Anda di halaman 1dari 9

Effect of High-Volume Fly Ash on

Shear Strength of Concrete Beams


Mahdi Arezoumandi, Jeffery S. Volz, and John J. Myers
ash as the finely divided residue that results from the combustion
of ground or powdered coal and that is transported by flue gasses.
ASTM further categorizes fly ash as Class C or Class F, a designation that depends on the type of coal burned in the boiler. Class F ash
is pozzolanic; Class C ash is both pozzolanic and cementitious as a
result of its higher calcium content and the presence of additional
reactive crystalline phases (5). The choice of which class to use
depends on both availability and the need for durability.
Fly ash has been used in the United States since 1930; in 1937
Davis and his colleagues were the first researchers to publish their
results about the use of fly ash in concrete (5). Initially, fly ash was
used in massive structures like the Thames Barrage in the United
Kingdom and the Upper Stillwater Dam in the United States, with
about 30% to 75% mass replacement of hydraulic cement to reduce
heat generation (5). Subsequent research has shown some beneficial
aspects of using fly ash in concrete, such as low permeability and
high durability (610).
Typically, fly ash used in structural concrete as a replacement or
supplementary material is limited to 15% to 25% cement replacement (11, 12). Exceptions include mass concrete applications (such as
foundations and piers) and high-strength concrete, in which replacement levels of portland cement at 35% are common to control peak
hydration temperature development (1315). The means by which
significant amounts of fly ash contribute to the strength development of the concrete and the hydration characteristics of this type
of material are of significant research interest. High-volume fly
ash concrete (HVFAC) is generally defined as a concrete with at
least 50% of the portland cement replaced with fly ash. In 1986, the
Canadian Centre for Mineral and Energy Technology developed
HVFAC for structural applications (16). These investigations and
studies by other researchers have shown that HVFAC has lower
shrinkage, creep, and water permeability and higher modulus of
elasticity than conventional concrete (CC) (1618).
Comprehensive research has been done on the fresh and hardened
properties of HVFAC, but very little research has been performed on
its structural behavior. Koyama et al. used fly ash as a replacement
of fine aggregate in their specimens and concluded that beams with
50% replacement of the fine aggregate had higher shear strength compared with CC (19). However, the beams had shear span-to-depth
values of 1.0, 1.5, and 2.0, classifying them all as deep beams with
respect to shear behavior. Rao et al. studied shear behavior of beams
with 50% Class F fly ash as a cement replacement with four longitudinal reinforcement ratios (ranging between 0.5% and 2.94%) (20).
They reported that the shear strength of HVFAC beams was slightly
lower than CC beams. However, the beam cross section measured
only 4 8 in.
This study presents the results of an investigation that compared the
shear strength of full-scale HVFAC beams and CC beams with shear
provisions of selected codes and a shear database of CC specimens.

Every year concrete production exceeds 6.5 billion tons worldwide,


nearly 1 ton for every man, woman, and child on the planet. In fact,
concrete is used 10 times more than any other building material in the
world, primarily because of its incredible versatility, availability, and
relatively low cost. Unfortunately, the production of portland cement
the key ingredient in concretegenerates approximately a pound of carbon dioxide for every pound of cement produced. One method of reducing
concretes carbon footprint is to replace a significant amount of the
cement with fly ash, a by-product of coal-burning thermal power stations. This paper presents the results of an investigation of the shear
strength of full-scale reinforced concrete beams constructed with both
high-volume fly ash concrete (concrete with at least 50% of the cement
replaced with fly ash) and conventional concrete. A test matrix of 16 beams
included 12 beams without shear reinforcing and four with shear reinforcing in the form of stirrups; eight beams were constructed from
each concrete type. Evaluation of the test results included a comparison with U.S. and international design codes, an examination based on
AASHTO load and resistance factor design provisions of the longitudinal strains and angle of the critical flexure-shear crack, a comparison
with an extensive database of shear tests on conventional concrete, and
a detailed statistical study. The results revealed that the high-volume fly
ash concrete beams had higher shear strengths than the conventional
concrete beams tested in this investigation.

Concrete is the most widely used manufactured material in the world,


and cement is an essential ingredient in the production of portland
cement concrete. The cement industry plays a key global role, both
economically and environmentally. In 2011, world cement output
was estimated at 3.4 billion metric tons (1). Cement production is
also a relatively significant source of global carbon dioxide (CO2)
emissions, accounting for approximately 4.5% of global CO2 emissions from industry in 2007 (2). According to the World Business
Council for Sustainable Development, the amount of cement emissions varies worldwide from 0.73 to 0.99 lb of CO2 for each pound
of cement produced (3).
One of the solutions for this global concern is the use of supplementary cementitious materials to replace cement. The most available supplementary cementitious material worldwide is fly ash, a by-product of
coal-burning thermal power stations (4). ASTM C618-08 defines fly
M. Arezoumandi, Room 212; J.S. Volz, Room 311; and J.J. Myers, Room 325,
Department of Civil, Architectural, and Environmental Engineering, Missouri University of Science and Technology, Butler Carlton Hall, 1401 North Pine Street,
Rolla, MO 65409. Corresponding author: M. Arezoumandi, ma526@mst.edu.
Transportation Research Record: Journal of the Transportation Research Board,
No. 2342, Transportation Research Board of the National Academies, Washington,
D.C., 2013, pp. 19.
DOI: 10.3141/2342-01
1

2

Transportation Research Record 2342

TABLE 1 Shear Beams Test Matrix

TABLE 2 Physical Properties of Cement and Fly Ash

No. and Type of Bar


Section

Property

Bottom
Reinforcement

Top
Reinforcement

4 #7
6 #7
8 #7
8 #7

2 #4
2 #4
2 #7
2 #7

NS-4
NS-6
NS-8
S-8

Stirrup

0.0127
0.0203
0.0271
0.0271

na
na
na
#3 at 7 in.

Fineness
Blaine, m2/kg
+325 mesh (+44 m)
Specific gravity

Component
SiO2
Al2O3
Fe2O3
CaO
MgO
SO3
Na2O
LOI

Experimental Program
Specimen Design

Materials
The cementitious materials used for this study were ASTM Type I
portland cement, ASTM Class C fly ash from the Ameren Labadie

Not measured
14.4%
2.73

Class C Fly Ash (%)

21.98
4.35
3.42
63.97
1.87
2.73
0.52 equivalent
0.60

33.46
19.53
6.28
26.28
5.54
2.40
1.43 equivalent
0.34

Note: SiO2 = silicon dioxide; Al2O3 = aluminum oxide; Fe2O3 =


ferric oxide; CaO = calcium oxide; MgO = magnesium oxide;
SO3= sulfur trioxide; Na2O = sodium oxide; LOI = loss of ignition.

Power Plant, gypsum from USA Gypsum, and calcium hydroxide


from the Mississippi Lime Company. Tables 2 and 3 show the physical properties and chemical compositions, respectively, of the cement
and fly ash.
Crushed limestone with a maximum nominal aggregate size
of 0.75 in. from Jefferson City Dolomite was used as the coarse
aggregate. The fine aggregate was natural sand from Missouri
River Sand.
The longitudinal steel consisted of ASTM A615-12, Grade 60
material. The shear reinforcement was ASTM A615-12, Grade 40 to
ensure a shear failure before a flexural failure.

4 ft.

4 ft.

347
4.1%
3.15

Type I Cement (%)

4 ft.
4 ft.

Class C Fly Ash

TABLE 3 Chemical Composition of Cement and Fly Ash

Note: No. = number; NS = no stirrup; na = not applicable.

Sixteen beams (12 without shear reinforcing and four with shear
reinforcing in the form of stirrups) with three longitudinal reinforcement ratios were designed to preclude flexural failure and
satisfy the minimum and maximum longitudinal reinforcement
requirements of American Concrete Institute (ACI) 318-11. All
beams tested in this program had a rectangular cross section with
a width of 12 in., a height of 18 in., and shear span-to-depth ratios
of 3.0 or greater (Table 1 and Figure 1). The beam designation
included a combination of letters and numbers. NS and S stand for
no stirrups and stirrups, respectively, and the numbers 4, 6, and 8
indicate the number of #7 longitudinal reinforcement bars within
the tension area of the beam section. For example, NS-6 indicates
a beam with no stirrups and six #7 bars within the bottom of the
beam (Table 1).

Type I Cement

4 ft.

#3@7 in.

(a)

4 ft.
#3@7 in.

(b)
2#7

2#4

2#4

8#7

6#7

4#7
NS-4

NS-6
(c)

NS-8 & S-8

FIGURE 1 Cross sections, load pattern, and location of strain gauges in beams (a) without and (b) with stirrups in test region and
(c) strain gauge (#3 U.S. 5 #10 SI; #4 U.S. 5 #13 SI; #7 U.S. 5 #22 SI; 1 ft 5 304.8 mm).

Arezoumandi, Volz, and Myers

TABLE 4 Mixture Proportions of Concrete

Mix
CC
HVFAC

Water
(lb/yd3)

Cement
(lb/yd3)

Fly Ash
(lb/yd3)

Fine
Aggregate
(lb/yd3)

Coarse
Aggregate
(lb/yd3)

Gypsum
(lb/yd3)

Calcium
Hydroxide
(lb/yd3)

HRWR
(fluid oz)

226
226

565
155

na
360

1,240
1,240

1,860
1,860

na
16

na
39

15.5
15.5

Note: HRWR = high-range water reducer; 1 lb/yd3 = 0.593 kg/m3; 1 fluid oz = 33.814 L/m3.

Mixture Proportions

Test Setup and Procedure

The concrete mixtures with a target compressive strength of 4,000 psi


were delivered by a local ready-mix concrete supplier. The purpose
of using the ready-mix supplier was to validate the HVFAC concept
in actual concrete production runs. The mixture proportions are given
in Table 4. The HVFAC mix used a 70% replacement of cement with
fly ash. Gypsum was used in the HVFAC mix to maintain the initial hydration stage by preventing sulfate depletion, and the calcium
hydroxide ensured a more complete hydration of the fly ash with the
low content of cement in the mix (21). The drums were charged at
the ready-mix facility with the required amounts of cement, fly ash,
sand, coarse aggregate, and water. The powder activators (gypsum
and lime) were added when the truck arrived at the lab, approximately
5 min. later. After the gypsum and lime were added, the HVFAC was
mixed at high speed for 10 min. All of the constituents for the CC
were added to the CC mixes at the ready-mix facility.

A load frame was assembled and equipped with two 110-kip servo
hydraulic actuators intended to apply the two-point loads to the beams.
The load was applied using a displacement control method with a
rate of 0.02 in./min. The shear beams were supported on a roller and
a pin support, 1 ft from each end of the beam, creating a four-point
loading situation with the two actuators. Linear variable differential
transformers and strain gauges were used to measure the deflection at
the beam center and strain in the reinforcement. The strain gauges
were installed on the lower layer of the bottom longitudinal reinforcement at midspan (maximum flexural moment location) and quarterpoint along the span (middle of the shear test region). For the sections
with stirrups, 10 additional strain gauges were installed on the stirrups. Figure 1 shows both the beam loading pattern and the location
of the strain gauges. During the test, any cracks that formed on the
surface of the beam were marked at 5-kip load increments, and both
the deformation and strains were monitored until the beam reached
failure.

Fabrication and Curing of Test Specimens


Specimens were constructed, cured, and tested in the laboratory.
After casting, the beam specimens and the quality control and quality
assurance companion cylinders (ASTM C39-12 and C496-11) and
beams (ASTM C78-10) were covered with wet burlap and a plastic
sheet. All of the beams and companion cylinders were moist cured
for 3 days and, after formwork removal, were stored in the laboratory
until they were tested.
Fresh and Hardened Properties
Table 5 presents the fresh and hardened strength properties of the
CC and HVFAC mixes.

Property

CC

HVFAC

Slump (in.)
Air content (%)
Unit weight (lb/ft3)
Split cylinder strengtha (psi)
Compressive strengtha (psi)

4.5
5.5
143.9
420
4,200

5.5
3.5
152.9
405
4,450

Note: 1 lb/ft = 16.02 kg/m ; 1 psi = 0.00689 MPa.


a
Values represent the average of three cylinders (ASTM
C39-12 and C496-11).
3

Table 6 summarizes the compressive strength at time of testing, shear


force at failure (Vtest), average shear stress at failure (Vtest/bwd), ratio
of the average shear stress to compressive strength (vtest/f c), and ratio
of the average shear stress to square root of the compressive strength
(vtest/f c). The average shear stress of the CC beams varied from 3.4%
to 4.8% of the compressive strength, and the average shear stress of the
HVFAC beams varied from 3.6% to 8.5% of the compressive strength.
It is useful to compare the far-right column in Table 6 with ACI 318-11
Equation 11-3, rewritten in terms of average shear stress for normalweight concrete and shown as Equation 1. The ratio of experimental
shear stress to the square root of compressive strength for the beams
without stirrups exceeded the ACI value of 2.0 for all of the beams.
vc = 2 fc

TABLE 5 Fresh and Hardened


Concrete Properties

Test Results and Discussion

(1)

where vc is the nominal shear stress provided by concrete, and f c


is the specified compressive strength of concrete (both values are
given in pounds per square inch).
General Behavior: Cracking and Failure Mode
In terms of crack morphology, crack progression, and load-deflection
response, the behavior of the HVFAC and CC beams was virtually
identical. All of the beams failed in shear. For the beams without
shear reinforcing, failure occurred when the inclined flexure-shear

4

Transportation Research Record 2342

TABLE 6 Test Results Summary


Section

f c
(psi)

V test
(kips)a

vtest = V test /bw d


(psi)

vtest /f c
(%)

an increase in aggregate interlock, and a larger concrete compression


zone as a result of a downward shift of the neutral axis.
vtest/f c

Comparison of Test Results with Shear


Provisions of Applicable Standards

CC
NS-4
1
2
NS-6
1
2
NS-8
1
2
S-8
1
2

4,200
3,840

26.9
25.6

142.8
135.9

3.4
3.5

2.2
2.2

4,200
3,840

34.5
32.5

194.9
183.6

4.6
4.8

3.0
3.0

4,200
3,840

33.2
32.3

187.5
182.5

4.5
4.8

2.9
2.9

4,400
4,400

67.4
71.9

380.8
406.2

na
na

na
na

4,450
3,000

30.2
27.6

160.3
146.5

3.6
4.9

2.4
2.7

4,450
3,000

33.8
37.8

191.0
213.6

4.3
7.1

2.9
3.9

4,450
3,000

36.5
45.3

206.2
255.9

4.6
8.5

3.1
4.7

5,030
5,030

73.9
75.8

417.5
428.5

na
na

na
na

HVFAC
NS-4
1
2
NS-6
1
2
NS-8
1
2
S-8
1
2

Note: 1 kip = 4.448 kN.


a
Includes part of the load frame not registered by the load cells and also the beam
self-weight at a distance d from the interior face of the support plate.

crack penetrated to the compression zone of the beam near the loading plate before yielding of the longitudinal reinforcement (Figure 2).
For the beams with shear reinforcing, failure occurred when the stirrups crossing the critical flexure-shear crack reached yield. According to data collected from the strain gauges, none of the longitudinal
reinforcement reached yield at failure, but all of the stirrups yielded.
Crack progression in the beams began with the appearance of flexural cracks in the maximum moment region, followed by additional
flexural cracks forming between the load and support regions as the
load was increased. As the applied load was increased, the majority
of the flexural cracks developed vertically, and after that, inclined
flexure-shear cracks began to appear. As the load increased further,
the inclined cracks progressed both upward toward the applied load
plate and downward along the longitudinal reinforcement toward
the support. Figure 2 offers a direct visual comparison of the crack
shape and distribution at failure for both the HVFAC and CC beams,
which are indistinguishable from each other.
Figure 3 shows the load-deflection behavior for the beams with different longitudinal reinforcement ratios (the deflection was measured
at midspan). Before the first flexural cracks occurred (Point A),
all of the beams displayed a steep linear elastic behavior. After additional application of load, the beams eventually developed the critical
flexure-shear crack, which resulted in a drop in load and redistribution of the internal shear (e.g., Point B). After this redistribution, the
beams were able to support additional load until reaching failure. As
expected, sections with a higher percentage of longitudinal reinforcement had a higher shear capacity, which can be attributed to a combination of additional dowel action (22), tighter shear cracks and thus

The experimental shear strengths of the beams were compared with


the shear provisions of standards from AASHTO (LRFD-10), ACI
(318-11), Australia (AS 3600-09), the Canadian Standards Association
(CSA-04), Europe (Eurocode 2-05), and the Japan Society of Civil
Engineers (JSCE-07). Most design codes consider the shear strength
of reinforced concrete beams as consisting of two parts (a concrete
contribution and a steel contribution), with the total capacity of the
section equal to the sum of these two values. Although the actual
behavior consists of a complex interaction of these aspects, the codes
treat each contribution independently, with one equation given for the
concrete contribution and another equation for the steel contribution.
For this comparison of test results to code-predicted values, all of
the safety factors of the standards were set equal to one, and all ultimate moments and shear forces were calculated without load factors.
Table 7 presents the ratio of experimental predicted capacity to
code-predicted capacity (Vtest/Vcode) for each of the selected design
standards. In general, for a given standard, the ratios for the HVFAC
beams are higher than the CC beams with the same amount of longitudinal reinforcement. Overall, for the beams without stirrups,
the ratios range from 0.91 to 1.41 and 1.06 to 2.15 for the CC and
HVFAC beams, respectively. With regard to the ratios that fell below
1.0 (an unconservative result), most standards do not allow sections
without stirrups unless the factored shear force is significantly less
than the concrete capacity in shear. For the beams with stirrups, the
ratios were in better agreement between the two concrete types, most
likely as a result of the greater predictability of the stirrup capacity
portion of the shear strength. Ratios range from 1.13 to 1.41 and 1.24
to 1.45 for the CC and HVFAC beams, respectively.
Comparison of Test Results
with Shear Test Database
Figure 4 presents the normalized shear strength versus longitudinal
reinforcement ratio for the beams of this study and the wealth of shear
test data available in the literature (23). Given the significant scatter of
the database shear test results, it is difficult to draw definitive conclusions on the current test values. Nonetheless, visually, Figure 4 seems
to indicate that the HVFAC test results fall within the central portion of
the data and follow the same general trend of increasing shear strength
as a function of the longitudinal reinforcement ratio. Furthermore, statistical analysis of the data indicates that the test results fall within
a 95% confidence interval of a nonlinear regression curve fit of the
database. This result indicates that the test values are consistent with
the shear test data available in the literature, even given the significant
scatter, and that the results do indeed indicate that for the specimens
tested, the HVFAC shear strength is higher than the CC shear strength.
Statistical Data Analysis
Parametric and nonparametric statistical tests were used to evaluate
whether there is any statistically significant difference between the
shear strength of the HVFAC and CC beams.

Arezoumandi, Volz, and Myers

1 ft.

4 ft.

4 ft.

4 ft.

1 ft.

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)
FIGURE 2 Crack pattern of beams on shear failure: (a) HVFAC-NS-4, (b) CC-NS-4,
(c) HVFAC-NS-6, (d) CC-NS-6, (e) HVFAC-NS-8, (f) CC-NS-8, (g) HVFAC-S-8, and (h) CC-S-8.

Transportation Research Record 2342

160

160

140

140

120

120
Load (kips)

Load (kips)

6

100
80
60

NS-4-1
NS-6-1
NS-8-1
S-8-1

40
20
0

A
0

0.2

0.4

0.6

0.8

100
80
60
20
0

NS-4-2
NS-6-2
NS-8-2
S-8-2

40
A
0

0.2

Deflection (in.)
(a)

0.4

0.6

0.8

Deflection (in.)
(b)

FIGURE 3 Load deflections of (a) CC and (b) HVFAC beams.

Parametric Test
The paired t-test is used to compare two population means. This
test assumes that the differences between pairs are normally distributed. If this assumption is violated, the paired t-test may not
be the most powerful test. The hypothesis for the paired t-test is
as follows:

Ho. The means of the shear capacity values of the HVFAC beams
are higher than those of the CC beams and
Ha. Not Ho.
The statistical computer program SAS 9.2 was employed to
perform these statistical tests. Both KolmogorovSmirnov and
AndersonDarling tests showed the data (the differences between

TABLE 7 V test/V code for Selected Codes


Section

AASHTO

ACI

AS-3600

CSA

Eurocode 2

JSCE

CC
NS-4
1
2
NS-6
1
2
NS-8
1
2
Avg.
CV (%)
S-8
1
2
Avg.
CV (%)

0.93
0.91

1.04
1.02

1.04
1.02

0.94
0.91

0.96
0.94

1.18
1.15

1.19
1.15

1.41
1.38

1.19
1.16

1.20
1.15

1.11
1.08

1.35
1.31

1.02
1.03
1.04
10.87

1.33
1.34
1.25
14.02

1.04
1.04
1.08
6.68

1.03
1.03
1.04
10.86

1.07
1.07
1.04
6.78

1.18
1.18
1.23
6.80

1.20
1.31
1.25
6.30

1.32
1.41
1.36
4.70

1.16
1.24
1.20
4.57

1.13
1.23
1.18
6.00

1.21
1.29
1.25
4.57

1.30
1.39
1.35
4.57

1.07
1.14

1.14
1.25

1.15
1.19

1.08
1.15

1.06
1.10

1.30
1.35

1.13
1.60

1.34
1.82

1.14
1.46

1.13
1.61

1.07
1.36

1.30
1.65

1.12
1.84
1.32
24.49

1.43
2.15
1.52
25.36

1.12
1.59
1.28
15.63

1.13
1.85
1.33
24.42

1.16
1.64
1.23
18.56

1.27
1.80
1.45
15.47

1.32
1.37
1.35
2.63

1.40
1.44
1.42
1.99

1.24
1.27
1.26
1.69

1.24
1.28
1.26
2.24

1.32
1.35
1.34
1.59

1.42
1.45
1.44
1.48

HVFAC
NS-4
1
2
NS-6
1
2
NS-8
1
2
Avg.
CV (%)
S-8
1
2
Avg.
CV (%)

Note: Avg. = average; CV = coefficient of variation.

Arezoumandi, Volz, and Myers

0.60

Database

Vtest / (bwd fc)

0.50

CC
HVFAC

0.40
0.30
0.20

Fit

0.10
0.00

L 95
U 95

Longitudinal reinforcement ratio (l %)


FIGURE 4 Shear strength versus longitudinal reinforcement ratios [Reineck et al.
(23) and present study].

the shear capacities of the HVFAC and CC beams) followed a normal distribution, which meant that paired t-tests could be used. The
result of the paired t-test showed that the p-value was .96 (>.05),
which confirmed the null hypothesis at the .05 significance level. In
other words, the means of the shear capacity of the HVFAC beams
are statistically higher than the means of the CC beams.
Nonparametric Test
Unlike parametric tests, nonparametric tests are referred to as
distribution-free tests. These tests have the advantage of requiring no assumption of normality, and they usually compare medians rather than means. The Wilcoxon signed-rank test is usually
identified as a nonparametric alternative to the paired t-test. The
hypothesis for this test is the same as that for the paired t-test. The
Wilcoxon signed-rank test assumes that the distribution of the difference of pairs is symmetrical. This assumption can be checked;
if the distribution is normal, it is also symmetrical. The data follow a normal distribution, and the Wilcoxon signed-rank test can
be used. The p-value for the Wilcoxon signed rank was .99 (>.05),
which confirmed the null hypothesis at the .05 significance level.
The p-values for both the paired t-tests (parametric test) and the
Wilcoxon signed rank test (nonparametric test) are very close to
each other.
Overall, the statistical data analyses showed that the HVFAC
beams had higher shear capacity than the CC beams.
Comparison of Test Results with AASHTO Load
and Resistance Factor Design
According to the AASHTO load and resistance factor design standard
(LRFD-10 standard), strain in the longitudinal tension reinforcement
can be determined by
Mu

d + Vu
v
s =
Es As
where
As = area of nonprestressed tension reinforcement,
dv = effective shear depth,
Es = modulus of elasticity of reinforcing bars,

(2)

Mu = factored moment at section,


Vu = factored shear force at section, and
s = strain in nonprestressed longitudinal tension reinforcement.
Table 8 presents the tensile strain in the longitudinal tension reinforcement at the quarter-point of the span (middle of the shear test
region) obtained from both the experiments (strain gauges) and the
AASHTO LRFD-10 equation. The AASHTO LRFD-10 equation
estimates the strain for both the HVFAC and CC beams very well
for low and medium reinforcement ratios (NS-4 and NS-6), but it
underestimates the strain for the sections with higher reinforcement
ratios (NS-8 and S-8).
The angle of the critical shear crack is an important design parameter in the AASHTO LRFD sectional design method, although it is
difficult to determine precisely as it is open to interpretation. The
procedure used to determine this angle consisted of measuring the
angle of a portion of the critical shear crack between two reference
points, with the points corresponding to immediately after crossing
the alignment of the longitudinal reinforcement and before entering
the compression zone (Figure 5). The diagonal shear crack angle is
calculated in AASHTO LRFD by Equation 3:
= 29 + 3,500 s

(3)

where is the angle of inclination of diagonal compressive stresses


(degrees).
Table 8 compares the measured diagonal shear crack angle from
the test specimens with the calculated angle from the AASHTO
LRFD equation. As Table 8 shows, the AASHTO LRFD equation
accurately predicted the diagonal shear crack angle for both the CC
and HVFAC beams without a stirrup, but it overestimated for beams
with a stirrup.
Conclusions
To evaluate the shear strength of chemically based HVFAC, 16
full-scale beams (eight for each concrete type) constructed with different longitudinal reinforcement ratios were tested to failure. The
following conclusions are presented on the basis of the results of
this study:
In terms of crack morphology, crack progression, and loaddeflection response, the behavior of the HVFAC and CC beams was
virtually identical.

8

Transportation Research Record 2342

TABLE 8 Comparison of Longitudinal Reinforcement Strain and Diagonal Shear Crack Angle from Experiment and AASHTO LRFD Equations
CC

Section
NS-4
1
2
NS-6
1
2
NS-8
1
2
S-8
1
2
Avg.

HVFAC

s quarter-point
Equationa

s quarter-point
Experimenta

CC

HVFAC

s -Eq.
s -Ex.

s quarter-point
Equationa

s quarter-point
Experimenta

s -Eq.
s -Ex.

AASHTO
Equation

measured
Experiment

AASHTO
measured

AASHTO
Equation

measured
Experiment

AASHTO
measured

1,004
954

844

1.13

1,127
1,029

1,211
730

0.93
1.41

34.7
34.4

40
34

0.87
1.01

35.4
34.8

36
45

0.98
0.77

892
840

989
906

0.90
0.93

875
977

943
1,148

0.93
0.85

34.1
33.8

41
35

0.83
0.97

34.0
34.6

35
35

0.97
0.99

645
626

726
818

0.89
0.77

707
878

780
1,483

0.91
0.59

32.7
32.6

40
29

0.82
1.12

33.0
34.0

35
34

0.94
1.00

1,305
1,392

1,648
1,791

0.79
0.78
0.88

1,431
1,468

1,700
1,847

0.84
0.79
0.91

36.5
37.0

27
33

1.35
1.12
1.01

37.2
37.4

27
28

1.38
1.33
1.04

Note: = not usable data.


a
Microstrain.

In general, for a given standard, the ratios of experimental


predicted capacity to code-predicted capacity are higher for the
HVFAC than for the CC beams with the same amount of longitudinal
reinforcement.
Statistical data analyses (parametric and nonparametric) showed
that the HVFAC beams had higher shear strength than the CC beams
tested in this study.
The HVFAC and CC test results fall within a 95% confidence
interval of a nonlinear regression curve fit of the CC shear test
database.
From a comparison of the test results of this study with the
shear database available for CC specimens, it can be inferred the
HVFAC beams had greater shear capacity than the CC beams.
The AASHTO LRFD estimations of the longitudinal tensile
strain of the reinforcements were close to the actual strain for the
low and medium longitudinal reinforcement ratio beams. However,
the AASHTO LRFD equations underestimated the strain of high
longitudinal reinforcement ratio beams.
The AASHTO LRFD equation predicts the diagonal shear crack
angle of both the CC and HVFAC beams very well for beams with-

FIGURE 5 Diagonal shear crack angle.

out shear reinforcement, but it overestimates the values for beams


with shear reinforcement.
Because of the limited nature of the data set regarding aspect
ratio, mix designs, and aggregate type and content investigated, the
researchers recommend further testing to increase the database of
test results.
Acknowledgments
The authors gratefully acknowledge the financial support provided
by the Missouri Department of Transportation and the National University Transportation Center at Missouri University of Science and
Technology.
References
1. Minerals Yearbook: Cement. U.S. Geological Survey, U.S. Department
of the Interior, 2012, pp. 3839.
2. Marland, G., T.A. Boden, and R.J. Andres. Global, Regional, and
National Fossil Fuel CO2 Emissions. In Trends: A Compendium of Data on
Global Change, Carbon Dioxide Information Analysis Center, Oak Ridge
National Laboratory, U.S. Department of Energy, Oak Ridge, Tenn., 2008.
http://cdiac.ornl.gov/trends/emis/overview.html. Accessed March 2012.
3. Hanle, L., K. Jayaraman, and J. Smith. CO2 Emissions Profile of the
U.S. Cement Industry. 2012. http://infohouse.p2ric.org/ref/43/42552.
pdf. Accessed March 2012.
4. Bilodeau, A., and V.M. Malhotra. High-Volume Fly Ash System: Concrete Solution for Sustainable Development. ACI Materials Journal,
Vol. 97, No. 1, 2000, pp. 4148.
5. ACI Committee 232. Use of Fly Ash in Concrete (ACI 232.2R-03).
American Concrete Institute, Farmington Hills, Mich., 2003.
6. Dunstan, E. Fly Ash and Fly Ash Concrete. Report No. REC-ERC-82-1.
Bureau of Reclamation, U.S. Department of the Interior, Denver, Colo.,
1984.
7. Dunstan, E.R., Jr. Performance of Lignite and Sub-Bituminous Fly Ash
in Concrete. Report No. REC-ERC-76. Bureau of Reclamation, U.S.
Department of the Interior, Denver, Colo., 1976.
8. Dunstan, E.R., Jr. A Possible Method for Identifying Fly Ashes That
Will Improve the Sulfate Resistance of Concretes. Cement, Concrete
and Aggregates, Vol. 2, No. 1, 1980, pp. 2030.

Arezoumandi, Volz, and Myers

9. Dunstan, M.R.H. Rolled Concrete for Dams: A Laboratory Study of


High Fly Ash Concrete. Technical Note No. 105. Construction Industry
Research and Information Association, London, 1981.
10. Dunstan, M.R.H. Rolled Concrete for Dams: Construction Trials
Using High Fly Ash Content Concrete. Technical Note No. 106. Construction Industry Research and Information Association, London,
1981.
11. Berry, E.E., R.T. Hemmings, M.H. Zhang, B.J. Cornelious, and D.M.
Golden. Hydration in High-Volume Fly Ash Binders. ACI Materials
Journal, Vol. 91, No. 4, 1994, pp. 382389.
12. ACI Committee 211. Guide for Selecting Proportions for High-Strength
Concrete with Portland Cement and Fly Ash: ACI 226.4R. ACI Materials
Journal, Vol. 90, No. 3, 1993, pp. 272283.
13. Myers, J.J., and R.L. Carrasquillo. Mix Proportioning for High-Strength
HPC Bridge Beams. Special Publication 189. American Concrete Institute,
Detroit, Mich., 1999, pp. 3756.
14. Poon, C.S., L. Lam, and Y.L. Wong. A Study on High Strength Concrete Prepared with Large Volumes of Low-Calcium Fly Ash. Cement
and Concrete Research, Vol. 30, No. 3, 2000, pp. 447455.
15. Li, G. Properties of High-Volume Fly Ash Concrete Incorporating
Nano-SiO. Cement and Concrete Research, Vol. 34, No. 6, 2004,
pp. 10431049.
16. Malhotra, V.M. Superplasticized Fly Ash Concrete for Structural
Applications. Concrete International, Vol. 8, No. 12, 1986, pp. 2831.
17. S. Li, D.M. Roy, and A. Kumer. Quantitative Determination of
Pozzolanas in Hydrated System of Cement or Ca(OH)2 with Fly Ash

or Silica Fume. Cement and Concrete Research, Vol. 15, No. 6, 1985,
pp. 10791086.
18. Gopalan, M.K. Nucleation and Pozzolanic Factors in Strength Development of Class F Fly Ash Concrete. ACI Materials Journal, Vol. 90,
No. 2, 1993, pp. 117121.
19. Koyama, T., Y.P. Sun, T. Fujinaga, H. Koyamada, and F. Ogata.
Mechanical Properties of Concrete Beam Made of a Large Amount of
Fine Fly Ash. Proc., 14th World Conference on Earthquake Engineering,
Beijing, 2008. http://www.iitk.ac.in/nicee/wcee/article/14_05-03-0040.
PDF. Accessed March 2012.
20. Rao, R.M., S. Mohan, and S.K. Sekar. Shear Resistance of HighVolume Fly Ash Reinforced Concrete Beams Without Web Reinforcement. International Journal of Civil and Structural Engineering, Vol. 1,
No. 4, 2011, pp. 986993.
21. Bentz, D.P. Powder Additions to Mitigate Retardation in High-Volume
Fly Ash Mixtures. ACI Materials Journal, Vol. 107, No. 5, 2010,
pp. 508514.
22. Taylor, H.P.J. Investigation of the Forces Carried Across Cracks in Reinforced Concrete Beams in Shear by Interlock of Aggregate. Technical
Report 42.447. Cement and Concrete Association, London, 1970.
23. Reineck, K.H., D.A. Kuchma, K.S. Kim, and S. Marx. Shear Database
for Reinforced Concrete Members Without Shear Reinforcement. ACI
Structural Journal, Vol. 100, No. 2, 2003, pp. 240249.
The Basic Research and Emerging Technologies Related to Concrete Committee
peer-reviewed this paper.

Anda mungkin juga menyukai