Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Residual Prestress Forces and Shear Capacity of Salvaged

Prestressed Concrete Bridge Girders


Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

G. Parry Osborn, S.M.ASCE1; Paul J. Barr, M.ASCE2; David A. Petty, S.M.ASCE3;


Marvin W. Halling, F.ASCE4; and Travis R. Brackus5

Abstract: Seven prestressed concrete bridge girders that had been in service for 42 years, and represented two span lengths and reinforce-
ment designs, were tested to determine their effective prestress force and ultimate shear capacity. A cracking moment test was used to
determine the effective prestress force in the girders. The measured effective prestress force was compared with calculated values according
to the AASHTO LRFD prestress loss equations to investigate their adequacy. The AASHTO refined method was shown to provide the most
accurate results to within 10% of the measured values. An ultimate shear test was also performed on two of the girders. An external load
was applied near the support and increased until the girder failed in shear. The various procedures in the AASHTO LRFD specifications
were compared with the measured results. The AASHTO simplified procedure predicted only 51% and 39% of the average measured shear
capacity for the short and long span girders, respectively. The strut-and-tie models were found to estimate the shear capacity more accurately.
The AASHTO refined method was shown to provide the most accurate results. DOI: 10.1061/(ASCE)BE.1943-5592.0000212. © 2012
American Society of Civil Engineers.
CE Database subject headings: Prestressed concrete; Finite element method; Experimentation; Concrete bridges; Girder bridges.
Author keywords: Prestress concrete; Shear capacity; Prestress losses; Finite-element modeling; Experimental testing.

Introduction steel in prestressed concrete beams affected the overall shear behav-
ior of girders. They found that as the area of the prestressing strand
The in-service structural properties of prestressed concrete girders increased, the shear capacity increased even when the total
are difficult to accurately determine, especially when the girders prestress force was held constant. By increasing the area of the
have been in the field for an extended period of time. While mild flexural steel reinforcement, the overall shear capacity also
there are several, two key factors that affect the performance of increased. Oh and Kim (2004) tested the shear capacities of large-
prestressed concrete girders are the following: (1) the effective pre- scale, prestressed concrete beams and concluded that beams with
stress force and (2) the shear capacity near the supports. Since these higher concrete strength exhibited more diagonal cracks with a
properties are very difficult to measure in situ, accurate analytical smaller crack width and that the principal directions decreased
methods are required. as the load increased. They also concluded that the concept of aver-
Current design methodologies provide some guidance on how to age strains and the changing of principal directions according to the
determine these two properties. According to the AASHTO LRFD applied load can be used for a more realistic shear analysis and that
specifications (2009), total prestress losses can be determined by the girders with higher-strength concrete showed an increase in the
using a time-dependent or lump-sum procedure. The AASHTO ultimate shear capacity. Kaufman and Ramirez (1988) tested six
LRFD specifications also provide two sectional procedures and AASHTO girders made with high-strength concrete to gain a better
one strut-and-tie approach for calculating the shear capacity. understanding of both the flexural and the shear behavior of pre-
Various researchers have studied both of these issues. Saqan and stressed concrete girders. They found that the higher-strength con-
Frosch (2009) performed a study to determine how the flexural crete increased the ultimate capacities for both shear and flexure.
1
MacGregor et al. (1965) tested 104 prestressed concrete beams to
Graduate Research Assistant, Dept. of Civil and Environmental Engi- determine their ultimate shear capacities and to understand their
neering, Utah State Univ., 4110 Old Main Hill, Logan, UT 84332-4110. overall shear behavior. Many different beam configurations were
2
Associate Professor, Dept. of Civil and Environmental Engineering,
Utah State Univ., 4110 Old Main Hill, Logan, UT 84332-4110 (corre-
tested to quantify important variables that affect the shear behavior
sponding author). E-mail: paul.barr@usu.edu of prestressed concrete beams and, on the basis of the findings,
3
Graduate Research Assistant, Dept. of Civil and Environmental Engi- proposed relationships that have been used over the years. Other
neering, Utah State Univ., 4110 Old Main Hill, Logan, UT 84332-4110. studies include Naito et al. (2010), Kordina et al. (1989), and
4
Associate Professor, Dept. of Civil and Environmental Engineering, Elzanaty et al. (1986).
Utah State Univ., 4110 Old Main Hill, Logan, UT 84332-4110. E-mail: The effective prestress force for a prestressed concrete girder is
marv.halling@usu.edu an important property required to determine the service behavior
5
Graduate Research Assistant, Dept. of Civil and Environmental Engi- and the ultimate shear and flexural capacity of a girder. Several
neering, Utah State Univ., 4110 Old Main Hill, Logan, UT 84332-4110.
researchers have performed studies to determine the effective pre-
Note. This manuscript was submitted on July 15, 2010; approved on
December 22, 2010; published online on December 27, 2010. Discussion
stress force in a girder. Barr et al. (2008) instrumented five pre-
period open until August 1, 2012; separate discussions must be submitted stressed, high-performance concrete bridge girders and recorded
for individual papers. This paper is part of the Journal of Bridge Engineer- the changes in prestress for 3 years from the time of casting. They
ing, Vol. 17, No. 2, March 1, 2012. ©ASCE, ISSN 1084-0702/2012/2- found that methods provided in previous editions of the AASHTO
302–309/$25.00. LRFD specifications for predicting prestress losses overpredicted

302 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012

J. Bridge Eng., 2012, 17(2): 302-309


the losses by 20%. They also compared measured results with
the Report 496 (Tadros et al. 2003) procedures, which were more
adaptable, and predicted losses within 10% of the average mea-
sured losses. Onyemelukwe et al. (2003) studied a newly con-
structed bridge to compare code-predicted prestress losses to
measured values. Time-dependant losses were measured from
the time of construction. Losses were found to be nonuniform
through the depth of the girder as is implied in most codes.
Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

Azizinamini et al. (1996) developed a method of nondestructively


testing for the effective prestress force in prestressed concrete gird-
ers and compared their proposed method with a standard cracking
test. The researchers found their proposed method advantageous
because it was on the basis of stresses, not time, therefore elimi-
nating problems with displacements, creep, and relaxation. The
researchers also found that the measured prestress losses were
on average about 20% lower than most code-predicted values after
Fig. 1. I-215/45th South bridge during demolition
25 years. Halsey and Miller (1996) used 40-year-old girders to de-
termine ultimate capacities by using destructive tests. During the
testing, the researchers employed three methods to determine
the effective prestress force. First was a cracking moment test, sec-
ond was a crack opening test, and third was a strand cut test. They
compared all three methods with code predictions that were used
when the girders were designed and found the measured values
were within 7 to 14% of predicted values. Other studies include
Trautner et al. (2010) and Civjan et al. (1998).
With several new bridges being built in Utah, researchers at
Utah State University obtained several of the decommissioned
bridge girders and tested them for their ultimate shear capacities
and effective prestress force. These girders had been in service
for over 40 years and represented two span lengths and reinforce-
ment designs. The girders provided researchers with a unique op-
portunity to compare measured and predicted values of in-service,
full-scale girders.

Bridge and Girder Description

A total of seven bridge girders were tested as part of this research.


Girders 1 through 6 were salvaged from Interstate 215 (I-215) in
Salt Lake City, Utah, at 45th South. The bridge was built in 1968 as
a four-span bridge with span lengths of 7.1, 14.5, 14.5, and 20.4 m
(22.3, 74.5, 74.5, and 67 ft). The bridge had a change of elevation
of about 13.1 m (43.0 ft) from one end to the other. Girders 1
through 6 obtained for this research were from the shortest span
of this bridge. The bearing center-to-center span length of these Fig. 2. Girder cross-section and stirrup spacing
six girders was 7.1 m (22.3 ft) with an outside-to-outside dimension
of 7.2 m (23.6 ft). The girders were spaced at 2.7 m (9.0 ft) on
center. In the field, the girders were made composite with a actual compressive strength according to ASTM standards. These
20.3 cm (8.0 in.) thick reinforced concrete deck. At the time the tests resulted in an average measured concrete compressive strength
girders were delivered to the researchers, only a portion of the deck (f 0c ) of 48.9 MPa (7,100 psi).
directly above the top flange was still intact. To prepare the girders According to the bridge plans, the prestressing force at jacking
for testing, the concrete decking was squared to provide a more was 964 kN (217 kip). The prestressing force after losses was speci-
uniform test specimen. Fig. 1 shows a photo of the bridge during fied as 783 kN (176 kip) at an eccentricity (cgp) of 27.9 cm
the demolition process. (11.0 in.) as measured from the bottom of the girder (Fig. 2).
All six girders were AASHTO Type II and were reinforced for The prestressing strands that were used in Girders 1 through 6 were
shear with 13 mm (No. 4) bars used as stirrups (Fig. 2). The first 12, 11.1 mm (7=16 in:) diameter, 7-wire straight strands. These
stirrup spacing (Si ) was placed 19.1 cm (7.5 in.) from the center of strands were tested in the lab to establish their ultimate stress,
the bearing and then 58.4 cm (23.0 in.) on center afterward (S). which was measured as 1,780 MPa (259 ksi). From these results
Representative samples of shear-reinforcing steel were removed and discussions with state officials, it was concluded that the speci-
from the girder and tested. The yield stress of the shear reinforce- fied grade for the strands used in these girders was 1,720 MPa
ment was measured as 230 MPa (33.4 ksi). (250 ksi) stress-relieved strands, which were commonly used
The concrete compressive strength of the girder at transfer (f 0ci ) during this time period.
was specified as 27.6 MPa (4,000 psi). Two concrete samples were In addition to the first six girders, one longer AASHTO Type II
removed from a girder and tested in compression to determine the girder (Girder 7) was also tested. This girder was obtained from the

JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012 / 303

J. Bridge Eng., 2012, 17(2): 302-309


10th West/1400 North Bridge in Salt Lake City, Utah. This bridge decompression load was defined as the magnitude of load at which
had also been in service for approximately 40 years. The span the strain initiated a nonlinear response that deviated from the
length of Girder 7 was 10.5 m (34.5 ft) and was in a similar con- initially straight line. This methodology is modeled on findings of
dition as the first six girders. previous researchers (i.e., Pessiki et al. 1996). Fig. 3 shows a typ-
The prestressing force for Girder 7 was applied with 14, ical strain and load relationship obtained from the cracking test
11.1 mm (7=16 in:) diameter, 7-wire prestressing strands, which data. As illustrated, the decompression load was determined as
imparted a total prestressing force of 1,177 kN (264.6 kip) onto 136 kN (30,500 lb) for Girder 2.
the beams at a centroidal distance (cgp) of 24.0 cm (9.5 in.) as mea- Once the decompression load of each girder was experimentally
Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

sured from the bottom of the beam (Fig. 2). The ultimate capacity determined, it was then used to calculate the effective prestress
of the strands was the same as Girders 1 through 6 [1,720 MPa force in the girder. At service, the three factors that contribute
(250 ksi)]. The compressive strength of this girder was specified to the stress in the bottom of a prestressed girder are as follows:
as 34.5 MPa (5,000 psi) but was experimentally determined to (1) self-weight of the girder, (2) prestress force in the girder,
be 64.1 MPa (9,300 psi). The shear-reinforcing steel also had and (3) externally applied load. The equation to calculate this stress
the same properties as Girders 1 through 6 with a yield stress of can be written as Eq. (1). Each of the girders had slightly different
230 MPa (33.4 ksi). For Girder 7, the shear reinforcement initiated cross-sectional properties owing to the amount of decking that re-
(Si ) at 15.2 cm (6.0 in.) from the center of the support and was then mained above the top flange of the girder. Therefore, detailed sec-
subsequently spaced (S) at 43.2 cm (17.0 in.) on center (S) through- tion properties were measured to accurately calculate the prestress
out the remaining length of the girder (Fig. 2). force for each girder

Pe Pe eyg M sw yg M max yc
σ¼ þ   ð1Þ
Cracking Moment Test Ag Ig Ig Ic

Accurate knowledge of the effective prestress force is vital to many


where σ = stress at the crack location; Pe = effective prestress force
aspects of the overall structural health and performance of a pre-
in the beam; Ag = cross-sectional area at the crack location; e =
stressed concrete girder, including the cracking load, ultimate mo-
eccentricity of the prestressing force at the crack location; yg =
ment strength, and shear capacity. Several proposed methodologies
neutral axis location of the girder measured from the bottom of
have been suggested by researchers to experimentally determine
the beam at the crack location; I g = moment of inertia of the girder
the effective prestress force in a prestressed concrete girder. Of
at the crack location; M sw = moment at the crack location owing to
these methods, one of the most commonly accepted is a cracking
self-weight of the girder; M max = maximum moment in the beam
moment test.
owing to externally applied load; yc = neutral axis location of the
Cracking moment tests were performed on Girders 1 through 5
composite section measured from the bottom of the beam at the
to determine the effective prestress force after 42 years of service.
crack location; and I c = moment of inertial of the composite section
Each girder was positioned under a reaction frame attached to a at the crack location.
strong floor at the Utah State University’s Systems, Materials, Table 1 lists each of the girders’ externally applied loads with
and Structural Health Laboratory (SMASH Lab). The girder ends their corresponding calculated decompression loads as determined
were simply supported at the bearing locations. This setup allowed by using Eq. (1). The first load listed is on the basis of the initial test
for an overall span length of 6.8 m (22.5 ft). A single-point load was in which an externally applied load of 310 kN (70,000 lb) was
applied at midspan of each girder with a hydraulic ram. The mag- reached before the test was concluded, and the second load listed
nitude of the load was slowly increased at the midspan until a is on the basis of the test in which an externally applied load of
clearly visible crack had propagated across the bottom flange of 360 kN (80,000 lb) was reached before the test was concluded.
the prestressed girder. Once the crack was identified, the load The average calculated residual prestress force for each girder is
was held constant, and the crack was traced with a permanent listed in the fourth column and is the force that was considered
marker for easy identification once the load was removed and as the residual prestressing force in the girders. The average effec-
the prestress force closed the crack. The crack locations were iden- tive prestress force for all girders was 727.3 kN (163.5 kip). The
tified for Girders 1 through 5 following this same procedure. differences between the two load tests were within 5% for three of
After each girder had undergone this initial cracking load, a
2-in. foil-strain gauge was attached to the bottom flange of the gird-
ers directly over the crack. This gauge was used to quantify the load
at which the crack reopened, and therefore the stress at the extreme
tension fiber of the concrete girder was equal to zero.
Two series of postcrack load tests were performed on each
girder to allow the crack to dilate sufficiently yet not permanently
yield the strain gauges. The initial test applied a monotonic load on
the girder up to 310 kN (70,000 lb), and the second was to apply the
monotonic load on the girder up to 360 kN (80,000 lb). During each
test, changes in strain were recorded from the foil-strain gauge on
the bottom flange along with corresponding values of the load,
which were measured with a load cell and a pressure transducer
in line with the hydraulic ram. All data were recorded at a sampling
rate of 10 Hz throughout each test.
At the conclusion of each test, the recorded data were plotted as
microstrain versus load at the crack location. The experimentally
Fig. 3. Typical load versus microstrain plot during cracking test
determined decompression load was obtained by fitting a straight
(Girder 2 shown at 310 kN)
line to the linear portion of the load versus strain diagram. The

304 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012

J. Bridge Eng., 2012, 17(2): 302-309


Table 1. Decompression Loads and Calculated Residual Prestress Force and the stress at the centroid of the prestressing steel can be calcu-
Residual Average residual lated at different stages of construction by using this method
Decompression prestressing prestressing
load force force Δf pLT ¼ ðΔf pSR þ Δf pCR þ Δf pR1 Þid
Percent
Girder kN kip kN kip kN kip loss
þ ðΔf pSD þ Δf pCD þ Δf pR2  Δf pSS Þdf ð3Þ
1 157.9 35.5 717.1 161.2
704.0 158.3 27.1
1 151.2 34.0 690.9 155.3
Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

where Δf pSR = prestress loss attributable to shrinkage of girder con-


2 135.7 30.5 616.7 138.6
646.2 145.3 33.0 crete pbetween
ffiffiffiffiffiffi transfer and deck placement; Δf pCR ¼ V cw ¼
2 151.2 34.0 675.6 151.9 ð0:06 f 0 c þ 0:30f pc Þbv d v þ V p prestress loss attributable to creep
3 162.4 36.5 793.2 178.3
783.5 176.1 18.8 of girder concrete between transfer and deck placement; Δf pR1 =
3 157.9 35.5 773.8 174.0 prestress loss attributable to relaxation of prestressing strands be-
4 166.8 37.5 756.1 170.0 tween time of transfer and deck placement; Δf pR2 = prestress loss
769.2 172.9 20.3
4 173.5 39.0 782.3 175.9 attributable to relaxation of prestressing strands in the composite
5 155.7 35.0 707.6 159.1 section between time of deck placement and final time; Δf pSD =
733.8 165.0 24.0 prestress loss attributable to shrinkage of girder concrete between
5 169.0 38.0 760.1 170.9
Average prestress force 727.3 163.5 24.6 time of deck placement and final time; Δf pCD = prestress loss
attributable to creep of girder concrete between time of deck place-
ment and final time; Δf pSS = prestress gain attributable to shrinkage
of deck in the composite section.
the girders. The largest difference was approximately 10% for Table 2 lists the various calculated and measured effective pre-
Girder 2. stress forces. In general, each of the estimated prestress forces was
higher than the measured results. The calculated effective prestress
force for Girders 1 through 5 was calculated as 838.1 kN (188.4
Comparison of Residual Prestress Force kip) by using the AASHTO LRFD approximate method. Using this
method, the calculated effective prestress force was 15% higher
The AASHTO LRFD specifications (2009) provide two methods of than the average measured prestress force in the girders. When
calculating the long-term prestress losses for prestressed concrete the effective prestress force was calculated by using the AASHTO
bridge girders. The first is an approximate method that lumps the LRFD refined method, an effective prestress force of 799.6 kN
long-term losses into three simplified terms. The second method is (179.8 kip) was obtained. This calculated value was 10% higher
a refined method that provides individual expressions for calculat- than the average measured prestress force from the cracking tests.
ing the interdependent long-term losses. For both methods, the in- One explanation of the difference between the AASHTO LRFD
stantaneous elastic shortening loss at transfer is calculated directly values and measured results may be that the AASHTO LRFD
by using transformed section properties. methods were developed on the basis of high-strength concrete,
The long-term prestress losses for the approximate method which is expected to have lower prestress losses (Tadros et al.
(Δf pLT ) are calculated from Eq. (2). These long-term losses include 2003). The maximum difference was 110.8 kN (24.9 kip) from
losses that result from creep, shrinkage of the concrete, and relax- the AASHTO approximate method, with the bridge plan specifica-
ation of the prestressing steel tions, which was the AASHTO standard specifications at the time
when the bridge was designed—the closest at a difference of only
f pi Aps
Δf pLT ¼ 10:0 γ γ þ 12:0γh γst þ Δf pR ð2Þ 55.6 kN (12.5 kip). This difference has implications for service load
Ag h st stresses, deflections, and shear capacities for girders with harped
strands.
where γh = humidity factor; γst = concrete compressive strength
factor; f pi = prestressing immediately before transfer; Aps = total
area of the prestressing steel; Ag = gross area; H = average ambient Shear Test Description
humidity as a percent; f 0ci = specified initial concrete compressive
strength; and Δf pR = relaxation loss. In addition to the cracking tests, shear capacity tests were
This loss in strand stress is applied at the centroid of the pre- performed for Girders 6 and 7. The goal of the shear tests was
stressing steel and the transformed concrete section resulting in to quantify the ultimate shear capacity, when the load is applied
an effective prestress force at service. Using the effective prestress near the critical shear location. Ultimate shear tests were performed
force at service, concrete service stresses can be calculated by using on each end of each girder independently, providing a total of four
the transformed section properties and compared with the design shear tests.
stress limits. Before testing, a high-strength grout pad was cast above the
The refined method is a slightly more involved process, but is decking on each girder centered at a distance of 122 cm (48.0 in.)
intended to lead to more accurate estimates of prestress losses. This from the center of the support. This distance was equal to the depth
method involves calculating creep, shrinkage, and relaxation losses of the AASHTO Type II (d) girder plus 30.5 cm (12.0 in.)
before and after casting independently and then summing their (a=d ¼ 1:5). This pad was cast to provide a uniform surface for
values to obtain the total loss. The elastic shortening losses are cal- the hydraulic ram to apply the load.
culated the same way as for the approximate method. Eq. (3) is the For each test, the girders were simply supported under a reaction
general equation to calculate the long-term losses. The subscript id frame as shown in Figs. 2 and 4. The span lengths of each girder
is used to denote calculated losses from the time of strand transfer varied owing to the fact that as soon as one girder end was tested to
until the time the deck is made composite, and the subscript df is failure it became necessary to move the support to a location
used to denote losses that occur after the deck has been made outside the failure region. The shear spans (a in Fig. 2) for each
composite until the end of service life. The bottom fiber stress test were kept constant at 122 cm (48.0 in.) with the exception

JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012 / 305

J. Bridge Eng., 2012, 17(2): 302-309


Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Typical shear test setup Fig. 5. Typical shear test failure

of one end of Girder 7. This shear span was slightly larger at vertical surface strain on the concrete was found to be 0.004,
131 cm (51.5 in.) because of the deck material. indicating that the stress in the reinforcing steel at failure was above
The externally applied load was measured in two ways. This the yield stress of the steel showing that yielding likely occurred.
redundancy was desirable since the ultimate applied load was
the critical value for this testing. The load was measured by a Calculation of Shear Capacity
strain-gauge-based load cell and with a pressure transducer in line
with the hydraulic ram. In addition to the load, deflections were For this research, three predictive methods were used to compare
recorded with an LVDT, and strain gauges attached to the concrete current code practices with the measured results. The first method-
surface of the girder measured changes in strain. The LVDT was ology was the simplified method provided in the AASHTO LRFD
positioned on the top of the girder adjacent to the load cell and ram. specifications (2009). This is the preferred method for most state
The hydraulic ram was used to gradually apply an increasing DOTs when designing or analyzing bridges. The second predictive
load through failure. The applied load was continuously monitored method was the nominal method in the AASHTO LRFD specifi-
and recorded during the test. Once the applied maximum load had cation (2009). The third method used was a strut-and-tie model
decreased by more than 20%, the test was terminated by completely (STM). This method is described in the appendix of the American
removing the load, after which the girder and data were examined. Concrete Institute (ACI) building code (2008). The AASHTO
Table 3 lists the recorded ultimate shear capacities from each test. LRFD specifications also allow for a STM to be used in the design
The ultimate shear capacity was obtained on the basis of the and analysis of prestressed concrete girders for shear capacity.
maximum load and the load configuration. The ultimate load In general, the shear capacity (V n ) of a member is determined
was defined as the maximum recorded load and was used to cal- as a combination of the shear contribution of individual parts. In a
culate the ultimate shear force. The average shear capacity was de- prestressed concrete girder there are three main contributors to the
termined as 727.7 kN (163.6 kip) for Girder 6 and 1,163.2 kN shear strength, namely, transverse steel (V s ), the vertical component
(261.5 kip) for Girder 7. As all testing parameters were constant, of the prestressing force (V p ), and the shear resistance of the con-
the difference in capacities is believed to be because of the smaller crete (V c ). The general governing equation is provided as Eq. (4)
stirrup spacing and larger concrete strength of Girder 7.
For Girder 6, the failure mechanism was flexural shear where Vn ¼ Vc þ Vs þ VP ð4Þ
the cracks first developed at the bottom of the girder at an angle The simplified and nominal methods in the AASHTO LRFD
of 90° from the longitudinal axis. After the cracks propagated specifications (2009) allow the shear capacity of prestressed con-
up through the bottom flange, the orientation changed as the shear crete girders to be calculated on the basis of a sectional analysis.
forces dominated the flexural effects, and the cracks propagated For the simplified method, the calculations for determining V c are
at an angle of approximately 42°. The primary shear crack was very similar to the ACI procedure. The value of V c is calculated in
accompanied by other shear cracks but was smaller in size. The two different ways, depending on the manner the shear cracks de-
accompanying cracks were generally parallel to the primary shear velop, namely, flexure-shear cracking or web-shear cracking. If
crack but with less dilation. The cracks then opened until there flexure-shear cracks control the design, the value V ci [Eq. (5)]
was not enough aggregate interlock or friction to hold the girder should be used as V c . However, if web-shear cracks control the
together, at which point the girder experienced a sudden failure. design, V cw is to be used [Eq. (6)]. V c is defined to be the lesser
Girder 7 failed in a similar manner but at a larger load. More of V ci and V cw
information regarding the cracking behavior of the girders can
be found in Osborn (2010). Fig. 5 shows a picture of a typical pffiffiffiffi VM pffiffiffiffi
V ci ¼ 0:02 f 0 cbv d v þ V d þ i cre ≥ 0:06 f 0 cbv d v ð5Þ
web-cracking failure observed for Girder 7. M max
The strain was measured at the concrete surface. The strain mea-
pffiffiffiffiffiffi
surements on the concrete surface were used to estimate the strains in V cw ¼ ð0:06 f 0 c þ 0:30f pc Þbv d v þ V p ð6Þ
the shear-reinforcing steel at failure. On the basis of the tension test
results, the yield stress of the vertical stirrups was 230 MPa (33.4 ksi). where V d = shear force at section attributable to unfactored
This yield stress occurred at a strain of 0.001, yet the maximum dead load, including both the dead load of the structural

306 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012

J. Bridge Eng., 2012, 17(2): 302-309


components (DC) and the dead load of the wearing surface (DW); values. Fig. 6 shows the STM used to analyze Girder 7. More in-
V i = factored shear force at section attributable to externally applied formation regarding the STMs can be found in Osborn (2010).
loads occurring simultaneously with M max ; f 0 c = concrete compres-
sive strength; bv = effective web width; d v = effective shear depth;
M cre = moment causing flexural cracking at section attributable to Shear Capacity Comparison
external loads; M max = maximum factored moment at section attrib-
utable to externally applied loads; f cpe = compressive stress in con- The ultimate shear capacities calculated for this research by using
crete attributable to effective prestress forces only (after all losses) each of the described analytical methods are presented in Table 3.
Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

at extreme fiber of section where tensile stress is caused by exter- The calculated values are also compared to the measured values and
nally applied loads; M dnc = total unfactored dead load moment act- shown as a percentage of the measured values.
ing on the monolithic or noncomposite section; Sc = section The average measured shear capacity was 727.7 kN (163.6 kip)
modulus for the extreme fiber of the composite section where ten- for Girder 6 and 1,163.2 kN (261.5 kip) for Girder 7. Using
sile stress is caused by externally applied loads; Snc = section the AASHTO simplified methodology, calculated shear values
modulus for the extreme fiber of the monolithic or noncomposite of 366.0 kN (82.3 kip) and 446.1 kN (100.3 kip) were obtained
section where tensile stress is caused by externally applied loads; for Girders 6 and 7, respectively. The AASHTO equations pre-
f pc = compressive stress in concrete (after allowance for all pre- dicted only 51% of the average measured shear capacity for Girder
stress losses) at centroid of cross-section-resisting externally ap- 6 and 39% for Girder 7. The AASHTO nominal method predicted a
plied loads or at junction of web and flange. calculated shear capacity of 274.4 kN (61.7 kip) and 323.3 kN
The AASHTO nominal method is on the basis of the modified (72.7 kip) for Girders 6 and 7, respectively. These values are
compression field theory. Once cracking has occurred, according to 38% and 28% of the average measured values for Girders 6 and 7,
this method, the web reinforcement resists the tensile stresses, and respectively. Calculated shear capacities from the STM were
the concrete struts carry the compressive stresses. The nominal 616.3 kN (138.6 kip) and 1,150.8 kN (258.7 kip) for Girders 6
capacity of the concrete (V c ) is calculated by using Eq. (7) and 7, respectively. The STM was accurate to within 85% of the
pffiffiffiffi average measured shear capacity for Girder 6 and 98% of the
V c ¼ 0:0316β f 0 cbv d v ð7Þ average measured shear capacity for Girder 7.
In all cases, the calculated values using the sectional analyses in
where β = factor indicating ability of diagonally cracked concrete the AASHTO specifications were very conservative. The AASHTO
to transmit tension and shear. procedures are both on the basis of bending theory, implying that
Both the AASHTO methods are on the basis of bending theory the bending stresses are distributed linearly through the depth of the
with the assumption that plane sections remain plane. The D-region beam, which is typically the case as long as the section being
is defined as any section of a beam within a distance d (depth of the considered is not near the support or an applied concentrated load.
beam) of a concentrated applied load or a support. For this research, The girders that were tested for this research had loads that were
the shear capacity was examined near the D-region where stresses applied near the D-region (a=d ¼ 1:5), where plane sections do not
are not typically distributed linearly through the cross section of the typically remain plane. The STMs developed for this research were
girder. able to much more accurately predict the ultimate shear capacity of
To investigate the predictive shear capacity of the prestressed the girders. This accuracy is attributed to the strut-and-tie method-
concrete beams used in this research, a simple STM was developed ology more effectively predicting the behavior of members that are
following the AASHTO and ACI recommendations. STMs are loaded within the D-region. Applying this methodology may
rarely used in the design of new girders, but this methodology remove restricted load capacities of bridges where shear is the
can prove very useful in design and in the analysis of prestressed controlling failure mechanism.
concrete girders. In addition to comparing the experimentally determined capac-
The STM is an idealized model of a girder’s load paths consist- ity with the code-predicted values, the experimental values were
ing of struts that are compression members made of concrete par- also compared with a detailed finite-element model by using
allel to the expected cracks, ties, or stirrups, which are tension ANSYS. The concrete girder, shear reinforcement, and prestressing
members made of steel analogous to the reinforcement, and nodes strands were modeled by using solid, shell, and link elements,
made of concrete, which are connecting members. respectively. The properties assigned to the concrete and shear
For this research a STM was developed by using two struts and reinforcement were on the basis of experimentally determined
one tie connected at three nodes. This configuration formed a sim- values. The magnitude of residual prestress force for the model
ple triangular truss, which was analyzed to obtain ultimate shear was on the basis of the results of the cracking tests (Table 2). Each

Fig. 6. Strut-and-tie model for Girder 7

JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012 / 307

J. Bridge Eng., 2012, 17(2): 302-309


Table 2. Comparison of Effective Prestress Forces 7 [1,163.2 kN (261.5 kip)]. The detailed finite-element models
Difference were used to perform a parametric study to investigate the sources
Percent
Pe from measured of this difference. The model for Girder 6 was adjusted to evaluate
of
Method kN kip kN kip measured various combinations of concrete compressive strength and shear-
steel reinforcement ratios. When the concrete compressive strength
AASHTO approximate 838.1 188.4 110.8 24.9 115% for this model was increased from 48.9 MPa (7,100 psi) to
AASHTO refined 799.6 179.8 72.3 16.3 110% 64.1 MPa (9,300 psi), the shear capacity increased to 1,104.5 kN
Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

Bridge plan specifications 782.9 176.0 55.6 12.5 108% (248.3 kip). This represented an increase of 52%. Subsequently, the
Cracking test (average) 727.3 163.5 0.0 0.0 100% shear-steel spacing for the Girder 6 model was decreased from
58.4 cm (23.0 in.) to 43.2 cm (17.0 in.), and the ultimate shear
capacity was additionally increased to 1,174.4 kN (264.2 kip). This
Table 3. Measured and Estimated Shear Capacities for Girder 6 new capacity represented an increase of approximately 10% in
(48.9 MPa) and Girder 7 (64.1 MPa) comparison with the original capacity. These results show that
Girder Girder the difference in concrete compressive strength between the two
Percentage Percentage girders had a much larger influence on shear capacity in compari-
6A shear 6B shear
of of son with the differences in stirrup spacing.
Method kN kip measured kN kip measured
AASHTO 366 82.3 55% 366 82.3 47%
simplified Summary and Conclusion
AASHTO 274.4 61.7 41% 274.4 61.7 35%
nominal
Seven decommissioned bridge girders were tested to investigate
the effective prestress force and shear capacity of 42-year-old,
Strut-and-tie 616.3 138.6 92% 616.3 138.6 78%
prestressed concrete, bridge girders. This set of girders was com-
Measured value 668.3 150.3 786.7 176.9
posed of two span lengths and reinforcing designs. A cracking test
was performed on Girders 1 through 5 to determine the residual
finite-element model was loaded incrementally up through failure. prestressing force in the prestressing strands. The experimentally
The calculated deflections at each load step were compared with determined prestress losses were then compared with estimated val-
the measured values and are shown in Fig. 7. ues that were calculated on the basis of procedures recommended in
Fig. 7 shows a very close comparison between experimental and the AASHTO LRFD specifications. Ultimate shear tests were also
analytical values. In all cases, the load-deflection plots showed performed on Girders 6 and 7 to quantify the shear capacity of the
an initially stiff region that was reduced once cracking occurred. girders with the applied loading near the D-region (a=d ¼ 1:5). The
The shear capacity from the analytical models was selected as measured shear capacities were compared with estimated values
the largest applied load and is shown on the figure as the last plotted that were calculated by using sectional procedures recommended
value. For both girders, the maximum difference in ultimate capac- in the AASHTO LRFD specifications. In addition, the measured
ity between experimental and analytical was less than 5%. shear capacities were compared with estimated values by using
The experimental results showed a large difference in shear strut-and-tie methodologies. The following conclusions were ob-
capacities between Girder 6 [727.7 kN (163.6 kip)] and Girder tained on the basis of the research results:

Fig. 7. Comparison of load versus deflection data for Girders 6 and 7

308 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012

J. Bridge Eng., 2012, 17(2): 302-309


• The approximate method in the AASHTO LRFD specifications “Instrument to evaluate remaining prestress in damaged prestressed
predicted 15% larger residual prestress forces in comparison concrete bridge girders.” PCI J., 43(2), 62–71.
with the average measured values. The detailed method resulted Elzanaty, A. H., Nilson, A. H., and Slate, F. O. (1986). “Shear capacity of
in residual forces that were 10% larger than the measured re- reinforced concrete beams using high-strength concrete.” J. Am. Concr.
sults. The current prestress loss equations were developed on Inst., 83(2), 290–296.
the basis of high-strength concrete, which would likely produce Halsey, T. J., and Miller, R. (1996). “Destructive testing of two forty-year
old prestressed concrete bridge beams.” PCI J., 41(5), 84–93.
smaller losses and could be a reason for the discrepancy.
Kaufman, K., and Ramirez, J. A. (1988). “Re-evaluation of the ultimate
• The simplified and nominal shear design procedures in the
Downloaded from ascelibrary.org by The State University of New York at Buffalo (SUNY - Buffalo) on 06/27/24. Copyright ASCE. For personal use only; all rights reserved.

shear behavior of high-strength concrete prestressed I-beams.” ACI


AASHTO LRFD specifications were very conservative and Struct. J.85(3), 295–303.
predicted between 28 and 55% of the measured shear capacity Kordina, K., Hegger, J., and Tuetsch, M. (1989). “Shear strength of
of the prestressed AASHTO Type II girders. These methodol- prestressed concrete beams with unbonded tendons.” ACI Struct. J.,
ogies are on the basis of bending theory and, since the load was 86(2), 143–149.
applied near the support (a=d ¼ 1:5), likely resulted in the very MacGregor, J. G., Sozen, M. A., and Siess, C. P. (1965). “Strength of pre-
conservative estimates. stressed concrete beams with web reinforcement.” J. Am. Concr. Inst.,
• Strut-and-tie models predicted shear capacities to within 2 and 1503–1519.
22% of the respective measured shear values. While less Naito, C., Sause, R., Hodgson, I., Pessiki, S., and Macioce, T. (2010).
frequently used, the strut-and-tie methodology is more applic- “Forensic examination of a noncomposite adjacent precast prestressed
able for D-region shear calculations and produced more accurate concrete box beam bridge.” J. Bridge Eng., 15(4), 408–418.
results, for this loading, in comparison with the AASHTO Oh, B. H., and Kim, K. S. (2004). “Shear behavior of full-scale
simplified and nominal procedures. post-tensioned prestressed concrete bridge girders.” ACI Struct. J.,
• Detailed finite-element models were constructed for the two 101(2), 176–182.
girder types. The parametric study showed that the differences Onyemelukwe, O. U., Issa, M., and Mills, C. J. (2003). “Field measured
prestress concrete losses versus design code estimates.” Exp. Mech.,
in concrete compressive strength (from 48.9 MPa to 64.1 MPa)
43(2), 201–215.
had a larger effect on shear capacity in comparison with the
Osborn, G. P. (2010). “Ultimate shear capacity and residual prestress force
reduction in stirrup spacing (from 58.4 cm to 43.2 cm). of full scale forty one year old prestressed concrete girders.” M.S. thesis,
Utah State Univ., 162.
Pessiki, S., Kaczinski, M., and Wescott, H. H. (1996). “Evaluation of
References
effective prestress force in 28-year-old prestressed concrete bridge
AASHTO. (2009). AASHTO LRFD bridge design specifications, 4th Ed., beams.” PCI J., 41(6), 78–89.
Interim, Washington, DC. Saqan, E. I., and Frosch, R. J. (2009). “Influence of flexural reinforcement
American Concrete Institute (ACI) Committee 318. (2008). Building code on shear strength of prestressed concrete beams.” ACI Struct. J., 106(1),
requirements for structural concrete (ACI 318-08) and commentary, 60–68.
American Concrete Institute, Farmington Hills, MI. Tadros, M. K., Al-Omaishi, N., Seguirant, S. J., and Gallt, J. G. (2003).
Azizinamini, A., Keeler, B. J., Rohde, J., and Mehrabi, A. B. (1996). “Prestress losses in pretensioned high-strength concrete bridge girders.”
“Application of a new nondestructive evaluation technique to a NCHRP Rep. 496, National Cooperative Highway Research Program,
25-year-old prestressed concrete girder.” PCI J., 41(3), 82–95. Transportation Research Board, National Research Council.
Barr, P. J., Kukay, B. M., and Halling, M. W. (2008). “Comparison of pre- Trautner, C., McGinnis, M., and Pessiki, S. (2010). “Analytical and numeri-
stress losses for a prestress concrete bridge made with high-performance cal development of the incremental core drilling method of non-
concrete.” J. Bridge Eng., 468–475. destructive determination of in-situ stresses in concrete structures.”
Civjan, S. A., Jirsa, J. O., Carrasquillo, R. L., and Fowler, D. W. (1998). J. Strain Anal. Eng. Des., 45(8), 647–658.

JOURNAL OF BRIDGE ENGINEERING © ASCE / MARCH/APRIL 2012 / 309

J. Bridge Eng., 2012, 17(2): 302-309

You might also like