Sie sind auf Seite 1von 20

Prestress losses

and camber growth


in wing-shaped
structural members

Marco Breccolotti and Annibale L. Materazzi

An experimental program on pretensioned wing-shaped members has investigated the evolution of prestressing stresses and
camber during the different stages of concrete hardening.
The result allowed the careful evaluation of prestress losses in
the first three weeks during concrete placement and curing,
emphasizing the influence of temperature history on prestress
losses and calibrating parameters for the prediction of camber
growth.
Calculations based on Eurocode2 equations compared
favorably with the data, indicating that these equations can
be applied to prestressed wing-shaped members made of selfconsolidating concrete as investigated in the present work.

98

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

restressed concrete wing-shaped members are frequently used in the construction of long-span roofs
in precast concrete buildings. Roofs spanning over
30m (100ft) are built by assembling a number of these
thin-walled members, which may be placed side by side or
with a certain transverse spacing, often rather wide, covered by ribbed slabs with curved shape. Figure1 shows
an example of a roof comprising curved ribbed slabs
supported by adjacent wing-shaped members resting on a
main I-girder. The ease of fabrication of the wing-shaped
members is comparable to that of a double tee but more
difficult than an I-girder. Nevertheless, the use of wingshaped members allows for better aesthetic results and for
roofs having smaller mass with the consequent benefits in
relation to seismic risk. To optimize the design of these
thin-walled members and take advantage of their loadcarrying capacity, prestressing is extensively used.
The exploitation of the mechanical properties of the materials, the prestress force necessary to ensure an appropriate
load-carrying capacity, and the requirement for in-service
camber for aesthetic and functional reasons necessitate
monitoring of prestress and camber. Inaccurate prediction
of the camber can lead to unsatisfactory service conditions.
Checking and monitoring of prestress loss and camber
growth and investigations of the parameters that could affect them have primarily focused on bridge girders.

Figure 1. Example of roof with wing-shaped members.

Stallings et al.1 investigated the accuracy of methods for


calculating the camber at the time of erection of standard AASHTO girders fabricated with high-performance
concrete. According to their findings, current analytical
techniques, such as the time-step method, can correctly
predict the camber provided that the material properties
are accurately known. Al-Omaishi et al.2 proposed an
extension to the provisions of current standards, which
were based on data obtained from concrete with strengths
ranging from 4 to 6ksi (28 to 41MPa), for estimating
prestress losses in concretes with strengths up to 15ksi
(100MPa). Tadros et al.3 emphasized the random nature
of camber that depends on several stochastic quantities, such as concrete modulus of elasticity, differential
temperature at prestress release, effective prestressing
force, debonding length, and transfer length. Storm et
al.4 focused on the effects of production practice, which
involves the curing method and other aspects, on the
camber of prestressed bridge girders, developing a refined
prediction method based on the 2010 American Association of State Highway and Transportation Officials
AASHTO LRFD Bridge Design Specifications5 that compared well with the measured camber of 382prestressed
concrete bridge girders.

Special attention has also been paid to the effect of curing


and temperature evolution on prestressed concrete bridge
girders. Roller et al.6 investigated the effects of curing
temperature on high-strength concrete in five AASHTO
TypeIII bridge girders with bonded pretensioned strands.
Their findings highlighted an 11% reduction in the average
prestressing force due to thermal expansion of the steel
during cement hydration. Barr et al.7 monitored a precast,
prestressed concrete bridge girder during fabrication and
the first three years of service. They found that high curing
temperatures during fabrication can reduce the calculated
prestress by up to 7%, reduce the initial camber by up
to 40%, and increase the in-service bottom tensile stress
by up to 27%. The structural behavior of high-strength
concrete bridge girders was also monitored by Roller et
al.,8 who compared the measured and calculated prestress
losses. They found that the measured prestress losses derived from concrete strains, corrected for the temperature
and load effects, were lower than the corresponding values
calculated using the AASHTO LRFD specifications.9
Prestress losses, camber growth, and other parameters
were monitored by Dwairi et al.10 in several prestressed
high-performance concrete bridge girders. Among other
interesting data, a temperature increase T of almost 72F
PCI Journal | J a n u a r y Fe b r u a r y 2015

99

Figure 2. Axonometric view of prestressed concrete wing-shaped member.

(40C) was recorded by the researchers during cement


hydration. Barr and Angomas11 highlighted the effect on
camber of the temperature gradient during the curing of
prestressed concrete girders and presented a procedure to
calculate it.
To carefully evaluate the effective prestress losses during
and after concrete casting and hardening, to monitor the
camber growth, and to evaluate the effect of temperature
variation during concrete hardening in wing-shaped members, an experimental program on full-scale pretensioned
specimens has recently been completed. The structural behavior of the prestressed concrete elements was investigated at various stages of their early life. Attention was also
placed on the use of self-consolidating concrete (SCC).
In fact, while most of the properties of hardened SCC are
comparable with those of conventional concrete, the modulus of elasticity at release Ecj, which plays an important
role in the evaluation of prestress losses, has been found
to be less than that of conventional concrete mixtures with
comparable compressive strength of concrete at release fcj.12
Electrical-resistance surface-bonded strain gauges, load
Table 1. Self-consolidating concrete mixture proportions
Material

Prestressed concrete wingshaped members


The wing-shaped members studied in this research project
have a nominal width of 2500mm (98in.), a depth of
710mm (28in.), and lengths between 17 and 19m (56 and
62ft). Their V-shaped section permits drainage of rainwater and provides the necessary moment of inertia, while
their end portions facilitate installation and guarantee
overall stability (Fig.2).
Self-consolidating concrete was used in the manufacture
of the wing-shaped members. The specified minimum concrete compressive strength at release fcj was 29 to 33MPa
(4210 to 4790psi), and the specified minimum concrete
compressive strength at 28days fc was 45MPa (6530psi).
Table1 shows the mixture proportions.

Amount

Portland cement (type CEM I 52,5 R), kg/m

460

Fine aggregate, kg/m3

992

Coarse aggregate, kg/m3

747

Water, kg/m3

185

High-range water-reducing admixture, L/m3

0.7

Watercementitious material ratio

0.40

Note: 1L/m3 = 25.8fl oz/yd3; 1kg/m3 = 1.6875lb/yd3.


100

cells, thermocouples, and displacement transducers were


used to measure the actual prestress losses and the growth
of camber in the first three weeks during fabrication and
curing. The results of the tests are presented in this paper,
andwhenever possiblethe experimental measurements
are critically compared with the theoretical predictions
to improve the design provisions and the construction
process.

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

Tests on companion cylinders made during casting were


used to evaluate the effective concrete compressive strength
and modulus of elasticity. Table2 reports the results.
Low-relaxation seven-wire strands with an ultimate tensile strength of prestressing steel fpu equal to 1860MPa
(270ksi) were used. The initial prestress tension pi
varied from 1400 to 1500MPa (203 to 217ksi) to limit
the concrete tensile stresses in the top fiber at prestress
release. For similar reasons several strands were debonded

Table 2. Experimental concrete compressive strength and modulus of elasticity


Concrete property

Sample

Compressive strength, MPa

Modulus of elasticity, MPa

Concrete age
16 hours (release)

1 day

28 day

35.1

50.0

54.7

46.1

49.5

49.3

45.4

50.0

n/a

Average

42.2

49.8

52.0

33,048

n/a

n/a

33,199

n/a

n/a

33,786

n/a

n/a

Average

33,344

n/a

n/a

Note: n/a = not applicable. 1MPa = 0.145ksi.

Abutment

Wing-shaped member

Casting bed

Abutment

Hydraulic jacks

Casting bed

Hydraulic power unit


Figure 3. Pretensioning system.

at various distances from the element ends. In accordance


with Italian standard D.M. 14.01.200813, B450C deformed
reinforcing bars and steel welded-wire mesh B450A were
used as reinforcement.

In particular, the following sensors were used (Fig.5):


twelve electrical-resistance surface-bonded strain


gauges for measuring the deformation in the strands,
appropriately protected from the fresh concrete by a
waterproof coating (Fig.6)

The experiments involved three pretensioned specimens


of different lengths cast in different seasons using a jig
composed of a movable abutment, two casting beds, hydraulic jacks for strand detensioning, and a fixed abutment
(Fig.3). The anchor heads are blocked in the abutments
by means of conical keys. At midspan the steel formwork,
which provides the reaction forces required by the abutments to counterbalance the pretensioning of the strands,
holds a special device to allow for strand detensioning
before prestress transfer. The specimens differ in overall
length, prestressing reinforcement, and initial prestress
(Table3). Figure4 shows the cross sections with the different strand configurations.

two strain gauges embedded in the fresh concrete for


the measurement of the deformation of the concrete
inside the bulb

three electrical-resistance surface-bonded strain


gauges for measuring the concrete deformation on

The members were instrumented with various sensors to


monitor the prestress forces and the growth of the camber
after tensioning of the strands and after prestress transfer.

Experimental setup

Table 3. Length, prestressing strands, and initial prestress of specimens


Element

Length, m

Strands

pi, MPa

T1

16.9

1 0.6in. + 12 0.5in.

1400

T2

19.8

1 0.6in. + 14 0.5in.

1500

T3

18.9

11 0.6in. + 2 0.5in.

1450

Note: pi = initial prestressing tension. 1mm = 0.0394in.;


1m = 3.28ft; 1MPa = 0.145ksi.
PCI Journal | J a n u a r y Fe b r u a r y 2015

101

710 mm

2490 mm

Strands debonding
2.5 m
2.0 m
1.5 m
1.0 m

3 2 1

3 2 1

upp
4

e
fac

r
r su

3 2 1

0.5 in. strands


0.6 in. strands
Concrete bulb

Figure 4. Cross sections and prestressing strands of the wing-shaped members tested in this project from top to bottom: tests1, 2, and 3. Note: 1mm = 0.0394in.;
1m = 3.28ft.

the upper surface (the installation was performed after


removal of the upper formwork)

102

two displacement transducers for measuring the vertical displacement, placed in the mean longitudinal
plane, at 1/4 and 1/2 of the span length

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

load cells for measuring the prestress force in two


strands, placed between each anchoring head and the
abutment holding all the anchoring heads (Fig.7)

three thermocouples, two embedded in the fresh concrete and the third to measure ambient temperature.

Channel 8

Section 4

Section 1

Section 5

Section 3

Section 2
Channel 10
Channel 7
Strand 7
Strand 5
Strand 3
Strand 1

Channel 12
Channel 11

Channel 4
Channel 2

Channel 15
Channel 14

Channel 9

Channel 1

Channel 13

Strand 8
Strand 6
Strand 4
Strand 2

Section 4

Section 1

Section 5

Section 3

Section 2

Strands strain gauges

Channel 19
Channel 18
Channel 17

Strand 8
Strand 6
Strand 4
Strand 2

Strand 7
Strand 5
Strand 3
Strand 1
Channel 22

Channel 21

Channel 34

Section 4

Section 1

Section 5

Section 3

Section 2

Concrete strain gauges

Channel 33
Strand 8
Strand 6
Strand 4
Strand 2

Strand 7
Strand 5
Strand 3
Strand 1

Section 4

Section 1

Section 5

Section 3

Section 2

Displacement transducers

Ambient

Strand 8
Strand 6
Strand 4
Strand 2

Strand 7
Strand 5
Strand 3
Strand 1

0.4 m

0.8 m
L/4

Concrete bulb
=

Concrete bulb
=

L/4

1.5 m
L/2

Thermocouples

Figure 5. Positions of sensors. Note: L = total length of prestredded element. 1 mm = 0.0394 in.; 1 m = 3.28 ft.
PCI Journal | J a n u a r y Fe b r u a r y 2015

103

Figure 6. Electrical-resistance surface-bonded strain gauges glued to prestressing strands after application of waterproof coatings.

Results of monitoring tests


Prestress
The tension in the strands was measured using electricalresistance surface-bonded strain gauges glued to one wire
of the strands and load cells placed between the anchoring heads of two strands and the abutment holding all the
anchoring heads.
Figure8 shows the values of the tensile stress measured on
four strands during the tensioning in test2. In this figure
the loss of prestressing caused by the wedge draw-in and
by the elastic deformation of the steel formwork produced
by the sequence of strand tensioning is plotted as a function of time.
Similar results hold for test3 (Fig.9). In this case the close
correspondence between the measurements made with
electrical-resistance surface-bonded strain gauges and load
cells is remarkable. In the same figure, it is also possible to
observe the reduction of the prestress losses caused by the
nonsimultaneous strand stretching, achievable by means of

104

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

a second tensioning. For example, strand1 gained a prestress of roughly 40MPa (6ksi) from 1325MPa (192ksi)
at the end of the first tensioning phase to about 1365MPa
(198ksi) at the end of the second tensioning phase.
Figure10 shows the history of the tensile stress measured
during the hardening phase up to prestress transfer of test3.
Concrete deformation
Two strain gauges embedded in the concrete close to
strand1 made it possible to monitor the evolution of deformation in the concrete at sections1 and 3 (Fig.11).
During the first phase (about 2hours) a rather limited
shortening occurred, followed by a longer phase of elongation, resulting from shrinkage and thermal effects. It is
possible to observe a different behavior between the strain
gauge placed at midspan (section1) and that placed close
to the anchoring head (section3). In the first case the elongation increased for the first 9hours and then remained
almost constant. The elongation measured by the second
strain gauge started to slowly decrease after 5hours. After-

Figure 7. Anchoring heads and load cells mounted on prestressing strands1 and 6.

ward, the release of the strands produced an abrupt shortening due to elastic deformation of the concrete on transfer
of prestress. As expected, the shortening was greater at
section3 due to the smaller bending moment produced by
the self-weight of the element.
For longer periods, the effect on concrete deformation of the
daily temperature variation was superimposed on the effects
of creep and shrinkage; the combined effect produced a
gradual but continuous increase of the deformation.
After the removal of the upper formwork, it was also possible to install the electrical-resistance surface-bonded strain
gauges on the extrados of the concrete surface. Figure11
shows the deformation measured by the uppermost strain
gauge (channel19 of Fig.5) and confirms that the section
is fully in compression at prestress transfer.
Camber monitoring
Some camber is always desirable at the serviceability limit
state in slender members such as these. Camber can be easily measured in the factory and can also be used to check

the prestress. The growth of the camber was monitored


continuously at midspan (section1) and at L/4 (section5)
during test2 for 21days after release (where L is the
length of the wing-shaped element). Figure12 shows the
measurements. For sections1 and 5, starting from an elastic deformation at release equal to 34 and 20mm (1.33 and
0.79in.), respectively, the camber increased in 21days to
approximately 42 and 27mm (1.65 and 1.06in.). Even in
this case the effect of the nonuniform diurnal temperature
variation can be observed superimposed on the creep and
shrinkage deformations.
Temperature
The variation of the concrete temperature adjacent to
strand1 was detected both at midspan (section1) and at
the end of the prestress transfer zone (section3). Figure13
illustrates the temperature histories, together with the ambient temperature measured during test3. In both sections
cement hydration was responsible for a temperature rise of
about 22C (40F), whichadded to the ambient temperature of 15C (59F)led to a maximum temperature of
37C (99F).

PCI Journal | J a n u a r y Fe b r u a r y 2015

105

Figure 8. Test2: stresses in strands at tensioning (top); tension losses of


strand1 due to wedge slip and subsequent tensioning of strands (bottom).
Note: 1MPa = 0.145ksi.

Furthermore, about 12hours after casting, when the concrete


temperature reached its maximum and the concrete started
cooling, the tensile stress of the strands as measured by the load
cells began to slowly increase (Fig.10). Between the 12th and
18th hour, the concrete temperature decreased by 5C (9F) and
the stress in the strand increased by 2MPa (0.29ksi). The stress
increase would have been much higher without the relaxation in
the pretensioned reinforcement. That is, heat of hydration was
responsible for a non-negligible prestress loss.

Calibration of design
parameters
The adjustment of some nominal parameters assumed in the
design phase, such as the factors that influence the evaluation of the prestress loss in the short- and long term and those
involved in the assessment of camber growth, was conducted,
taking into account the experimental results. These parameters are treated separately in the following sections.
Prestress loss
Immediate losses In the case of prestressing strands

106

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

Figure 9. Test3: stresses in strands after tensioning (top); tension losses in


strand1 due to wedge slip and subsequent tensioning of strands and stress
recover achieved by a second tensioning (bottom). Note: 1MPa = 0.145ksi.

tensioned individually between the abutments interconnected by the casting bed, the elastic deformations of the whole
pretensioning system produced a prestress loss that cannot
be evaluated with sufficient precision. The experimental
measurements provided the data to evaluate this prestress
loss. In fact, the prestress loss of the first tensioned strand
caused by the elastic deformation of the casting bed produced by the tensioning of the remaining strands was measured by load cells and electrical-resistance surface-bonded
strain gauges. Table4 reports the values of these losses.
After the tensioning of a number n of strands, each having nominal area Ai, with initial prestressing tension pi,
Eq.(1) can estimate the final tension of the first strand (assuming no load eccentricity), similar to the prestress loss
produced by elastic shortening in posttensioned beams for
the sequential jacking steps.

where

(1)

Figure 10. Test3: tensile stress in strands as measured in the first 20hours by load cells (top and middle); and magnification of stress-time curves measured in first
hour after tensioning by electrical-resistance surface-bonded strain gauges (bottom). Note: 1MPa = 0.145ksi.
PCI Journal | J a n u a r y Fe b r u a r y 2015

107

The experimental findings allowed the evaluation of the summation in the right-hand-side member of Eq.(1), thus permitting the evaluation of the average prestress loss of Eq.(2).
In tests1, 2, and 3, the average prestress losses were equal
to 29.5, 31.1, and 25.3MPa (4.3, 4.5, and 3.7ksi), respectively, close to 2% of the initial prestress. This loss can be
introduced directly in the calculations or can be significantly reduced by a second tensioning of the strands as
done, for example, in test3, where about 40MPa (5.8ksi)
was recovered in strand1 (Fig.9).
By observing the force variations measured by load cells
and electrical-resistance surface-bonded strain gauges
glued to the strands, it was possible to quantify the prestress loss due to wedge draw-in at the fixed anchorage.
The two measurements were in perfect agreement. Table5
shows the measured values of these losses. In particular,
although new wedges and new cylinders were used during the tests, the following different average losses were
observed for the 0.5 and 0.6in (13and 15mm). diameter
strands fwed,0.5 and fwed,0.6:
fwed,0.5 = 40.3MPa (5.84ksi)
fwed,0.6 = 24.5MPa (3.55ksi)
The prestress losses for the relaxation of the strands were
quantified following EuroCode215 for low-relaxation
strands (Class2):
Figure 11. Test2: short-term (top) and long-term (bottom) concrete deformation as a function of time (hours and days, respectively).

1,fin = final tension of first-tensioned strand after tensioning n strands

where

Afw = cross section of the casting bed

fpR = relaxation prestress loss

Equation(2) calculates the average prestress loss due to


elastic deformation of the casting bed during tensioning of
n strands m,fin.14

1000 = relaxation prestress loss after 1000hours = 2.5%

= ratio between the initial prestressing tension and


the ultimate tensile strength of prestressing steel

(2)

Table 4. Experimental prestress loss due to elastic deformation of casting bed and abutments used for pretensioning strands
Pretension strand

pi, MPa

Strand number

Strand area,
mm2

Ai, mm2

Prestress loss,
MPa

T1

1400

139

12

99

64.4

T2

1500

139

14

99

66.9

T3

1450

139

10

139

56.2

Element

Note: Ai = nominal area of ith strand; n = number of strands; pi = initial prestressing tension. 1mm = 0.0394in.; 1MPa = 0.145ksi.
108

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

According to ACI435R, the prestress losses for the relaxation of the strands can also be evaluated as follows:

where
t = time from initial tensioning
fpy = yielding tensile strength of prestressing steel
These equations refer to the losses that would occur at
20C (68F). To take into account actual test conditions,
where the hydration reaction of the cement produced a
temperature increase in the strands, an equivalent time teq
was calculated on the basis of the following equation15 with
temperature expressed in C:

Figure 12. Test2: camber history. Note: 1mm = 0.0394in.

where
teq = equivalent time for temperatures above 20C (68F)
Tmax = maximum temperature
T(ti) = constant temperature occurred for a duration ti
ti = duration of the ith constant temperature T(ti)
Table6 shows the theoretical values of the prestress losses
due to steel relaxation.
Prestress losses can also take place due to thermal variations in the steel due to heat of hydration. EuroCode2
provides the following simplified formula for steam curing:
Ntemp = 0.5ApEpc(Tmax T0)
where

Figure 13. Test 3: temperature history. Note: C = 5/9(F 32)

hardening (Fig.13). Consequently, the prestress loss due to


the heating of both concrete and prestressing steel turned
out to be considerable. Table7 summarizes these losses
for the second test, conducted in August, and the third test,
conducted in December.
Table 5. Experimental prestress loss due to wedge draw-in
Element

Ntemp = prestress loss due to steam curing

Strand
number

Strand area, mm2

Prestress loss,
MPa

T1

139

29.4

T2

139

21.8

Ap

= total area of prestressing steel

Ep

= elastic modulus of prestressing steel

T2

99

40.3

= coefficient of thermal expansion of concrete

T2

99

37.5

T0

= ambient temperature

T2

99

43.5

T3

139

25.1

T3

139

21.9

Although in this case no steam curing was used, significant temperature variations were recorded during concrete

Note: 1mm = 0.0394in.; 1MPa = 0.145ksi


PCI Journal | J a n u a r y Fe b r u a r y 2015

109

At release, the prestressing force was transmitted to the


hardened concrete, which shortened elastically, depending
on the modulus of elasticity of the concrete at that time.
The prestress losses measured in strand1 were 83, 90, and
113MPa (12.0, 13.0, and 16.4ksi) for the first, second,
and third test, respectively (Table8). The numerically
predicted losses were 91 and 106MPa (13.2 and 15.4ksi)
in tests2 and 3, respectively; hence, agreement with the
experimental results was good. The elastic modulus of the
concrete at release was evaluated by tests on cube specimens. Its design value, derived from the concrete compressive strength by means of a standard correlation formula,15
was equal to 32.8GPa (4757ksi), in good agreement with
the value 33.3GPa (4836ksi) determined experimentally.
The total instantaneous loss can be evaluated by assumingat
least in the short termthe absence of any interdependence
among the individual prestress losses, by the summation of
the losses due to elastic deformation of the self-stressing casting bed, wedge draw-in that depends on strand diameter, steel
relaxation that depends on initial pretensioning, temperature
of the concrete, and elastic concrete shortening.

as = time-dependent coefficient for autogenous shrinkage


Assuming a concrete characteristic compressive strength
fck equal to 45.7MPa (6630psi), a notional size of the concrete section h0 equal to 70mm (2.76in.), kh equal to 1.0,
and t equal to 21days, the prestress loss due to shrinkage
fpSH is 38.0MPa (5510psi).
The same prestress loss could be evaluated according to
ACI435R as:

where

SH = shrinkage strain of concrete


Eps = modulus of elasticity of prestressing steel

The theoretical values of the total instantaneous prestress


loss range from 213.2 to 244.5MPa (30.9 to 35.5ksi)
(Table9). These values appear to be in good agreement
with the experimental values for the specimens tested in
this project (Table10).

V = volume of concrete element

Time-dependent losses The long-term prestress


losses due to strand relaxation, as well as to concrete shrinkage and creep, were analyzed. At the end of test2, after
21days, the prestress loss due to relaxation was 53.6MPa
(7.8ksi). The average temperature from the 2nd day to the
21st day was equal to 27C (81F).

The prestress loss due to concrete creep fpCR at 21days


has been evaluated according to EuroCode2 as follows:

The shrinkage-induced loss at 21days, according to the


provisions of EuroCode2 for normal concrete cs, was
evaluated as the sum of the losses due to drying shrinkage cd and autogenous shrinkage ca because no specific
indications are provided for SCC.

where

S = surface of exposed concrete


RH = relative humidity

Eci = modulus of elasticity of concrete at time of initial


prestress

cs = cd + ca

c = stress in concrete at level of centroid of prestressing


tendon

where

t0 = age of concrete at time of loading

cd(t) = ds(t,ts)khcd,0

= creep coefficient

ds = time-dependent coefficient for drying shrinkage

Assuming a concrete mean compressive strength fcm of


54.6MPa (7920psi), h0 of 70mm (2.76in.), and t0 of
0.75days and taking into account the effect of temperature
history (as required for steel relaxation), the prestress loss
due to creep is fpCR equal to 31.8MPa (4610psi).

ts

= age of concrete (days) at beginning of drying


shrinkage

kh

= coefficient depending on notional size h0

cd,0 = basic drying shrinkage strain

110

ca(t) = as(t)ca()

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

Similarly, this prestress loss could be evaluated according


to the ACI435R-95 standard14 as:

where
KCR = coefficient equal to 2.0 for pretensioned members
Ec = elastic modulus of concrete
fcs = stress in concrete at centroid of reinforcement immediately after transfer

from Fig.11. The experimentally measured strain variation,


from cutting of the strands up to the 21stday, was approximately equal to 230. Taking into account the elastic
modulus of the steel, this figure corresponds to a prestress
variation of 46.9MPa (6.8ksi). Because the theoretical
value of the prestress loss determined from Eq.(3) and
Eq.(4) is 46.5MPa (6.74ksi), there is no practical difference between the prestress loss derived experimentally and
the calculated prestress loss. The total long-term theoretical
losses amount to 106.7MPa (15.5ksi).
Camber growth

fcsd = stress in concrete at centroid of reinforcement due to


all superimposed dead loads applied after prestressing is accomplished

During test2, the elastic camber induced at midspan by


the combined action of the prestressing force and the selfweight just after release and at 21days was found to be 35
and 41mm (1.38 and 1.61in.), respectively (Fig.12).

To take into account the mutual interaction among the


individual prestress losses, Eq.(3) provided by EuroCode2
was used to calculate the total prestress loss:

The theoretical values of the deflection due to the selfweight sw and to the prestressing force with unbonded
tendons pr were also calculated using Eq.(5) and (6).

Ep

p ,c + s + r =

( t , t0 ) c ,Qp + cs E p + 0.8 pr
Ecm
E p Ap
Ac 2
zcp 1 + 0.8 ( t , t0 )
1+
1 +
Ecm Ac
Ic

(3)

(5)

where

where

w = lineal weight of prestressed element

p,c+s+r = absolute value of variation of stress in tendons


due to creep, shrinkage, and relaxation

L = total length of the prestressed element

c,Qp

= stress in concrete at level of prestressing steel


due to quasi-permanent loads

Ecm

= 28-day elastic modulus of concrete

pr
= absolute value of variation of stress in tendons
due to relaxation of prestressing steel

Ac

= area of concrete section

Ic

= moment of inertia of concrete section

zcp

= distance between center of gravity of concrete


section and tendons

I = section moment of inertia




l 2 l xin2 , dt ,i x 2fin , dt ,i (6)
xin2
pr
n
pr =

+
x
A
e

p
,
i
pg
,
i
,
,
fin
dt
i
Ecj I i =1
2l
2
8 2

2
2
2
2

pr
x
l
l xin , dt ,i x fin , dt ,i
n
+ x fin , dt ,i + in , dt ,i
=
Ap,i e pg ,i 8 2

Ecj I i =1
2
l
2


where

pr = prestressing tension
Ap,i = area of ith strand
epg,i = eccentricity of ith strand

It is now simple to calculate the variation of the strain at


the centroid of the prestressing steel using Eq.(4).

xin,dt,i = abscissa of ith strand after initial debonding and


transfer length

xfin,dt,i = abscissa of ith strand after final debonding and


transfer length

(4)

An indication of the soundness of the evaluation of the prestressing losses following prestress transfer can be obtained

The permanent deflection due to creep caused by the selfweight, which was the only sustained load acting on the
wing-shaped element during the experimental tests, has

PCI Journal | J a n u a r y Fe b r u a r y 2015

111

Table 6. Computed prestress loss due to strand relaxation in 18 hours

pi, MPa

Tmax, C

Tmed, C

teq, hours

T2

1500

56

43

T3

1450

37

26

Element

Prestress loss, MPa


EC2

ACI435R

1286

38.9

26.9

59

18.5

13.1

Note: teq = equivalent time; Tmax = maximum temperature; Tmed = mean temperature; pi = initial prestressing tension. 1MPa = 0.145ksi;
C = 5/9(F 32).

no significant external loads have been applied to the member


during the monitoring phase, the value of the elastic modulus
at release Ecj has also been used to predict the camber at later
ages. The effect of concrete hardening on the deformation
produced by prestress loss can be evaluated more accurately
using the refined method proposed by Storm et al.4

Table 7. Computed prestress loss due to thermal variations


Tmax T0, C

Prestress loss, MPa

T2

26

26.5

T3

22

22.4

Test

Note: T0 = ambient temperature; Tmax = maximum temperature.


1MPa = 0.145ksi; C = 5/9(F 32).

The theoretical values obtained by Eq.(5) and (6) overestimated both displacements: 38.8 and 45.6mm (1.53 and
1.79in.), at release and after 12days, or +11% in both
cases. Camber growth was measured on 10elements with
different lengths and prestressing forces (Table11) at
transfer and at 14days.

Table 8. Experimental prestress loss due to concrete elastic deformation


Strand
number

Strand area,
mm2

T1

139

82.7

T2

139

90.0

T3

139

113.0

Element

Prestress
loss, MPa

Calibration of the corrective


terms for camber prediction
The numerical procedure to evaluate the total camber at release and at time t was adjusted to take into account the effective material properties and the effective prestress losses that
depend on the ambient temperature, on the strand tensioning
procedure, and so on, but also on the uncertainties due to variation in relative humidity, temperature distribution, effective
surface area in contact with the environment, and manufacturing tolerances (mainly referring to the actual geometry of the
sections). Starting from the design value, the results of the
monitoring permitted refinement of the coefficients c1 and c2
used in Eq.(7) and (8) to correct the camber due to prestress
and prestress loss at release rel and at time t (t):

Note: 1mm = 0.0394in.; 1MPa = 0.145 ksi.

Table 9. Theoretical total immediate prestress loss


Season

Prestress loss, MPa


Strand 0.5in.

Strand 0.6in.

Summer

244.5

228.7

Winter

229.0

213.2

Note: 1mm = 0.0394in.; 1MPa = 0.145 ksi.

been assumed negligible.

rel = c1pr sw (7)

(t) = c1pr c2pr sw (8)

where
Figure14 shows the debonding and transfer lengths. Because
Table 10. Experimental total immediate prestress loss
Element

Strand number

Initial stress, MPa

Stress at transfer, MPa

, MPa

T2

1500

1260

240

T3

1450

1216

234

T3

1450

1206

244

Note: = Experimental total immediate prestress loss. 1MPa = 0.145 ksi.


112

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

Ap,i

i th strand

x in,deb,i x in,tr,i
x in,dt,i

x fin,tr,i

x fin,deb,i

x fin,dt,i

Figure 14. Debonding length and transfer length. Note: Ap,i = area of ith strand; pg,i = eccentricity of ith strand; xfin,deb,i = abscissa of ith strand after final debonding;
xfin,dt,i = abscissa of ith strand after final debonding and transfer length; xfin,tr,i = final transfer length of ith strand; xin,deb,i = abscissa of ith strand after initial debonding;
xin,dt,i = abscissa of ith strand after initial debonding and transfer length; xin,tr,i = initial transfer length of ith strand.

pr = deformation due to prestress loss at time t, evaluated


similarly to pr (Eq.[6])
No corrective term has been used for the deflection due
to self-weight because the theoretical estimate can be
considered reliable. The optimal values of the coefficients c1 and c2 are 0.80 and 0.83, respectively. The
use of these coefficients is, nevertheless, valid only
for the specific wing-shaped elements analyzed in this
study. Figure15 plots the ratios between the experimental camber and the camber predicted by Eq.(7)
and Eq.(8). Good agreement has been found for the
camber at 14 days, with a ratio between the experimental and the theoretical values between 0.95 and 1.10.
Slightly worse results have been found for the camber
at release.

Conclusion
In this research project, an experimental investigation and
a number of theoretical analyses were conducted to gain
firsthand information on the actual prestress losses and
camber growth in the first weeks following casting of prestressed, pretensioned wing-shaped concrete members.
The results of the analysis provided the following information:

Prestress losses, roughly equal to 2% of the initial


prestressing, occurred for tensioning the strands in
successive steps and for the elastic deformation of
the casting bed and of the steel trestle holding the anchored heads of the strands. Appropriate countermeasures, such as a second tensioning of the individual

Table 11. Length, prestressing strands, initial prestressing, concrete strength at release, camber at release, and camber at 14days
Length, m

Strands

pi, MPa

fcj, MPa

rel, mm

14, mm

M1

17.8

1 0.6in. + 12 0.5in.

1500

45.1

22.1

35.1

M2

17.8

1 0.6in. + 12 0.5in.

1500

45.1

22.1

32.0

M3

17.8

1 0.6in. + 12 0.5in.

1500

46.4

25.9

32.0

M4

17.8

1 0.6in. + 12 0.5in.

1500

46.4

25.9

32.0

M5

17.8

1 0.6in. + 12 0.5in.

1500

45.6

26.9

32.0

M6

17.8

1 0.6in. + 12 0.5in.

1500

45.6

29.0

34.0

M7

17.8

1 0.6in. + 12 0.5in.

1500

50.7

22.1

35.1

M8

17.8

1 0.6in. + 12 0.5in.

1500

50.7

22.9

33.0

M9

17.8

1 0.6in. + 12 0.5in.

1500

47.7

22.1

32.0

M10

11.9

1 0.6in. + 12 0.5in.

1500

46.8

17.0

21.1

Element

Note: fcj = compressive strength of concrete at release; rel = camber due to prestress and prestress loss at release; 14 = camber at 14days;
pi = initial prestressing tension. 1mm = 0.0394in.; 1m = 3.28ft; 1MPa = 0.145ksi.
PCI Journal | J a n u a r y Fe b r u a r y 2015

113

2009. Estimating Prestress Loss in Pretensioned,


High-Strength Concrete Members. PCI Journal 54
(4): 132159.
3. Tadros, M. K., F. Fawzy, and K. E. Hanna. 2011.
Precast, Prestressed Girder Camber Variability. PCI
Journal 56 (1): 135154.
4. Storm, T. K., S. H. Rizkalla, and P. Z. Zia. 2013.
Effect of Production Practices on Camber of Prestressed Concrete Bridge Girders. PCI Journal 58
(4): 96111.

Figure 15. Ratio between measured and calculated camber.

strands, have been suggested to reduce this prestress


loss.

Different prestress losses, equal to 40.3 and 24.5MPa


(5.84 and 3.55ksi), have been measured for the 0.5
and 0.6in. strand for the wedge draw-in at the anchored extremities.
The temperature history has a non-negligible influence on the development of prestress losses for steel
relaxation and concrete creep.
The predictive formulas about prestress losses provided by EuroCode2 can provide accurate indications
also in the case of special structural members, such as
the prestressed, pretensioned wing-shaped members
made of self-consolidating concrete investigated in the
present work.

Appropriate corrective actions are proposed as well to reduce as much as possible the prestress loss during concrete
casting, curing, and hardening, and suitable values for the
design parameters have been calibrated to correctly predict
prestress losses and camber growth.

Acknowledgments
The Authors are grateful to Manini Prefabbricati S.p.A.
(S. Maria degli Angeli, Perugia, Italy) and in particular
to Paolo Manni and Graziano Baldograni for the valuable
support in organizing the experimental phase of this work.

References
1. Stallings, J. M., R. W. Barnes, and S. Eskildsen.
2003. Camber and Prestress Losses in Alabama HPC
Bridge Girders. PCI Journal 48 (5): 216.
2. Al-Omaishi, N., M. K. Tadros, and S. J. Seguirant.

114

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

5. AASHTO (American Association of State Highway


and Transportation Officials). 2010. AASHTO LRFD
Bridge Design Specifications. 5th ed. Washington DC:
AASHTO.
6. Roller, J. J., H. G. Russell, R. N. Bruce, and B. Hassett. 2003. Effects of Curing Temperatures on High
Strength Concrete Bridge Girders. PCI Journal 48
(5): 7279.
7. Barr, P. J., J. F. Stanton, and M. O. Eberhard. 2005.
Effects of Temperature Variations on Precast, Prestressed Concrete Bridge Girders. Journal of Bridge
Engineering 10 (2): 186194.
8. Roller, J. J., H. G. Russell, R. N. Bruce, and W. R.
Alaywan. 2011. Evaluation of Prestress Losses in
High-Strength Concrete Bulb-Tee Girders for the
Rigolets Pass Bridge. PCI Journal 56 (1): 110134.
9. AASHTO. 2008. AASHTO LRFD Bridge Design
Specifications, 4th Edition2008 Interim Revisions.
Washington, DC: AASHTO.
10. Dwairi, H., M. Wagner, M. Kowalsky, and P. Zia.
2010. Behavior of Instrumented Prestressed High
Performance Concrete Bridge Girders. Construction
and Building Materials 24 (11): 22942311.
11. Barr, P. J., and F. Angomas. 2010. Differences
between Calculated and Measured Long-Term Deflections in a Prestressed Concrete Girder Bridge.
Journal of Performance of Constructed Facilities 24
(6): 603609.
12. Schindler, A., R. Barnes, J. Roberts, and S. Rodrguez.
2007. Properties of Self-Consolidating Concrete for
Prestressed Members. ACI Materials Journal 104 (1):
5361.
13. Ministero delle Infrastrutture. 2008. Decreto Ministeriale 14 Gennaio 2008Nuove Norme Tecniche per le
Costruzioni. GU n.29, February 4, 2008, Rome, Italy:
Ministero delle Infrastrutture.

14. ACI (American Concrete Institute) Committee 435.


1995. Control of Deflection in Concrete Structures.
ACI 435R-95. (Reapproved 2000.) Farmington Hills,
MI: ACI.

fpu

= ultimate tensile strength of prestressing steel

fpy

= yielding tensile strength of prestressing steel

h0

= notional size of concrete section

= section moment of inertia

Ic

= moment of inertia of concrete section

Notation

k h

= coefficient depending on notional size h0

Ac

= area of concrete section

KCR

= coefficient equal to 2.0 for pretensioned members

Afw

= cross section of casting bed

ldeb

= debonding length

A i

= nominal area of ith strand

= total length of prestressed element

Ap

= total area of prestressing steel

= number of strands

Ap,i

= area of ith strand

RH

= relative humidity

c1

= coefficient for correcting camber due to initial


prestress

= surface of exposed concrete

= time from initial tensioning

c2

= coefficient for correcting deformaiton due to


prestress loss

t0

= age of concrete at loading

epg,i

= eccentricity of ith strand

teq

= equivalent time for temperatures above 20C


(68F)

Ec

= elastic modulus of concrete

Eci

= modulus of elasticity of concrete at time of


initial prestress

ts

= age of concrete (days) at beginning of drying


shrinkage

= elastic modulus of concrete at release

T0

= ambient temperature

Ecj

= 28-day elastic modulus of concrete

Tmax

= maximum temperature

Ecm

= elastic modulus of prestressing steel

Tmed

= mean temperature

Ep
Eps

= modulus of elasticity of prestressing steel

T(ti)

= constant temperature occurred for duration ti

= volume of concrete element

fc

= 28-day compressive strength of concrete


w

= lineal weight of prestressed element

fck

= concrete characteristic compressive strength

fcj

= compressive strength of concrete at release

fcm

= concrete mean compressive strength

fcs

= stress in concrete at centroid of reinforcement


immediately after transfer

fcsd

= stress in concrete at centroid of reinforcement


due to all superimposed dead loads applied after
prestressing is accomplished

15. CEN (European Committee for Standardization).


2010. Eurocode 2: Design of Concrete Structures
Part 11: General RulesGeneral Rules and Rules for
Buildings. EN 199211. Brussels, Belgium: CEN.

xfin,deb,i = abscissa of ith strand after final debonding;


xfin,dt,i

= abscissa of ith strand after final debonding and


transfer length

xfin,tr,i

= final transfer length of ith strand

xin,deb,i = abscissa of ith strand after initial debonding


xin,dt,i

= abscissa of ith strand after initial debonding and


transfer length

PCI Journal | J a n u a r y Fe b r u a r y 2015

115

xin,tr,i

=initial transfer length of ith strand

cd

= drying shrinkage

zcp

= distance between center of gravity of concrete


section and tendons

cd,0

= basic drying shrinkage strain

= coefficient of thermal expansion of concrete

cs

= shrinkage-induced loss at 21 days for normal


concrete

as

= time-dependent coefficient for autogenous


shrinkage

SH

= shrinkage strain of concrete

ds


= time-dependent coefficient for drying shrinkage

= ratio between initial prestressing tension and


ultimate tensile strength of prestressing steel

14

= camber at age 14days

pr

= theoretical values of deflection due to prestressing force with unbonded tendons

1000

= relaxation prestress loss after 1000hours

rel

= camber due to prestress and prestress loss at


release

1,fin

= final tension of first-tensioned strand after tensioning n strands

sw

= stress in concrete at centroid of prestress

= theoretical value of deflection due to self-weight

pr

= deformation due to prestress loss at time t

c,Qp

= stress in concrete at level of prestressing steel


due to quasi-permanent loads

fpCR

= prestress loss due to creep

pi

= initial prestressing tension

fpR

= prestress loss due to relaxation

pr

= prestressing tension

fpSH

= prestress loss due to shrinkage

= creep coefficient

fwed,0.5 = average prestress loss for 0.5in. diameter


strands
fwed,0.6 = average prestress loss for 0.6in. diameter
strands
Ntemp = prestress loss due to steam curing
ti

= duration of ith constant temperature T(ti)

= temperature increase

p,c+s+r = variation of strain at centroid of prestressing


tendons due to creep, shrinkage, and relaxation

= experimental total immediate prestress loss

m,fin = average prestress loss due to elastic deformation


of casting bed during tensioning of n strands
p,c+s+r = absolute value of variation of stress in tendons
due to creep, shrinkage, and relaxation

116

pr

= absolute value of variation of stress in tendons


due to relaxation of prestressing steel

ca

= autogenous shrinkage

Jan u a r y F e b r u a r y 2 0 1 5 | PCI Journal

About the authors


Marco Breccolotti, PhD, is
assistant professor in structural
engineering in the Department of
Civil and Environmental Engineering of the University of
Perugia in Italy. His research
interests include the safety
assessment of existing reinforced concrete structures
by means of nondestructive testing, the evaluation of
the fire resistance of reinforced concrete structures, the
identification of damage in reinforced concrete
structures through dynamic tests, and prestressing in
reinforced concrete elements.
Annibale Luigi Materazzi, PhD, is
professor and chair of structural
analysis and design in the
Department of Civil and Environmental Engineering of the
University of Perugia. His
research interests include durability and in-service behavior of reinforced concrete
structures, fire engineering, and vibration-induced
damage on materials and structures.

Abstract
Prestressed concrete wing-shaped members are frequently used in the construction of long-span roofs in
precast concrete buildings. Roofs spanning more than
30m (100ft) are built by assembling a number of these
thin-walled members. To optimize their design and take
advantage of their potential load-carrying capacity, the
transfer of large forces from the pretensioned strands to
the concrete is critical. Information on the evolution of
these forces and of the ensuing deformations is necessary, not only in the long run, but also immediately after
concrete casting and during the first few weeks.

The Department of Civil and Environmental Engineering of the University of Perugia in Italy and a major
prefabrication firm of central Italy have completed an
experimental program on pretensioned wing-shaped
members to investigate the evolution of prestressing
stresses and of camber during the different stages of
concrete hardening.
The result allowed the careful evaluation of the prestress losses that occur in the first three weeks during
concrete placement and curing to emphasize the influence of the temperature history on the development of
prestress losses and to calibrate suitable parameters for
the prediction of camber growth.
Numerical comparisons based on Eurocode2 are also
performed to check the reliability of the numerical
procedure adopted by the authors and to ascertain
whether using self-consolidating concrete (as in the
manufacture of the specimens tested in this project)
makes any sizeable difference with respect to the predictions based on Eurocode2, which refers to ordinary
concrete.

Keywords
Camber, prestress loss, self-consolidating concrete,
temperature, wing-shaped members.

Review policy
This paper was reviewed in accordance with the
Precast/Prestressed Concrete Institutes peer-review
process.

Reader comments
Please address and reader comments to journal@pci
.org or Precast/Prestressed Concrete Institute, c/o PCI
Journal, 200 W. Adams St., Suite 2100, Chicago, IL
60606. J

PCI Journal | J a n u a r y Fe b r u a r y 2015

117

Das könnte Ihnen auch gefallen