Sunteți pe pagina 1din 12

ACI MATERIALS JOURNAL

TECHNICAL PAPER

Title no. 101-M04

Properties of Polyester Polymer Concrete with Glass and


Carbon Fibers
by Kallol Sett and C. Vipulanandan
In this study, the effect of adding glass and carbon fibers on the
compressive and tensile behavior of polyester polymer concrete
(PC) was investigated. The compressive and tensile strengths of
the optimum PC were 55 and 7 MPa, respectively. PC systems with
fibers were optimized based on the workability and mechanical
properties. Fibers improved the properties of the PC system to
varying degrees, which depended on the type and amount of fibers.
The tensile-to-compressive modular ratio for the PC was 0.75.
Strength, stiffness, and stress-strain models were used to predict
the observed behavior of PC. The pulse velocity method and the
impact resonance method were sensitive to determine the changes
in nondestructive properties of PC systems with the addition of
fibers. The pulse velocity method predicted the static moduli more
closely than the impact resonance method. The damping properties
of polymer concrete systems have been quantified.
Keywords: fiber; glass fiber; Poissons ratio; polymer concrete; test.

INTRODUCTION
Polymer concrete (PC) is a mixture of aggregate and
monomer that hardens through polymerization of the
monomer. Due to its rapid setting, high strength-to-density
ratio, and ability to withstand corrosive environment, PC is
increasingly being used as an alternate to cement concrete in
many applications.1 To minimize material cost, it is imperative
to use the least possible amount of polymer in PC formulations
to achieve desired properties depending on its applications. This
has led to the study of the effect of fiber addition in the PC
system. Steel fibers are one of the most popular fibers for matrix
reinforcement. Ohama and Nishimura2 studied the effect of steel
fibers and reported increases in compressive, flexural and impact
strength. Glass fibers are also very popular because of their close
proximity (chemically) to fine aggregates (sand) used in PC
systems. Mebarkia and Vipulanandan3 and Vipulanandan
and Dharmarajan 4 studied the effect of glass fibers and
reported compressive strength enhancement. Carbon
fibers are increasingly becoming popular among
researchers because of their high tensile strength, high
tensile modulus, and unique characteristics like electrical
conductivity and thermal insulation.
In this study, the mechanical properties of carbon fiberreinforced polymer concrete (CFRPC) were investigated and
the performance was compared with glass fiber systems
(GFRPC). The fiber content varied up to 6% by weight. The
applicability of two popular nondestructive testing methods,
the impact resonance method and the pulse velocity method,
was studied to characterize the PC systems.5-7 Both the tests
were performed to determine dynamic Youngs modulus,
dynamic shear modulus, and dynamic Poissons ratio. In
both nondestructive methods, the effects of the shape of the
specimens on the dynamic properties of the PC system were
studied. Damping properties of the PC system were also
30

studied by varying the fiber and polymer contents for their


potential application in dynamic loading design.
OBJECTIVE
The objective of this study is to 1) investigate the effect of
glass and carbon fibers on the compressive and tensile
behavior of the PC system; 2) verify the applicability of
strength, stiffness, and stress-strain models to predict the
behavior of PC, GFRPC, and CFRPC systems; 3) determine
the sensitivity of nondestructive test methods to characterize
the compressive and tensile properties of the PC system with
and without fibers; and 4) determine the nondestructive
properties, low strain moduli and damping ratios, of the PC
system with and without fibers.
RESEARCH SIGNIFICANCE
This study establishes a basis to better understand the
effect of glass and carbon fibers on the mechanical and
nondestructive properties of PC systems. PC systems with
fibers were optimized based on the mechanical properties
and workability. The suitability of previously published
strength, stiffness, and stress-strain models to predict the
behavior of different PC systems was investigated. The
applicability of the pulse velocity and the impact resonance
method to characterize the polymer concrete systems with
fibers was investigated.
MATERIALS AND METHODS
Materials
The compositions of the PC system are summarized in
Table 1. The isopthalic thermosetting polymer resin used in
this study had a viscosity of 40 to 50 poise (4 to 5 Pa-s) at
room temperature and a specific gravity of 1.07. The sand
was constituted by mixing five grades of commercially
available blasting sand of equal weight.8-11 The sand was
well graded and had a coefficient of uniformity Cu of 5.8 and
a coefficient of concavity Cc of 0.9. Based on manufacturers
literature, glass fibers 6.35 mm in length and 13 m in diameter
had a tensile strength of 2500 MPa, tensile modulus of 70 GPa,
and density of 2550 kg/m3. Polyacrylonitrile (PAN) based
chopped carbon fibers 6.35 mm in length and 7.2 m in
diameter were used in this study. Carbon fibers had a tensile
strength of 3800 MPa, tensile modulus of 228 GPa, and
density of 1810 kg/m3. The carbon fibers were 50%
stronger and more than 200% stiffer than the glass fibers but
ACI Materials Journal, V. 101, No. 1, January-February 2004.
MS No. 02-404 received October 16, 2002, and reviewed under Institute publication
policies. Copyright 2004, American Concrete Institute. All rights reserved, including
the making of copies unless permission is obtained from the copyright proprietors.
Pertinent discussion including authors closure, if any, will be published in the NovemberDecember 2004 ACI Materials Journal if the discussion is received by August 1, 2004.

ACI Materials Journal/January-February 2004

ACI member Kallol Sett is an MS student in the Department of Civil and Environmental
Engineering, University of Houston, Houston, Tex. He received his BS in civil engineering
from Jadavpur University, Calcutta, India.
ACI member C. Vipulanandan is Chairman and Professor of the Department of Civil
and Environmental Engineering, University of Houston. He is also Director of the
Center of Innovative Grouting Materials and Technology (CIGMAT) at the University of
Houston. He received his PhD and MS from Northwestern University, Evanston, Ill. He is a
member of ACI Committees 446, Fracture Mechanics, and 548, Polymers in Concrete.

the aspect ratio of glass fibers were 80% lesser than that of
carbon fibers. Also, SEM and visual inspection showed rougher
surface texture of glass fibers compared with carbon fibers.
In preparing the PC specimens, cobalt napthanate was first
added to the polyester resin and the solution was mixed
thoroughly; methyl ethyl ketone peroxide was then added.
After further mixing, sand and fibers were added slowly and
mixed long enough to obtain a uniform mixture.12 Teflon
molds were used to cast the specimens. Cylindrical specimens,
100 mm in length and 38 mm in diameter, were used for
destructive compression tests, and 200 mm in length and
60 mm in diameter were used for nondestructive tests. The
prism specimens used for the nondestructive tests were 300
x 50 x 50 mm. For the direct tensile tests, dog bone-shaped
specimens were used with proper metal grips13 (for detailed
dimensions of the tension test specimen refer to CIGMAT
PC-214 ). Specimens were cured at room temperature
(about 25 C) for 24 h followed by 80 C for 24 h in an oven.
Test procedures and principles
To optimize the composition of PC based on workability
and performance, various resin contents were investigated.
Resin content varied up to 20% by weight of composite.3 For
fiber-reinforced polymer concrete (FRPC) systems, specimens
were prepared with different fiber content. Fiber content was
varied up to 6% by weight of polymer concrete (polymer +
sand) for both GFRPC and CFRPC systems. But it should be
noted that the volume fraction of carbon fibers in the CFRPC
system was 1.5 times more than that of glass fibers in a
comparable GFRPC system.
For the FRPC systems, resin contents were chosen based
on workability. The resin contents chosen for this study were
18 and 20% (by weight of composite) for GFRPC and
CFRPC, respectively.
Destructive tests
At least three cylindrical specimens were tested under
uniaxial compression to obtain the mechanical properties
under each condition. Specimens were loaded using a
2000 kN-capacity universal testing machine at a displacement rate of 0.05 mm/min (CIGMAT PC-5).13,15-17 Direct
tension tests were performed using a 50 kN-capacity screwtype machine at a displacement rate of 0.05 mm/min
(CIGMAT PC-2).14 Commercially available 12 mm strain
gages having a least-count of 106 strain/strain were used to
measure strains during compressive and tensile loading.
Strain gages were attached to the specimens directly. Test
data was analyzed to obtain the compressive and tensile
strengths, moduli, and stress-strain relationships.
Nondestructive tests
Impact resonance testImpact resonance tests were
performed as per ASTM C 215.5 The test specimens were
made to vibrate as a whole in one of their natural frequency
modes: transverse, longitudinal, or torsional.7,18-20 The
ACI Materials Journal/January-February 2004

corresponding fundamental frequencies of these modes were


obtained by proper location of the support, impact point, and
accelerometer, and the response was monitored using a
dynamic signal analyzer (HP 35665 A). The accelerometer
used in this study had a range of 500 g and a reference sensitivity
of 10.4 mV/g; it weighed 2.5 g and had a resonant frequency
of 54 kHz. The operating range of the frequency response
was 2 Hz to 10 kHz with an error of 5%. The accelerometer was
mounted on a small steel disc and glued to the specimen surface
using grease.7 The specimen was supported at nodes on small
wooden blocks and struck by a standard hammer. A plotter was
used to plot the output from the dynamic signal analyzer.
Ultrasonic pulse velocity testElastic, homogeneous, and
isotropic material can be characterized by two material
constants. These two elastic moduli control the velocity of
the P-wave and the shear wave.7,21 The deformation
resulting from a shear wave is represented in Fig. 1(a). The
only deformation in this case is shear deformation , so the shear
modulus G controls the wave velocity. The equation relating
shear wave velocity Vs and shear modulus is as follows
2

G = V S

(1)

where is mass density of the material. The deformation


caused by a P-wave is represented in Fig. 1(b), the deformation
is axial , and the wave velocity is controlled by constrained
modulus M. The equation relating constrained modulus and
P-wave velocity Vp s as follows
2

M = V S

(2)

For an elastic, homogeneous and isotropic material

Fig. 1Deformation during: (a) shear wave; and (b)


constrained P-wave.
Table 1Composition of PC, GFRPC, and CFRPC
Relative proportions (by weight)

Constituent
material

Function

PC

Polyester
resin

Binder

Up to 20%
*

GFRPC

CFRPC

18%

20%

MEKPO

Initiator

1.5%

1.5%

1.5%*

CN

Promoter

0.3%*

0.3%*

0.3%*

Blasting sand

Aggregate

Up to 86%

82%

Glass fiber

Microreinforcement

Up to 6%

Carbon fiber

Microreinforcement

80%

Up to 6%

By weight of polymer.
By weight of (polymer + blasting sand).

31

(1 v )
M = ------------------------------------- E
( 1 + v ) ( 1 2v )

(3)

where E is the Youngs modulus. At small strain levels, the


PC system can be assumed to be linearly elastic, homogeneous,
and isotropic, and the dynamic Poissons ratio of different
PC systems were calculated by combining Eq. (1) and (3) as
Vp
2( 1 v)
----- = ------------------( 1 2v )
Vs

12

(4)

and Youngs modulus (E) as

( 1 + v ) ( 1 2v ) 2
E = ------------------------------------- V p
(1 v)

(5)

Pulse velocities for the PC systems were measured using a


commercially available portable V-meter. Lead zirconate
titanate ceramic transducers having a natural frequency of
150 kHz were used to pass longitudinal or shear waves
through the specimens. Commercially available grease was
used to provide good coupling between the specimens and
the transducers. The travel time of the ultrasonic pulse
through the specimens under direct transmission (with the

transducers on opposite faces along the length) was recorded


up to an accuracy of 0.1 s. Pulse velocities (either P-wave
or shear wave) were calculated by dividing the length of the
specimen by the travel time and also the velocity ratio
defined as the ratio of shear wave and P-wave velocity. This
data was analyzed to obtain the dynamic shear modulus,
dynamic Youngs modulus, and dynamic Poissons ratio.
TEST RESULTS AND DISCUSSIONS
Destructive tests
Polymer concreteCompressive properties: Compressive
stress-strain relationships of PC with varying polymer
contents are compared in Fig. 2. The variations of strength,
modulus, and failure strain with polymer content are shown
in Fig. 3 and are summarized in Table 2. Though failure
strain increased with increasing polymer content, strength
remained constant while the modulus decreased after
reaching a peak value at 14% polymer content and, hence,
14% PC was considered the optimum system. Because
increasing polymer content in the PC system reduced the
aggregate content, it resulted in lower modulus. Behavior
models have been proposed by Cohen and Isahi22 and
Hirsch 23 for optimum combination, and modified by
Vipulanandan and Dharmarajan4 for excess polymer
system. Optimum PC systems can be idealized as either
aggregate and porous polymer distributed in a series (Fig. 4(a))
or as a combination of both series and parallel distribution
(Fig. 4(b)). In excess polymer system aggregate, optimum
porous polymer and excess polymer can either be distributed
in a series (Fig. 5(a)) or as a combination of both series and
parallel distribution (Fig. 5(b)). In the case of optimum
system, if the constituents are distributed in a series, the
modular ratio can be represented as22
PC

E
1
--------- = --------------------------------------------P

1
X
A
E
---------------------- + mX A
23

1 CV

(6)

Fig. 2Compressive stress-strain relationships of polyester


PC systems.

Fig. 4Behavior models for optimum PC system: (a) series;


and (b) series-parallel.

Fig. 3Variation of properties with polymer content for


polyester PC systems.
32

Fig. 5Behavior models for excess polymer system: (a)


series; and (b) series-parallel.
ACI Materials Journal/January-February 2004

where CV = XV /(XV + XP) is the porosity of polymer matrix


and XV and XP are the void content and polymer content by
volume; the ratio of polymer modulus (EP = 6.8 GPa) to
aggregate modulus (EA = 700 GPa) is denoted by m; and XA
is the solid content (ratio of aggregate volume to total volume
of polymer concrete). If the constituents are distributed in a
combination of both series and parallel then the modular
ratio can be represented as (Model 1)22

PC

E
1
--------- = -----------------------------------------------------P
( 1 XA )
E
+ mX A
Z --------------------------23
(1 CV )

(7)

-------------------------------------------------------------------------------------------X
23
+ ( 1 Z ) ( 1 X A ) ( 1 C V ) + -----Am

Table 2Static test results of PC, GFRPC, and CFRPC


Model value
Specimen no. Peak strength, MPa
14% polymer

Compression
test

18% polymer

PC
20% polymer

Tension test

14% polymer

0% fiber

Compression
test

4% fiber

GFRPC
6% fiber

Tension test

6% fiber

0% fiber

Compression
test

4% fiber

6% fiber

CFRPC

0% fiber

2.5% fiber
Tension test
4% fiber

6% fiber

Peak strain, %

Modulus, GPa

55

0.45

19.44

56

0.45

19.5

55

0.45

19.5

55

0.56

19.9

50

0.55

19.6

52

0.54

19.6

61

0.78

18.57

59

0.77

18.21

57

0.78

18.29

0.055

15

7.5

0.05

14.5

0.055

15

55

0.56

19.9

50

0.55

19.6

52

0.54

19.6

65

0.8

18.57

63

0.78

18.2

68

0.81

18.5

80

0.85

21

82

0.88

21.6

58

0.38

21.8

13

0.105

18.5

12.5

0.1

18.4

13.5

0.11

18.6

61

0.78

18.57

59

0.77

18.21

57

0.78

18.29

60

0.9

17.8

55

0.79

17.1

61

0.88

17.9

65

0.95

17.5

68

0.92

17.6

65

0.98

17

7.33

0.063

13.5

7.5

0.071

13.1

7.1

0.059

13.3

9.5

0.08

14.1

9.27

0.075

14.3

9.2

0.072

13.9

10.55

0.095

14.4

10.6

0.095

14.4

10.7

0.095

14.3

11.5

0.19

14

11.3

0.195

13.8

11.5

0.19

14

ACI Materials Journal/January-February 2004

55

0.45

0.04

0.63

55

0.56

0.07

0.49

61

0.85

0.08

0.39

0.055 0.015

0.85

55

0.56

0.07

0.49

65

0.8

0.06

0.44

80

0.85

0.3

0.45

13

0.105 0.095

0.55

61

0.85

0.05

0.39

60

0.9

0.05

0.37

65

0.95

0.05

0.39

7.3

0.063

0.05

0.86

9.3

0.075 0.065

0.86

10.6

0.095 0.065

0.85

11.5

0.19

0.82

0.07

33

where Z and (1 Z) are the relative proportions of a series


and parallel elements. The series-parallel model predicted
the PC modulus reasonably well for optimum system (14%
polymer content) with Z = 0.625. For excess polymer system,
the effect of excess polymer is taken into account, if distributed
in a series, the modular ratio can be represented as4
PC

E
1
--------- = -----------------------------------------------------------P

1
X
A
E
--------------------- + mX A + X EP
23

1 Cp

(8)

where XA and CP are the solid content and porosity considering


optimum polymer content (14%), and XEP is the excess
polymer content over and above the optimum content; and,
if distributed in a series and parallel combination, the
relationship is as follows, (Model 2)4
PC

1
E
-------- = ------------------------------------------------P
Z [ ZA ] + ( 1 Z ) [ Z B ]
E

(9)

1 XA
- + mX A + X EP and
where Z A = --------------------23

1 Cp
23

ZB = ( 1 XA ) ( 1 C p

X
) + -----A- + X EP
m

Fig. 6Comparison of predicted and experimental compressive modulus of PC.

Fig. 7Comparison of tensile and compressive behavior of


14% PC system.
34

This excess polymer model predicted the compressive


moduli of 18% PC and 20% PC reasonably well with the
same relative proportion of a series and parallel elements (Z
= 0.625) (Fig. 6). By combining the optimum polymer model
(Eq. 7) and excess polymer model (Eq. 9), the theoretical
peak value of the compressive modulus of the PC system
was obtained (Fig. 6). The theoretical peak was in good
agreement with the experimental data and hence, 14% PC
was considered as the optimum system.
Tensile propertiesThe tensile and compressive stressstrain relationships for 14% PC are compared in Fig. 7. The
tensile-to-compressive strength ratio for 14% PC system was
0.13, the tensile-to-compressive modular ratio was 0.75, and
the tensile-to-compressive toughness (area under stress-strain
curve) ratio at failure was 0.01. Based on the modular ratio, PC
systems can be characterized as bimodulus material.
Polymer concrete with glass fibersCompressive properties:
Typical compressive stress-strain relationships of GFRPC with
varying glass fiber contents are shown in Fig. 8. The variations
in strength, failure strain, and modulus with glass fiber content
are shown in Fig. 9(a), (c), and (e). Maximum fiber content in the
GFRPC system was limited to 6% by weight based on
workability. The addition of 6% glass fibers in PC systems
increased the compressive strength; strain at peak stress and
compressive modulus by 45, 50, and 10% respectively.
Tensile properties: Tensile properties were also improved
with the addition of glass fiber, where an 85% increase in the
tensile strength of the PC system was observed with the addition
of 6% glass fiber. The tensile behavior of the optimum
combination of the GFRPC system (6% glass fiber in 18%
polymer) is shown in Fig. 10.
Bimodulus behavior was observed for the GFRPC system
also. The tensile-to-compressive strength ratio for a 6%
GFRPC system was 0.16, the tensile-to-compressive
modular ratio was 0.85, and the tensile-to-compressive
toughness ratio at failure was 0.018.
Polymer concrete with carbon fibersCompressive
properties: Figure 11 shows the typical compressive stressstrain relationship of the CFRPC with varying carbon fiber
contents. The variation in strength, failure strain, and
modulus with carbon fiber content is shown in Fig. 9(b),
(d), and (f) and is summarized in Table 2. The test results
showed that the addition of carbon fibers did not improve
the compressive properties of the PC system. High
modulus PAN-based carbon fibers in the polymer matrix
(fiber-to-matrix modular ratio of 33.5) were not effective
in improving the compressive properties of the PC

Fig. 8Compressive
GFRPC systems.

stress-strain

relationships

of

ACI Materials Journal/January-February 2004

system. This might be because of high aspect ratios of the


carbon fibers.
Tensile properties: The typical tensile stress-strain
relationship of the CFRPC system is shown in Fig. 10.
The variations of the tensile properties with fiber content for
the CFRPC systems are shown in Fig. 12(a), (b), and (c) and are
summarized in Table 2. The addition of 6% carbon fibers to
the PC system (20% polymer) increased the tensile
strength, strain at peak stress, and tensile modulus by 60,
100, and 5% respectively. Based on the tensile properties
and workability, the fiber content in the CFRPC system
was limited to 6% by weight.
Bimodulus behavior was observed for the CFRPC system
also. The tensile-to-compressive strength ratio for the 6%
CFRPC system was 0.18, the tensile-to-compressive
modular ratio was 0.78, and the tensile-to-compressive
toughness ratio at failure was 0.029.
Strength model
Inspection of failure surfaces of the fiber-reinforced
polymer concrete in tension indicated adhesive and cohesive
failure of polymer with mainly fiber pullout; hence, the strength
model, as suggested by Vipulanandan and Dharmarajan,4 was
used to predict the tensile strength of the CFRPC
systems. The tensile strength of FRPC systems can be
represented as

FRPC

AT

= ( ) ( 4 ) + ( 1 4 ) ( 1 C V ) t

(10)

AS

+ Xf
and considering the change in strength (Model 3)
FRPC

( t

PC

t ) t

PC

= K1 Xf

AS

K1 = t

(11)

where tFRPC = tensile strength of fiber-reinforced polymer


concrete; tPC = tensile strength of polymer concrete without
fibers; tP = tensile strength of polymer; AT = adhesive tensile
strength between aggregate and polymer; AS = adhesive
shear strength between fiber and polymer; C V = porosity of
the polymer matrix; = fiber efficiency factor; and Xf =
fiber content by volume. Hence, incremental tensile strength
ratio of the CFRPC system was directly proportional to the
volume fraction of fibers. The comparison of model
predictions with experimental data for both GFRPC and
CFRPC systems are shown in Fig. 13. The factor K1 for
GFRPC systems was 1.42, whereas for CFRPC systems it
was 0.86. The deficient performance of carbon fibers as
compared with glass fibers may be attributed to the smoother
surface texture of carbon fibers, which may have reduced
both the adhesive shear strength between fiber and polymer
AS and fiber efficiency factor .

Fig. 10Tensile stress-strain properties of optimum


combination of PC, GFRPC, and CFRPC system.

Fig. 9Variation of compressive properties with fiber


content for GFRPC and CFRPC.
ACI Materials Journal/January-February 2004

Fig. 11Compressive stress-strain relationships of


CFRPC systems.
35

Stress-strain model
A model8,10,11 that has been used in the past to describe
PC behavior is proposed to predict the stress-strain behavior
of PC, GFRPC, and CFRPC systems in both tension and
compression. The proposed model is (Reference 9)

---c
- c
= -----------------------------------------------------------------------(p + q )

q + ( 1 p q ) ---- + p ----
c
c

(12)

----------------p

where p and q = material parameters; c and c = peak stress


and strain at peak stress; and and = stress and corresponding
strain. By satisfying the condition of initial slope of the stressstrain curve ( = 0), the following relationship can be obtained

E
q = -----s
Ei

(13)

where Es is the secant modulus at peak, and Ei is the initial


tangent modulus. The parameter p can be calculated using
regression analysis (least square method). A typical way of
calculating the material constant p is shown in Fig. 14. From
the known values of c , c , q, and the assumed value of p,
stress was predicted at a particular strain level and the square
of error of predicted stress with the experimental stress was
calculated. This procedure was repeated assuming different
values of p, and the value of p corresponding to the least
square of error was assumed to be the material constant p for
that material. The model parameter q signifies the ratio of
secant modulus at peak to initial tangent modulus, and q = 1
signifies linear material up to peak stress; hence, the lower
value of q represents more nonlinear material. On the other
hand, the parameter p controls mainly the postpeak behavior,
though it influences the prepeak behavior also. The parameter
p can be related to the toughness of the material, and in Fig. 15,
the variation of prepeak normalized toughness

Fig. 13Experimental and predicted tensile strength of


CFRPC and GFRPC.

Fig. 12Variation of tensile properties with fiber content


for CFRPC systems.
36

Fig. 14Predicted and measured compressive stress-strain


relationship of 4% CFRPC system.
ACI Materials Journal/January-February 2004

Dynamic G

---
----d
c c

---
----d
c c

with the material parameter p is compared. Both prepeak and


postpeak toughness increased with an increase in the material
parameters p and q. But q has the greatest influence on the
prepeak toughness, whereas p controls the postpeak toughness,
and also the influence of p on the prepeak toughness of the
material is negligible in the range of p obtained for the PC,
GFRPC, and CFRPC systems studied; hence, for PC
systems, material constant p was used to compare the
postpeak toughness. The typical prediction of stress-strain
relationship is compared with the experimental data in Fig. 15.
The modulus of elasticity, strength, failure strain, and model
values c, c, p, and q for different PC, GFRPC, and CFRPC
systems are summarized in Table 2.
For PC systems, the increase in polymer content above
14% w/w (weight/weight) did not increase the peak strength,
but material parameter q decreased and p increased. Reduction
in q indicates the fact that material is becoming more
nonlinear with an increase in polymer content. An increase
in parameter p with polymer content indicates that postpeak
toughness increased with polymer content. Similarly, for the
GFRPC systems, the initial stiffness remained more or less
the same with an increase in fiber content and, hence, q
remain unchanged but postpeak toughness increased
substantially (increase in parameter p) with the addition of
6% glass fibers. For CFRPC systems, there was not much
change in initial stiffness as well as toughness with the addition
of carbon fibers because both the parameters p and q
remained more or less constant. Similarly, comparing the tensile
model values p and q of optimum combinations of PC, GFRPC,
and CFRPC systems, it can be concluded that GFRPC was the
stiffest (comparing q) and also the toughest (comparing p).
Nondestructive testsImpact resonance method: Impact
resonance tests were performed on both cylindrical and
prismatic specimens to study the effect of shape of the
specimens on the measured material properties. A typical
frequency spectrum X-f obtained from longitudinal mode of
vibration of 6% GFRPC cylindrical specimen is shown in
Fig. 17. From the measured resonant frequencies, the dynamic
Youngs modulus, dynamic torsional modulus, and dynamic
Poissons ratio were calculated as per ASTM C 215 as follows:
From transverse mode of vibration
Dynamic E

= CM ( n )

E
Dynamic = ------- 1
2G

and postpeak normalized toughness


1.5

= BM ( n )

(16)

(17)

where n, n, n are the respective resonance frequencies in


transverse, longitudinal, and torsional mode, and C, D, and
B are shape correction factors (Reference 11). The results are
summarized in Table 3 for prismatic specimens and in Table 4
for cylindrical specimens. The dynamic Youngs moduli
calculated from the longitudinal mode of vibration were 16
to 20% higher than the moduli from the transverse mode of
vibration for all the PC systems studied. This reinforces the
earlier observation that PC, GFRPC, and CFRPC systems are
bimodulus materials having different moduli in compression
and tension, where direct compression moduli were 15 to
25% higher than the direct tension moduli for the PC,
GFRPC, and CFRPC systems studied. Though the shape of
specimens did not have significant effect on the Youngs
modulus, it had substantial effect on the shear modulus, with
the prism specimens showing about 20% higher values than
cylindrical specimens; also, very large variations in the
dynamic Poissons ratio were observed (Table 3 and 4), with
the cylindrical specimens yielding higher values.
Because resonant frequencies are not material property,
resonant wave velocities (material property) for PC,
GFRPC, and CFRPC systems were calculated and later
compared with P-wave and shear wave velocities obtained
from the pulse velocity test to study the shape effect of the
specimens. The wave velocity V can be determined by
V = f n n

(18)

Fig. 15Variation of toughness with material constant p.

(14)

From longitudinal mode of vibration


Dynamic E

= DM ( n )

From torsional mode of vibration


ACI Materials Journal/January-February 2004

(15)
Fig. 16Typical frequency spectrum (X-f) of 6% GFRPC
cylinder for longitudinal mode of vibration.
37

where wavelength n = 2 L for longitudinal and torsional


fundamental modes. Obtaining the resonant frequencies in
longitudinal and torsional modes of vibration from the
impact resonance tests, and using Eq. (18) resonant P-wave
and shear wave velocities of PC, GFRPC, and CFRPC
systems were calculated and summarized in Table 3 for prism
specimens and in Table 4 for cylindrical specimens. The typical
time-decay X-t plot obtained from the longitudinal mode
of vibration of a 6% GFRPC cylindrical specimen is shown
in Fig. 17. From the time-decay plot the damping ratio was
calculated for each system for each mode of vibration and is

summarized in Table 5 for both prismatic and cylindrical


specimens. The damping ratio was calculated by the logarithmic
decrement method as follows

= 2 where = ln ( y n y n + 1 )

(19)

The average damping ratios for each mode of vibrations were


calculated from the X-t plots for both GFRPC and CFRPC
systems by considering at least 10 cycles. Overall damping ratio
increased with increasing polymer content and in general, glass
fiber PC had a higher damping ratio than the carbon fiber PC

Table 3Nondestructive test results of prismatic specimens


Impact resonance

PC

Pulse velocity

No.

EiL,
GPa

Longitudinal
wave velocity,
m/s

EiTr,
GPa

Gi ,
GPa

Shear wave
velocity, m/s

vi

P-wave
velocity,
Vp, m/s

14%
polymer

21.92

3210

17.27

9.20

1932

0.19

3226

2027

0.63

0.17

8.92

20.93

21.90

3199

17.86

9.99

2027

0.10

3215

2020

0.63

0.17

8.91

20.92

18%
polymer

20.94

3114

18.18

10.22

1980

0.02

3297

2041

0.62

0.19

9.18

21.83

22.47

3234

17.57

10.90

2010

0.03

3279

2069

0.63

0.17

9.39

21.95

20%
polymer

21.13

3250

17.79

8.84

1993

0.20

3268

2041

0.62

0.18

8.5

20.07

21.79

3234

17.14

10.22

2016

0.07

3261

2055

0.63

0.17

8.98

21.02

20.94

3114

18.18

10.22

1980

0.02

3297

2041

0.62

0.19

9.18

21.83

22.47

3234

17.57

10.90

2010

0.03

3279

2069

0.63

0.17

9.39

21.95

20.32

3227

17.10

9.47

1987

0.07

3233

2020

0.62

0.18

8.13

19.17

21.77

3246

17.58

9.93

2016

0.10

3293

2013

0.61

0.2

8.55

20.54

22.37

3332

18.97

10.42

2091

0.07

3389

2095

0.62

0.19

9.03

21.49

22.48

3334

18.72

9.79

2066

0.15

3393

2097

0.62

0.19

9.08

21.62

21.13

3250

17.79

8.84

1993

0.20

3268

2041

0.62

0.18

8.5

20.07

21.79

3234

17.14

10.22

2016

0.07

3261

2055

0.63

0.17

8.98

21.02

21.56

3210

17.39

10.24

2034

0.05

3257

2041

0.63

0.18

8.99

20.93

21.06

3256

17.82

10.25

2023

0.03

3272

2041

0.62

0.18

8.44

19.95

20.29

3210

17.95

9.61

1992

0.06

3243

2027

0.63

0.18

8.25

19.47

21.08

3257

17.89

9.89

2011

0.07

3279

2076

0.63

0.17

8.74

20.37

0% fiber
GFRPC 4% fiber
6% fiber
0% fiber
CFRPC 4% fiber
6% fiber

Shear wave
velocity, Vs

Vs /Vp

vp

Gp, GPa

Ep, GPa

Table 4Nondestructive test results of cylindrical specimens


Impact resonance

PC

vi

Shear wave
velocity, Vs

Vs /Vp

vp

Gp,
GPa

Ep, GPa

No.

EiL,
GPa

14%
polymer

21.09

3198

18.7

8.16

1989

0.29

3250

2031

0.63

0.18

8.69

20.49

21.37

3224

18.7

8.19

1995

0.31

3262

2052

0.63

0.17

8.84

20.73

18%
polymer

21.75

3234

17.9

8.06

1968

0.35

3339

2074

0.62

0.19

9.13

21.65

22.5

3235

18.9

8.49

1986

0.33

3397

2096

0.62

0.19

9.64

22.99

20%
polymer

20.82

3186

17.1

7.87

1959

0.32

3275

2026

0.62

0.19

8.59

20.44

20.98

3206

17.3

8.01

1980

0.31

3257

2020

0.62

0.19

8.5

20.19

21.75

3234

17.9

8.06

1968

0.35

3339

2074

0.62

0.19

9.13

21.65

22.5

3235

18.9

8.49

1986

0.33

3397

2096

0.62

0.19

9.64

22.99

21.54

3182

18.8

8.55

2003

0.26

3210

2031

0.63

0.17

8.95

20.89

21.64

3176

18.9

8.5

1990

0.27

3236

2031

0.63

0.18

9.03

21.22

24.05

3328

20

8.57

1985

0.4

3363

2111

0.63

0.17

9.88

23.22

23.51

3302

20.4

9.42

2090

0.25

3361

2105

0.63

0.18

9.76

22.97

20.82

3186

17.1

7.87

1959

0.32

3275

2026

0.62

0.19

8.59

20.44

20.98

3206

17.3

8.01

1980

0.31

3257

2020

0.62

0.19

8.5

20.19

21.84

3217

18.4

8.46

2001

0.29

3273

2010

0.61

0.2

8.7

20.84

20.29

3193

17.6

7.88

1990

0.29

3257

2031

0.62

0.18

8.37

19.79

21.05

3197

18.2

7.98

1968

0.32

3230

2031

0.63

0.17

8.67

20.34

21.26

3196

17.9

8.07

1969

0.32

3248

2036

0.63

0.18

8.81

20.72

0% fiber
GFRPC 4% fiber
6% fiber
0% fiber
CFRPC 4% fiber
6% fiber

38

Pulse velocity
P-wave
velocity,
Vp, m/s

Longitudinal
wave velocity,
m/s

EiTr,
GPa

Gi ,
GPa

Shear wave
velocity, m/s

ACI Materials Journal/January-February 2004

systems. For the CFRPC systems, though the longitudinal


damping ratio increased with addition of carbon fibers, it had a
negative effect on the transverse and torsional damping ratio.
The damping ratio in the longitudinal mode was 2% for 6%
CFRPC system, compared with 1.3% for 6% GFRPC system.
The transverse mode damping ratio was 1.8% for 6% GFRPC
system and in torsional mode 4.96% for 6% GFRPC system.
Pulse velocity method: Pulse velocity tests were also
performed on both cylindrical and prismatic specimens. Wavelengths of the test P-wave pulses were in the range of 20 to
23 mm, ten times the largest aggregate size used in test
specimens, having a minimum dimension of 50 mm.
Wavelengths of the test shear wave pulses were in the
range of 13 to 14 mm. Measuring the wave velocities,
wavelengths were calculated from the known test pulse
frequency using Eq. (18). From the P-wave and shear
wave velocity, dynamic Poissons ratio was first calculated
from Eq. (4) and then dynamic Youngs modulus and dynamic
shear modulus were calculated from Eq. (5) and (1),
respectively. This was another way to calculate the
Youngs modulus from the pulse velocity method other than
ASTM C 597, where calculation of Youngs modulus
requires assumption of Poissons ratio. Shear modulus
can also be calculated from this method. From the pulse
velocity test, P-wave velocities Vp and shear wave velocities Vs
for PC, GFRPC, and CFRPC systems were measured and
shear wave velocity to P-wave velocity ratios Vs /Vp,
dynamic Youngs moduli, dynamic shear moduli, and dynamic
Poissons ratios were calculated and are summarized in
Table 3 for prism specimens and in Table 4 for cylindrical
specimens. The test results showed that the dynamic Youngs
modulus, dynamic shear modulus, as well as the dynamic
Poissons ratio was independent of specimen shape. The
velocity ratio Vs/Vp for 14% PC system was 0.63, and this ratio
was almost independent to polymer and fiber addition in the
PC system as well as to shape of the specimen.
The longitudinal and shear wave velocities obtained from
impact resonance tests for both cylindrical and prism
specimens for all the PC, GFRPC, and CFRPC systems

studied were on an average 2 to 3% less than those calculated


from pulse velocity tests; this was partly because the test
wavelengths in impact resonance tests were higher than
those of pulse velocity test. While the test wavelengths in the
pulse velocity tests were between 13 and 23 mm, the test
wavelengths in impact resonance tests were in the range
of 400 to 600 mm. Because longitudinal wave velocity is
directly proportional to dynamic Youngs modulus (Eq. (5)), and
shear wave velocity is directly proportional to dynamic shear
modulus (Eq. (1)), dynamic Youngs moduli, and dynamic shear
moduli, those calculated from the impact resonance test should
be less than those calculated from pulse velocity tests. But the
dynamic Youngs moduli and dynamic shear moduli, calculated
from impact resonance tests using ASTM C 215 correction
factors for the PC, GFRPC, and CFRPC systems, studied were
on an average 10% higher than respective moduli calculated
from pulse velocity tests, except for shear moduli for cylindrical
specimens, where impact resonance method predicted a
lower value. This is contradicting but partly explainable
because the correction factors used in ASTM C 215 are
approximate.7 The equation relating dynamic Youngs
modulus and longitudinal wave velocity for impact resonance
test (combining Eq. (15) and (18)) and the equation relating
dynamic Youngs modulus and longitudinal wave velocity
for pulse velocity tests (Eq. (5)) are compared in Fig. 18 and
19 for prism and cylindrical specimens, respectively, along
with experimental results. Similarly, the equation relating

Fig. 17Typical time-decay (X-t) relationship of 6% GFRPC


cylinder for longitudinal mode of vibration.

Table 5Damping ratios of PC, GFRPC, and CFRPC


Longitudinal damping ratio, %

PC

Cylinder

Prism

Cylinder

Prism

14%
polymer

0.80

0.92

1.15

1.32

2.74

2.84

0.88

0.92

1.03

1.32

2.82

2.79

18%
polymer

0.92

0.95

1.3

1.41

2.9

2.95

0.86

0.95

1.2

1.38

2.83

2.89

20%
polymer

1.18

1.2

1.78

1.65

3.32

3.51

1.1

1.1

1.82

1.79

3.41

3.39

0.92

0.95

1.3

1.41

2.9

2.95

0.86

0.95

1.2

1.38

2.83

2.89

4% fiber
6% fiber
0% fiber

CFRPC

Torsional damping ratio, %

Prism

0% fiber
GFRPC

Flexural damping ratio, %

Specimen

4% fiber
6% fiber

Cylinder

1.01

1.13

1.6

1.65

3.81

3.99

1.05

1.19

1.52

1.49

3.72

4.01

1.17

1.26

1.75

1.8

4.96

4.9

1.23

1.3

1.7

1.76

4.91

4.85

1.18

1.2

1.78

1.65

3.32

3.51

1.1

1.1

1.82

1.79

3.41

3.39

1.71

1.82

0.3

0.23

2.2

2.41

1.69

1.77

0.27

0.38

2.29

2.53

1.98

2.01

0.35

0.3

2.42

2.53

1.89

1.92

0.32

0.33

2.3

2.44

ACI Materials Journal/January-February 2004

39

dynamic shear modulus and shear wave velocity for impact


resonance tests (combining Eq. (16) and (18)) and the equation
relating dynamic shear modulus and shear wave velocity for
pulse velocity tests (Eq. (1)) are compared in Fig. 20 and 21
for prism and cylindrical specimens respectively, along
with experimental results. These plots showed that for the
same wave velocity, impact resonance tests predict higher
moduli (except for shear moduli for cylindrical specimen)
and pulse velocity tests predict lower moduli. Though the
discrepancies in Youngs moduli and shear moduli were within
an acceptable limit, the Poissons ratios were much more
sensitive. The dynamic Poissons ratios calculated from the
impact resonance method for prism specimens were far less
than those calculated from the pulse velocity method and for
cylindrical specimens they were about 30% higher. The
dynamic Poissons ratios calculated from pulse velocity tests
matched reasonably well with static Poissons ratio reported in
the literature3,10 for PC systems.
While comparing the Youngs moduli obtained from static
tests with those obtained from dynamic tests, the dynamic
moduli for all the PC systems were higher than the respective
static moduli; this was due to the fact that moduli were
determined at low strain (order of 106) in the dynamic tests.
The moduli obtained from the pulse velocity method
predicted the static moduli more closely than the moduli
from the impact resonance method.

CONCLUSIONS
For PC systems, the optimum combination (lowest
polymer content at which strength and/or modulus are
maximum) was obtained at 14% polymer content. For the
GFRPC system, the mechanical properties improved with
the addition of glass fibers, but, based on workability, the
optimum combination was 6% glass fibers with 18%
polymer. For the CFRPC system, though the compressive
properties were almost independent to fiber addition, tensile
properties improved significantly. Based on the tensile
properties and workability, for the CFRPC systems, the
optimum system was with 6% carbon fibers and 20%
polymer. Based on the destructive and nondestructive tests
performed on over 45 PC, GFRPC, and CFRPC specimens, the
following conclusions are advanced:
1. Polymer concrete is a bimodulus material with a tensileto-compressive modular ratio of 0.75. A combination of
series and parallel iso-stress models predicted the modulus
of PC systems reasonably well;
2. The tensile strength of the PC system improved by 85
and 60% with the addition of 6% glass fibers and 6% carbon
fibers (by weight), respectively. The tensile strength model
predicted the strength increase in PC systems with fibers.
Glass fiber addition improved the compressive strength of
polymer concrete, but carbon fibers did not. Polymer
concrete with glass and carbon fibers also behaved as a
bimodulus material with a tensile-to-compressive modular
ratio of 0.85 and 0.78, respectively;

Fig. 18Comparison of Youngs moduli obtained from pulse


velocity and impact resonance tests for prism specimens.

Fig. 20Comparison of shear moduli obtained from pulse


velocity and impact resonance tests for prism specimens.

Fig. 19Comparison of Youngs moduli obtained from pulse


velocity and impact resonance tests for cylindrical specimens.

Fig. 21Comparison of shear moduli obtained from pulse


velocity and impact resonance tests for cylindrical specimens

40

ACI Materials Journal/January-February 2004

3. The p-q model for stress-strain relationships predicted


both the prepeak and postpeak behavior with a single function
for both tension and compression. While the parameter q
influenced the stress-strain relationships before peak, the
parameter p influenced the toughness, mainly the postpeak toughness;
4. The impact resonance and the pulse velocity methods
can be used to characterize the behavior of PC with and
without fibers. The ratio of shear wave velocity to P-wave
velocity for 14% PC was 0.63. The addition of polymer or
fibers did not influence the velocity ratio. While the pulse
velocity method was independent of specimen shape, the
impact resonance method was dependent on specimen
shape, with dynamic shear modulus and dynamic Poissons
ratio having the greatest effect. Dynamic Youngs moduli
and dynamic shear moduli obtained from the pulse velocity
method were within 10% of the respective static moduli,
whereas those from the impact resonance test showed larger
variations. Dynamic Poissons ratio obtained from the pulse
velocity method was closer to static Poissons ratio. Wave
velocities measured from the pulse velocity test were higher
than those determined from the impact resonance method; and
5. The damping ratio increased with an increase in
polymer and fiber contents in the PC systems. Glass fibers
increased the damping ratio in the longitudinal, flexural, and
torsional modes. Carbon fibers increased the longitudinal
damping ratio but it also reduced the flexural and torsional
damping ratios.
ACKNOWLEDGMENTS
This work was supported by the Center for Innovative Grouting Materials
and Technology (CIGMAT) at the University of Houston under grants from
the National Science Foundation (CMS-9634685) and Advanced Research
Program (ARP)-Texas Higher Education.

REFERENCES
1. ACI Committee 548, Guide for the Use of Polymers in Concrete
(ACI 548.1R-97), American Concrete Institute, Farmington Hills, Mich.,
1997, 29 pp.
2. Ohama, Y., and Nishimura, T., Properties of Steel Fiber Reinforced
Polyester Resin Concrete, Proceedings of the 22nd Congress on Material
Research, Society of Material Sciences, Kyoto, Japan, 1979, pp. 364-367.
3. Mebarkia, S., and Vipulanandan, C., Compressive Behavior of Glass
Fiber Reinforced Polymer Concrete, Journal of Materials in Civil
Engineering, V. 4, No. 1, Feb. 1992, pp. 91-105.
4. Vipulanandan, C., and Dharmarajan N., Flexural Behavior of Polyester
Polymer Concrete, Cement and Concrete Research, V. 17, 1987, pp. 219-230.

ACI Materials Journal/January-February 2004

5. ASTM C 215-91, Standard Test Method for Fundamental Transverse,


Longitudinal and Torsional Frequencies of Concrete Specimens,
ASTM International, West Conshohocken, Pa., 1997, 6 pp.
6. ASTM C 597, Standard Test Method for Pulse Velocity Through
Concrete, ASTM International, West Conshohocken, Pa., 10 pp.
7. Mantrala, S. K., and Vipulanandan, C., Nondestructive Evaluation of
Polyester Polymer Concrete, ACI Materials Journal, V. 92, No. 6, Nov.Dec. 1995, pp. 660-668.
8. Vipulanandan, C., and Paul, E., Characterization of Polyester Polymer
and Polymer Concrete, Journal of Materials in Civil Engineering, V. 5,
No. 1, Feb. 1993, pp. 62-82.
9. Mebarkia, S., Mechanical and Fracture Properties of High Strength
Polymer Concrete Under Various Loading Conditions and Corrosive
Environments, PhD thesis, University of Houston, Tex., 1993.
10. Mantrala, S. K., Role of Core Material on Structural Behavior of
Sandwich Pipes, PhD thesis, University of Houston, Tex., 1996.
11. Vipulanandan, C., and Paul, E., Performance of Epoxy and Polyester
Polymer Concrete, ACI Materials Journal, V. 87, No. 3, May-June 1990,
pp. 241-251.
12. CIGMAT PC 1-00, Standard Practice for Making and Curing Polymer
Concrete Test Specimens in Laboratory, University of Houston, Houston,
Tex., 8 pp.
13. CIGMAT PC 5-00, Standard Test Method for Compressive Properties
of Polymer Concrete, University of Houston, 12 pp.
14. CIGMAT PC 2-00, Standard Test Method for Tensile Properties of
Polymer Concrete, University of Houston, 4 pp.
15. ASTM C 469-94, Standard Test Method for Static Modulus of
Elasticity and Poissons Ratio of Concrete in Compression, ASTM
International, West Conshohocken, Pa., 1994, 4 pp.
16. ASTM C 39M-01, Test Method for Compressive Strength of
Cylindrical Concrete Specimens, ASTM International, West Conshohocken,
Pa., 2001, 5 pp.
17. ASTM C 617-98, Standard Practice for Capping Cylindrical
Concrete Specimens, ASTM International, West Conshohocken, Pa.,
1998, 6 pp.
18. Obert, L., and Duvall, W. I., Discussion of Dynamic Methods of
Testing Concrete with Suggestions for Standardization, Proceedings of
ASTM, V. 41, 1941, pp. 1053-1070.
19. Picket, G., Equations for Computing Elastic Constants from
Flexural and Torsional Resonant Frequencies of Vibration of Prisms and
Cylinders, Proceedings, ASTM International, V. 45, 1945, pp. 846-865.
20. Bay, J. A., and Stokoe, K. H., II, Field and Laboratory Determination of
Elastic Properties of Portland Cement Concrete Using Seismic Techniques,
Transportation Research Record 1355, 1992, pp. 67-74.
21. Leslie, J. R., and Cheesman, W. J., An Ultrasonic Method of Studying
Deterioration and Cracking in Concrete Structures, ACI JOURNAL,
Proceedings V. 24, No. 1, Sept. 1949, pp. 17-35.
22. Cohen, L. J., and Isahi, O., The Elastic Properties of Three-Phase
Composites, Journal of Composite Materials, V. 1, 1967, pp. 390-403.
23. Hirsch, T. J., Modulus of Elasticity of Concrete Affected by Elastic
Moduli of Cement Paste Matrix and Aggregate, ACI JOURNAL, Proceedings
V. 59, Mar. 1962, pp. 427-452.
24. Jones, R. M., Apparent Flexural Modulus and Strength of Multimodulus
Material, Journal of Composite Materials, V. 10, Oct. 1976, pp. 342-354.

41

S-ar putea să vă placă și