening may be
suitable to produce reinforcement for use in members where yielding is not
expected. For current design practice in the United States, work hardening is
likely not suitable to produce reinforcement for members resisting
earthquake effects or where yielding is expected.
The processes by which producers will manufacture high-strength
reinforcement will be driven by minimizing the costs relative to producing
Grade 60 reinforcement. This will likely mean minimizing changes to
current methods of producing reinforcing bars. For example, using the same
bar deformations as those for Grade 60 bars and eliminating bar marks will
allow the same rolls to be used. However, doing so may increase the
likelihood of the wrong reinforcement being used. Also, some producers
within the reinforcing steel industry oppose the elimination of bar marks
because it is a deviation from long-established practice.
Some producers are interested in eliminating the marks from bars because
bar marks must be chosen before the bars are produced and tested.
Occasionally, bars do not meet the required mechanical properties. Because
they have a mark that they do not comply with, these bars must be scrapped.
One idea to address this is that the bar could be rolled without the marks and
laser printing could be used to indicate the heat on the bar. The test results
for the heats could then be placed on an accessible database. This approach
would allow the producer to sell the reinforcement as complying with a
different specification or grade. The downside of this approach is that
inspectors and engineers checking bar placement in the field likely will not
have access to the database on site, so they may not have access to the type
and grade of reinforcement while on the site. There has also been an
indication that color coding the ends of bars could be used, but this would not
be effective for bars that are shop cut. Another idea that could be explored
is to laser print the specification and grade that the bar meets after testing is
complete; however, this would require additional bar handling. An added
ATC-115

2: Production, Fabrication, and Construction Issues

2-3

95% Draft

benefit of not rolling the marks into the bar is elimination of the known
weakness that bars tend to fail in low and high cycle fatigue at these marks.
Also, in epoxy coated bars, coating damage tends to occur at sharp edges of
bar marks during delivery and handling in the field.
Reinforcing bars are most commonly produced in straight lengths at the mill.
However, coiling smaller sized bars (i.e., No. 3 to No. 6), is becoming more
common, and currently represents nearly 5% of the market for bars of all
sizes. Bars are coiled soon after rolling and while hot, which traps heat in the
coil. The trapped heat can lead to unintentional annealing. Annealing is the
process of heating steel between 700 and 800C for several hours and
allowing it to cool slowly. Annealing softens the steel and increases ductility
and workability of steel. It leads to a lower yield strength and lower tensile
strength with a rounded post-yield stress-strain curve. To overcome
unintentional annealing, high-strength reinforcement that is coiled will likely
require higher quantities of micro-alloying elements than for bundled straight
bars, which have more opportunity to cool before being bundled.
In addition to the production methods affecting the tensile-to-yield strength
ratio, they will also affect bar elongation. From an engineering perspective,
reporting uniform elongation, i.e., the elongation that exists in the bar as the
tensile strength is reached, is preferable to elongation at fracture, which is the
increase in length over the 8- in. gage length including the necked-down
region of the bar. U.S. producers are considering the impacts of reporting
uniform elongation. They do not have a history of reporting uniform
elongation, so they are currently trying to determine if and how uniform
elongation is affected by chemistry and production processes. It is also
recognized that uniform elongation is likely reduced with an increase in bar
size. Although one or more producers are willing to roll all bar sizes, at least
initially, producers may want to roll a limited number of bar sizes. Doing so
will allow the producers to minimize extra storage space required with the
availability of additional grades of bars. At least one producer is rolling No.
20 bars and has plans for rolling No. 24 bars.
Producers have expressed an interest in having Grade 100 included in the
ASTM A615 specification, or equivalent, in the near future. If such a
specification were to exist, producers believe there would be demand for the
bar. Producers would use the production of this bar as an opportunity to
explore means of producing more ductile Grade 100 bars. This will help
determine realistic material properties to include in a future specification for
Grade 100 reinforcement targeted for use in earthquake-resistant construction
(i.e., special moment frames and special structural walls).

2-4

2: Production, Fabrication, and Construction Issues

ATC-115

95% Draft

As producers refine how they will be producing high-strength steel bars, an


initial test program is recommended to define structurally acceptable
properties of high-strength bars to assist with the development of ASTM
specifications. In addition to tensile and bend tests, consideration should be
given to performing less common tests such as low-cycle fatigue and bendrebend tests. To evaluate the effects of bar bending on the ductile-to-brittle
transition temperature of steel, careful consideration should be given to
performing tests similar to those conducted by Hopkins and Poole (2005) on
AS/NZ 500E reinforcement (AS/NZS 4671 2001). Data collected from
Charpy V-notch toughness test are also relevant for assessing the expected
performance of the bar if used in structures exposed to colder climates. In
addition, the effects of strain aging need to be studied. An initial test program
will account for most of these issues.
Once the higher grade reinforcement is being reliably produced, a second
more detailed test program of reinforcement properties should be performed.
Recommendations for Studies of Bar Mechanical Properties
An initial test program as described in Section 4.3.1 is needed to explore
issues related to the mechanical properties of high-strength bar. A detailed
test program of reinforcement properties as described in Section 4.3.2 should
be performed once the high-strength reinforcement is being reliably
produced.
2.1.2

Specification of High-Strength Reinforcement

The development of specifications for high-strength reinforcement will


likely involve the following actions:

Expand ASTM A615 (carbon steel bars) to cover higher grades for
reinforcing bars, possibly from 80 to 120 ksi with 20-ksi increments.
This specification would not apply for bars in members that are part of
special seismic-force-resisting systems;

Create a new ASTM specification to cover all grades for reinforcing bars
to be used in members that are part of special seismic-force-resisting
systems;

Leave ASTM A706 (low-alloy steel bars) to cover only Grades 60 and
80.

The approach of having Grade 100 in ASTM A615 will give producers room
to experiment with the production of more ductile Grade 100 bars for the
future seismic specification with reduced risk of scrapping out entire heats.

ATC-115

2: Production, Fabrication, and Construction Issues

2-5

95% Draft

Specification overlap should be maximized. For example, 100% of the bars


meeting the specification for seismic applications should be adequate for
meeting the specification for ASTM A615, though not all ASTM A615 bars
would be adequate for seismic applications. The specification overlap should
be performed at all strength levels, not just higher grades. Taking this
approach would allow producers and fabricators to minimize bar inventories
and storage.
The chemical restrictions in ASTM A706 (weldable steel) make it difficult to
obtain high-strength steel bars with tensile-to-yield strength ratios and
uniform elongations similar in magnitude to those routinely obtained for
Grades 60 and 80 bars. Imposing limits on carbon content (or carbon
equivalency) restricts achievable strengths. Therefore, including grades
higher than 80 in ASTM A706 is not recommended at this time.
Specifications for bars with specific ranges of chemical composition or for
specific uses are likely to be maintained, but perhaps with some
modifications. For example, Grades 100 and 120 bars are already produced
to specification ASTM A1035. Changes are likely to be made to these
specifications that would provide for increased acceptance and perhaps to
meet requirements for use in earthquake-resistant structures. The producers
of bars to these specifications may also be willing to consider using high
relative rib area deformations to improve bond characteristics. However,
potential reduced fatigue resistance of bars with high relative rib area should
be explored.
Regardless of how the high-strength reinforcement is specified, yield
strength will be reported as the 0.2% offset (Paulson, et al., 2013) required in
ACI 318-14, ASTM A615, ASTM A706, and ASTM A1035. For
elongation, consideration should be given to reporting uniform elongation at
peak stress rather than total elongation (typically measured over a gage
length that includes the necked-down region), as recommended in NIST
GCR 14-917-30.
Producers are requesting to increase the bend test radii in the ASTM
specification for Grade 100 bars to match the bend radii specified for
standard hooks in ACI 318-14. The reason is to reduce the risk of having
heats rejected (by failing the bend test) while they are still refining how highstrength steel bars should be produced. Increasing the radii for the bend test
will increase the chance of bar fracture as it is shop bent. This issue should
be resolved as higher grade bars are added to the ASTM A615 specification.

2-6

2: Production, Fabrication, and Construction Issues

ATC-115

95% Draft

There will be some applications in which bars will be used in structures


exposed to cold weather. ASTM A706 bar is sometimes specified for these
applications because of its more stringent ductility requirements. Because
ASTM A706 will not include requirements for Grade 100 bar, supplemental
requirements could be added to the future specification of ASTM A615
Grade 100 bars and to specifications of high-strength bars for use in
earthquake-resistant structures.. The supplemental requirements would likely
include passing Charpy V-notch tests conducted on the reinforcement at cold
temperatures.
Producers are the ones who should take the lead in determining how bars are
specified and how they will be produced. As the challenges of
manufacturing bars to meet specified properties will be resolved by
individual producers, and the producers will be most affected by how highstrength bars are specified, they should have the most input on how to
resolve these issues. They will need input from the engineering and research
communities on the requirements to include in the specifications assuming
current design procedures are continued. The requirements must be
achievable by the producers. If not, design requirements could be revised so
that acceptable designs can be obtained using the bars with achievable
properties. Although funded research is not required to resolve these issues,
resolving them is on the critical path so that bars can be produced for the
research that is necessary for accepting the general use of high-strength steel
bars as concrete reinforcement. Producers need to develop a database for
measured mechanical properties associated with different manufacturing
processes and conduct a cost-benefit analysis based on the cost of process
modifications to improve key mechanical properties.
Recommendations for Specification Development
The reinforcing bar producers need to take the lead in developing
specifications containing the minimum acceptable properties of high-strength
reinforcement. The specifications are likely to include an expansion of
ASTM A615 to incorporate Grades 100 and 120. In addition, a new ASTM
specification should be developed for high-strength reinforcement with
controlled mechanical properties for use in earthquake-resistant structures.
2.2

Fabrication Issues

Issues with fabrication of high-strength reinforcement are related to the


following two general categories:

ATC-115

2: Production, Fabrication, and Construction Issues

2-7

95% Draft

The introduction of multiple grades of reinforcing bars that need to be


scheduled, received, and stored at a fabricating facility prior to their use;
and

The actual fabricating processes required for high-strength bars.

Fabrication facilities will need more space to store multiple bar sizes of
multiple grades prior to, and during, their fabrication. This problem may be
reduced by standardization around the use of fewer bar sizes in higher
strengths.
The actual fabrication processes of shearing the bars and bending the
reinforcement will be impacted when using higher strength reinforcement.
High-strength reinforcement requires greater force to shear and bend the
same size bar, and involves increased rebound at bends after the bending
force is removed. This leads to issues and concerns regarding:

wear and tear on bending equipment;

safety of workers when bending bar; and

complying with fabrication tolerances for bars.

Fabricating bends of high-strength reinforcement requires greater force to


make the bends and causes more wear and tear on equipment used to make
the bends. Bending high-strength reinforcing bars imparts higher force on
the central pin that holds the disc about which the bar is bent. Some
fabrication shops have reported that these pins break more frequently and
release more energy upon breaking when high-strength reinforcement is
fabricated. Safety concerns are heightened with this increase in breakage. In
cases where bars have defects that cause them to fracture when bending, the
release of energy is also greater, which creates a safety concern. Extra
precautions may be necessary to maintain a safe work environment. Worker
protection cages are more often required, which leads to lower efficiency in
the shop.
Bending higher strength bars will require larger equipment that is capable of
imparting larger forces on the bars. Larger and stronger equipment will
likely require more floor space in a shop.
ACI 117 (ACI, 2010b) provides specific tolerances for the tails of hooks and
the resulting angle between straight portions of bent bars. Increased rebound
of bars at bends makes it harder to control the final bar configuration in
higher strength reinforcement. Consideration needs to be given to whether
the tolerances should be adjusted for high-strength reinforcement.

2-8

2: Production, Fabrication, and Construction Issues

ATC-115

95% Draft

Issues related to fabrication and concerns for wear and tear on bending
equipment, safety of workers when bending bars, and compliance with
fabrication tolerances for bars are best addressed by fabrication shops.
Concerns of wear and tear on the equipment and worker safety may require
modifications to bending equipment.
Issues related to bar fracture and tolerances may be resolved by increasing
the inside bend diameter. Engineers and researchers can assist with
determining whether changes to tolerances are acceptable, and if larger
diameter bends are necessary, engineers and researchers can determine the
impacts on the design requirements. However, doing so will require
additional research to determine the effects on development length of
standard hooks.
The difficulty in bending bars may lead to a greater use of headed bars
involving the threading of bars or otherwise attaching the head to the end of
the bar.
In general, the fabricating challenges associated with the use of higher
strength bars will be lessened with greater acceptance and use of the bars.
Adaptation of equipment to deal with high-strength steel bars will be a
natural consequence of the increased use of the bars.
Recommendations for Resolving Fabrication Issues
Fabricators should determine whether bend diameters and tolerances at bends
for high-strength reinforcement need to be adjusted in relation to
conventional strength reinforcement. Any changes required for detailing of
high-strength reinforcement will need a close examination of their impact on
current design and construction practices.
2.3

Constructability Issues and Construction Efficiencies

From a construction and cost effectiveness perspective, the benefit of higher


strength reinforcement is that it can allow the use of a lower volume of
reinforcing steel to accomplish the same design goals as those obtained with
conventional reinforcement. This will generally result in lower construction
time and lower cost for concrete structures. The challenge for the design
community will be to integrate the use of high-strength reinforcement into
concrete structures in ways that optimize and utilize the higher yield strength
of the steel bars. As there will be some cost premium for the higher strength
steel on a per pound basis, the reduction in overall reinforcing steel weight
within a structure will need to be sufficiently large to offset the premium in
cost per pound. Minimum spacing, minimum reinforcement ratios, and other

ATC-115

2: Production, Fabrication, and Construction Issues

2-9

95% Draft

non-strength related requirements in ACI 318-14 will have to be evaluated to


determine their impact on obtaining the full benefit of high-strength bars.
The use of higher strength concrete may be necessary to completely utilize
the strength of the bars.
If substituting Grades 80, 100, or 120 for Grade 60, while maintaining the
member cross section, the use of higher strength reinforcement can be
implemented in two ways:

Using bars of the same size as Grade 60, but spacing them at greater
intervals; or

Using smaller size bars but maintaining spacing similar to that of Grade
60.

Using bars spaced at greater intervals means that construction and cost
efficiencies are achieved through lower placement costs, less congestion and
better consolidation potential for concrete placement. Fewer bars also mean
fewer lap splices or mechanical splices required. An unfavorable aspect of
reinforcement cages built with bars at a greater spacing is that they are less
stable when in a free-standing condition. Also, higher strength bars are more
difficult to bend, couple, and terminate.
Using smaller size bars at a spacing similar to that required of Grade 60 bars
means that, although bar weight will still be reduced, there will be less
placement cost savings. Smaller diameter bars are easier to bend, terminate,
and couple as compared to larger diameter bars. Cages built with a greater
number of smaller diameter bars will likely be more stable in a free-standing
condition during construction.
Higher strength bars will require longer splice and development lengths than
Grade 60; this will impact construction and cost efficiencies. If greater bar
length is needed to develop higher strength bars, this will negate some of the
benefit of reduced steel volume. Splice and development length issues may
be reduced by exploring higher relative rib area designs for the bars to
increase their force transfer to the concrete.
If higher strength is used in a slab, this may lead to a greater concern for slab
deflections. Larger slab deflections impact such things as exterior wall
attachments and floor levelness. Post-tensioning of slabs might be a way to
overcome this concern.
Although the use of high-strength reinforcement is generally associated with
reducing congestion, increased congestion is possible. Designers may be
tempted to use smaller concrete member dimensions in which case

2-10

2: Production, Fabrication, and Construction Issues

ATC-115

95% Draft

congestion will be similar to what occurs in members reinforced with Grade


60 bars. Also, the spacing of transverse reinforcement may need to be
reduced to better resist longitudinal bar buckling.
During construction, contractors frequently have to order additional bars on
short notice from fabricators. In the near term, as plants are still ramping up
production capabilities, the availability of reinforcement on short notice
could be a problem. If the bars needed are high strength and the fabricator
has not stored extra bars of this grade, construction delays could ensue.
More likely, Grade 60 bars will be substituted for the higher strength bars,
which could increase congestion, especially if member sizes have been
reduced based on the assumed use of higher strength reinforcement.
If multiple grades of the same size bars are used on a construction project,
the contractor may unintentionally install bars of the incorrect grade. This is
more likely to occur if the bars are straight and of similar lengths. Bar
markings or other identifiers will have to be clear to prevent this from
occurring.
In the early stages of a transition to higher strength reinforcement, the use of
all bar sizes for multiple grades of reinforcement will lead to supply chain,
inventory, and inspection complexities for producers, fabricators, and
installers. These types of issues are best considered early in the design
process by focusing on using a limited quantity of bar sizes for the higher
strength bars.
The unit material cost of high-strength reinforcement has been high in recent
years because demand has been low. With increased demand, prices will
decrease. In addition, cost savings in labor costs associated with placing
lower amounts of steel are likely to exceed the difference in material costs.
A cost study performed for NIST GCR 14-917-30 determined that cost
savings associated with the use of Grade 80 reinforcement (in lieu of Grade
60) was on the order of 4% of the cost of the concrete structure. The study by
Price, et al., (2014) evaluated the potential for high-strength reinforcement to
reduce the volume of reinforcement in typical concrete buildings and reduce
construction time, ultimately leading to a reduction in the total cost of
building construction. Price et al. concluded that if the ACI Code limits on
steel yield strength are ignored, high-strength reinforcement can be
effectively used to further reduce steel volumes for reinforced concrete
building components. The study suggested future research activities to
support ACI Code changes and enable the full potential of high-strength
reinforcement.

ATC-115

2: Production, Fabrication, and Construction Issues

2-11

95% Draft

Thoughtful consideration of the construction and cost efficiencies of the


aspects of the reinforcing steel design identified in this section will be
necessary to maximize the benefit from using higher strength reinforcing
bars.
Recommendations for Addressing Construction Issues and Efficiencies
No studies are necessary to address construction issues.

2-12

2: Production, Fabrication, and Construction Issues

ATC-115

95% Draft

Chapter 3

Key Design Issues

Design issues associated with the use of high-strength reinforcement could


potentially affect many design provisions in ACI 318, as well as other codes
and standards for reinforced concrete construction. Issues related to design
for gravity and wind forces acting on the structure can broadly be grouped as
affecting provisions related to:

Design strength and ductility issues

Serviceability

Reinforcement Limits

Detailing

Analysis

Resolution of issues or concerns related to the use of high-strength


reinforcement first requires an understanding of the effects of high-strength
reinforcement on structural behavior. Once the effects of high-strength
reinforcement on behavior are established, revisions to design provisions can
be considered.
Serviceability is more likely to control the design of members with highstrength reinforcement than with Grade 60 reinforcement. Some
serviceability provisions in ACI 318 explicitly considered that Grade 80
reinforcement could be used, and likely require modification only for Grade
100 or 120 bars. For example, the provision for minimum beam side-face
reinforcement is based on research that considered the use of Grade 80
reinforcement for the beam. In contrast, recommendations for effective
moment of inertia for determination of lateral drift are based on the opinion
of designers that were using Grade 60 reinforcement for design of structures.
Use of high-strength reinforcement in beams, columns, and walls of seismicforce-resisting systems has a high potential to take full advantage of the
increase in bar strength because strength provisions generally control over
serviceability issues in these members. However, detailing requirements
may lessen the advantage of using high-strength reinforcement, as closer
spacing of transverse reinforcement may be required.

ATC-115

3: Key Design Issues

3-1

95% Draft
This chapter identifies the current state-of-knowledge on key design issues
associated with the use of high-strength reinforcement.
3.1

Strength and Ductility Issues

Design provisions for computing flexural strength, axial load capacity, and
shear strength may require adjustments for members reinforced with highstrength reinforcement. Other miscellaneous provisions and approaches to
strength computation may also be affected.
3.1.1

Flexural and Axial Load Strength

Accurately determining the strength of members under flexural or combined


flexural and axial loading is key to the design of reinforced concrete
members because these loads generally establish the size of the members.
Design for other load effects, such as shear and torsion, necessarily follow.
Thus, changes in member behavior that result from the use of high-strength
reinforcing steel must be well understood to allow for the safe and
economical use of this new material.
3.1.1.1

Strain Limits

In determining the design strength of a member, it is necessary to reduce the


nominal strength through the use of a strength reduction factor, . The
strength reduction factor for flexure, axial load, or both, is computed based
on the strain conditions of the cross-section at nominal strength and in
particular, the extreme tensile strain, t, as shown in Figure 3-1. This
approach to computing has been in the main body of ACI 318 since 2002,
and is based on the Unified Design Provisions for Reinforced and
Prestressed Concrete Flexural and Compression Members recommended by
Mast (1992).

Figure 3-1

3-2

Strength reduction factor based on strain (ACI 318-14).

3: Key Design Issues

ATC-115

95% Draft
As evident in Figure 3-1, is dependent on the region in which the
computed strain lies, and two limits are used to control the value, the
compression controlled limit (ty) and the tension controlled limit (0.005). In
considering the use of high-strength reinforcement, it is important to address
whether these two limits remain appropriate.
Compression-controlled strain limit
For the compression-controlled strain limit, ACI 318 defines ty as fy/Es
which is considered the yield strain of the reinforcement. The use of the
yield strain to define this limit is appropriate regardless of reinforcement
strength, and this computation is correct for use with reinforcement that
remains linear-elastic up to the yield plateau. However, for stress-strain
curves with a gradual reduction in stiffness and rounded post yield strength
that becomes nonlinear before reaching the yield strength, the definition of
the yield strain as fy/Es may not be appropriate especially if fy is being defined
through the offset method or other alternate procedure. Therefore, changes
to ACI 318-14 may be required in the definition depending on the variation
of the actual yield strain to the computed value using the elastic computation.
Alternately, specific values of ty for a given reinforcement grade may be
needed similar to that provided for prestressed reinforcement.
Tension-controlled strain limit
For the tension-controlled strain limit, a fixed value of 0.005 is used. This
value, which is designed to provide ductility, is approximately 2.5 times the
yield strain of about 0.002 for A615 Grade 60 reinforcement, and is higher
than required in ACI 318 prior to 2002. According to Mast (1992), the value
of 0.005 was selected because:

the net tensile strain is measured at dt instead of d (or dp), and the limit
should be slightly higher when more than one layer of tension steel is
used; and

use of 0.005 produces interaction diagrams that look reasonable.

For high-strength reinforcement, the yield strain will be greater; therefore, it


appears that the tension controlled limit should be increased to provide levels
of ductility similar to that designed with A615 Grade 60 reinforcement.
Consequently, it may be appropriate to provide a limit that is a function of
the reinforcement yield strain.
The shape of the stress-strain curve may affect the strain that should be used
for this limit. Similar to the discussion for the compression-controlled limit,
the use of fy/Es to define the yield strain, which may be used to define the

ATC-115

3: Key Design Issues

3-3

95% Draft
tension-controlled limit, may not always be appropriate. ACI ITG-6 (2010)
points out the tensile strain corresponding to a tension-controlled member for
ASTM A1035 bars is 0.0066, based on realistic stress-strain curves, and
0.009, based on a simplified representation of the highly nonlinear ASTM
A1035 stress-strain curve idealized as elastic-plastic.
Stress-strain curves for high-strength reinforcement will likely exhibit greater
variation in shape than those of ASTM A615 Grade 60 reinforcement. As
illustrated in Figure 3-2, three distinct shapes for the stress-strain curves of
bars likely to be available are: (1) linear-elastic behavior up to the yield
strength, with a sharp bend representing the yield strength, a relatively flat
yield plateau, and a rounded strain hardening region (curves designated S3 in
the figure); (2) a gradual reduction in stiffness with a rounded post-yield
shape that becomes nonlinear before reaching the yield strength, which is
defined by the 0.2% offset method, followed by gradual softening up to the
tensile strength (curve designated S1 in the figure); and (3) a nearly bi-linear
curve in which linear-elastic behavior occurs almost to the yield strength,
followed by linear strain-hardening behavior in the inelastic range until the
tensile strength is reached (curve designated S2 in the figure).
160
140

Stress, ksi

120
100
80
100S1
100S2
100S3
80S3
60S3

60
40
20
0
0.00

0.01

0.02

0.03

0.04

0.05

0.06

Strain

Figure 3-2

Types of stress-strain curves with distinct shapes.

Maximum strain limit


With the use of high-strength reinforcement, there is likely to be a reduction
in the elongation capacity of the reinforcement relative to that of Grade 60.
While the current code does not include a limit on the tensile strain, there is
the potential that a limit will need to be established so that fracture of the
reinforcement does not occur prior to crushing of the concrete in

3-4

3: Key Design Issues

ATC-115

95% Draft
compression. Studies should be conducted to evaluate if a maximum strain
limit is required.
Recommendations for Strain Limits
It is recommended that the strain limits defining the boundaries between
compression-controlled, transition, and tension controlled sections for the
selection of factor values be examined. Considering that the current limit is
partially based on the appearance of interaction diagrams, this study should
not only investigate flexural response, but also evaluate the interaction
between flexure and axial loads. In addition, it should be examined if a
maximum strain limit is required. The proposed study is outlined in 4.4.1.
3.1.1.2

Beams

Flexure strength provisions for tension-controlled members, which include


most slabs and beams, are intended to provide an acceptable level of
reliability that the computed design strength will be achieved at a reasonable
deflection. These provisions are also intended to result in designs that
provide ample warning of a potential failure if the member is inadvertently
overloaded. The warning includes substantial cracking and large deflections
forming prior to complete failure. The provisions were developed assuming
reinforcing steel has a yield plateau and based on confirmation of adequate
performance of beam specimens tested in laboratories.
In determining whether existing flexural strength design provisions are
adequate or require change for members reinforced with high-strength
reinforcement, the following should be considered:

The shape of the stress-strain curve for the reinforcement affects the
deflection at which the design strength is developed and potentially the
spread of plasticity as the member is loaded monotonically to failure.

A minimum extreme tensile strain t of 0.004 is required by ACI 318 for


flexural members.

The yield strain of compression reinforcement may exceed the assumed


maximum concrete strain of 0.003.

Stress-strain curve
The shape of the load-deflection curve of tension-controlled members
constructed with high-strength reinforcement will be affected by the shape of
the stress-strain curve. Current provisions for computing flexural strength
are based on the assumption that the stress-strain curve for the bar has a yield
plateau. It is not yet clear whether the flexural strength provisions need to

ATC-115

3: Key Design Issues

3-5

95% Draft
explicitly account for the shape of the stress-strain curve if the bar does not
have a yield plateau. The effects of the stress-strain curve shape should be
explored to determine whether changes to the flexural design provisions are
required, which could be accomplished using moment-curvature studies and
tests to validate the analyses.
The desire for large deflections to occur prior to failure, which would warn
building occupants of the potential for failure of gravity loaded flexural
members, is considered in the provisions of ACI 318. The ability to develop
these large deflections is dependent in part on the mechanical properties of
the reinforcement. In particular, bars in flexural tension with a yield plateau
and a large value of the tensile-to-yield strength ratio will result in members
that deflect more prior to failure. About 95% of ASTM A615 Grade 60 bars
currently produced have a yield plateau (Paulson, et al., 2013). Based on the
study of reinforcing bar data by Bournonville et al. (2004), the average
tensile-to-yield strength ratio is about 1.5 with the average ratio for larger
bars being as low as 1.4. Although ASTM A615 does not set minimum
acceptable values for this ratio, the reinforcement produced has acceptable
values for this ratio in most cases. An experimental program should be
performed to confirm that slabs and beams reinforced with Grade 100 bars
with low tensile-to-yield ratios, and low elongations, will deflect sufficiently
to provide adequate warning prior to failure.
Minimum tensile strain
According to ACI 318, a minimum tensile strain is required to ensure a
minimum level of ductility. The current limit of 0.004 is approximately
twice the yield strain of about 0.002 for A615 Grade 60 reinforcement.
Although this limit was not part of the recommendations in Mast (1992), this
limit was included in ACI 318-02 to provide consistency with past practice,
which included a maximum longitudinal reinforcement ratio of 0.75b. This
reinforcement ratio produces a net tensile strain ty of 0.00376 for Grade 60
reinforcement, which is the reason ACI 318 adopted the slightly more
conservative value of 0.004. Considering that this ratio was selected based
on Grade 60 reinforcement, an adjustment may be required for high-strength
reinforcement. Another possibility is to delete this requirement and address
only the tension-controlled limit, as was originally recommended by Mast
(1992).
Compression reinforcement
Compression reinforcement is commonly used in beams. In some instances,
compression reinforcement is provided in beams with smaller cross sections

3-6

3: Key Design Issues

ATC-115

95% Draft
to allow the use of more tension reinforcement. For high-strength
reinforcement used in compression, the yield strain in the reinforcement is
likely greater than the assumed maximum concrete strain of 0.003.
Therefore, the design stress in compression reinforcement will be
automatically limited by the strain profile of the cross-section, or by a
prescribed limit in the maximum design stress, such the current 80 ksi limit
in ACI 318. Considering that a strain of 0.003 would result in a steel stress
of 87 ksi, it would seem that no change is needed in ACI 318 to address this
issue. However, for practical considerations, the stress in the compression
reinforcement will be limited by the assumed maximum concrete
compressive strain, and the current 80 ksi limit on the stress in the
reinforcing steel could be removed without complication, if desired.
Recommendations for Flexural Strength of Slabs and Beams
An analytical study should be performed to evaluate the impact of the
roundhouse stress-strain curve on flexural strength. It is recommended that
a moment-curvature analysis be performed that incorporates various stressstrain relationships for high-strength reinforcement. These computed
flexural strengths should be compared with code-predicted nominal strengths
to determine if any changes are needed to the code calculation procedure.
This study is outlined in Section 4.4.1.
In addition, a series of tests should be performed to confirm that beams
reinforced with Grade 100 reinforcement, but with lower tensile-to-yield
strength ratios, and the low end of elongations that can realistically be
produced, will deflect an adequate amount prior to failure. This study is
presented in Section 4.4.2.
3.1.1.3

Columns

The yield strength of compression reinforcement is limited to 80 ksi in ACI


318-14. This limit is imposed because bars with yield strengths much above
80 ksi will not contribute to higher column capacity considering that higher
yield strength can only be achieved at strains above 0.003 (the strain assumed
for concrete crushing). At an assumed maximum strain of 0.003, the
maximum usable stress is 87 ksi, considering linear-elastic behavior of the
reinforcement.
Richart and Brown (1934) reported the results of a test program in which 564
concentrically loaded columns were tested. Eight of the tests were of
spirally-reinforced columns that were longitudinally reinforced with bars
having a yield strength of 96 ksi, tensile strength of 133 ksi, and fracture
elongation of 11% (8-in. gauge length). The stress-strain curves for the

ATC-115

3: Key Design Issues

3-7

95% Draft
longitudinal bars had well-defined yield plateaus. Richart and Brown (1934)
concluded that the high-strength longitudinal bars were fully effective in
producing the column strength. All columns had 4% vertical reinforcement
and 1.2% spiral (volumetric reinforcement ratio) spaced at one-sixth of the
core diameter. The spiral had a yield strength of 64 ksi and tensile strength of
100 ksi.
Pfister and Mattock (1963) reported the results of 16 concentrically-loaded
circular and rectangular columns reinforced with high-strength steel bars
with and without splices. The circular columns (12-in. diameter) were
spirally reinforced and the rectangular columns (1012 in.) were tied. The
longitudinal reinforcement consisted of six No. 8 bars, for a reinforcement
ratio of about 4%, with yield strengths between 82 and 93 ksi. The yield
strength of the transverse reinforcement was 65 ksi for the spiral and 59 ksi
for the ties. The volumetric ratio of spiral reinforcement provided in the
circular columns was 1.3%, while the area ratio of lateral ties in the
rectangular columns was 0.1% (very low but compliant with the minimum
permitted by ACI 318-63 (ACI, 1963)). Based on their test results, the
following conclusion was reached, If the specified yield point of
longitudinal reinforcement in tied columns is to be developed at ultimate
strength of the columns, then it is necessary that the yield point be reached at
or before a strain of 0.003 in./in. This condition will normally be more
readily complied with by bars having a clearly defined yield point and a
nearly linear stress-strain curve up to yield than by bars having a gradually
curving stress-strain curve with no clearly defined yield point. The report
also states, In future usage of 90 ksi column reinforcement, lapped splices
will probably be impractical even with spiral reinforcement.
Todeschini et al. (1964) tested eccentrically loaded tied columns reinforced
with high-strength steel with a specified yield strength of 75 ksi. Reinforcing
bars both with and without yield plateaus were used. Reinforcement with a
rounded stress-strain curve reached 90 ksi at a strain of 0.006. They found
that the shape of the steel stress-strain curve affected the stress that can be
utilized in the reinforcement. For reinforcement with a rounded stress-strain
curve, the stress in the reinforcement that can be utilized was 70 to 80 ksi.
For reinforcement with a relatively flat yield plateau, stresses up to 90 ksi
could be developed.
References to more recent test programs and findings are presented in NIST
GCR 14-917-30, with emphasis on members reinforced with high-strength
steel bars and subjected to reverse cyclic loading. The test results indicate
that replacing Grade 60 longitudinal reinforcement with reduced amounts of
high-strength reinforcement (reduced in proportion to the yield strength of
3-8

3: Key Design Issues

ATC-115

95% Draft
reinforcement) leads to comparable flexural strength and deformation
capacity. The test specimens included reinforcement with limited variations
in the shapes of the steel stress-strain curve.
Considering that the maximum usable concrete strain is limited to 0.003, and
considering strain compatibility, the maximum design stress of the
reinforcement is limited to 87 ksi. Therefore, it does not seem necessary to
conduct further research in this area. The current limit of 80 ksi can be
maintained, or this limit could be removed to allow stresses in the reinforcing
steel consistent with the maximum strain limit. It should be realized that the
maximum stress of 87 ksi is only realized under pure axial load conditions.
Strain gradients will decrease the strain in the compression reinforcement.
Most of the discussion up to this point has concentrated on pure axial load.
However, most practical columns include an interaction between axial load
and flexure. Overall, the computation of the P-M interaction curve is
controlled by the maximum concrete strain and the strain gradient. The
reinforcement is directly accounted for through strain compatibility;
therefore, it does not appear that high-strength reinforcement will cause any
additional issues. The more important issue, however, will be the selection
of the compression-controlled and tension-controlled strain limits as these
significantly impact the design strength curves. As previously noted in
Section 3.1.1.1, the selection of the tension-controlled strain limit was
influenced by the appearance of the interaction diagrams.
Recommendations for Columns
It is recommended that interaction diagrams be considered as part of the
selection of the tension-controlled strain limit as discussed in 3.1.1.1. A
study investigating this topic is outlined in 4.4.1.
A code change is not considered required, but it appears that the maximum
stress limit of 80 ksi could be removed. This minor change would result in
the maximum strain limit of concrete controlling design. The stress of the
reinforcement can then be computed directly from the stress-strain
relationship of the reinforcement. If the stress-strains relationship of the
high-strength reinforcement is linear up to 87 ksi, then the maximum usable
stress of the reinforcement would be 87 ksi. If the reinforcement has a stressstrain curve with a significantly rounded shape and nonlinear behavior before
the yield strength is reached, then the usable stress would be reduced. A
study investigating the elimination of the 80 ksi limit is outlined in 4.4.3.

ATC-115

3: Key Design Issues

3-9

95% Draft
3.1.2

Shear Strength

The proper calculation of shear strength is essential for the design and safety
of structural concrete members. The overall philosophy in design is to
provide for shear strength greater than flexural strength allowing for
complete flexural along with signs of warning prior to failure. In accordance
with ACI 318, shear strength consists of the addition of a concrete
component Vc and a shear reinforcement component Vs (ACI 326 1962). This
approach is used throughout ACI 318 and is provided for both one-way and
two-way shear strengths. While the simplest and most commonly used shear
design expressions in ACI for both one-way and two-way shear are based on
the assumption that shear strength is independent of flexure, research has
demonstrated that the reinforcement ratio significantly affects shear strength.
Consequently, the effects of the use of high-strength flexural reinforcement
need to be considered. The use of high-strength transverse reinforcement is
also expected to influence shear strength; therefore, the effects of the use of
this reinforcement also need to be considered.
3.1.2.1

Shear Strength of Members without Shear


Reinforcement

A number of members are designed without the use of shear reinforcement


and typically include slabs and foundations. Many beams often also do not
require shear reinforcement in certain regions of the span or in any part of the
span depending upon size and geometry. The use of high-strength
longitudinal reinforcement used to develop flexural strength, however, may
affect the concrete contribution to shear strength. Considering that the use of
higher strength reinforcement is likely to produce a member with a reduced
amount of longitudinal reinforcement; higher strains levels than provided by
Grade 60 reinforcement will occur for the same loading. Higher longitudinal
strains result in larger crack widths which have the potential to reduce shear
transferred across cracks (Angelakos et al. 2001, Brown et al. 2006, Collins
and Kuchma 1999, Lubell et al. 2004, Reineck et al. 2003, Tureyen and
Frosch 2003). In addition, a reduction in the amount of longitudinal
reinforcement reduces the depth of the compression zone which is considered
a primary region contributing to shear transfer. Overall, it is important to
determine if adjustments in commonly used shear expressions are required.
3.1.2.1.1

Beams , One-Way Slabs, and One-Way Foundations

The common expression used to compute one-way shear strength is


2 f c' bw d . As previously noted, this expression is independent of the
flexural reinforcement. This expression was developed considering the
lower-bound of test results available in the 1960s (ACI 326 1962). While

3-10

3: Key Design Issues

ATC-115

95% Draft
this expression has been in use for decades, it has been known that there is an
interaction between flexure and shear. In fact, ACI 318 provides an
alternative equation that includes this interaction.
Research by Tureyen and Frosch (2003) clearly illustrates the influence of
the reinforcement ratio. In evaluating shear tests available in the literature
along with more recent tests, the significant influence of the reinforcement
ratio and thus the interaction between flexure and shear is evident as shown
in Figure 3-3. As the reinforcement ratio decreases below 1%, shear
strengths significantly below 2 f c' occur. If the reinforcement ratio is
maintained constant between different reinforcement grades, a reduction in
shear strength is not expected. In a research study by Tureyen and Frosch
(2002), a specimen was reinforced with high-strength steel reinforcing bars
having a yield strength of 108 ksi. It was found that the behavior of this
specimen was essentially identical to that of a companion FRP reinforced
specimen for the same effective reinforcement ratio. Therefore, for the same
axial stiffness of the reinforcement, the same overall behavior occurred and
the same shear strengths developed. Therefore, the use of high-strength
reinforcement alone is not expected to change overall shear behavior. It is
important to note that shear tests are designed such that the longitudinal
reinforcement does not yield, and maintaining the longitudinal reinforcement
in the elastic region is fundamental in clearly identifying shear failures and
subsequently shear strength.
6.0

Steel
FRP

5.0
4.0

Trendline
Vtest
3.0
bwd f c
2.0
1.0
0.0
0.0

Figure 3-3

ATC-115

1.0

2.0
3.0
4.0
5.0
Effective Reinforcement Ratio, eff (%)

6.0

7.0

Influence of reinforcement ratio on shear strength (Tureyen and


Frosch, 2003).

3: Key Design Issues

3-11

95% Draft
While shear behavior is not expected to change with the use of high-strength
reinforcement, it is expected that the use of high-strength reinforcement will
decrease the amount of longitudinal reinforcement ratio required which will
consequently decrease shear strengths. Furthermore, it is expected that
reinforcement ratios below 1% will be common. Therefore, a change in the
expression used to calculate shear strength is needed.
The shear expression proposed by Tureyen and Frosch (2003), 5 f c' bw c , can
potentially provide a solution in this regard, and this expression has been
adopted for the design of FRP reinforced members (ACI 440 2006) for the
same reason. As noted in the figure, FRP reinforced structures result in low
effective reinforcement ratios (<1%) and require a different expression for
the calculation of shear strength. Low effective reinforcement ratios occur in
FRP reinforced structure due to the relatively low modulus of elasticity of
these materials (often 1/4 to 1/5 of steel).
Recommendation for Further Study

It is clear that further research is needed regarding the shear strength of


beams and one-way members without shear reinforcement. It is important to
evaluate the shear strength of one-way members reinforced with highstrength reinforcement and determine an appropriate design expression for
the calculation of strength Vc,. A proposed research program is outlined in
4.4.5 and in particular for one-way shear in 4.4.5.1.
3.1.2.1.2

Two-Way Slabs and Two-Way Foundations

For two-way slabs and foundations, punching shear strength typically


controls design, and this strength has many of the same issues as outlined for
one-way shear strength. Punching shear strength in ACI 318 is considered
independent of flexure with the base punching shear strength computed as
4 f c' bo d . Again, experimental results indicate that the reinforcement ratio
has a significant influence on shear strength. Ospina (2005) evaluated
punching shear strength from tests available in the literature. As shown in
Figure 3-4, shear capacity ratios (Vtest/Vcalc) below 1.0 result as the
reinforcement ratio is decreased.
With the use of high-strength reinforcement, it is expected that lower
reinforcement ratios would be used. Therefore, a design approach that
accounts for the reduction in reinforcement ratio is important. Ospina (2005)
recommended a modification of the Tureyen and Frosch expression used for
one-way shear, 10 f c' bo c . This approach, which has also been adopted by
ACI 440.1R-06, Guide for the Design and Construction of Structural

3-12

3: Key Design Issues

ATC-115

95% Draft
Concrete Reinforced with FRP Bars (ACI, 2006), may also be appropriate
for the use of high-strength reinforcement.

Figure 3-4

Ratios of tested to predicted punching shear capacity using ACI


318-05 (Ospina, 2005).

Recommendation for Further Study

Further research is needed regarding the shear strength of two-ways


members. It is important to evaluate the punching shear strength of two-way
members reinforced with high-strength reinforcement and determine an
appropriate design expression for the calculation of strength Vc,. A proposed
research program is outlined in 4.4.5 and in particular for two-way shear in
4.4.5.2.
3.1.2.2

Shear Strength of Members with Shear Reinforcement

The shear strength of members containing shear reinforcement is computed


through the addition of the concrete contribution to shear strength Vc and the
shear reinforcement contribution Vs. The contribution due to concrete Vc is
identical to that for specimens without shear reinforcement and is based on
tests results from those specimens. The shear contribution is based on the
strength of an individual stirrup multiplied by the number of stirrups assumed
to cross a shear crack. The strength of a stirrup is assumed to be the area of
the legs of the stirrup multiplied by the yield strength of the stirrup.

ATC-115

3: Key Design Issues

3-13

95% Draft
Although tests (ACI-ASCE Committee 326, 1962; ACI-ASCE Committee
426, 1973) support the use of yield strength; these tests were focused on
yield strengths up to 60 ksi. Therefore, in considering the use of highstrength reinforcement, it is important to determine if the full yield strength
can be utilized.
ACI 318 currently limits the design yield strength of shear reinforcement to
60 ksi for reinforcing bars and to 80 ksi for welded deformed wire
reinforcement. The commentary states that the limit of 60 ksi provides a
control on diagonal crack width while the higher yield strength of 80 ksi can
be used for deformed wire reinforcement because the widths of inclined
shear cracks at service load levels were less for beams reinforced with
smaller-diameter welded deformed wire reinforcement cages designed on the
basis of a yield strength of 75,000 psi than beams reinforced with deformed
Grade 60 stirrups. Considering that the ACI 318 limit on yield strength is
based on serviceability rather than strength, it may not be possible to fully
utilize high-strength reinforcement as shear reinforcement. It may be
possible, however, to control diagonal crack widths through detailing such as
the use of smaller transverse bars with a decreased spacing.
Overall, there are two primary issues that need to be addressed in considering
the use of high-strength reinforcement as shear reinforcement. First,
determine whether the yield strength of the reinforcement can be fully
utilized in the development of shear strength. In this regard, it is also
important to determine if the addition of the concrete contribution Vc and the
shear reinforcement Vs remains applicable. Second, whether it is possible to
acceptably control diagonal crack widths when reaching stresses in the
transverse reinforcement greater 60 ksi.
Several studies have been conducted investigating the use of high-strength
reinforcement as shear reinforcement (Aoyama (2001), Budek et al. (2002),
Munikrishna (2008), Ou et. al. (2012), Sumpter et al 2009)). In addition, a
National Cooperative Highway Research Program (NCHRP) research project
was conducted to evaluate the use of high-strength reinforcement (NCHRP
2011) and tests were included specifically evaluating the use of high-strength
reinforcement as shear reinforcement. In general, these studies indicate that
an increase above the current 60 ksi limit for shear reinforcement is possible.
In the NCHRP study, it was indicated that only small differences in crack
widths between Grade 60 and Grade 100 stirrups were observed. Therefore,
crack control may be possible even at higher stress levels.
As a secondary issue, ACI 318 requires a minimum area of shear
reinforcement. This reinforcement area is currently based on preventing

3-14

3: Key Design Issues

ATC-115

95% Draft
sudden shear failures when include cracking occurs and for that reason is a
function of f c' . The total force provided, however, is based on the yield
strength of the transverse reinforcement ( Av ,min f vt ). Therefore, if the
transverse reinforcement is not capable of reaching yield, this minimum
reinforcement requirement may need to be modified.
Recommendation for Further Study

Further research is recommended to fully address the two primary issues


identified above. In particular, investigation of service level crack widths is
essential to ensure proper serviceability when stresses greater than 60 ksi in
the stirrups are developed. This study should also evaluate if the current
minimum reinforcement requirement is appropriate. A proposed research
program is outlined in 4.4.6.
3.1.2.3

Shear Friction

Design for shear forces that are transferred across a given plane such as an
existing or potential crack, an interface between dissimilar materials, or an
interface between two concretes cast at different times is handled using the
shear friction design method. In this method, shear strength is calculated as
the normal force across the plane multiplied by a coefficient of friction, .
As shear is applied across the interface, one surface slips or slides along the
other. Assuming that the surfaces are irregular such as cracking concrete, the
slip is accompanied by vertical separation of the surfaces. This vertical
separation produces strains and resulting stresses in the reinforcement. At
peak strength, this separation is sufficient to yield the reinforcement.
Therefore, the normal force is computed as Avf f y . The coefficient of friction
is different depending on the condition of the shear plane and if different
materials are connected at this location.
Considering a shear plane with a given amount of reinforcement crossing the
plane, the lowest force is transferred across the plane when concrete is placed
against hardened concrete that is not intentionally roughed. In this case, is
assumed to be 0.6. It can be assumed that vertical separation is negligible
due to the smooth interface and that the shear reinforcement is acting
primarily as a dowel. Considering von-Mises yield criterion, the yield
strength of a dowel is computed as approximately 0.6fy which is consistent
with the commonly assumed shear yield strength of steel (AISC 360).
Therefore, the coefficient of friction in this case may be considered as shear
yielding of the shear reinforcement.
Considering the same shear plane and mount of reinforcement, the greatest
force is transferred applies when concrete is placed monolithically. In this

ATC-115

3: Key Design Issues

3-15

95% Draft
case, is assumed to be 1.4. Because a jagged and irregular interface will be
formed when the concrete cracks along the interface, maximum vertical
separation will occur. Therefore, shear strength is provided by a combination
of shear yielding of the reinforcement (0.6fy) and friction along the interface
caused by the normal force. In a testing program by Valluvan (1993),
specimens were constructed that applied a vertical force across the shear
plane, but did not include reinforcement across the plane. In this testing
program, a coefficient of friction of 0.8 was observed. Similar results were
obtained by Mattock and Hawkins (1972), and it is for this reason that the
alternative shear friction equation presented in the commentary Section
R11.6.3 of ACI 318-11 uses a coefficient of friction of 0.8. Therefore, shear
strength provided from the normal force may be considered to be 0.8fy. By
adding the two components, a total coefficient of friction of 1.4 is developed
which is consistent with the ACI 318 value.
ACI 318 currently limits the value of fy used for the design of shear-friction
reinforcement to 60,000 psi. This limit is primarily based on the fact that test
results are not available with higher yield strengths. Therefore, it is not
known if adequate separation will develop to induce sufficient strains to
produce yield in the reinforcement. In considering higher strength
reinforcement, it is important to determine if sufficient strains will be
developed. For the lower bound strength of 0.6fy for concrete placed against
hardened concrete not intentionally roughened, it appears that no
modification will be needed as this is controlled by the shear yield strength of
the reinforcement and that the full yield strength of the reinforcement will be
able to be utilized. Testing, however, will be required to justify if the full
yield strength of the reinforcement can be realized in producing the normal
force across the shear transfer plane.
Recommendation for Further Study

Further research is needed to fully evaluate interface shear strength, and a


proposed research program is outlined in 4.4.7.
3.1.2.4

Deep Beams

Deep beams require special consideration in ACI 318 due to their span-todepth ratio which results in the member being a discontinuity region where
plane-sections do not remain plane. In general, there are two approaches to
their design: (1) nonlinear finite element analysis which accounts for the
nonlinear distribution of longitudinal strain over the depth of the beam or (2)
strut-and-tie models.

3-16

3: Key Design Issues

ATC-115

95% Draft
In general, a deep beam acts as a tied arch with the longitudinal
reinforcement acting as a tension tie. Based on this force mechanism, higher
strength reinforcement should be able to be used effectively assuming that
the reinforcement is properly anchored (anchorage of the reinforcement will
become more critical with the use of high-strength reinforcement as
anchorage forces will increase). However, several of the design provisions
for deep beams are focused on crack control. First, the shear stress is deep
beams is limited to 10 f c' . This limit imposes a dimensional restriction to
control cracking under service loads while also protecting against diagonal
compression failures at ultimate. Second, distributed reinforcement is
required along the side faces and a minimum reinforcement ratio of 0.25% is
required both perpendicular and parallel to the axis of the beam.
Birrcher et al. (2013) conducted an experimental study to evaluate the
minimum web reinforcement requirement of deep beams. Twelve full-scale
tests were conducted in which the shear span-depth ratio (a/d) was 1.2, 1.85,
or 2.5. At each a/d, the quantity of web reinforcement was the primary
variable. Web reinforcement ratios ranged from 0 to 0.3% in the vertical and
horizontal directions. The compressive strength of then concrete in the test
specimens ranged from 3200 to 5000 psi. The results indicated that a larger
quantity of web reinforcement was needed to adequately restrain the width of
diagonal cracks than to provide adequate deep beam shear capacity. Based on
the strength and serviceability results, a minimum web reinforcement ratio of
0.3% in each orthogonal direction was recommended for deep beams. This
amount is slightly higher than that provided in ACI 318. While Grade 60
reinforcement was used in the testing program, it is expected that higher
yield reinforcement would not impact this minimum reinforcement amount.
Therefore, it does not appear that modification of the deep beam
requirements will be required for the use of high-strength reinforcement.
Recommendation

No further study is recommended or required, and a code change is not


considered necessary.
3.1.2.5

Ordinary Structural Walls

Ordinary structural walls are commonly used as all or part of a buildings


lateral-force-resisting system providing resistance to wind and earthquake
loads, if located in a region of low or moderate seismicity. For wind loads,
elastic response is intended. For earthquake loads, inelastic yielding is
anticipated because the design in accordance with ASCE 7-10 includes a
response modification factor, R, in the denominator of the earthquake load
equation that reduces the earthquake load below the magnitude required for
ATC-115

3: Key Design Issues

3-17

95% Draft
elastic behavior. R equals 4 if part of a bearing wall system and 5 if part of a
frame system.
The behavior and design of ordinary structural walls depends on the aspect
ratio of the wall with tall slender walls being flexure-dominated and short
squat walls being shear-dominated. The design of horizontal wall
reinforcement is typically controlled by the required strength to resist inplane shear forces. For slender walls, vertical reinforcement of ordinary
walls is typically controlled by the design for combined axial force and inplane flexure. For squat walls, the required vertical reinforcement can be
controlled by (1) the combined axial force and in-plane flexure, which is
more common for taller squat walls, or (2) the requirement for vertical
reinforcement to match or be a function of the horizontal reinforcement. The
combined axial and flexural strength is computed assuming plane sections
remain plane, with extreme fiber concrete compression strain equal to 0.003.
For walls with aspect ratios less than 2.5, minimum requirements for vertical
reinforcement relative to horizontal reinforcement ratios reflect the
participation of vertical reinforcement in shear resistance.
Shear in ordinary structural walls is designed in accordance with Sections
11.5.4, 11.6, and 11.7 of ACI 318-14. Nominal shear strength calculations
include both concrete and reinforcement contributions to resistance, with a
combined limit equal to 10 fc' hd. In this equation, h is the wall thickness
and d is the flexural depth, often taken as 80% of the wall length. The
concrete contribution to shear strength, Vc, can be computed with either the
simplified formula of 2 fc' hd or the more detailed formula in Table
11.5.4.6, which includes the interactive effects of shear with both axial and
flexural forces.
Ordinary structural walls are commonly included in tall buildings with
height-to-length aspect ratios typically more than 3-to-1. Walls with aspect
ratios of 3 and greater are considered slender behaving according to beam
theory. Although most of these walls are 12 to 18 in. thick reinforced with
conventional bar sizes, some of these walls are now as much as 30 in. thick
reinforced with No. 18 and larger high-strength reinforcing bars and concrete
with strength up to 15,000 psi. Although ordinary slender structural walls
may experience yielding and nonlinear response, the ends of the walls are not
required to have transverse confining reinforcement, which provides stable
post-yield concrete compressive strength as well as buckling restraint for
longitudinal reinforcement. This may be
Issues with the use of high-strength reinforcement in structural walls include:

3-18

3: Key Design Issues

ATC-115

95% Draft

ACI 318-14 limits the yield stress used for the design of shear
reinforcement to 60 ksi. This is intended to control service level
cracking.

Grade 100 and 120 reinforcement yields in compression at a strain


greater than 0.003, assumed to be the limit state for concrete crushing.

The higher bond stresses for high-strength reinforcement are more likely
to split the cover concrete at splices.

The behavior of walls with No. 18 and larger bars with high-strength
concrete may be different than walls with smaller bars and more
conventional strength concrete.

Recommendations for Ordinary Structural Wall Studies

Tests of ordinary structural walls with high-strength reinforcement are


recommended. Such tests are described in Section 4.9.8.
3.1.3

Strut and Tie Modeling

ACI 318 allows for members to be designed using strut-and-tie models.


While the approach is available for members in general, the approach is most
commonly used for the design of discontinuous regions which are also
known as D-regions. Strut-and-tie modeling is a plasticity approach;
therefore, it provides a safe solution for the load capacity of the structure.
This approach, however, does not address serviceability issues such as
deflections or cracking (Schlaich et al. 1987). Deflections are typically
estimated using elastic analysis to analyze the strut-and-tie model. Crack
widths are typically controlled as follows: 1) crack widths of ties are
controlled using the crack control provisions for flexural members and 2)
crack widths within the member that cross struts are controlled through
distributed reinforcement (0.3% reinforcement ratio) provided across the
strut.
Considering that strut-and-tie modeling is a plasticity approach, its use
remains applicable for high-strength reinforcement. However, it is important
that the reinforcement provide adequate ductility so that the strut-and-tie
approach remains applicable. With the use of higher-strength reinforcement,
it is expected that deflections and crack widths will increase. Therefore, it
will become more important that these serviceability issues are addressed in
design.
Recommendation

ATC-115

3: Key Design Issues

3-19

95% Draft
Assuming that adequate reinforcement ductility is provided, further testing
does not seem warranted, and a code change is not considered necessary.
3.1.4

Other Strength Issues

Several other miscellaneous issues should also be considered before adopting


high-strength reinforcement into ACI 318. These include establishing
appropriate concrete strengths to use, whether higher bar stresses may be
used for tension regions of shell and folded plates, and whether the higher
yield strengths can be used when computing required quantities of bonded
reinforcement for members constructed with unbonded post-tensioning
tendons.
3.1.4.1

Concrete Strength

It is advantageous to use high-strength concrete in members that contain


high-strength reinforcement. High concrete strength reduces the required
development and splice lengths of reinforcement, improves deformation
capacity of flexural members, increases the shear strength of members,
improves the strength of columns with high axial loads and those with
combined axial load and flexure, increases the shear strength of joints of
special moment frames, and reduces deflections. Existing literature includes
tests combining high-strength concrete with high-strength reinforcement
(e.g., Aoyama, 2001; Okamoto et al., 2004; Nishiyama, 2009; Restrepo et al.,
2006).
Use of high-strength concrete improves the deformation capacity of flexural
members. Consider two beams close to their ultimate deformation capacity
with the only difference being that one has higher strength concrete than the
other. The beam with the higher strength concrete will have a shallower
neutral axis depth, slightly increased peak moment strength, higher
curvature, higher bar tensile strain, and higher hinge rotation.
Similarly, higher strength concrete improves the strength of columns with
combined axial load and flexure because a smaller Whitney stress block
supports the same compressive force. For the same axial load, the smaller
Whitney stress block for the column with higher strength concrete increases
the distance between the tension and compression couple resulting in higher
flexural strength.
Raising the minimum concrete compressive strengths allowed in Section
19.2.1.1 of ACI 318, e.g. 2500 psi for members except those of special
moment frames and special shear walls, which require a minimum concrete
compressive strength of 3000 psi, is likely not necessary for members

3-20

3: Key Design Issues

ATC-115

95% Draft
designed and constructed using high-strength reinforcement, and could be
left to ACI Committee opinion to determine whether a change is required.
However, it may not be practical to construct such members without the
higher strength concrete.
Exploring the use of high-strength concrete in members reinforced with highstrength bar is beyond the scope of this roadmap. However, some testing
with concrete compressive strengths on the order of 8000 to 15,000 psi are
appropriate as strengths this high are becoming more common and are likely
to become even more common in the future. One reason that higher strength
concrete is being used in slabs and beams is to control deflections.
Recommendation Regarding Concrete Strength

A test program specifically targeted to explore the combination of highstrength concrete with high-strength reinforcement is not required; rather the
use of higher strength concrete should be incorporated into the majority of
the test programs identified in Chapter 4. ACI Committee 318 can decide
whether to increase the minimum concrete strength without additional
testing.
3.1.4.2

Tension Regions of Shells and Folded Plates

Shells and folded plates are configured to take advantage of the concrete
acting in compression so tension regions are often minimized. An example of
a tension region of a dome is the circumferential direction for portions of the
dome below 52 degrees measured from a vertical axis passing through the
center of the dome. An example of a tension element is a ring beam at the
base of the dome.
Using a smaller area of higher-strength reinforcement in tension regions and
elements will increase the strain of the reinforcement, increase the level of
concrete cracking, and reduce the stiffness. The increased cracking and
reduced stiffness may be objectionable, thus limiting the use of high-strength
reinforcement for tension regions of shells and folded plates. The effect of
the reduced stiffness and whether the reduction in stiffness is acceptable from
a structural standpoint requires investigation. In addition, shells and folded
plates are often used as roof coverings without a roofing membrane, and
increased cracking might lead to less durability if the cracks are wide.
Recommendation for Tensile Regions of Folded Plates and Shells

Investigating structural concerns could be addressed with a combination of


analysis and testing. The analysis would consist of finite element analysis
that accounts for the effects of the reduced stiffness. The level of stiffness

ATC-115

3: Key Design Issues

3-21

95% Draft
reduction and the impact of cracking could be assessed with tension tests on
panels reinforced with high-strength reinforcement. The analysis and test
program is described in Section 4.4.4.
3.1.4.3

Bonded Reinforcement Ratios for Members with


Unbonded Post-Tensioning

Minimum bonded reinforcement is required by ACI 318 in prestressed beams


with unbonded tendons. According to the ACI 318 commentary, this
reinforcement is provided to ensure flexural behavior at ultimate beam
strength, rather than tied arch behavior, and to limit crack width and spacing
at service load when concrete tensile stresses exceed the modulus of
rupture. The commentary further indicates that Providing minimum
bonded reinforcement helps to ensure acceptable behavior at all loading
stages. This required reinforcement is based on research by Mattock et al.
(1971) that compared the behavior of bonded and unbonded post-tensioned
beams and is independent of the yield strength of the reinforcement.
Considering the basis of this requirement, it does not seem necessary to alter
this minimum requirement when high-strength reinforcement is used.
Recommendation

It is recommended that no further study is needed for the use of high-strength


reinforcement, and a code change is not required.
3.1.4.4

High-Cycle Elastic Fatigue of Reinforcement

There is a potential for reduced high-cycle elastic fatigue resistance of highstrength reinforcing bar relative to that of Grade 60 bar. ACI Committee 215
report (ACI, 1997) addresses fatigue of reinforced-concrete members.
Although fatigue resistance is generally not a concern for building structures,
this is an important material issue for bridges and perhaps select non-building
structures. Testing for fatigue resistance is not needed for adoption of highstrength reinforcement into ACI 318 but it is required for adoption into
AASHTO LRFD Bridge Design Specifications. Research that supports the
AASHTO design provisions for elastic fatigue of steel reinforcing bars is
reported in NCHRP Report 164 (Helgason, et al., 1976). For high-strength
reinforcing bar of Grade 100 and 120, testing is likely required for bars
manufactured by various processes such as (1) micro-alloying and (2)
quenching and self-tempering. It would also be required for various bar
deformation arrangements. For example, higher relative rib area bars may be
more susceptible to fatigue, although no problems were observed for Grade
60 high relative rib area bars (Fei and Darwin, 1999). The high-cycle elastic
fatigue resistance is likely a function of the radius at the base of the
deformation and defects at this location. Mechanical splicing of high3-22

3: Key Design Issues

ATC-115

95% Draft
strength bar might also affect the elastic fatigue resistance. Fatigue of
mechanical splices was studied as part of NCHRP Project 10-35 (Paulson
and Hanson, 1991). Testing for high-cycle fatigue resistance of reinforcedconcrete members is something that the Transportation Research Board
(TRB) of the National Academies and the Federal Highway Administration
(FHWA) should consider sponsoring as a project under the National
Cooperative Highway Research Program (NCHRP).
Recommendation to Study High-Cycle Elastic Fatigue

Perform tests in accordance with Section 4.4.8 to determine the high-cycle


elastic fatigue resistance of members reinforced with high-strength
reinforcement.
3.2

Serviceability

Serviceability performance of a structure is an essential complement to the


ultimate load-carrying capacity. The serviceability limit states include
deflections, lateral drifts, crack control, and vibrations. These limit states are
intended to avoid disruption of the intended use of the structure.
3.2.1

Deflections

ACI provisions for deflections, specific to flexural members, are intended to


provide adequate stiffness to limit deflections that adversely affect
serviceability and strength of a structure. Deflection limits are provided as a
fraction of the member span, based on use (floor or roof) and brittleness of
supported materials. Member span-to-depth limits are provided with the
intent of automatically satisfying these deflection limits. Methods of
estimating member stiffness, and hence deflections, are also provided for
instances when span-to-depth requirements are not met. Finally, estimates
for additional long-term deflections, commonly referred to as creep
deflections, are provided by ACI 318-14.
Some users of ACI 318 have questioned how deflections can be controlled
by member depth as a proportion of span length alone, without considering
the magnitude of the load. The best description is considering it in terms of
allowable stress design. One-way slab and beam depth limits were first
introduced in ACI 318-63, and were intended for allowable stress design
using Grade 40 reinforcement. Deflections are directly related to curvature.
Assuming plane sections remain plane and concrete cracks, curvature equals
the change in strain divided by the depth. The change in strain varied over a
small range because: (1) the concrete stress was limited to 0.45f'c; (2) the
variation in concrete strengths typically specified was narrow at that time,

ATC-115

3: Key Design Issues

3-23

95% Draft
about 3,000 psi to 5,000 psi; and (3) the reinforcement stress and thus strain
was essentially constant for all members (i.e. the stress limit of 20 ksi was
used to design most members). The change in strain varied between
approximately 0.001 and 0.0015, regardless of the magnitude of the load.
Members with greater design loads required proportionally greater depth.
The concept of using depth as a proportion of span length to control
deflection is less easily explained in terms of strength design. The change in
strain, thus curvature, is greater, and not all members have the same
reinforcement stress at service level loads. The greater range of stress and
strain makes the use of depth only to control deflection less valid, but not
excessively so. In ACI 318-63, the tables could only be used for designs in
accordance with strength design if the net reinforcing ratio was less than
0.18fc/fy and the reinforcement was Grade 40 or less.
In ACI 318-71 (ACI, 1971a), the table for minimum beam and one-way slab
depths increased the required depths by about 25% and a footnote was added
that required adjustments to depths for members with specified yield
strengths not equal to 60 ksi. The adjustment for Grade 80 reinforcement
required 20% greater depth, and the adjustment for Grade 40 reinforcement
reduced the depth to be essentially the same as in ACI 318-63. The required
depths and footnote have remained essentially the same in all subsequent
editions of ACI 318 up to and including the ACI 318-14 (Table 7.3.1.1 for
one-way slabs and Table 9.3.1.1 for beams) even though reductions in load
factors have increased service level reinforcement stresses leading to greater
curvatures.
In lieu of using the minimum depth limits for controlling deflections,
deflections can be computed and compared to allowable deflections.
Equation 24.2.3.5a of ACI 318-14, repeated here as Equation 3-1, is
permitted to be used as the effective moment of inertia for computing
immediate deflections. This equation was developed by Branson (1977) and
Branson and Trost (1982) and appears in ACI 435R-95, Control of
Deflection in Concrete Structures (ACI, 1995).

M 3
M cr 3
Ie (
) I g 1 cr I cr
Ma
M a

(3-1)

Equation 3-1 yields a single effective moment of inertia which was calibrated
to represent the entire length of a simply supported flexural member
containing both uncracked and cracked regions and allow for deflection
calculation using elastic beam equations. For continuous spans, the effective
moment of inertia used for the entire span is typically taken as the average of

3-24

3: Key Design Issues

ATC-115

95% Draft
the moment of inertia for the negative moment region(s) and the positive
moment region.
The current methodology for controlling or computing deflections presents
several questions when using high-strength reinforcement:

Are the minimum span-to-depth tables of ACI 318 still appropriate?

Is the effective moment of inertia equation (24.2.3.5a shown here as Eq.


3-1) from ACI 318-14 valid with high-strength reinforcement?

Is the time-dependent deflection factor (24.2.4.1.1) in ACI 318-14


which is not reflective of reinforcement yield strength valid for highstrength reinforcement?

With the use of high-strength reinforcement, it is expected that the amount of


required flexural reinforcement will decrease resulting in lower
reinforcement ratios . The current design tables provide a multiplier,
0.4 f y / 100,000 to account for Grades other than 60; however, this
expression is currently limited to Grade 80 reinforcement. It is expected that
flexural members designed using high-strength reinforcement will be
subjected to higher reinforcement stresses and greater curvatures. Therefore,
an extension of the current approach may not be valid and requires further
investigation.
While Equation 3-1 accounts for the reinforcement ratio through Icr, the
overall expression also considers tension stiffening which is the blending of
Icr and Ig to provide an effective stiffness somewhere between these two
values. For low reinforcement ratios, it has been shown that the Bransons
equation overestimates tension stiffening (Bischoff 2007). Therefore, an
alternative expression has been developed (Bischoff 2007, Bischoff and
Scanlon 2007) to more accurately account for this effect. The equation
developed by Bischoff (Equation 3-2) has been shown to provide much more
accurate results, even with very low effective reinforcement ratios that are
produced by beams reinforced with fiber reinforced polymer (FRP) bars. For
this reason, ACI 440 has adopted this equation for the estimation of
deflection of FRP reinforcement members and will appear in the next version
of the report (ACI 440.1R). Specifically for high strength steel, ACI ITG6R-10, also presents Equation 3-2 as an appropriate approach, and this
expression will also be included in the next version of the ACI 435 report on
control of deflections. Because of the issues related to the use of low
reinforcement ratios with high-strength reinforcement, it is expected that a
change from the current 318 effective stiffness equation will be required, and
it appears that the Bischoff expression will be suitable as its replacement.

ATC-115

3: Key Design Issues

3-25

95% Draft


Ie

I cr
2
M cr

1
Ma

(3-2)

where:
1

I cr
Ig

Regardless of the procedure used to estimate deflections (Equation 3-1 or 32), deflection calculations are dependent on the computation of the cracking
moment Mcr. While the cracking moment is typically computed using a
modulus of rupture of 7.5 f c' , studies have indicated that improved
deflection calculations are possible using a reduced value as low as half the
typically computed value. Lower values have been considered to be
necessary to account for shrinkage as well as restraint which can reduce the
cracking moment of a flexural member. In some research, the use of a lower
modulus of rupture has been provided to offset the over-estimation of tension
stiffening that results from the use of the Branson equation.
The appropriateness of the long-term deflection factor applied to the
immediate deflection of members reinforced with higher grade reinforcement
should also be considered. Because the primary factor influencing this factor
is creep of the concrete, it is not felt that changes in this factor will be
required. However, with the use of potentially lower reinforcement ratios
made possible through the use of high-strength reinforcement, higher
curvatures will be possible which potentially may increase stresses in the
compression zone leading to higher overall creep deflections. It seems
appropriate that long-term deflections be evaluated to ensure appropriateness
of the current time-dependent factor approach.
While ACI 318 provides requirements for the calculation and control of
flexural deflections, it is silent on the axial shortening of compression
members. Although often taken simply as PL/AE, the interaction of concrete
and reinforcing bars strained equally but with different yield strains creates a
more complex behavior. Furthermore, the time-dependent creep strain of
column concrete is not addressed in ACI 318. The creep time-dependent
factor for flexural members (24.2.4.1.1) is often used for compression
members for lack of other guidance. Improved understanding regarding the
shortening of compression members seems warranted, especially with the use
of high-strength reinforcement. High-strength reinforcement will make it

3-26

3: Key Design Issues

ATC-115

95% Draft
possible that column cross-sections will be reduced which has the potential to
increase axial shortening, especially long-term shortening.
Recommendation for Studies of Deflections of Flexural Members

An analytical study is recommended to determine whether the deflection


control or computational methods require revision for flexural members
reinforced with high-strength reinforcement. The analytical models should
be compared to tests of beams performed for other purposes, such as the tests
of beams described in Section 4.4.1 that are being tested for adequate
ductility and deflection warning.
3.2.2

Drift

Acceleration and displacement response of structures due to service level


lateral loads can affect the comfort of building occupants and damage
finishes, if excessive. In general, this is only a concern for taller reinforcedconcrete buildings loaded by wind because the service level earthquake
occurs too infrequently to be a concern. Drift is computed as part of the
structural analysis and is influenced primarily through the selection of the
flexural stiffness of the members. With the use of high-strength
reinforcement, a reduction in the flexural stiffness of the members is
expected not only at factored loads, but also service loads, which can
influence drift.
Recommendation

Detailed discussion on the influence of high-strength reinforcement on


flexural stiffness is provided in 3.5.1.
3.2.3

Crack Control

Since 1971, ACI 318 has required the control of crack widths for flexural
members. Explicit calculation to evaluate crack widths was added due to the
introduction of Grade 60 reinforcement. At the time, the z-factor method
was incorporated, which was based on a crack width calculation equation
developed by Gergely and Lutz (1968). The test results from which this
equation was derived focused on service stresses that may develop using
Grade 60 reinforcement, and the maximum stress considered was 50 ksi.
With the introduction of ACI 318-99, the approach used to control cracking
was changed and is based on a physical model developed by Frosch (1999).
The physical model considers bar strains; therefore, it is expected that this
approach is applicable for high-strength reinforcement (Frosch 2001). It
should be noted, however, that the maximum stress considered in the
evaluation of the model was 50 ksi.

ATC-115

3: Key Design Issues

3-27

95% Draft
Cracking along the side faces of beams also requires control. The approach
used by ACI is consistent with that used along the tension face and is based
upon the same physical model (Frosch 2002). Therefore, it is expected that
this approach is also applicable for high-strength reinforcement. It should be
noted, however, that the maximum stress considered in the evaluating the
accuracy of the model was 40 ksi. Therefore, evaluating the expression with
higher reinforcement stresses is warranted.
The control of cracking is expected to become even more important with the
use of higher strength reinforcement. Therefore, it is important that the
physical model from which the ACI expression for control of flexural
cracking is derived be evaluated considering test results from higher stress
ranges. In addition, it will likely be desirable to reformat the ACI design
expression so that either the default reinforcement service stress (40 ksi) for
which the equation was developed changes or the expression is reformatted
so that the equation is not presented based on a default stress. Overall, it is
expected that increased services stresses will require reduced bar spacings,
which can be achieved with the same amount of reinforcement through
decreasing the bar size.
Crack control is also required in ACI 318 for members designed using strutand-tie modeling. Specifically, a minimum of 0.3% distributed
reinforcement that crosses the strut axis is required. According to ACI
Committee 318, This reinforcement will help control cracking in a bottleshaped strut and result in a larger strut capacity than if this distributed
reinforcement was not included.
As previously mentioned in Section 3.1.2.4, Birrcher et al. (2013) conducted
twelve full-scale tests to evaluate the minimum web reinforcement
requirements for deep beams. While this study was focused on deep beams,
the results can be considered appropriate for considering minimum
distributed reinforcement for strut-and-tie modeling. The results indicated
that a larger amount of web reinforcement is needed to control diagonal
crack widths than to provide shear capacity. This study concluded based on
both strength and serviceability that a minimum reinforcement ratio of 0.3%
in each orthogonal direction is appropriate, which is consistent with the
amount of reinforcement currently specified for distributed reinforcement
when strut-and-tie models are used. While Grade 60 reinforcement was used
in the testing program, it is expected that higher yield reinforcement would
not impact this minimum reinforcement amount.
Recommendation for Further Study

3-28

3: Key Design Issues

ATC-115

95% Draft
Further research is recommended to evaluate cracking behavior in flexural
members when high-strength reinforcement is utilized, and a proposed
research program is outlined in 4.5.2. Considering crack control for
structures design using strut-and-tie models, the current minimum
reinforcement amounts are considered appropriate for use with high-strength
reinforcement, and no further research is deemed necessary. Furthermore, a
code change of this minimum requirement is not required.
3.3

Reinforcement Limits

Reinforcement limits for concrete members are based on objectives of


strength, serviceability, and constructability. The influence of these three
objectives is handled differently for each primary member type discussed.
3.3.1

Beams

Current ACI 318 provisions for minimum reinforcement in flexural members


are intended to provide cracked section flexural strength greater than
uncracked section flexural strength. This is a critical limit state, which
protects against sudden collapse of flexural members in the event of loading
beyond cracked section strength.
The current As,min equation in Section 9.6.1.2 of ACI 318-14 already
addresses reinforcement with variable yield strength:

3 f 'c bw d

(3-1)

fy

However, this equation is only applicable to rectangular beams and was


intended to be conservative for rectangular beams. ACI 318 is expected to
eliminate this equation in a future edition, and directly specify a minimum
value for the ratio Mn/Mcr ratio, where Mn is the design flexural strength
and Mcr is the theoretical moment that cracks the section. This would
resemble current requirements for prestressed members in which the cracked
section flexural capacity is 120% of the uncracked flexural capacity,
(Sections 7.6.2.1, 8.6.2.2, 9.6.2.1 of ACI 318-14). These generalized
requirements inherently include and consider high strength reinforcement in
the limit state. No strength-based testing would be required for high strength
reinforcement in this application should this code change take place.
ACI Committee 318 may decide that entirely eliminating direct solutions
within ACI 318 is undesirable, in which case verifying the use of the existing
approach in Section 9.6.1.2 of ACI 318-14, or a modified version of it, may
be required.

ATC-115

3: Key Design Issues

3-29

95% Draft

Recommendations for Study of Beam Minimum Flexural Reinforcement

The need for analytical studies of beam minimum reinforcement is pending


ACI 318 modifications. If Equation 3-1 for minimum reinforcement is
replaced with provisions that directly address compare design strength to
cracking moment no further study is needed. If Equation 3-1 remains either
as the base requirement or as a simplified option, analytical studies will be
required as described in Section 4.6.1.
3.3.2

Slabs and Footings

ACI 318-14 provisions exempt slabs and footings of uniform thickness from
satisfying strength-based minimum flexural reinforcement requirements
intended to ensure that Mn,cracked > Mn,uncracked. This exemption is a
committee-based decision reflecting the perceived redistribution capacity of
these elements. Instead, the minimum reinforcement for slabs and footings is
based on shrinkage and temperature requirements, but with the requirement
that this reinforcement be provided entirely on the tension face of the
member.
The ACI 318 shrinkage and temperature provisions for slabs, Section
(7.12.2.1), already account for reinforcement with yield strength that exceeds
60 ksi. The minimum reinforcement area to gross concrete area is as
follows:
0.0018 x60
fy

(3-2)

where fy is in ksi.
The yield strength in this formula is based on the extension under load
method for a strain of 0.35 percent. This requirement on how the yield
strength is measured recognizes that reinforcement that behaves nonlinearly
prior to reaching yield does not provide as much resistance to crack control.
ACI 318 eliminated this approach to compute yield strength and went solely
to the 0.2% offset method for computing the yield strength of high-strength
reinforcement. The existing provision for minimum slab reinforcement will
have to be revisited and revised in consideration that yield strength is no
longer required to be measured using the extension under load method,
which results in no means of separating reinforcement with and without
nonlinear behavior prior to the defined yield strength using the 0.2% offset
method.

3-30

3: Key Design Issues

ATC-115

95% Draft

Recommendation for Reassessment of Minimum Flexural


Reinforcement for Slabs and Footings

The incorporation of strength-based reinforcement limits for slabs and


footings is pending ACI 318 decision. Until such a change in approach is
made, no additional testing is required. If a strength approach is
incorporated, analytical studies such as those in Sections 4.6.1 may be
required.
3.3.3

Columns

ACI 318 requires a minimum amount of reinforcement in all columns. The


minimum amount is specified as 1% based on the gross section, and this
minimum has been constant going back to the 1936 edition of the code, ACI
501-36T, Building Regulations for Reinforced Concrete (ACI, 1936). The
commentary to ACI 318-11 indicates that Reinforcement is necessary to
provide resistance to bending, which may exist regardless of analytical
results, and to reduce the effects of creep and shrinkage of the concrete under
sustained compressive stresses. Creep and shrinkage tend to transfer load
from the concrete to the reinforcement, and the resultant increase in
reinforcement stress becomes greater as the reinforcement ratio decreases.
Therefore, a minimum limit is placed on the reinforcement ratio to prevent
reinforcement from yielding under sustained service loads (Richart, 1933).
It should be noted that it is possible for a column to have less than 1%
reinforcement. When a column is oversized due to architectural
considerations, the code permits consideration of a reduced area section, but
not less than one-half the total area. Consequently, reinforcement ratios as
low as 0.5% are possible.
A research project was conducted by Ziehl et al. (1998; 2004) to assess the
feasibility of reducing the current ACI 318 and AASHTO Bridge
Specification requirements for minimum longitudinal reinforcing steel in
columns. Considering that the current code requirement is based primarily
on research conducted in the 1920s and 1930s on concrete with a
compressive strength generally less than 5000 psi, the investigation was
conducted to determine the effects of present-day construction materials on
minimum reinforcement requirements. The research program included: (1)
fabrication and long-term loading and monitoring of 24, 8-inch diameter by
4-ft long reinforced concrete column specimens; (2) fabrication of reducedhumidity enclosures for storage of all specimens throughout the test program;
and (3) long-term monitoring of 14 unloaded companion specimens. Test
variables included concrete strength and reinforcement ratio, and Grade 60
reinforcement was used in all specimens. All loaded specimens were

ATC-115

3: Key Design Issues

3-31

95% Draft
subjected to a nominal compressive force of 0.40 f c' /Ag. It was found that
for reinforcement ratios ranging from 0 to 0.72%, all of the specimens with
high-strength concrete and most with normal strength concrete experienced
time-dependent axial strains that exceed the yield strain of the reinforcement
(0.00207). Therefore, it was concluded that a reduction of the current
ACI 318 requirement for minimum longitudinal reinforcement is not justified
because passive yielding, based on the nominal yield strength of the
longitudinal reinforcement, occurred in all high-strength specimens, and
would likely have occurred in most of the other specimens had loading
continued for another year.
Considering that the yield strain of higher strength reinforcement will be
higher, there is the potential that the minimum reinforcement ratio can be
decreased. However, the economic benefit of reductions below 1% may not
be significant. Furthermore, decreased reinforcement ratios likely will
increase creep deformation and has the potential to increase column
shortening. Overall, it seems appropriate to maintain the current minimum
reinforcement ratio.
Recommendation

No additional research is recommended, and a code change is not considered


needed.
3.3.4

Walls

Reinforcement requirements for walls in ACI 318 reflect minimum ratios


dependent upon fy as a step function. One limit exists for reinforcement
having fy less than 60 ksi, and a smaller minimum ratio is allowed for
reinforcement having fy not less than 60 ksi. Although this trend could be
extrapolated to much higher strength reinforcement and corresponding lower
reinforcement ratios, the current limits are deemed sufficiently low.
As discussed by Wood (1989), very low reinforcement ratios, particularly at
the base of buildings, can lead to a localization of strain in longitudinal
reinforcement when walls crack. Brittle failures may occur at low drift
ratios, particularly when the cracked flexural capacity of the wall is less than
its uncracked flexural capacity (Davey and Blaikie, 2005). Such failures are
possible when a single flexural crack occurs, and high-strain low-cycle
fatigue initiates longitudinal bar fracture.
Recommendations for Minimum Reinforcement in Ordinary Walls

At this time no analytical or experimental tests are recommended which


focus on the minimum reinforcement ratio to use for high-strength

3-32

3: Key Design Issues

ATC-115

95% Draft
reinforcement for ordinary walls. Current minimum reinforcement ratios for
walls are deemed sufficiently low.
3.4

Detailing and Other Design Considerations

Proper location and arrangement of reinforcement and member connectivity


(joints) are essential for obtaining adequate strength and ductility in
reinforced concrete structures. Detailing requirements include bar bending,
bar spacing, development and splice lengths of straight bars, development
lengths of hooked and headed bars, continuity and termination of bars, and
concrete cover for reinforcement. Structural detailing and member
connectivity need to be examined when using high-strength steel bars as
concrete reinforcement.
3.4.1

Bend Test Requirements and Minimum Bend Diameters

Reinforcing bars produced according to ASTM specifications must conform


to specific bend test requirements. The bending requirements in ASTM
Specifications define the degree of bending and the pin diameters about
which reinforcing bars must be bent without cracking on the outside of the
bent portion.
The minimum percentages of elongation at fracture in an 8 inch gauge length
required by ASTM A615, A706, and A1035 are summarized in Table 3-1,
and the minimum bend diameters of hooks required by ACI 318, and the pin
diameters for bend tests required by the ASTM specifications, are
summarized in Table 3-2.
Kudder and Gustafson (1983) reported the results of 945 bend tests and 540
tension tests performed on five sizes of Grade 60 reinforcing bars. The bend
test data showed that for bars with a percentage of elongation at least equal to
the ASTM minimum, the probability of a bend test failure was very small
(0.02 percent or less). They concluded that there is a very low likelihood of
a failure occurring at a bend diameter equal to, or larger than the minimum
bend diameter required by the ACI Building Code.
Stecich et al. (1984) reported the results of bending and straightening tests
conducted on 254 Grade 60 reinforcing bars. The major variables were bend
diameter, bar size and supplier, bend axis and temperature. Three bar sizes
were included in the tests, No. 5, No. 8, and No. 11. They observed that the
smaller bars performed better and that breakage and cracking is more likely
to occur when straightening is done at cold temperatures. They concluded
that field bending and straightening of reinforcing bars up to No. 11 should
be permitted as long as the bars are bent about the weak axis up to 90 degrees

ATC-115

3: Key Design Issues

3-33

95% Draft
using the minimum bending bar diameters of ACI 318 and straightened at
normal temperatures.
Statistical analysis of bend test data is needed for high-strength steel bars. It
is important to mention that the current use of ASTM A1035 Grade 100 and
Grade 120 steel bars as transverse reinforcement adopts the minimum bend
diameters commonly used for Grade 60 reinforcement.
Table 3-1

Minimum Percentage of Fracture Elongation for ASTM A615,


A706, and A1035 Reinforcing Bars
A615

A615

A615

A706

A706

A1035

A1035

Grade
40

Grade
60

Grade
75

Grade
60

Grade
80

Grade
100

Grade
120

11*

14

12

4 and 5

12

14

12

14

12

7 and 8

12

12

9 to 11

12

12

14 and
18

10

10

Bar Size,
No.

NOTES: * All values correspond to percentage of elongation in 8-in. gauge length

Table 3-2

Minimum Bend Diameter of Hooks in ACI 318 and Bend Test Requirements for ASTM
A615, A706, and A1035 Reinforcing Bars
Minimum Bend
Diameter of Hooks
per ACI 318-14*

Bar Size,
No.

Pin Diameter for Bend Test


per ASTM Specification

Stirrups,
A615
A615
A615
A706
A706
Ties, and All Other
Hoops
Grade 40 Grade 60 Grade 75 Grade 60 Grade 80
Bars

A1035

A1035

Grade
100

Grade
120

4db**

6db

3db

3db

3db

3db

3db

3db

6db

6db

5db

5db

5db

4db

5db

5db

5db

7 and 8

6db

6db

5db

5db

4db

5db

5db

5db

9 to 11

8db

7db

7db

6db

7db

7db

7db

14 and 18

10db

9db

9db

8db

9db

9db

9db

3, 4, and 5

NOTES: * ACI 318-14 allows up to Grade 100 for stirrups, ties, and hoops, all other bars are limited to Grade
80.
** db: nominal bar diameter

3-34

3: Key Design Issues

ATC-115

95% Draft

Recommendation

Collection of test data, performing tests, and statistical analyses of test data
of the mechanical properties of high-strength reinforcing bars, including data
from bend tests, are described in Section 4.3.2
3.4.2

Transverse Reinforcement Spacing

Transverse reinforcement serves different purposes for various members and


systems. For the purposes of this report the transverse reinforcement spacing
requirements are addressed separately for (1) members that are designed for
strength to resist only gravity and wind loads or ordinary seismic systems
and (2) members of intermediate and special seismic systems and members
in SDCs D, E and F that are not specifically designated as part of the
seismic-force-resisting system.
3.4.2.1

Spacing of Transverse Reinforcement in Members


Resisting Gravity and Wind Loads or in Ordinary
Seismic-Force-Resisting Systems

Transverse reinforcement spacing can be controlled by shear strength


requirements, shear reinforcement spacing requirements, requirements for
spirals to provide confinement to the column core, or detailing requirements
for spirals or ties. This section focuses on the detailing requirements for
spirals and ties. The spacing for shear and confinement of the core are not
considered in this section.
The design of spirally reinforced columns is based on the assumption that
when the cover spalls, the core concrete will be confined so it will maintain
strength at least equal to the strength prior to concrete cover spalling. The
spirals must also prevent the longitudinal bars from buckling to meet this
intent. The maximum clear spacing of 3 inches for spiral reinforcement in
accordance with Section 25.7.3.1 of ACI 318-14 is likely adequate for bars
with strength up to 120 ksi for gravity and wind loading. For comparison
purposes, the recommendation of NIST GCR 14-917-30 for hoop spacing of
longitudinal bars of Grade 100 and 120 is 4db, where db is the diameter of the
longitudinal bar. This would result in 3-inch center-to-center spacing for a
No. 6 bar. Given that the bars are not cyclically loaded, the 3-inch maximum
clear spacing requirement for spirals is likely adequate for supporting Grade
120 longitudinal bars.
The center-to-center spacing required for column ties placed in accordance
with Section 25.7.2.1 of ACI 318-14 is no greater than the minimum of 16db
where db is the smallest sized longitudinal bar in the column, 48db where db is
the diameter of the tie bar, and the smallest dimension of the column section.

ATC-115

3: Key Design Issues

3-35

95% Draft
These tie spacing requirements have been unchanged since the 1936 version
of the code, ACI 501-36T. The commentary to ACI 318-14 does not provide
the basis for the spacing of transverse reinforcement of tied columns; instead,
it appears to be based on committee opinion. An explanation for tie
requirements that are close to the requirements in ACI 318-14 appears in
NACU (1910), which states that longitudinal bars need to be sufficiently held
in place until the concrete is set.
The commentary to ACI 318-14 indicates that Pfister (1964) performed tests
that showed that there was no appreciable strength difference between
columns with and without ties at the tie spacing used. Following this
research, the tie spacing did not change but a change increased the required
diameter of the ties and related it to the longitudinal bar size and the
requirement that each longitudinal bar be supported by transverse
reinforcement was relaxed in ACI 318-71. Bars within 6 inches of laterally
supported bars no longer had to be supported by transverse reinforcement.
The requirements for ties in ACI 318-14 are the same as those in ACI
318-71.
The maximum allowed transverse reinforcement spacing for columns with
high-strength longitudinal reinforcement likely does not require change for
high-strength reinforcement though there could be some desire to reduce the
spacing.
Consideration should also be given to imposing new spacing requirements
for transverse bars at the end regions of ordinary walls and ordinary moment
frame beams and columns.
Recommendation for Studies of Transverse Reinforcement Spacing of
Members Resisting Gravity and Wind Loads or Part of an Ordinary
Seismic-Force-Resisting System

The requirements for tie spacing to use with longitudinal bars of Grades 100
and 120 could be based on committee opinion including those for end
regions of members of ordinary seismic systems.
3.4.2.2

Transverse Reinforcement Spacing for Members of


Intermediate and Special Seismic-Force-Resisting
Systems

Members of intermediate and special seismic systems and members not


specifically designated as part of the seismic-force-resisting system but in
structures assigned to SDCs D, E and F require transverse reinforcement to
maintain an appropriate level of ductility when the members are loaded
repeatedly beyond yield. The transverse reinforcement in the form of hoops

3-36

3: Key Design Issues

ATC-115

95% Draft
or spirals provides the ductility by confining the concrete and providing
resistance to longitudinal bar buckling.
The use of high-strength reinforcement to confine concrete has been studied
(Budek, et al., 2002; Muguruma, et al., 1990; Sugano, et al., 1990), and ACI
318-14 allows the use of Grade 100 reinforcement to provide confinement of
columns and wall boundary elements for special seismic systems. Grade 100
is also allowed for steel spirals providing confinement of columns. The
existing research could be reevaluated to determine whether Grade 120
reinforcement could be used to confine concrete columns and wall boundary
elements and as spiral reinforcement. If necessary, additional tests could be
performed to confirm the effectiveness of Grade 120 reinforcement to
confine concrete. Transverse reinforcement used to confine the core concrete
of members generally requires spacing that it is closer than the maximum
allowed spacing, so these tests would be different than tests used to
determine the maximum allowed transverse reinforcement spacing.
The maximum spacing of transverse reinforcement for columns and
boundary elements of special seismic systems exists to control bar buckling.
ACI 318-14 allows a maximum spacing of transverse reinforcement in the
zone of expected yielding of columns equal to the smallest of (a) one-quarter
the minimum member dimension, (b) six times the diameter of the smallest
longitudinal bar, 6db, and (c) so, which depends on the spacing of the vertical
bars, but not more than 6 inches. For wall boundary elements item (a) is
changed to one-third the thickness of the boundary element. The spacing so
in Item (c) is required for confinement but the maximum limit of 6 inches
provides additional bar buckling restraint for bars larger than No. 8. For
high-strength reinforcement, Item (b), the 6db maximum spacing limit, likely
requires adjustment for restraining high-strength longitudinal reinforcement.
Maximum transverse reinforcement spacing requirements for support of
high-strength longitudinal bars were studied and recommendations were
made as reported in NIST GCR 14-917-30. Both the spacing and stiffness of
the transverse reinforcement are important for restraining the longitudinal bar
(Tanaka, 1990; Restrepo-Posada, 1992; Rodriguez, et al., 1999; Wang and
Restrepo, 2001; Moyer and Kowalsky, 2003). An analytical approach was
used to determine the maximum transverse reinforcement spacing required to
obtain equivalent bar buckling restraint to that provided by transverse
reinforcement spaced at 6db to restrain Grade 60 longitudinal reinforcement.
The recommendations are to tighten transverse bar spacing to 5db for Grade
80 longitudinal reinforcement and 4db for Grade 100 and 120 longitudinal
bars. Tests are necessary either to confirm the results of the analytical study
or to determine alternative maximum spacing requirements.
ATC-115

3: Key Design Issues

3-37

95% Draft
Additionally, tests should also explore whether every longitudinal bar should
be supported. Currently, every other bar within 6 inches of supported bars do
not need to be inside of a transverse bar hook. Several tests have
demonstrated that supporting every bar with a transverse bar hook provides
better confinement and may be necessary for buckling restraint.
Several limits are imposed for the maximum spacing requirements of
transverse reinforcement for beams and columns of intermediate moment
frames. These limits are based on committee opinion rather than tests. The
maximum spacing limits of 8db and 12 inches likely require change with the
use of high-strength longitudinal reinforcement. The amount of longitudinal
bar buckling restraint required is less than that required for members of
special moment frames because the inelastic rotation demands are generally
not as large. However, tests rather than committee opinion are likely
required to support a reduced maximum transverse bar spacing if Grade 100
and 120 longitudinal bars are allowed for intermediate moment frames.
Consideration should also be given to imposing new or revised maximum
spacing requirements for transverse bars at the end regions of special
structural walls where special boundary elements are not required.
Recommendations for Spacing of Transverse Reinforcement in
Members Resisting Seismic Forces

Testing to determine the required spacing of transverse reinforcement for


beams, columns, and walls should be included as part of the test programs for
cyclic testing of beams, columns, and walls in Sections 4.9.3, 4.9.4.1, 4.9.4.2,
4.9.7 and 4.9.10.
3.4.3

Head Size and Attachment for Headed Deformed Bars

There are a number of questions that arise in dealing with the fabrication of
headed deformed bars using high-strength reinforcement: Are restrictions on
head type needed, are larger heads needed for special moment frames and
shear walls resisting earthquake effects, and do bars with nonlinear behavior
prior to yield and rounded post-yield stress-strain curves exhibited by a
number of high-strength steels require improved techniques for attaching the
head?
Ongoing structural tests at the University of Kansas indicate that headed bars
can be fabricated that will provide satisfactory performance for bars with
yield strengths of at least 120 ksi. For monotonic loading, there have been no
problems with the type of head used, any need for larger heads, or any
difficulties for high-strength bars with stress-strain curves that are rounded.

3-38

3: Key Design Issues

ATC-115

95% Draft
The behavior of headed bars with high-strength reinforcement is not
markedly different from that of bars with conventional reinforcement.
Producers indicate that the material used for the head will need to be changed
to achieve tensile strengths above 125 ksi.
Requirement for larger head sizes and the need for head attachments that
develop the actual tensile strength or a defined strain in the bar for special
moment frames and shear walls will become apparent when members
containing headed reinforcement are tested.
Recommendation for Study of Attachment or Size of Head

No separate study appears to be necessary to test potential changes to


required head attachments or size for monotonic loading. For cyclic loading,
refer to Section 3.6.3.5 on beam-column joints.
3.4.4

Development and Splice Lengths

Reinforced concrete members are designed based on the assumption that,


when loaded, reinforcing steel and concrete have the same nominal strain.
Within practical bounds, this assumption is true because the reinforcing bars
are deformed which limits slip and provides attachment to the concrete.
Attachment is also achieved with devices such as hooks and heads if the
available length of straight bars is insufficient. Design procedures used for
conventional reinforcement do not necessarily reflect the development and
splice length needs associated with high-strength reinforcement.
3.4.4.1

Straight Bar Development Lengths

Test results on the development and splice strengths of reinforcing bars with
conventional deformation patterns are available for No. 3 through No. 11
bars. The database also includes two No. 14 bars not confined by transverse
reinforcement. As shown in Figures 3-5 and 3-6, most of the tests are for
developed bar stresses below 80 ksi and concrete compressive strengths
below 10,000 psi.
A significant number of tests, however, are available for bar stresses in
excess of 80 ksi and for concrete strengths below 10,000 psi enough to
establish the adequacy of design expressions for concrete compressive
strengths below 10,000 psi. There are insufficient data, however, for bars
above 80 ksi in high-strength concrete.
The specimens used in splice tests of high-strength reinforcement have had
covers of at least one bar diameter and spacing between splices of at least

ATC-115

3: Key Design Issues

3-39

95% Draft
two bar diameters. To complete the results, test specimens with bar spacing
down to the greater of one bar diameter or 1 inch are needed.
160

Developedstress,ksi

140
120
100
80
60
40
20
0
0

2000

4000

6000

8000

10000

12000

14000

16000

18000

Compressivestrength,psi
Figure 3-5

Developed bar stresses and concrete compressive strengths for


development and splice tests on bars without confining
transverse reinforcement.

160
140

Developed stress, ksi

120
100
80
60
40
20
0
0

2000

4000

6000

8000

10000

12000

14000

16000

18000

Compressive strength, psi

Figure 3-6

3-40

Developed bar stresses and concrete compressive strengths for


development and splice tests on bars with confining transverse
reinforcement.

3: Key Design Issues

ATC-115

95% Draft
The development and splice length provisions of ACI 318 were developed
based on tests in which bar stresses were, for the most part, less than 60 ksi
(Orangun et al., 1975). ACI 318 equations for development length are
simplified in that development/splice length is proportional to bar stress. That
simplification is conservative for bars with yield strengths below 60 ksi. The
relationship becomes progressively less conservative, and even
unconservative, as the yield strength increases above 60 ksi. This is
especially true for bars not confined by transverse reinforcement. There is,
however, a linear relationship between the amount of confining transverse
reinforcement and the increase in stress developed in a spliced or developed
bar due to the addition of the confining reinforcement (Zuo and Darwin
2000; ACI, 2003).
This generally results in adequately conservative results when the ACI
provisions are applied to bars confined by transverse reinforcement and
suggests that transverse reinforcement should be required for high-strength
reinforcement if the current ACI equations are retained.
It has been demonstrated (Seliem et al., 2009; ACI, 2010; Choi et al., 2014)
that the design expressions developed by ACI Committee 408 (ACI, 2003)
are appropriately accurate and conservative for development and splicing of
bars with stresses as high as 150 ksi.
No tests have been reported on development and splice tests of epoxy-coated
bars at stresses above 80 ksi. Tests in this range are needed.
Recommendations for Testing of Straight Bar Development and Splice
Length

Additional development and splice length testing is needed for high-strength


reinforcement for the following conditions:

Bars in concrete with strengths greater than 10,000 psi,

Splices spaced no more than the greater of 1 in. or one bar diameter
apart, and

Epoxy coated bars.

Recommendations for these tests appear in Section 4.7.1.


3.4.4.2

Hooked Bar Development Length

Prior to industry-supported research now underway at the University of


Kansas, the total available data on the development of hooked bars was
limited. Early tests involved simulated beam-column connections in which
one or two reinforcing bars with standard or nonstandard hooks were pulled

ATC-115

3: Key Design Issues

3-41

95% Draft
out of a concrete block in most cases simulating a beam-column
connection. In the test, compression is applied to the surface of a column
specimen to simulate the compression region of a beam framing into the
column. Prior to the current study, no tests have been performed on closely
spaced or staggered hooked bars and only limited tests have been performed
on high-strength hooked reinforcing bars. The one test series on highstrength hooked bars (Shahrooz et al. 2011) was unrealistic in that individual
hooked bars were cast well away from member boundaries. Thus, the results
of those tests are appropriate only for hooked bars anchored in deep
foundations, well away from the exterior of the concrete.
To date, tests at the University of Kansas indicate that the current ACI
provisions for hooked bars become progressively less conservative (and even
unconservative) as the size of the bar and as the compressive strength of the
concrete increase. In addition, tests of closely spaced hooked bars indicate
that the current provisions become less conservative as the bar spacing
decreases.
No data are available on the bond of epoxy-coated high-strength hooked
reinforcement.
Recommendations for Study of Hooked Bar Development Lengths

Tests of closely spaced hooks are needed, and currently planned to be


performed at the University of Kansas. Tests of epoxy-coated hooked bars at
high stresses would be useful in the long term. Recommendations for these
tests are provided in Section 4.7.2.
3.4.4.3

Headed Deformed Bar Development Length

Until recently, only limited information has been available on the


development properties of high-strength headed reinforcement. Test results
are now just becoming available from Korea (Bang, 2014) and additional test
results will be available soon from a project at the University of Kansas. The
Korean research, which is largely connected to the construction of nuclear
power plants, has resulted in recommendations for modifications in
development lengths. Both deceases and increases in the development length
currently used in the ACI 318-14 are recommended, depending on the
application. Work at the University of Kansas indicates that some portions of
the Korean recommendations may be unconservative for beam-column joints
with closely spaced headed bars.

3-42

3: Key Design Issues

ATC-115

95% Draft
A discussion of headed deformed bars loaded by earthquake shaking is
provided in Section 3.6.3.5. The potential need for stronger head attachment
requirements or larger heads is discussed.
Recommendation for Study of Development Length for Headed
Deformed Bars

Further testing underway for required development of high-strength headed


deformed bars is described in Section 4.7.3.
3.4.4.4

Mechanical Splice

As part of adopting high-strength reinforcement into ACI 318, mechanical


splices will have to be tested with higher grade bar. Before manufacturers
embark on such testing, existing mechanical splice requirements should be
revisited and if necessary changes should be made. NIST GCR 14-917-30
recommends introduction of a Type 3 mechanical splice that would either
allow the bar to fracture away from the splice or be capable of developing a
preset uniform strain. The reason for introducing a Type 3 coupler is that
Type 2 couplers are only required to develop the specified tensile strength of
the bar. The actual tensile strength may be significantly greater, in which
case an adequate strain may not be developed in the bar if the splice is
located where ductility is needed in plastic hinge regions of members. The
current approach to achieve increased ductility is to heat treat the steel in the
region of the coupler, which also reduces strength. Such an approach will
likely not work for high-strength reinforcement.
Generally, couplers manufactured for Grade 60 reinforcement can carry bar
stresses up to 120 ksi. For bars with higher tensile strengths, current coupler
geometry can be maintained, but higher-strength alloys will be needed.
Coupler manufacturers feel they can adapt their current designs, but to do so
they need to know the specifications of the high-strength reinforcement. A
new specification or modifications to an existing specification will be
required for mechanical splices for high-strength reinforcement, particularly
for Type 3 couplers.
A study does not seem to be needed at this time, but producers need to
demonstrate the capacity of couplers in compression as well as in tension, in
accordance with ASTM A1034.
Recommendations for Mechanical Couplers

ACI Committee 318 needs to decide what requirements are needed for
splicing high-strength reinforcement and whether a new mechanical splice
designation is needed for splices that develop an established uniform bar

ATC-115

3: Key Design Issues

3-43

95% Draft
strain away from the splice or the actual tensile strength of the bar. No
further study is recommended at this time.
3.4.5

Bar Extensions in Slabs

Typically, longitudinal bars resisting moments extend beyond the inflection


point a length equal to the maximum of the depth of the member, a portion of
the span length, and a multiple of the longitudinal bar diameter. Higher yield
strengths require longer development lengths so bar extensions for one-way
and two-way slabs could possibly need to be increased so that the bars are
adequately developed.
3.4.5.1

Bar Extensions in One-Way Slabs

For one-way slabs, Sections 7.7.3.8.3 and 7.7.3.8.4 of ACI 318-14 provide
the requirements for minimum bar extensions beyond points of inflection for
positive and negative moment reinforcement, respectively. For a continuous
slab, the positive moment reinforcement must extend beyond a point of
inflection the greater of d and 12db. One-third of the negative moment
reinforcement must extend past the point of inflection by the greater of d,
12db and ln/16.
ACI 318R-11 Section R12.10.3 gives the reason for the extensions, The
moment diagrams customarily used in design are approximate; some shifting
of the location of maximum moments may occur due to changes in loading,
settlement of supports, lateral loads, or other causes. A diagonal tension
crack in a flexural member without stirrups may shift the location of the
calculated tensile stress approximately a distance d towards a point of zero
moment.
Recommendation for Study of Bar Extensions in One-Way Slabs

Based on this commentary statement, change to the bar extension


requirements is likely not needed for the use of high-strength
reinforcement. At most the 12db extension could be increased in proportion
to the yield strength. This could be addressed by committee opinion with no
further engineering studies or testing.
3.4.5.2

Bar Extensions in Two-Way Slabs

Section 8.7.4.1.3 and Figure 8.7.4.1.3(a) of ACI 318-14 includes minimum


slab bar extensions for two-way slabs. These bar extensions were established
considering dead and live loads acting on the slab. If the slab is part of a
slab-column frame designed to resist lateral loads, the minimum extensions
still apply but minimum extensions must also be computed based on the
distribution of moments to resist all applicable load combinations. Bar
3-44

3: Key Design Issues

ATC-115

95% Draft
extensions for two-way first appeared in the 1956 edition of ACI 318, a time
when Grade 60 reinforcement was not yet specified (Erlemann, 1997 and
1999; Meinheit and Felder, 2014). Although some minimum extensions in
the 1956 and 1963 editions were longer than currently required, the
minimum extensions have been unchanged since the 1971 version of ACI
318, a time when Grade 60 bar was being specified.
The commentaries to the 1963 and 1971 versions of ACI 318 do not indicate
how the minimum extensions were established. Engineering studies by
committee members likely were performed to establish them. The minimum
extensions are likely based on estimates of inflection point locations due to
pattern loading with added length to account for the tensile stress shifting due
to a shear crack as described in Section 3.4.5.1 for one-way slabs.
Recommendation for Study of Bar Extensions in Two-Way Slabs

Code change is likely not required with the use of high-strength


reinforcement but this could be decided by ACI Committee 318 opinion
without the benefit of an engineering study or research.
3.4.6

Horizontal Support of Offset Column Reinforcement

Longitudinal bars for square and rectangular columns are typically offset at
splices, especially the corner bars. ACI 318-14 Section 10.7.4.1 requires that
the slope of the inclined portion of an offset bar shall not exceed 1 in 6
relative to the longitudinal axis of the column. The force in the inclined
portion of a bar has a horizontal component that must be resolved near its
bends. Section 10.7.6.4.1 requires that horizontal support be provided to
resolve 1-1/2 times the horizontal component of force in the inclined portion
of the bar. This support can be from parts of the floor construction or from
transverse ties or spirals located within 6 inches of the bends as Section
10.7.6.4.2 requires. ACI 315-99 provides guidance in that it states, For
practical purposes, three closely spaced ties are usually used, one of which
may be part of the regularly spaced ties, plus two extra ties.
A similar requirement appears in the first edition of ACI 318 (ACI, 1941),
except that the ties needed to be within eight longitudinal bar diameters of
the bends. The eight longitudinal bar diameter (8db) requirement was
changed to the 6 inch requirement with the 1971 edition of ACI 318 (ACI,
1971a). For commonly used column longitudinal bar sizes, the 6 inch limit is
more restrictive. The change occurred soon after Grade 60 reinforcement
became widely available so it could have been in recognition of the transition
from structural and intermediate grade reinforcement with yield strengths of
33 and 40 ksi, respectively, to Grade 60 reinforcement. Unfortunately, the

ATC-115

3: Key Design Issues

3-45

95% Draft
commentary to the 1971 edition provides not guidance for the change (ACI,
1971b).
References that address the requirement for resisting the horizontal
component of force in the inclined longitudinal bar were not identified. It is
likely that the requirement was based on the opinion of the committee with
the possibility of concrete cover spalling having been observed where the
horizontal component was not resolved.
For equal sized bars with the same incline and bend geometry, the bar with
higher yield strength will develop a greater horizontal component to be
resisted. The increase in the horizontal component that must be resolved will
be in proportion to its increase in yield strength. It will also increase the
tensile stress in the concrete, which will make spalling of concrete cover at
the bends if the same limit of 6 inches is maintained. For higher strength
reinforcement, reducing the proximity of the ties that resist the horizontal
force component appears to be appropriate.
Recommendation for Study of Support of Offset Bars

The reduction in distance to supporting ties should be based on ACI 318


Committee opinion.
3.4.7

Cover for Fire Protection

The mechanical properties of all common building materials generally


decrease with elevation of temperature. A rise in temperature of concrete and
steel due to fire causes a decrease in the strength and modulus of elasticity of
the concrete and steel reinforcement. Some of these changes are not
recoverable after subsequent cooling. The International Building Code (IBC,
2012) contains prescriptive requirements for building elements and contains
tables describing various assemblies of building materials and finishes that
meet specific fire ratings. Determining the fire rating requirements in
building codes is based on the type of occupancy and the building height and
area. The fire resistance prescribed in building codes may require member
dimensions (including concrete cover for reinforcement) that are
significantly different from those based only on ACI 318-14 strength design
criteria.
Figure 3-7 shows the yield strength and tensile strength for various
reinforcing bars after exposure to high temperature (heating and cooling).
The data correspond to reinforcing bars commonly used in Japan. The figure
indicates that there is a greater reduction in both the yield and tensile
strengths for higher grade bars. Nonetheless, at temperatures approaching

3-46

3: Key Design Issues

ATC-115

95% Draft
1000 degrees Fahrenheit (537 degrees Celsius) relative changes in the yield
and tensile strengths are comparable to those experienced by conventional
reinforcing bars. A more pronounced reduction occurs if the yield and tensile
strengths are measured while the bars are being heated as shown in Figure
3-8.
Temperature, Fahrenheit
600
800
1000 1200

1400

1600

1400

400

Temperature, Fahrenheit
600
800
1000 1200

200

1200
160
1000
120

600

80

400

120

800
600

80

400

USD980
USD785
USD685A
USD685B
USD345

40
200

200

0
100

200

300
400
500
600
Temperature, Celsius

700

800

900

100

200

(a) Yield Strength


Figure 3-7

300
400
500
600
Temperature, Celsius

700

800

40

0
900

(b) Tensile Strength

Yield strength and tensile strength of reinforcing bars after


heating and cooling (adapted from Aoyama, 2001).

200

400

Temperature, Fahrenheit
600
800
1000 1200

1400

1600

1400
USD980
USD685A
USD685B
USD345

400

Temperature, Fahrenheit
600
800
1000 1200

600

80

Stress,MPa

1000
120

1600
200

1200

800

1400

USD980
USD685A
USD685B
USD345

160

Stress,ksi

1000

200

1400

200

1200

Stress,MPa

1600

160

800

1400

160

120

800
600

80

Stress,ksi

Stress,MPa

1000

200

1400

200

USD980
USD785
USD685A
USD685B
USD345

1200

Stress,ksi

400

Stress,MPa

200

Stress,ksi

400

400

40

40
200

200
0

0
0

100

200

300
400
500
600
Temperature, Celsius

700

800

900

0
0

100

200

(a) Yield Strength


Figure 3-8

300
400
500
600
Temperature, Celsius

700

800

900

(b) Tensile Strength

Yield strength and tensile strength of reinforcing bars at high


temperature (adapted from Aoyama, 2001).

As higher strength materials are impacted to a greater degree than lower


strength materials by a rise in temperature associated with a fire, studies are
needed to determine how higher strength reinforcement encased in a concrete
ATC-115

3: Key Design Issues

3-47

95% Draft
element is affected by fire. Consideration should be given to the effects of
bar size and ratio of tensile-to-yield strengths since exposure to high
temperatures tend to more severely affect the yield strength of small bars
(such as No. 4 bars) and the tensile strength of large bars (such as No. 11
bars) as indicated by Edwards and Gamble (1986).
Recommendation

Although a study to evaluate the effects of fire on concrete members


reinforced with high-strength steel bars is required, it is not included in
Chapter 4 because it is beyond the scope of ACI 318-14.
3.4.8

Beam-Column and Slab-Column Joints (Type 1)

Beam-column and slab-column joint requirements are in Chapter 15 of ACI


318-14. The requirements for interior and exterior joints are generally
treated separately.
3.4.8.1

Interior Beam-Column Joints

For interior joints, ACI 318-14 does not require minimum joint depths or
specify the maximum bar size that can pass through a joint of a moment
frame resisting wind and other non-seismic lateral loads, an ordinary moment
frames resisting seismic forces, and an intermediate moment frames resisting
seismic forces. ACI 352R-02 (ACI, 2002b) refers to these moment frame
joints as Type 1 joints and does not make any recommendations for
minimum joint depth or maximum size bar that can pass through the joint,
refer to Section 4.5.5 of ACI 352R-02.
The bond stresses developed along the bars passing through these joints will
increase with the use of high-strength reinforcement. The bars passing
through the joint may be in compression on one side of the joint and tension
on the other side of the joint. This change in bar force through the joint depth
causes bond stresses in the concrete that can be high. Assuming the same
number of smaller high-strength bars are used as compared to Grade 60 bars
for joints that are otherwise unchanged, the higher strength reinforcement
will cause higher bond stress in the concrete because the change in bar force
will be distributed over a smaller bar surface area and thus to smaller rib
area.
Bars passing through Type 1 joints resisting wind loads and other lateral
loads besides earthquake loads are not expected to undergo inelastic
deformations, but the bars passing through Type 1 joints of ordinary and
intermediate moment frames designed for earthquake forces are expected to
undergo inelastic deformations. The higher axial stresses and strains

3-48

3: Key Design Issues

ATC-115

95% Draft
resulting from the bars yielding as compared to those that are not yielding
will results in higher bond stresses. Therefore, the bond stresses for bars
passing through joints of ordinary and intermediate moment frames will be
larger than those for the bars resisting gravity, wind and other non-seismic
lateral loads.
ACI Committee 318 should consider the need for minimum joint depth or
maximum bar size for a given joint depth. Separate consideration is
appropriate for joints with bars not expected to yield versus those expected to
undergo inelastic yielding.
Type1 interior joints for which bars passing through are not expected to yield
are unlikely to require restrictions on its depth. For the case in which gravity
loads dominates the design of the beams and columns, the change in bar
stress through the joint is usually small. For frames supporting wind and
other non-seismic lateral loads that will not yield the bars, the deterioration
within the joint due to high bond stresses will be substantially less than
expected for joints resisting seismic loads. Therefore, requirements for
minimum joint depths or maximum bar size passing through the joint of a
given depth are probably unnecessary.
Bars of ordinary and intermediate moment frames designed for earthquake
forces are expected to yield even though they are not expected to experience
the degree of yielding of bars in a special moment frame. Thus, restrictions
on joint depth or maximum size of bar passing through a joint of a given
depth may be prudent. A combination of a smaller number of tests combined
with ACI 318 Committee opinion is likely sufficient to address this issue.
Although, the shear stress through these joints can be high, ACI 318-14 does
not address the shear stress in Type 1 joints directly. It does require
minimum transverse reinforcement except for joints confined on four sides
by beams. ACI 352R-02 does provide recommendations for an explicit
check of the joint shear. Such a requirement would be required regardless of
the yield strength of the reinforcement, so it is not addressed further in this
roadmap.
Recommendations for Study of Type 1 Interior Joints

Required joint depth or limit on the bar size passing through a joint of given
depth are likely unnecessary for Type 1 interior joints with bars passing
through them that are not expected to yield; however, ACI Committee 318
should consider this and, if deemed necessary, base the requirements on
committee opinion. No tests are recommended for these joints.

ATC-115

3: Key Design Issues

3-49

95% Draft
ACI Committee 318 should separately consider whether Type 1 joints of
ordinary and intermediate moment frames designed to resist earthquake loads
requires provisions for minimum joint depth or maximum bar size passing
through a joint of given depth. A small number of beam-column joint tests
representative of intermediate moment frames should be performed as part of
the test program recommended in Section 4.9.4.3 The test results along with
ACI 318 Committee opinion should be sufficient to develop code
requirements, if necessary.
Study for a maximum allowed joint shear stress is not considered as part of
this roadmap as this is an issue for all reinforcement grades.
3.4.8.2

Exterior Beam-Column Joints

For exterior joints separate treatment should be given to joints with beam
bars terminating in the joint that are not expected to yield and those with
beam bars that are expected to yield. Frames that resist gravity loads, wind
loads and other non-seismic lateral loads usually have bars that terminating
in the joint that are not expected to yield. The minimum depth of these
joints, regardless of whether the bars are or are not expected to yield, is
controlled by the depth needed to develop the hooked or headed beam bars in
the joint. ACI 318-14 has no other specific requirements that might control
the joint depth such as a check on the joint shear. The need for such
requirements should be considered with the adoption of high-strength
reinforcement.
For exterior joints resisting only gravity and wind loads or other non-seismic
lateral loads, the behavior is likely to be similar to be similar regardless of
the strength of the bar. This could be established with a small number of
tests. For exterior joints of ordinary and intermediate moment frames
designed to resist earthquake loads, the need for adjustments to joint depth,
the need for more joint reinforcement, and the limits on joint shear should be
explored.
Another issue that has recently become apparent but is yet unpublished is
tests of closely spaced hooked bars in joints can result in a multi-bar breakout
failure. This issue will likely be addressed by adding additional requirements
for the design of transverse reinforcement in these joints. Although this is
not an issue limited only to high-strength reinforcement, addressing it may be
necessary to allow adoption of high-strength reinforcement into ACI 318.

3-50

3: Key Design Issues

ATC-115

95% Draft

Recommendation for Study of Type 1 Exterior Joints

A small test program of beam-column exterior joints designed for gravity and
wind loads is recommended to confirm equivalent performance to joints
reinforced with Grade 60 bar. Exterior joints of intermediate moment frames
should be tested as part of the test program described in Section 4.9.4.4.
3.4.8.3

Slab-Column Joints

Slab-column joints can be used for gravity framing and for joints of slabcolumn ordinary and intermediate moment frames. For gravity frames, the
joints are likely to perform similar to the joints of frames reinforced with
Grade 60 reinforcement. These are covered with two-way slabs described in
Section 3.1.2.1.2.
For slab-column ordinary and intermediate moment frames, the frame should
be capable of undergoing the expected drift for the maximum considered
earthquake defined in ASCE 7-10 without a shear failure resulting in
collapse. Discussion is provided in Section 3.6.2 for Intermediate slabcolumn moment frames.
Recommendations for Study of Slab-Column Joints

Slab-column joints for gravity frames can be studied with two-way slabs.
The test program is described in Section 4.4.5.2. The joints of intermediate
moment frames should be studied as described in 4.9.4.5. The results of the
tests for intermediate moment frame joints can be used to inform ACI
Committee 318 regarding the likely performance of slab-column joints of
ordinary moment frames.
3.5

General Considerations for Analysis

3.5.1

Flexural Stiffness

For structural analysis, the selection of the moment of inertia used to


represent the cross-section is an important consideration. While designers
are permitted to perform a rigorous analysis to select these values, ACI 31814 presents in Section 6.6.3 moment of inertias that are considered
appropriate. These values are based on research by MacGregor and Hage
(1977) and were selected from the results of frame tests and analyses. In
general, the recommended moments of inertia are presented as a fraction of
the gross moment of inertia Ig for various members such as beams and
columns. These moments of inertia reflect a number of variables including
the amount of reinforcement, the extent of cracking, the variation of cracking
along the member, axial loads, creep, and inelastic behavior of steel and

ATC-115

3: Key Design Issues

3-51

95% Draft
concrete. It is for this reason that different values are presented for factored
loads and service loads (ACI 318-14, Sections 6.6.3.1 and 6.6.3.2). Overall,
it is important that the flexural stiffness of the members is appropriate to
reflect the overall behavior of the structure considering gravity and lateral
analysis. These studies have been carried out by Korinda (1972) and Hage
(1974).
With the use of higher strength reinforcement, there is the potential that
increased cracking will occur at a given section and that the extent of
cracking over the length of the member will increase. An increase in
cracking will result in a reduction in the moment of inertia, which suggests
that modification of the moments of inertia may be needed. The
modification required will be greater as the grade of the reinforcement is
increased.
The moment of inertia is also used for consideration of second-order effects.
While general analysis of a structure considers average moments of inertia
across the structure, the moment of inertia used for consideration of secondorder effects depends on the forces being calculated. The behavior of the
structure is controlled by sway when subjected to lateral loads, for which the
calculation of second-order end moments in columns depends on the average
stiffness of the structure. However, for calculation of second-order moments
within a column (between the ends), the stiffness of the specific column
being investigated is required. It is for this reason that lower values of
stiffness, EI, are used (ACI 318-14, Section 6.4.4.4). For columns, it is not
expected that increased cracking levels will be developed through the use of
higher strength reinforcement. Therefore, effective EI values should be
similar. However, it is likely that decreased amounts of reinforcement may
be provided or column sizes may be decreased. These parameters are
expected to have a more significant influence than changes in cracking.
Recommendation for Further Study

Further research to evaluate the influence of high-strength reinforcement on


flexural stiffness, EI, is recommended, and a proposed research program is
outlined in 4.8.1. In addition, further research to evaluate the influence of
high-strength reinforcement on the flexural stiffness used in second-order
analysis is recommended, considering that decreased columns sizes will be
possible through the use of high-strength reinforcement. A proposed
research program is outlined in 4.8.2.

3-52

3: Key Design Issues

ATC-115

95% Draft
3.5.2

Moment Redistribution

ACI 318 allows for redistribution of moments calculated by elastic analysis.


The amount of redistribution permitted depends on the tensile strain of the
longitudinal reinforcement at the extreme layer with the maximum amount
being 20 percent. Because moment redistribution depends on adequate
ductility in the plastic hinge regions, adequate ductility must be provided.
Using the ACI approach, ductility is accounted through the tensile strain
achieved at ultimate, and a minimum strain of 0.0075, which is
approximately 3.6 times the yield strain of Grade 60 reinforcement, is
required. The redistribution permitted is based upon analysis of flexural
members with small rotation capacity using conservative values of limiting
concrete strains and lengths of plastic hinges derived from extensive tests
(ACI 318). As shown in Figure 3-9, the current permissible redistribution is
conservative for yield strengths up to 80 ksi.

25

Percentchangeinmoment

/d =23
b/d =1/5
20
Calculated
percentage
available

15

Permissibleredistribution
allowedby6.6.5.3

10

5
Minimumpermissible
nettensilestrain=0.0075

0
0

Figure 3-9

.005

.010
.015
Nettensilestrain,t

.020

.025

Permissible redistribution in accordance with ACI 318.

With the use of higher strength reinforcement, the yield strain of the
reinforcement will increase and the curves shown in the figure will become
closer to the permissible redistribution line. For instance, if Grade 100 is
used, maintaining the minimum strain as 0.0075 will produce a reduction in
the ratio of 2.2 times yield. Depending on the stress-strain response of
higher-strength materials, this reduced ratio may not be sufficient to

ATC-115

3: Key Design Issues

3-53

95% Draft
adequately provide for redistribution of moments. Consequently, either
redistribution may not be allowable for higher-yield materials or a
modification in the redistribution allowances may be warranted. Regardless
of code-permitted redistribution, it would be useful to understand the
redistribution capacity of members reinforced with high-strength
reinforcement as this capability often provides for reserve capacity of
structures especially in the event of overloads.
Recommendation for Further Study

Further research regarding redistribution seems warranted for structures


reinforced with high-strength reinforcement and essential if redistribution is
explicitly permitted for these higher grades, and a proposed research program
is outlined in 4.8.3.
3.6

Seismic-Force-Resisting Systems

The response of moment frames and shear walls to earthquake ground


motions is dependent on many factors. The response of the components that
comprise these framing systems depends on the mechanical properties of the
reinforcement and the concrete, rotational capacity, the relative strengths in
shear and flexure, configuration and geometry, and the detailing of
longitudinal and transverse reinforcement.
The appropriateness of high-strength reinforcement for members of seismicforce-resisting systems can be evaluated two ways. The first approach is to
establish equivalent response to that provided by members and systems
reinforced with ASTM A706 Grade 60 reinforcement, which assumes that
systems and members reinforced with ASTM A706 Grade 60 reinforcement
provide adequate performance, including less than 10% probability of
collapse conditioned on risk targeted maximum considered earthquake
shaking. The second approach is to directly evaluate the performance of
moment frames and shear wall systems for adequate performance. Although
this approach is conceptually more appealing, current modelling approaches
are not sufficiently reliable to predict the behavior of reinforced concrete
structures. However, current analysis capabilities are likely sufficient to
establish equivalent performance.
In determining whether equivalent performance is provided, both system and
member performance must be assessed. Differences in member stiffness,
energy dissipation, and backbone curves for strength will likely result in
differences in overall performance of systems reinforced with high-strength
reinforcement and those reinforced with Grade 60 reinforcement. The
differences in overall system performance should be used to determine

3-54

3: Key Design Issues

ATC-115

95% Draft
whether member performance can be compared directly or whether
adjustments are needed for the comparison. For example, based on the
analysis results, do members with high-strength reinforcement require greater
rotational capacity?
The use of high-strength reinforcement for seismic systems is discussed in
NIST GCR 14-917-30. The focus of NIST GCR 14-917-30 is special
seismic systems and many of the issues and concerns identified in the
sections that follow are introduced and discussed in greater detail in that
report. It identifies and summarizes research completed to date on members
reinforced with high-strength reinforcement. ACI 318-14 limits bar for use
in special seismic systems to Grade 60. NIST 14-917-30 makes
recommendations regarding code changes required to allow the use of Grade
80 reinforcement for special seismic systems and makes recommendations
for the research likely needed and types of code changes required for
adoption of Grade 100 and 120 bar.
The following sections identify issues and concerns with seismic force
resisting systems and identify whether further study in the form of
engineering studies, analyses, and testing are needed to address these issues
and concerns.
3.6.1

Shape of Stress Strain Relationship

The shape of the reinforcement stress-strain curve and response of bars to


cyclic loading affects several factors that influence the response of beams,
columns, and walls of seismic-force-resisting systems. The influence of the
stress-strain curve was partially explored in NIST GCR 14-917-30. Momentcurvature relationships were computed for beams, columns, and shear walls
using stress-strain curves representative of A706 Grade 60 and Grade 80
reinforcement, and Grade 100 reinforcement with three different types of
stress-strain curves. These curves are shown in Figure 3-2, and additional
information about the curves is provided in Table 3-3. Stress-strain
parameters used in this study were established based on only a few actual
stress-strain curves, so they are not considered representative of the full
potential variation that is likely to exist in high-strength reinforcement. For
example, useable elongations, ratios of tensile-to-yield strengths, and lengths
of the yield plateau are likely to vary substantially from the values
considered in NIST 14-917-30.
Additional information regarding stress-strain curves of Grade 80 and 100
bars will soon be available. Moment-curvature relationships could be
computed for additional stress-strain relationships that more closely represent

ATC-115

3: Key Design Issues

3-55

95% Draft
the stress-strain curves of reinforcement that is or will become available in
the United States. The moment-curvature relationships should be computed
for stress-strain relationships that represent the minimum properties likely to
be required in specifications, as well as for average and maximum likely
properties.
Table 3-3

Stress-Strain Parameters for Reinforcing Steel Used to Create


Moment-Curvature Relationships in NIST GCR 14-917-30

Reinf. Type1

fy 2

sh 3

su 4

fu / fy 5

100-S16

100 ksi

0.06

1.5

100-S2

100 ksi

0.00207

0.06

1.2

100-S3

100 ksi

0.0134

0.06

1.5

80-S3

80 ksi

0.0128

0.06

1.5

60-S3

60 ksi

0.0121

0.06

1.5

Notes:
1
Refer to Figure 3-2.
2
Yield strength of reinforcement based on the 0.2% offset method.
3
Strain defining the onset of strain hardening.
4
Maximum usable strain. For reinforcing steel models in Figure 3-2, maximum
usable strain coincides with the tensile strength. A value of 6% is a representative
lower-bound strain associated with the peak stress (tensile strength). In contrast,
the strain associated with fracture elongation may be up to two times the strain
associated with peak stress.
5
Ratio of tensile strength to yield strength.
6
Refer to NIST GCR 14-917-30 for the stress-strain relationship.

The shape of the stress-strain relationship is also likely to affect the spread of
plasticity and the contribution of bar slip in joints and foundations to
rotational capacity of members. In the study for NIST GCR 14-917-30,
rotational capacities were computed assuming that no bar slip in joints and
foundations occurred, the useable elongation was 6%, and the plastic hinge
length equals one-half of the overall member depth. Ignoring the bar slip is
conservative. Future analyses could include estimations of the effects of bar
slip but the variation in slip should account for the possibility that highstrength bar may have greater relative rib area and mechanical anchoring in
joints that may reduce bar slip. The useable elongation of 6% may not be
achievable for all high-strength bars to be produced, so the impact of this
being smaller should be explored. The plastic hinge length will depend on
the shape of the stress-strain relationship and member proportions rather than
equal for all cases. Correlation to tests is likely required to better understand
this variation.
Analytical models alone are not capable of accurately addressing the
contribution of bar slip, the spread of yielding, and the overall influence of

3-56

3: Key Design Issues

ATC-115

95% Draft
the shape of the stress-strain curve and cyclic loading behavior of bars on
rotational capacity. Tests are required to explore the influence.
A series of beam tests that explored the effect of tensile-strength-to-yieldstrength ratio on cyclic behavior was performed in Japan as part of the New
RC Project (Aoyama, 2001). The results of these tests are only briefly
discussed in Aoyama (2001) and a reference is not provided to the original
source of the work. The information provided, however, clearly
demonstrates the importance of this ratio on how well strength is maintained
to drift levels of 5% and more. Figure 3-10 shows a representation of the test
specimens and Figure 3-11 shows the two load-deflection curves that appear
in Aoyama. This topic is also discussed in NIST GCR 14-917-30.

Figure 3-10

Beam specimens used to study the effect of the tensile-strength-to-yield-strength


ratio (Aoyama, 2001).

(a)
Figure 3-11

(b)

Load-deflection curves for beam specimens with: (a) yield ratio of 75% with
splice; and (b) yield ratio of 90% without splice. Yield ratio is the inverse of the
tensile-strength-to-yield-strength ratio (Aoyama, 2001).

Tests that explore the effects of various reinforcement mechanical properties


on the cyclic response of beams and columns would be valuable for

ATC-115

3: Key Design Issues

3-57

95% Draft
establishing required mechanical properties that should appear in bar
specifications. Tests should consist of beams, columns, and walls
constructed with reinforcement having different stress-strain curve
relationships and should cover the range of mechanical properties of bars that
mills are likely to produce.
The purpose for the additional analytical studies and testing would be to
establish equivalency of members reinforced with high-strength
reinforcement to those with Grade 60 reinforcement. They would also be
used to establish minimum required reinforcement properties such as the
uniform elongation, ratio of actual tensile-to-yield strength, and maximum
yield strength.
Recommendations for Study of Effects of Reinforcement Properties

Tests and numerical studies on the effect of stress-strain curve shape are
described in Sections 4.*.* and 4.*.*.
3.6.2

Intermediate Moment Frames

Provisions for intermediate moment frames were introduced into ACI 318 in
the 1983 edition (ACI, 1983). This edition also included provisions for slabcolumn, intermediate moment frames. The provisions for detailing typical
beams and columns were based primarily on committee opinion. The
spacing of hoops in the end regions of beams and columns has not changed
since 1983. The 2002 edition of ACI 318 (ACI, 2002a) introduced specific
requirements for punching shear, including that two-way shear caused by
gravity load not exceed 0.4Vc. The 2011 edition (ACI, 2011) increased the
design shear of columns from 2 times the shear determined from code level
earthquake forces to the shear associated with development of the nominal
flexural strength at each end of the column loaded in reverse curvature
bending or the overstrength factor, o, times the shear determined from code
level forces. For intermediate moment frames the need to change the spacing
of hoops, the maximum gravity shear, and the design shear strength of
columns and beams should be explored for members constructed with highstrength reinforcement.
Beams and Columns

Few studies have investigated the specific provisions of intermediate frames


Hwang and Hsu, 1993; Hwang and Hsu, 1993; Panahshahi and Lu, 1997;
Sheth, 2003). Cyclic tests have been performed on columns reinforced with
Grade 60 longitudinal bars with transverse reinforcement similar to that
required for intermediate moment frames (Han and Jee, 2005), but the results

3-58

3: Key Design Issues

ATC-115

95% Draft
of these tests have not lead to code changes. Cyclic tests of beams and
columns reinforced with high-strength longitudinal bars and transverse
reinforcement similar to that required for intermediate moment frames have
not been identified (if they are available) but are needed to support the use of
high-strength reinforcement.
Collapse Probability

Although the study performed on one archetype beam-column, intermediate


moment frame building indicated the collapse probability was sufficiently
low (Richard, et al., 2010), more buildings must be evaluated before
definitive conclusions can be drawn. Collapse probability studies of
intermediate moment frames are likely to be required to support a code
change proposal. Testing of beams, columns and joints, and nonlinear
response-history analyses in accordance with the FEMA P695 methodology
will be required to achieve this.
Slab-Column Frames

Adequate performance of a slab-column moment frame requires that a twoway shear failure not occur. The limit of 0.4Vc for two-way shear caused by
gravity load is based on tests with Grade 60 longitudinal slab bars (Pan and
Moehle, 1989). Keeping the gravity load shear below this value provides
protection against shear failure at story drifts up to 1.5%. Although these
provisions appear independent of reinforcement ratio and bar strain, two-way
shear strength depends on both as described in Section 3.1.2.1.2. Use of
high-strength slab bar will reduce the reinforcing ratio and increase the strain
in the bar. Therefore, tests are required to confirm that the current limit of
0.4Vc is still appropriate to provide protection of slabs constructed with
high-strength bar to drifts expected for MCE shaking. The slab shrinkage
should be restrained to replicate field conditions. Also, nonlinear response
history analyses should be performed to establish the expected drifts of codecompliant slab-column moment frames.
Recommendations for Intermediate Moment Frame Studies

Tests and numerical studies on intermediate moment frames, and


components thereof, constructed with high-strength reinforcement are
included in studies focused on special moment frames described in Sections
4.9.4.1, 4.9.4.2, 4.9.4.5 and 4.9.4.6.
3.6.3

Special Moment Frames

The performance of special moment frames depends on not only the strength
and deformation capacities of members but also the relative strengths of

ATC-115

3: Key Design Issues

3-59

95% Draft
members and the strength and detailing of joints at member intersections.
Strength and deformation capacity of beams, columns, and joints are
addressed in the following sections. The relative strength of beams and
columns to obtain adequate strong column-weak beam behavior is discussed.
3.6.3.1

Beams

A summary of cyclically loaded test beams with high-strength reinforcement


are presented in NIST GCR 14-917-30. Among the research papers available
in English are Sugano et al. (1990) and Kimura et al. (1993), both of which
focus on the cyclic response of concrete beams. Two of the eight beams
(part of beam-column subassemblages) tested by Sugano et al. were
reinforced with longitudinal and transverse reinforcement having yield
strengths of 85 ksi and 125 ksi, respectively. The compressive strength of
concrete was 12 ksi, and the beam cross section was 12 inches wide by 16
inches deep, with a shear span to effective depth ratio of about 3.5. Kimura
et al. tested 14 cantilever beams, seven of which were reinforced with
longitudinal and transverse reinforcement that had yield strength of 115 ksi.
Each specimen had a compressive strength of concrete of either 5.5 ksi or 11
ksi and the beam cross section was 8 inches wide by 12 inches deep, with a
shear span to effective depth ratio of 4.7. The beams tested by Sugano et al.
and Kimura et al. reached drift ratios of 5% with no more that 20% loss of
peak strength.
Tavallali (2011) performed cyclic tests of concrete beams with high-strength
reinforcing bars. The tensile strength and total elongation of the bars were
98 ksi and 16%, respectively, for Grade 60 bars, and 117 ksi and 10%,
respectively, for Grade 97 bars. Experimental data for two beam specimens,
CC4-X and UC4-X, are presented here. The geometry and reinforcement
details of the test specimens are shown in Figure 3-12. The specimens
consisted of two identical beams connected to a central stub through which
load was applied, subjecting the specimens to single curvature bending. All
transverse reinforcing bars were Grade 60, and the nominal compressive
strength of the concrete was 6 ksi. The layout of the longitudinal
reinforcement was symmetrical, with identical top and bottom layers. The
amount of longitudinal reinforcement was chosen so that both specimens
reached nearly identical flexural strength while limiting the shear stress,
V/bd, to values approaching 6 f c' (psi).The flexural reinforcing ratio,, is
1.9% for the Grade 60 specimen and 1.4% for the Grade 97 specimen.
The measured shear-drift response for CC4-X is shown in Figure 3-13a,
excluding the final monotonic loading event. The north beam tolerated two

3-60

3: Key Design Issues

ATC-115

95% Draft
cycles of 5% drift while maintaining a load-carrying capacity similar to the
peak load resisted in previous cycles.

Figure 3-12

Details for beam specimens with high-strength reinforcing bars (adapted from
Tavallali, 2011).

Specimen UC4-X had similar properties to Specimen CC4-X, with the


exception that it was reinforced longitudinally with Grade 97 (SAS 670) steel
bars. Figure 3-13b shows its measured shear-drift response.
Compared with Specimen CC4-X, Specimen UC4-X demonstrated reduced
post-cracking stiffness, increased yield deformation, and slightly narrower
hysteresis loops. The south beam of specimen UC4-X controlled the loading
protocol and tolerated two cycles of 5% drift while maintaining a loadcarrying capacity similar to the peak load resisted in previous cycles.

(a)
Figure 3-13

(b)

Measured shear versus drift ratio in beam tests: (a) Specimen CC4-X, with Grade 60
reinforcing bars; and (b) Specimen UC4-X, with Grade 97 reinforcing bars (Tavallali, 2011).

After loading to 1% drift, Specimen CC4-X resisted small increases in shear


to cycles of 4% drift and maintained the shear to drifts of 5%. Specimen
UC4-X reached a peak shear at about 2% drift and resisted slightly
decreasing shear to cycles of 5% drift. The Grade 60 bars were characterized
by a tensile strength to yield strength ratio (fu / fy) of 1.5, whereas the Grade
97 bars had a ratio of 1.2.

ATC-115

3: Key Design Issues

3-61

95% Draft
The above test data indicate that replacing Grade 60 longitudinal
reinforcement with Grade 97 reinforcement, reduced in proportion to the
yield strength of the reinforcement, leads to comparable flexural strength and
deformation capacity. These findings were corroborated by Pfund (2012)
using similar beam specimens reinforced with ASTM A1035 Grade 120
longitudinal bars. These observations suggest that concrete beams reinforced
with high-strength reinforcement are a viable option in earthquake-resistant
construction.
Recommendation for Studies of Beams of Special Moment Resisting
Frames

Additional testing of beams reinforced with high-strength reinforcement is


recommended in Section 4.*.*. Also, the adequacy of the rotational capacity
of beams should be confirmed using nonlinear response history analyses.
3.6.3.2

Columns

Cyclic tests of columns constructed with high-strength reinforcement have


been performed in the United States in recent years. Many such tests are
summarized in detail in NIST GCR 14-917-30. Restrepo et al. (2006) reports
the testing of two circular columns that were scale models for the Oakland
approach of the new San Francisco-Oakland Bay Bridge. The columns were
3 feet in diameter and 9-6 tall.
Unit 1 was constructed with ASTM A706 Grade 60 longitudinal and
transverse reinforcement. Two cages, each containing 42 No. 5 bars were
tied to No. 3 fuse-welded hoops spaced at 1.56 inches on center. The
longitudinal reinforcement ratio was = 2.54%, and the volumetric
transverse reinforcement ratio was s = 1.74%. The measured concrete
compressive strength was 9.3 ksi
Unit 2 incorporated ASTM A1035 Grade 100 longitudinal and transverse
reinforcement. A single cage with 42 No. 5 longitudinal bars was tied to
No. 3 fuse-welded hoops spaced at 1.56 inches on center. The measured
concrete compressive strength was 8.2 ksi.
Both units were tested with an axial load of 0.07 f'cAg and subjected to
reverse cyclic lateral loading. Figure 3-14 shows the hysteretic response of
the two columns.
For the levels of story drift normally expected for special moment frames, the
two columns exhibited effectively identical behavior. Unit 1 was cycled to
drift ratios greater than 6%, when yielding of the hoops led to longitudinal
bar buckling followed by fracture. Unit 2 failed at 3.9% drift on the first

3-62

3: Key Design Issues

ATC-115

95% Draft
cycle to a target drift of 6% via hoop fracture in the heat-affected region
adjacent to the fuse weld.

Figure 3-14

Hysteretic response of two circular columns based on Restrepo


et al. (2006): Unit 1 including Grade 60 reinforcement, and
Unit 2 including Grade 100 reinforcement (courtesy of J.
Restrepo).

Rautenberg (2011) tested columns with high-strength reinforcing bars.


Experimental data are presented for two column specimens, CC-3.3-20 and
UC-1.6-20. Specimen CC-3.3-20 was designed with 60 ksi reinforcement
according to ACI 318-08 requirements for special moment frame columns.
Specimen UC-1.6-20 had properties that were similar to Specimen CC-3.320, with the exception that UC-1.6-20 was reinforced longitudinally with
ASTM A1035 Grade 120 steel bars, achieving nearly the same fy of CC3.3-20. The yield strength, tensile strength, tensile-to-yield strength ratio,
and total elongation were 64 ksi, 92 ksi, 1.44, and 20% for the Grade 60 bars,
respectively, and 133 ksi, 168 ksi, 1.26, and 8.6% for the Grade 120 bars.
The nominal compressive strength of the concrete was 6 ksi. All transverse
reinforcement was Grade 60. The test setup and loading protocol were
similar to those used by Tavallali (2011) presented previously. Both
specimens were subjected a constant axial load of 0.2 fcAg. The typical
geometry and reinforcement details of the test specimens are shown in Figure
3-15. The amounts of longitudinal reinforcement was chosen so that the
columns would reach nearly identical flexural strengths but limit the average
shear stress to approximately 8 f c (psi).

ATC-115

3: Key Design Issues

3-63

95% Draft

Figure 3-15

Details for column specimens with high-strength reinforcing bars (adapted from
Rautenberg, 2011).

The measured shear drift response for CC-3.3-20 is shown in Figure 3-16a.
The controlling column completed the first cycle to 5% drift, but the
longitudinal bars buckled during the second cycle at that drift ratio. Figure
3-16b shows the measured shear-drift response of UC-1.6-20. The
controlling column of Specimen UC-1.6-20 completed the first half-cycle to
5% drift, but the longitudinal bars buckled during the second half-cycle at
that drift ratio.

(a)
Figure 3-16

(b)

Measured shear versus drift ratio in column tests: (a) Specimen CC-3.3-20,
with Grade 60 reinforcement; and (b) Specimen UC-1.6-20, with Grade 120
reinforcement (Rautenberg, 2011).

Similar findings were obtained by Tretiakova (2013) on cyclic tests of


concrete columns reinforced with SAS 670 Grade 97 steel bars. The
columns tested by Tretiakova were nearly identical to those tested by
Rautenberg (2011).
Many tests have been conducted in Japan on columns with high-strength
reinforcement. Data from tests of 115 columns has been used to calibrate
statistical models for predicting ultimate deformation capacity (Ishikawa et al.,
2008).

3-64

3: Key Design Issues

ATC-115

95% Draft

Recommendation for Studies of Columns of Special Moment Resisting


Frames

Additional testing of columns reinforced with high-strength reinforcement is


recommended in Section 4.*.*. Also, the adequacy of the rotational capacity
of columns should be confirmed using nonlinear response history analyses.
3.6.3.3

Shear Demand on Beams and Columns

Shear demand on beams and columns in special moment-resisting frames is


based in part on the flexural strength of the beams at levels of deformation
expected for shaking more intense than design basis. The commentary to
Section R18.6.5 of ACI 318-14 notes that shear demand be calculated using a
reinforcement stress of at least 1.25fy Because the actual yield strength of
the longitudinal reinforcement may exceed the specified yield strength and
because strain hardening of the reinforcement is likely to take place at a joint
subjected to large rotations. The presence of slabs reinforced in the same
direction as the beam longitudinal reinforcement will also increase flexural
strength of the cross section over the nominal strength of the rectangular
beam cross section. If the ratios of expected-to-specified yield strength,
expected tensile strength-to-expected yield strength, and measured tensile-toyield strengths are substantially greater for high-strength reinforcement, the
factor of 1.25 should be adjusted upwards.
At the time that the probable moment strength was established as the strength
assuming a reinforcement stress of 1.25fy, buildings typically had more
frames resisting seismic forces and the size of those members were smaller.
Part of the reason that the factor was 1.25 and not greater was that spalling of
concrete cover reduced the member strength in excess of the nominal value.
Nowadays, buildings are generally designed with fewer seismic frames and
larger members (as a consequence of the smaller number of frames). The
spalling of the cover concrete results in less of a reduction in the overstrength
because the cover is a smaller fraction of the member depth. For this reason
alone, the factor of 1.25 should be reassessed for both Grade 60 and higher
strength reinforcements.
Recommendation for Reassessment of Shear Demand on Beams and
Columns

The reassessment of shear demand on beams and columns requires a review


of test results and performing a moment curvature study such as those in
Sections 4.*.* and 4.*.*.

ATC-115

3: Key Design Issues

3-65

95% Draft
3.6.3.4

Strong Column-Weak Beam Behavior

Section 18.7.3.2 of AC1 318-14 includes provisions for strong column-weak


beam behavior. It requires that the sum of the nominal flexural strengths of
columns framing into a joint be greater than six-fifths times the sum of the
nominal flexural strengths of beams framing into the joint. The required
ratio was originally determined largely through engineering judgment, at a
time when the yield strength of flexural reinforcement used in concrete
moment frames was 60 ksi. The purpose of the provision is to promote
yielding and formation of plastic hinges in beams before yielding and
hinging takes place in columns. Hinging of beams spreads story drifts more
uniformly throughout the building height than hinging of columns; moreover,
hinging of columns can create story collapse mechanisms. A story collapse
mechanism would not arise if plastic hinges formed at the ends of all beams
at the top and bottom of the same story.
FEMA P695 (FEMA, 2009) indicates that the current strong column-weak
beam requirements do not preclude column hinging. Although the provisions
protect against a single story mechanism forming, multi-story collapse
mechanism will form that include only a fraction of the total number of
stories in a building. Eliminating multi-story mechanisms and thus column
hinge, except at the base and underside of roof, is unrealistic and
unnecessary. However, the collapse mechanisms that do form need to spread
yielding over enough stories that the collapse resistance to maximum
considered earthquake ground shaking meets the goal intended by ASCE 710, i.e. less than 10% probability of collapse for maximum considered
earthquake using FEMA P695 analyses.
The effect of high-strength reinforcement on the strong column-weak beam
provisions in ACI 318-14 should be investigated further because the
mechanical properties of high-strength reinforcement may be less favorable
to the spread of yielding than the properties of Grade 60 bar are. In
particular, the tensile-to-yield strength ratio is likely to be less. Response
history nonlinear analyses could be used to explore whether the existing sixfifths ratio is at least as affective for high-strength reinforcement as it is for
Grade 60 reinforcement.
Recommendation to Study Strong Column-Weak Beam Behavior

Analytical study of special moment frames include exploring strong columnweak beam behavior is recommended in Section 4.*.*. Ideally, the analytical
models would be calibrated to results of beam, column, and joint reverse
cyclic load tests in Section 4.9.8.

3-66

3: Key Design Issues

ATC-115

95% Draft
3.6.3.5

Beam-Column Joints

The depth of a beam-column joint should be sufficient to transfer load from


yielding reinforcement to adjacent concrete via bond. Current rules and
guidance are empirical and based on tests of conventionally reinforced
components and frames. High-strength reinforcement will develop greater
axial load for a given bar diameter and deeper joints will be required than
those at present if concrete tensile strength limits the bond stresses that can
be developed.
Depth of Interior Joints

Section 18.8.2.3 of ACI 318-14 requires that the depth of interior beamcolumn joints be 20 times the diameter of the largest sized beam bar passing
through a joint (20db) of normal weight concrete. This is a relatively shallow
depth when considering that a beam bar may be simultaneously yielding in
tension on one side of the joint and in compression on the opposite side. The
intention of ACI 318 is not to prevent slip of bar at and even through the
joint but rather to ensure that the performance of the building is consistent
with that intended by ASCE/SEI 7-10. In joint tested by Zhu and Jirsa,
(1983) and by Lin, et al. (2000), bond failure of longitudinal beam bars
occurred at drift levels that today can be equated to those expected for
maximum considered earthquake ground motion, where the ability to repair
the joints may no longer be an issue.
The required depth for use of high-strength reinforcement was explored and
recommendations are made in NIST GCR 14-917-30. For Grade 80 and 100
bars, minimum joint depths of 26db and 35db, respectively, are
recommended. Formulae of varying complexity are provided to account for
the major factors that influence the required joint depth, including bar yield
strength, top bar effects, column axial load, and concrete compressive
strength.
Thirteen tests of interior beam-column joints were performed as part of
Japans New RC Project and a formula for joint depth was developed from
the results of these tests. Similar factors are considered in the Japanese
formula as those in the NIST GCR 14-917-30 but the factors are treated
differently enough that further study is warranted.
A review of existing tests of joints with high-strength concrete should be
performed and additional tests should be performed to confirm the
recommendations in NIST GCR 14-917-30 or to make new
recommendations for the required joint depth. The joint depth requirements
should help deliver the expected performance in earthquake shaking.

ATC-115

3: Key Design Issues

3-67

95% Draft

Depth of Exterior Joints

At exterior joints, beam longitudinal bars are usually terminated within the
joint with 90-degree hooks or headed bars located near the exterior side of
the joint. The depth of the joint needs to be sufficient to develop these beam
bars within it, and the hooks or heads must terminate near the exterior side of
the joint so that a strut can develop diagonally through the joint to resist the
joint shear.
A concern associated with terminating high-strength bars with standard 90degree hooks in exterior joints is the potential for increased bar slip as a
result of crushing of the concrete, due to the larger local contact stresses
around the bend in the hooked bars. In Japan, this and the possibility that a
joint will split as described in the following section is addressed by requiring
concrete compressive strength of at least 11 ksi for joints in which USD685,
Grade 100 reinforcement is terminated with 90-degree hooks (Aoyama,
2001; NIST GCR 14-917-30). This requirement is based on the results of
tests that used rectangular hoops around the perimeter of the beam-column
joint, which is noncompliant with ACI 318 requirements because the spacing
between legs of the hoops and cross ties exceed the maximum allowed
distance and the area hoop reinforcement was less than required by ACI 318.
Generally, Japanese buildings are stiffer and stronger than U.S. buildings
designed for the same earthquake shaking, so they are only trying to achieve
drifts of 1.5% (Otani, 1991) compared to drift levels of 2.5 to 4% for U.S.
buildings. For this reason, the results of the tests performed in Japan cannot
be used to draw conclusions for practice in the United States.
No exterior joint tests with Grade 80 or higher longitudinal bar and
compliant with ACI 318-14 joint detailing have been identified. However as
reported in NIST GCR 14-917-30, a shake table test of a five-story building
built at full-scale (Chen et al., 2012) demonstrated good performance of
exterior joints in which high-strength beam bars terminated with 90-degree
hooks. Special moment frames, consisting of 12-inch wide by 30-inch deep
beams and 18-inch wide by 26-inch deep columns, provided the lateral
resistance in this building. The beams in the lower two floors were
reinforced with ASTM A1035 Grade 120 steel bars, consisting of two No. 7
bars, top and bottom. The measured yield and ultimate tensile strength of
these bars were 130 ksi and 160 ksi, respectively. The compressive strength
of the concrete averaged 7.5 ksi on the day of testing. The joint had
transverse reinforcement compliant with ACI 318 and the hooked bars were
extended to near the back of the joint for an embedment of 21.5 inches,
which slightly exceeds the hooked bar development length, ldh, equal to 20
inches calculated in accordance with ACI 318-11 (ACI, 2011). The building
3-68

3: Key Design Issues

ATC-115

95% Draft
was subjected to input ground motions that caused rotation demand on the
beams to exceed 0.06 radians and fractured some of the No. 7 top and bottom
bars. Inspections indicated no distress in the beam-column joints in the
building.
Although the results of the shake table tests appear promising, controlled
tests of cyclically loaded exterior joint assemblies are needed such as those in
Section 4.*.*.
ACI 318-14 restricts the use of headed deformed bars to Grade 60 bar but
there is interest to increase the maximum grade to reduce congestion in
exterior joints. ACI 318-14 only requires Class HA heads, that is, the bar-tohead connection only need develop the minimum specified tensile strength of
the bar. For special moment frames, the bar may slip within the joint that
could result in the bar-to-head connection being more heavily loaded. Thus,
a head that develops the specified elongation of the bar may be required.
ASTM A970 includes a Class B head that requires this, but a new Class HB
may need to be introduced into ASTM A970 Annex A1 to comply with other
deformation obstructions and bearing face feature limitations. Further, ACI
318-14 requires that heads have a net bearing area of four times the area of
the bar (4Ab). Heads with a net bearing area of up to nine times the area of
the bar are available and these may be needed if bar slip occurs. Cyclic tests
of exterior joints constructed with headed bars are needed such those in
Section 4.*.*.
Shear Reinforcement for Exterior Joints

The potential for splitting failure in a joint was identified by Marques and
Jirsa (1975). Beam-column exterior joints anchoring high-strength beam and
column reinforcement have an increased potential for splitting as a result of
increased bar forces. Although inspection of the exterior joints in the shake
table test reported by Chen et al. (2012) showed that the use of ASTM A1035
Grade 120 longitudinal reinforcement did not result in joint splitting, the
joint only had two No. 7 bars top and bottom for a 12-inch wide beam.
Joints in which more closely placed bars are terminated will develop larger
splitting forces and so further research is required. Such tests should be used
to determine whether a change in the amount of joint transverse
reinforcement is required if high-strength longitudinal reinforcement is used.
Recommendations for Beam-Column Joint Studies

Tests and numerical studies on joints of special moment frames with beams
and columns constructed with high-strength reinforcement are described in
Section 4.*.*.

ATC-115

3: Key Design Issues

3-69

95% Draft
3.6.4

Flexure-Critical Special Structural Walls

ACI 318 provisions for special structural walls include requirements for
shear strength, axial load-moment (P-M) interaction, boundary elements,
coupling beams, and wall piers. These provisions are intended to protect
against brittle or limited ductility states, and to encourage energy-dissipating
ductile mechanisms. Other than for confinement, all reinforcement in special
structural walls is limited to Grade 60.
Slender structural walls are flexure-critical, with height-to-length aspect
ratios generally of 3 and greater, but for walls specifically design and
detailed for flexural behavior this ratio can be as low as 2. Flexure-critical
walls are expected to be ductile, which is possible if shear failure is
prevented.
Axial load-moment (P-M) design for vertical reinforcement of special
structural walls follows the same provisions applicable to other compression
members. However, the end regions of special structural walls are required
to have transverse reinforcement to resist vertical bar buckling and in end
regions expected to experience large concrete compressive strains, transverse
reinforcement is required to provide compression strain ductility similar to
special moment frame columns.
Diagonally reinforced coupling beams of special structural walls are highly
ductile if confined in accordance with ACI 318-14 and often required for
coupled walls. These beams are difficult to construct due to reinforcement
congestion.
Wall piers are vertical segments of a shear wall system with geometric aspect
ratios similar to columns. As such, many of the requirements for special
moment frame columns are applied for the purpose of providing ductile
behavior in these elements. These piers are often heavily reinforced.
The use of high-strength reinforcement has the potential to reduce the area of
reinforcement required for special structural walls and the number of bars
that must be handled. It also has the potential to reduce the congestion of
boundary elements and wall piers. Reinforcement congestion can limit the
design strength that can be provided with coupling beams. Performance
verification of high strength reinforcement for diagonals in diagonally
reinforced beams and for longitudinal bars and hoops in conventionally
reinforced beams would allow for greater nominal shear strength for given
coupling girder dimensions.

3-70

3: Key Design Issues

ATC-115

95% Draft
There are few tests of walls with high-strength reinforcement that are directly
applicable to slender wall design common in the U.S. Tests of structural
walls with high-strength reinforcement are identified and summarized in
NIST GCR 14-917-30. Only three wall tests with Grade 100 reinforcement
by Kimura and Ishikawa (2008) are similar in configuration to walls
designed in the U.S. Kimura and Ishikawa tested three 1/5-scale,
rectangular, slender wall specimens. These walls were reinforced with
SD685 bars (nominal yield strength of 100 ksi) in the boundary elements and
the horizontal and vertical web reinforcement. The shear span ratio of the
tests was 2.0. Although the transverse reinforcement area did not comply
with the requirements of ACI 318 for special boundary elements, the spacing
and configuration requirements did. The transverse reinforcement spacing to
longitudinal bar diameter ratio, s/db, was 5. Hysteresis loops showed no
degradation in strength up to the ultimate drift. The ultimate drift was 1.5%
for the wall with an axial stress ratio of 0.15 (axial load ratio is the wall
compression load divide by the horizontal cross-sectional area of the wall
times the wall compressive strength, P/Agfc), and 2% for the walls with an
axial stress ratio of 0.10. All walls exhibited a flexural compression failure
mode in which the concrete crushed and the boundary bars buckled. The
length of bar yielding at the base of the specimens was about half the wall
length. The drift levels achieved in the Kimura and Ishikawa tests are on the
low end of what is expected of walls with Grade 60 reinforcement. Slender
walls reinforced with Grade 60 bar generally reach an ultimate drift of 2%.
Two sets of tests of walls with high-strength reinforcement, yield strength
close to 80 ksi, are reported by Dazio et al. (2009) and Lowes et al. (2012).
Dazio et al. tested six walls that, in comparison with the averages found in
the survey of shear walls constructed on the West Coast (Lowes et al., 2012),
had low reinforcing ratios in the boundary elements and low-to-moderate
reinforcement ratios of distributed longitudinal reinforcement. In the wall
specimens the axial load, quantity of reinforcing steel, steel elongation, and
transverse reinforcement at the ends of the walls varied. The transverse
reinforcement for the boundary elements did not comply with the ACI 318
requirements of special boundary elements, but the test results of four
specimens demonstrate trends in how walls will respond to earthquakes.
The need for ductile reinforcement that strain harden was identified because
in a wall Dazio, et al., tested in which typical vertical bars with uniform
elongation of 2.3% ruptured at 0.6% drift and boundary element bars with
uniform elongation of 4.6% ruptured at 1% drift. Rupture of these bars also
appears influenced by low reinforcing ratios of 0.3% for typical vertical wall
bars and 1.3% for boundary bars. Another Dazio, et al., test demonstrated

ATC-115

3: Key Design Issues

3-71

95% Draft
that a ratio of transverse bar spacing to longitudinal bar diameter of 7.5 does
not adequately resist buckling of longitudinal boundary element bars. All six
longitudinal bars in a boundary element buckled at a drift of 1.4% and
fractured during loading in the reverse direction.
A pair of tests by Dazio, et al., further highlights the benefit of closer spaced
hoops in boundary elements. The two walls had the same vertical and
horizontal wall reinforcement. One wall had a ratio of transverse
reinforcement spacing to longitudinal bar diameter, s/db, of 6.3 and an axial
load ratio, P/(Agfc), of 0.058. The other wall had a ratio of transverse
reinforcement spacing to longitudinal bar diameter, s/db, of 4.2 and an axial
load ratio of 0.11. Higher axial load ratios reduce deformation capacity if
everything else is equal. Even though the second wall had nearly twice the
axial load ratio, the deformation capacities were similar. The onset of
boundary element bar buckling for both walls occurred at a drift of 1.7%, and
the maximum drift was about 2% for both walls demonstrating improvement
in performance with a closer spacing of transverse reinforcement in the
boundary members.
Two walls had low vertical bar reinforcement ratios, 1.3% for the boundary
element bars and 0.3% for typical vertical reinforcement. Rather than
spreading the bar yielding, a wide horizontal crack formed resulting in
fracture of vertical bars. Similar poor behavior in walls was observed
following the Chilean Earthquake in 1985 (Wood, 1989) and Christchurch,
New Zealand earthquake in 2011. The poor performance of the Dazio et al.
walls in the laboratory, walls in the field, and the research by Wood
demonstrate that a minimum amount of vertical reinforcement is required,
noting that the formation of a single crack and the localization of inelastic
action is a concern for all grades of reinforcement. The required minimum
reinforcement ratio needs to be studied for conventional and high-strength
reinforcement.
Lowes et al. (2012) tested four walls, two of which had longitudinal
boundary element reinforcement with yield strength of 85 ksi. The boundary
element details complied with the requirements of ACI 318-11 (ACI, 2011).
The bases of the two walls with Grade 85 boundary element reinforcement
included splices of the vertical boundary element and web reinforcement.
For one of the walls, boundary element vertical bar buckling began during
the second cycle to 1% drift, and these bars fractured below the splice during
the second cycle to 1.5% drift. For the other wall, boundary element vertical
bar buckling occurred during the third cycle to 0.75% drift, core damage was
observed above the splices at 1.5% drift, and the concrete crushed
extensively and bars buckled above the splice at 1.05% drift during the
3-72

3: Key Design Issues

ATC-115

95% Draft
loading cycle to 1.5% drift. The results of these tests show that the yielding
of bars can occur at both the top and bottom of a splice and that strain can be
concentrated at the base of the wall, resulting in bar fracture.
Recent studies by Luna et al. (2014) on low aspect ratio (squat) rectangular
walls with Grade 60 reinforcement and normal-strength, normal-weight
concrete have shown the effects of bidirectional loading to be important in
terms of peak in-plane shear strength, in-plane stiffness and the deterioration
of strength and stiffness with repeated cycling at large story drifts. The outof-plane loadings, which included displacement and twisting, were not
controlled. The inelastic cyclic response of rectangular and flanged walls
under bidirectional horizontal loading is not well understood and a family of
tests focused on the effect of controlled bidirectional loading histories is
needed. Part of that testing program would address the use of high-strength
reinforcement and its use to a) reduce bar size for a given spacing, and/or b)
increase bar spacing for a given bar size.
The adequacy of walls reinforced with Grade 80 bars cannot be proven
because there are little available data. Although the minimum specified yield
strength of Grade 80 reinforcement is not appreciably greater than the yield
strength of the Grade 60 bar delivered to construction sites nowadays, the
expected yield strength of Grade 80 bars will likely exceed 90 ksi for which
there are little-to-no data. Tests of flexure-critical walls, including coupled
walls are needed to support a code change to allow the use of high-strength
reinforcement in boundary elements, wall webs and coupling girders.
Recommendation for Flexure-Critical Wall Studies

Tests and numerical studies on flexure-critical walls constructed with highstrength reinforcement are described in Section 4.*.*.
3.6.5

Shear-Critical Walls

The seismic response of low aspect ratio walls, also referred to as squat or
shear-critical walls, is poorly understood, irrespective of reinforcement and
concrete strengths, wall aspect ratio, and rebar distributions.
Reinforcement serves multiple purposes in squat reinforced concrete walls:
horizontal web reinforcement delays diagonal tension failure; vertical web
reinforcement anchors the diagonal compression struts; vertical web
reinforcement delays sliding at horizontal construction joints; web
reinforcement confines the diagonal compression struts; and vertical
boundary reinforcement anchors compression struts, improves the integrity
of the struts at the toes of the wall, and controls crack widths at the wall

ATC-115

3: Key Design Issues

3-73

95% Draft
boundaries (Luna et al. 2014). The roles of reinforcement in squat walls will
vary as a function of drift ratio, with roles changing at drift ratios greater than
those associated with peak shearing strength. These roles are not understood
for conventional reinforcement in walls with rectangular cross sections and
uniform distributions of web reinforcement. The hysteretic response of walls
equipped with boundary columns or flanges, which are commonplace in
building construction, will be more complicated than for rectangular walls
given the number of additional cross section and rebar variables, and the
unknown effects of seismic loading perpendicular to the plane of the web
(and in plane of flanges) that will introduce out-of-plane flexure and shear.
The impact of replacing conventional reinforcement with high strength
reinforcement will depend in part on whether it is used to reduce bar size for
a given spacing and/or increase bar spacing for a given bar size, and on the
anticipated range of response. If the goal is to increase bar spacing, the
impact could be significant.
Reverse cyclic testing of large-size squat reinforced concrete walls will be
needed to characterize the effect of replacing conventional reinforcement
with high strength reinforcement. Walls with different aspect ratios, concrete
strengths, reinforcement grades, and plan geometries (rectangular, barbell
and flanged) should be evaluated for the purpose of ensuring that the use of
high strength reinforcement does not lead to poorer performance than walls
with an equivalent shearing capacity (as measured by ACI 318-14 Chapter
18) constructed with conventional reinforcement. Benchmark tests should
first be performed on walls constructed with Grade 60 reinforcement and
prescriptive details consistent with ordinary and special construction.
Although squat structural wall specimens with Grade 100 reinforcement have
been tested in Japan (Aoyama, 2001; Kabeyasawa and Hiraishi, 1998), these
walls had very thin webs and column-type boundary elements. Tests were
performed to study the: (1) strength of walls that first yield in flexure but fail
due to shear-compression (web crushing), (2) impact of bi-directional
loading, and (3) shear strength of walls. The walls failed predominately due
to concrete crushing in the web, a brittle mode, at relatively low drifts. Few
squat walls are constructed in the United States with thin webs and barbells
and so the Japanese test results cannot directly support a code change to ACI
318 to permit the use of high strength reinforcement.
Recommendation for Shear-Critical Wall Studies

Tests and numerical studies on shear-critical and flexure-shear-critical walls


constructed with high-strength reinforcement are described in Section 4.*.*.

3-74

3: Key Design Issues

ATC-115

95% Draft
3.6.6

Diaphragms

The issues associated with use of high strength reinforcement in diaphragms


are similar to those discussed previously for flexure- and shear-critical walls.
High strains in the reinforcement of diaphragm collectors and chords can
lead to excessive cracking of the diaphragm. ACI 318-14 Section 18.12.7.2
recognizes this when it limits the stress from design earthquake forces to
60,000 psi for bonded tendons, which are often used as diaphragm chord
reinforcement. NIST (2010a) writes
Although stress in other collector and chord reinforcement is not
limited, consideration should be given to deformation compatibility
between tension chords, collectors, and the floor slab. High tensile
stress and strains in collectors and chords can result in excessive
cracking that will migrate into the slab.
However, given that deformation demands on diaphragms are likely smaller
than those in the critical regions of shear walls, there is a basis for accepting
Grade 80 reinforcement for diaphragms, especially recognizing that the
expected yield strength of Grade 60 reinforcement is not appreciably less
than 80 ksi. Tests of diaphragms are not considered to be a high priority and
none are proposed in Chapter 4. That said, data from the test of flexure- and
shear-critical walls should be mined for the purpose of assessing implications
for diaphragms, noting that a) tests of walls normally produce peak shear and
bending moment at one location, and b) diaphragms may experience peak
shears and bending moments at different locations along its span(s).
Recommendation for Studies of Diaphragms

At this time no additional engineering studies specifically addressing


diaphragms is required. Tests of low-aspect ratio walls can be reviewed to
obtain information about the likely behavior. ACI Committee 318 should
decide whether high-strength reinforcement can be used for diaphragms.

ATC-115

3: Key Design Issues

3-75

95% Draft

Chapter 4

Research Studies

In most cases, issues related to production and fabrication of high-strength


reinforcement, and design and construction using high-strength
reinforcement, require further study or research to resolve. This chapter
outlines the experimental research and engineering studies needed to
adequately investigate the use of high-strength reinforcement for use in
general reinforced concrete design.
4.1

Objectives

The objectives of this research plan are to: (1) further investigate the use of
high-strength reinforcement; and (2) support the development of code change
proposals for ACI 318, Building Code Requirements for Structural Concrete
and Commentary (ACI, 2014) to allow the general use of reinforcing
steel bars in excess of Grade 60.
4.2

Overview

Studies are related to the issues identified and discussed in Chapters 2 and 3.
They are intended to resolve the issues in sufficient detail to substantiate a
code change, where recommended. Each research or study topic includes a
description of; (1) the purpose for the work; (2) details of the study; (3) the
recommended team; and (4) the anticipated timeline. Study topics are
prioritized in Chapter 5.
4.3

Bar Production and Specification

Studies related to bar production and specification are summarized in Table


4-1.
Table 4-1
Reference
Section

Summary of Studies Related to Bar Production and Specification


Research or Engineering Study

Estimated
Timeline

4.3.1

Mechanical Properties of Recent Heats of High Strength Reinforcement

6 months

4.3.2

Detailed Mechanical Property Tests of Grade 100 and Grade 120


Reinforcement

18 months

ATC-115

4: Research Plan

4-1

95% Draft
4.3.1

Mechanical Properties of Recent Heats of High Strength


Reinforcement

Collect and present mechanical properties of recently rolled heats of highstrength reinforcement.
Purpose of Study:
The primary objective of this study is to determine the range of values for
key mechanical properties of reinforcing bars that mills are currently capable
of producing. The results from this study will help identify the mechanical
properties of reinforcement that need to be included as variables in a testing
program involving beam, column, and wall specimens. The data will help
determine the minimum acceptable values for mechanical properties of
reinforcing bars to specify in future ASTM standards, especially the tensileto-yield strength ratio and the uniform elongation. Other data of interest
include evaluating the effects of bend radius, bar deformations, and bar
marks on bend tests.
Details of Study:
A researcher with the assistance of the Concrete Reinforcing Steel Institute
(CRSI) and some of its member producers will collect and analyze the data.
This will allow the producer of each heat of bar to remain anonymous while
at the same time allowing the process of establishing reinforcing bar
properties to include in an ASTM International specification to move
forward. The data will be presented in a series of charts and tables. The data
collected should include yield strength and length of yield plateau, tensile
strength, measured tensile-to-yield strength ratio, uniform elongation and the
means used to measure the uniform elongation, and fracture elongation.
Perform bend tests on a sample of bars from various producers to gain insight
as to whether bars are likely to fracture at bends. Perform bend-rebend tests
on a sample of bars and evaluate strain-aging characteristics.
Team:
One researcher, one part-time graduate student, and one part-time laboratory
technician to assist.
Timeline:
Approximately 6 months.

4-2

4: Research Plan

ATC-115

95% Draft
4.3.2

Detailed Mechanical Property Tests of Grade 100 and


Grade 120 Reinforcement

Collect test data from producers and perform tests and statistical analyses of
mechanical property data of heats of high-strength reinforcement.
Purpose of Study:
Obtain data based on tests of reinforcing bars conforming to the ASTM
specifications currently in development in which deformed high-strength
steel bars will be specified. The main goal is to determine properties of
primary interest to structural designers specifying the use of high-strength
reinforcement. The study is necessary to ensure that bars are produced with
adequate strength and ductility.
Details of Study:
Perform tests of No. 4 to No. 18 bars and larger bars, if being produced, to
collect the following data:

Complete stress-strain curve with values for modulus of elasticity, yield


strength, length of yield plateau, tensile strength, uniform elongation, and
total elongation (across fracture) measured during standard tensile tests;

Low cycle fatigue characteristics at ambient and low temperature


conditions measured during large strain amplitude cyclic tests;

Toughness of bars at various temperatures and transition temperature


using Charpy V-notch tests;

Bend tests;

Strain-aging characteristics using bend-rebend tests; and

The tests should be conducted on reinforcing bars from a minimum of three


mills and three heats per mill. Other measurements of interest include bar
deformations and chemical composition, and whether the bending
requirements are met. These may be collected and analyzed with the
assistance of the Concrete Reinforcing Steel Institute (CRSI) and some of its
member producers. The data will be presented in a series of charts and table
including results from statistical analysis.
Team:
One researcher, one part-time graduate student, and one part-time laboratory
technician to assist.

ATC-115

4: Research Plan

4-3

95% Draft

Timeline:
Approximately 18 months.
4.4

Strength of Members

Studies related to strength of members are summarized in Table 4-2.


Table 4-2

Summary of Studies Related to Strength of Members

Reference
Section

Research or Engineering Study

Estimated
Timeline

4.4.1

Flexural Strength and Tensile Strain Limits

4.4.2

Deflection Warning for Flexural Members Subjected to Gravity Loads

4.4.3

Column Strength

4.4.4

Tension Regions of Shells and Folded Plates

3 years

4.4.5.1

One-Way Shear in Beams without Shear Reinforcement

3 years

4.4.5.2

Two-Way Shear in Slabs without Shear Reinforcement

3 years

4.4.6

Shear Strength of Beams with Shear Reinforcement

3 years

4.4.7

Shear Friction

3 years

4.4.8

High-Cycle Elastic Fatigue of High-Strength Reinforcing Bars

3 years

4.4.1

18 months
2 years
3 months

Flexural Strength and Tensile Strain Limits

Perform an analytical study in which moment-curvature relationships are


developed for beams, columns, and walls using various stress-strain curve
shapes and mechanical properties for the bars and concrete.
Purpose of Study:
Determine the effects of different shapes of stress-strain curves for highstrength reinforcing bars on moment-curvature relationships of concrete
members. The analytical study will (1) evaluate the compression-controlled
and tension-controlled strain limits and (2) evaluate whether the traditional
design assumptions in ACI 318 to calculate flexural strength (stress block for
concrete and elastic-plastic stress-strain curve for steel) are appropriate with
the use of high-strength reinforcement. Using results from the momentcurvature analysis, the study should include an evaluation of the effects of
the maximum reinforcement tensile strain on deformation capacity of
reinforced concrete members and determine whether a maximum strain limit
is required to avoid fracture of the reinforcement prior to concrete crushing.
Details of Study:

4-4

4: Research Plan

ATC-115

95% Draft

An analytical study should be conducted to evaluate the influence of the


various parameters that define the steel stress-strain curves on momentcurvature relationships and on the monotonic deformation capacity of
reinforced concrete members, including beams, columns, and walls. The
study should include review of existing experimental data to allow for
comparison with results from the analytical study. The following variables
should be included:

Compressive strength of concrete varying from 4,000 to 12,000 psi

Confined and unconfined concrete

Yield strength of reinforcement

Reinforcement tensile-to-yield strength ratios

Reinforcement post-yield modulus

Shear-span ratios

Reinforcement ratios

Axial load ratios

Based on results of the study, recommendations should be provided


regarding the compression-controlled and tension-controlled strain limits as
well as the addition of a maximum tensile strain limit. In addition,
recommendations should be provided if changes are required to the
traditional methods of calculating flexural strength.
Team:
One researcher and one part-time graduate student.
Timeline:
Approximately 18 months.
4.4.2

Deflection Warning for Flexural Members Subjected to


Gravity Loads

Perform an analytical study and confirmation testing to determine whether


adequate deflection warning is provided by a beam reinforced with highstrength reinforcement that has a tensile-to-yield strength ratio less than 1.25
and determine the effect of the shape of the stress-strain curve, when a beam
is overloaded.

ATC-115

4: Research Plan

4-5

95% Draft

Purpose of Study:
The purpose of this study is to determine whether overloaded flexural
members reinforced with high-strength reinforcement that has a tensile-toyield strength ratio less than 1.25 develop sufficient deflections to warn of
impending failure. The importance of the shape of the stress-strain curve,
and in particular whether a yield plateau is necessary and what uniform
elongation for bars is needed for gravity loaded flexural members, should
also be explored as part of this study. Tests of beams should be performed to
confirm the conclusions of the analytical study.
Details of Analytical Study:
Moment curvature analyses and nonlinear beam deflection analyses should
be developed for a range of beams and slabs consistent with current practice.
This study can complement that outlined in 4.4.1. The study should start
with a literature search to determine whether applicable tests are available for
the range of bar properties being included in this analytical study. Initial
models should be calibrated to test results, which will assist in establishing
the level of detail for beam modeling. The members and sections analyzed
should vary approximately as follows:

Member depths should range from one to two times the minimum depths.

Bars should be up to Grade 120.

Bar stress-strain curves should have a range of shapes with varying


values of tensile-to-yield strength ratio, uniform elongation, and length
of yield plateau.

Details of Test Program:


Tests should be designed to determine whether the results of the analytical
study are valid. The number and range of specimens should target the limits
of the bar properties and member sections studied analytically. The tests
should attempt to track bar elongations as the beam is loaded to failure.
These tests should also explore the net tensile strain needed for various bar
stress-strain curve shapes and mechanical properties. The number of test
specimens anticipated is on the order of 10.
In addition to the main focus of this study, the tests should be used to track
deflections at service load levels and, if possible, to explore effectiveness of
crack control requirements.

4-6

4: Research Plan

ATC-115

95% Draft

Team:
One researcher, two full-time graduate students (two years), and one parttime laboratory technician to assist.
Timeline:
Approximately 2 years.
4.4.3

Column Strength

Perform an analytical study in which moment-curvature relationships are


developed for columns using various stress-strain curve shapes and
mechanical properties for the bars and concrete.
Purpose of Study:
Determine the effects of high-strength reinforcement on columns and
evaluate whether the current code limit for the compressive strength of the
reinforcement remains applicable or if this limit can be removed. In
addition, this study should evaluate whether traditional design assumptions in
the development of interaction-curves are appropriate with the use of highstrength reinforcement.
Details of Study:
An analytical study using moment-curvature should be conducted to evaluate
the influence of the various parameters that define the steel stress-strain
curves. This study should be a complement to the study outlined in 4.4.1. In
addition, this study should investigate the effects of creep of the concrete and
in particular, creep transfer of stress from the concrete to the reinforcement.
The following variables should be included:

Compressive strength of concrete varying from 4,000 to 12,000 psi

Confined and unconfined concrete

Yield strength of reinforcement

Reinforcement tensile-to-yield strength ratios

Shape of the stress-strain curve

Reinforcement ratios

Axial load ratios

Based on results of the study, recommendations should be provided


regarding the elimination of the current code maximum stress limit for
compression reinforcement.

ATC-115

4: Research Plan

4-7

95% Draft

Team:
One researcher and one part-time graduate student.
Timeline:
Approximately 3 months (assumed to occur in combination with the study
outlined in Section 4.4.1).
4.4.4

Tension Regions of Shells and Folded Plates

Perform analytical and experimental studies to evaluate effects of reduced


stiffness and associated increases in tensile strain and cracking with the use
of high-strength reinforcement for shells and folded plates.
Purpose of Study:
The purpose of this study is to determine whether the reduced stiffness,
increased bar strains, and increased cracking are acceptable for tension
regions of shells and folded plates.
Details of Study:
This study should consist of tension testing of reinforced concrete panels to
quantify the reduced tension stiffness, increase in bar strains, and increase in
cracking. Using the results of the tests, the effect of reduced stiffness should
be simulated in finite element analyses that incorporate the reduced stiffness
in tension regions.
The test program should explore the effect of tensile strength of concrete,
shrinkage of concrete, bar size, and duration of loading. The tensile strength
and shrinkage can be assessed by using three different concrete mix designs.
At least two different bar sizes should be used in the panels to determine
whether a bar size effect exists. At least two panels should be placed in
constant tension for a period of about one year to determine the effect of
long-term loading. The number of tests should be established to strategically
test for different parameters. Approximately 10 test panels should be
targeted. The analyses should consist of shells and folded plates in which
tension stiffness is varied based on the test results to determine the effects on
the behavior of shells and folded plates. Approximately five models of
various geometries should be analyzed.

4-8

4: Research Plan

ATC-115

95% Draft

Team:
One researcher, one full-time graduate student (three years), one part-time
graduate student (one year), and one part-time engineering practitioner for
structural advice and consultation.
Timeline:
Approximately 3 years.
4.4.5

Shear Strength of Beams and Slabs without Shear


Reinforcement

Perform an experimental study to determine whether shear strength design


expressions are appropriate for calculation of Vc for flexural members
incorporating high strength reinforcement.
Purpose of Study:
Evaluate whether the current shear strength equations are appropriate with
the use of high-strength reinforcement. The use of high strength
reinforcement provides the opportunity for the use of lower longitudinal
reinforcement ratios. However, low reinforcement ratios have the potential
to result in shear strengths lower than that calculated using current ACI 318
design expressions for both one-way and two-way shear. For both one-way
and two-way shear testing, it is anticipated that only targeted testing will be
required to verify that the test results from specimens utilizing high-strength
reinforcement fit within the band of existing test results.
4.4.5.1

One-Way Shear in Beams without Shear Reinforcement

Details of Study:
An experimental study should be conducted to evaluate the one-way shear
strength of reinforced concrete beams incorporating high-strength
longitudinal reinforcement. For this testing program, it is considered that
rectangular, simply supported beams subjected to a region of constant shear
(a/d =3) would be appropriate. In addition, a limited number of continuous
beam tests would be useful to evaluate the influence of the M/V ratio and
provide perspective on the realistic situation in frame structures of high shear
accompanied by high moment. A number of variables should be considered,
including:

Reinforcement ratio: 0.25%, 0.5%, 0.75%, 1%

Reinforcement grade: 60, 100, and, if available, 120

Concrete strength: 5ksi, 10 ksi

ATC-115

4: Research Plan

4-9

95% Draft

Member depth: 12 in., 24 in., 48 in.

M/V ratio

The number of test specimens anticipated is on the order of 30.


Team:
One researcher, one full-time graduate student (three years), one part-time
graduate student (18 months), and one part-time laboratory technician to
assist.
Timeline:
Approximately 3 years.
4.4.5.2

Two-Way Shear in Slabs without Shear Reinforcement

Details of Study:
An experimental study should be conducted to evaluate the two-way shear
strength of reinforced concrete slabs incorporating high-strength longitudinal
reinforcement. For this testing program, it is considered that a standard
punching shear specimen with a concentrated load applied through a column
stub would be appropriate. Both concentric axial loading and combination
moment-axial loading should be considered to evaluate the influence of high
shear accompanied by high moment. A number of variables should be
considered, including:

Reinforcement ratio: 0.25%, 0.5%, 0.75%, 1%

Reinforcement grade: 60, 100, and, if available, 120

Concrete strength: 5 ksi, 10 ksi

Member depth: 4 in., 12 in., 24 in.

M/V ratio

The number of test specimens anticipated is on the order of 20.


Team:
One researcher, one full-time graduate student (three years), one part-time
graduate student (18 months), and one part-time laboratory technician to
assist.
Timeline:
Approximately 3 years.

4-10

4: Research Plan

ATC-115

95% Draft
4.4.6

Shear Strength of Beams with Shear Reinforcement

Perform an experimental study to determine whether shear strength design


expressions are appropriate for calculation of Vn for flexural members
incorporating both longitudinal and transverse high-strength reinforcement
and address the minimum shear reinforcement requirement.
Purpose of Study:
Evaluate (1)whether the yield strength is the appropriate parameter to use in
the design equation for Vs and determine whether the addition of the concrete
contribution Vc and the shear reinforcement Vs remains applicable, (2)
investigate serviceability issues regarding the width of shear cracks, and (3)
evaluate whether the minimum reinforcement ratio for shear is adequate.
Details of Study:
An experimental study should be conducted to evaluate whether the yield
strength of the transverse reinforcement can be fully developed at ultimate
shear strength. The use of yield strength is also important as it factors into the
minimum shear reinforcement requirements. In addition, one of the primary
reasons stated in ACI 318 to limit the yield strength of the reinforcement is to
control cracking. Therefore, it is important that this study also evaluate crack
widths for various stress levels to determine whether serviceability issues
will develop with the full utilization of the yield strength of high strength
bars.
For this testing program, it is considered that rectangular, simply supported
beams subjected to a region of constant shear would be appropriate. In
addition, a limited number of continuous beam tests would be useful to
evaluate the influence of the M/V ratio and provide perspective on the
realistic situation in frame structures of high shear accompanied by high
moment. This study should build upon the results of the shear tests of beams
without shear reinforcement discussed in 4.4.5.1. A number of variables
would be considered, including:

Transverse reinforcement spacing

Transverse reinforcement grade: 60, 100, and, if available, 120

Transverse reinforcement stress-strain shape

Longitudinal reinforcement ratio: 0.5%, 1%

Longitudinal reinforcement grade: 60, 100, and, if available, 120

Concrete strength: 5 ksi, 10 ksi

ATC-115

4: Research Plan

4-11

95% Draft

Member Depth: 12 in., 24 in., 48 in.

a/d ratio: greater than 3

M/V ratio

The number of test specimens anticipated is on the order of 40.


Team:
One researcher, one full-time graduate student (three years), one part-time
graduate student (18 months), and one part-time laboratory technician to
assist.
Timeline:
Approximately 3 years.
4.4.7

Shear Friction

Perform an experimental study to determine whether the full yield strength of


shear-friction reinforcement can be realized in producing the normal force
across the shear transfer plane.
Purpose of Study:
Evaluate whether the yield strength of the transverse reinforcement crossing
the shear plane is the appropriate parameter to use in the design equation for
shear strength across an interface or whether a limit on the stress developed
by the reinforcement is required.
Details of Study:
An experimental study should be conducted to evaluate the interface shear
strength of concrete structures reinforced with high-strength reinforcement.
It is recommend that a direct shear test be performed on suitably sized
specimens so that realistic bar spacings can be provided as well as full
anchorage of the transverse reinforcement. A variety of specimens are
recommended to be tested to evaluate the range of interfaces that are covered
by the ACI code. In particular, the following interfaces are considered
needed: concrete placed monolithically, concrete placed against hardened
concrete with surface intentionally roughened, concrete placed against
smooth hardened concrete, and concrete placed against structural steel. In all
of these cases, various grades of reinforcing bars are provided to cross the
shear plane. For this testing program, it is only considered required that the
reinforcement be placed perpendicular to the shear plane consistent with the

4-12

4: Research Plan

ATC-115

95% Draft

majority of usage of shear friction reinforcement. The following additional


variables should be considered:

Interface shear reinforcement percentage

Shear reinforcement bar size

Grade: 60, 100, and, if available, 120

Concrete strength: 5 ksi, 10 ksi

The number of test specimens anticipated is on the order of 30.


Team:
One researcher, one full-time graduate student (two years), and one part-time
laboratory technician to assist.
Timeline:
Approximately 3 years.
4.4.8

High-Cycle Elastic Fatigue of Flexural Members

Perform tests to determine the high-cycle fatigue resistance of high-strength


reinforcement.
Purpose of Study:
The purpose of this study is to determine whether changes are required to the
AASHTO Bridge Design Specifications design provisions for high-cycle
elastic fatigue resistance.
Details of Study:
This study consists of testing for high-cycle fatigue resistance of highstrength reinforcing bars using in-air axial tension specimens (Jhamb and
MacGregor, 1974; Paulson and Hanson, 1991; Fei and Darwin, 1999; Zheng
and Abel, 1999).
The effect of bar production method should be explored. Specimens should
include bars produced by both quenching and tempering and micro-alloying,
with bars obtained from multiple producers and in multiple sizes for each
production method. The radius at the base of deformations, where the
transverse lug meets the barrel of the bar, should be measured (Helgason et
al., 1976). Representative bars should be examined to identify the presence
and severity of defects at the base of the deformations. The range of bars
sizes to be tested should consider both historical precedence and also the size

ATC-115

4: Research Plan

4-13

95% Draft

of bars likely to be commonly specified for bridge flexural members with


high strength reinforcement subjected to elastic fatigue stresses in tension.
Approximately 100 tension specimens of unspliced high-strength reinforcing
bars are recommended for testing.
Team:
One researcher and two full-time graduate students.
Timeline:
Approximately 3 years.
4.5

Serviceability

Studies related to serviceability are summarized in Table 4-3.


Table 4-3

Summary of Studies Related to Serviceability

Reference
Section

Research or Engineering Study

Estimated
Timeline

4.5.1

Deflection of Flexural Members

3 years

4.5.2

Crack Control of Flexural Members

2 years

4.5.1

Deflection of Flexural Members

Perform an analytical study on the deflection of flexural members reinforced


with high-strength reinforcement as well as an experimental study on the
long-term deflection of flexural members.
Purpose of Study:
The purpose of this study is to determine what changes are necessary to ACI
318 procedures for controlling deflections of flexural members. Specifically,
potential adjustments to minimum thickness tables in ACI 318-14 for oneway and two-way flexural members (Tables 7.3.1.1, 8.3.1.1, 8.3.1.2, and
9.3.1.1) will be evaluated for beams and slabs reinforced with high-strength
bars. Additionally, Equations 24.2.3.5a and 24.2.4.1.1 of 318-14 will be
evaluated for calculating deflections using an effective moment of inertia, Ie.
Details of Study:
Task 1: Short-Term Deflections
An analytical study should be conducted to predict the deflections of a
variety of flexural members including beams, one-way slabs, and two-way
slabs. For the one-way slab study, it is recommended that deflections be

4-14

4: Research Plan

ATC-115

95% Draft

calculated using moment-curvature analysis which considers the stress-strain


relationship of the high-strength reinforcement. For the two-way slab study,
a finite element analysis is recommended. It is important, however, that the
finite element analysis appropriately incorporate the moment-curvature
behavior of the slab reinforced with high-strength reinforcement. The
following variables are recommended for study:

Reinforcement grade: 60, 100, 120 ksi

Concrete strength: 4000 psi, 8000 psi, 12,000 psi

Span length

Member depth and width

Reinforcement percentage

For two way-slabs: with and without drop-caps

Loading: range of service loads

A study was conducted for the Charles Pankow Foundation by the University
of Washington and Magnusson Klemencic Associates (MKA) in 2013 which
computed deflections using a finite element analysis in which section
properties were varied. The results of this study can be used as a starting
point for member sizes and reinforcement. In addition to parametric
analyses, the analytical models should also be used to calculate deflections of
the beams that will be tested for one-way members as described in Section
4.4.1 and two-way slabs that will be tested as described in Section 4.4.4.2.
Comparison of the experimental and analytical results will enable evaluation
of the accuracy of the models.
Minimum Depth Tables:
Using the results of the analytical study and considering code-defined
deflection limits, appropriate span-to-depth ratios should be developed and
compared with the values presented in the ACI 318-14 minimum depth
tables. Recommended changes to Tables 7.3.1.1, 8.3.1.1, 8.3.1.2, and 9.3.1.1
of ACI 318 should be developed to address high-strength reinforcement.
Effective Moment of Inertia:
The results of the analytical study should be compared with results from
elastic theory using the effective moment of inertia method. Different
approaches to computation of the effective moment of inertia should be
considered and include (1) Bransons equation, (2) Bischoffs equation, and
(3) the cracked moment of inertia Icr. For both the Branson and Bischoff
equations, the use of a reduced cracking moment Mcr through reduction of
ATC-115

4: Research Plan

4-15

95% Draft

the modulus of rupture fr should be considered as previous research has


indicated that this parameter can be used to improve the estimation of
deflections computed by the Branson equation. This approach may not be
useful for the Bischoff equation but should be considered. For two-way
slabs, recommended finite element modeling techniques for the calculation
of deflections should be provided.
Task 2: Long-Term Deflections:
Long-term deflections are dominated by creep. While analytical models can
be used to estimate creep effects, experimental testing provides the best
information regarding this behavior. The design of the specimens for this
study should be influenced by the results of the analytical study, and beam
specimens are recommended. Test specimens should be loaded to service
levels which will allow for comparison and verification of analytical models
recommended in Task 1. In addition, these same specimens should be
subjected to sustained loads for one to two years allowing investigation of
long-term deflections. It is recommended that six specimens be tested at two
different levels of sustained loading using three different longitudinal
reinforcement ratios (0.2%, 0.8 % and 1.5%). The results of this study
should be compared with the results of the recommended design expression
from Part 1 amplified using the time-dependent factor from ACI 318.
Appropriate changes to the time-dependent factor, if required, should be
developed.
Team:
One researcher, one full-time graduate student (three years), one part-time
graduate student (18 months), one part-time laboratory technician to assist,
and one part-time engineering practitioner for structural advice and
consultation.
Timeline:
Approximately 3 years.
4.5.2

Crack Control of Flexural Members

Perform an analytical study of crack control of flexural members with


targeted confirmation testing.
Purpose of Study:
The purpose of the study is evaluate whether the current design approach for
control of cracking is appropriate for service stresses higher than 40 ksi, the
current default service stress in ACI 318.
4-16

4: Research Plan

ATC-115

95% Draft

Details of Study:
A combination analytical and experimental study should be conducted to
evaluate crack widths developed for service stresses level greater than 40 ksi.
It would be appropriate to consider crack widths developed up to the linear
limit of the reinforcement to provide complete understanding of the range of
behavior in the linear range. The analytical study should consider the
physical model developed by Frosch (1999) which is the basis of the current
318 design approach and evaluate for a number of parameters including
service stress level, concrete cover, and spacing of the reinforcement. A
limited experimental study should be conducted incorporating high strength
reinforcement to evaluate the accuracy of the analytical model. Testing
should consist of simply supported beams subjected to a constant moment
region. As loading is increased, crack widths can be measured for a variety
of stress levels. Primary variables considered should include:

Reinforcement spacing

Reinforcement grade: 60, 100, 120

Specimen type: Beam (16 in. depth) and slab (6 in. depth)

The number of test specimens anticipated is on the order of 10.


Team:
One researcher, one full-time graduate student (two years), and one part-time
laboratory technician to assist.
Timeline:
Approximately 2 years.
4.6

Reinforcing Limits

Studies related to reinforcing limits are summarized in Table 4-4.


Table 4-4
Reference
Section

Summary of Studies Related to Reinforcing Limits


Research or Engineering Study

Estimated
Timeline

4.6.1

Minimum Reinforcement Ratio for Beams

2 months

4.6.2

Minimum Reinforcement Ratio for Slabs and Footings

2 months

4.6.1

Minimum Reinforcement Ratio for Beams

Provisions for minimum flexural reinforcement for beams are currently being
reworked by ACI 318. If ACI Committee 318 decides that a direct solution

ATC-115

4: Research Plan

4-17

95% Draft

such as the existing approach in Section 9.6.1.2 of ACI 318-14 or a modified


version of it should remain in the code this analytical study is required. If
instead, the final change resembles current requirements for prestressed
members in which the cracked section flexural capacity is 120% of the
uncracked flexural capacity, (Sections 7.6.2.1, 8.6.2.2, 9.6.2.1 of ACI 31814) this study is not required.
Purpose of Study:
To determine flexural reinforcement required to provide a minimum
acceptable ratio of cracked section flexural strength to uncracked flexural
strength for beams.
Details of Study:
Flexural section analyses will be conducted to identify As,min, corresponding
to 1.2*Mn,cracked/Mn,uncracked. These trial designs will consider a range of beam
depth/width, rectangular and T-sections, reinforcement depth / beam depth,
concrete strength, and flexural reinforcement ranging in yield strength from
80 to 120 ksi.
Team:
One lead engineering practitioner and one part-time staff engineer to assist in
trial designs and flexural section analyses.
Timeline:
Approximately 2 months.
4.6.2

Minimum Reinforcement Ratio for Slabs and Footings

Current ACI provisions place no limit on fy in the calculation of minimum


reinforcement for slabs and footings. A limit needs to be identified.
If ACI 318 decides to implement a strength-based lower bound on flexural
reinforcement in slabs and footings, a set of figures similar to those discussed
in Section 4.6.1 would be useful and would follow a nearly identical
analytical study.
Team:
One researcher and one part-time graduate student.
Timeline:
Approximately 2 months.

4-18

4: Research Plan

ATC-115

95% Draft
4.7

Detailing of Members

Studies related to detailing of members are summarized in Table 4-5.


Table 4-5

Summary of Studies Related to Detailing of Members

Reference
Section

Research or Engineering Study

Estimated
Timeline

4.7.1

Development and Splice Lengths

2.5 years

4.7.2

Hooked Bar Development Length

3 years

4.7.3

Headed Bar Development Length

2 years

4.7.1

Development and Splice Lengths

Background:
Major areas in which there is a lack of information on the development and
splice strength of high-strength reinforcing bars involve bars cast in concrete
with strengths above 10,000 psi bars, bars with clear spacing of less than two
bar diameters, and bars with epoxy coating. All three of these subject areas
can be efficiently evaluated in a single study.
Purpose of study:
The purpose of the study is to perform tests to determine the performance of
high-strength reinforcement considering the parameters of concrete
compressive strength, bar spacing, and the effect of epoxy coating on splice
and development capacity, and to use the results to develop code provisions.
Details of Study:
Splice tests will be run at bar stresses of 80 to 140 ksi. Three series of tests
will be performed to determine, respectively, (1) the effect of concrete
compressive strengths ranging from 8,000 to 15,000 psi, (2) the effect of bar
spacing as low as one bar diameter, and (3) the effect of epoxy coatings.
Tests will involve splices both with and without confining transverse
reinforcement. The study will emphasize the effect of low bar spacing. Even
a few tests with clear spacing between splices of only one db are desirable to
demonstrate that the reduction in stress developed.
Team:
One researcher, two full-time graduate students, and one part-time laboratory
technician to assist.
Timeline:

ATC-115

4: Research Plan

4-19

95% Draft

Approximately 2.5 years.


4.7.2

Hooked Bar Development Length

Background:
A three year study to determine the development length of high-strength
hooked-bars is now in its final six months at the University of Kansas. This
study addresses bar size, concrete strength, cover, bend radius, and confining
reinforcement. Initial tests, not originally proposed for the study have also
involved the use of multiple hooks with close spacing. The use of staggered
hooks will not be addressed in the current study. The performance of closely
spaced and staggered hooks is the area of principal interest in the proposed
study.
There is a possibility that reinforcement producers will request that the
standard hook bend diameters be increased for fabrication purposes,
particularly for larger bars. Testing of hooked bars with larger bend
diameters is included in this study. If standard hook diameters are not
increased, the scope of this study can be reduced.
Purpose of Study:
The purpose of this study is to perform tests to determine the performance of
conventional and high-strength hooked reinforcement when the hooks are
closely spaced, staggered, or both, as permitted by ACI 318 and to determine
the effect of larger bend diameters on anchorage strength. The results will be
used to modify Code design provisions.
Details of Study:
Tests will involve placement of hooks in simulated beam column joints with
bar stresses between 80 and 140 ksi. Bar sizes will range from No. 5 to No.
11, and concrete compressive strengths will range from 5,000 to 15,000 psi.
The anchorage strength of standard hooks with the new larger diameter bends
will be compared with that of conventional standard hooks. Bar spacing as
low as one bar diameter and hooks staggered behind each other in the same
plane in beam column joints both with and without confining transverse
reinforcement will be evaluated. Limited tests will also be performed in deep
beams with staggered hooks providing anchorage at CCT nodes.
Team:
One researcher, two full-time graduate students, and one part-time laboratory
technician to assist.

4-20

4: Research Plan

ATC-115

95% Draft

Timeline:
Approximately 3 years (two years if hooks with larger bend diameters are not
tested).
4.7.3

Headed Bar Development Length

Background:
A three-year study on the development length of high strength headed bars is
now underway at the University of Kansas. Variables include bar size, bar
spacing, concrete compressive strength, transverse reinforcement, and
member type. The use of staggered headed bars in various applications is not
within the scope of this study.
This study does not include exploration of using larger head sizes. The
exterior joint test program does include the possibility of testing larger head
sizes for applications in which the heads are resisting earthquake induced
forces.
Purpose of this Study:
The purpose of this study is to perform tests to determine the behavior of
closely spaced staggered headed bars such as may be used in beam column
joints with high and low amounts of transverse reinforcement and in
members, such as transfer girders, that require closely spaced reinforcement
that must be anchored within a short distance.
Details of Study:
The anchorage capacity of closely spaced staggered headed bars will be
evaluated at bar stresses between 80 and 140 ksi in beam column joints and
deep beams, in regions both confined and unconfined by transverse
reinforcement. No. 5 through No. 11 bars cast in concrete with compressive
strengths between 5,000 and 15,000 psi will be studied. Heads with bearing
areas ranging from 4 to 14 times the bar area will be evaluated. The results
will be used to develop code provisions.
Team:
One researcher, two full-time graduate students, and one part-time laboratory
technician to assist.
Timeline:
Approximately 2 years.

ATC-115

4: Research Plan

4-21

95% Draft
4.8

General Considerations for Analysis

Studies related to general considerations for analysis are summarized in


Table 4-6.
Table 4-6

Summary of Studies Related to General Considerations for Analysis

Reference
Section

Research or Engineering Study

Estimated
Timeline

4.8.1

Flexural Stiffness

2 years

4.8.2

Effective Stiffness for Column Slenderness

2 years

4.8.3

Moment Redistribution

2 years

4.8.1

Flexural Stiffness

Perform an analytical study to develop appropriate guidance on the selection


of flexural stiffness, EI, for use by engineers when analyzing structures.
Guidance is required for analysis of a structure at the ultimate limit state for
the computation of design forces and at the serviceability limit state for the
estimation of deflections and lateral drifts.
Purpose of Study:
To develop guidance for design engineers on the selection of flexural
stiffness, EI, for use in structural analysis.
Details of Study:
An analytical study should be performed to evaluate the flexural stiffness, EI,
required to reasonably estimate: (1) design forces at the ultimate limit state;
and (2) deflections and drifts at the service limit state. As noted in the
commentary of ACI 318, EI values should not be based totally on the
moment-curvature relationship for the most highly loaded section along the
length of each member. Instead, they should correspond to the moment-end
rotation relationship for a complete member. Structural analysis at the
ultimate limit state should consider the extent of cracking at a section and
along the length of a member so that reasonable relative stiffness for
columns, walls, and beams can be realized. Similarly, the level of cracking
anticipated at service levels should be considered to develop guidance at this
limit state. Overall, a single value of EI for each member type is needed to
allow for simplified analysis of an entire structural system. The study should
evaluate the differences in stiffness required for Grade 100 and Grade 120
reinforcement, as compared to Grade 60 reinforcement, and test results from
experimental studies outlined in this Roadmap should be used to calibrate the
analytical work.

4-22

4: Research Plan

ATC-115

95% Draft

Team:
One researcher, one full-time graduate student (three years), and one parttime engineering practitioner for structural advice, consultation, and design
assistance.
Timeline:
Approximately 2 years.
4.8.2

Effective Stiffness for Column Slenderness

Evaluate whether current values of effective stiffness, EI, used to estimate


column second-order effects are appropriate.
Purpose of Study:
The purpose of this study is to determine whether effective stiffness values
used to design for column slenderness effects require adjustment, and if so,
how to adjust them for members reinforced with high-strength reinforcement.
Details of Study:
An analytical study should be performed to evaluate the effective stiffness,
EI, required to properly estimate second-order effects in slender columns.
The study should use moment-curvature analysis to characterize the column
in combination with second-order analysis of slender-columns. Previous
research has found that a critical column configuration consists of a small
(i.e., 12 in. by 12 in.) column incorporating a reinforcement ratio of 1%.
Primary variables should include the grade of reinforcement, slenderness
ratio of the column, and strength of the concrete. A small experimental
study, on the order of five slender column test specimens, is recommended to
provide confirmation of the analytical results.
Team:
One researcher, one full-time graduate student (two years), and one part-time
engineering practitioner for structural advice, consultation, and design
assistance.
Timeline:
Approximately 2 years.
4.8.3

Moment Redistribution

Perform an analytical study to evaluate the capability of structures reinforced


with high-strength reinforcement to redistribute moments.
ATC-115

4: Research Plan

4-23

95% Draft

Purpose of Study:
The purpose of the study is evaluate the capacity of structures containing
high-strength reinforcement to redistribute moments and to determine
whether changes are required to the current design requirements which
permit a specified maximum redistribution percentage.
Details of Study:
An analytical study should be conducted to determine the amount of
redistribution that can be achieved when high-strength longitudinal
reinforcement is provided. The analytical study can benefit from the results
of the tests outlined in 4.4.1 as the plastic capacity of a beams reinforced
with different reinforcement materials will be developed. The analytical
study should evaluate a number of variables including:

Longitudinal reinforcement ratio: 0.5% to 2.0%

Reinforcement grade: 60, 80, 100, 120

Structural geometry including consideration of beam aspect ratio and


span-to-depth ratios.

Plastic hinge length

Team:
One researcher and one full-time graduate student.
Timeline:
Approximately 2 years.
4.9

Earthquake-Resistant Structures

Studies related to earthquake-resistant structures are summarized in Table


4-7.
4.9.1

Moment Curvature and Rotational Capacity

Perform moment curvature studies of beam, column, and wall elements used
in seismic force-resisting systems, and calibrate with test results.
Purpose of Study:
The purpose of this study is to determine reinforcement strain levels and
rotational capacities for various seismic force-resisting members. Results
will help define requirements for bar elongation and tensile-to-yield strength
ratio.

4-24

4: Research Plan

ATC-115

95% Draft
Table 4-7

Summary of Studies Related to Earthquake-Resistant Structures

Reference
Section

Research or Engineering Study

Estimated
Timeline

4.9.1

Moment Curvature and Rotational Capacity

18 months

4.9.2

Factor for Estimating Expected Flexural Strength

18 months

4.9.3

Cyclically Loaded Beams and Columns Initial Tests and Analytical Studies

2 years

4.9.4.1

Cyclically Loaded Beams

3 years

4.9.4.2

Cyclically Loaded Columns

3 years

4.9.4.3

Cyclically Loaded Interior Joints

3 years

4.9.4.4

Cyclically Loaded Exterior Joints

3 years

4.9.4.5

Two-Way Shear in Slab-Column Intermediate Moment Frames

3 years

4.9.5

Performance of Moment Frames Systems

2 years

4.9.6

Multi-Bay, Multi-Story Frames

3 years

4.9.7

Ordinary Flexure-Critical Walls

3 years

4.9.8

Special and Ordinary Shear-Critical Walls

5 years

4.9.9

Special Flexure-Critical Walls Initial Tests

2 years

4.9.10

Special Flexure-Critical Walls

4 years

4.9.11

Performance of Flexure-Critical Wall Systems

2 years

Details of Study:
This study will be an extension of moment curvature and rotational capacity
studies presented in NIST GCR 14-917-30 to investigate possible variation in
how plasticity spreads for different reinforcement stress-strain curve shapes
and controlling parameters. The shape of the stress-strain curves used in this
study should include low tensile-to-yield strength ratios and different values
for useable strain. Plastic hinge lengths should be varied to account for the
effects of tensile-to-yield strength ratio on yield penetration in the members.
The strain gradient within the plastic hinge length should also be considered.
As a minimum, this study should consider the results of tests described in
Section 4.9.3 to determine plastic hinge length, bar strain, and spread of
plasticity within the plastic hinge region.
The following variations in the shape of the stress-strain curves for Grade 80,
Grade 100, and Grade 120 reinforcement are recommended:

Bilinear curves with ft/fy values of 1.10, 1.15, and 1.20, and uniform
elongation of 5%.

Curves with a yield plateau of 0.5% length, ft/fy values of 1.10, 1.15,
1.20, 1.25, and 1.35, and uniform elongation of 5%, 7% and 8%.

ATC-115

4: Research Plan

4-25

95% Draft

Rounded curves with yield strength based on 2% offset, fu/fy values of


1.15, 1.20, 1.25, 1.35, and 1.50, and uniform elongation of 5%, 7% and
8%.

For consistency with typical U.S. practice, beams and slabs should have
unequal top and bottom areas of reinforcement.
Team:
One researcher and one part-time graduate student.
Timeline:
Approximately 18 months.
4.9.2

Factor for Estimating Expected Flexural Strength

Perform moment-curvature analyses and numerical simulations on typical


moment frame beam sections to evaluate nominal and plastic (expected)
flexural strengths.
Purpose of Study:
The purpose of this study is to establish an appropriate decimal fraction,
currently 1.25 for Grade 60 reinforcement, to estimate plastic moment
capacity from the nominal flexural strength of members reinforced with
high-strength reinforcement. This ratio is used to compute capacity-based
shear force demands on beams and beam-column joints, and flexural and
shear force demands on supporting columns, avoiding the need for momentcurvature calculations for which there is no guidance in ACI 318.
Details of Study:
A two-phase study is recommended to establish the ratio between plastic
flexural capacity and nominal flexural strength of typical beams.
In Phase 1, engineering practitioners will be used to assemble a database of
moment-frame beam details, organized by reinforcement grade, concrete
strength, beam span-to-depth ratio, effective slab width and depth, and
transverse reinforcement details, based on a survey of past projects. Beam
sections in the database will then be redesigned using two grades of highstrength reinforcement, with the goal of achieving the same nominal strength,
while other details remain unchanged. Moment curvature analyses for
positive and negative bending will be performed on all beam sections in the
database to establish the ratio of plastic (maximum) flexural strength to
nominal strength. The effects of confinement and rebar strain hardening

4-26

4: Research Plan

ATC-115

95% Draft

should be considered explicitly, and nominal strength should be based on the


rectangular cross section only.
In Phase 2, rotation demands on beams in special moment frames will be
estimated, and the maximum likely curvatures (for the design basis and
maximum considered earthquake hazard levels) will be back-calculated. The
resulting estimates of curvature, which will vary with span-to-depth ratio,
should be used as the basis for calculating maximum considered strength
(less than plastic flexural strength) and the multiplier on nominal strength
based on a rectangular cross section only.
Team:
Two engineering practitioners, one researcher, and one part-time graduate
student to assist.
Timeline:
Approximately 18 months, assuming 12 months for Phase 1, and an
additional 6 months for Phase 2.
4.9.3

Cyclically Loaded Beams and Columns Initial Tests and


Analytical Studies

Perform cyclic load tests on beam and column test specimens for establishing
required bar mechanical properties. Test the mechanical properties of bars
used in the tests for properties that are important to seismic performance of
bars. Perform nonlinear response history analyses to establish the rotational
demands for the beams and columns just prior to a frame becoming unstable.
Purpose of Experimental and Analytical Program:
The purpose of performing these tests is to determine rotational capacity of
beams and columns reinforced with Grade 100 and possibly Grade120 bar
that has tensile-to-yield ratios and elongations less than the minimum
required for ASTM A706 Grade 80 bar. The purpose of the analytical
program is to determine the rotational demands on beams and columns of
special concrete moment frames reinforced with high-strength reinforcement.
Details of Testing:
The specimens should be loaded cyclically to failure with data recorded
sufficiently so that it can be used for modeling members in FEMA P695
analyses. Bar strains and length of yielding should also be monitored in
sufficient detail to determine equivalent plastic hinge lengths and so that the
distribution of yielding and the strain levels along the bar length in the zone

ATC-115

4: Research Plan

4-27

95% Draft

of yielding can be used with computed moment curvatures described in


Section 4.9.1 to estimate the rotational capacities of beams and columns. Bar
slip into the joint will also need to be monitored. Mechanical properties of
bars and concrete must be well documented. For example, the extensometer
shall not be removed from bar until after it is reached its peak strength, i.e.
beyond the uniform elongation. As a minimum, the following information
regarding the bar properties should be determined: 1) the tensile-to-yield
strength ratio (T/Y ratio), 2) the ultimate and uniform elongations, and 3)
the low-cycle fatigue performance relevant to seismic applications.
The tests should focus on determining the minimum tensile-to-yield strength
ratio of high-strength reinforcement required to obtain sufficient spread of
yielding, the uniform elongation required, and the low cycle fatigue
performance of the bar relative to seismic performance.
Tests of beams should consist of two pairs of tests, one with Grade 100
reinforcement and the second pair with either Grade 100 or Grade 120
reinforcement. The number of column specimens should also be four tests
with similar yield strengths as used for the beams.
Team size:
Two teams (separate research facilities assumed), each consisting of one
researcher, one full-time graduate student, and one part-time laboratory
technician to assist.
Timeline:
Approximately 2 years, assuming 1 year to conduct testing, and 1 year to
process results.
4.9.4

Cyclically Loaded Beams, Columns, and Joints

After establishing the required material properties for high-strength


reinforcement, test series should be performed on beams, columns, and
joints. Both interior and exterior joints should be tested.
4.9.4.1

Beams

Cyclically load test beams using high-strength reinforcement.


Purpose of Study:
The purpose of this study is to establish strength and rotational capacities of
beams of special and intermediate moment frames. Tests should extend to
complete failure so that nonlinear response can be established and used for
analytically modeling beams in response history analyses.
4-28

4: Research Plan

ATC-115

95% Draft

Details of Study:
The tests should account for an appropriate range of ratios of moment-toshear, explore the effect of hoop spacing and layout, account for fy values
ranging from 80 ksi to about 135 ksi, assuming Grade 120 is available, to
account for overstrength and the effect of the shape of the stress-strain curve.
The majority of the tests should be of beams detailed as special moment
frames. Approximately, 20 to 30% of the tests should be of beams detailed
as intermediate moment frames. Some specimens should be constructed
with bar that has the minimum specified tensile-to-yield strength ratio to
confirm that the spread of plasticity is adequate. Other specimens should be
constructed with bar that has the maximum tensile-to-yield strength ratio
likely to be produced so that data are available to establish an appropriate
factor to use to compute the shear demand and to compute the expected beam
strength for strong column-weak beam provisions. The effect of concrete
strength and reinforcement ratio should also be explored. The concrete
strength should be from 5 ksi to 10 ksi. The specimens should include an
appropriate range of shear spans. The specimens should vary the spacing of
the transverse reinforcement within a limited range close to the expected
maximum spacing for special and intermediate moment frame beams. The
size of the specimens should be representative of the size of beams being
used today. A couple of large beam specimens should be tests to explore the
effect if any of beam depth. The specimens should be instrumented so that
bar strains are recorded during the duration of the tests. This information
will be used in part to establish the required useable strain of the bar. The
study should include enough test specimens to cover the variations described
herein. The number of tests is likely to be on the order of 16 beam
specimens.
Team size:
Two teams (separate research facilities assumed), each consisting of one
researcher, two full-time graduate students, and one part-time laboratory
technician to assist.
Timeline:
Approximately 3 years.
4.9.4.2

Columns

Cyclically load test columns using high-strength reinforcement.


Purpose of Study:

ATC-115

4: Research Plan

4-29

95% Draft

The purpose of this study is to establish strength and rotational capacities of


columns of special and intermediate moment frames. Tests should extend to
complete failure so that nonlinear response can be established and used for
analytically modeling columns in response history analyses.
Details of Study:
The tests should account for an appropriate range of ratios of moment-toshear, explore the effect of hoop spacing and layout, account for fy values
ranging from 80 ksi to about 135 ksi, assuming Grade 120 reinforcement is
available, to account for overstrength, axial load, the effect of the shape of
the stress-strain curve, and the concrete strength. The concrete strength
should for specimens should vary from 5 to 12 ksi. The results of FEMA P695 studies on special moment frames should be used to establish the range
of compressive stress level and may be appropriate for establishing some
member sizes; however, exact sizes may not be appropriate due to changes in
required joint depth and the possibility of using smaller beam sizes. Some
tests should include the minimum allowed post-yield stiffness as these will
have the least reserve strength relative to the beams. The study is likely to
test about 16 column specimens. Approximately 20% to 30% of the tests
should be columns of intermediate moment frames.
Team size:
Two teams (separate research facilities assumed), each consisting of one
researcher, two full-time graduate students, one part-time laboratory
technician to assist, and one engineering practitioner to provide structural
advice and consultation.
Timeline:
Approximately 3 years.
4.9.4.3

Interior Joints

Test beam-column sub-assemblages of interior joints of special moment


frames. These are referred to as Type 2 joints in accordance with ACI
352-02.
Purpose of Study:
The purpose of this study is to: (1) establish the depth of joint required; and
(2) confirm that the performance of interior beam-column joints is consistent
with that of joints of moment frames with Grade 60 bars. Tests should be
primarily of joints of special moment frames. A limited number of
intermediate moment frame joints should be tested.

4-30

4: Research Plan

ATC-115

95% Draft

Details of Study:
The specimens should be cross-shaped with beam and column sizes
consistent with those used for beam and column tests in the studies described
in Sections 4.9.4.1 and 4.9.4.2. Prior to testing, the joint depth
recommendations of NIST GCR 14-917-30 should be studied along with the
recommendations established by Japans New RC Project (Aoyama, 2001).
The depths should be selected to explore whether the required depths from
these sources are consistent with the performance of joints with Grade 60
bars passing through them. Specimens should include a range of concrete
strengths and a range of axial loads applied to the columns as these two
factors are likely to affect bar slip and degradation due to bar slip through the
joint. Concrete strength, fc, variation of 5,000, 8,000 and 12,000 psi is
suggested. The axial load ratios, P/ fc Ag, applied to the column should be
0.2, 0.4 and the maximum ratio at large drifts from FEMA P-695 analyses.
Bars should be Grade 80, 100, and 120 (if available). Joint depths should
vary with the grade of bar. For Grade 80 bar, the joint depth should be 26db;
for Grade 100 bar, joint depths should be 32db to 36db; and for Grade 120,
the joint depths should be 32db to 40db. The number of tests should be
sufficient to establish a joint depth criterion applicable to multiple grades of
reinforcement, concrete strengths, and column axial load. A couple of
intermediate moment frame joints should be tested. The number of tests is
likely to be on the order of twelve.
Team size:
Two teams (separate research facilities assumed), one team consisting of one
researcher, two full-time graduate students, and one part-time laboratory
technician to assist, and the other team consisting of one researcher, one fulltime graduate student, and one part-time laboratory technician to assist.
Timeline:
Approximately 3 years, assuming 2 years to conduct testing, and 1 year to
process results.
4.9.4.4

Exterior Joints

Test beam-column sub-assemblages of exterior joints of special and


intermediate moment frames.
Purpose of Study:
The purpose of this study is to establish that joint depths based on
development length are adequate for hooked and headed bars terminating at

ATC-115

4: Research Plan

4-31

95% Draft

the joint, to confirm that the amount of joint reinforcement does not need to
be increased to prevent joint splitting, and to establish whether the 1.25
factor on longitudinal bar yield strength used to compute the joint shear is
adequate for high-strength reinforcement.
Details of Study:
The exterior joint specimens should be T-shaped with the beam and the stem.
The beams, columns, and joints should cover a sufficient range of sizes to be
representative of frames used in practice, and to provide results that can be
used for FEMA P-695 nonlinear response history analyses. Required
development lengths should be computed consistent with results of
development length for high-strength bar. Bars should have tensile-to-yield
strength ratios similar to the maximum expected for high-strength bar to
increase the likelihood of bar slip. Some specimens should be constructed
with longitudinal beam bars that terminate with 90 degree hooks and others
with heads. Tests should include: a) bars terminated with both Class HA and
Class HB bar-to-head connections to explore whether the stronger bar-tohead connections are necessary; and b) head areas between 4Ab and 9Ab. The
number and spacing of beam and column bars should be consistent with that
which is likely to be used in actual structures, e.g., 3-inch center-to-center
spacing. Joints depths should be determined for at least two different bar
sizes, e.g. No. 6 and No. 9 bars. Bars should be Grade 100 and Grade 120 (if
available). Concrete strength, fc, should vary, and values of 6,000 and
12,000 psi are suggested. Selected specimens should include bars with the
greatest actual tensile-to-yield strength ratio (fu/fy) expected to be produced
for use in moment frames. These specimens should be used to specifically
study the applicability of the 1.25 factor applied to the yield strength of
longitudinal bars to compute joint shear. These specimens can also be used
to determine whether splitting of the joint might be a problem for these joints
with high-strength longitudinal bars and whether the required transverse
reinforcement should be increased. The test program should include enough
tests to confirm that adequate performance of these joints will be achieved.
This is likely to require approximately twelve tests.
Team size:
Two teams (separate research facilities assumed), one team consisting of one
researcher, two full-time graduate students, and one part-time laboratory
technician to assist, and the other team consisting of one researcher, one fulltime graduate student, and one part-time laboratory technician to assist.
Timeline:

4-32

4: Research Plan

ATC-115

95% Draft

Approximately 3 years, assuming 2 years to conduct testing, and 1 year to


process results.
4.9.4.5

Two-Way Shear in Slab-Column Intermediate


Moment Frames

Perform cyclic load tests of slab-column frame joints.


Purpose of Test Program:
The purpose of this test program is to determine whether the 0.4Vc limit on
the two-way shear from gravity loads is still appropriate or, if not, what it
should be changed to.
Details of Study:
The specimens should be similar to the four tested by Pan and Moehle
(1989), which were used, along with other tests reported in the reference, to
establish the 0.4Vc limit originally. The longitudinal bar should be Grade
120, assuming its use is appropriate for two-way slabs, as this will minimize
the reinforcement ratio and maximize the strain in the bars. The momentshear ratio at the joint should be consistent with that expected in an actual
structure.
Team:
One researcher, two full-time graduate students, and one part-time laboratory
technician to assist.
Timeline:
Approximately 3 years.
4.9.5

Performance of Moment Frame Systems

Identify the collapse resistance of moment frame systems constructed with


high strength reinforcement.
Purpose of Study:
The purpose of this study is to determine whether moment frame systems
constructed with high-strength reinforcement will provide the intended
collapse resistance described in the commentary to ASCE/SEI 7, namely,
less than 10% probability of collapse due to risk targeted maximum
considered earthquake (MCE) ground shaking. This study is needed to
demonstrate that the use of high-strength reinforcement is acceptable for use
in seismic-force-resisting systems that have traditionally been designed and
constructed using Grade 60 reinforcement that is ductile, such as ASTM
ATC-115

4: Research Plan

4-33

95% Draft

A706 reinforcement. This study will answer questions regarding the


appropriateness of using high-strength reinforcement that ASCE 7 Seismic
Code Committee and the Provisions Update Committee of the National
Earthquake Hazard Reduction Program (NEHRP) are likely to raise when
referencing the first edition of ACI 318 that allows the use of high-strength
reinforcement for special seismic systems. One such question that should
specifically be addressed is whether the strong column-weak beam
provisions require adjustment with the use of high-strength reinforcement.
Details of Study:
Nonlinear models for beams, columns and joints, developed previously for
Grade 60 reinforcement, will be extended to address high strength
reinforcement based on the component tests of Section 4.9.4. The models
will be verified and validated.
FEMA P-695 analyses will be performed for prototype buildings reinforced
with high-strength bars. Buildings will be sited in regions of moderate and
high seismic hazard. Nonlinear models for components constructed with high
strength reinforcement will be used for the analysis. Building archetypes
from existing studies (FEMA, 2009; NIST, 2010; and NIST, 2012), including
low-rise, medium-rise and high-rise frames, will be redesigned with Grade
80, Grade 100, and Grade 120 reinforcement, with the goal of optimizing the
cross sections per standard design practice. The study will determine whether
acceptable collapse resistance is achieved for prototype buildings with
special and intermediate moment frame systems. Beam-column and slabcolumn frames will be considered for the intermediate moment frame
buildings.
If the prototype buildings equipped with high strength reinforcement do not
provide acceptable resistance to collapse, the study will identify those
measures needed to improve resistance, revise the models and rerun the
analyses. The deliverable would include design and detailing rules for special
and intermediate moment frame systems constructed with high strength
reinforcement that would a) provide acceptable collapse resistance following
FEMA P695-type calculations, and b) be suitable for inclusion in the next
edition of ACI 318.
Team:
One researcher, one full-time graduate student, and two part-time
engineering practitioners to provide advice, consultation, and design
assistance.

4-34

4: Research Plan

ATC-115

95% Draft

Timeline:
Approximately 2 years.
4.9.6

Multi-Bay, Multi-Story Frames

Cyclically load test two-bay, two- or three-story special moment frames


using Grade 100 reinforcement.
Purpose of Study:
The purpose of these tests is to confirm expected global performance of
provisions for special moment frames. The effectiveness of specific
provisions for strong column-weak beam behavior, probable moment
strength, shear demand, joint depth, and joint shear strength should be
monitored.
Details of Study:
The frame specimens should be relatively large scale and representative of
possible frame configurations. The specimens should be two-dimensional
but with partial width slabs to capture the increase in strength from the slab.
Reinforcement should be Grade 100 and concrete compressive strengths
should be 6,000 and 12,000. The specimens should be designed without
being conservative. Two frames are expected to be tested.
Team:
One researcher, two full-time graduate students, one part-time laboratory
technician, and one part-time engineering practitioner to provide advice,
consultation, and design assistance.
Timeline:
Approximately 3 years.
4.9.7

Ordinary Flexure-Critical Walls

Perform cyclic tests of ordinary flexure-critical shear walls to evaluate


performance of high strength reinforcement and applicability to current code
provisions for shear, P-M interaction, and possibly confinement.
Purpose of Test Program:
To evaluate the performance of high strength reinforcement in flexurecritical shear walls subject to minor/moderate ductility demands. Current
ACI provisions limit shear reinforcement to Grade 60 and P-M reinforcement
to Grade 80. Current provisions could be validated for high strength
ATC-115

4: Research Plan

4-35

95% Draft

reinforcement, or reworked with additional variables to accommodate


differing behavior.
Details of Study:
Shear. Perform cyclic lateral loading for ordinary shear walls with aspect
ratio of 3:1 to 4:1. The primary variable of the study is horizontal
reinforcement quantity (equating to the range of Vs = 2 to 8). Reinforcement
yield strength should be Grades 100 and 120. Vertical reinforcement will be
sized to ensure shear is the primary mechanism. The goal is to explore the
validity of Vs calculations in sections 11.9.5, 11.9.6, and 11.9.9 (318-11) for
high strength reinforcement.
P-M Interaction. Perform cyclic lateral loading for ordinary shear walls with
aspect ratio in the range of 3:1 to 4:1. The primary variable of the study is
vertical reinforcement ratio (0.0025 to 0.03). Reinforcement yield strength
should be in the range of 100 ksi to 120 ksi. An additional variable could be
axial load ratio (P/( fc Ag) in the range of 0.05 to 0.30. Horizontal
reinforcement will be sized to ensure P-M interaction is the primary
mechanism. The goal is to verify the validity of standard Pn-Mn equations
for HSR. Additionally, grade 60 specimens will serve as the benchmarks for
evaluating P-M ductility and hysteretic energy absorption, to which the high
strength reinforcement specimens will be compared. Ductility should also be
compared to that intended by ACI 318 and ASCE 7 for ordinary walls.
Confinement. Current provisions require ties for portions of walls with
vertical reinforcement greater than 0.01 to prevent bar buckling. This trigger
needs to be validated with physical testing using vertical bars stronger than
the current limit of 80 ksi, which potentially are more critical for bar
buckling. This could also be conducted as a column-only test.
Approximately 10 specimens have been assumed, distributed between shear,
P-M interaction, and confinement studies to optimize coverage of the
recommended parameters.
Team:
Two teams (separate research facilities assumed), each consisting of one
researcher, one full-time graduate student, one part-time laboratory
technician to assist, and one engineering practitioner to provide structural
advice, consultation, and design assistance.
Timeline:
Approximately 3 years.

4-36

4: Research Plan

ATC-115

95% Draft
4.9.8

Special and Ordinary Shear-Critical Walls

Test a number of low aspect ratio, shear-critical, walls, with and without
boundary columns and flanges.
Purpose of Study:
The purpose of performing these tests is to establish whether the prescriptive
details in ACI 318 for ordinary and special shear-critical walls will be
acceptable if high strength reinforcement is used.
Details of Study:
A two-phase study consisting of reverse-cyclic, pseudo-static tests on walls
with aspect ratios of 0.33 and 0.67, concrete strengths of 3500 psi and
6000 psi, vertical and web reinforcement ratios of 0.25%, and Grade 60
(benchmark) and high-strength reinforcement. Testing will be preceded by a
review of the literature and analysis of tests of prior large-scale walls (e.g.,
testing conducted under the NEESR program). Phase 1 tests will examine
walls with rectangular cross sections. Phase 2 tests will examine walls with
boundary columns and flanges. Work will include numerical simulation
studies using OpenSees (OpenSees, 2011), LS-DYNA (LSTC, 2013), or
DIANA (TNO, 2013).
A goal of the testing program will be to determine whether seismic
performance is poorer, measured in terms of hysteretic behavior, if Grade 60
reinforcement is replaced by high strength reinforcement. Determining what
is acceptable hysteretic behavior for a shear-critical wall is not part of the
scope of this study. Walls will be designed to have the same nominal
shearing strength with the Grade 60 and high strength reinforcement, where
rebar area and rebar spacing will be varied.
Approximately 16 wall specimens will be tested in Phase 1 to examine
ordinary and special walls. If appreciable differences are seen in the
behavior of nominally identical (shearing strength) Phase 1 walls with Grade
60 and high strength reinforcement, six to eight specimens with boundary
columns and flanges should be tested in Phase 2.
Team:
Phase 1 Two teams (separate research facilities assumed), one team
consisting of one researcher and three full-time graduate students (three
years) to conduct testing and numerical studies, and the other team consisting
of one researcher and two full-time graduate students (three years) to conduct

ATC-115

4: Research Plan

4-37

95% Draft

testing, along with one part-time engineering practitioner to provide


structural advice, consultation, and design assistance on each team.
Phase 2 One team (at a third research facility), consisting of one researcher,
two full-time graduate students (three years) to conduct testing and numerical
studies, and one part-time engineering practitioner to provide structural
advice, consultation, and design assistance.
Timeline:
Approximately 5 years, assuming 3 years to complete Phase 1 tests and 3
years to complete Phase 2 tests, in a staggered schedule.
4.9.9

Special Flexure-Critical Walls Initial Tests

Purpose of Study:
To perform tests of flexure-critical concrete shear walls constructed with
high-strength reinforcement to determine the minimum uniform elongation
and tensile-to-yield strength ratio required in high-strength steel bars for their
safe use in earthquake-resistant construction. The test program should
include large-scale flanged walls subjected to inelastic displacement
reversals.
Details of Study:
Perform cyclic tests to investigate the strength, stiffness, deformation
capacity, and hysteretic response of concrete walls reinforced longitudinally
and transversely with high-strength steel bars. Variables in this study should
include the specified yield strength of the wall reinforcement, Grade 60
(control specimen) and Grade 100; and the tensile-to-yield strength ratio of
the wall reinforcement, varying from 1.1 to 1.5. Design of the specimens
should comply with ACI 318-14 requirements for special structural walls and
the special additional detailing requirements identified in NIST GCR 14-91730 (2014). All specimens should have height-to-length ratio not less than 3
and the applied axial load (if any) should be held constant throughout the
test.
The study should address: (1) strength, especially the influence of the shape
of the stress-strain curves of the wall reinforcement on flexural strength and
on the variation of concrete and steel contributions to shear strength as a
function of deformation demands; (2) stiffness, with emphasis on
characterizing the reduction of unloading and reloading lateral stiffness with
increased deformation; (3) deformation capacity, including the effects of
reinforcement strains on strength reduction; and (4) hysteretic response,

4-38

4: Research Plan

ATC-115

95% Draft

identifying key parameters that influence the measured cyclic response and
the variation in energy dissipation in successive cycles. It is envisioned that
four wall specimens will be part of this initial test program.
Team:
One researcher, two full-time graduate students, and one part-time laboratory
technician to assist.
Timeline:
Approximately 2 years, assuming 1 year to conduct testing, and 1 year to
process results.
4.9.10 Special Flexure-Critical Walls

Perform tests of flexure-critical concrete shear walls reinforced with highstrength reinforcement that meet newly developed ASTM specifications for
Grade 100 steel bars. The study will evaluate the applicability of current ACI
318-14 provisions to walls reinforced with high-strength steel bars subjected
to combined shear, flexural, and axial loads. Wall specimens should include
confinement details at boundary elements consistent with findings from the
initial column tests of Section 4.9.3. Testing of reinforced concrete coupling
beams is included here, as these elements are used in conjunction with
special structural walls to create coupled structural wall systems. Design of
the specimens should be in compliance with ACI 318 Building Code
requirements for special structural walls and the additional detailing
requirements identified in NIST GCR 14-917-30.
Purpose of Test Program:
To evaluate the performance of high-strength reinforcement in flexurecritical shear walls subjected to inelastic displacement reversals. Current
ACI 318-14 provisions limit shear, flexural, and axial reinforcement to
Grade 60. Code provisions could be extended for high strength
reinforcement, or reworked with additional variables to accommodate
differing behavior.
Details of Study:
Shear. Perform cyclic tests of special shear walls with height-to-length ratio
not less than 3. The primary variable of the study is the amount of flexural
reinforcement to induce shear stresses in the range of 4 to 6 f c' . Additional
variables could be concrete strength and perhaps wall flanges. The study
should explore the validity of ACI 318-11 Equation 21-7 when using highstrength reinforcement.
ATC-115

4: Research Plan

4-39

95% Draft

P-M Interaction. Perform cyclic lateral loading for special shear walls with
height-to-length ratios not less than 3. The primary variable of the study is
vertical reinforcement ratio, ranging from 0.0025 to 0.03. Axial load ratio,
P/(f'c Ag), ranging from 0.05 to 0.30 should also be considered. The goal is to
verify the validity of standard Pn Mn equations for high-strength
reinforcement. Grade 60 specimens will serve as the benchmark for
evaluating deformation capacity and hysteretic response, to which the high
strength reinforcement specimens will be compared. Response should also
be compared to that intended by ACI 318-14 and ASCE 7-10 for special
walls.
Confinement. As in previous ACI 318 provisions, confinement in special
boundary elements will be considered similar to those required in columns of
special moment frames, avoiding the need for supplementary testing.
Outside special boundary elements, current provisions require ties for
portions of walls with vertical reinforcement ratios greater than 400/fy to
prevent bar buckling. This trigger needs to be validated with physical testing
using vertical bars stronger than the current limit of Grade 60. This could
also be conducted as a column-only test.
Coupling Beams. Perform cyclic testing of coupling beams (with and
without diagonal reinforcement) subjected to large displacement reversals.
The study should include shear stresses up to 8.5 f c' for diagonallyreinforced beams and 6 f c' for coupling beams designed according to the
provisions that apply to special moment frames. Span-to-depth ratios
between 1 and 3 should be considered. The goal is to explore the validity of
the current Vn equation using high-strength reinforcement. Also to be studied
is the potential need for a greater quantity of confinement for coupling beams
reinforced with high-strength steel bars.
Approximately ten wall specimens (six coupling beam tests and two coupled
wall tests) have been assumed. Specimens should be distributed across
Grade 80, Grade 100 and Grade 120 reinforcement, optimizing coverage of
the remaining parameters.
Team:
Three teams (separate research facilities assumed), two teams consisting of
one researcher, one full-time graduate student (three years), and one parttime laboratory technician to assist, and one team consisting of one
researcher, two full-time graduate students (three years), and one part-time
laboratory technician to assist, along with one part-time engineering

4-40

4: Research Plan

ATC-115

95% Draft

practitioner to provide structural advice, consultation, and design assistance


on each team.
Timeline:
Approximately 4 years, assuming 3 years for each team to complete their
assigned tests in parallel, but with a staggered schedule.
4.9.11 Performance of Wall Systems

Quantify the seismic performance of structural wall systems constructed


using high strength reinforcement.
Purpose of Study:
The purpose of this study is to determine whether flexure-critical and shearcritical structural wall systems constructed with high-strength reinforcement
will provide the intended resistance to collapse consistent with the seismic
performance factors given in ASCE 7-10, namely, to have a probability of
collapse of 10% or less when subjected to risk targeted maximum
considered earthquake (MCE) ground shaking. This study is needed to
demonstrate that high-strength reinforcement is acceptable for use in seismicforce-resisting systems that have traditionally been designed and constructed
using ductile Grade 60 reinforcement. This study will answer questions
regarding the appropriateness of using high-strength reinforcement that
ASCE 7 Seismic Code Committee and the Provisions Update Committee of
the National Earthquake Hazard Reduction Program (NEHRP) are likely to
raise when referencing the first edition of ACI 318 that allows the use of
high-strength reinforcement for special seismic-force-resisting systems.
Details of Study:
Nonlinear models for structural walls, developed previously for Grade 60
reinforcement, will be extended to address high strength reinforcement. The
model parameters will be calibrated to results from component and system
tests outlined in Sections 4.9.8, 4.9.9, and 4.9.10.
FEMA P695-type analyses will be performed on prototype buildings with
structural walls designed using high-strength reinforcement. Buildings will
be sited in regions of moderate and high seismic hazard. Nonlinear models
for components constructed with high strength reinforcement will be used for
the analysis. Building archetypes from existing studies (FEMA, 2009; NIST,
2010; and NIST, 2012), including medium and high-rise walls, will be
redesigned with Grade 80, Grade 100, and Grade 120 reinforcement. The

ATC-115

4: Research Plan

4-41

95% Draft

study will determine whether acceptable collapse resistance is achieved for


the prototype buildings.
If the prototype buildings equipped with high strength reinforcement do not
provide acceptable resistance to collapse, the study will identify those
measures needed to improve resistance. Those measures, and likely others,
would form the basis of a Phase II study involving the development of design
and detailing rules for structural walls constructed with high strength
reinforcement that would: (1) provide acceptable resistance to collapse
following FEMA P695-type evaluations; and (2) be suitable for inclusion in
the next edition of ACI 318.
Team:
One researcher, one full-time graduate student, and one part-time engineering
practitioner to provide advice, consultation, and design assistance.
Timeline:
Approximately 2 years.
4.10

Engineering Design Studies

Studies related to trial engineering designs are summarized in Table 4-8.


Table 4-8
Reference
Section

Summary of Studies Related to Trial Engineering Designs


Research or Engineering Study

Estimated
Timeline

4.10.1

Trial Engineering Designs for Use of Grade 80 Reinforcement in Special


Seismic Systems

1 year

4.10.2

Trial Engineering Designs for Use of Grade 100 Reinforcement in General


Applications

1 year

4.10.3

Trial Engineering Designs for Use of Grade 100 Reinforcement in Special


Seismic Systems

1 year

4.10.1 Trial Engineering Designs for Use of Grade 80


Reinforcement in Special Seismic Systems

Perform trial designs of example structures that incorporate and fully utilize
Grade 80 high-strength reinforcement in special seismic systems.
Purpose of Study:
The purpose of this study is to: (1) evaluate the critical/controlling design
issues that are encountered with the use of high-strength reinforcement; (2)
provide comparison of the resulting structural design relative to Grade 60

4-42

4: Research Plan

ATC-115

95% Draft

reinforcement; and (3) evaluate the effects of changes in design requirements


on the resulting structural design.
Details of Study:
Trial designs to evaluate and illustrate the influence of incorporating highstrength reinforcement into special seismic systems are intended to
understand how the combined effects of new strength, serviceability, and
detailing requirements control the resulting reinforced concrete section
designs. It is anticipated that such information will be of interest to ACI
Committee 318 in the adoption of revised code requirements.
Two example structures in the five- to ten-story range are recommended for
study. As a baseline for comparison, each structure should be designed using
Grade 60 reinforcement and the requirements of the current version of
ACI 318 (ACI 318-14). To incrementally study the effects of high-strength
reinforcement and the interaction of new design provisions, each structure
should be redesigned as follows:

Using Grade 80 reinforcement and the requirements of the current


version of ACI 318 (ACI 318-14), ignoring all limits restricting the
strength of reinforcement (fy and fyt).

Using Grade 80 reinforcement and any new design requirements for


high-strength reinforcement developed based on research outlined in this
Roadmap.

The trial designs should investigate the following:

Design requirements that control the resulting design.

Code limits, if any, that control the resulting design.

Differences in controlling factors that are keyed to the use of highstrength reinforcement.

Differences in member size and reinforcing requirements (e.g., crosssectional area and details of reinforcement).

Team:
Two teams (separate engineering firms assumed), each consisting of one lead
engineering practitioner and one part-time staff engineer to assist, and one
researcher to provide advice and consultation on results and interpretation of
recent and ongoing tests.

ATC-115

4: Research Plan

4-43

95% Draft

Timeline:
Approximately 1 year.
4.10.2 Trial Engineering Designs for Use of Grade 100
Reinforcement in General Applications

Perform trial designs of example structures that incorporate and fully utilize
Grade 100 (or higher) high-strength reinforcement in gravity, wind, and
ordinary seismic systems.
Purpose of Study:
The purpose of this study is to: (1) evaluate the critical/controlling design
issues that are encountered with the use of high-strength reinforcement; (2)
provide comparison of the resulting structural design relative to Grade 60 and
Grade 80 reinforcement; and (3) evaluate the effects of changes in design
requirements on the resulting structural design.
Details of Study:
Trial designs to evaluate and illustrate the influence of incorporating highstrength reinforcement into gravity, wind, and ordinary seismic systems are
intended to understand how the combined effects of any additional strength,
serviceability, and detailing requirements for Grade 100 (or higher)
reinforcement control the resulting reinforced concrete sections. It is
anticipated that such information will be of interest to ACI Committee 318 in
the adoption of revised code requirements.
Two example structures in the five- to ten-story range are recommended for
study. Structures designed under Section 4.10.1 using Grade 60
reinforcement and the requirements of the current version of ACI 318 (ACI
318-14) can be used as a baseline for comparison. To incrementally study
the effects of Grade 100 (or higher) high-strength reinforcement and the
interaction of new design provisions, each structure should be redesigned as
follows:

Using Grade 100 (or higher) reinforcement and the requirements of the
current version of ACI 318 (ACI 318-14), ignoring all limits restricting
the strength of reinforcement (fy and fyt).

Using Grade 100 (or higher) reinforcement and any new design
requirements for high-strength reinforcement developed based on
research outlined in this Roadmap.

The trial designs should investigate the following:

4-44

4: Research Plan

ATC-115

95% Draft

Design requirements that control the resulting design.

Code limits, if any, that control the resulting design.

Differences in controlling factors that are keyed to the use of highstrength reinforcement.

Differences in member size and reinforcing requirements (e.g., crosssectional area and details of reinforcement).

Team:
Two teams (separate engineering firms assumed), each consisting of one lead
engineering practitioner and one part-time staff engineer to assist, and one
researcher to provide advice and consultation on results and interpretation of
recent and ongoing tests.
Timeline:
Approximately 1 year.
4.10.3 Trial Engineering Designs for Use of Grade 100
Reinforcement in Special Seismic Systems

Perform trial designs of example structures that incorporate and fully utilize
Grade 100 (or higher) high-strength reinforcement in special seismic
systems.
Purpose of Study:
The purpose of this study is to: (1) evaluate the critical/controlling design
issues that are encountered with the use of high-strength reinforcement; (2)
provide comparison of the resulting structural design relative to Grade 60 and
Grade 80 reinforcement; and (3) evaluate the effects of changes in design
requirements on the resulting structural design.
Details of Study:
Trial designs to evaluate and illustrate the influence of incorporating highstrength reinforcement into special seismic systems are intended to
understand how the combined effects of any additional strength,
serviceability, and detailing requirements for Grade 100 (or higher)
reinforcement control the resulting reinforced concrete sections. It is
anticipated that such information will be of interest to ACI Committee 318 in
the adoption of revised code requirements.
Two example structures in the five- to ten-story range are recommended for
study. Structures designed under Section 4.10.1 using Grade 60

ATC-115

4: Research Plan

4-45

95% Draft

reinforcement and the requirements of the current version of ACI 318 (ACI
318-14) can be used as a baseline for comparison. To incrementally study
the effects of Grade 100 (or higher) high-strength reinforcement and the
interaction of new design provisions, each structure should be redesigned as
follows:

Using Grade 100 (or higher) reinforcement and the requirements of the
current version of ACI 318 (ACI 318-14), ignoring all limits restricting
the strength of reinforcement (fy and fyt).

Using Grade 100 (or higher) reinforcement and any new design
requirements for high-strength reinforcement developed based on
research outlined in this Roadmap.

The trial designs should investigate the following:

Design requirements that control the resulting design.

Code limits, if any, that control the resulting design.

Differences in controlling factors that are keyed to the use of highstrength reinforcement.

Differences in member size and reinforcing requirements (e.g., crosssectional area and details of reinforcement).

Team:
Two teams (separate engineering firms assumed), each consisting of one lead
engineering practitioner and one part-time staff engineer to assist, and one
researcher to provide advice and consultation on results and interpretation of
recent and ongoing tests.
Timeline:
Approximately 1 year.

4-46

4: Research Plan

ATC-115

95% Draft

Chapter 5

Program Recommendations

This chapter summarizes recommended research and engineering studies,


provides estimated budget requirements, identifies possible priorities, and
provides recommendations for the implementation of a program to encourage
adoption of high-strength reinforcement into future editions of ACI 318.
5.1

Summary of Program

In this Roadmap, a series of research and engineering studies are proposed to


address key production, fabrication and engineering issues related to the use
of high-strength reinforcement in reinforced concrete construction. The
proposed research and engineering studies are intended to: (1) further
investigate the use of high-strength reinforcement; and (2) support the
development of code change proposals for ACI 318 to allow the general use
of reinforcement in excess of Grade 60. The list of research and engineering
studies is summarized in Table 5-1. Detailed information on the purpose of
each study, the scope of recommended testing, the anticipated team, and
estimated timeline is provided in Chapter 4.
The overall program is necessarily comprehensive. Together, the proposed
research and engineering studies cover the major areas of production and
fabrication, strength of members, serviceability, reinforcing limits, detailing,
analysis, and earthquake resistance, and affect nearly every chapter of
ACI 318.
5.2

Estimated Budget Requirements

The total estimated budget required to implement the program outlined in


this Roadmap is summarized in Table 5-1. In order to comprehensively
investigate all parameters of interest, the number of required test specimens
quickly rises to levels that are impractical to implement. As a result, the
number of test specimens, and the number of parameters investigated, has
been intentionally limited in each study. The resulting recommendations are
a compromise between the need to cover variation in important controlling
parameters, and what can reasonably be accomplished in the environment of
a traditional university research program.

ATC-115

5: Program Recommendations

5-1

95% Draft
Table 5-1

Summary of Proposed Research and Engineering Studies

Reference
Section
4.3

Estimated Budget

Bar Production and Specification

4.3.1

Mechanical Properties of Recent Heats of High Strength Reinforcement

$100,000

4.3.2

Detailed Mechanical Property Tests of Grade 100 and Grade 120


Reinforcement

$439,000

4.4

Strength of Members

4.4.1

Flexural Strength and Tensile Strain Limits

$149,000

4.4.2

Deflection Warning for Flexural Members Subjected to Gravity Loads

$518,000

4.4.3

Column Strength

4.4.4

Tension Regions of Shells and Folded Plates

$834,000

4.4.5.1

One-Way Shear in Beams without Shear Reinforcement

$692,000

4.4.5.2

Two-Way Shear in Slabs without Shear Reinforcement

$716,000

4.4.6

Shear Strength of Beams with Shear Reinforcement

$716,000

4.4.7

Shear Friction

$384,000

4.4.8

High-Cycle Elastic Fatigue of High-Strength Reinforcing Bars

$550,000

4.5

$26,000

Serviceability

4.5.1

Deflection of Flexural Members

$611,000

4.5.2

Crack Control of Flexural Members

$341,000

4.6

Reinforcing Limits

4.6.1

Minimum Reinforcement Ratio for Beams

$43,000

4.6.2

Minimum Reinforcement Ratio for Slabs and Footings

$43,000

4.7

Detailing of Members

4.7.1

Development and Splice Lengths

$738,000

4.7.2

Hooked Bar Development Length

$988,000

4.7.3

Headed Bar Development Length

$623,000

4.8

General Considerations for Analysis

4.8.1

Flexural Stiffness

$209,000

4.8.2

Effective Stiffness for Column Slenderness

$383,000

4.8.3

Moment Redistribution

$209,000

4.9

5-2

Research or Engineering Study

Earthquake-Resistant Structures

4.9.1

Moment Curvature and Rotational Capacity

$151,000

4.9.2

Factor for Estimating Expected Flexural Strength

$197,000

4.9.3

Cyclically Loaded Beams and Columns Initial Tests and Analytical Studies

$749,000

4.9.4.1

Cyclically Loaded Beams

$1,654,000

4.9.4.2

Cyclically Loaded Columns

$1,734,000

4.9.4.3

Cyclically Loaded Interior Joints

$1,541,000

5: Program Recommendations

ATC-115

95% Draft
Table 5-1

Summary of Proposed Research and Engineering Studies (continued)

Reference
Section

Research or Engineering Study

Estimated Budget

4.9.4.4

Cyclically Loaded Exterior Joints

4.9.4.5

Two-Way Shear in Slab-Column Intermediate Moment Frames

$809,000

4.9.5

Performance of Moment Frames Systems

$265,000

4.9.6

Multi-Bay, Multi-Story Frames

$1,040,000

4.9.7

Ordinary Flexure-Critical Walls

$1,226,000

4.9.8

Special and Ordinary Shear-Critical Walls

$3,105,000

4.9.9

Special Flexure-Critical Walls Initial Tests

4.9.10

Special Flexure-Critical Walls

4.9.11

Performance of Flexure-Critical Wall Systems

4.10

$1,429,000

$591,000
$2,109,000
$265,000

Engineering Design Studies

4.10.1

Trial Engineering Designs for Use of Grade 80 Reinforcement in Special


Seismic Systems

$268,000

4.10.2

Trial Engineering Designs for Use of Grade 100 Reinforcement in General


Applications

$268,000

4.10.3

Trial Engineering Designs for Use of Grade 100 Reinforcement in Special


Seismic Systems

$268,000

Total Estimated Program Budget

$26,981,000

Estimated budget requirements by program area (production and fabrication,


strength of members, serviceability, reinforcing limits, detailing, analysis,
and earthquake resistance) are summarized in Table 5-2.
Table 5-2

Estimated Budget Requirements by Program Area

Reference
Section

Study

4.3

Bar Production and Specification

4.4

Strength of Members

4.5

Serviceability

4.6

Reinforcing Limits

4.7

Detailing of Members

4.8

General Considerations for Analysis

4.9

Earthquake-Resistant Structures

4.10

Engineering Design Studies

Estimated Budget
$539,000
$4,585,000
$952,000
$86,000
$2,349,000
$801,000
$16,865,000
$804,000

The effects of high-strength reinforcement on strength, serviceability,


detailing, and member behavior are interrelated, and testing in one area can
impact the results of testing in another area. In preparing budget estimates,

ATC-115

5: Program Recommendations

5-3

95% Draft

however, individual research and engineering studies have been priced


individually. Some consideration has been given to combining studies and
using the results from a series of tests in one area to serve the needs of
another area. It is anticipated that program recommendations will evolve as
implementation progresses, and that additional efficiencies will be realized as
test results become available, lessons are learned, and the state of knowledge
improves on which parameters are most important.
5.2.1

Budget Assumptions

Budgeting assumptions have been based on the traditional university research


program model, which consists of a lead researcher assisted by one or more
graduate students performing the work. In this model, the lead researcher
serves as a part-time mentor and advisor, and the graduate students work
essentially full-time in fabricating and testing the specimens, processing
results, performing necessary analytical studies, and extracting conclusions.
Additional participants in this model include part-time laboratory technicians
assisting in the fabrication and testing of specimens, and engineering
practitioners advising on structural design and construction issues.
As the estimates for individual research and engineering studies were
developed, the timing of a typical degree cycle (Ph.D. or M.S. program), and
the amount of work (fabrication and testing of specimens) that could
reasonably be accomplished at a typical research facility within a typical
degree cycle, were considered. The following basic budget categories were
included in the estimates:

Salary (lead researcher and graduate students)

Tuition (graduate students)

Travel (all participants)

Specimen material and fabrication costs

Laboratory costs

Laboratory labor costs (technicians)

Consultant costs (engineering advisors)

Unit direct costs for the above budget categories were based on typical (yet
conservative) values taken from major research institutions across the United
States. In addition to itemized direct costs, budget estimates included an
allowance for typical indirect costs associated with the university research
program model. These included an assumed fringe benefit rate of 1.15 on
salaries, and an overhead rate of 1.6 on direct charges (except tuition).

5-4

5: Program Recommendations

ATC-115

95% Draft
5.3

Priority and Schedule Recommendations

The recommended program of research and engineering study is ambitious,


and practical implementation will require prioritization. Practical
considerations for prioritization include the need to define material properties
for high-strength reinforcement, the timing of future code update cycles, the
time it takes to fabricate and test large numbers of specimens, the relative
costs and potential benefits associated with the use of high-strength
reinforcement in selected areas of reinforced concrete design and
construction (i.e., where it is most likely to be used, and where it provides the
most advantageous result). Based on the above considerations, the objectives
identified in Table 5-3 are recommended as a basis for prioritization of the
program.
Table 5-3
Priority
Level

Recommended Priorities Based on Implementation Objectives and Target Milestone


Dates
Objective

Target
Milestone

Approximate
Timeline

Revise ASTM A615 to include Grade 100 reinforcement

2016 Interim
Provisions for
ACI 318

Research completed
by December 2015

Modify ACI 318 to allow the use of ASTM A706 Grade 80


reinforcement in special seismic systems

2016 Interim
Provisions for
ACI 318

Research completed
by December 2015

Modify ACI 318 to allow the use of ASTM A615 Grade


100 reinforcement in general applications (gravity, wind,
and ordinary seismic systems)

2019 edition of
ACI 318

Research completed
by December 2017

Develop a new ASTM specification for Grade 100


reinforcement for use in special seismic systems

2022 (or later)


edition of
ACI 318

Research completed
by December 2020
(or later)

Modify ACI 318 to allow the use of Grade 100


reinforcement in special seismic systems

2022 (or later)


edition of
ACI 318

Research completed
by December 2020
(or later)

To date, ACI Committee 318 has operated on a three-year code update cycle,
but the schedule for future update cycles is in transition. At present, ACI
Committee 318 is considering the publication of Interim Provisions in 2016,
followed by a 2019 edition of ACI 318. Beyond the 2019 edition, it is
unclear if future updates will occur on a three-year, five-year, or six-year
cycle.
The milestones listed in Table 5-3 represent an idealized timeline that targets
future editions of ACI 318 in the relative near-term, assuming the availability
of adequate funding, sufficient production and availability of high-strength
reinforcing materials, and the availability of appropriately equipped

ATC-115

5: Program Recommendations

5-5

95% Draft

laboratory facilities capable of performing the necessary testing. Individual


research or engineering studies contributing to each objective at each priority
level are summarized in the sections that follow.
5.3.1

Priority Level 1 Revise ASTM A615 to Include Grade 100


Reinforcement

Revising ASTM A615 to include Grade 100 reinforcement is considered a


Priority Level 1 objective because the existence of a specification for the
material will allow its use in certain applications. Demand for the material
will then provide an opportunity for material producers to refine their
metallurgy and manufacturing processes to better meet requirements (e.g.,
ductility and tensile-to-yield strength ratios) related to engineering needs
associated with the development of Grade 100 reinforcement for seismic and
other applications.
The ASTM committee responsible for specification A615 has begun the
process of adding Grade 100 to its specification. Data from recent heats has
already been collected and analyzed. Test results from additional heats are
required and confirmation testing by an independent laboratory is needed.
Studies related to Priority Level 1 objectives are summarized in Table 5-4.
Table 5-4

Priority Level 1 Revise ASTM A615 to Include Grade 100 Reinforcement

Reference
Section
4.3.1

Research or Engineering Study


Mechanical Properties of Recent Heats of High Strength Reinforcement
Total Estimated Budget Priority Level 1

Estimated
Budget
$100,000
$100,000

The first version of ASTM A615 to include Grade 100 reinforcement may
require adjustment before ACI Committee 318 is willing to adopt it. Studies
to facilitate adoption of ASTM A615 Grade 100 reinforcement by ACI
Committee 318 are included under Priority Level 3.
5.3.2

Priority Level 2 Modify ACI 318 to Allow the use of ASTM


A706 Grade 80 Reinforcement in Special Seismic Systems

Modifying ACI 318 to allow use ASTM A706 Grade 80 reinforcement in


special seismic systems is considered a Priority Level 2 objective because
there is a strong demand for use of high-strength reinforcement in this area,
and the research and engineering study to make a change to Grade 80
reinforcement is relatively limited. Studies related to Priority Level 2
objectives are summarized in Table 5-5. To meet the target milestone of the
2016 Interim Provisions of ACI 318, testing associated with these studies
would need to be completed by December 2015.

5-6

5: Program Recommendations

ATC-115

95% Draft
Table 5-5
Reference
Section

Priority Level 2 Modify ACI 318 to Allow the use of ASTM A706 Grade 80
Reinforcement in Special Seismic Systems
Research or Engineering Study

4.9.1

Moment Curvature and Rotational Capacity (Grade 80 reinforcement only)

4.9.4.3

Cyclically Loaded Interior Joints (Grade 80 reinforcement only)

4.9.5

Performance of Moment Frames Systems (Grade 80 reinforcement only)

4.9.10

Special Flexure-Critical Walls (Grade 80 reinforcement only)

4.9.11

Performance of Flexure-Critical Wall Systems (Grade 80 reinforcement only)

4.10.1

Trial Engineering Designs for Use of Grade 80 Reinforcement in Special


Seismic Systems
Total Estimated Budget Priority Level 2

Estimated
Budget(1)
$50,000
$509,000
$88,000
$696,000
$88,000
$268,000
$1,699,000

Notes: (1) Budgeted amounts have been prorated for the conduct of tests or engineering studies on Grade 80
reinforcement only.

Studies listed in Table 5-5 address knowledge gaps related to members


reinforced with Grade 80 reinforcement. These include the performance of
moment frames and structural wall systems, and the required depth of
interior joints in special moment frames. These studies also include
investigation of configurations with higher strength reinforcement (i.e.,
Grade 100 and Grade 120 reinforcement). To achieve Priority Level 2
objectives, only the subset of work related to Grade 80 reinforcement need to
be completed within this timeline. Tests on higher grade reinforcement can
be deferred to Priority Level 5.
5.3.3

Priority Level 3 Modify ACI 318 to Allow the use of ASTM


A615 Grade 100 Reinforcement in General Applications
(Gravity, Wind, and Ordinary Seismic Systems)

Modifying ACI 318 to allow use ASTM A615 Grade 100 reinforcement for
general application in gravity, wind, and ordinary seismic systems is
considered a Priority Level 3 objective because the level of research and
engineering study required to validate its use is more significant than can be
realistically be accomplished in a shorter timeframe. Studies related to
Priority Level 3 objectives have been further subdivided into relative
priorities of high, moderate, and low, as summarized in Table 5-6.
It should be noted that use of Grade 100 reinforcement may not be
appropriate in all applications. Elements with stringent serviceability
requirements (e.g., slabs and beams) may experience limited benefit from
the use of high-strength reinforcement, because serviceability concerns are
expected to control, while elements without stringent serviceability

ATC-115

5: Program Recommendations

5-7

95% Draft

requirements (e.g., foundations, columns, and walls) may potentially see


greater benefit.
Table 5-6

Priority Level 3 Modify ACI 318 to Allow the use of ASTM A615 Grade 100
Reinforcement in General Applications (Gravity, Wind, and Ordinary Seismic Systems)

Reference
Section

Research or Engineering Study

Estimated
Budget

High (relative priority)


4.3.2

Detailed Mechanical Property Tests of Grade 100 and Grade 120


Reinforcement (1)

$219,000

4.4.1

Flexural Strength and Tensile Strain Limits

$149,000

4.4.2

Deflection Warning for Flexural Members Subjected to Gravity Loads

$518,000

4.4.5.1

One-Way Shear in Beams without Shear Reinforcement

$692,000

4.4.5.2

Two-Way Shear in Slabs without Shear Reinforcement

$716,000

4.4.7

Shear Friction

$384,000

4.6.2

Minimum Reinforcement Ratio for Slabs and Footings

4.7.1

Development and Splice Lengths

$738,000

4.7.2

Hooked Bar Development Length

$988,000

4.7.3

Headed Bar Development Length

$623,000

4.9.7

Ordinary Flexure-Critical Walls (1)

$1,226,000

4.10.2

Trial Engineering Designs for Use of Grade 100 Reinforcement in General


Applications

$43,000

$268,000

Moderate (relative priority)


4.4.3

Column Strength

$26,000

4.5.1

Deflection of Flexural Members

$611,000

4.5.2

Crack Control of Flexural Members

$341,000

4.6.1

Minimum Reinforcement Ratio for Beams

4.8.1

Flexural Stiffness (1)

$105,000

4.8.2

Effective Stiffness for Column Slenderness

$383,000

$43,000

Low (relative priority)


4.4.6

Shear Strength of Beams with Shear Reinforcement

$716,000

4.8.3

Moment Redistribution

$209,000

4.9.8

Special and Ordinary Shear-Critical Walls

(1)

Total Estimated Budget Priority Level 3

$3,105,000
$9,937,000

Notes: (1) Budgeted amount has been prorated to account for the conduct of tests on Grade 100 reinforcement
only.

5-8

5: Program Recommendations

ATC-115

95% Draft
5.3.4

Priority Level 4 Develop a new ASTM specification for


Grade 100 Reinforcement for use in Special Seismic
Systems

Development of a new ASTM specification that includes Grade 100


reinforcement for use in special seismic systems is considered a Priority
Level 4 objective. Prior to developing such a specification, research that
establishes the required mechanical properties, and what elongation and
tensile-to-yield strength ratios can be achieved for this grade of
reinforcement, must be completed. Studies related to Priority Level 4
objectives are summarized in Table 5-7.
Table 5-7
Reference
Section

Priority Level 4 Develop a new ASTM specification for Grade 100 Reinforcement for
use in Special Seismic Systems
Research or Engineering Study

Estimated
Budget

4.3.2

Detailed Mechanical Property Tests of Grade 100 and Grade 120


Reinforcement (1)

$219,000

4.9.3

Cyclically Loaded Beams and Columns Initial Tests and Analytical Studies

$749,000

4.9.9

Special Flexure-Critical Walls Initial Tests

$591,000

Total Estimated Budget Priority Level 4

$1,559,000

Notes: (1) Budgeted amount has been prorated to account for the portion of tests conducted under Priority
Level 3.

5.3.5

Priority Level 5 Modify ACI 318 to Allow the use of Grade


100 Reinforcement in Special Seismic Systems

Modifying ACI 318 to allow the use of Grade 100 reinforcement in special
seismic systems is considered a Priority Level 5 objective. It is anticipated
that there will high demand for the use of Grade 100 reinforcement in
seismic applications; however, the level of effort needed to justify its use is
significantly greater than for other priorities. Studies related to Priority
Level 5 objectives have been further subdivided into relative priorities of
high and moderate, as summarized in Table 5-8.
5.4

Other Conclusions and Recommendations

Several issues raised during the preparation of this Roadmap were judged
capable of resulting in a potential code change without the need for additional
research or engineering study. Some issues were judged as not needing
further action or consideration, either in the form of additional research or
code change. Finally, some issues were judged as needing further research,
but the resulting studies were not considered critical for implementation of
high-strength reinforcement in ACI 318. These are included here for
reference, and might warrant additional study at a later time.
ATC-115

5: Program Recommendations

5-9

95% Draft
Table 5-8

Priority Level 5 Modify ACI 318 to Allow the use of Grade 100 Reinforcement in
Special Seismic Systems

Reference
Section

Research or Engineering Study

Estimated
Budget

High (relative priority)


4.9.8

Special and Ordinary Shear-Critical Walls (1)

4.9.10

Special Flexure-Critical Walls (2)

4.9.11

$1,552,000
$1,413,000

Performance of Flexure-Critical Wall Systems

(2)

$178,000

Moderate (relative priority)


4.8.1

Flexural Stiffness (1)

$105,000

4.9.1

Moment Curvature and Rotational Capacity (2)

$101,000

4.9.2

Factor for Estimating Expected Flexural Strength

$197,000

4.9.4.1

Cyclically Loaded Beams

4.9.4.2

Cyclically Loaded Columns

$1,654,000
$1,734,000
(2)

4.9.4.3

Cyclically Loaded Interior Joints

$1,033,000

4.9.4.4

Cyclically Loaded Exterior Joints

4.9.4.5

Two-Way Shear in Slab-Column Intermediate Moment Frames

$809,000

4.9.5

Performance of Moment Frames Systems (2)

$178,000

4.9.6

Multi-Bay, Multi-Story Frames

$1,429,000

$1,040,000
(1)

4.9.7

Ordinary Flexure-Critical Walls

$613,000

4.10.3

Trial Engineering Designs for Use of Grade 100 Reinforcement in Special


Seismic Systems
Total Estimated Budget Priority Level 5

$268,000
$12,304,000

Notes: (1) Budgeted amount has been prorated to account for the portion of tests or engineering studies conducted
under Priority Level 3.
(2) Budgeted amount has been prorated to account for the portion of tests or engineering studies conducted
under Priority Level 2.

5.4.1

Potential Code Changes that can be Implemented without


Additional Research

Issues that were judged capable of resulting in a potential code change


without the need for additional research or engineering study are summarized
in Table 5-9. These issues are applicable to Priority Level 3 objectives to
allow the use of Grade 100 reinforcement for general application in gravity,
wind, and ordinary seismic systems.
In each case, it was felt that code change proposals could be developed, and
approved, based on the consensus opinion of ACI Committee 318 without
additional information. To facilitate adoption, it may be useful to develop a
published source (e.g., a summary paper in Concrete International or journal
article) that can serve as a reference for the proposed change.

5-10

5: Program Recommendations

ATC-115

95% Draft
Table 5-9

Potential Code Changes that can be Implemented without Additional Research

Reference
Section

Recommended
Action

Issue

3.4.4.4

Mechanical splices that develop the actual tensile


strength of high-strength reinforcment

ACI Committee 318 to determine the


need for such a mechanical splice
based on consensus opinion

3.4.5.1

Bar extensions in one-way slabs

ACI Committee 318 to consider the


need for bar extensions proportional to
yield strength based on committee
opinion

3.4.3

Spacing of transverse reinforcement in members


resisting gravity and wind loads or in ordinary
seismic-force-resisting systems

ACI Committee 318 to determine


required spacing of transverse
reinforcement based on existing tests
using Grade 60 reinforcement

3.4.6

Horizontal support of offset column reinforcement

ACI Committee 318 to determine the


need for additional horizontal support
based on consensus opinion

5.4.2

Issues Not Requiring Further Action or Code Change

Several issues considered during the preparation of this roadmap were judged
as not warranting further action or a code change. These findings should be
presented to ACI Committee 318 for consideration and confirmation, but no
further action is expected or recommended. Issues not requiring further
action or code change are summarized in Table 5-10.
Table 5-10 Issues Not Requiring Further Action or Code Change
Reference
Section

Issue

Recommended
Action

3.1.1.2

Longitudinal bar maximum useable compressive stress


limited to 80 ksi

None

3.1.2.4

Minimum reinforcement ratio for deep beams

None

3.1.3

Strut and tie modeling using high-strength reinforcement

None

3.1.4.1

Minimum concrete strength for use with high-strength


reinforcement

None

3.1.4.3

Bonded reinforcement ratios for members with


unbonded post-tensioning

None

3.3.3

Minimum reinforcement ratio for columns

None

3.4.4.3

Headed reinforcement head size and attachment to bars


(non-seismic applications)

None

3.4.5.2

Bar extensions in two-way slabs

None

5.4.3

Other Potential Studies

Two studies were judged not essential for the adoption of high-strength
reinforcement into ACI 318 for gravity, wind, or seismic applications. These

ATC-115

5: Program Recommendations

5-11

95% Draft

are listed in Table 5-11 for reference, and might warrant additional study at a
later time.
Table 5-11

Other Potential Studies

Reference
Section

Research or Engineering Study

Estimated
Budget

4.4.4

Tension Regions of Shells and Folded Plates

$834,000

4.4.8

High-Cycle Elastic Fatigue of High-Strength Reinforcing Bars

$550,000

Total Estimated Budget Other Potential Studies

5.5

$1,384,000

Implementation Recommendations

Implementation of the program outlined in this Roadmap will likely involve


the following activities:

Identification of key collaborators and potential funding partners

Technical oversight, including ongoing monitoring of results and


adjustment of priorities

Technical synthesis and development of products

Advocacy and adoption into codes and standards

The path from research results to code change is not always direct. Given the
magnitude of the program, and the complexity and extent of code changes
needed to incorporate the use of high-strength reinforcement into ACI 318, a
plan should be implemented to manage the research effort, coordinate
between studies, continually monitor results, and adjust the scope and
priorities of future studies as new information becomes available.
5.5.1

Key Collaborators

Use of high-strength reinforcement in reinforced concrete design and


construction has attracted the attention of a number of stakeholders who are
potential collaborators on the implementation of this Roadmap. These
include material producers, industry associations, and governmental
agencies, some of which will be providers of necessary information, or
sources of supplemental funding.
Potential key collaborators include the American Concrete Institute (ACI),
Portland Cement Association (PCA), and Concrete Reinforcing Steel
Institute (CRSI). Potential Federal partners include the National Institute for
Standards and Technology (NIST), Federal Emergency Management Agency
(FEMA), National Science Foundation (NSF), and Department of Energy
(DOE).

5-12

5: Program Recommendations

ATC-115

95% Draft

In some cases, logical partners can be identified by the nature of the studies.
One such example is fatigue testing of beams high-strength reinforcement, in
which case logical partners might be the Transportation Research Board
(TRB) of the National Academies, and the Federal Highway Administration
(FHWA).
Other potential partners include industry groups that are likely to benefit
from the use of high-strength reinforcement, such as the Electric Power
Research Institute (EPRI) and Nuclear Energy Institute (NEI).
5.5.2

Technical Oversight

The purpose of a technical oversight function is to make sure that studies are
properly focused, and implemented as efficiently as possible. It also
provides a mechanism to adjust the scope and priorities of future tests based
on results of tests completed to date.
Technical oversight can be used to: (1) monitor how each study fits within
the overall program; (2) provide advice regarding technical aspects of the
testing, including specimen design, parametric coverage of variables, loading
protocols, and necessary data; (3) evaluate whether a single series of tests can
be used to addresses more than one issue; (4) recommend priorities based on
the timing of future ACI 318 update cycles and relative impacts to design and
construction; (5) help ensure that research stays on target and will effectively
lead to code changes; (6) provide quality assurance reviews of test results for
accuracy and reasonableness of conclusions; and (7) provide
recommendations for modifications to the program as intermediate results are
obtained and reviewed.
Ideally, technical oversight should be performed by a panel consisting of
engineering practitioners and researchers, with representation from ACI
Committee 318 and relevant subcommittees.
5.5.3

Technical Synthesis

Given the number and complexity of the research and engineering studies
outlined in this Roadmap, and the extent of code changes necessary to
incorporate the use of high-strength reinforcement into ACI 318, there is a
need to process information from individual research studies into a coherent
set of results (i.e., synthesize). A technical synthesis function would serve to
bring consistency to the interpretation of results, and would use the resulting
test data to formulate coordinated recommendations to be sent to ACI
Committee 318 for consideration.

ATC-115

5: Program Recommendations

5-13

95% Draft

A technical synthesis strategy has been used in other major research and
development programs, and was instrumental in the development of new
steel moment frame guidelines following the 1994 Northridge earthquake.
FEMA 350, Recommended Seismic Design Criteria for New Steel MomentFrame Buildings (FEMA, 2000) is one in a series of products developed
based on technical synthesis of the results of a multi-year, multi-million
dollar experimental research effort funded by FEMA.
Ideally, technical synthesis should be performed by a group of experts on the
topic under consideration, including of engineering practitioners and
researchers, with representation from ACI Committee 318 and relevant
subcommittees.
5.5.4

Adoption into Codes and Standards

Ideas for changes to ACI 318 can originate from many sources, including
engineers not associated with the ACI 318 committee process, responses
from the public review process, and the work of other committees. Most,
however, originate from members of ACI Committee 318. Change is
initiated with a proposal that includes an explanation of why the change is
being proposed, the proposed change to the code provisions and commentary
sections, and supporting information. The reason for the proposed change
sets the stage with background information, and is generally relatively brief.
The code and commentary changes are prepared in strike-through and
underline form. Supporting information is often elaborate, especially for
highly technical changes, and can be similar to a technical paper. For the
research and engineering studies outlined in this Roadmap, resulting code
change proposals will require substantial supporting information.
Proposals are most often prepared by one or more subcommittee members,
but first drafts can come from other sources and assigned to the
subcommittee. Once the draft is available, the subcommittee vets the
proposal through a balloting process. After the proposal is modified and
approved by the subcommittee, it is balloted by the main committee. Ballot
results are addressed by the subcommittee and sent back to the main
committee for re-balloting. This process can iterate between the main
committees and the subcommittee multiple times.
To address code change proposals related to the use of high-strength
reinforcement, ACI 318 Committee has formed a subcommittee, designated
ACI 318R. Results from the research and engineering studies outlined in this
Roadmap can be package for use by ACI 318 in multiple ways:

5-14

5: Program Recommendations

ATC-115

95% Draft

1. Raw data for individual studies can be provided to ACI 318R, and its
members can interpret the results and create code change proposals.
2. Data from studies can be processed by individual research teams, and
general recommendations for change can be provided.
3. Data from studies can be processed by individual research teams, and
formal code change proposals can be provided.
4. Data from studies can be processed by individual research teams, but
results across multiple studies are synthesized by a technical synthesis
group administered as part of the program.
5. Results from the overall program can synthesized by a technical
synthesis group administered as part of the program and charged with the
development of coordinated packages of formal code change proposals.
ACI Committee 318 is a committee of volunteers that could potentially
become overwhelmed by a substantial number of complex and interrelated
change proposals. For that reason, the approaches identified in items 4 and 5
are likely to be the preferred approach for Subcommittee ACI 318R, and will
likely speed the process of adoption within ACI Committee 318. In some
cases, coordination between studies is not needed, and approaches identified
in 1, 2 or 3 could be used. Of these options, 2 is the most desirable because it
will allow for some thought to be given on how to approach the change
proposal, and the time to generate general recommendations is less than the
time it could take to produce formal code change proposals. The approach in
item 3 is the least desirable because of potential delays in considering the
information, and lack of coordination with other changes. The approach in
item 1 places the greatest burden on Subcommittee ACI 318R, but it may be
necessary in some cases, particularly when code update cycle deadlines are
approaching.

ATC-115

5: Program Recommendations

5-15

References
ACI, 1936, Building Regulations for Reinforced Concrete, ACI 501-36T,
ACI Committee 501, American Concrete Institute, Detroit,
Michigan.
ACI, 1941, Building Code Requirements for Reinforced Concrete, ACI 31841, ACI Committee 318, American Concrete Institute, Detroit,
Michigan.
ACI, 1963, Building Code Requirements for Reinforced Concrete, ACI 31863, ACI Committee 318, American Concrete Institute, Detroit,
Michigan.
ACI, 1971a, Building Code Requirements for Reinforced Concrete, ACI 31871, ACI Committee 318, American Concrete Institute, Detroit,
Michigan.
ACI, 1971b, Commentary on Building Code Requirements for Reinforced
Concrete (ACI 318-71), ACI Committee 318, American Concrete
Institute, Detroit, Michigan.
ACI, 1983, Building Code Requirements for Reinforced Concrete, ACI 31883, ACI Committee 318, American Concrete Institute, Detroit,
Michigan.
ACI, 1995, Control of Deflection in Concrete Structures, ACI 435R-95, ACI
Committee 435, American Concrete Institute, Detroit, Michigan.
ACI, 1997, Considerations for Design of Concrete Structures Subjected to
Fatigue Loading, ACI 215R-92, ACI Committee 215, American
Concrete Institute, Farmington Hills, Michigan.
ACI, 1999, Building Code Requirements for Structural Concrete and
Commentary, ACI 318-99 and ACI 318R-99, ACI Committee 318,
American Concrete Institute, Farmington Hills, Michigan.
ACI, 2002a, Building Code Requirements for Structural Concrete and
Commentary, ACI 318-02 and ACI 318R-02, ACI Committee 318,
American Concrete Institute, Farmington Hills, Michigan.
ACI, 2002b, Recommendations for Design of Beam-Column Connections in
Monolithic Reinforced Concrete Structures, ACI 352R-02, Joint

ATC-115

References

A-1

ACI-ASCE Committee 352, American Concrete Institute,


Farmington Hills, Michigan.
ACI, 2003, Bond and Development of Straight Reinforcing Bars in Tension,
ACI 408R-03, ACI Committee 408, American Concrete Institute,
Farmington Hills, Michigan.
ACI, 2006, Guide for the Design and Construction of Structural Concrete
Reinforced with FRP Bars, ACI 440.1R-06, ACI Committee 440,
American Concrete Institute, Farmington Hills, Michigan.
ACI, 2010a, Design Guide for the Use of ASTM A1035/A1035M Grade 100
Steel Bars for Structural Concrete, ACI ITG-6R-10, ACI Innovation
Task Group 6, American Concrete Institute, Farmington Hills,
Michigan.
ACI, 2010b, Specification for Tolerances for Concrete Construction and
Materials and Commentary, ACI 117-10, ACI Committee 117,
American Concrete Institute, Farmington Hills, Michigan.
ACI, 2011, Building Code Requirements for Structural Concrete and
Commentary, ACI 318-11, ACI Committee 318, American Concrete
Institute, Farmington Hills, Michigan.
ACI, 2014a, Building Code Requirements for Structural Concrete and
Commentary, ACI 318-14, ACI Committee 318, American Concrete
Institute, Farmington Hills, Michigan.
ACI, 2014b, Code Requirements for Nuclear Safety-Related Concrete
Structures and Commentary, ACI 349-13, ACI Committee 349,
American Concrete Institute, Farmington Hills, Michigan.
ACI-ASCE Committee 326, 1962, Shear and diagonal tension, ACI
Journal Proceedings, Vol. 59, No. 1, pp. 1-30; No. 2, pp. 277-334;
and No. 3, pp. 352-396.
ACI-ASCE Committee 426, 1973, Shear Strength of Reinforced Concrete
Members, ACI 426R-74, Journal of the Structural Division, ASCE,
Vol. 99, No. 6, pp. 1148-1157.
Angelakos, D., Bentz, E., and Collins, M., 2001, Effect of concrete strength
and minimum stirrups on shear strength of large members, ACI
Structural Journal, Vol. 98, No. 3, pp. 290-300.
Aoyama, H., 2001, Design of Modern Highrise Reinforced Concrete
Structures, Imperial College Press, London, United Kingdom.

A-2

References

ATC-115

ASCE, 2010, Minimum Design Loads for Buildings and Other Structures,
ASCE/SEI 7-10, American Society of Civil Engineers, Structural
Engineering Institute, Reston, Virginia.
AS/NZS, 2001, AS/NZS 4671 Steel Reinforcing Materials, Joint Standards
Australia/Standards New Zealand, Wellington, New Zealand and
Sydney, Australia.
ASTM, 1974, Standard Specification for Low-Alloy Steel Deformed and
Plain Bars for Concrete Reinforcement, ASTM A706, ASTM
International, West Conshohocken, Pennsylvania.
ASTM, 2009a, Standard Specification for Deformed and Plain Carbon-Steel
Bars for Concrete Reinforcement, ASTM A615-09b, ASTM
International, West Conshohocken, Pennsylvania.
ASTM, 2009b, Standard Specification for Low-Alloy Steel Deformed and
Plain Bars for Concrete Reinforcement, ASTM A706-09b, ASTM
International, West Conshohocken, Pennsylvania.
ASTM, 2011, Standard Specification for Deformed and Plain, Low-Carbon,
Chromium, Steel Bars for Concrete Reinforcement, ASTM A1035,
ASTM International, West Conshohocken, Pennsylvania.
ASTM, 2013, Standard Specification for Headed Steel Bars for Concrete
Reinforcement, ASTM A970 including Annex A1, ASTM
International, West Conshohocken, Pennsylvania.
Bang, C.-J., 2014, Application of High-Strength/Large-Diameter Headed
Deformed Bars in Construction of Nuclear Power Plant Structures,
presentation to ACI Committees 349 and 359.
Bilow, D.N., and Kamara, M.E., 2008, Fire and concrete structures,
Proceedings, ASCE Structures Congress, Vancouver, Canada.
Birrcher, D.B., Tuchscherer, R.G., Huizinga, M., and Bayrak, O., 2013,
Minimum web reinforcement in deep beams, ACI Structural
Journal, Vol. 110, No. 2, pp. 297-306.
Bischoff, P.H., 2007, Deflection calculation of FRP reinforced concrete
beams based on modifications to the existing Branson equation,
Journal of Composites for Construction, Vol. 11, No. 1, pp. 4-14.
Bischoff, P.H., and Scanlon, A., 2007, Effective moment of inertia for
calculating deflections of concrete members containing steel
reinforcement and FRP reinforcement, ACI Structural Journal, Vol.
104, No. 1, pp. 68-75.

ATC-115

References

A-3

Bournonville, M., Dahnke, J., and Darwin, D., 2004, Statistical Analysis of
the Mechanical Properties and Weight of Reinforcing Bars, SL
Report 04-1, Structural Engineering and Materials Laboratory,
University of Kansas, Lawrence, Kansas.
Branson, 1977, Deformation of Concrete Structures, McGraw Hill Book Co.,
Advanced Book Program, New York, New York.
Branson, D.E., and Trost, H., 1982, Unified Procedures for Predicting the
Deflection and Centroidal Axis Location of Partially Cracked NonPrestressed Members, ACI Journal Proceedings, Vol. 79, No. 2,
pp. 119-130.
Brown, M., Bayrak, O., and Jirsa, J., 2006, Design for shear based on
loading conditions, ACI Structural Journal, Vol. 103, No. 4, pp.
541-550.
Budek, A., Priestley, M., and Lee, C., 2002, Seismic design of columns with
high-strength wire and strand as spiral reinforcement, ACI
Structural Journal, Vol. 99, No. 5, pp. 660-670.
Chen, M., Pantoli, E., Wang, X., Espino, E., Mintz, S., Conte, J., Hutchinson,
T., Marin, C., Meacham, B., Restrepo, J.I., Walsh, K., Englekirk,
R., Faghihi, M., and Hoehler, M., 2012, Design and construction of
a full-scale 5-story base isolated building outfitted with nonstructural
components for earthquake testing at the UCSD-NEES facility,
Proceedings, ASCE Structures Congress 2012, pp. 1349-1360.
Choi, W.-S., Park, H.-G. , Chung, L., and Kim, J.-K., 2014, Experimental
study for Class B lap splice of 600 MPa (87 ksi) reinforcing bars,
ACI Structural Journal, Vol. 111, No. 1, pp. 49-58.
Collins, M., and Kuchma, D., 1999, How safe are our large, lightly
reinforced, concrete beams, slabs, and footings? ACI Structural
Journal, Vol. 96, No. 4, pp. 482-490.
CRSI, 2013, data from an unpublished database of the Concrete Reinforcing
Steel Institute, Schaumburg, Illinois.
Davey, R.A., and Blaikie, E.L., 2005, On the flexural ductility of very
lightly reinforced concrete sections, Proceedings, 2005 NZSEE
Conference, Taupo, New Zealand.
Dazio, A.D.A., Beyer, K., and Bachmann, H., 2009, Quasi-static cyclic tests
and plastic hinge analysis of RC structural walls, Engineering
Structures, Vol. 31, pp. 1556-1571.

A-4

References

ATC-115

Edwards, W.T., and Gamble, W.L., 1986, Strength of Grade 60 reinforcing


bars after exposure to fire temperatures, Concrete International,
Vol. 8, No. 10, pp. 17-19.
Erlemann, G.G., 1997, Reinforcing Bar Specifications 1911 through
1968, Engineered Concrete Structures, Portland Cement
Association, Skokie, Illinois.
Erlemann, G.G., 1999, Steel reinforcing bar specifications in old
structures, Concrete International, Vol. 21, No. 4, pp. 49-50.
Fei, J., and Darwin, D., 1999, Fatigue of High Relative Rib Area Reinforcing
Bars, SM Report No. 54, University of Kansas Center for Research,
Inc., Lawrence, Kansas.
FEMA, 2000, Recommended Seismic Design Criteria for New Steel MomentFrame Buildings, FEMA 350, prepared by the SAC Joint Venture for
the Federal Emergency Management Agency, Washington, D.C.
FEMA, 2009, Quantification of Building Seismic Performance Factors,
FEMA P-695, prepared by the Applied Technology Council for the
Federal Emergency Management Agency, Washington, D.C.
Frosch, R.J., 1999, Another look at cracking and crack control in reinforced
concrete, ACI Structural Journal, Vol. 96, No. 3, pp. 437-442.
Frosch, R.J., 2001, Flexural crack control in reinforced concrete, Design
and Construction Practices to Mitigate Cracking, ACI Special
Publication 204, American Concrete Institute, Farmington Hills,
Michigan.
Frosch, R.J., 2002, Modeling and control of side face beam cracking, ACI
Structural Journal, Vol. 99, No. 3, pp. 376-385.
Gamble, W.L., 2011, Loads, Codes, and Fire Endurance, ACI Special
Publication 279, American Concrete Institute, Farmington Hills,
Michigan.
Garay-Moran, J.D., and Lubell, A.S., 2008, Behavior of concrete deep
beams with high strength reinforcement, Proceedings, ASCE
Structures Congress, Reston, Virginia.
Gaston, J.R., and Hognestad, E., 1962, High strength bars as concrete
reinforcement, Part 3 Tests of full-scale roof girder, Journal of the
PCA Research and Development Laboratories, Vol. 4, No. 2, pp.
10-23.

ATC-115

References

A-5

Gergely, P., and Lutz, L.A., 1968, Maximum crack width in reinforced
concrete flexural members, Causes, Mechanism, and Control of
Cracking in Concrete, ACI Special Publication 20, American
Concrete Institute, Detroit, Michigan.
Hage, S.E., 1974, The Second-Order Analysis of Reinforced Concrete
Frames, M.S. Thesis, University of Alberta, Edmonton, Canada.
Han, S.W., and Jee, N.Y., 2005, Seismic behaviors of columns in ordinary
and intermediate moment resisting concrete frames, Engineering
Structures, Vol. 27, No. 6, pp. 951-962.
Helgason, T., Hanson, J.M., Somes, N.F., Corley, W.G., and Hognestad, E.,
1976, Fatigue Strength of High-Yield Reinforcing Bars, National
Cooperative Highway Research Program Report 164, Transportation
Research Board of the National Academies, Washington, D.C.
Hognestad, E., 1961, High strength bars as concrete reinforcement, Part 1
Introduction to a series of experimental reports, Journal of the PCA
Research and Development Laboratories, Vol. 3, No. 3, pp. 23-29.
Hognestad, E., 1962, High strength bars as concrete reinforcement, Part 2
Control of flexural cracking, Journal of the PCA Research and
Development Laboratories, Vol. 4, No. 1, pp. 46-63.
Hopkins, D., and Poole, R., 2005, Grade 500E Reinforcing Steel, Tests on
Micro-Alloy and Quenched and Tempered Samples Available in New
Zealand, Department of Building and Housing, Wellington, New
Zealand.
Hwang, H.H.M., and Hsu, H.M., 1993, Seismic LRFD Criteria for RC
Moment-Resisting Frame Buildings, Journal of Structural
Engineering, Vol. 119, No. 6, pp. 1807-1824.
Hwang, H.H.M., and Hsu, H.M., 1994, A study of reliability-based seismic
design criteria, Structural Safety & Reliability, ISBN 90 5410 357 4,
Balkema, Rotterdam.
ICC, 2012, International Building Code, International Code Council,
Washington, D.C.
Ishikawa, Y., Kimura, H., Takatsu, H., and Ousalem, H., 2008, Ultimate
deformation of R/C columns using high-strength concrete and highstrength steel bars under earthquake loading, Proceedings,
8th International Symposium on Utilization of High-Strength and
High-Performance Concrete, S1-5-4, Tokyo, Japan.

A-6

References

ATC-115

Jhamb, I.C., and MacGregor, J. G., 1974, Effect of Surface Characteristics


On Fatigue Strength of Reinforcing Steel, Abeles Symposium on
Fatigue of Concrete, ACI Special Publication 41, American
Concrete Institute, Detroit, pp. 139-167.
Kaar, P.H., 1966, High strength bars as concrete reinforcement, Part 8
Similitude in flexural cracking of T-beam flanges, Journal of the
PCA Research and Development Laboratories, Vol. 8, No. 2, pp.
2-12.
Kaar, P.H., and Hognestad, E., 1965, High strength bars as concrete
reinforcement, Part 7 Control of cracking in T-beam flanges,
Journal of the PCA Research and Development Laboratories, Vol. 7,
No. 1, pp. 42-53.
Kaar, P.H., and Mattock, A.H., 1963, High strength bars as concrete
reinforcement, Part 4 Control of cracking, Journal of the PCA
Research and Development Laboratories, Vol. 5, No. 1, pp. 15-38.
Kabeyasawa, T., and Hiraishi, H., 1998, Tests and analyses of high-strength
reinforced concrete shear walls in Japan, High-Strength Concrete in
Seismic Regions, ACI Special Publication 176, American Concrete
Institute, Farmington Hills, Michigan.
Kimura, H., Sugano, S., Nagashima, T., and Ichikawa, A., 1993, Seismic
loading tests of reinforced concrete beams using high strength
concrete and high strength steel bars, Proceedings, 3rd International
Symposium on Utilization of High-Strength Concrete, pp. 377-384,
Tokyo, Japan.
Kimura, H., and Ishikawa, Y., 2008, Seismic performance of high-strength
reinforced concrete slender walls subjected to high axial loading,
Proceedings, 8th International Symposium on Utilization of HighStrength and High-Performance Concrete, pp. 945-950, Tokyo,
Japan.
Korinda, K., 1972, Discussion No. 3 Cracking and crack control,
Proceedings, International Conference on Planning and Design of
Tall Buildings, Lehigh University, Bethlehem, Pennsylvania.
Kudder, R.J., and Gustafson, D.P., 1983, Bend tests of Grade 60 reinforcing
bars, ACI Journal Proceedings, Vol. 80, No. 3, pp. 202-209.
Larkins, T., 2004, Shear Design of Reinforced Concrete Beams:
Investigation of an Alternative Approach, M.S. Thesis, Purdue
University, West Lafayette, Indiana.

ATC-115

References

A-7

Lin, C.M., Restrepo, J.I., and Park, R., 2000, Seismic Behaviour and Design
of Reinforced Concrete Interior Beam Column Joints, Research
Report 2000-1, Department of Civil Engineering, University of
Canterbury, New Zealand.
Lowes, L.N., Lehman, D.E., Birely, A.C., Kuchma, D.A., Marley, K.P., and
Hart, C.R., 2012, Earthquake response of slender planar concrete
walls with modern detailing, Engineering Structures, Vol. 43, pp.
31-47.
LSTC, 2013, LS-DYNA, Livermore Software Technology Corporation,
Livermore, California, http://www.lstc.com/products/ls-dyna.
Lubell, A., Sherwood, T., Bentz, E., and Collins, M.P., 2004, Safe shear
design of large, wide beams, ACI Concrete International, Vol. 26,
No. 1, pp. 66-78.
Luna, B.N., 2014 (in preparation), Seismic Response of Low Aspect Ratio
Reinforced Concrete Shear Walls for Buildings and Safety-Related
Nuclear Structures, Ph.D. Dissertation, University at Buffalo,
Buffalo, New York.
MacGregor, J.G., and Hage, S.E., 1977, Stability analysis and design of
concrete frames, Journal of the Structural Division, ASCE,
Vol. 103, No. 10, pp. 1953-1970.
Marques, J., and Jirsa, J., 1975, A study of hooked bar anchorages in beamcolumn joints, ACI Structural Journal, Vol. 72, No. 5, pp. 198-209.
Mast, R.F., 1968, Auxiliary reinforcement in concrete connections,
Journal of the Structural Division, ASCE, Vol. 94, No. 6, pp. 14851504.
Mast, R.F., 1992, Unified design provisions for reinforced and prestressed
concrete flexural and compression members, ACI Structural
Journal, Vol. 89, No. 2, pp. 185-199
Mattock, A.H., and Hawkins, N.M., 1972, Shear transfer in reinforced
concreteRecent research, PCI Journal, Vol. 17, No. 2, pp. 55-75.
Mattock, A.H., Yamazaki, J., and Kattula, B.T., 1971, Comparative study of
prestressed concrete beams, with and without bond, ACI Journal
Proceedings, Vol. 68, No. 2, pp. 116-125.
Meinheit, D.F., and Felder, A.L., 2014, Vintage Steel Reinforcement in
Concrete Structures, Concrete Reinforcing Steel Institute,
Schaumburg, Illinois.

A-8

References

ATC-115

Moyer, M.J., and Kowalsky, M.J., 2003, Influence of tension strain on


buckling of reinforcement in concrete columns, ACI Structural
Journal, Vol. 100, No. 1, pp. 75-85.
Muguruma, H., and Watanabe, F., 1990, Ductility improvement of highstrength concrete columns with lateral confinement, High-Strength
Concrete: Second International Symposium, ACI Special Publication
121, American Concrete Institute, Detroit, Michigan.
Munikrishna, A., 2008, Shear Behavior of Concrete Beams Reinforced with
High Performance Steel Shear Reinforcement, M.S. Thesis, North
Carolina State University, Raleigh, North Carolina.
NACU, 1910, Standard Building Regulations for the Use of Reinforced
Concrete, Standard 4, National Association of Cement Users,
Philadelphia, Pennsylvania.
Nishiyama, M., 2009, Mechanical properties of concrete and reinforcement
state-of-the-art report on HSC and HSS in Japan, Journal of
Advanced Concrete Technology, Vol. 7, No. 2, pp. 157-182.
NIST, 2010a, Seismic Design of Cast-in-Place Concrete Diaphragms,
Chords, and CollectorsA Guide for Practicing Engineers, NIST
GCR 10-917-4 Report, prepared by the NEHRP Consultants Joint
Venture, a partnership of the Applied Technology Council and the
Consortium of Universities for Research in Earthquake Engineering,
for the National Institute of Standards and Technology,
Gaithersburg, Maryland.
NIST, 2010, Evaluation of the FEMA P-695 Methodology for Quantification
of Building Seismic Performance Factors, GCR 10-917-8, prepared
by the NEHRP Consultants Joint Venture, a partnership of the
Applied Technology Council and the Consortium for Universities for
Research in Earthquake Engineering, for the National Institute of
Standards and Technology, Gaithersburg, Maryland.
NIST, 2012, Tentative Framework for Development of Advanced Seismic
Design Criteria for New Buildings, GCR 12-917-20, prepared by the
NEHRP Consultants Joint Venture, a partnership of the Applied
Technology Council and the Consortium for Universities for
Research in Earthquake Engineering, for the National Institute of
Standards and Technology, Gaithersburg, Maryland.
NIST, 2014, Use of High-Strength Reinforcement in Earthquake-Resistant
Concrete Structures, GCR 14-917-30, prepared by the NEHRP
Consultants Joint Venture, a partnership of the Applied Technology

ATC-115

References

A-9

Council and the Consortium for Universities for Research in


Earthquake Engineering, for the National Institute of Standards and
Technology, Gaithersburg, Maryland.
Okamoto, M., Sato, T., Tanimura, Y., and Kuroiwa, T., 2004, Experimental
study on seismic performance verification method for RC members
using high-strength materials, Journal of Advanced Concrete
Technology, Vol. 2, No. 2, pp. 223-231.
OpenSees, 2011, Open System for Earthquake Engineering Simulation:
OpenSees, University of California, Berkeley, http://opensees
.berkeley.edu.
Orangun, C.O., Jirsa, J.O., and Breen, J.E., 1975, The Strength of Anchored
Bars: A Reevaluation of Test Data on Development Length and
Splices, Research Report No. 154-3F, Center for Highway Research,
University of Texas, Austin, Texas.
Ospina, C.E., 2005, Alternative for concentric punching capacity evaluation
of reinforced concrete two-way slabs, Concrete International,
Vol. 27, No. 9, pp. 53-57.
Otani, S., 1991, AIJ proposal of ultimate strength design requirements for
RC buildings with emphasis on beam-column joints, Design of
Beam-Column Joints for Seismic Resistance, ACI Special
Publication 123, American Concrete Institute, Detroit, Michigan.
Ou, Y., Kurniawan, D., and Handika, N., 2012, Shear behavior of reinforced
concrete columns with high strength steel and concrete under low
axial load, Proceedings, ACI Fall Convention, Toronto, Ontario,
Canada.
Pan, A., and Moehle, J.P., 1989, Lateral displacement ductility of reinforced
concrete flat plates, Structural Journal, Vol. 86, No. 3, pp. 250-258.
Panahshahi, N., and Lu, C., 1997, Inelastic Response of RC Intermediate
and Special Moment Resisting Frames, Proceedings, ASCE
Structures Congress, Portland, Oregon.
Paulson, C., and Hanson, J.M., 1991, Final Report: Fatigue Strength of
Welded and Mechanical Splices in Reinforcing Steel, National
Cooperative Highway Research Program, Project 10-35,
Transportation Research Board, Washington, D.C.
Paulson, C., Graham, S.K., and Rautenberg, J.M., 2013, Determination of
Yield Strength for Nonprestressed Steel Reinforcement, prepared by

A-10

References

ATC-115

Wiss, Janney, Elstner Associates, Inc., for the Charles Pankow


Foundation, RGA 04-13, WJE No. 2013.4171, Pasadena, California.
Pfister, J.F., 1964, Influence of ties on the behavior of reinforced concrete
columns, ACI Journal Proceedings, Vol. 61, No. 5, pp. 521-538.
Pfister, J.F., and Hognestad, E., 1964, High strength bars as concrete
reinforcement, Part 6 Fatigue tests, Journal of the PCA Research
and Development Laboratories, Vol. 6, No. 1, pp. 65-84.
Pfister, J.F., and Mattock, A.H., 1963, High strength bars as concrete
reinforcement, Part 5 Lapped splices in concentrically loaded
columns, Journal of the PCA Research and Development
Laboratories, Vol. 5, No. 2, pp. 27-40.
Pfund, S.J., 2012, Cyclic Response of Concrete Beams Reinforced with
ASTM A1035 Grade-120 Steel Bars, M.S. Thesis, The Pennsylvania
State University, University Park, Pennsylvania.
Price, K.R., Fields, D., and Lowes, L.N., 2014, The Impact of High-Strength
Reinforcing Steel on Current Design Practice, Research Grant
Agreement #01-13, Charles Pankow Foundation, Vancouver,
Washington.
Rautenberg, J.M., 2011, Drift Capacity of Concrete Columns Reinforced with
High Strength Steel, Ph.D. Thesis, Purdue University, West
Lafayette, Indiana.
Reineck, K.H., Kuchma, D.A., Kim, K.S., and Marx, S., 2003, Shear
database for reinforced concrete members without shear
reinforcement, ACI Structural Journal, Vol. 100, No. 2, pp. 240249.
Restrepo, J.I., Seible, F., Stephan, B., and Schoettler, M.J., 2006, Seismic
testing of bridge columns incorporating high-performance
materials, ACI Structural Journal, Vol. 103, No. 4, pp. 496-504.
Restrepo-Posada, J.I., 1992, Seismic Behaviour of Connections Between
Precast Concrete Elements, Ph.D. Thesis, University of Canterbury,
Christchurch, New Zealand.
Richard, M.J., Albano, L.D., Kelly, D.J., and Liel, A., 2010, Case study on
the seismic performance of reinforced-concrete intermediate-moment
frames using ACI design provisions, Proceedings, NASCC Steel
Conference/ASCE Structures Congress, Orlando, Florida.
Richart, F.E., 1933, Reinforced concrete column investigation, ACI
Journal Proceedings, Vol. 29, No. 2, pp. 275-284.

ATC-115

References

A-11

Richart, F.E., and Brown, R.L., 1934, An Investigation of Reinforced


Concrete Columns, Engineering Experiment Station Bulletin No.
267, University of Illinois, Urbana, Illinois.
Rodriguez, M., Botero, J., and Villa, J., 1999, Cyclic stress-strain behavior
of reinforcing steel including effect of buckling, Journal of
Structural Engineering, Vol. 125, No. 6, pp. 605612.
Schlaich, J., Schafer, K., and Jennewein, M., 1987, Toward a consistent
design of structural concrete, PCI Journal, Vol. 32, No. 3, pp. 74150.
Seliem, H.M., Hosny, A., Rizkalla, S., Zia, P., Briggs, M., Miller, S.,
Darwin, D., Browning, J., Glass, G.M., Hoyt, K., Donnelly, K., and
Jirsa, J.O., 2009, Bond characteristics of high-strength ASTM
A1035 steel reinforcing bars, ACI Structural Journal, Vol. 106, No.
4, pp. 530-539.
Shahrooz, B.M., Miller, R.A., Harris, K.A., and Russell, H.G., 2011, Design
of Concrete Structures Using High-Strength Steel Reinforcement,
National Cooperative Highway Research Program (NCHRP) Report
679, Transportation Research Board of the National Academies,
Washington, D.C.
Sheth, A., 2003, Use of intermediate RC moment frames in moderate
seismic zones, The Indian Concrete Journal, Vol. 77, No. 11,
pp. 1431-1435.
Sozen, M.A., Siess, C.P., 1963 Investigation of multiple panel reinforced
concrete floor slabs, ACI Journal Proceedings, Vol. 60, No. 8, pp.
999-1028.
Stecich, J.P., Hanson, J.M., and Rice, P.F., 1984, Bending and straightening
of Grade 60 reinforcing bars, Concrete International, Vol. 6, No. 8,
pp. 14-23.
Sugano, S., Nagashima, T., Kimura, H., Tamura, A., and Ichikawa, A., 1990,
Experimental studies on seismic behavior of reinforced concrete
members of high-strength concrete, High-Strength Concrete:
Second International Symposium, ACI Special Publication 121,
American Concrete Institute, Detroit, Michigan.
Sumpter, M.S., Rizkalla, S.H., and Zia, P., 2009, Behavior of highperformance steel as shear reinforcement for concrete beams, ACI
Structural Journal, Vol. 106, No. 2, pp. 171-177.

A-12

References

ATC-115

Tanaka, H., 1990, Effect of Lateral Confining Reinforcement on the Ductile


Behaviour of Reinforced Concrete Columns, Ph.D. Thesis,
University of Canterbury, Christchurch, New Zealand.
Tavallali, H., 2011, Cyclic Response of Concrete Beams Reinforced with
Ultrahigh Strength Steel, Ph.D. Thesis, Pennsylvania State
University, University Park, Pennsylvania.
Thomas, K., and Sozen, M.A., 1965, A Study of the Inelastic Rotation
Mechanism of Reinforced Concrete Connections, Civil Engineering
Studies, Structural Research Series No. 301, University of Illinois,
Urbana, Illinois.
TNO, 2013, Displacement Analyzer, DIANA, TNO DIANA BV, Delft, The
Netherlands, http://tnodiana.com/content/DIANA.
Todeschini, C.E., Bianchini, A.C., and Kesler, C.E., 1964, Behavior of
concrete columns reinforced with high strength steels, ACI Journal
Proceedings, Vol. 61, No. 6, pp. 701-716.
Topcu, I.B., and Karakurt, C., 2008, Properties of reinforced concrete steel
rebars exposed to high temperatures, Research Letters in Materials
Science, Vol. 2008, Article ID 814137.
Tretiakova, K., 2013, Cyclic Response of Concrete Columns Reinforced with
SAS 670 Grade-97 Steel Bars, M.S. Thesis, Pennsylvania State
University, University Park, Pennsylvania.
Tureyen, A.K., and Frosch, R.J., 2002, Shear tests of FRP-reinforced
concrete beams without stirrups, ACI Structural Journal, Vol. 99,
No. 4, pp. 427-434.
Tureyen, A.K., and Frosch, R.J., 2003, Concrete shear strength: Another
perspective, ACI Structural Journal, Vol. 100, No. 5, pp. 609-615.
Valluvan, R., 1993, Issues Involved In Seismic Retrofit of Reinforced
Concrete Frames Using Infilled Walls, Ph.D. Thesis, University of
Texas, Austin, Texas.
Wang, Y., and Restrepo, J.I., 2001, Investigation of concentrically loaded
reinforced concrete columns confined with glass fiber-reinforced
polymer jackets, ACI Structural Journal, Vol. 98, No. 3, pp. 377385.
Wood, S.L., 1989, Minimum tensile reinforcement requirements in walls,
ACI Structural Journal, Vol. 86, No. 5, pp. 582-591.

ATC-115

References

A-13

Zheng, H., and Abel, A., 1999, Fatigue Properties of Reinforcing Steel
Produced by TEMPCORE Process, Journal of Materials in Civil
Engineering, ASCE, Vol. 11, No. 2, pp. 158-165.
Zhu, S., and Jirsa, J.O., 1983, A Study of Bond Deterioration in Reinforced
Concrete Beam-Column Joints, PMFSEL Report No. 83-1,
Department of Civil Engineering, University of Texas, Austin,
Texas.
Ziehl, P.H., Cloyd, J.E., and Kreger, M.E., 1998, Evaluation of Minimum
Longitudinal Reinforcement Requirements for Reinforced Concrete
Columns, Research Report 1473-S, Center for Transportation
Research, University of Texas, Austin, Texas.
Ziehl, P.H., Cloyd, J.E., and Kreger, M.E., 2004, Investigation of minimum
longitudinal reinforcement requirements for concrete columns using
present-day construction materials, ACI Structural Journal, Vol.
101, No. 2, pp. 165-175.
Zuo, J., and Darwin, D., 2000, Splice strength of conventional and high
relative rib area bars in normal and high-strength concrete, ACI
Structural Journal, Vol. 97, No. 4, pp. 630-641.

A-14

References

ATC-115

Project Participants
Charles Pankow Foundation

Applied Technology Council

Mark J. Perniconi (Executive Director)


Charles Pankow Foundation
P.O. Box 820631
Vancouver, Washington 98682

Jon A. Heintz (Project Manager)


Applied Technology Council
201 Redwood Shores Parkway, Suite 240
Redwood City, California 94065

Project Management Committee


Andres Lepage
University of Kansas
Dept. of Civil, Environmental, and Arch. Engin.
2150 Learned Hall
1530 W. 15th Street
Lawrence, Kansas 66045

Dominic J. Kelly
(Project Technical Director)
Simpson Gumpertz & Heger, Inc.
41 Seyon Street, Bldg. 1, Suite 500
Waltham, Massachusetts 02453
David Darwin
University of Kansas
Dept. of Civil, Environmental, and Arch. Engin.
2150 Learned Hall
1530 W. 15th Street
Lawrence, Kansas 66045

Joseph C. Sanders
Charles Pankow Builders, Ltd. (retired)
1168 North Chester Avenue
Pasadena, California 91104
Andrew Whittaker
University at Buffalo
Dept. of Civil, Structural, and Environ. Engin.
230 Ketter Hall
Buffalo, New York 14260

David C. Fields
Magnusson Klemencic Associates
1301 Fifth Avenue, Suite 3200
Seattle, Washington 98101
Robert J. Frosch
Purdue University
Lyles School of Civil Engineering
550 Stadium Mall Drive
West Lafayette, Indiana 47907
Project Review Panel
Wassim Ghannoum
University of Texas at Austin
Dept. of Civil, Architectural, and Environ. Engin.
301 E. Dean Keeton, Stop C1700
Austin, Texas 78712

ATC-115

S.K. Ghosh
S.K. Ghosh Associates, Inc.
334 E. Colfax St
Palatine, Illinois 60067

B: Project Participants

B-1

Ramon Gilsanz
Gilsanz Murray Steficek LLP
129 W. 27th Street, 5th Floor
New York, New York 10001

Thomas C. Schaeffer
Structural Design Group
220 Great Circle Road, Suite 106
Nashville, Tennessee 37228

James O. Jirsa
University of Texas at Austin
Dept. of Civil, Architectural, and Environ. Engin.
ECJ Hall, Suite 4.726
Austin, Texas 78712

Loring A. Wyllie, Jr.


Degenkolb Engineers
235 Montgomery Street, Suite 500
San Francisco, California 94104

Mike Mota
Concrete Reinforcing Steel Institute
933 N. Plum Grove Road
Schaumburg, Illinois 60173

B-2

B: Project Participants

ATC-115

You might also like