You are on page 1of 11

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/259641909

Structural Design Methodologies for Concrete


Pipes with Steel and Synthetic Fiber
Reinforcement
Article in Aci Structural Journal January 2014

CITATIONS

READS

866

3 authors, including:
Amirpasha Peyvandi
Stantec
30 PUBLICATIONS 170 CITATIONS
SEE PROFILE

All content following this page was uploaded by Amirpasha Peyvandi on 10 January 2014.
The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.

ACI STRUCTURAL JOURNAL

TECHNICAL PAPER

Title No. 111-S08

Structural Design Methodologies for Concrete Pipes with


Steel and Synthetic Fiber Reinforcement
by Amirpasha Peyvandi, Parviz Soroushian, and Shervin Jahangirnejad
Improved structural designs were developed and experimentally
verified for concrete pipes. These designs make effective use of
discrete synthetic fiber reinforcement to lower the reinforcing
steel ratio, and thus allow for increasing the protective cover
of concrete on steel for improved durability. The new concrete
pipe design also enhances the toughness and damage resistance
of pipes. The work reported herein covers theoretical modeling,
design, and experimental verification of concrete pipes with
synthetic fiber and conventional steel reinforcement. The focus of
the theoretical models was on the flexural strength and load-carrying capacity of concrete pipes. These models account for the
contributions of fibers to the tensile behavior of concrete via
fiber pullout or rupture. These models were used to develop new
concrete pipe designs, which made complementary use of synthetic
fiber (polyvinyl alcohol [PVA]) and conventional steel reinforcement. Full-scale pipes embodying the new design were fabricated
and experimentally evaluated. The experimental results were used
to refine the theoretical models and design procedures. Test results
confirmed that the number of steel reinforcement layers in concrete
pipes can be reduced with the use of synthetic fibers. This allows
for increasing the protective cover of concrete on steel, which is a
major advantage towards increasing the service life of concrete in
aggressive environments, including sanitary sewers, where microbial-induced corrosion is a major concern.

Fig. 1Concrete pipe external loads, flexural cracks, and


flexural reinforcement system: (a) vertical and horizontal
loads; (b) flexural cracks; and (c) reinforcement system.

Keywords: concrete cover; concrete pipes; steel reinforcement; structural


design; synthetic fibers.

INTRODUCTION
Sewer systems account for approximately half of the
infrastructure investment in the United States.1 Concrete
is the primary material used for construction of the sewer
infrastructure due to its satisfactory performance over
long-term periods. It offers a desired balanced of structural performance, barrier qualities, durability, and cost.
Concrete, however, provides relatively small tensile
strength and toughness. Steel reinforcement of concrete is
thus necessary for meeting structural performance requirements. Susceptibility of steel to corrosion and other degradation phenomena limit the service life of the concrete-based
sewerinfrastructure.
Advances in fiber-reinforced concrete provide a new
basis for design of more efficient concrete pipes with
reduced amount of steel reinforcement.2-9 Existing reinforced concrete pipe designs have not yet taken advantage of discrete fibers as complementary reinforcement for
enhancing structural efficiency and resolving some critical
performance issues.9-13 Pipes are subjected to earth pressures
(Fig. 1(a)) which generate transverse bending moments
in pipe walls (Fig. 1(b)). The transverse (circumferential) reinforcement is introduced in concrete pipes to resist
ACI Structural Journal/January-February 2014

Fig. 2Location of crown, invert, and springlines.


bending moments (Fig.1(c)). Under external loads, pipes
first develop vertical cracks (on interior surfaces [crown and
invert]), and then horizontal cracks (at mid-height on exterior surfaces [spring lines]). Figure 2 shows the locations of
crown, invert, and spring lines. Under the effect of transverse bending moments, pipes behave as rectangular reinforced concrete sections.
The use of two layers of steel reinforcement in traditional
pipe designs (Fig. 3(a)) reduces the concrete cover thickness
over the reinforcement, which compromises the durability of
pipes in the highly corrosive environment of a sanitary sewer.
The new design developed here (Fig. 3(b)) makes effective
ACI Structural Journal, V. 111, No. 1, January-February 2014.
MS No. S-2012-017.R1, doi:10.14359.51686432, was received October 7, 2012, and
reviewed under Institute publication policies. Copyright 2014, 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 ten months from this journals date if the discussion
is received within four months of the papers print publication.

83

Fig. 3Reinforcement configurations for two different design methods: (a) traditional design; and (b) new design.
towards experimental verification of the new pipe materials
and systems. Industrial-scale production and structural evaluation of concrete pipes embodying the new materials and
structural design principles were also implemented.
DESIGN EQUATIONS FOR ULTIMATE
FLEXURAL STRENGTH OF STEEL REINFORCED
CONCRETEPIPES
The moments generated in pipes under load in the threeedge bearing test (Fig. 4) are related to the weight of pipe Wp
and the total three-edge bearing load Wt as follows14-16

Fig. 4Schematics three-edge bearing test setup.


use of corrosion-resistant synthetic fiber reinforcement to
reduce the number of reinforcement layers from two to one.
This approach is enabled by the structural value offered by
synthetic reinforcement to compensate for the loss of steel
ratio in concrete pipes. This would increase the protective
concrete cover thickness over steel without compromising
the pipe structural performance.
The increased cover thickness would enhance the service
life of pipes in aggressive sewer environment, and would
reduce their maintenance requirements. Polyvinyl alcohol
(PVA) fibers suit the targeted application in concrete pipes,
due to their relatively high acid resistance, elastic modulus,
and bond strength to concrete.
RESEARCH SIGNIFICANCE
The objective of this research was to investigate the value
of synthetic fibers in concrete pipes towards reduction of
the number of steel reinforcement layers. The resulting rise
in concrete cover thickness over reinforcement benefits the
durability of the concrete pipes. To achieve the objective
of the research, new design procedures were developed for
making effective use of corrosion-resistant synthetic fibers
to reduce the number of reinforcement layers from two to
one. Comprehensive experimental studies were undertaken
84

M = (Cm1 Wp + Cm 2 Wt )

S
2

(1)

where S = Di + h; Di is pipe internal diameter; h is the pipe


wall thickness; and Cm1 and Cm2 are constant coefficients
which are introduced in the following sections.
For circular pipes, the moments generated in the threeedge bearing test vary as shown in Fig. 5. The critical conditions for design in flexure occur at the crown and the invert,
with the invert moment used for development of design
equations. While the invert moment caused by the test load
is slightly less than the crown moment, the moment caused
by the pipe weight is substantially greater at invert, making
the combined moment approximately the same at crown
andinvert.
At invert, coefficients Cm1 and Cm2 for calculating moment
in the elastic range are14-16
Cm1 = 0.75Cm2, Cm2 = 0.07
The moment equation can thus be simplified as follows

M=

0.14
(0.75 Wp + Wt )( Di + h )
40

(2)

Defining DL as (Wt /Di) 12, the moment equation can be


expressed as follows

9Wp
D
M = 0.14
+ DL i ( Di + h )
480
Di

(3)

ACI Structural Journal/January-February 2014

Fig. 5Moment distributions in three-edge bearing test loading: (a) weight of pipe; (b) uncracked; (c) first-stage cracking;
(d)second-stage cracking; and (e) ultimate flexural capacity.
Equation (3) can be expressed in the following simplified
form

9Wp
M = 2.925 10 4
+ DL Di ( Di + h )
D

(4)

Equation (4) can be used towards calculation of the ultimate load of concrete pipes in three-edge bearing tests as
described in the following equations.
The flexural strength of reinforced concrete sections can
be expressed as
a

M U = As f y d

(5)

Using the above equation for flexural strength in the


moment Eq. (4) yields

9Wp
a

As f y d = 2.925 10 4
+ DL Di ( Di + h ) (6)

Di

Hence,

3419 As f y d
9Wp

2
DL + D =
Di ( Di + h)
i

of Cmp over the load, which causes first yield in the inner
reinforcement. Cmp is equal to 1.10 for circular pipes with
two concentric cages and 1.25 for pipes with one cage.14-16
Cm is the coefficient corresponding to flexural first-yield and
is equal to 1.0 for circular cages. The load DL = (Wt /Di) 12
corresponding to ultimate flexural strength can be calculated
using the following equation

9Wp

DL + D =
i

3419Cm Cmp As f y d

2
Di ( Di + h)

DESIGN EQUATIONS FOR 0.025 cm (0.01 in.)


CRACK WIDTH FLEXURAL STRENGTH OF STEELREINFORCED CONCRETE PIPES
The 0.025 cm (0.01 in.) crack width is an arbitrarily chosen
test criterion commonly used in design and evaluation of
concrete pipes. It is not a serviceability criterion, but can
be viewed as a design criterion other than ultimate strength
which reflects on the cracking behavior of concrete pipes.
The nominal stress in transverse steel reinforcement under
moment M at a critical section can be calculated asfollows
fs =

(7)

Before a pipe reaches its collapse state in the three-edge


bearing test, its (inner) transverse steel reinforcement yields
crown and invert, and its outer reinforcement yields spring
lines. The final moment distribution at a pipe cross section
is shown in Fig. 5(e). The load capacity increases by a factor
ACI Structural Journal/January-February 2014

(8)

M
As j d

(9)

where j is the coefficient for moment arm at service load


and is considered to be 0.9. The moment corresponding to
0.025cm (0.01 in.) crack width (M.01) can be calculated
asfollows
fs = l fs.01 (10)

85

where l is a reduction factor for variability in test results


and is 0.9. Equations (9) and (10) yield the following value
for M.01
M.01 = 0.9fs.01 As 0.9d (11)
The expression for M.01 in terms of applied loads is

9Wp
M.01 = 2.925 10 4
+ DL Di ( Di + h )

Di

(12)

Equations (11) and (12) yield

0.9 fs.01

9Wp

4
2.925 10 D + DL Di ( Di + h )
i

(13)
=

0.9 As d

Solving the aforementioned equation for DL at 0.025 cm


(0.01 in.) crack width (DL.01) yields

9Wp 2769 As fs.01 d

DL.01 + D = D ( D + h)
i
i
i

(14)

DESIGN EQUATIONS FOR CONCRETE PIPES


WITHOUT STEEL REINFORCEMENT
Flexural design of plain concrete pipes is governed by the
tensile (flexural) strength limit state. Plain pipe must be designed
so that the highest flexural stress around the pipe circumference produced by a combination of moment and thrust is less
than the flexural tensile strength of concrete (usually taken as
the modulus of rupture of concrete fmr). The governing location for the highest combined flexural tension is commonly at
the invert. The modulus of rupture can be obtained using two
alternative methods. One method determines the modulus of
rupture from results of a three-edge bearing test

4.8 mr DLut Di ( Di + h )
fmr =

h2

(15)

where DLut is the test D-load that cracks the pipe.


The second method estimates the modulus of rupture
using the following equation

fmr = kmr mr

fc

(16)

The maximum value of fmr (0.85) produces service load


conditions which are close to those obtained from the threeedge bearing test results. The coefficient kmr in Eq. (16)
varies with wall thickness,15 and the minimum value of kmr
for any wall thickness is 201. Based on many three-edge
bearing tests, it has been concluded that kmr increases with
decreasing wall thickness, varying from approximately 228
86

Fig. 6Flexural stress distribution at ultimate condition in


singly reinforced fiber concrete section with predominant
fiber rupture.
for 10 cm (4 in.) walls, up to 322 or more for 5 cm (2 in.)
walls. The modulus of rupture of concrete pipe without steel
reinforcement was calculated using Eq. (15).
MODIFICATION OF DESIGN EQUATIONS TO
ACCOUNT FOR EFFECTS OF SYNTHETIC FIBERS
The ability of fiber-reinforced concrete to carry tensile
stress increases the flexural load-carrying capacity of
fiber-reinforced concrete at different stages of behavior.
Hence, the equations for the flexural strength of pipes have
to be modified to account for the structural contributions of
fibers. A semi-empirical approach was followed to derive
design equations for concrete pipes with fiber reinforcement.
They were developed based on theoretical concepts which
were modified to account for experimental results.
Ultimate flexural strength of steel-reinforced
concrete pipes with fiber reinforcement
Fibers in a cementitious matrix can exhibit two types of
behavior at cracks: pullout or rupture. The synthetic (PVA)
fibers used in this investigation show a stronger tendency
towards rupture, especially for fibers of finer diameter
and higher aspect ratio. The hydrophilic nature and strong
bonding of PVA fibers to the cementitious matrix explain
this behavior. The tendency towards fiber pullout increases
with increasing fiber diameter and decreasing fiber aspect
ratio. Other methods to promote fiber pullout involve the use
of an oiling agent and modifying the matrix to reduce chemical and frictional bonds to PVA fibers.
Design equations when fiber rupture dominates
The fiber concrete tensile stress distribution at flexural
failure is assumed to be triangular,17,18 with the tension zone
starting at the neutral axis and ending at the level of steel
reinforcement. It is assumed that, at the ultimate stage, the
crack width below steel reinforcement is wide enough to
have already caused rupture of fibers. These assumptions,
together with the equilibrium considerations in Fig. 6, yield
ACI Structural Journal/January-February 2014

the nominal flexural strength of a singly reinforced fiberreinforced concrete section as follows

a (d c)
a 2(d c)

M n = As f y d + t
bc +
(17)

2
2
2
3



where a is the depth of the compressive stress block (refer to
Fig. 6) and the fibrous concrete tensile strength (st) is calculated as follows

st = NU

where N is the number of fibers per unit area; and U is the


ultimate tensile force of a single fiber. An expression for st
is as follows

0.5V f
t =
2
d f

fu d 2f

(18)

t =

Hence,

st = 0.5Vfsfu (19)

where Vf, df, and sfu are fiber-volume fraction, diameter, and
ultimate tensile stress, respectively.
Design equations when fiber pullout dominates
When fiber pullout dominates, the fiber concrete tensile
stress distribution is assumed to be uniform.19,20 This assumption, together with equilibrium considerations in Fig.7,
yields the following equation for nominal flexural strength
of singly reinforced fibrous concrete sections

a
a (d c)

M n = As f y d + ( t (d c])b) c +
(20)

2
2
2

The tensile stress in fibrous concrete (st) is equal to


st=NF, where F can be calculated as follows when fiber
pullout dominates

f d f l f
F =
4

(21)

where tf is shear strength in fibrous concrete.


Hence,

Fig. 7Flexural stress distribution at ultimate condition in


singly reinforced fiber concrete section with predominant
fiber pullout.

0.5V f
t =
2
d f

lf
f d f
4

or
ACI Structural Journal/January-February 2014

df

(23)

At 0.025 cm (0.01 in.) crack width, considering the contribution of fibers, the stress level in steel reinforcement should
be less than that of the nonfibrous concrete. The stress in
steel reinforcement can be determined by the aforementioned
equation for fs.01, and multiplying by a decreasing index that
accounts for fiber effects. Using geometric relationships
for strain distribution, the maximum concrete stress can be
determined, and subsequently, the moment corresponding to
0.025 cm (0.01 in.) crack width can be determined.
For the pipes under consideration (Class IV wall C with
Di = 68 cm [27 in.], defined in the experimental section),
the steel reinforcement will develop stresses beyond its yield
stress, implying that the 0.025 cm (0.01 in.) crack strength
will be equal to the ultimate strength. This prediction was
confirmed by test results.
Flexural strength of fiber-reinforced concrete
pipes without steel reinforcement
The stress distribution under ultimate moment in concrete
pipes with fiber reinforcement (but no steel reinforcement) is
assumed to be linear19 (Fig. 8), with tensile stresses covering
the full area below the neutral axis (the expressions for st
were introduced in a previous section). Simple equilibrium
considerations yield the expression for flexural strength in
Fig. 8.
EXPERIMENTAL PROGRAM

(22)

0.5V f f l f

Materials
The materials used in concrete mixture were: 19 mm
(0.7in.) maximum size natural stone, natural sand (2NS),
portland cement (Type I), fly ash (ASTM Class F), plasticizer, and set-retarding admixture. The concrete mixture
design considered in this investigation, which is commonly
used in dry-mixed concrete pipe production, is shown in
87

Table 1Concrete mixture formulations


Materials common to all mixture
formulations

Proportion*

Aggregate (all natural stone)

2.21

Sand (2NS)

2.42

Plasticizer

6.83

Retarder

6.16

Fly ash

0.26

Water

0.44

Cement

0.74

Proportions are per unit weight of cementitious materials (cement + fly ash), except
for additives (plasticizer and retarder), which are in mL/kg of cementitious materials.

Table 2Properties of produced pipes Series I


Pipe

Steel reinforcement

Fiber content

Two layers

0%

One layer

0%

One layer

1.0% fine fiber*

One layer

1.5% fine fiber

One layer

0.75% coarse fiber

One layer

2% coarser fiber

One layer

2% coarser fiber

Fine PVA fiber (length = 6 mm [0.25 in.]; diameter = 0.026 mm [0.001 in.]).

Coarse PVA fiber (length = 12 mm [0.5 in.]; diameter = 0.1 mm [0.004 in.]).

Coarser PVA fiber (length = 15 mm [0.59 in.]; diameter = 0.3 mm [0.012 in.]).

Table 3Properties of produced pipes Series II


Pipe

Steel reinforcement

Fiber content

Two layers

0%

One layer

0%

No steel
reinforcement

0.5% fine fiber*

No steel
reinforcement

0.5% coarse fiber

No steel
reinforcement

2% coarser fiber

Fine PVA fiber (length = 6 mm [0.25 in.]; diameter = 0.026 mm [0.001 in.]).

Coarse PVA fiber (length = 12 mm [0.5 in.]; diameter = 0.1 mm [0.004 in.]).

Coarser PVA fiber (length = 15 mm [0.59 in.]; diameter = 0.3 mm [0.012 in.]).

Table1.21 Concrete materials were mixed following ASTM


C192/C192M22 recommendations, though steam curing
was used in lieu of moist curing in these series of tests.
The concrete used in the experimental program provided
34.5MPa (5000psi) compressive strength. The steel
reinforcement used in the project had a yield strength of
491MPa (71 ksi).
Concrete pipes were produced with different PVA fiber
types and volume fractions. The primary PVA fibers used
in the experimental work had an elastic modulus of 43GPa
(6237 ksi) and were relatively fine with a 0.026mm
(0.001in.) diameter, 6 mm (0.2362 in.) length and 1600MPa
(232 ksi) tensile strength. Then fibers were used at 0.5, 1.0,
and 1.5% volume fraction. Two other PVA fiber types
were also considered. One was relatively coarse with 0.1
mm (0.0039 in.) diameter and 12 mm (0.4724 in.) length
of 0.5 and 0.75% volume fraction with tensile strength of
88

Fig. 8Stress distribution at flexural failure in fiber-reinforced concrete section.


1100MPa (159 ksi). The other PVA fiber was even coarser,
with a 0.3 mm (0.01181 in.) diameter, 15 mm (0.5905 in.)
length and 900 MPa (130 ksi) tensile strength. This coarser
fiber was used in concrete at a 2% volume fraction.
Pipe production and experimental evaluation
Two distinct categories of pipes were produced: 1) with
one layer of steel reinforcement and fiber reinforcement; and
2) without steel reinforcement, but with fiber reinforcement.
Control pipe with two layers of steel reinforcement and no
fibers was also produced and tested. In the first category,
synthetic fibers were used to lower the steel ratio (from
two layers to one layer), and thus increase the protective
cover of concrete over steel while preserving the load-carrying capacity of the pipe. In the second category, synthetic
fibers fully replaced the structural steel. All pipes fabricated
and tested in the project had an internal diameter of 68 cm
(27in.) (ASTM C7623 Class IV, C-wall pipe), and a wall
thickness of 10 cm (4 in.). The pipes were produced in 2.8m
(8.5 ft) lengths. The reinforcing steel cage was 3 x 6 - W 2 x
W 2.5 (W 2 circumferential and W 2.5 longitudinal) welded
wire fabric. The pipes produced and tested in the first and
second series of experiments are introduced in Tables 2 and
3, respectively (the control pipe with two layers of steel
reinforcement is also included in these tests).
Three-edge bearing test procedures
In the three-edge bearing test, the pipe is supported on
a lower bearing of two parallel longitudinal strips, and the
load is applied uniformly along the pipe length using an
upper bearing strip. Both lower and upper bearing strips
are extended the full length of the pipe. Figure 4 shows the
three-edge bearing test configuration.
In this test, the load corresponding to 0.025 cm (0.01 in.)
crack width and also the peak load-carrying capacity of
concrete pipes are determined.
Experimental results
Figure 9 presents the load deflection relationships (in threeedge bearing tests) for the first series. In this figure, all pipes
have one layer of steel reinforcement, except for control pipe 1,
ACI Structural Journal/January-February 2014

Table 4Loads corresponding to 0.025 cm


(0.01in.) crack widths and peak capacity of
different pipes of first series
Pipe

Fig. 9Load-deflection curves in three-edge bearing tests


on first series of pipes with steel reinforcement (refer to
Table 2). (Note: 1 kg = 2.2 lb; 1 mm = 0.039 in.)
which has two layers of steel. Pipes 1 and 2 with two and one
layer of steel, respectively, were not fiber-reinforced. Adding
fiber with certain fiber types and volume fractions is observed
to improve ductility and load-carrying capacity of pipes. Table
4 presents loads corresponding to 0.025 cm (0.01in.) crack
width and also the peak load of the pipes in series I. The ultimate load capacity of the control pipe (pipe 2) with one layer of
steel was 21,000 kg (46,297 lb), which increased to 28,000 kg
(61,729 lb) for pipe 4 with 1% volume fine PVA fibers (33%
improvement). Fiber-reinforced pipe 4 also exhibited improved
ductility when compared with the control pipe 2. Although
lower ductility was observed for pipes 6 and 7, which used
coarser fiber (2% volume), improvement in load capacity for
these two pipes was observed. The coarser PVA fibers used in
pipes 6 and 7 at a relatively high volume fraction (2%) actually lowered the ductility of the concrete pipes. This could be
attributed to the damage for workability of fresh concrete and
dispersion of fibers introduced by the high volume fraction of
coarser fibers. These coarser fibers also had a relatively low
aspect (length-diameter) ratio (50 for coarser fibers versus 230
for fine fibers), which lowers their reinforcement efficiency in
concrete. The best balances of properties (strength and ductility)
were produced by pipe 4 (with 1.5% volume fine fiber) and pipe
5 (with 0.75% volume coarse fiber, which had an intermediate
coarseness with an aspect ratio of 120). It seems that 0.75%
volume of PVA fiber with an aspect ratio of 120 produces a
deserved balance of reinforcing effects and economy.
This preferred fiber reinforcement condition yields
concrete pipes with one layer of steel reinforcement that
compare well with pipe 2, with two layers of steel reinforcement in terms of load at 0.025 cm (0.01 in.) crack width
and ductility. Pipe 1 actually exhibits better ductility when
compared with pipe 1 at large deflection. It should be noted
that the peak load (generally neglected in designs) of pipe 1
is somewhat greater than that of pipe 4. Figure 10 presents
the load-deflections curve for pipes of series II without steel
reinforcement (with fibers) as well as control pipes 1 and 2
with two layers of steel reinforcement (and no fibers). The
load at 0.025 cm (0.01 in.) crack width as well as the ultimate load test results for pipes of series II is presented in
Table 5. Reinforcement with 0.5% vol. fine and coarse fibers
(with aspect ratios of 230 and 120) produced pipes with
ACI Structural Journal/January-February 2014

Load at 0.025 cm (0.01 in.)


crack width, kg (lb)

Load-carrying capacity,
kg (lb)

20,792 (45,838)

33,448 (73,740)

20,340 (44,842)

20,792 (45,838)

25,312 (55,803)

25,312 (55,803)

28,476 (62,779)

28,476 (62,779)

26,668 (58,793)

26,668 (58,792)

27,255 (60,088)

27,346 (60,287)

25,990 (57,298)

26,080 (57,497)

Table 5Loads corresponding to 0.025 cm


(0.01in.) crack width and peak capacity of
different pipes of second series
Pipe

Load at 0.025 cm (0.01 in.)


crack width, kg (lb)

Load-carrying capacity,
kg (lb)

20,792 (45,838)

33,448 (73,740)

20,340 (44,841)

20,566 (45,340)

23,504 (51,817)

23,504 (51,817)

22,600 (49,824)

22,600 (49,824)

24,182 (53,312)

24,182 (53,312)

load-carrying capacity and ductility levels which compared


well with those of control pipe 2 with one layer of steel reinforcement (and no fibers). The coarse fiber with a relatively
low aspect ratio of 50 and a relatively high volume fraction of 2% did not produce favorable results. This could be
attributed to the damaging effect of such a high fiber volume
fraction on fresh mixture workability and fiber dispersion.
Pipes 3 and 4 with 0.5% volume of higher aspect ratio fibers
and no steel reinforcement compared well against the control
pipes with two layers of steel, in terms of ductility and loads
at 0.025cm (0.01 in.) crack width.
Comparing theoretical predictions with
experimental results
Figures 11 and 12 compare the experimental and theoretical values of ultimate load for different pipes considered in
the experimental program. Figures 13 and 14 also compare
the theoretical value of loads at 0.025 cm (0.01 in.) crack
width for pipe series I and II, respectively. Theoretical
predictions were made based on observations of the predominance of either fiber rupture or pullout, as described below.
In series I tests, pipes 3 and 4 with finer fibers with an
aspect ratio of 230, fiber rupture dominated, and thus equations corresponding to fiber rupture were used in theoretical
investigation. In the case of pipes 5 and 7 (series I) and pipe
3 (series II), with fibers of medium aspect ratio (120), some
fiber pullout was observed but fiber rupture was still prevalent. In pipe 4 (series II), which incorporated coarse fibers
of relatively low (120) aspect ratio, a bond strength t of
10 MPa (1450 psi) yielded a satisfactory theoretical prediction of experimental results. (Using Eq. (20) and (22) with
the experimental results could provide basis for the value
of t.) Pipe 5 (series II) also contained coarse fibers which
exhibited predominantly pullout behavior. Fiber rupture was
assumed to be prevalent in pipe 6 (Series I).
89

Fig. 10Load-deflection curves in three-edge bearing tests


on the second series of pipes without steel reinforcement
(refer to Table 3). (Note: 1 kg = 2.2 lb; 1 mm = 0.039 in.)

Fig. 11Experimental and analytical predictions of ultimate loads for tested pipes Series I (refer to Table 2). (Note:
1 kg.m/m = 2.2 lb.in./in.)
The experimental results are observed in Fig. 11 and 12 to
occur within 10% of theoretical predictions. This finding
suggests that the theoretical models provide a reasonable
basis for predicting the ultimate load-carrying capacity of
concrete pipes with synthetic fiber reinforcement used alone
or in combination with conventional steel reinforcement.
CONCLUSIONS
To explore the potential for enhancement of service life
through refinement of structural design, PVA fibers of high
elastic modulus and desirable bonding to concrete were chosen
for use in concrete pipes. Concrete pipes were produced at
an industrial scale with different combinations of PVA fiber
and steel reinforcement. Fibers with different aspect ratios
(50-230) were used with volume fractions varying from 0.5%
to 2%. Three-edge bearing tests were conducted on pipes, and
their load-deflection behavior and those loads corresponding
to 0.025 cm (0.01 in.) crack width were obtained. Proper use
of synthetic fibers enabled reduction of steel reinforcement
and thus increased the protective cover of concrete over steel.
Depending on load-bearing requirements, synthetic fibers can
reduce the amount of steel reinforcement in concrete pipes by
50% or more. PVA fibers with aspect ratios of 120 to 230,
when used at a volume fraction of 0.5% to 0.75% in concrete
produced particularly desirable gains in the load-carrying
characteristics and durability of concrete pipes. Theoretical
models were developed for predicting the flexural strength
90

Fig. 12Experimental and analytical predictions of ultimate


loads for tested pipes Series II (refer to Table 3). (Note:
1kg.m/m = 2.2 lb.in./in.)

Fig. 13Experimental and analytical predictions of loads


at 0.025 cm (0.01 in.) crack width for tested pipes Series I
(refer to Table 2). (Note: 1 kg.m/m = 2.2 lb.in./in.)
and load-carrying capacity of concrete pipes reinforced with
secondary synthetic fibers, with or without conventional steel
reinforcement. The models developed here account for the
contribution of fibers to the tensile load-carrying capacity of
concrete through pullout or rupture modes of failure at cracks.
The models were verified using outcomes of experiments on
pipes with different combinations of steel and synthetic fiber
reinforcement, and were found to provide a reasonable basis
for prediction of the load-carrying capacity of concrete pipes.
AUTHOR BIOS

Amirpasha Peyvandi is currently with HNTB Corporation as a Structural/Bridge Engineer. He received his BS and MS in civil and structural
engineering from the University of Tehran, Tehran, Iran. He also received
his PhD in civil and environmental engineering at Michigan State University, East Lansing, MI. His research interests include application of nanotechnology in cementitious material and development of environmentally
friendly and energy-efficient construction materials and systems.
Parviz Soroushian, FACI, is a Professor of civil and environmental engineering at Michigan State University. He received his BS from the University of Tehran and his MS and PhD from Cornell University, Ithaca, NY. His
research interests include materials science and engineering and environmentally friendly and energy-efficient construction materials and systems.
Shervin Jahangirnejad is a PhD Candidate in the Civil and Environmental Engineering Department at Michigan State University. He received
his BS in civil engineering from Azad University, Najafabad, Isfahan, Iran,
in 2000, and his MS in civil engineering from Michigan State University in
2005. His research interests include civil engineering materials and pavement engineering.

ACI Structural Journal/January-February 2014

Fig. 14Experimental and analytical predictions of loads


at 0.025 cm (0.01 in.) crack width for tested pipes Series II
(refer to Table 3). (Note: 1 kg.m/m = 2.2 lb.in./in.)
ACKNOWLEDGMENTS

The authors wish to acknowledge the support of Northern Concrete Pipe,


Inc. in Charlotte, MI, toward performance of the pipe tests. The support of
the U.S. EPA (Contract 68-0-03-065) for the project is also acknowledged.

As
a
C
Cm1
Cm2
Di
DLut
d
d f
F
fs.01

=
=
=
=
=
=
=
=
=
=
=

NOTATION

steel reinforcement area


depth of compressive stress block
neutral axis depth in concrete section
defined constant coefficient
defined constant coefficient
internal diameter of pipe
test load that cracks pipe
effective depth of concrete section
fiber diameter
mean fiber frictional bond resistance
maximum stress in reinforcement when maximum crack width
is 0.025 cm (0.01 in.)
fy
= reinforced steel yield stress
h
= pipe wall thickness
j
= coefficient for moment arm at service load stress
Kmr = coefficient that varies with wall thickness of pipe
Lf = length of fiber
M = moment at different pipes section
MU = ultimate flexural strength of reinforced concrete
N
= number of fibers per unit area
S
= internal diameter of pipe
U = ultimate tensile force of single fiber
Vf = fiber volume fraction
Wp = weight of pipe
Wt = total load of three-edge bearing test
fmr = strength reduction factor for flexural tension
l
= redaction factor for variability in test results
sfu = ultimate tensile stress tolerated by fibers
st = tensile stress in fibrous concrete
tf
= shear strength in fibrous concrete

REFERENCES

1. Shook, W. E., and Bell, L. W., Corrosion Control in Concrete Pipe


and Manholes, Technical Presentation, Water Environmental Federation,
Orlando, FL, 1998, pp. 1-5.
2. Fuente, A.; Escariz, R. C.; Figueiredo, A. D.; Molins, C.; and Aguado,
A., A New Design Method for Steel Fibre Reinforced Concrete Pipes,
Construction and Building Materials, V. 30, 2012, pp. 547-555.

ACI Structural Journal/January-February 2014

3. Mirsayah, A. A., and Banthia, N., Shear Strength of Steel Fiber-Reinforced Concrete, ACI Materials Journal, V. 99, No. 5, Sept.-Oct. 2002,
pp. 473-479.
4. Naaman, A. E., Pull-Out Mechanism in Steel Fiber-Reinforced
Concrete, Journal of the Structural Division, ASCE, V.102, No. 8, 1976,
pp. 1537-1548.
5. Batson, G., Steel Fiber Reinforced Concrete, Materials Science and
Engineering, V. 25, 1976, pp. 53-58.
6. Soroushian, P., and Lee, C. D., Distribution and Orientation of Fibers
in Steel Fiber Reinforced Concrete, ACI Materials Journal, V. 87, No. 5,
Sept.-Oct. 1990, pp. 433-439.
7. Olivito, R. S., and Zuccarello, F. A., An Experimental Study on the
Tensile Strength of Steel Fiber Reinforced Concrete, Composites. Part B,
Engineering, V. 41, No. 3, 2010, pp. 246-255.
8. Bencardino, F.; Rizzuti, L.; Spadea, G.; and Ramnath, N.S., StressStrain Behavior of Steel Fiber-Reinforced Concrete in Compression,
Journal of Materials in Civil Engineering, ASCE, V. 20, No. 3, 2008,
pp.255-263.
9. Peyvandi, A.; Soroushian, P.; Balachandra, A. M.; and Sobolev, K.,
Enhancement of the Durability Characteristics of Concrete Nanocomposite
Pipes with Modified Graphite Nanoplatelets, Construction and Building
Materials, V. 47, 2013, pp. 111-117.
10. Heger, F. J., Structural Behavior of Circular Reinforced Concrete
Pipe-Development of Theory, ACI Journal, V. 60, No.11, Nov. 1963,
pp. 1567-1614.
11. Peyvandi, A., and Soroushian, P., Structural Performance of
Dry-Cast Concrete Nanocomposite Pipes, Materials and Structures, Oct.
2013, DOI: 10.617/s11527-013-0196-0.
12. Haktanir, T.; Ari, K.; Altun, F.; and Karahan, O., A Comparative
Experimental Investigation of Concrete, Reinforced-Concrete and SteelFibre Concrete Pipes Under Three-Edge-Bearing Test, Construction and
Building Materials, V. 21, No. 8, Aug. 2007, pp.1702-1708.
13. Banthia, N.; Bindiganavile, V.; Jones, J.; and Novak, J., Fiber Reinforced Concrete in Precast Concrete Applications: Research Leads to Innovative Products, PCI Journal, Summer 2012, pp. 33-46.
14. American Concrete Pipe Association, Standard Installations and
Bedding Factors for the Indirect Design Method, Design Data9, Oct.
2007, pp. 1-11.
15. ASCE 27-00, Standard Practice for Direct Design of Precast
Concrete Pipe for Jacking in Trenchless Construction, ASCE, Reston, VA,
2000, 62 pp.
16. ASCE 15-98, Standard Practice for Direct Design of Buried Precast
Concrete Pipe Using Standard Installations, ASCE, Reston, VA, 2000,
50pp.
17. Hulatt, J.; Hollaway, L.; and Thorne, A., The Use of Advanced
Polymer Composites to Form an Economic Structural Unit, Construction and
Building Materials, V. 17, No. 1, 2003, pp. 55-68.
18. Prudencio, L.; Austin, S.; Jones, P.; Armelin, H.; and Robins,P.,
Prediction of Steel Fibre Reinforced Concrete Under Flexure From an
Inferred Fibre Pull-Out Response, Materials and Structures, V. 39, No.6,
2006, pp. 601-610.
19. Ezeldin, A. S., Optimum Design of Reinforced Fiber Concrete
Subjected to Bending and Geometrical Constraints, Computers & Structures, V. 41, No. 5, 1991, pp. 1095-1100.
20. Yang, J. M.; Min, K. H.; Shin, H. O.; and Yoon, Y. S., Effect of
Steel and Synthetic Fibers on Flexural Behavior of High-Strength Concrete
Beams Reinforced with FRP Bars, Composites Part B: Engineering,
V.43, No. 3, Apr. 2012, pp. 1077-1086.
21. Soroushian, P.; Chowdhury, H.; and Tewodros, G., Evaluation
of Water-Repelling Additives for Use in Concrete-Based Sanitary Sewer
Infrastructure, Journal of Infrastructure Systems, V. 15, No. 2, pp. 2009,
pp. 106-110.
22. ASTM C192/C192M-07, Standard Practice for Making and Curing
Concrete Test Specimens in the Laboratory, ASTM International, West
Conshohocken, PA, 2007, 8 pp.
23. ASTM C76-12, Standard Specification for Reinforced Concrete
Culvert, Storm Drain, and Sewer Pipe, ASTM International, West Conshohocken, PA, 2012, 11 pp.

91

NOTES:

92

ACI Structural Journal/January-February 2014


View publication stats

You might also like