You are on page 1of 116

CHAPTER ONE: INTRODUCTION

1.1

Research Background
Portland cement concrete is made with coarse aggregate, fine aggregate

(sand), portland cement, water and, in some cases, selected admixtures such as fly
ash, air-entraining agents, water-reducing agents, retarders, etc. Each constituent
influences the characteristics of the concrete and must be controlled as to
composition and quantity if the end product is to be within acceptable limits of
uniformity, workability, and strength.
With the possible exception of water, coarse and fine aggregate materials
are normally the least expensive materials in concrete and provide the greatest
volume. Typically, in the concrete, the coarse aggregate and sand will occupy
approximately 80 percent of the total volume of the finished mixture. Obviously,
the most expensive component in the concrete mix is cement. A typical value of
the cement for the same volume of concrete is 60 percent of the cost of all of the
raw materials. The amount of cement primarily depends on the volume of the
aggregates in the concrete mix.
Specifications for the fine aggregate fraction of concrete have been
developed almost exclusively on the basis of experience with natural sand for
many years, since it was virtually the only type utilized. Clean, natural sands have
rounded particles that provide good workability in concrete without the addition of
1

2
excessive quantities of either water or cement. When mineral fines are present in
natural sand, the particles are frequently clay or silt particles that may be
deleterious particles.

These have an undesirable influence on water

requirements, workability, and strength characteristics of concrete mixtures. As a


means of controlling the presence of such fines in fine aggregate, specifications
have been developed which limit the amount of minus No. 200 mesh (75m)
material. Hereinafter, minus No. 200 mesh (75m) material will be designated as
micro fines. Current specifications, including ASTM C33 and those of most state
Departments of Transportation, limit the percentage of micro fines allowed in fine
aggregate used for portland cement concrete.

Such specifications have been

utilized throughout the concrete industry for the past half-century but during that
same period the nations supply of aggregate material has undergone gradual
change [J. Fowler, 1997].
As the deposits of natural sands have slowly been depleted, it has become
necessary and economical to produce manufactured fine aggregate (MFA). MFA
is a fine aggregate processed from quarried stone that is crushed and classified to
obtain a controlled gradation and a cubical to angular particle shape.
The first commercial use of MFA was made in the early 1930s. Current
usage of MFA is over 100 times the amount utilized in 1930s and represents
approximately 20% of total concrete fine aggregate requirements. This percentage
is expected to increase in the years ahead.

3
MFA differs from natural sands in gradation, particle shape and texture.
Each of these characteristics has some influence on mixture design and influences
cement requirements, water requirements, additive requirements, workability, and
finishing characteristics of the concrete [McKeagney, 1984].

1.2

Problem Statement
The requirement of the properties of fine aggregate will be different

according to types of structure, parts of using, the condition of mixture, and


environment of casting. Therefore the typical requirement of the fine aggregate
should be provided by specifications.
Standard specifications for fine aggregate for concrete contained in ASTM
C33 permit a maximum of 7 percent finer than the No. 200 sieve (75m), if the
fines consist of dust-of-fracture essentially free of clay or shale. When the fines
are not dust-of-fracture, the limit on the minus No. 200 sieve (75m) for concrete
subject to abrasion is 5 percent. Such specifications severely limit the amount of
fine particles that can be used, even though higher fines contents can improve the
properties and the performance of the resulting concrete and reduce the cost of the
mixture in which they are used.
The production process for MFA normally generates 10 to 20 percent of
micro fines that is more than permitted by specifications. As a consequence,
excess fines must be separated from the desired sizes by screening or washing

4
operations or both. It is estimated that over four billion tons of by-products have
been stockpiled annually at quarry sites around the US country. The amount of
by-products will continue to grow due to production increases and environmental
discharge restrictions [Wood, 1995].
Previous studies indicate an improvement in the properties of both fresh
and hardened concrete when the MFA included a higher percentage of micro fines.
This material consists of dust-of-fracture and essentially is free from deleterious
material such as clay or shell. In other words, producers have spent both time and
resources to remove a portion of the fines that, if left in the sand, would improve
its quality. Additionally, it would reduce the amount of waste material that has to
be handled and disposed of in the sand producing process [J. Fowler, 1997].
Use of increased amounts of fines in concrete should be carefully
investigated with the objective of producing concrete with equal or better
performance, while using aggregates more efficiently to decrease aggregate cost.
MFA containing from ten to 20 percent of micro fines can have a beneficial effect
on the properties of portland cement concrete [Hudson 1997, Nichols 1982, and
Marek 1995]. These fines may be able to fill the void spaces between the coarser
particles and produce concrete having similar workability without significantly
increasing water requirement. This would result in an increase in density and
greater compressive strength.

5
1.3

Research Objectives
The main objectives of this research study to determine what practical

limits can be established for proportioning concrete with higher levels of crushed
fines are the following:
1) Determine the state-of-the-art regarding the use of higher crushed fines
contents (ten to twenty percent of total fine aggregate) in the production of
concrete for different applications in construction,
2) Determine the most relevant characteristics and composition of crushed
fines that govern and/or control their use in the production of concrete for
different applications in construction,
3) Develop a classification of crushed fines based on their suitability for use
in producing concrete for different applications in construction,
4) Develop guidelines for mix proportioning of concrete incorporating higher
fines contents, wherever compensation is needed from sand or rock due to
higher surface area contributed by crusher fines,
5) Determine the effect of higher amounts for several types of crusher fines
on concrete including fresh and hardened properties, durability and
serviceability,
6) Develop

modifications

to

existing

construction

specifications

incorporate the use of higher levels of fines in concrete construction,

to

6
7) Conduct a cost comparison between concrete containing fines at current
fines limits and concrete incorporating higher fines contents.

1.4

Scope of Research Program


To provide guidance in the conduct of the research a Project Advisory

Committee (PAC) was established including members of industry, federal and


state agencies, researchers and academia. The PAC was instrumental in ensuring
that the research work is conducted effectively, efficiently and in the best interest
of the sponsor.
This research study consisted of four main stages. The first stage was the
selection of aggregates that are used to this project.

There were over 110

aggregate sources available for this project representing 22 states and seven rock
types. The aggregates from sixteen sources were crushed and screened by Svedala
Barmac in Birmingham, Alabama. All sixteen source aggregates were used in this
research study.
The aggregate property-testing program was the second stage. The basic
aggregate characteristic tests were conducted to select the aggregates that are to be
tested for the next stage. Five tests were performed for 112 sands and seven rock
types. Based on the aggregate characterization tests, 50 sands were selected for
the mortar testing program.

7
The third stage was the mortar testing program to evaluate the use of MFA
in mortar. Two variables (the cement-sand ratio and the flow rate of mortar) were
considered to investigate the characteristics for each type of aggregate in mortar.
The fourth stage was the concrete testing program to investigate concrete
performance and the properties of concrete. This concrete evaluation included
mixes with fixed water-cement ratio and fixed slump.
After the third and fourth stages were finished, a statistical analysis was
conducted to evaluate the effect of the characteristics of fine aggregate on the
properties of mortar and concrete. Based on the analysis, the guidelines for using
higher amounts of micro fines in portland cement concrete were developed. In
addition, classification of manufactured fine aggregate based on suitability for use
in portland cement concrete was developed and a cost analysis was performed.

CHAPTER TWO: REVIEW OF STUDIES ON MANUFACTURED FINE


AGGREGATES

2.1

Introduction
A literature review of the characteristics and effects of manufactured fine

aggregates on concrete properties is presented in this chapter. Many studies have


recently been conducted on manufactured fine aggregate in portland cement
concrete. The topics of their research included basic characteristics of MFA,
effect of MFA on fresh concrete properties, and effect of MFA on hardened
concrete properties.

2.2

Definitions

2.2.1 Manufactured Fine Aggregates


Manufactured fine aggregate (MFA) is a process-controlled, crushed
aggregate produced from quarried stone by crushing or grinding and classification
to obtain a controlled gradation product that completely passes the 3/8-in. (9.5mm) sieve [NCSA, 1976]. Due to variations in the rock type and crushing process
the physical characteristics of the MFA, such as gradation, shape, dust-of-fracture
content and texture, can change significantly and influence the performance of
concrete. MFA is also referred to as stone sand, crusher sand, crushed fine
aggregate, specification sand or manufactured sand [NCSA, 1976].
8

9
2.2.2 Dust-of-Fracture
The dust-of-fracture is the by-product of the deliberate fracturing of rock
for aggregate production. As the shape of the MFA becomes more spherical, the
corners of the aggregates are removed creating the dust, which typically passes the
No. 200 sieve (75 m). The content of dust-offracture in MFA due to the
crushing process can be expected to exceed 10% without having a detrimental
effect on most concrete [Hudson, 1997]. Without deleterious materials such as
clay, shale, coal, lignite or other impurities the dust should be considered clean and
acceptable for use in concrete [NCSA, 1976].

2.3

Characteristics of Manufactured Fine Aggregates


The characterization of fine aggregates for concrete is important due to the

new performance requirements from increasingly technical placement methods.


Fine aggregate that prevents segregation is easy to finish and provides equal
hardened properties that are desirable for high volume applications.

2.3.1 Particle Shape


Particle shape, roundness, and sphericity are not usually determined for
natural sands. The particle shape is influenced by the physical properties of the
parent rock and by the method of production [McKeagney, 1984].

The

workability, flow, yield, air content, water requirement, bleeding and finishability

10
of concrete are all influenced by the particle shape of the fine aggregate in the
mortar [NCSA, 1976]. Crushed aggregates contain more angular particles with
rougher surface textures and flatter faces than natural sands that are more rounded
as a result of weathering experienced over time.
Researchers have become interested in quantifying particle shape as a way
to explain variations in mixing water requirements and compressive strength for
identically proportioned mixtures. In the 1960s, Wills [Wills, 1967] investigated
the effects of both fine and coarse aggregate on water demand in concrete. The
fine aggregate was found to have a more significant impact on water demand than
the coarse aggregate.
The relationship observed by Wills between 7-day compressive strength of
2-in. (50.8-mm) mortar cubes and the orifice flow rate of the fine aggregate is
shown in Figure 2.1. The fine aggregate with the highest flow rate was said to
possess the more flaky and elongated particle shape characteristics. The figure
shows that, as the flow rate increases, the 7-day compressive strength of the mortar
decreases.
Figure 2.2 shows that the same relationship between the average 28-day
compressive strength and the orifice flow rate. The variation in strength suggests
that the orifice flow rate cannot be used to predict the compressive strength, but
there is a general improvement in the compressive strength as the orifice flow rate
decreases.

11
7500
7250

50 M Pa

7000

48 M Pa

6750

46 M Pa

6500

44 M Pa

6250
42 M Pa

6000

40 M Pa

5750
5500

38 M Pa

5250

36 M Pa

5000
0.15

0.16

0.17
0.18
0.19
O rifice Flow Rate (sec/cm 3 )

0.2

0.21

Figure 2.1: Seven-day Compressive Strength vs. Orifice Flow Rate


[Wills, 1967]
6 5 00

44 M Pa

6 2 50
42 M Pa

6 0 00
5 7 50

40 M Pa

5 5 00

38 M Pa

5 2 50

36 M Pa

5 0 00

34 M Pa

4 7 50

32 M Pa

4 5 00
4 2 50
0 .1 5

30 M Pa
0 .1 6

0 .1 7
0 .1 8
0 .1 9
O rifice F lo w R a te (se c/c m 3 )

0 .2

0 .2 1

Figure 2.2: Twenty-eight-day Compressive Strength vs. Orifice Flow Rate


[Wills, 1967]

12
Roundness measures the relative sharpness or angularity of the edges and
corners of a particle [Neville, 1996]. Roundness can be defined numerically as the
ratio of the average radius of curvature of the corners and edges of the particle to
the radius of the maximum inscribed circle, but descriptive terms are more
commonly used [Popovics, 1992]. A classification used in the U.S. is as follows:
[Popovics, 1992].
Angular

: little evidence of wear on the particle surface

Subangular

: evidence of some wear, but faces untouched

Subrounded

: considerable wear, faces reduced in area

Rounded

: faces almost gone

Well rounded

: no original faces left

Sphericity is the property that measures, depends on, or varies with the
ratio of the surface area of the particle to its volume, the relative lengths of its
principal axes or those of the circumscribing rectangular prism, the relative settling
velocity, and the volume of the particle to that of the circumscribing sphere [Harr,
1977]. For instance, if two of the principal axes are much shorter than the third
axis, the particle is called elongated; if two of these axes are much longer than the
third one, the particle is called flat. Figure 2.3 provides two comparable charts for
the visual assessment of particle shape.

13

(a)

Figure 2.3: Visual Assessment of Particle Shape


[Powers, 1953; Krumbein, 1963]
(a) Derived from Measurements of Sphericity and Roundness
(b) Based upon Morphological Observations

14
Nichols [1982] reported that, as the angularity of the particles increased,
the voids content increased and water-cement ratios were greater than comparable
mixtures with less angular fine aggregate. As shown in Figure 2.4, the water
demand increases for concrete with a given slump as the particle shape index
increases. The water demand increases significantly when the shape index is
greater than 53 for both cement contents. The increase in water demand above the
53-shape index is attributed to flaky particles in the aggregate which require more
water to obtain the same slump.

0.8
4.2 sks./cu . yd.
0.75

5.0 sks./cu . yd.

0.7
0.65
0.6
0.55
0.5
48

49

50

51

52

53

54

55

N C S A P article S hap e In dex

Figure 2.4: Influence of Particle Shape Index on Water Demand


[Nichols, 1982]

56

15
2.3.2 Particle Surface Texture
The surface texture, also called surface roughness, of particles is the sum of
its minute surface features [Dolar-Mantuani, 1983]. It is an inherent and specific
property that depends on the texture, the structure, and the degree of weathering of
the source rock. The surface texture influences the workability, the quantity of
cement needed to produce satisfactory mortar mixtures, and the bond between the
particles and the cement paste in hardened mortar.

In most cases, some

improvement in bond is obtained as surface roughness increases [Barksdale,


1993].
After investigating the influence of surface texture of particles on the
workability and cement content of a concrete mixture, Mather [1966] stated that as
surface smoothness increases, the contact area with the cement paste decreases.
Hence, a highly polished particle will have less bond area with the matrix than will
a rough particle of the same volume. A smooth particle, however, will require a
thinner layer of paste to lubricate its movement with respect to other aggregate
particles.

It will, therefore, permit tighter packing for equal workability and,

hence, will require lower paste content than a rough particle of similar roundness
and sphericity.
Rhoades and Mielenz [1946] explained the complex interrelation among
the main textural features that influence the quality of the aggregate bond to the
cement paste. Although rugosity increases the bond to the cement paste, even

16
more important aspects of surface texture are the porosity, absorption and
permeability of the zone immediately underlying the surface. Penetration of the
aggregate by cement slurry is conducive to good bond, but the porosity implied by
very high penetrability may involve low tensile and shearing strength of the
aggregate, with the loss in strength of the concrete.
Other investigators have written that fine aggregates with very low
absorption generally develop lower strength bonds and produce less durable
mortars than those with slightly higher absorption. The interrelation between bond
and absorption may account in part for the poor correlation between the durability
of concrete and absorption, because the strength of bond increases as absorption
increases, whereas the durability of concrete tends to decrease as absorption
increases.

Thus, the absorption characteristics of aggregates alone cannot be

considered a reliable indication of bonding characteristics, for capillaries of


extremely small size may not permit penetration of the slurry into the aggregate
particles but may permit considerable penetration of water. From the standpoint of
durability and bond, penetrable voids of very small size are the least desirable
[Dolar-Mantuani, 1983].
The strength and permanence of the bond between the cement and
aggregate are functions not only of the surface texture, but also of the chemical
characteristics of the aggregate. The integrity of bond will be lost if chemical
reactions, such as that between high-alkali cement and reactive aggregates,

17
subsequently take place. On the other hand, some types of chemical surficial
interactions between the aggregate and the cement paste may be beneficial in
effecting a more intimate and stronger union. At present, however, there is no
good test method that can be used to evaluate the texture of manufactured sand
effectively.

2.3.3 Grading
The significance of the aggregate grading is that it influences directly many
important properties of fresh concrete, such as consistency and segregation, and to
a certain extent the properties of hardened concrete as well [Baker, 1973].
Generally, within the permitted standard limits, the gradation of fine aggregate has
a greater influence on the properties of concrete than that of coarse aggregate
[Hewlett, 1998]. At one end of the range, unusually coarse sand tends to produce
a harsh mix of low workability and with a greater liability to bleeding, segregation
of water during mixing and/or placing of the concrete. At the other end, unusually
fine sand can significantly increase the water demand of a concrete mix, because
of its much greater particle surface area, but it can improve cohesiveness.
In a study for the British Standards Institution, Pike [Pike, 1989] concluded
that: A controlled content of mineral flour other than clay minerals may be
tolerable and occasionally useful, but clay should be avoided; smectites are
particularly harmful and their content shall be severely restricted. Pike suggested

18
that fines limits perhaps could sometimes be relaxed, if the methylene blue
absorption test is available for harmful clay fines. Ramirez et al. [1990] have
similarly proposed controls for calcareous sands based upon both fines content and
methylene blue absorption.

However, the strength performance effects for a

particular methylene blue value depend upon the type of clay, and thus the test is
not definitive [Pike, 1992].
Marek [1995] investigated the effect of fine aggregate shape and grading
on properties of concrete. He stated that a fine aggregate grading specified by the
Portland Cement Association (PCA) is suitable for relatively spherical particles
but is not suitable for grading containing highly non-spherical particles. The test
results indicated the use of micro fines reduced the void content of the aggregate,
thereby lubricating the aggregate system without increasing the water requirement
of the mixture. He recommended a fines content in excess of 5 percent, and up to
and even exceeding 10 percent, be considered when the fines are dust of fracture
without clay or silt.
ASTM C33 includes fine aggregate specifications for concrete and is the
basis for many of the specifications in use today. Clelland [1980], writing in the
New Zealand Standards Bulletin, pointed out that some sands that comply with
fine aggregate specifications do not make good concrete due to poor physical
characteristics. At the same time, sands that do not fit into the grading envelope
have been used successfully in concrete.

19

100

North Carolina
Specification
Georgia
Secification

80

ASTM C33
Specification

60
Coarse
Gradations

40

Fine Gradations

20

0
3/8"

No. 4

No. 8

No. 16 No. 30 No. 50 No. 100 No. 200

US Standard Sieve Size

Figure 2.5: Comparison of Aggregate Grading Specifications


Shown in Figure 2.5 is a comparison of the ASTM C33, Georgia DOT and
North Carolina DOT specifications for fine aggregate grading. Figure 2.5 shows
that for the finest allowable grading, the ASTM C33 allows less of each size
fraction below the No. 8 (2.36-mm) sieve than the North Carolina and Georgia
specifications, resulting in a coarser blend of aggregate particles. The Georgia
specification allows the finest blend by requiring more of the smaller size fractions
than the other specifications.
The effect of changing the grading of MFA depends on the proportioning
of the concrete. Nichols [1982] and Kalcheff [1977] reported that changing the

20
grading of a given MFA had little or no effect on the water demand. However,
as the content of micro fines increased, the plastic concrete bled significantly
less than concrete conforming to ASTM C33.

The effect on the strength

properties varied with the cement content and not the changes in grading.
It was suggested by Nichols that the increase in micro fines lowered the
fineness modulus (FM) and the total quantity of fine aggregate needed to produce
a workable mixture. The reduction in fine aggregate should offset the increase in
water demand resulting from the presence of micro fines. In the end, the grading
of the fine aggregate does not affect the compressive strength, flexural strength
and freeze-thaw resistance as much as mixture proportioning.

2.3.4 Clay Content and Deleterious Components


Particles finer than 75m may be present in three different forms: clay, silt,
or dust. Clay is a natural mineral material having plastic properties and composed
of very fine particles; the clay mineral fraction of a soil is usually considered to be
the portion consisting of particles finer than 2 m. Clay minerals are essentially
hydrous aluminum silicates or occasionally hydrous magnesium silicates [ACI
116, 1985].
Grading of four aggregate dusts is shown in Figure 2.6. The dust has a
maximum size of 75 m. The graph shows dust grading from granite, quartzite,
limestone, and traprock source. The vertical dashed line indicates the 2-m size

21
fraction. The maximum allowable clay content in fine aggregate is 3 percent of
the fine aggregate by weight according to the ASTM C33 specification. If 20
percent of the minus 75-m-size fraction is smaller than 2 m and 15 percent of
the fine aggregate is minus 75-m material, the clay content of the total fine
aggregate will be 3 percent.
100
90
80
70
60
50

Limestone

40
30

Granite

20

Traprock

10

Quartzite

Sieve Size (mm)

Figure 2.6: Particle Size Distribution of Crushed Aggregate Fines


[Wood, 1995]
As shown in Figure 2.6, the dust fraction of the four MFA types typically
contains less than 20 percent finer than 2 m. The rock type with the highest
content of clay-size particles is quartzite, with approximately 40 percent of the
dust finer than 2 m. Qualifying the clay content by size penalizes fine aggregate

22
since the particles may be the same size as clay, but not deleterious to concrete
[Wood, 1995].
Ramirez [1990] investigated the problem of calcareous and clay fines in
fine aggregate and their influence on concrete properties. The methylene blue test
(P 18-592 AFNOR) was used to test 21 fine aggregates with varying contents of
clay and micro fines. The methylene blue test is widely accepted in Europe as an
effective method to quantify the clay content both in natural sand and
manufactured sand. The procedure for this test is as follows [Maldonado, 1996]: a
small amount of water containing the sample material and titrated methylene blue
is removed via a glass rod and dropped onto filter paper. If a blue ring is observed
on the paper, the sample is stirred and tested again without adding additional
methylene blue solution. The final amount of methylene blue is recorded.
In Ramirez study clay content (i) is the percent clay/sand where the clay is
finer than 40-m and composed of 90/10 illyte/kaolinite. The clay contents used
for the test program are shown in Table 2.1. The clay was added to the fine
aggregate in order to determine the sensitivity of the methylene blue test
procedure.
Ramirez reported compressive strength when the MB value is put in terms
of grams per 100-g sand and grams per 100-g fines. This relationship does not
reflect the effect of increasing the content of micro fines in fine aggregate on the
compressive strength as shown Figure 2.7 and 2.8.

23
Table 2.1: Mixture Proportioning used in Ramirez Tests [Ramirez, 1990]
Total fines in

Clay/sand, i(%)

sand, f (%)

25

20

10

3.5

25

3600 psi
i=0

24

i=1

23

i=2

22

3400 psi
3200 psi

i=4

21

3000 psi

20
2800 psi

19
18

2600 psi

17

2400 psi

16

2000 psi

15
0.000

0.020

0.040

0.060

0.080

0.100

0.120

0.140

Methylene Blue (g/100g sand)

Figure 2.7: Compressive Strength vs. Methylene Blue Value (g/100-g sand)
[Ramirez, 1990]

24

25

3600 psi

24

i
i
i
i

23
22

=
=
=
=

21

0
1
2
4

3400 psi
3200 psi
3000 psi

20
2800 psi

19
18

2600 psi

17

2400 psi

16
2200 psi

15
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.1

1.2

Methylene blue (g/100g fines)

Figure 2.8: Compressive Strength vs. Methylene Blue Value (g/100-g fines)
[Ramirez, 1990]
2.4

Effect of Manufactured Fine Aggregates on Fresh Concrete Properties

2.4.1 Mixing Water Requirement


Many researchers have noted increased water demand as the dust of
fracture percentage is increased in the fine aggregate in concrete. The observation
seems to go against the earlier claim that the minus 75-m-size fraction helps to
lubricate the plastic concrete. Celik and Marar [1996] noted that without adjusting
the proportions of coarse and fine aggregate, the specific surface of the aggregate
particles increased as the dust of fracture percentage increased. They attributed the
increase in water demand to the increased specific surface.

25
Ahmed and El-Kourd [1989] tested concrete with constant slump and
concrete with a constant w/c. The concrete made to have a constant slump of 4.0
0.5-in. (100 15-mm) required more water as the content of dust was increased.
Shown in Figure 2.9, the required w/c to maintain a constant slump was greater for
the natural sand than for MFA with the same dust content. Concrete batched with
a constant w/c had decreasing slump as the dust content increased.
0.8
Natural Sand
Limestone MFA
0.75

0.7

0.65

0.6
0

10
Dust Content (Percent of FA)

15

20

Figure 2.9: Influence of Aggregate Type and Dust Content on W/C


[Ahmed, 1989]
2.4.2 Air Content
Celik and Marar [1996] batched concrete specimens with constant w/c, fine
aggregate, coarse aggregate and cement contents. The only variation was the
percentage of the fine aggregate replaced with dust of fracture.

Air content

26
measurements were taken on the various mixtures with the results shown below in
Figure 2.10.
3
2 .8
2 .6
2 .4
2 .2
2
1 .8
1 .6
1 .4
1 .2
0

10

15

20

25

30

D u st C o n te n t (% )

Figure 2.10: Effect of Increasing Dust of Fracture Content on Air Content in


Concrete [Celik, Marar 1996]

2.5

Effect of Manufactured Fine Aggregates on Hardened Concrete Properties

2.5.1 Compressive Strength


The compressive strength of concrete containing higher contents of dust of
fracture has been shown comparable to concrete conforming to ASTM C33 in
many studies. Dukatz and Marek [1985] reported that concrete made with MFA
containing up to 7 percent dust of fracture obtains compressive strength equal to or
better than concrete made with natural sand with the same w/c. Malhotra and
Carette [1985] reported that for concrete with a w/c of 0.70 and increasing

27
contents of micro fines, the compressive strength was as good as or better than the
control concrete at all ages.

Figure 2.11 shows the relationship between the

amount of sand replaced with limestone dust, age of the concrete and the
compressive strength of the concrete.

For concrete with a 0.70 w/c, the

compressive strength increased with the amount of dust and the age for all
specimens. Concrete with a 0.53 w/c showed constant strength for all levels of
dust replacement and increasing strength with age. The concrete made with a 0.40
w/c had a decrease in strength when the dust content in the MFA was 10 percent as
compared to the control and the same strength for concrete made with MFA
containing 20 percent dust.
7000
45 MPa
6000

40 MPa
35 MPa

5000

30 MPa

4000

25 MPa

3000

20 MPa

2000

15 MPa
0

10

15

20

Dust Content
7 day, w/c = 0.70
7 day, w/c = 0.53
7 day, w/c = 0.40

28 day, w/c = 0.70


28 day, w/c = 0.53
28 day, w/c = 0.40

90 day, w/c = 0.70


90 day, w/c = 0.53
90 day, w/c = 0.40

Figure 2.11: Relationship between Age, W/C, Dust Content and Strength
[Celik, 1996]

28
Shown in Figure 2.12, Ahmed and El-Kourd [1989] reported that, for
concrete of constant slump, the compressive strength decreased linearly with
increasing percentages of dust of fracture in the fine aggregate.

In contrast,

concrete made with a constant w/c of 0.70 showed increasing compressive strength
as the dust of fracture content in the fine aggregate was increased. The difference
is a result of increased water demand in the mixtures requiring a constant slump.

5750
Constant Slump
Constant w/c

5500

39 MPa

37 MPa
5250
35 MPa

5000
4750

33 MPa

4500

31 MPa

4250
0

10

15

20

Dust of Fracture Content (% Fine Aggregate)

Figure 2.12: Comparison of the 28-day Compressive Strength of Concrete


made with MFA Proportioned to a Constant Slump and a Constant W/C
[Ahmed, El-Kourd 1989]

29
Ahmed [1989] reports that the concrete made with MFA was stronger at all
ages than concrete made with blended natural sand. Figure 2.13 shows the 28-day
compressive strength of concrete made with limestone MFA and natural sand as
reported by Ahmed. Both concrete mixtures had a constant w/c. The strength
increased for every level of replacement except the 3 percent mixture as compared
to the mixture with no dust included.
5100

35 MPa

Natural Sand
Crushed Limestone

5000

34 MPa

4900
4800

33 MPa

4700
32 MPa

4600
4500

31 MPa

4400

30 MPa

4300
29 MPa

4200
0

10

15

20

Percent Dust Replacement

Figure 2.13: Influence of Fine Aggregate Type on 28-day Compressive


Strength [Ahmed, 1989]
The natural sand concrete showed a decrease in compressive strength for
the 3, 5 and 15 percent replacement levels as compared to 0 percent replacement

30
mixture. The concrete containing 7 and 10 percent of natural sand dust showed an
increase in compressive strength at 28 days.
The decrease in compressive strength in the natural sand concrete is
anticipated by current specifications, which include the limitations on the micro
fines.

The increase in compressive strength by the MFA concrete is not

anticipated by the current specifications, which would prevent the benefits of the
material from being realized.
Bonavetti and Irassar [1976] investigated the effect of stone dust content in
mortar. The tests were conducted with limestone, quartz and granite dust of
fracture combined with natural sand. The crushed fine aggregate was not used due
to difference in the grain size distribution, shape and texture of the particles.
Mortar cubes were batched to have a constant flow with a 1:3 cement-to-sand ratio
combined with 5, 10, 15 and 20 percent of the sand replaced with dust of fracture.
The cubes tested containing quartz dust had increased compressive strength
at all percentages of sand replacement and ages. The cubes containing granite had
mixed results. The cubes with 5 percent replacement had greater compressive
strength at all ages as compared to the control mortar cubes. The cubes containing
10 percent dust replacement had below average strength at seven days, but above
average strength at 28, 90 and 180 days. The specimens containing 15 and 20
percent dust replacement were weaker in compression at all ages than the control
mortar specimens.

31
The limestone specimens had rapid early strength gain, which over time
fell into line with the strength of the control group. The mortar specimens with 5,
10 and 15 percent replacement were stronger than the control mortar while the
mortar with 20 percent replacement had a compressive strength equal to 95 percent
of the control at 180 days.

2.5.2 Flexural Strength


Celik and Marar [1996] reported that as the dust of fracture content was
increased to 10 percent the 28-day flexural strength increased 10 percent as
compared to the control specimens with no dust of fracture. As the dust of fracture
content was increased to 30 percent, the 28-day flexural strength decreased 10
percent as compared to the control concrete mixture.
Malhotra and Carette [1985] tested the flexural strength of concrete
containing varying percentages of dust with water-cement ratios of 0.70, 0.53 and
0.40 at 14 and 70 5 days. The flexural strength for each w/c was as good as or
better than the flexural strength of the control mixture for the corresponding w/c as
shown in Figure 2.14. The increase in dust content did not negatively effect the
flexural strength. The most significant influence on the flexural strength is the
w/c.

32

1100

w/c=0.70

w/c=0.53

w/c=0.40
7.0 MPa

1000
900

4.5 MPa

800
4.0 MPa

700
600

3.5 MPa

500
0

10

15

20

Percent Dust Content

Figure 2.14: Percent Dust Content vs. Flexural Strength [Malhotra, 1985]
Bonavetti and Irassar (1994) tested the flexural strength of mortars
containing up to 20 percent dust of fracture. For all levels of replacement, the
concrete showed increased flexural strength at all ages as compared to the control
concrete.

2.5.3 Shrinkage
The shrinkage of concrete typically increased with the percentage of the
minus 75-m-size fraction. Ahmed and El-Kourd [1989] monitored the drying
shrinkage of seven concrete mixtures with different percentages of dust of fracture.

33
Concrete prisms measuring 2-in.2-in.11.25-in. (50-mm50-mm285-mm) were
cast for the test program. The specimens were water cured for 3 days and then air
cured at 73.4 3.6F (23 2C) and 45 5 percent relative humidity.
Measurements were taken at 7, 28, 56, 100 and 330 days as shown Figure 2.15.
0.07
0.0675
0.065
0.0625
0.06
0.0575
0.055
0.0525
0.05
0

10

15

20

Percent of MFA Replaced by Dust

Figure 2.15: Effect of Increasing Dust of Fracture Content


on 330 day Drying Shrinkage [Ahmed, El-Kourd 1989]

Malhotra and Carette [1985] used 3-in.4-in.16-in. (76-mm102mm406-mm) prisms for their shrinkage monitoring program. The specimens
were tested according to ASTM C157 procedures. The specimens were cured in
water at 73 3F (23 1.7C) for 7 days and then exposed to air-drying at 73

34
3F (23 1.7C) and 50 percent relative humidity. The drying shrinkage strains
were monitored for a period of 217 days.
As shown in Figure 2.16, the concrete shrinkage increased with increasing
dust of fracture content.

Shrinkage effects were more pronounced for lean

concrete containing more than 10 percent dust-of-fracture. Factors attributed to


influencing the test results are accelerated hydration, carboaluminate formation
and large superplasticizer dosages in the specimens incorporating 15 and 20
percent limestone dust of fracture.
700
650
600
550
500
450

w/c=0.70
w/c=0.53

400

w/c=0.40

350
300
0

10

15

20

Percent Replacement of Sand by Limestone Dust

Figure 2.16: Influence of Dust of Fracture Content and W/C [Ahmed, 1989]

35
2.5.4 Permeability
Permeability of concrete can be reduced by using MFA with high
percentages of dust of fracture. Inclusion of MFA with good particle shape and
gradation allows for efficient aggregate packing.

The particle arrangement

decreases permeability by blocking capillary passages formed during the hydration


of the cement [Hudson, 1997].
Bonavetti and Irassar [1994] tested mortars with varying w/c and dust
contents. As the w/c increased, the permeability increased, as would be expected.
For every w/c, as the dust content increased the permeability of the mortar
decreased as compared to the control mixture

2.5.5 Impact Resistance


The impact resistance of concrete is typically thought to be directly related
to the compressive strength of the concrete. Celik and Marar [1996] reported that
the impact resistance of the concrete was greatest with the addition of 5 percent
dust of fracture. Impact resistance was improved with 10 percent dust content but
decreased below the impact resistance of the control concrete for the mixtures
tested containing more than 10 percent dust of fracture. Impact resistance did not
increase with increasing compressive strength as would normally be expected.
The greatest compressive strength was observed with the concrete containing 10

36
percent dust of fracture while the greatest impact resistance was observed at 5
percent dust replacement.

2.5.6 Absorption
According to Celik and Marar [1996] the absorption percentage of concrete
containing up to 15 percent dust of fracture decreased. The decrease in absorption
was attributed to the dust acting as filler in the concrete. However, concrete
containing more than 15 percent dust of fracture caused an increase in absorption,
which they related to the observed decrease in compressive strength. Since the w/c
was constant for the test batches, the increase in absorption may have been related
to the pore structure of the concrete. Celik stated that the concrete made with fine
aggregate containing more than 15 percent dust did not have enough paste to cover
all of the aggregate particles in the mixture.

2.5.7 Creep
Malhotra and Carette [1985] tested the creep of concrete with a 0.53 w/c
and dust of fracture contents ranging from zero to 20 percent of the weight of the
fine aggregate. The creep of the concrete containing 5 percent dust was less than
the control while the concrete containing 10, 15 and 20 percent dust was 22 to 26
percent greater than the control after 200 days of loading. The increase in creep

37
was attributed to the formation of carboaluminates and the increased rate of
hydration.

2.5.8 Resistance to Freezing and Thawing


Kalcheff [1977] tested concrete specimens according to ASTM C666
Procedure B. There were no conclusions drawn as to the effect of fine aggregate
on the freezing and thawing resistance of the concrete. A constant air content of
5.5 0.5 percent was used in all of the concrete mixtures. The constant air content
was achieved using a vinsol resin solution with the content varying due to the
quantity of cement, the quantity of fine aggregate and the gradation of the fine
aggregate including the dust of fracture content.

CHAPTER THREE: AGGREGATE SELECTION

3.1

Introduction
Some properties of the types of rocks in the continental United States and

the method of aggregate selection are presented in this chapter. The aggregates
chosen for testing in this study represent a broad cross section of the crushed
aggregates produced in the U.S. Limestone, dolomite, granite, and traprock are
crushed aggregates most commonly used in concrete and account for 94% of the
crushed stone produced in the U.S. in 1996 [Tepordei, 1996]. Sandstone, quartzite
and crushed river gravel sources were also identified due to their potential for use
in producing MFA. Each potential source was evaluated by its rock type and
geographic location. There were over 110 aggregate sources available to ICAR for
this project representing 22 states and seven rock types.

3.2

Properties of the Types of Rocks


Waddell [1993] reported some physical and engineering properties of the

types of rocks according to the class of rocks. Table 3.1 and 3.2 show the physical
and engineering properties of the types of rocks used in this study. As shown in
the tables, the absorption (%) of sedimentary rocks is relatively higher than that of
other rocks.

38

39
Table 3.1: Average Values for Physical Properties of the Types of Rocks
Investigated in This Study [Waddell, 1993]
Loss by abrasion, %

Bulk specific
gravity

Absorption, %

Igneous
Granite
Basalt
Diabase

2.65
2.86
2.96

0.3
0.5
0.3

38
14
18

Sedimentary
Limestone
Dolomite
Sandstone

2.66
2.70
2.54

0.9
1.1
1.8

26
25
38

Metamorphic
Quartzite

2.69

0.3

28

Type of rock

Los Angeles

After immersion in water at atmospheric temperature and pressure.

ASTM C535.

Table 3.2: Summary of Engineering Properties of Rocks [Waddell, 1993]


Type of rock

Mechanical Durastrength
bility

Presence of
Crushed
Chemical
Surface
undesirable
shape
stability characteristics
impurities

Igneous
Granite
Basalt, Diabase

Good
Good

Good
Good

Good
Good

Good
Good

Possible
Seldom

Good
Fair

Sedimentary
Limestone,
Dolomite
Sandstone

Good
Good
Fair

Fair
Fair
Fair

Good
Good
Good

Good
Good
Good

Possible
Possible
Seldom

Good
Good
Good

Metamorphic
Quartzite

Good

Good

Good

Good

Seldom

Fair

40
3.3

Manufactured Fine Aggregate Selection


The U.S. Geological Survey (USGS) reported the amount of MFA, or stone

sand, sold or used by producers in the U.S. [Tepordei, 1996]. Limestone and
granite are the dominant aggregates used for MFA production accounting for 86%
of the reported totals. Traprock, dolomite, sandstone and quartzite make up the
other 14% with less than 1% of aggregates classified as miscellaneous by the
USGS also being used as MFA for concrete.

3.3.1 Limestone
Crushed limestone is produced in every state in the continental U.S. except
North Dakota, Louisiana and Delaware [Tepordei, 1996]. The USGS reports 8%
of crushed limestone is used in concrete and 2% of crushed limestone is used as
MFA. Figure 3.1 shows by dots the location of the limestone sources used for this
study with the shaded states indicating the top five producing states of crushed
limestone in the U.S.
The top five producing states account for 39% of crushed limestone
production in the U.S. Of the other states represented, Alabama, Tennessee and
Pennsylvania are among the top ten producers of crushed limestone.

In the

reporting of crushed limestone production some states did not differentiate


between limestone and dolomite production that might affect the reported
production levels in each state.

41

Figure 3.1: Crushed Limestone Sources for ICAR Research

Table 3.3: Location of Limestone Sources and Percentage of Total US Output


Number of Sources

Percentage of Total

evaluated in ICAR 102

U.S. Crushed

Study

Limestone Production

Texas

Florida

Pennsylvania

Tennessee

Alabama

California

Virginia

Nevada

0.2

Total

18

39

State Represented
in ICAR 102 Study

42
3.3.2 Dolomite
Dolomite production was reported in 25 states [Tepordei, 1996].

The

reported production volume of dolomite does not reflect the true amount of
dolomite produced due to some states not differentiating between limestone and
dolomite.

According to USGS data, 10% of crushed dolomite was used in

concrete and 1% was used as MFA. Figure 3.2 shows the location of the dolomite
sources used for this study.
The top five producing states are shaded representing 56% of U.S.
production. The production levels for Tennessee and Alabama were withheld to
avoid disclosing company proprietary data. The production in California and
Oklahoma accounts for 4% of U.S. production.

43
Table 3.4: Location of Dolomite Sources and Percentage of Total US Output
State

Number of

Percentage of

Represented

Sources

Total U.S.

in ICAR 102

evaluated in ICAR

Crushed Dolomite

Study

102 Study

Production

Ohio

18

California

0.4

Oklahoma

Tennessee

Withheld

Alabama

Withheld

Illinois

10

Iowa

N/A

Total

>32

3.3.3 Granite
Crushed granite was produced in 37 states in 1996 [Tepordei, 1996]. The
USGS reports that 7% of crushed granite produced was used in concrete and 2% of
crushed granite was used as MFA. Figure 3.3 indicates the location of twelve
sources from seven states identified for classification and testing in this study.
The top five producing states, which are shaded, account for 73% of US
production. The top granite producing states are represented by ten of twelve
granite sources. Table 3.5 indicates the number of sources in each state and the
percentage of US production that each state represents. The granite production
volume in Minnesota was withheld for proprietary reasons, but is less than 5% of
US production.

44

Figure 3.3: Crushed Granite Sources for ICAR Research

Table 3.5: Location of Granite Sources and Percentage of Total US Output


State

Number of

Percentage of

Represented in

Sources

Total U.S.

ICAR 102 Study

evaluated in ICAR

Crushed Granite

102 Study

Production

Georgia

26

North Carolina

21

South Carolina

California

Minnesota

Withheld

Virginia

12

Connecticut

N/A

Wyoming

N/A

Total

14

>71

45
3.3.4 Traprock
The USGS reported crushed traprock production in 27 states [Tepordei,
1996]. Seven percent of the traprock produced was used in concrete with 1% of
the crushed traprock used as MFA. Figure 3.4 shows by dots the location of the
sources of crushed traprock used in this study.
The top five producing states, shaded in the map below, crushed 64% of
the traprock processed in the US [Tepordei, 1996].

The volume of traprock

crushed in New York was withheld for proprietary reasons, but accounted for less
than 8% of the total US production volume.

Figure 3.4: Crushed Traprock Sources for ICAR Research

46

Table 3.6: Location of Traprock Sources and Percentage of Total US Output


State

Number of

Percentage of

Represented in

Sources

Total U.S.

ICAR 102 Study

evaluated in ICAR

Crushed Granite

102 Study

Production

Washington

26

California

21

New York

Virginia

12

Total

12

>71

3.3.5 Sandstone, Quartzite and Crushed River Gravel


The output of sandstone and quartzite was combined in the USGS Crushed
Stone report in 1996 while production data for crushed river gravel are not
specified.
production.

The total quartzite production in the US is 35% of the sandstone


Three percent of crushed sandstone is used in concrete and two

percent is used as MFA, while 6% of crushed quartzite is used in concrete and 1%


is used as MFA. Figure 3.5 indicates by dots the location of the sandstone,
quartzite and crushed river gravel sources used in the ICAR research.
The top five producing states of sandstone and quartzite, shaded in Figure
3.5, account for 53% of US output. The quartzite sources are located in South
Dakota and Montana, the sandstone sources are in Pennsylvania and Arkansas and

47
the crushed river gravel source is in Michigan. There are no specific data on the
production of crushed river gravel in the USGS Survey.

Figure 3.5: Crushed Sandstone, Quartzite and Gravel Sources


for ICAR Research

3.4

Summary
After aggregate producers were contacted, the sources were delivered to be

crushed and screened in a Svedala-Barmac crusher in Birmingham, Alabama.


Table 3.7 indicates the selected aggregates and location for this study.
All sixteen source aggregates were used in this research study. Since the
production of limestone and granite is approximately 86% of all type of MFA in
U.S., three limestone aggregates and five granite aggregates were selected for this

48
study. The number of sands indicates the number of different screening sands that
were crushed by different crushing speed in the Svedala-Barmac crusher.

Table 3.7: Selected Aggregates and Location for ICAR Research


Quarry
ID No.

Rock
Type

Location
(State)

No. of
sands

51

Dolomite

Oklahoma

54

Limestone

55

Sandstone

56

Dolomite

Tennessee

57

Limestone

Virginia

62

Granite

Minnesota

63

Quartzite

South Dakota

65

Diabase

66

Granite

67

Granite

68

Granite

69

Basalt

70

Granite

Virginia

71

Limestone

Missouri

72

Dolomite

Iowa

73

Sandstone

Arkansas

Pennsylvania

Virginia
Wyoming
Connecticut

8
7

8
8
8
8
8

After being crushed and screened by the Svedala-Barmac crusher in


Birmingham, Alabama, approximately 800 boxes and 120 drums of crushed MFA
were shipped to Austin, Texas for the testing.

49

Figure 3.6: Rock Types and Corresponding Locations Used in This Study

Granite (GR) from Virginia (VA), Minnesota (MN),


Connecticut (CT), Wyoming (WY)

Basalt (BA) from Connecticut (CT)

Diabase (DI) from Virginia (VA)

Sandstone (SS) from Pennsylvania (PA), Arkansas (AR)

Limestone (LS) from Pennsylvania (PA), Virginia (VA),


Missouri (MO)

Dolomite (DO) from Oklahoma (OK), Tennessee (TN), Iowa


(IA)

Quartzite (QZ) from South Dakota (SD)

CHAPTER FOUR: AGGREGATE CHRACTERIZATION PROGRAM

4.1

Introduction
The physical property tests of MFA are presented in this chapter. Since

the aggregates for the mortar and concrete tests were selected based on the
aggregate properties, this testing program was very crucial.
The aggregate property tests consisted of specific gravity and absorption
tests, wet sieve analysis, uncompacted void content test, hydrometer particle size
analysis, chemical analysis using ICP technology, laser diffraction particle size
analysis, and methylene blue test.

4.2

Materials
Table 4.1 shows sixteen source aggregates (seven types of rock) that were

tested for aggregate characterization tests. A total of 112 sands were tested. After
the aggregates were crushed in a Svedala-Barmac crusher, they were screened as
shown in Table 4.2. The control variables were aggregate size (two or three
different sizes from each aggregate producer) and crushing speed (as received [no
crushing], 35, and 68 rpm). Aggregates from quarries 51, 55, 56, 57, 72 and 73
were crushed to produce a different number of sands as shown in Table 4.1.

50

51

Table 4.1: Tested Aggregates and Location in This Study


Quarry
ID No.

Rock
Type

Location
(State)

No. of
sands

51

Dolomite

Oklahoma

54

Limestone

55

Sandstone

56

Dolomite

Tennessee

57

Limestone

Virginia

62

Granite

Minnesota

63

Quartzite

South Dakota

65

Diabase

66

Granite

67

Granite

68

Granite

69

Basalt

70

Granite

Virginia

71

Limestone

Missouri

72

Dolomite

Iowa

73

Sandstone

Arkansas

Pennsylvania

Virginia
Wyoming
Connecticut

8
7

8
8
8
8
8

52
Table 4.2 shows all samples used in the aggregate characterization tests.
According to the screened sands, identification number (ID No.) was determined
for each sample to differentiate all samples. In the table, location is the state in
which the quarry is located and size is the maximum-minimum size of the
aggregate as delivered by the producer. Speed is the impact speed of crusher
head (00 means no crushing), cycle is the aggregate screening process
(product is final material after crushing and as-received is the material
received from the producer with no further crushing).
One type of sand usually consisted of eight crushed sands (two as-received
sands and six crushed products) except Oklahoma dolomite, Pennsylvania
sandstone, Tennessee dolomite, Virginia limestone, Iowa dolomite and Arkansas
Sandstone.

53
Table 4.2: Manufactured Sands Used in This Study
ID No.
PA/LS/13-02/68
PA/LS/13-02/36
PA/LS/06-00/00
PA/LS/06-00/65
PA/LS/06-00/36
PA/LS/05-00/36
PA/LS/05-00/65
PA/LS/05-00/00

Location

Type

Pennsylvania Limestone
(PA)
(LS)

VA/70/GT/19-02/35
VA/70/GT/19-02/68
VA/70/GT/06-00/68
Virginia
Granite
VA/70/GT/06-00/35
(VA)
(GT)
VA/70/GT/05-00/68
VA/70/GT/06-00/00
VA/70/GT/05-00/35
VA/70/GT/05-00/00
SD/QZ/19-02/68
SD/QZ/19-02/36
SD/QZ/06-00/00
South Dakota Quartzite
SD/QZ/06-00/65
(SD)
(QZ)
SD/QZ/06-00/36
SD/QZ/05-00/00
SD/QZ/05-00/65
SD/QZ/05-00/36
VA/DI/06-00/00
VA/DI/13-02/68
VA/DI/13-02/36
Virginia
Diabase
VA/DI/06-00/65
(VA)
(DI)
VA/DI/06-00/36
VA/DI/05-00/00
VA/DI/05-00/65
VA/DI/05-00/36

Size, mm Speed, m/s

Cycle

13-02
13-02
06-00
06-00
06-00
05-00
05-00
05-00

68
36
00
65
36
36
65
00

product
product
as-received
product
product
product
product
as-received

19-02
19-02
06-00
06-00
05-00
06-00
05-00
05-00
19-02
19-02
06-00
06-00
06-00
05-00
05-00
05-00
06-00
13-02
13-02
06-00
06-00
05-00
05-00
05-00

35
68
68
35
68
00
35
00
68
36
00
65
36
00
65
36
00
68
36
65
36
00
65
36

product
product
product
product
product
as-received
product
as-received
product
product
as-received
product
product
as-received
product
product
as-received
product
product
product
product
as-received
product
product

54
Table 4.2: Manufactured Sands Used in This Study (Continued)
ID No.
OK/DO/19-02/68
OK/DO/19-02/36
OK/DO/06-00/00
OK/DO/06-00/65
OK/DO/06-00/36
VA/LS/05-00/00
VA/LS/05-00/65
VA/LS/05-00/36

Location

Type

Oklahoma
(OK)

Dolomite
(DO)

Virginia
(VA)

Limestone
(LS)

CT/BA/05-00/36
CT/BA/05-00/65
CT/BA/05-00/00
CT/BA/06-00/36 Connecticut
Basalt
(CT)
(BA)
CT/BA/06-00/65
CT/BA/06-00/00
CT/BA/19-02/36
CT/BA/19-02/68
PA/SS/09-00/68
PA/SS/09-00/35
PA/SS/06-00/68 Pennsylvania Sandstone
PA/SS/06-00/35
(PA)
(SS)
PA/SS/05-00/68
PA/SS/05-00/35
PA/SS/05-00/00
MO/LS/19-02/68
MO/LS/19-02/35
MO/LS/06-00/68
MO/LS/06-00/00
Missouri
Limestone
(MO)
(LS)
MO/LS/06-00/35
MO/LS/05-00/00
MO/LS/05-00/68
MO/LS/05-00/35

Size, mm Speed, m/s

Cycle

19-02
19-02
06-00
06-00
06-00
05-00
05-00
05-00

68
36
00
65
36
00
65
36

product
product
as-received
product
product
as-received
product
product

05-00
05-00
05-00
06-00
06-00
06-00
19-02
19-02
09-00
09-00
06-00
06-00
05-00
05-00
05-00
19-02
19-02
06-00
06-00
06-00
05-00
05-00
05-00

36
65
00
36
65
00
36
68
68
35
68
35
68
35
00
68
35
68
00
35
00
68
35

product
product
as-received
product
product
as-received
product
product
product
product
product
product
product
product
as-received
product
product
product
as-received
product
as-received
product
product

55
Table 4.2: Manufactured Sands Used in This Study (Continued)
ID No.
VA/66/GT/13-02/35
VA/66/GT/13-02/68
VA/66/GT/06-00/68
VA/66/GT/06-00/35
VA/66/GT/05-00/68
VA/66/GT/06-00/00
VA/66/GT/05-00/35
VA/66/GT/05-00/00

Location

Virginia
(VA)

MN/GT/19-02/68
MN/GT/19-02/35
MN/GT/06-00/68
MN/GT/06-00/35 Minnesota
(MN)
MN/GT/05-00/68
MN/GT/05-00/00
MN/GT/05-00/35
MN/GT/06-00/00
CT/GT/19-02/65
CT/GT/19-02/36
CT/GT/06-00/65
CT/GT/06-00/36 Connecticut
(CT)
CT/GT/05-00/65
CT/GT/06-00/00
CT/GT/05-00/36
CT/GT/05-00/00
WY/GT/19-02/65
WY/GT/19-02/36
WY/GT/06-00/65
WY/GT/06-00/36
Wyoming
(WY)
WY/GT/05-00/65
WY/GT/06-00/00
WY/GT/05-00/36
WY/GT/05-00/00

Type

Granite

Granite

Granite

Granite

Size, mm Speed, m/s

Cycle

13-02
13-02
06-00
06-00
05-00
06-00
05-00
05-00

35
68
68
35
68
00
35
00

product
product
product
product
product
as-received
product
as-recieved

19-02
19-02
06-00
06-00
05-00
05-00
05-00
06-00
19-02
19-02
06-00
06-00
05-00
06-00
05-00
05-00
19-02
19-02
06-00
06-00
05-00
06-00
05-00
05-00

68
35
68
35
68
00
35
00
65
36
65
36
65
00
36
00
65
36
65
36
65
00
36
00

product
product
product
product
product
as-received
product
as-recieved
product
product
product
product
product
as-received
product
as-recieved
product
product
product
product
product
as-received
product
as-recieved

56
Table 4.2: Manufactured Sands Used in This Study (Continued)
ID No.
TN/DO/13-02/68
TN/DO/13-02/36
TN/DO/06-00/00
TN/DO/06-00/65
TN/DO/06-00/36
IA/DO/25-02/68
IA/DO/25-02/45
IA/DO/06-06/35
IA/DO/06-06/68
IA/DO/05-00/00
IA/DO/05-00/68
IA/DO/05-00/35
AR/SS/25-02/68
AR/SS/25-02/45
AR/SS/06-00/68
AR/SS/06-00/00
AR/SS/06-00/35

4.3

Location

Tennessee
(TN)

Type

Dolomite

Iowa
(IA)

Dolomite

Arkansas
(AR)

Sandstone

Size, mm Speed, m/s

Cycle

13-02
13-02
06-00
06-00
06-00

68
36
00
65
36

product
product
as-received
product
product

25-02
25-02
06-06
06-06
05-00
05-00
05-00
25-02
25-02
06-00
06-00
06-00

68
45
35
68
00
68
35
68
45
68
00
35

product
product
product
product
as-received
product
product
product
product
product
as-received
product

Testing Procedures for Aggregate Characterization

4.3.1 Specific Gravity and Absorption Capacity Test


The specific gravity and absorption capacity of each aggregate are
important because they affect the other characterization tests and are required for
mortar and mixture proportioning of a concrete batch. The testing was conducted
in accordance with ASTM C 128-97, Standard Test Method for Specific Gravity
and Absorption of Fine Aggregate. The bulk, SSD (surface-saturated dry) and

57
apparent specific gravity and absorption capacity of the aggregate were determined
by this test.
The bulk specific gravity of the aggregate is used to calculate the volume
occupied by the aggregate according to its weight. The absorption capacity of the
fine aggregate is an indication of the weight of water that can be absorbed into its
pores as the aggregate is in SSD condition.

4.3.2 Sieve Analysis by Washing


The gradation, fineness modulus, and micro fines contents were determined
by washing and sieving each aggregate sample.

The material was washed

according to ASTM C 117-95, Standard Test Method for Materials Finer than
No. 200 (75 m) Sieve in Mineral Aggregates by Washing. The gradation of the
washed sample was determined in accordance with ASTM C 136-96a, Standard
Test Method for Sieve Analysis of Fine and Coarse Aggregates.
In the study the aggregates were tested in the as-received condition,
meaning the aggregates were not recombined to achieve a specific gradation.
Based on the results of the sieve analysis each aggregate was classified according
to the percentage of material passing the No. 200 (75 m) sieve.

58
4.3.3 Uncompacted Void Content Test
The particle shape of fine aggregate is important since it influences the
workability, finishability and water demand of fresh concrete which, in turn,
influences hardened properties of the concrete. The uncompacted void content of
a sample of fine aggregate can indicate the angularity, sphericity and surface
texture of the aggregate relative to another aggregate of the same gradation. The
influence in fresh and hardened properties is due to changes in water demand by
the fine aggregate. ASTM C 1252-93, Standard Test Methods for Uncompacted
Void Content of Fine Aggregate (as Influenced by Particle Shape, Surface
Texture and Grading) methods A, B, and C were preformed on all aggregate
samples.
In the specification title, ASTM C 1252 is introduced as a test for
determining the particle shape and surface texture of fine aggregates. However,
the problem with this type of volumetric test is that the test will not distinguish
between the two separate characteristics, since particles can have similar shapes
and volumes, but differing surface textures. Hudson has investigated test methods
that will distinguish between these characteristics [Hudson, 1999].

59
4.3.4 Hydrometer Particle Size Analysis
The use of hydrometers to determine the particle size distribution is
primarily a test for soils. Only material passing the No. 200 (75 m) sieve was
used in the test. This test was conducted in accordance with ASTM D 422-90,
Standard Test Method for Particle Size Analysis of Soils.

4.3.5 Methylene Blue Test


The methylene blue test is one of the most important tests in this study.
Since the test samples contain high quantities of micro fines, the test of the
properties of the micro fines is definitely a crucial test. The methylene blue test
was introduced from France as an effective method to evaluate the presence of
potential harmful materials in an aggregate finer than No. 200 (75 m) sieve. The
test procedure used in this study was developed in NCHRP Project 4-19 [NCHRP,
1998] for determining the presence of potentially deleterious materials in
aggregate. ASTM C 837-92 is another test used to measure methylene blue
absorption; however it is intended more for clay than for crushed aggregate fines.
The test was performed by saturating a 10.0g ( 0.05g) sample of material
passing the No. 200 (75m) sieve with 30g of distilled water. One gram of
methylene blue is dissolved in enough distilled water to produce 200 ml of
solution, with each 1 ml of solution containing 5 mg of methylene blue. With the
slurry still mixing, 0.5 ml of methylene blue solution is added into the slurry and

60
stirred for one minute. A small amount of water, which contains the sample
material and titrated methylene blue, is removed via a glass rod and dropped onto
filter paper. If a light blue halo is not observed, an additional 0.5 ml of MB is
added and stirred. The procedure continues until the halo is observed.

The

amount of MB required to produce the halo is the amount recorded.

4.3.6 Particle Size Analysis using Coutler LS130 Laser Diffraction Analyzer
Since the hydrometer particle size analysis shows the percent passing of
micro fines by weight according to particle size, it is difficult to know the actual
volume distribution of the micro fines. Hence, another particle size analysis was
performed by Paul Lessard at Granite Rock Company in California. Twenty-nine
samples (mortar test samples) were tested using the Coutler LS130 laser
diffraction analyzer.

4.3.7 Chemical Analysis


The chemical analysis of the aggregates employed inductively coupled
plasma (ICP) technology that was performed in the Geochemistry Analytical Lab
at The University of Texas at Austin. First, ten grams of selected manufactured
sand samples were ground finer than 75m. The 10-gram sample was immersed
in water and sprayed into a high temperature plasma as shown in Figure 4.1. Then
excited free atoms were produced (e.g., high temperature gas that is a mixture of

61
ions and electrons in a magnetic field). The wavelength of the resulting atomic
emission is a function of a particular element, and the intensity is a function of
concentration [Ingamells, 1986]. Through this process, the chemical composition
of a sample was obtained. A typical ICP spectrometer configuration is shown in
Figure 4.2.

Figure 4.1: Illustration of Typical Inductively Coupled Plasma (ICP)


Torch used as Source in Emission Spectrometry [Ingamells, 1986]

62

Figure 4.2: Illustration of Typical ICP Configuration [Ingamells, 1986]

4.4

Test Results and Discussion

4.4.1 Specific Gravity and Absorption Capacity Test


Table 4.3 shows the test results for specific gravity, absorption capacity,
micro fines content, and methylene blue for 112 sands. The complete test results
of specific gravity and absorption capacity test are shown in Appendix A (Table
A.1).

As shown in Table 4.3, South Dakota quartzite, 66-Virginia granite,

Connecticut granite, and some of the Pennsylvania sandstone had lower absorption
capacity (bold items).

63
Table 4.3: Test Results of Bulk Specific Gravity, Absorption Capacity, Micro
Fines Content, and Methylene Blue
ID No.
PA/LS/13-02/68
PA/LS/13-02/36
PA/LS/06-00/00
PA/LS/06-00/65
PA/LS/06-00/36
PA/LS/05-00/36
PA/LS/05-00/65
PA/LS/05-00/00
VA/70/GT/19-02/35
VA/70/GT/19-02/68
VA/70/GT/06-00/68
VA/70/GT/06-00/35
VA/70/GT/05-00/68
VA/70/GT/06-00/00
VA/70/GT/05-00/35
VA/70/GT/05-00/00
SD/QZ/19-02/68
SD/QZ/19-02/36
SD/QZ/06-00/00
SD/QZ/06-00/65
SD/QZ/06-00/36
SD/QZ/05-00/00
SD/QZ/05-00/65
SD/QZ/05-00/36
VA/DI/06-00/00
VA/DI/13-02/68
VA/DI/13-02/36
VA/DI/06-00/65
VA/DI/06-00/36
VA/DI/05-00/00
VA/DI/05-00/65
VA/DI/05-00/36

Bulk Specific Absorption Micro fines Methylene Blue


Gravity (OD)
(%)
Content (%)
Value
2.62
1.6
7.9
1.50
2.63
1.1
5.5
1.50
2.63
1.3
3.6
2.25
2.57
2.8
10.2
1.25
2.63
1.5
7.5
1.25
2.60
1.9
10.2
1.50
2.52
2.6
14.3
1.50
2.57
1.7
10.0
1.50
2.71
1.5
7.8
1.25
2.71
2.0
11.7
1.25
2.72
1.7
18.4
3.50
2.74
1.5
16.3
3.00
2.73
1.6
13.3
1.25
2.68
2.4
15.2
2.75
2.72
1.5
7.1
1.50
2.74
1.2
5.6
1.25
2.64
0 .1 *
11.6
0.50
2.63
0 .3
8.1
0.75
0 .6
2.61
7.4
1.25
2.64
13.5
0.75
0 .3
0 .2
2.64
11.0
1.00
2.62
0 .4
1.7
1.50
2.64
0 .1
8.4
1.00
0 .3
2.63
3.9
1.50
2.77
1.1
11.3
3.50
2.76
1.3
9.6
3.25
2.70
2.3
8.3
3.75
2.80
0.8
15.8
4.00
2.79
1.1
13.5
3.25
2.77
1.6
4.2
3.50
2.75
1.9
9.6
3.75
2.77
1.5
5.8
3.25

64
Table 4.3: Test Results of Bulk Specific Gravity, Absorption Capacity, Micro
Fines Content, and Methylene Blue (Continued)
ID No.
OK/DO/19-02/68
OK/DO/19-02/36
OK/DO/06-00/00
OK/DO/06-00/65
OK/DO/06-00/36
VA/LS/05-00/00
VA/LS/05-00/65
VA/LS/05-00/36
CT/BA/05-00/36
CT/BA/05-00/65
CT/BA/05-00/00
CT/BA/06-00/36
CT/BA/06-00/65
CT/BA/06-00/00
CT/BA/19-02/36
CT/BA/19-02/68
PA/SS/09-00/68
PA/SS/09-00/35
PA/SS/06-00/68
PA/SS/06-00/35
PA/SS/05-00/68
PA/SS/05-00/35
PA/SS/05-00/00
MO/LS/19-02/68
MO/LS/19-02/35
MO/LS/06-00/68
MO/LS/06-00/00
MO/LS/06-00/35
MO/LS/05-00/00
MO/LS/05-00/68
MO/LS/05-00/35

Bulk Specific Absorption Micro fines Methylene Blue


Gravity (OD)
(%)
Content (%)
Value
2.69
2.69
2.70
2.71
2.67

1.6
1.8
1.3
1.2
1.8

2 .8 2
2 .8 3
2 .8 2

0 .5
0 .3
0 .4

2.78
2.83
2.81
2.79
2.76
2.77
2.79
2.82
2.63
2.57
2.60
2.56
2.64
2.59
2.60

2.7
1.6
2.4
2.6
2.9
2.8
2.3
1.2

2.65
2.64
2.57
2.56
2.45
2.53
2.53
2.54

0 .3
1.1
0.8
1.3

0 .1
0.7
0.8
0.6
0.8
1.4
1.9
3.5
2.8
2.3
2.5

14.8
10.4
11.2
16.7
13.4
4.5
13.3
7.4
8.0
10.9
4.7
15.0
16.1
12.7
12.9
14.3
10.3
7.5
9.2
5.3
9.6
4.0
2.2

1.25
1.50
2.75
2.25
2.25

0 .2 5
0 .2 5
0 .2 5
3.00
3.00
3.00
3.25
3.00
3.25
3.00
2.00
0.75
1.00
0.75
0.75
0.75
0.75
0.75

15.8
13.5

2.75
4.00

3 0 .2
1 8 .4
2 9 .3

1 2 .0 0
1 0 .0 0
1 1 .5 0

4.2
15.8
8.3

7.50
5.00
4.50

65
Table 4.3: Test Results of Bulk Specific Gravity, Absorption Capacity, Micro
Fines Content, and Methylene Blue (Continued)
ID No.
VA/66/GT/13-02/35
VA/66/GT/13-02/68
VA/66/GT/06-00/68
VA/66/GT/06-00/35
VA/66/GT/05-00/68
VA/66/GT/06-00/00
VA/66/GT/05-00/35
VA/66/GT/05-00/00
MN/GT/19-02/68
MN/GT/19-02/35
MN/GT/06-00/68
MN/GT/06-00/35
MN/GT/05-00/68
MN/GT/05-00/00
MN/GT/05-00/35
MN/GT/06-00/00
CT/GT/19-02/65
CT/GT/19-02/36
CT/GT/06-00/65
CT/GT/06-00/00
CT/GT/06-00/36
CT/GT/05-00/00
CT/GT/05-00/65
CT/GT/05-00/36
WY/GT/19-02/65
WY/GT/19-02/36
WY/GT/06-00/00
WY/GT/06-00/65
WY/GT/06-00/36
WY/GT/05-00/00
WY/GT/05-00/65
WY/GT/05-00/36

Bulk Specific Absorption Micro fines Methylene Blue


Gravity (OD)
(%)
Content (%)
Value
2.66
9.6
1.50
0.3
2.67
13.8
1.75
0.2
2.69
18.7
1.50
0.1
2.68
13.6
1.50
0.2
2.65
11.9
1.25
0.3
2.70
12.4
1.50
0.4
2.68
4.9
1.25
0.4
2.65
2.3
1.25
N/A
2.65
10.8
2.25
0.3
2.65
7.7
2.00
0.5
2.66
14.3
2.50
0.3
2.65
4.5
2.00
0.6
2.63
10.8
2.00
0.5
2.63
2.0
3.00
0.9
2.64
4.1
2.50
0.8
2.66
5.3
2.50
0.6
13.5
1.00
2.82
0 .2
14.7
1.00
2.82
0 .2
2.82
17.1
1.50
0 .4
13.5
1.50
2.77
0 .6
20.7
1.50
2.80
0 .6
3.5
1.00
2.79
0 .8
6.5
1.25
2.77
0 .2
15.7
1.25
2.79
0 .5
2.72
15.6
3.00
0.4
2.70
11.8
3.50
0.5
2.68
9.0
7.00
1.2
2.67
14.3
5.50
0.3
2.65
11.1
5.00
0.7
2.67
2.8
3.50
1.2
2.67
14.8
3.00
0.4
2.66
6.4
3.00
0.9

66
Table 4.3: Test Results of Bulk Specific Gravity, Absorption Capacity, Micro
Fines Content, and Methylene Blue (Continued)
ID No.

Bulk Specific Absorption Micro fines Methylene Blue


Gravity (OD)
(%)
Content (%)
Value

TN/DO/13-02/68
2.83
0.4
15.7
0.50
TN/DO/13-02/36
2.81
0.7
11.9
0.50
TN/DO/06-00/00
2.83
0.8
11.1
0.50
TN/DO/06-00/65
2.82
0.4
15.5
0.50
TN/DO/06-00/36
2.79
0.8
13.1
0.50
IA/DO/25-02/68
2.79
0.2
26.5
0.50
IA/DO/25-02/45
2.80
0.6
22.2
0.25
IA/DO/06-06/35
2.82
1.2
13.3
0.75
IA/DO/06-06/68
2.84
0.4
23.8
0.50
IA/DO/05-00/00
2.76
2.8
5.2
1.00
IA/DO/05-00/68
2.79
0.2
24.3
0.50
IA/DO/05-00/35
2.83
0.9
12.0
0.75
1 .4
AR/SS/25-02/68
2.74
7.6
6.00
AR/SS/25-02/45
2.73
1.4
4.3
5.50
AR/SS/06-00/68
2.75
1.9
9.6
6.00
AR/SS/06-00/00
2.75
2.0
6.8
6.00
AR/SS/06-00/35
2.73
1.8
7.7
6.50
*Items in bold indicate values found much higher or lower than typical.

Quartzite and sandstone have high quantities (about 80%) of quartz


[Waddell, 1993]; as a result, the absorption is lower. Virginia limestone, however,
had low absorption capacity even though the materials have little quartz. As a
result, it was concluded that if a sample has a high quantity of quartz, the sample
usually has low absorption capacity; however, a low quantity of quartz in a sample
does not necessarily result in high absorption capacity.

67
In addition Virginia limestone, Tennessee dolomite, and some Iowa
dolomite showed higher specific gravity values (at oven dry condition) as well as
lower absorption capacity, which seems to indicate that the molecular structures of
these particles is pretty tight and dense.
The higher specific gravity and lower absorption capacity of Virginia
limestone, Tennessee dolomite and some Iowa dolomite may improve the
properties of the concrete. The test results for specific gravity and absorption
capacity were used in the mix proportioning of mortar and concrete batches.

4.4.2

Sieve Analysis by Washing


Table 4.3 shows the micro fines content for each sample. It is noted that

some of the Missouri limestone has the highest (up to 30%) micro fines content.
On the other hand some of South Dakota quartzite has the lowest (1.7%) micro
fines content. As shown in the table, the samples usually have 5 to 20% of micro
fines content. The percentage increased as crushing speed was increased. All of
the wet sieve analysis results are shown in Appendix A (Table A.2). Table A.2
shows the amount of cumulative passing for each sieve size as a percentage of
total sample by weight and fineness modulus of the sand based on ASTM C 33
(calculation up to No. 100 sieve).
The gradation graphs of each sample are shown in Figures 4.3 through
4.18. For clarity only the graphs of the samples used in mortar tests are presented.

68
As shown in the figures, the gradations of most MFA do not satisfy the ASTM C
33 grading specification. Under current practice, aggregate producers compensate
the insufficient portion and remove excess fines to meet ASTM C 33. However, in
this study nearly all crushed samples were used in mortar and concrete tests as
they were produced even if they did not meet ASTM C 33, since the main
objective of this study was the development of guidelines for mix proportioning of
concrete and modifications to existing construction specifications incorporating
higher fines contents.
100
PA/LS/13-02/68

Percent Passing by Weight

90

PA/LS/06-00/65
PA/LS/05-00/65

80

PA/LS/05-00/00
ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

Figure 4.3: Gradations of Pennsylvania Limestone

#200

69

Percent Passing by Weight

100
VA/GT/06-00/68

90

VA/GT/05-00/68
VA/GT/06-00/00

80

ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.4: Gradations of 70-Virginia Granite


100

SD/QZ/19-02/68

Percent Passing by Weight

90

SD/QZ/06-00/65
SD/QZ/05-00/00

80

SD/QZ/05-00/65

70

ASTM Upper
ASTM Lower

60
50
40
30
20
10
0

3/8"

#4

#8

#16

#30

#50

#100

Sieve Sizes
Figure 4.5: Gradations of South Dakota Quartz

#200

70

Percent Passing by Weight

100
VA/DI/06-00/00

90

VA/DI/06-00/65
VA/DI/05-00/65

80

ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.6: Gradations of Virginia Diabase


100

Percent Passing by Weight

90

O K /D O /06-00/00
O K /D O /06-00/65

80

O K /D O /06-00/36
A S TM U pper

70

A S TM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

S ieve S izes

# 50

#100

Figure 4.7: Gradations of Oklahoma Dolomite

#200

71

Percent Passing by Weight

100
VA/LS/05-00/00

90

VA/LS/05-00/65
VA/LS/05-00/36

80

ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.8: Gradations of Virginia Limestone

Percent Passing by Weight

100
CT/BA/06-00/36

90

CT/BA/06-00/00
CT/BA/19-02/68

80

ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

Figure 4.9: Gradations of Connecticut Basalt

#200

72
100

Percent Passing by Weight

90

PA/SS/09-00/68
PA/SS/06-00/68
PA/SS/05-00/00
ASTM Upper
ASTM Lower

80
70
60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.10: Gradations of Pennsylvania Sandstone


100

Percent Passing by Weight

90

MO/LS/19-02/68
MO/LS/06-00/00
MO/LS/05-00/68
ASTM Upper
ASTM Lower

80
70
60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

Figure 4.11: Gradations of Missouri Limestone

#200

73

100
VA/66/GT/13-02/68

Percent Passing by Weight

90

VA/66/GT/06-00/68

80

VA/66/GT/06-00/00
ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.12: Gradations of 66-Virginia Granite


100
MN/62/GT/06-00/68

Percent Passing by Weight

90

MN/62/GT/05-00/68
MN/62/GT/06-00/00

80

ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

Figure 4.13: Gradations of Minnesota Granite

#200

74

100
CT/68/GT/19-02/65

Percent Passing by Weight

90

CT/68/GT/06-00/00

80

CT/68/GT/05-00/36
ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.14: Gradations of Connecticut Granite


100
W Y/67/GT/19-02/65

Percent Passing by Weight

90

W Y/67/GT/06-00/00

80

W Y/67/GT/05-00/65
ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

Figure 4.15: Gradations of Wyoming Granite

#200

75

100
TN/56/DO/13-02/68

Percent Passing by Weight

90

TN/56/DO/06-00/00

80

TN/56/DO/06-00/65
ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.16: Gradations of Tennessee Dolomite


100
IA/72/DO/25-02/68

Percent Passing by Weight

90

IA/72/DO/06-06/68
IA/72/DO/05-00/00

80

IA/72/DO/05-00/68

70

ASTM Upper
ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

Figure 4.17: Gradations of Iowa Dolomite

#100

#200

76

Percent Passing by Weight

100
AR/73/SS/06-00/68

90

AR/73/SS/06-00/00

80

ASTM Upper

70

ASTM Lower

60
50
40
30
20
10
0
3/8"

#4

#8

#16

#30

Sieve Sizes

#50

#100

#200

Figure 4.18: Gradations of Arkansas Sandstone

4.4.3 Uncompacted Void Content Test


The uncompacted void content test results are shown in Appendix A (Table
A.3) for selected aggregates. As shown in the table, the void contents of the
samples are similar except for method C. The method C gave low void contents
due to using the whole sample instead of a part of sample. In Table A.3, the
results using ASTM C 1252-93 do not appear to present adequate definition of the
particle shape and texture of crushed fine aggregate.

77
Hudson, of Aggregate Research Industries, and Ownby, of Vulcan
Materials Company, developed a related method that uses the modified
uncomapacted void content concept to characterize the particle shape and texture
of fine aggregate [Hudson, personal communication]. The apparatus measures the
angle and height of repose and flow rate of a fine aggregate dropped from a
standard height. The angle will indicate the amount of internal friction in the
aggregate and the particle shape can be known by measuring the flow rate and the
height of repose. However, the apparatus was not developed in time to be used for
this study.

4.4.4 Hydrometer Particle Size Analysis


The test results of hydrometer particle size analysis for selected aggregates
are included in Appendix A (Figures A.1 through A.63). Since this test could
show the distribution of particle size up to only 30 m, another particle size
analysis was needed to get more accurate information for the samples.

4.4.5 Methylene Blue Test


Table 4.3 shows methylene blue test results. The methylene blue value
(MBV) was usually 0.25 to 4.00 except for Missouri limestone, Arkansas
sandstone, and some Wyoming granite.

Virginia limestone, Pennsylvania

sandstone, South Dakota quartzite, Connecticut granite, Tennessee and Iowa

78
dolomite showed lower MBV (less than 1.50) compared to other samples. On the
other hand Missouri limestone had the highest MBV ranging up to 12.00. The
highest MBV for the Missouri limestone was nearly 50 times higher than for
Virginia limestone. The reason is not clear, but is thought to be due to either the
presence of clay or silt, possibly from overburdening and/or the higher amount of
very fine particles in the Missouri limestone.

4.4.6 Particle Size Analysis using Coutler LS130 Laser Diffraction Analyzer
Using a Coutler LS130 laser diffraction analyzer, a selected 29 samples
(used also in mortar tests) were tested for particle size analysis. The analysis time
was 60 seconds, obscuration was typically between 8 to 12%, and pump speed was
51%. Figures 4.19 through 4.27 show the test results of particle size analysis
using Coutler LS130 laser diffraction analyzer for each rock type. The particle
sizes in the distribution of most samples show that the highest volume is for the
particle size range from 56 to 73m. As shown in Figure 4.19, however, the
particle size distribution of Pennsylvania limestone has the highest volumes in the
vicinity of particle size 10m (1st peak) and 60m (2nd peak).

79

3.5
PA/LS/13-02/68

PA/LS/06-00/65
PA/LS/05-00/65

Volume (%)

2.5

PA/LS/05-00/00

2
1.5
1
0.5
0
0.3

0.7

1.8

4.6

11.4

28.3

70.3

Particle Size (um)

Figure 4. 19: Particle Size Analysis of Pennsylvania Limestone by Volume


6

VA/GT/06-00/68
VA/GT/05-00/68
VA/GT/06-00/00

Volume (%)

0
1.0

4.2

17.9

77.0

Particle Size (um)

Figure 4. 20: Particle Size Analysis of Virginia Granite by Volume

80

SD/QZ/19-02/68
SD/QZ/06-00/65
SD/QZ/05-00/00
SD/QZ/05-00/65

Volume (%)

5
4
3
2
1
0
1.0

4.2

17.9

77.0

Particle Size (um)

Figure 4.21: Particle Size Analysis of South Dakota Quartzite by Volume


7

VA/DI/06-00/00
VA/DI/06-00/65
VA/DI/05-00/65

Volume (%)

5
4
3
2
1
0
1.0

4.2

17.9

77.0

Particle Size (um)

Figure 4.22: Particle Size Analysis of Virginia Diabase by Volume

81

OK/DO/06-00/00
OK/DO/06-00/65
OK/DO/06-00/36

Volume (%)

5
4
3
2
1
0
1.0

4.2

17.9

77.0

Particle Size (um)

Figure 4.23: Particle Size Analysis of Oklahoma Dolomite by Volume


7

VA/LS/05-00/00
VA/LS/05-00/65
VA/LS/05-00/36

Volume (%)

5
4
3
2
1
0
1.0

4.2

17.9

77.0

Particle Size (um)

Figure 4.24: Particle Size Analysis of Virginia Limestone by Volume

82

CT/BA/06-00/36
CT/BA/06-00/00
CT/BA/19-02/68

Volume (%)

5
4
3
2
1
0
1.0

4.2

17.9

77.0

Particle Size (um)

Figure 4.25: Particle Size Analysis of Connecticut Basalt by Volume


7

PA/SS/09-00/68
PA/SS/06-00/68
PA/SS/05-00/00

Volume (%)

5
4
3
2
1
0
1.0

4.2

17.9

77.0

Particle Size (um)

Figure 4.26: Particle Size Analysis of Pennsylvania Limestone by Volume

83

MO/LS/19-02/68
MO/LS/06-00/00
MO/LS/05-00/68

3.5

Volume (%)

3
2.5
2
1.5
1
0.5
0
0.3

0.7

1.8

4.6

11.4

28.3

70.3

Particle Size (um)

Figure 4.27: Particle Size Analysis of Missouri Limestone by Volume


In addition, as shown in Figure 4.27, the particle size distribution of the
Missouri limestone as-received sample (MO/LS/06-00/00) shows high volume
around particle size 18m. After crushing the distribution was changed to high
volumes around particle size 60m (1st peak) and 15m (2nd peak). The remainder
of the selected aggregate test results for particle size analyses is in Appendix A
(Figures A.64 through A.92).

4.4.7 Chemical Analysis


There were two series of chemical analyses in this study. Series 1 used the
selected aggregate samples with micro fines directly delivered from the quarry,

84
and series 2 used material obtained from crushing size larger than 4.75mm. This
test was conducted by the ICP technology in the Geochemistry Analytical Lab at
The University of Texas at Austin on selected aggregate types used in this study.
The chemical compositions examined in this study included CaO, MgO, Fe2O3,
Na2O, K2O, MnO, TiO2, SiO2, Al2O3, and ignition. The results of the chemical
analyses are shown in Tables 4.4 and 4.5. It is noted that there is a difference in
results between series 1 and 2 for Missouri limestone (bold items). The lower
amount of K2O, SiO2, and Al2O3 in series 2 may indicate that clay, probably from
overburdening, was present in the as-received materials (series 1).

Table 4.4: Chemical Composition (%) of Series 1 (As-Received)


ID No.

CaO

MgO

Fe2O3

PA/LS/05-00/00
VA/GT/06-00/00
SD/QZ/05-00/00
VA/DI/06-00/00
OK/DO/06-00/00
VA/LS/05-00/00
CT/BA/06-00/00
PA/SS/05-00/00
MO/LS/06-00/00

45.1
5.8
0.4
7.4
25.0
30.1
9.5
0.3

3.3
4.3
0.0
5.2
15.3
21.1
5.7
0.3
1.3

0.8
7.7
0.4
8.7
1.1
0.3
12.3
0.9
1.6

37.6

Na2O

K2O

MnO

TiO2

SiO2

Al2O3

Ignition

0.1
2.8
0.6
2.7
0.4
0.3
2.4
0.5
0.5

0.0
1.4
0.4
1.9
3.0
0.4
0.3
1.5

0.0
0.2
0.1
0.1
0.0
0.0
0.2
0.0
0.1

0.1
0.9
0.3
1.1
0.2
0.0
1.0
0.4
0.3

11.0
57.3
92.4
52.6
15.5
2.0
51.3
87.7

1.5
16.4
3.7
17.4
3.4
0.6
14.5
5.7

38.8
2.9
1.3
2.1
36.0
45.7
2.5
2.0

22.0

5.3

30.2

1.1

Table 4.5: Chemical Composition (%) of Series 2 (Crushed from Size > 4.75mm)
ID No.

CaO

MgO

Fe2O3

Na2O

PA/LS/05-00/00
VA/GT/06-00/00
SD/QZ/05-00/00
VA/DI/06-00/00
OK/DO/06-00/00
VA/LS/05-00/00
CT/BA/06-00/00
PA/SS/05-00/00
MO/LS/06-00/00

44.4
6.1
0.0
6.2
24.5
27.0
9.6
0.4

2.5
2.2
0.0
5.0
16.1
19.9
5.8
0.3
0.8

0.5
5.2
0.1
8.2
0.8
0.2
12.3
0.1
0.8

0.0
4.5
0.0
3.2
0.0
0.0
2.4
0.6
0.1

50.2

K2O
0.0
0.6
0.0
2.3
2.1
0.0
0.7
1.4

0.0

MnO

TiO2

SiO2

Al2O3

Ignition

0.0
0.1
0.0
0.1
0.0
0.0
0.2
0.0
0.1

0.1
0.6
0.1
1.0
0.1
0.0
1.0
0.1
0.1

12.4
60.3
97.1
57.3
17.5
6.7
51.8
89.4

1.4
16.7
0.9
15.6
2.7
0.4
13.8
5.0

39.6
2.7
1.0
1.6
36.5
45.8
1.6
1.7

8.9

1.4

38.3
85

86
4.5

Results of Other Aggregate Characterization Tests


Vulcan Materials Company and Svedala Barmac performed a companion

study on over 200 aggregate samples from 29 locations using 3 different types of
rock.

The results of these aggregate characterization tests can be found in

Appendix B.

4.6

Summary
Aggregate characterization tests were performed for sixteen sources of

aggregates (seven types of rock). Numerous sands were tested for seven physical
properties.
South Dakota quartzite, 66-Virginia granite, Connecticut granite, and some
of Pennsylvania sandstone had lower absorption capacity. Quartzite and sandstone
had high quantities (about 80%) of quartz; as a result the absorption was lower.
Virginia limestone, however, had low absorption capacity even though the
material has little quartz. As a result, it is concluded that if a sample has a high
quantity of quartz and the sample usually has low absorption capacity; however, a
low quantity of quartz in a sample does not always result in high absorption
capacity. In addition Virginia limestone, Tennessee and Iowa dolomite showed
higher specific gravity (at oven dry condition) as well as lower absorption
capacity.

87
Most samples had 5 to 20% micro fines content. The Missouri limestone
had the highest (up to 30%) micro fines content; on the other hand some South
Dakota quartzite had the lowest (1.7%).

The percentage of micro fines was

definitely increased as crushing speed was increased. The gradations of most


MFA in this study did not meet ASTM C33 specification.
The void contents of the samples are similar. It is concluded that ASTM
C1252 is not an adequate test to differentiate particle shape and texture of MFA.
Since the hydrometer particle size analysis test measures distribution of
size up to only 30m, another particle size analysis was needed to get more
accurate information of the samples.
The methylene blue value (MBV) was usually 0.25 to 4.00 except for
Missouri limestone, Arkansas sandstone and some Wyoming granite. MBVs of
less than 1.50 were obtained for Virginia limestone, Pennsylvania sandstone,
Connecticut granite, South Dakota quartzite and Tennessee and Iowa dolomite.
Missouri limestone had the highest MBV, up to 12.00. The reason is not clear, but
is thought to be due to either the presence of clay or silt and/or the higher amount
of very fine particles in the Missouri limestone.
Using a Coutler LS130 laser diffraction analyzer on a selected number of
samples, the particle size distribution of most of these samples shows that the
highest volume is for the particle size range from 56 to 73m. However, the
particle size distribution of Pennsylvania limestone showed the highest volumes in

88
the vicinity of particle size 10m (1st peak) and 60m (2nd peak). In addition, the
particle size distribution of the Missouri limestone as-received sample (MO/LS/0600/00) showed high volume around particle size 18m.

After crushing the

distribution was changed to high volumes around particle size 60m (1st peak) and
15m (2nd peak).
There were two series of chemical analyses performed on selected
aggregates in this study. Series 1 used micro fines as received from the quarry,
and series 2 used the same aggregates but the test samples resulted from crushing
of sizes larger than 4.75mm. It is noted that there is a difference in results between
series 1 and 2 for Missouri limestone (bold items). The lower amount of K2O,
SiO2, and Al2O3 in series 2 may indicate that overburden containing clay was
present in the as-received materials (Series 1).

CHAPTER FIVE: MORTAR TESTING PROGRAM

5.1

Introduction
After screening and performing aggregate characterization tests, mortar

tests were performed to investigate the effect of MFA on the mortar paste. The
following sections outline the materials, the test procedures, and the results and
discussion of mortar tests using the selected sands based on the aggregate
characterization test results. The materials and procedures used throughout this
study conform to procedures approved by an accepted standard except mixture
proportioning did not conform ASTM C 33.

With the mortar test results,

statistical analyses were conducted as described in Chapter 7. Based on the mortar


test results several sands were also selected for the concrete testing stage.

5.2

Materials

5.2.1 Portland Cement


Commercially available Type I cement was used throughout this study. It
conformed to ASTM C 150-94, Standard Specification for Portland Cement.

5.2.2 Mixing Water


Potable City of Austin water was used throughout the laboratory batching
series and was assumed to have a specific gravity of 37 kg/m3 (62.4 lbs/cy).
89

90
5.2.3 Fine Aggregate
Table 5.1 shows aggregate samples (seven types of rock) that were tested
for mortar tests.

Based on the characterization test results some sands were

selected on the following basis:


1)

At least one as-received sample per aggregate type

2)

A total of two to four samples per aggregate type

3)

Sands containing higher amounts of micro fines

4)

Sands with very low contents of micro fines

Some of the selected as-received samples were subjected to chemical


analysis.

Particular selected samples underwent particle size analysis using a

Coutler LS130 Laser Diffraction Analyzer.


At least one as-received sample per aggregate type was selected to compare
the results of other crushed aggregates. Since one of the objectives of this study
was the determination of the effect of higher amounts for several types of crusher
fines on mortar and concrete properties, two to four samples per aggregate type
were selected based on higher micro fines contents. They included the samples
with very low micro fines contents to compare to those with higher contents. As a
result 50 sands (see Table 5.1) were selected for mortar testing and 29 of those
samples were subjected to particle size analysis using a Coutler LS130 Laser
Diffraction Analyzer.

91

Table 5.1: Aggregates for Mortar Tests


ID No.

Location

Type

Size, mm Speed, m/s

Cycle

PA/LS/13-02/68

13-02

68

product

PA/LS/06-00/65
Pennsylvania Limestone
PA/LS/05-00/65

06-00

65

product

05-00

65

product

PA/LS/05-00/00

05-00

00

as-received

VA/GT/06-00/68

06-00

68

product

05-00

68

product

VA/GT/06-00/00

06-00

00

as-received

SD/QZ/19-02/68

19-02

68

product

06-00

65

product

05-00

00

as-received

SD/QZ/05-00/65

05-00

65

product

VA/DI/06-00/00

06-00
06-00

00
65

as-received
product

05-00

65

product

06-00

00

as-received

06-00

65

product

OK/DO/06-00/36

06-00

36

product

VA/LS/05-00/00

05-00
05-00

00
65

as-received
product

05-00

36

product

06-00

36

product

06-00

00

as-received

CT/BA/19-02/68

19-02

68

product

PA/SS/09-00/68

09-00
06-00

68
68

product
product

05-00

00

as-received

19-02

68

product

06-00

00

as-received

05-00

68

product

VA/GT/05-00/68

SD/QZ/06-00/65
SD/QZ/05-00/00

VA/DI/06-00/65
VA/DI/05-00/65

Virginia

Granite

South Dakota Quartzite

Virginia

Diabase

OK/DO/06-00/00
OK/DO/06-00/65

VA/LS/05-00/65
VA/LS/05-00/36

Oklahoma

Virginia

Dolomite

Limestone

CT/BA/06-00/36
CT/BA/06-00/00

Connecticut

Basalt

PA/SS/06-00/68 Pennsylvania Sandstone


PA/SS/05-00/00
MO/LS/19-02/68
MO/LS/06-00/00
MO/LS/05-00/68

Missouri

Limestone

92
Table 5.1: Aggregates for Mortar Tests (Continued)
ID No.

Location

VA/GT/13-02/68
Virginia
VA/GT/06-00/68
VA/GT/06-00/00
MN/GT/06-00/68
MN/GT/05-00/68 Minnesota
MN/GT/06-00/00
CT/GT/19-02/65
CT/GT/06-00/00 Connecticut
CT/GT/05-00/36

Type
Granite

Granite

Granite

WY/GT/19-02/65
WY/GT/06-00/00
WY/GT/05-00/65

Wyoming

Granite

TN/DO/13-02/68
TN/DO/06-00/00
TN/DO/06-00/65

Tennessee

Dolomite

Iowa

Dolomite

Arkansas

Sandstone

IA/DO/25-02/68
IA/DO/06-06/68
IA/DO/05-00/00
IA/DO/05-00/68
AR/SS/06-00/68
AR/SS/06-00/00

5.3

Size, mm Speed, m/s


13-02
06-00
06-00
06-00
05-00
06-00
19-02
06-00
05-00
19-02
06-00
05-00
13-02
06-00
06-00
25-02
06-06
05-00
05-00
06-00
06-00

68
68
00
68
68
00
65
00
36
65
00
65
68
00
65
68
68
00
68
68
00

Cycle
product
product
as-received
product
product
as-received
product
as-received
product
product
as-received
product
product
as-received
product
product
product
as-received
product
product
as-received

Testing Procedures for Mortar Tests


The mixing procedure for mortars was in accordance with ASTM C 305-

94, Standard Practice for Mechanical Mixing of Hydraulic Cement Pastes and
Mortars of Plastic Consistency.
following section.

The testing procedure for mortar is in the

93
5.3.1 Flow
The flow table test was conducted in accordance with ASTM C 109/C
109M-95, Standard Test Method for Compressive Strength of Hydraulic Cement
Mortars using the flow table in accordance with ASTM C 230-97, Standard
Specification for Flow Table for Use in Tests of Hydraulic Cement. After the
mortar paste was filled and tamped (two layers) into a flow mold, the flow table
plate was dropped 25 times in 15s. Using the calipers, the flow was determined by
measuring the diameter of mortar paste on the table. The standard flow of mortar
was 110 5. This test was used to evaluate workability of the mortar paste. Also
water demand of each sand could be known.

5.3.2 Compressive Strength


The test of compressive strength was conducted according to ASTM C
109/C 109M-95, Standard Test Method for Compressive Strength of Hydraulic
Cement Mortars (Using 50-mm [2-in.] cube specimens). This test was used to
characterize the compressive strength as well as the water demand of each fine
aggregate. Compression cubes were made for each mixture and tested at ages of 1
and 28 days. Before testing, the specimens were cured in lime-saturated water in
storage tanks. The results were correlated with various aggregate characterization
properties in order to illustrate trends in the data.

94
5.3.3 Drying Shrinkage
The drying shrinkage was obtained in accordance with ASTM C 157-93,
Standard Test Method for Length Change of Hardened Hydraulic-Cement Mortar
and Concrete. Three prisms, 1 inch x 1 inch x 11 inches, were prepared for each
mixture. The specimens were made and cured in the lime-saturated water for 3
days. After removal from the lime-saturated water, the specimens were stored in
the drying room and comparator readings for each specimen were taken at the
periods required by the standard test procedure.

5.4

Mixture Proportioning
Two control variables were used in the mortar test program: fixed water-

cement ratio and fixed flow rate (workability) of mortar. The first variable was
selected to investigate the effect of the sand on the properties of mortar at the same
water-cement ratio, and the last one was chosen to evaluate the effect of the
property change according to water demand.
The compressive strength of concrete was simulated as 28 MPa (4000 psi)
for this particular part of the study; hence, cement contents were targeted at a
typical cement volume for this grade of concrete. From a recent concrete project
by Calera, an optimal value of sand used in concrete was found to be 42 percent
based on the total aggregate volume of concrete. The mortar proportion was

95
simulated from the concrete mixture proportions after removal of coarse aggregate
[Hudson, personal communication].
Cement contents were set based on a control (250 kg/m3) and two
increments of 10%. Six mixtures per sand were tested. Mixtures of fixed watercement ratio with low, medium, and high cement content were labeled as LF, MF,
HF, respectively. Similarly low, medium, and high cement content mixtures with
fixed flow were labeled as LV, MV, HV, respectively.
The first three batches (LF, MF, HF) were made based on a fixed watercement ratio, 0.485, and the only difference was cement content, 250 kg/m3, 275
kg/m3, and 300 kg/m3, respectively.

This served as a two-part comparison;

obviously the extra cement contents introduced strength changes as well as


showing the effects of a lower sand percentage. The other three mixtures (LV,
MV, HV), although the water-cement ratio was the same initially, had used a
variable amount of water to achieve a flow of 110 5, and the resulting watercement ratio could change. Detailed mixture proportions are shown in Appendix
C (Table C.1).

5.5

Test Results and Discussion

5.5.1 Flow
Table 5.2 shows the average test results of mortar for three batches (LF,
MF, and HF). As shown in the table, the flows of Pennsylvania and Virginia

96
limestone, Tennessee Dolomite and Pennsylvania sandstone had higher values.
On the other hand, Missouri limestone, 66-Virginia Granite, Iowa Dolomite, and
Virginia diabase had lower flows that might cause higher water demand.
Specially, the reason that the as-received sample of Missouri limestone had lower
flow may be related to the higher methylene blue value as well as higher micro
fines contents.
It was noted that some of the as-received samples had higher flow than
those of product samples. Since the flow test is dynamic instead of static, micro
fines content as well as particle shape affected the flow. The flow of mortar with
fixed water-cement ratio depends on micro fines contents, particle shape, and
MBV. Detailed flow test results are shown in Appendix C (Table C.2).

5.5.2 Water Demand


The average water-cement ratios for three fixed-flow batches (LV, MV,
HV) are shown in Table 5.2. As shown in the table, the water-cement ratios were
inversely related to the flows, since low flow for fixed water-cement ratio
represents high water demand for fixed flow. Detailed water demand is shown in
Appendix C (Table C.1).

97
Table 5.2: Test Results of Mortar
Flow

W/C

for

for

W/C:

Flow:

0.485
PA/LS/13-02/68
PA/LS/06-00/65

ID No.

Compressive Strength
at 28days (psi)

Drying Shrinkage
at 28 days (%)

110

W/C:
0.485

Flow:
110

W/C:
0.485

Flow:
110

127

0.44

9100

9850

0.085

0.080

129

0.45

8370

8930

0.091

0.082

PA/LS/05-00/65

125

0.44

8350

9150

0.089

0.088

PA/LS/05-00/00

122

0.46

8500

9020

0.082

0.080

VA/GT/06-00/68

74

0.54

7570

6530

0.105

0.112

VA/GT/05-00/68

106

0.50

8290

8130

0.094

0.094

VA/GT/06-00/00

65

0.55

7200

6570

0.109

0.112

SD/QZ/19-02/68

104

0.50

9680

9230

0.072

0.073

SD/QZ/06-00/65

93

0.50

9730

9570

0.076

0.078

SD/QZ/05-00/00

114

0.47

8630

9340

0.066

0.068

SD/QZ/05-00/65

85

0.51

8350

9090

0.076

0.077

VA/DI/06-00/00

64

0.55

9570

8000

0.096

0.102

VA/DI/06-00/65

45

0.56

9450

7900

0.110

0.115

VA/DI/05-00/65

93

0.51

8720

8200

0.097

0.098

OK/DO/06-00/00

80

0.53

8900

7900

0.086

0.093

OK/DO/06-00/65

92

0.52

9460

8800

0.094

0.095

OK/DO/06-00/36

101

0.50

9020

8430

0.088

0.095

VA/LS/05-00/00

136

0.44

9390

10410

0.076

0.071

VA/LS/05-00/65

121

0.46

10370

11110

0.078

0.077

VA/LS/05-00/36

133

0.44

9600

11140

0.068

0.070

CT/BA/06-00/36

85

0.53

9510

8210

0.092

0.090

CT/BA/06-00/00

99

0.50

8350

7180

0.083

0.088

CT/BA/19-02/68

105

0.49

8800

8700

0.083

0.083

PA/SS/09-00/68

108

0.49

9060

9090

0.078

0.072

PA/SS/06-00/68

124

0.46

8930

9420

0.078

0.079

PA/SS/05-00/00

116

0.48

8630

9070

0.080

0.081

MO/LS/19-02/68

63

0.55

9100

7620

0.082

0.085

MO/LS/06-00/00

16

0.75

6710

3870

0.142

0.165

MO/LS/05-00/68

90

0.51

8380

7950

0.090

0.087

Natural Sand

143

0.41

7770

9420

0.072

0.066

98
Table 5.2: Test Results of Mortar (Continued)
Flow

W/C

for

for

W/C:

Flow:

0.485
VA/GT/13-02/68

ID No.

Compressive Strength
at 28days (psi)

Drying Shrinkage
at 28 days (%)

110

W/C:
0.485

Flow:
110

W/C:
0.485

Flow:
110

74

0.522

7880

8100

0.0793

0.0822

VA/GT/06-00/68

28

0.577

5710

6080

0.0828

0.0885

VA/GT/06-00/00

82

0.513

7920

7930

0.0745

0.0747

MN/GT/06-00/68

80

0.516

7840

7990

0.0702

0.0712

MN/GT/05-00/68

114

0.480

8450

8570

0.0748

0.0742

MN/GT/06-00/00

105

0.491

8450

8260

0.0628

0.0627

CT/GT/19-02/65

93

0.501

8480

8250

0.0813

0.0792

CT/GT/06-00/00

88

0.517

7860

7490

0.0787

0.0845

CT/GT/05-00/36

115

0.483

8380

8500

0.0783

0.0787

WY/GT/19-02/65

90

0.514

9440

9120

0.0835

0.0850

WY/GT/06-00/00

101

0.497

8720

8280

0.0837

0.0858

WY/GT/05-00/65

80

0.521

9110

8710

0.0897

0.0933

TN/DO/13-02/68

130

0.451

9490

10280

0.0690

0.0642

TN/DO/06-00/00

131

0.451

8540

9670

0.0580

0.0628

TN/DO/06-00/65

120

0.469

9210

9890

0.0717

0.0713

IA/DO/25-02/68

47

0.557

9180

8150

0.0780

0.0780

IA/DO/06-06/68

31

0.557

8860

8210

0.0727

0.0790

IA/DO/05-00/00

106

0.488

8350

8070

0.0702

0.0705

IA/DO/05-00/68

29

0.575

7920

7510

0.0740

0.0797

AR/SS/06-00/68

76

0.550

8880

7690

0.1058

0.1073

AR/SS/06-00/00

80

0.539

8890

7770

0.0950

0.0980

5.5.3 Compressive Strength


The average values of 28-day compressive strength are shown in Table 5.2.
Figures 5.1 through 5.17 show the test results of compressive strength for each

99
type of aggregate source in the form of compressive strength versus type of
mixture.
Test results of mixtures with sandstone, limestone, or quartzite show that
they have high compressive strengths. The Virginia limestone used in this study
showed consistent high compressive strength and good workability, even though
the percentage of micro fines was as high as 13.3 percent. Mortar made with
Missouri limestone MO/LS/06-00/00 had poor workability and the percentage of
reduction in compressive strength with fixed water-cement ratio and fixed flow
compared to Virginia limestone VA/LS/05-00/65 was 35 percent and 65 percent,
respectively. It should be noted that the former had the highest MBV of 10.0, and
the latter had the lowest MBV of 0.25. As shown in the Figure 5.2, the test results
of Virginia granite show that they are consistent in low compressive strength, even
when the cement content increases.
For a fixed water-cement ratio, aggregates can be separated into three
groups by comparing compressive strength versus cement content.

The first

group, 70-VA/GT/06-00/00, 70-VA/GT/06-00/68, PA/LS/13-02/68, MO/LS/0600/00, MO/LS/19-02/68, 66-VA/GT/13-02/68, 66-VA/GT/06-00/68, MN/GT/0600/68, WY/GT/19-02/65, and IA/DO/05-00/68, has increasing compressive
strength with increasing cement content.

100

Compressive Strength at 28 days (psi)

12000

10000

8000

PA/LS/13-02/68
PA/LS/06-00/65

6000

PA/LS/05-00/65
PA/LS/05-00/00

4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.1: Compressive Strengths of Mortar of Pennsylvania Limestone

Compressive Strength at 28 days (psi)

12000

10000

8000

VA/GT/06-00/68
VA/GT/05-00/68

6000

VA/GT/06-00/00
4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.2: Compressive Strengths of Mortar of Virginia Granite

101

Compressive Strength at 28 days (psi)

12000

10000

8000

SD/QZ/19-02/68
SD/QZ/06-00/65

6000

SD/QZ/05-00/00
SD/QZ/05-00/65

4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.3: Compressive Strengths of Mortar of South Dakota Quartzite

Compressive Strength at 28 days (psi)

12000

10000

8000

VA/DI/06-00/00
VA/DI/06-00/65

6000

VA/DI/05-00/65
4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.4: Compressive Strengths of Mortar of Virginia Diabase

102

Compressive Strength at 28 days (psi)

12000

10000

OK/DO/0600/00
OK/DO/0600/65
OK/DO/0600/36

8000

6000

4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.5: Compressive Strengths of Mortar of Oklahoma Dolomite

Compressive Strength at 28 days (psi0

12000

10000

8000

VA/LS/05-00/00
VA/LS/05-00/65

6000

VA/LS/05-00/36
4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.6: Compressive Strengths of Mortar of Virginia Limestone

103

Compressive Strength at 28 days (psi)

12000

10000

8000

CT/BA/06-00/36
CT/BA/06-00/00

6000

CT/BA/19-02/68
4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.7: Compressive Strengths of Mortar of Connecticut Basalt

Compressive Strength at 28 days (psi)

12000

10000

8000

PA/SS/09-00/68
PA/SS/06-00/68

6000

PA/SS/05-00/00
4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.8: Compressive Strengths of Mortar of Pennsylvania Sandstone

104

Compressive Strength at 28 days (psi)

12000

10000

8000

MO/LS/19-02/68
MO/LS/06-00/00

6000

MO/LS/05-00/68
4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.9: Compressive Strengths of Mortar of Missouri Limestone

Compressive Strength at 28 days (psi)

12000

10000

8000

Natural Sand
6000

4000

2000

0
LF

MF

HF

LV

MV

HV

Type of Mixure

Figure 5.10: Compressive Strengths of Mortar of Natural Sand

105

Compressive Strength at 28 days (psi)

12000

10000

8000
VA/GT/13-02/68

6000

VA/GT/06-00/68
VA/GT/06-00/00

4000

2000

LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.11: Compressive Strength of Mortar of Virginia Granite

Compressive Strength at 28 days (psi)

12000
10000
8000
MN/GT/06-00/68

6000

MN/GT/05-00/68
MN/GT/06-00/00

4000
2000
0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.12: Compressive Strength of Mortar of Minnesota Granite

106

Compressive Strength at 28 days (psi)

12000
10000
8000
CT/GT/19-02/65

6000

CT/GT/06-00/00
CT/GT/05-00/36

4000
2000
0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.13: Compressive Strength of Mortar of Connecticut Granite

Compressive Strength at 28 days (psi)

12000
10000
8000
WY/GT/19-02/65

6000

WY/GT/06-00/00
WY/GT/05-00/65

4000
2000
0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.14: Compressive Strength of Mortar of Wyoming Granite

107

Compressive Strength at 28 days (psi)

12000
10000
8000
TN/DO/13-02/68

6000

TN/DO/06-00/00
TN/DO/06-00/65

4000
2000
0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.15: Compressive Strength of Mortar of Tennessee Dolomite

Compressive Strength at 28 days (psi)

12000
10000
8000

IA/DO/25-02/68
IA/DO/06-06/68

6000

IA/DO/05-00/00
IA/DO/05-00/68

4000
2000
0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.16: Compressive Strength of Mortar of Iowa Dolomite

108

Compressive Strength at 28 days (psi)

12000
10000
8000
AR/SS/06-00/68

6000

AR/SS/06-00/00

4000
2000
0
LF

MF

HF

LV

MV

HV

Type of Mixture

Figure 5.17: Compressive Strength of Mortar of Arkansas Sandstone


The second group decreases in compressive strength when cement content
increases,

including

CT/BA/06-00/00,

PA/LS/05-00/65,

VA/LS/05-00/00,

VA/LS/05-00/65, and SD/QZ/19-02/68. Compressive strengths of the aggregates,


which do not show any correlation, belong to the third group that includes over
two-thirds of the aggregates examined in this study.
Although mixtures with fixed water-cement ratio and low cement content
did not have good workability, their strengths are comparable to those made with
natural sand except aggregates 70-VA/GT/06-00/00, 70-VA/GT/06-00/68, 66VA/GT/13-02/68,

66-VA/GT/06-00/68,

MN/GT/06-00/68,

CT/GT/06-00/68,

IA/DO/05-00/68, and MO/LS/06-00/00 which are all about 25% lower in

109
compressive strength except for 66-VA/GT/06-00/68 which is considerably less.
As a result, if a chemical admixture is used to improve the workability, MFA with
higher contents of micro fines can be acceptable for field performance.
For fixed flow, it is observed that compressive strengths of mortar made
with manufactured sands increase with increasing cement content, and mortars
with natural sand show the same trend. Detailed compressive strength test results
are shown in Appendix C (Table C.2).

5.5.4 Drying Shrinkage


Drying shrinkage is the strain resulting from the loss of moisture from the
specimen. The average values of 28-day drying shrinkage are shown in Table 5.2.
Generally drying shrinkage strain increased with increasing cement content, but
mixtures of aggregates PA/LS/06-00/65, MO/LS/06-00/00, AR/SS/06-00/68, and
AR/SS/06-00/00 with a fixed flow show less shrinkage when cement content
increased. Comparing the mixture with natural sand, Virginia granite, Connecticut
basalt, Virginia diabase, Oklahoma dolomite, and Arkansas sandstone showed
greater 28-day shrinkage strains in general, about 40 percent higher, while 28-day
shrinkage strains of the other aggregates were about 15 percent higher, except for
all the Tennessee dolomite and MN/GT/06-00/00 which actually proved to be
lower. Aggregate MO/LS/06-00/00 shows the greatest difference from the natural
sand, being nearly twice as much as the natural sand. It should be noted that the

110
MBV of those aggregates with higher shrinkage was greater than 1.25, and
aggregate MO/LS/06-00/00 with a MBV of 10 had the highest drying shrinkage
strain, which was twice that of natural sand mortar. Detailed drying shrinkage test
results are shown in Appendix C (Table C.3).

5.6

Results of Other Mortar Testing


Vulcan Materials Company performed a companion study on 70 aggregates

from four different sources and three types of rock. The tabulated results of this
study can be found in Appendix D.

5.7

Summary
After screening and performing aggregate characterization tests, mortar

tests were performed to investigate the effect of MFA on the mortar paste. Flow,
compressive strength, and drying shrinkage were investigated in accordance with
ASTM.
The flows of Pennsylvania and Virginia limestone, Pennsylvania
sandstone, and Tennessee dolomite were higher. On the other hand, Missouri
limestone, Virginia diabase, 66-Virginia granite, and Iowa dolomite had lower
flow that might cause higher water demand.

111
It should be noted that some of the as-received samples had higher flow
than those of product samples. Since the flow test is dynamic instead of static,
micro fines content as well as particle shape affected the flow rate. Hence, it is
concluded that the flow of mortar with fixed water-cement ratio depends on micro
fines content, particle shape, and methylene blue value.
Test results of mixtures with sandstone, limestone, or quartzite show that
they had high compressive strengths. It should be noted that they had low MBV
(0.25 to 1.0).
Although mixtures with fixed water-cement ratio and low cement content
do not have good workability, their strengths are comparable to those made with
natural sand except aggregates VA/GT/06-00/00, VA/GT/06-00/68, MO/LS/0600/00, 66-VA/GT/13-02/68, 66-VA/GT/06-00/68, MN/GT/06-00/68, CT/GT/0600/68, and IA/DO/05-00/68 which are about 25 percent lower in compressive
strength except for 66-VA/GT/06-00/68 which is considerably less. As a result if
a chemical admixture is used to improve the workability, MFA with higher
contents of micro fines can be acceptable for field performance. Figures 5.18 and
5.19 show the comparison of compressive strength for the representative of each
type of aggregate.
Generally drying shrinkage strain increased with increasing cement
content; however mixtures of aggregates PA/LS/06-00/65, MO/LS/06-00/00,
AR/SS/06-00/68, and AR/SS/06-00/00 with a fixed flow rate showed less

112
shrinkage when cement content increased. Comparing the mixture with natural
sand, Virginia granite, Connecticut basalt, Virginia diabase, Oklahoma dolomite
Arkansas sandstone showed greater 28-day shrinkage strains in general, about 40
percent higher, while 28-day shrinkage strains of the other aggregates were about
15 percent higher, except for Tennessee dolomite and MN/GT/06-00/00 which
showed less. Aggregate MO/LS/06-00/00 had the largest difference of all. It
should be noted that the MBV of those aggregates with higher shrinkage was
greater than 1.25, and aggregate MO/LS/06-00/00 with a MBV of 10 had the
highest drying shrinkage strain, which was twice that of natural sand mortar.
Figures 5.20 and 5.21 show the comparison of drying shrinkage for the
representative type of each aggregate.

10000

28-day Compressive Strength (psi)

9000
8000
7000
6000
5000
4000
3000
2000

SS(AR)

DO(IA)

DO(TN)

GT(WY)

GT(CT)

GT(MN)

GT(66-VA)

LS(MO)

SS(PA)

BA

LS(VA)

DO(OK)

DI

QZ

GT(70-VA)

LS(PA)

Nat

1000

Type of Aggregate

113

Figure 5.18: Twenty-eight-day Mortar Compressive Strength for Each Type of Aggregate (Fixed W/C)

10000

8000

6000

4000

SS(AR)

DO(IA)

DO(TN)

GT(WY)

GT(CT)

GT(MN)

GT(66-VA)

LS(MO)

SS(PA)

BA

LS(VA)

DO(OK)

DI

QZ

GT(70-VA)

LS(PA)

2000

Nat

28-day Compressive Strength (psi)

12000

Type of Aggregate

Figure 5.19: Twenty-eight-day Mortar Compressive Strength for Each Type of Aggregate (Fixed Flow)
114

115
0.14

0.12

Drying Shrinkage (%)

0.1

0.08

LS(PA)
GT(70-VA)
QZ
DI

0.06

DO(OK)
LS(VA)
BA
SS(PA)

0.04

LS(MO)
Natural
GT(66-VA)
GT(MN)
GT(CT)

0.02

GT(WY)
DO(TN)
DO(IA)
SS(AR)

0
7

14

28

56

112

Time (days)

Figure 5.20: Drying Shrinkage of Mortar for Each Type of Aggregate


(Fixed W/C)

116

0.14

0.12

Drying Shrinkage (%)

0.1

0.08
LS(PA)
GT(70-VA)
QZ
DI

0.06

DO(OK)
LS(VA)
BA
SS(PA)

0.04

LS(MO)
Natural
GT(66-VA)
GT(MN)
GT(CT)

0.02

GT(WY)
DO(TN)
DO(IA)
SS(AR)

0
7

14

28

56

112

Time (days)

Figure 5.21: Drying Shrinkage of Mortar for Each Type of Aggregate


(Fixed Flow)

You might also like