Sei sulla pagina 1di 11

Materials and Structures (2015) 48:30373047

DOI 10.1617/s11527-014-0377-5

ORIGINAL ARTICLE

Mechanical properties of prestressing steel after fire


exposure
Zhong Tao

Received: 26 January 2014 / Accepted: 8 July 2014 / Published online: 15 July 2014
RILEM 2014

Abstract Prestressed concrete structures have been


widely used all over the world, and there is a growing
need to study the postfire repairability that is relevant
whenever a fire occurs and no collapse happens after
cooling. To evaluate the damage to the structure after
fire exposure, the residual mechanical properties of
structural materials need to be evaluated first. A
literature review is conducted to analyse the major
factors influencing the post-fire properties of prestressing steel. Existing test data are collected from an
extensive survey of the available literature. Based on
statistical analysis, the effects of heat exposure on the
modulus of elasticity, yield strength and tensile
strength, as well as ultimate strain, are analysed. A
simplified stressstrain model is developed for prestressing steel in residual conditions (i.e. after heating
and cooling to room temperature). Measured stress
strain curves are used to verify the accuracy of the
proposed model.
Keywords Prestressing steel  Prestressed concrete 
Stressstrain curves  Post-fire  Temperature effects 
Damage evaluation

Z. Tao (&)
Institute for Infrastructure Engineering, University
of Western Sydney, Penrith, NSW 2751, Australia
e-mail: z.tao@uws.edu.au

1 Introduction
In prestressed concrete, the steel is tensioned and held
against the concrete, thus putting the concrete in
compression [1]. The prestressing technology helps to
overcome concretes natural weakness in tension, and
ensures high-strength concrete and high-strength prestressing steel working together in prestressed concrete.
In general, prestressing steel is more sensitive to
temperature than ordinary hot rolled steel. Therefore,
larger values of concrete cover are usually prescribed
for prestressing steel [2]. But this does not necessarily
guarantee the fire safety of prestressed concrete
structures since fire-induced spalling might take place
at the concrete cover. Six fires in real unbonded posttensioned concrete building structures were described
by Gales et al. [2]. It was found that some degree of
concrete spalling occurred in all cases, and tendon
rupture or release of prestress occurred in two thirds of
the case studies. The findings highlight the importance
of evaluating the residual safety of prestressed
concrete structures after fire-exposure before repair
works are undertaken. As reported also in [2],
unbonded post-tensioned slabs in a fire-damaged
building were reinstated by re-connecting and retensioning the ruptured tendons. To do this, the decay
of the mechanical properties of the prestressing steel
needs to be determined first.
In the past, many tests were carried out to evaluate
the residual mechanical properties of prestressing steel
after fire exposure. The earliest tests were carried out

3038

in France in 1940s on cold-drawn steel wires and


reported by Guyon [3]. However, due to the many
influencing factors, such as heating temperature,
exposure time, steel type, heating rate, cooling
method, and initial loading, it is not possible to
accurately investigate all influencing factors in a
particular condition. In particular, there were errors
due to uncertainties in the measurements themselves.
Preferably, all published test data should be collected
and used to make a comprehensive and reliable
evaluation. Furthermore, post-fire stressstrain models of materials are required to conduct an accurate
structural analysis for fire-damaged prestressed concrete members and structures. No such model for
prestressing steel, however, is available in any current
design codes, such as the Australian standard AS 3600
[4] and Eurocode 2 [5].
A literature review is presented in the next section
to describe the influence of several different factors on
the residual mechanical properties of prestressing
steel. Utilising the collected test data, a simplified
stressstrain model is developed; finally, the accuracy
of the proposed model is verified against the available
experimental stressstrain diagrams.

2 Literature review
2.1 Prestressing tendons
Three basic types of high-strength prestressing tendons, namely cold-drawn wires, strands and highstrength bars are used in concrete structures. Wires are
produced by drawing hot-rolled steel rods through
dies, and the drawing process takes place when the
steel is cold, thereby altering its mechanical properties
and increasing its strength [6]. Strands are produced by
spinning several individual wires around a central core
wire. Modern strands usually consist of seven wires
with overall diameters ranging from 8 to 18 mm [7].
High-strength steel bars are obtained by introducing
alloying elements in the manufacture of the steel and
by cold working (stretching) the bars [6].
Before they are used in structures, as-drawn wires
or strands are often heated for a short time, or heated
while subjected to high tension [7]. The former
process is called stress-relieving and the latter
stabilising. The corresponding products are called
stress-relieved and low-relaxation wires or

Materials and Structures (2015) 48:30373047

strands, respectively. Both stress-relieving and stabilising processes increase the elastic range of the wires
and strands with respect to the as-drawn condition.
Nowadays, strands are the most commonly used type
of prestressing steel. In North America, low-relaxation
strands have become the standard product [8].
Prestressing steel is manufactured from highcarbon steel with pearlitic (eutectoid composition) or
near-pearlitic microstructure, which is a two-phased,
lamellar structure consisting of alternating layers of
alpha-ferrite (pure iron) and cementite (Fe3C) [9]. The
carbon content by weight normally varies from 0.7 to
0.85 % [7]. The heavy drawing process severely
elongates individual grains in the direction of the
longitudinal axis of the steel. The high values of
strength of the steel are obtained thanks to the decrease
of interlamellar spacing during the drawing process,
which causes the dislocations at the ferrite/cementite
interphase [9] to block.
2.2 Influence of heating and cooling
Exposure to a certain temperature level may lead to a
decay of the mechanical properties of steel [10]. The
influence of heat exposure is no exception to prestressing steel. For normal hot-rolled reinforcing steel,
no obvious influence of heating can be observed below
500 C. For prestressing steel, on the contrary, the
influence becomes apparent after heating to temperatures higher than 300 C. Similar phenomena have
also been observed for cold-worked or heat-treated
reinforcing steel [10].
To compare the effects of high temperature on
prestressing steel and hot-rolled reinforcing steel,
typical stress (r)strain (e) curves of them are depicted
in Fig. 1, in which the test curves reported by Felicetti
et al. [11] are plotted. The yield strengths (fy) for the
prestressing and reinforcing steels at room temperature are 1,730 and 300 MPa, respectively. For comparison purposes, the stressstrain curves in Fig. 1 are
normalised with respect to the corresponding yield
strength at room temperature. The results clearly
indicate that the decrease in fy is only 15 % for the
reinforcing steel heated up to 700 C and cooled to
room temperature, whereas the decrease is 64.1 % for
the prestressing steel heated to the same temperature.
This comparison exemplifies the significant sensitivity
of prestressing steel to heating when compared to
reinforcing steel.

Materials and Structures (2015) 48:30373047

Fig. 1 Typical stress-strain curves for prestressing steel after


heating to 700 C and cooling to room temperature (data from
Felicetti et al. [11] )

Metallographic analysis has been carried out by Day


et al. [12], Abrams and Cruz [13], Abrams and Erlin
[14], and MacLean [15] to investigate the possible
influence of fire exposure on the microstructure of
prestressing steel. When heated to temperatures below
450 C, no obvious change in microstructure was found.
Therefore, the reduction in strength can be attributed to
the reduction of the dislocation density when the heating
temperature is between 300 and 450 C [11]. After
treatments at 450720 C, the cementite lamellae begin
to coalesce into globules that increase in size upon
prolonged heating or exposure to higher temperatures.
Concurrently, ferrite becomes more discernible and
begins to reform into more equi-axial grains [14]. If the
temperature exceeds 720 C, the microstructure reverts
to pearlite with equi-axial grains [15]. At about 800 C,
decarburisation begins at the surface of the steel and
progressively works inward upon prolonged heating or
exposure to higher temperature [14]. This metallographic analysis indicates that the strength reduction at a
temperature above 450 C results from the combined
effects of changes in microstructure and grain size, as
well as from the recovery of dislocations.
2.3 Influence of cooling methods
The cooling rate affects the microstructure of steel
[16]. To investigate the influence of the cooling rate on
the post-fire behaviour of prestressing steel, three
cooling methods, namely cooling in air (CIA), cooling
in furnace (CIF), and cooling with water jet (CWJ),
were adopted by some researchers [13, 1719]. The
cooling rate of CIF is the lowest, followed by that of
CIA, whereas the cooling rate of CWJ is the highest.

3039

Fig. 2 Influence of different cooling methods on the residual


tensile strength of prestressing steel

When heated below 700 C, no evident influence


on the mechanical properties of prestressing steel was
found for the different cooling methods [13, 20]. This
is also supported by other test data collected from the
literature. On the contrary, considerable variation in
test results are found for specimens heated above
700 C. After heating at temperatures above 700 C,
the steel generally has a slightly higher strength when
the CIA method is used, as shown in Fig. 2, where fp
and fpT are the ultimate tensile strength of the unheated
steel and residual tensile strength of the steel heated to
temperature T. The lowest values are obtained when
the steel is kept in the furnace to cool down. It should
be noted that in one case micro-cracking was reported
by Neves et al. [18] after the specimen was heated to
900 C and cooled by water jet. There was virtually no
residual strength left for this specimen as shown in
Fig. 2. Neves et al. [18] attributed this to the formation
of martensitic structure due to the rapid cooling,
leading to increased brittleness. The micro-cracking,
however, was not observed by other researchers, like
Shen et al. [17] and Fan [19], who did similar tests at a
temperature of 900 C and used the CWJ cooling
method. Further tests should be carried out in the
future to clarify this aspect.
2.4 Influence of heating rate and hold time period
of heating
In civil structures, prestressing steel is expected to be
insulated by a concrete cover. Its rate of heating in fire
depends on the fire intensity, location of the prestressing

3040

steel, concrete embedment, and many other factors. In


previous tests, the heating rate selected varied from 3 to
20 C/min. Statistical analysis indicates that the
mechanical properties of prestressing steel are not
obviously affected by the heating rate. This is consistent
with the behaviour observed for normal reinforcing
steel [10].
The hold time period (ts) of heating may affect the
performance of prestressing steel after fire exposure.
This influence was checked by Guyon [3], Abrams and
Cruz [13], Abrams and Erlin [14], MacLean [15], and
Moore [21]. To guarantee a uniform distribution of the
temperature, the specimen needs to be kept in the
furnace at the reference temperature level for a
minimum time, which depends mainly on the diameter
of the specimen. In previous tests, the minimum hold
time ts was 5 min, which was adopted by MacLean
[15] for steel wires with a diameter of 4.4 mm. On the
other hand, the maximum value of ts was 8 h, in the
case of 9.5 mm diameter steel strands tested by
Abrams and Erlin [14]. The test results reported by
different researchers consistently demonstrate that
increasing ts has a detrimental influence on the residual
yield strength and tensile strength of prestressing steel.
This can be attributed to the continuous influence of ts
on the dislocations and microstructure of the prestressing steel. The strength decrease, however, is
normally within 10 %. It is worth noting that strand
specimens tested by Moore [21] were kept for 35 and
60 min at temperatures of 538, 649 and 704 C,
respectively. Compared with the test results of specimens with ts = 35 min, a significant additional
strength decrease was found when ts was increased
to 60 min. For example, at a temperature of 649 C,
fpT decreased from 1,291 to 733 MPa when ts was
increased from 35 to 60 min. It is worth noting that the
diameter of the strands tested by Moore [21] was
12.7 mm, which is a relatively large value among
those of all tested specimens. The 12.7 mm diameter
strands might have not reached temperature equilibrium during the rest time of 35 min. For this reason, in
the following analysis, the test results presented by
Moore [21] for strands with a rest time of 35 min were
not taken into account.
2.5 Influence of steel type
No obvious difference was found between steel wires
and steel strands in terms of influence of fire exposure.

Materials and Structures (2015) 48:30373047

This is also true for different types of prestressing


steel, including as-drawn, stress-relieved and lowrelaxation steel. For as-drawn wires or strands,
however, an increase in 0.2 % proof strength (fp0.2)
up to 29 % was observed after heating to temperatures
of 200 or 300 C [12, 20]. Unheated as-drawn
prestressing steel exhibits significant nonlinearity. It
actually goes through a stress-relieving process when
heated to temperatures around 200300 C, and the
elastic range of the wires or strands can be increased in
comparison to the original as-drawn condition. This
explains the increase in fp0.2 for the as-drawn prestressing steel after the particular heat treatment.
Therefore, a generalised post-fire stressstrain relationship is suitable for different types of prestressing
steel without special consideration for as-drawn wires
and strands.
2.6 Influence of preloading
Most post-fire tests conducted on prestressing steel
were carried out starting from zero stress level. To
simulate the actual fire conditions, preloading was
applied to some specimens in the tests carried out by
Guyon [3], Day et al. [12], Abrams and Erlin [14],
Crook [20], MacLean [15], and Atienza and Elices
[22]. In accordance with the actual level of preloading
in engineering practice, the prestress levels selected in
the tests normally ranged from 40 to 70 % of the
tensile strength of the prestressing steel. Since the
strength of prestressing steel decreases very fast upon
fire exposure, a high constant preload can only be
maintained at a temperature below 300 C. Otherwise,
rupture of the steel will occur. For this reason, instead
of applying constant stress, relaxation tests were
usually conducted to consider the influence of preloading for prestressing steel heated at temperatures
above 300 C. In a relaxation test, the preload was
applied and then the deformation was maintained at a
constant level during heating. Such tests were carried
out by Abrams and Erlin [14], MacLean [15], and
Atienza and Elices [22].
No obvious detrimental effect of preloading on
prestressing steel was observed by all but one of the
researchers. MacLean [15] reported that the application of preloading at temperatures in excess of 400 C
had a mild detrimental effect on the residual tensile
and yield strengths of prestressing steel. When the
exposure temperatures were 400, 500 and 700 C, the

Materials and Structures (2015) 48:30373047

3041

and in particular the ultimate strain (eu) corresponding


to fpT. For the collected re curves, the majority of
them were reported to a strain level of 35 %, which is
significant for most structural applications.
It should be noted that all test data summarised in
Table 1 were obtained in testing steel wires and
strands in residual conditions. Although different
types of prestressing tendons demonstrate similar
behaviour, caution should be taken to extend the
current research results to prestressing bars.

With preload

1.2

Without preload

fpT/fp

0.8
0.6
0.4
0.2
0

200

400

600

800

1000

3.2 Stressstrain curve expression

Temperature T ( C)

Fig. 3 Influence of preloading on the residual tensile strength


of prestressing steel

reductions in tensile strength due to the presence of the


preload were 2.2, 9.8 and 6.8 %, respectively. If
compared with the combined test results of specimens
without preload, however, comparable residual
strengths were obtained for the specimens with
preload from different researchers, as shown in
Fig. 3. Future tests, especially at temperatures above
700 C, are needed before more definitive conclusions
can be drawn. In the following post-fire damage
evaluation, the influence of preload is ignored and all
test results with or without preload are used to work
out the model.

3 Post-fire stressstrain model


3.1 Test data
The aim of this research is to develop a simplified re
model for prestressing steel after heating and cooling
to room temperature. For this purpose, an extensive
survey of the available literature was carried out. A
total of 246 test data and 37 re curves were collected
from 16 studies [3, 1115, 1726]. The details of the
test data are summarised in Table 1. The tensile
strength at room temperature (fp) ranges from 1,248 to
2,089 MPa, whereas the maximum temperature
T ranges from 100 to 900 C. For the collected test
data, most researchers reported the residual tensile
strength and yield strength, but less information is
available for the residual modulus of elasticity (EpT)

Figure 4 shows a typical re curve reported by


Felicetti et al. [11] for prestressing steel at room
temperature, which has also been used in Fig. 1
before. The high-strength steel does not show a welldefined yield point. The initial response to the strain
increment is linear elastic (from point O to point A),
followed by non-linear work-hardening of the material
(from point A to point C). After reaching the tensile
strength at point C, the curve enters into necking and
failure stage.
There is not a well-accepted re model for
prestressing steel. A power formula was proposed by
Skogman et al. [27] for predicting the re relationship
of any type of prestressing steel. Its general form is
expressed as:
3
2
6
6
r eE6Q 
4

7
1Q
7
 fp

 R 1=R 7
5

eE
Kfp

where E, Q, K, and R are curve fitting constants.


Although this model has been used by some researchers, no general equations are available for determining
the constants E, Q, K, and R. This makes this model
less suitable for a wider application. In contrast, the
Australian Standard AS 3600 [4] suggests re curves
of prestressing steel to be determined from appropriate
test data.
A simple re model as expressed by Eq. (2) is
recommended in the PCI handbook [8] for 250 ksi
(1,724 MPa) strand and 270 ksi (1,862 MPa) strand.
(
r

196; 500e for e  0:0076


0:276
fp 
for e [ 0:0076
eA

3042

Materials and Structures (2015) 48:30373047

Table 1 Summary of test data for post-fire prestressing steel


Source

Number of
specimens

Number
of curves

Steel
type

Treatment

Diameter
(mm)

fp (MPa)

T (C)

Cooling
method

Guyon [3]

12

Wire

As-drawn

2.5, 5.1

1,3642,089

200600

CIF

Day et al. [12]

Abrams & Cruz [13]


Abrams & Erlin [14]
Crook [20]

Shen et al. [17]

Neves et al. [18]


Fan [19]

Falkner & Gerritzen [23]


Deng [24]
MacLean [15]
Zheng et al. [25]

Wire

As-drawn

5.1

1,448

300

Wire

5.1

1,489

300

CIF

Wire

Stress-relieved

5.1

1,448

300

CIF

Strand

Stress-relieved

9.5

1,248

450750

CIF

Strand

Stress-relieved

9.5

1,248

450750

CWJ

16

Strand

1,875

400800

CIF

Wire

As-drawn

1,797

100700

CIA

Wire

Stress-relieved

1,679

100700

CIA

Strand

Stress-relieved

9.3

1,940

100700

CIA

Wire

1,569

100900

CIF

Wire

1,569

400900

CIA

Wire

1,569

400900

CWJ

Wire

5.5

200900

CIA

8
18

Wire
Strand

5.5
15

1,967

200900
100900

CWJ
CIA

Strand

15

1,967

100900

CIF

Strand

15

1,967

100900

CWJ

18

Wire

1,717

100900

CIA

Wire

1,717

100900

CIF

Wire

1,717

100900

CWJ

1,860

300500

Strand

Stress-relieved

15.2

1,975

100900

CIA

14

14

Wire

Low-relaxation

4.4

2,020

200700

CIF

Wire

Low-relaxation

1,803

100700

CIA

22

Strand

Low-relaxation

9.5, 12.7

1,908, 1,960

260704

CIA, CIF

Atienza & Elices [22]

Wire

Stress-relieved

1,950

100600

CIA

Felicetti et al. [11]


Galvez et al. [26]

Strand

12.7

1,935

200850

CIF

Wire

Low-relaxation

1,856

200600

CIF

Moore [21]

Fig. 4 Typical re curve for prestressing steel at room


temperature [11]

where the unit for r and fp is MPa; fp is taken as


1,724 MPa for 250 ksi strand and 1,862 MPa for
270 ksi strand; and A is taken as 0.0064 for 250 ksi
strand and 0.007 for 270 ksi strand.
Figure 5 shows the comparison between the prediction using Eq. (2) and the test curve reported by
Felicetti et al. [11]. Reasonable prediction is obtained
when a nominal fp of 1,862 MPa is used in the
calculation rather than the actual fp of 1,935 MPa.
Although it is possible to use 1,935 MPa as fp in the
calculation to get a good match between the predicted
and test curves at point C, the strength will be greatly
overestimated for most of the strain-hardening stage
of the curve. PCI handbook does not make

Materials and Structures (2015) 48:30373047

3043

C
B
A

O fp0.1/Ep

EC2 model

Test curve [11]

PCI model

Fig. 5 Simplified re models for prestressing steel at room


temperature

recommendations regarding re curves for other


strands and wires.
A simple idealised bi-linear re relation is suggested in Eurocode 2 [5], as shown in Fig. 5. This
model, which is also recommended in the fib Model
Code 2010 [28], can be expressed as:
8
for e  fp0:1 =Ep
< Ep e
eu  efp  fp0:1
r
for fp0:1 =Ep \e  eu
: fp 
eu  fp0:1 =Ep
3
For simplicity, the above bi-linear model will be
referred to as EC2 model hereafter. In this model, the r
e relation can be determined by four parameters, i.e.,
fp0.1, fp, Ep and eu, where fp0.1 and Ep are the 0.1 % proof
stress and modulus of elasticity at room temperature,
respectively. Using experimental values for these
parameters, the prediction from EC2 model is compared
with the test curve shown in Fig. 5, which indicates a
very good agreement. For wires, Ep normally varies
from 195 to 210 GPa, whereas Ep can vary from 185 to
205 GPa for strands. If accurate values are not known,
Ep may be assumed equal to 205 GPa for wires and
195 GPa for strands [5]. eu normally ranges from 0.05 to
0.07 [8] and a minimum value of 0.02 is specified in the
fib Model Code 2010 [28]. If eu is not known, it may be
taken as 0.022 [5]. There is no direct relationship
between fp0.1 and fp [5]. fp0.1 may be taken as 0.85fp for
stress-relieved wire and strand, and 0.9fp for lowrelaxation wire and strand [5, 6].
Fire-damaged prestressing steel exhibits less strain
hardening during tensile deformation compared with
unheated prestressing steel. This is evident in Fig. 1.

where fpT and fp0.1T are the residual tensile strength and
0.1 % proof strength, respectively, for the steel heated
to a temperature T; EpT is the residual modulus of
elasticity; and euT is the ultimate strain corresponding
to fpT.
3.3 Determining fpT
The ratios of fpT/fp for the collected data are shown as a
function of T in Fig. 6. According to the literature
review and previous discussion in subsection 2.3, the
test data for specimens heated above 700 C are
discarded if the specimens were cooled in air or by
water jet. This is done for two reasons: first, if heated
over 700 C, specimens cooled in furnace have
statistically lower strength compared with specimens
cooled in other methods; and second, since prestressing steel is embedded in concrete, it is expected that
prestressing steel is more likely to cool down slowly as
in a furnace due to the heat sink effect of the concrete.

1.2
Hot-rolled reinforcing steel [10]

1
0.8

fpT/fp

fp
fp0.1

Therefore, it is decided that the bi-linear EC2 re


relation can be used for post-fire prestressing steel, and
possible revisions to the material parameters are
required to consider the influence of fire damage.
The revised re model is expressed as follows for
post-fire prestressing steel:
8
for efp0:1T =EpT
< EpT e
euT efpT fp0:1T
r
for fp0:1T =EpT \eeuT
: fpT 
euT fp0:1T =EpT

Cold-worked/heat-treated
reinforcing steel [10]

0.6
0.4
R2=0.966

0.2
0

fpT / fp
0

200

0.27
400

600

800

1000

Temperature T ( C)
Fig. 6 Ratio of fpT/fp as a function of temperature

3044

Materials and Structures (2015) 48:30373047

fpT =fp 0:27

Formulae to predict the residual tensile strength


were proposed by Tao et al. [10] for post-fire hotrolled and cold-worked/heat-treated reinforcing steel.
Using those formulae, the predictions for reinforcing
steel are shown in Fig. 6 as well. Compared with
reinforcing steel, much significant loss of tensile
strength is expected for prestressing steel after heating
to 400 C and above.
3.4 Determining fp0.1T
Various arbitrary methods have been proposed for
defining the yield point of high-tensile steel, such as
the 0.1 % proof stress, 0.2 % proof stress or 1.0 %
strain. In the US, the yield strength is defined to be
taken at 1 % elongation fp1 [8], whereas 0.2 % proof
stress fp0.2 is taken as yield strength in the Chinese
code GB 50010 [29]. In contrast, 0.1 % proof stress
fp0.1 is defined as yield strength by the Australian
Standard AS 3600 [4], EC2 [5] and the fib Model Code
2010 [28]. In general, fp0.1 is slightly smaller than fp0.2,
and the difference between them is about 24 %
according to the Australian/New Zealand Standard
AS/NZS 4672 [30].
For the test data collected, only 0.2 % proof stresses
were reported in the majority of the references, whilst
0.1 % proof stresses were presented by Day et al. [12]
and stresses at 1 % elongation given by Moore [21].
Although there are some discrepancies between the
yield strengths determined by using different definitions, it can be assumed that the percentages of
strength loss are the same regardless of different yield
strength definitions. In the following, fp0.1T/fp0.1 is
replaced by fp0.2T/fp0.2 or fp1T/fp1 if the ratios of fp0.1T/
fp0.1 were not reported. fp0.2T and fp1T are the residual

Hot-rolled reinforcing steel [10]

1
0.8
0.6
Cold-worked/heat-treated
reinforcing steel [10]

0.4
0.2
0

1
1:37 T=510

1.2

fp0.1T/fp0.1

A nonlinear regression analysis is performed, and


Eq. (5) is proposed to predict fpT. The coefficient of
determination R2 is 0.966, which indicates very high
correlation between the ratio of (fpT/fp) and T. According to Eq. (5), no strength loss occurs when T is below
200 C, and a sharp decrease in tensile strength takes
place when T is between 400 and 700 C. After that, a
relatively stable residual tensile strength is achieved. At
a heating temperature of 700 C, the predicted loss of
tensile strength is 63.5 %.

200

400

600

800

1000

Temperature T ( C)

Fig. 7 Ratio of fp0.1T/fp0.1 as a function of temperature

0.2 % proof stress and stress at 1 % elongation,


respectively, for prestressing steel heated to a temperature T.
Figure 7 shows the experimental values of (fp0.1T/
fp0.1). Once again, the test data for specimens heated
over 700 C are discarded if the specimens were
cooled in air or by water jet. The trend of the loss in
yield strength against temperature is similar to that of
tensile strength, which can be seen from Figs. 6 and 7.
Therefore, a formula similar to Eq. (5) is proposed to
predict fp0.1T:
fp0:1T =fp0:1 0:15

1
1:176 T=5605

Similarly, formulae have been presented by Tao


et al. [10] to predict the residual yield strength of postfire hot-rolled and cold-worked/heat-treated reinforcing steel. Compared with reinforcing steel, much
significant loss of yield strength is also expected for
prestressing steel after heating to 400 C and above, as
shown in Fig. 7.
3.5 Determining EpT
The experimental values of (EpT/Ep) are plotted in
Fig. 8. The variation increases once the exposed
temperature T is over 500 C. When T is below
700 C, the heat exposure has little effect on the
residual modulus of elasticity EpT. However, a reduction in EpT was reported by Fan [19] when T is over
700 C. Eq. (7) is proposed to predict EpT.

Materials and Structures (2015) 48:30373047


EpT

3045

Ep
0:254 2:5  103 T  2:05  106 T 2 Ep

3.6 Determining euT


Test data of euT are extremely scarce, which were only
reported by Galvez et al. [26]. Meanwhile, euT can be
retrieved from the full-range re curves reported by
Felicetti et al. [11]. When T varied from 200 to 600 C,
Galvez et al. [26] observe little influence of T on euT as
shown in Fig. 9. On the contrary, Felicetti et al. [11]
reported a decline in euT with increasing T up to
600 C. At a temperature of 700 C, a sharp increase
in euT was observed by Felicetti et al. [11], which was
about 1.6 times eu. After that, euT decreased again.
Values of elongation after fracture were reported by
Crook [20], Neves et al. [18], Fan [19] and MacLean
[15]. Due to the influence of necking, elongation after
fracture is larger than the ultimate strain. According to
the test data reported by Galvez et al. [26], elongation
is 30-60 % higher than euT. If euT is assumed to vary
proportionally to the elongation after fracture, the
latter may also be used to evaluate the influence of
heating on euT, as shown in Fig. 9. Test results
presented by Crook [20], Neves et al. [18] and
MacLean [15] indicate no obvious adverse effect of
heating on elongation. But Fan [19] reported a
declining trend in elongation when T increased from
100 to 600 C. The controversial observations on euT
highlight further research work in this area.

T  700  C
700 C\T  900  C

The minimum euT reported by Felicetti et al. [11]


was 0.048, whereas the minimum elongation of 0.045
was reported by Crook [20]. It can be inferred that the
existing test data of euT are all well above the minimum
eu of 0.02 as specified in the fib Model Code 2010 [28].
Before a solid conclusion can be reached, the ratio of
(euT/eu) may be taken as unity for the temperature
range 20600 C. Beyond this temperature range, it is
on the safe side for ductility if euT/eu is taken as unity.

3.7 Comparison between predicted and test curves


The collected test curves reported by Falkner and
Gerritzen [23], MacLean [15], Zheng et al. [25],
Moore [21], and Felicetti et al. [11], are further used to
verify the proposed re model for fire-damaged
prestressing steel. In the calculations, four parameters
including fp0.1, fp, Ep and eu need to be used. The value
of Ep was not reported by Felicetti et al. [11] and a
default value of 195 GPa for strands is used. fp0.1 can
be retrieved from the re curves presented, and it is
found that fp0.1 is quite close to fp0.2 or fp1. The
difference is negligible in terms of prediction accuracy
of re curves if fp0.2 or fp1 is used to replace fp0.1. Since
full-range re curves were not presented by Falkner
and Gerritzen [23], MacLean [15], Zheng et al. [25],

EpT/Ep=1

1.2

Eq. (7)

uT/ u

EpT/Ep

Crook [20]
Neves et al. [18]
Fan [19]
MacLean [15]
Felicetti et al. [11]
Glvez et al. [26]

2.5

0.8
0.6
Day et al. [12]
Fan [19]
Zheng et al. [25]
Moore [21]

0.4
0.2
0

Crook [20]
Deng [24]
MacLean [15]

1.5
1
0.5
0

200

400

600

Temperature T (C)

Fig. 8 Influence of temperature on EpT/Ep

800

1000

200

400

600

800

1000

Temperature T (C)

Fig. 9 Influence of temperature on euT/eu

3046

(a)

(b)

2400
2000

Stress (MPa)

Fig. 10 Comparison
between stress-strain
relation model and test
curves

Materials and Structures (2015) 48:30373047

1600

Test [21]

1200

Predicted

800

T=260C, Ep=178,600 MPa, fp=1,960 MPa,

400

fp1=1,760 MPa, u =0.08, 9.5 mm strand

0
0

(c)

(d)

(e)

(f)

and Moore [21], values of eu are taken as the maximum


strains of curves reported for steel at room temperature. In general, reasonably good agreement is
achieved between the predictions and measured re
curves, and the comparison is shown in Fig. 10.

4 Conclusions
Prestressing steel is very sensitive to temperature, and
heat exposure can bring in a significant decrease of the
mechanical properties. A bi-linear stressstrain model
has been developed for prestressing steel after exposure to elevated temperatures and cooling to room
temperature. In general, the predicted stressstrain
curves show good agreement with the available test
results. The models are valid for wires and strands with

0.03

0.06
Strain

0.09

0.12

a tensile strength ranging from 1,200 to 2,000 MPa,


and a heating temperature not higher than 900 C.
Further tests are still required to investigate the
influence of heat exposure on the ultimate strain and
modulus of elasticity for prestressing steel. In particular, more full-range stressstrain curves should be
reported. These test data can be used to further verify
the proposed model and improve the prediction
accuracy if required.
Acknowledgments The author would like to thank Mr Henry
Huynh for his assistance in collecting the test data.

References
1. Lin TY, Burns NH (1981) Design of prestressed concrete
structures. Wiley, New York

Materials and Structures (2015) 48:30373047


2. Gales J, Bisby LA, Gillie M (2011) Unbonded post tensioned concrete in fire: a review of data from furnace tests
and real fires. Fire Saf J 46(4):151163
3. Guyon Y (1953) Prestressed concrete. Wiley, New York
4. Standard Australia (2009) Concrete structures. Standards
Australia, AS 3600-2009, Sydney
5. British Standards Institution (2004) Design of concrete
structuresPart 1-1: general rules and rules for buildings.
British Standards Institution, Eurocode 2, BS EN 1992-11:2004, London
6. Gilbert RI, Mickleborough NC (1990) Design of prestressed
concrete. Unwin Hyman Ltd., London
7. Hurst MK (1998) Prestressed concrete design. E & FN
SPON, London
8. Precast/Prestressed Concrete Institute (Precast/PCI) (2010)
PCI design handbook, Precast and prestressed concrete, 7th
edn. Precast/Prestressed Concrete Institute, Chicago
9. Toribio J (2004) Relationship between microstructure and
strength in eutectoid steels. Mater Sci Eng, A 378389:227230
10. Tao Z, Wang XQ, Uy B (2013) Stress-strain curves of
structural and reinforcing steels after exposure to elevated
temperatures. J Mater Civ Eng 25(9):13061316
11. Felicetti R, Gambarova PG, Meda A (2009) Residual
behavior of steel rebars and R/C sections after a Fire. Constr
Build Mater 23(12):35463555
12. Day MF, Jenkinson EA, Smith AI (1960) Effect of elevated
temperatures on high-tensile-steel wires for prestressed
concrete. ICE Proc 16(1):5570
13. Abrams MS, Cruz CR (1961) Behavior of high temperatures
of steel strand for prestressed concrete. J PCA Res Dev Lab
3(3):819
14. Abrams MS, Erlin B (1967) Estimating post-fire strength
and exposure temperature of prestressing steel by a metallographic method. J PCA Res Dev Lab 9(3):2333
15. MacLean KJN (2007) Post-fire assessment of unbonded
post-tensioned concrete slabs: strand deterioration and
prestress loss. Masters thesis, Department of Civil Engineering, Queens University, Ontario
16. Avner SH (1974) Introduction to physical metallurgy.
McGraw-Hill, New York
17. Shen R, Rong K, Feng LY (1991) Mechanical properties of
reinforcing bars after exposure to elevated temperatures or
fire. Build Sci Res Sichuan 17(2):59 [in Chinese]

3047
18. Neves IC, Rodrigues JPC, Loureiro AP (1996) Mechanical
properties of reinforcing and prestressing steels after heating. J Mater Civ Eng 8(4):189194
19. Fan J (2001) Fire resistance of unbonded prestressed concrete structures. PhD thesis, School of Civil Engineering,
Southeast University, Nanjing [in Chinese]
20. Crook RN (1980) The Elevated Temperature properties of
reinforced concrete. PhD thesis, University of Aston,
Birmingham
21. Moore WL (2008) Performance of fire-damaged prestressed
concrete bridges. Masters thesis, Faculty of the Graduate
School, Missouri University of Science and Technology,
Rolla, Missouri
22. Atienza JM, Elices M (2009) Behavior of prestressing steels
after a simulated fire: fire-induced damages. Constr Build
Mater 23(8):29322940
23. Falkner H, Gerritzen D (2004) Prestressed slabs without
bond. Structural and serviceability behaviour after fire
exposure. Beton- und Stahlbetonbau 99(8):652656 [in
German]
24. Deng CN (2005) Studies of mechanics characters on main
beam of prestressed concrete undergone high temperature.
Masters thesis, College of Resources and Civil Engineering, Northeastern University, Shenyang [in Chinese]
25. Zheng WZ, Hu Q, Zhang HY (2007) Experimental research
on the mechanical property of prestressing steel wire during
and after heating. Front Archit Civ Eng China 1(2):247254
26. Galvez F, Atienza JM, Elices M (2011) Behaviour of steel
prestressing wires under extreme conditions of strain rate
and temperature. Struct Concr 12(4):255261
27. Skogman BC, Tadros MK, Grasmick R (1988) Flexural
strength of prestressed concrete members. PCI J
33(5):96123
28. Federation internationale du beton (fib) (2012) Model code
2010-final draft, vol 1. FIB Bulletin No 65, Lausanne
29. Ministry of Housing and Urban-Rural Development of the
Peoples Republic of China (2011) Code for design of
concrete structures. GB50010-2010, China Architecture &
Building Press, Beijing [in Chinese]
30. Standard Australia/Standards New Zealand (2009) Steel
prestressing materials, part 1: general requirements. Standard Australia, AS/NZS 4672.1:2007, Sydney

Potrebbero piacerti anche