You are on page 1of 13

Fire Safety Journal 60 (2013) 1–13

Contents lists available at SciVerse ScienceDirect

Fire Safety Journal


journal homepage: www.elsevier.com/locate/firesaf

Structural responses of reinforced concrete columns subjected


to uniaxial bending and restraint at elevated temperatures
Kang-Hai Tan, Truong-Thang Nguyen n
School of Civil and Environmental Engineering, Nanyang Technological University 50 Nanyang Avenue, Singapore 639798, Singapore

art ic l e i nf o a b s t r a c t

Article history: A total of six specimens were tested to failure to investigate the effects of uniaxial bending, axial
Received 29 May 2012 restraint, and initial load level on the structural responses of reinforced concrete columns at elevated
Received in revised form temperatures. The full-scale column specimens (nominal height of 3.3 m and square cross-section of
24 March 2013
300 mm) were cast with concrete of 55 MPa compressive strength and reinforcing steel of 550 MPa yield
Accepted 15 April 2013
strength, and were tested with different levels of initial load and uniaxial eccentricity. Temperature-
dependent axial deformations, lateral deflections, thermal-induced restraint forces, failure modes, and
Keywords: failure times of the test specimens are compared with those obtained from numerical analyses using
Reinforced concrete SAFIR program. It is experimentally shown that the lateral deflection at elevated temperatures is
Columns
adversely affected by uniaxial eccentricity and initial load level. Besides, the development of thermal-
Eccentricity
induced restraint forces increases with eccentricity but decreases with initial load level, and can be
Uniaxial bending
Restraint overpredicted by the numerical analysis that ignores concrete spalling and implicitly accounts for
Elevated temperatures concrete transient strain at elevated temperatures.
& 2013 Elsevier Ltd. All rights reserved.

1. Introduction used for verification of a model that was proposed to predict the
critical temperature of axially-restrained steel columns by extend-
Actual columns are rarely under pure compression due to ing the traditional Rankine formula [5]. In a fair number of fire
irregularity of structural layout, frame discontinuity, live load tests conducted on axially-restrained RC columns [6,7], the effect
patterns and construction imperfections. In reality, many columns of eccentric load was not considered. Besides, only small-scale
are designed to resist a combination of an axial load and a major specimens were conducted in these tests. Behaviour of full-scale
bending moment about a principal axis of their cross-section, so- restrained columns at elevated temperatures was investigated
called uniaxial bending. When a building is subjected to a localised with different levels of initial applied load and axial restraint
fire, a heated column is axially restrained since its thermal [8–12]. Again, no eccentric load had yet been used in these
elongation is constrained by surrounding structural members that research works. As a result, current codes of practice for fire
are at lower temperatures [1–3]. It is envisaged that uniaxial design of concrete structures [13,14] only provide tabulated data
bending and axial restraint are the two governing factors for the and simplified design methods for isolated columns under pure
fire resistance analysis of column elements. compression or uniaxial bending, without any consideration of
To date, there have been a considerable number of research restraint effect. Hence, there is an essential need to conduct such
works conducted on various types of restrained columns subjected experiments to enhance the understanding of structural responses
to fire. However, to the authors’ best knowledge, there are limited of RC columns under the combined effects of eccentric loads and
fire tests investigating the combined effects of eccentric load and axial restraint at elevated temperatures.
axial restraint on full-scale columns. Lie et al. [1] and Wang et al. This paper presents a programme of fire tests on six full-scale
[2] investigated reinforced concrete (RC) columns fully restrained axially-restrained RC columns subjected to various levels of uni-
against thermal expansion with the outcomes only applicable axial bending and initial applied load. Thermal and structural
for columns under pure compression. A small-scale experimental responses, as well as failure modes and failure times of the test
model was introduced by Rodrigues et al. [4] to study hinged steel specimens are discussed and compared with finite element (FE)
bars with restrained thermal elongation under uniaxial bending. predictions using computer software SAFIR [15].
However, without any testing of steel profiles at a realistic scale
the results of this experiment may not be so applicable to practice.
Besides, only specimens with a small eccentricity of 1 mm were
2. Initial applied load in uniaxial bending case

n
Corresponding author. Tel.: +65 6790 4577; fax: +65 6791 0676. Categorizing fire as an accidental action, most of current fire
E-mail addresses: NTThang@ntu.edu.sg, thangcee@gmail.com (T.-T. Nguyen). design codes [13,14] stipulate that when heating occurs there are

0379-7112/$ - see front matter & 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.firesaf.2013.04.005
2 K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13

only service loads acting on structures. This so-called initial 3. Experimental programme
applied load when a fire occurs can be determined by multiplying
the column design axial resistance at ambient condition with a 3.1. Column specimens
reduction factor μfi, which can also be referred to as the initial load
level [13]. However, there is a remarkable difference in axial In this experimental study, six column specimens of a test
resistance between the cases of pure compression and uniaxial series, namely, S1, were designed and tested with different levels
bending. of uniaxial eccentricity and initial load level (Table 1). The last two
Fig. 1 depicts an M–N interaction diagram at ambient condition numerals in a specimen nomenclature represent the first-order
of an RC column and the initial applied load when heating occurs. eccentricity (in mm) exerting onto the specimen during test.
If the column is stocky and is subjected to pure compression, it can Specimens C1-1-00 and C1-2-00 were loaded with zero eccen-
resist a maximum axial load of Nud, which is the design axial tricity whereas the remaining four specimens, namely, C1-3-25,
resistance of its cross-section. However, when this column is C1-4-40, C1-5-60, and C1-6-80, were subjected to uniaxial eccen-
subjected to an increasing axial load with a first-order eccentricity tricities of 25, 40, 60, and 80 mm, respectively. As listed in Table 1,
e along one principal axis of its cross-section, so-called uniaxial the initial loads N 0 used in the tests were based on the
eccentricity, it can only attain a design resistance represented by eccentricity-dependent axial resistances NR dEC2 of the columns
point P2 (NRd, MRd) which is the intersection between the column that were determined in accordance with EC2 Pt.1.1 [16]. The
M–N interaction diagram and the non-linear load–moment curve reduction factors μfi shown in Table 1 were within an allowable
that depends on e. In this case of uniaxial bending, the column range of 0.2–0.7 [13]. Apart from a value of 0.45 for C1-1-00,
axial resistance NRd is obviously lower than Nud. As a result, the reduction factor μfi was kept at 0.55 for C1-2-00 to C1-6-80. Hence,
conventional axial resistance Nud cannot be longer used to deter- the effect of initial load level was experimental observed from
mine the initial load for columns subjected to uniaxial bending. specimens C1-1-00 and C1-2-00, whereas the effect of eccentricity
The initial applied load N 0 of the column when a fire occurs can be was obtained from the test results of the remaining specimens.
represented by point I at N 0 ¼ μf iN R d. If this column is subjected to As shown in Fig. 2(a) and (b), all column specimens were 3.3 m
a larger first-order eccentricity, such that e′ 4 e, it has a lower long with 300 mm square cross-section and 40 mm concrete
ambient design axial resistance represented by point P2′ cover. They were all cast from concrete with mean cylinder
(N′R d o NR d). As a result, the initial applied load is reduced to compressive strength of fc ¼55.3 MPa and were stored in room
point I′ at N′0 ¼ μf iN′R d, and N′0 oN 0 . In other words, when condition for two years. The columns were all reinforced with
heating occurs, the initial applied load N0 of a column subjected six 20 mm-in-diameter reinforcing bars (6T20) and 8 mm-in-
to uniaxial bending is inversely proportional to the first-order diameter and 901-hooked stirrups at a spacing of 250 mm
eccentricity e. Besides the effect of eccentricity, the initial load (R8@250). The reinforcing bars and stirrups had a mean yield
level μfi is also a variable of interest in this study. strength (fy) of 550.2 and 496 MPa, respectively (Table 2). Two
It is also noted in Fig. 1 that at an elevated temperature T, the concrete blocks with the dimensions in mm of 500  500  300
M–N interaction diagram of the column regresses to a smaller were designed at both ends of the specimens (Fig. 2(a)). Hence,
shape. The reason is due to heat transfer, both concrete and uniaxial eccentricity ez was produced by shifting the centroidal
reinforcing bars within the cross-section experience a non uniform positions of these two end blocks, where the axial load was
distribution of elevated temperatures and undergo deterioration applied, relative to the column centroid (Fig. 2(c)).
in mechanical properties, leading to a reduction in the load In order to measure temperature distribution within a heated
resistance of the cross-section. column specimen, fourteen butt-welded chromel–alumel K-type
thermocouples were embedded at three cross-sections marked ‘A’,
‘B’, and ‘C’ along the column (Fig. 2(a) and (d)). At these cross-
sections, stand frames were formed by tightening 1 mm-diameter
heat-resistant nickel–chromium alloy wires to reinforcing bars.
Then, thermocouples were fixed to their designed positions using
the wires as well.
In this test programme, all the columns were exposed to a pre-
designed fire condition with the same axial restraint ratio αr,
which was the normalised restraint stiffness Kr with respect to the
column axial stiffness Kc, until failure.

3.2. Experimental set-up

An elevation of the experimental set-up for a pin-ended


column specimen subjected to uniaxial bending and restraint
under fire conditions is shown in Fig. 3. The column specimen
was connected to a steel bulkhead of a self-reaction test rig
through a knuckle bearing block and was loaded from a 5000 kN
Fig. 1. Initial applied load of columns under uniaxial bending. servo-controlled hydraulic actuator fixed to another bulkhead

Table 1
Test specimens.

Series S1—Column specimens C1-1-00 C1-2-00 C1-3-25 C1-4-40 C1-5-60 C1-6-80

Uniaxial eccentricity (mm) 00 00 25 40 60 80


Reduction factor μfi ¼N0/NRdEC2 0.45 0.55 0.55 0.55 0.55 0.55
Initial applied load N0 (kN) 1720 2100 1700 1400 1100 904
K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13 3

Fig. 2. Test specimen. (a) Column elevation, (b) column cross-section, (c) uniaxial eccentricity and (d) position of thermocouples.

Table 2 constant rate of 10 1C/min. It was then kept unchanged at 800 1C until
Material mechanical properties. the specimen failed. This non-standard heating curve was designed
based on the capacity of the electric furnace used, which has a
Concrete Main rebars (6T20) Stirrups (R8@250)
maximum heating rate of 12 1C/min and can be accidentally shut
fc (MPa) Ec (MPa) fy (MPa) fu (MPa) Es (MPa) fy (MPa) fu (MPa) Es (MPa)
down during test if either a significant amount of free water is
instantly released or temperature reaches 900 1C. However, this pre-
55.3 30417.8 550.2 648.8 196 496 567.4 218 designed curve can be used to obtain better understanding of the
specimen structural behaviour at elevated temperatures owing to
these reasons: (i) It was observed from the Cardington series of fire
tests that closely simulated real building fires during a compartment
located at the other end of the test rig. Axial restraint was fire, before flash over took place, gas temperature developed at a mean
generated by using a transverse steel beam supported by two rate of around 5 to 15 1C/min, which was much more slowly and
small A-frames at the ends of the beam. At the mid-span of this so- gently compared to the standard ISO 834 fire at a rate of around 60 1C/
called restraint beam, one beam flange was fixed to the actuator min [17]; and (ii) experimental data obtained can be useful for the
whereas the other flange was connected to the specimen through analysis of thermal and mechanical responses of columns in industrial
another knuckle bearing block. Hence, the column specimen could buildings which may allow for moderate elevated working tempera-
be tested in pinned and pinned-on-roller end conditions with axial tures, columns in normal buildings but are located far away from a
restraint exerting onto the pinned-on-roller end. During test, the localised fire, and fire-protected columns.
2700 mm-long middle part of the specimen was placed inside an Temperature within the furnace was measured by eight built-in
electric furnace whereas the protruded 300 mm-long portion at temperature sensors and another eight K-type thermocouples
each column end remained unheated (Fig. 3). located at interior positions. Temperatures at the top, bottom, left,
and right surfaces of the column mid-height cross-section were
3.3. Heating facility and temperature acquisitions measured by four thermocouples, namely, TC-T, TC-B, TC-L, and
TC-R, respectively (Fig. 4).
As shown in Fig. 4, an electric furnace assembled from four
individually-controlled heating panels, was used to generate pre- 3.4. Measurements of deflections
designed fire conditions. There were two temperature sensors
embedded inside the long and the short internal walls of each Deflections of specimens and movements of test facilities were
heating panel, namely, TS-iL and TS-iS, respectively (i¼1 to 4). measured by a total of fourteen linear variable differential trans-
These temperature sensors were connected to the input terminal formers (LVDTs) (Fig. 3). LVDTs L1, L2, L3, and L4 were installed at
of a temperature controller, which was plugged to a power column end A for the axial displacement, namely, uA, of the
controller. Based on the sensor recording, the temperature con- column centroid at this end. The axial displacement at column
troller dictated whether the power controller should be switched end B, namely, uB, was obtained from L5, L6, L12, and L13. The
on or off to achieve a pre-defined heating curve. difference between the centroidal axial displacements of the ends
Fig. 5 shows the typical heating curve used in this experiment was the measured axial deformation of the specimen (u ¼uA−uB).
programme. The room temperature was assumed to be 25 1C. At the Since conventional LVDTs could not work properly at a tempera-
first step, temperature was increased at a rate of 5 1C/min to 200 1C. ture higher than 100 1C, an intermediate measurement system con-
Temperature was subsequently held at 200 1C for 85 min in order to sisting of ceramic rods with a low thermal expansion coefficient was
gradually purge free water content out from the column to achieve a designed and fabricated to transmit the recorded lateral deflections
low level of relative humidity and to avoid the furnace from electric of specimens from within the furnace to external LVDTs. The actual
tripping. In the third step, temperature was increased up to 800 1C at a column lateral deflections were obtained by subtracting from L8
4 K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13

Fig. 3. Elevation view of experimental set-up.

Fig. 4. Heating facility and temperature acquisitions.

3.5. Axial restraint system

In this experimental study, the axial restraint generated from


surrounding structural members on a heated column in reality
was closely simulated by a steel beam with properties listed in
Table 3. This beam was supported by an A-frame at each end and
was perpendicularly connected to the column specimen at its mid-
span (Fig. 6). During test, the transverse beam was subjected to a
concentrated load at mid-span due to thermal elongation of the
heated column. Hence, the axial restraint stiffness exerting on the
column was arose from the flexural stiffness of the beam, which
was determined from the linear relationship between the imposed
load recorded by load cells LC1 and LC2 and the beam mid-span
deflection obtained from L9, L10, and L11.

3.6. Experimental procedure

Fig. 5. Heating curve used in fire tests. In this experimental programme, all the column tests were
conducted under an identical transient heating state condition. In
order to prevent the heating elements of the electric furnace from
being physically damaged due to spalled concrete debris, each test
measurements the calculated thermal expansion of ceramic rods specimen was covered by protection steel meshes (Fig. 4). Once
based on a measured coefficient of 7.74  10−6/1C [18] and its heated the specimen was installed into the test rig and thermocouples
length. The remaining LVDTs, namely, L7, L9, L10, L11, and L14, were were fixed in place, the furnace was closed, all the gaps were
used for the determination of axial restraint stiffness, which will be sealed with insulation wool and cloth, and LVDTs were positioned.
explained in Sections 3.5 and 5.2. Preload up to 15% of test load N0 listed in Table 1 was applied and
K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13 5

Table 3
Restraint steel beam.

Beam no. Width (mm) Height (mm) tflange (mm) tweb (mm) Area (cm2) I (cm4) Lb (m) Kb (kN/mm)

Joist 258  146  33 146 258.3 9.1 6.1 41.8 4917 2.48 31.72

Restraint beam

L9

LC-1

A-Frame

Fig. 6. Axial restraint system. (a) Three-dimensional view and (b) elevation view.

Fig. 7. SAFIR numerical models. (a) Modeling of column cross-section (961 models, 906 elements), (b) 3-D beam element and (c) modeling of column height (32 noises,
15 elements).

released for two times in order (i) to eliminate undesirable slacks the actuator in order to reach the N0 value. Since the actuator
within the test system due to closing of gaps; (ii) to check for could no longer maintain the load, the column was deemed to
proper working condition of test instrumentation; and (iii) to have failed.
predetermine the suitable rate for displacement of the actuator.
The value of 15%N0, which is about 8% of the specimen axial
resistance, is to ensure that the column still behaves in elastic 4. Numerical simulation
manner during preloading stage. Subsequently, the specimen was
loaded at the constant rate determined up to 100% of N0. This load Thermal and structural analyses were conducted on S1 speci-
level was maintained constant for 30 min before the electric mens using computer program SAFIR [15]. This software has been
furnace was switched on to expose the specimen to the pre- well verified by a fair number of column test results carried out at
designed heating curve shown in Fig. 5. While the applied load N0 different laboratories [19,20]. A mesh sensitivity study was con-
was still kept unchanged, readings of thermocouples, LVDTs, and ducted by the authors to obtain a proper discretization of 900 two-
load cells were recorded at 30-s intervals. The test was terminated dimensional (2-D) rectangular solid elements with 961 nodes for
when the hydraulic actuator, which was operated in a heat transfer analysis (Fig. 7(a)), in which the EC2 Pt.1.2 [13]
displacement-controlled mode, attainted a sudden drop of the specifications of thermal conductivity, specific heat, and density
reading of N0 and a significant displacement had to be provided for for siliceous concrete were adopted. A concrete moisture content
6 K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13

of 5.4% of weight was assumed based on the local highly humid


tropical climate of Singapore. The relative emissivity were set
0.5 and the convection coefficients on hot and cold surfaces were
assumed the respective default values of 25 and 9 W/m2 K. These
values were validated by fire tests conducted on columns in the
same laboratory conditions [10]. A model including fifteen 3-D
beam elements and 32 nodes was also determined from mesh
sensitivity study for structural analysis of the test specimens (Fig. 7
(b) and (c)). Axial restraint was simulated by an elastic 3-D beam
element with an axial stiffness obtained from experiment
(Sections 2.5 and 5.2). In order to account for the column initial
imperfection, a half-sine curve with a typical relative crookedness
of 1/400 of the column length Lc according to EC2 Pt.1.1 [16] was
assumed for the numerical simulations. This assumption was
made since the coarseness of concrete surfaces prevented accurate
measurements of initial crookedness of the specimens. This was
also to account for the minor role of the gravitation effects of the
horizontally-tested specimens. Material tests were conducted on
heated concrete cylinders using tube electric furnace. It was
observed that when average temperatures within the samples
were 340, 420, and 670 1C, strength reductions factors obtained
were of 0.83, 0.76, and 0.30, respectively. These test values were
close to the specification of EC2 Pt.1.2 [13], in which strength
reductions factors of siliceous concrete at 300, 400, and 700 1C are
0.85, 0.75, and 0.30, respectively. Hence, elevated-temperature
material models for concrete specified in EC2 Pt.1.2 [13] were
adopted in the numerical models. The Eurocode properties at
elevated temperatures for reinforcing steel were also adopted
since there was no test result available.

5. Test results and discussions

5.1. Temperature distribution

Fig. 8(a) shows the performance of heating facility obtained


from built-in temperature sensors, which was identical for all
specimens. Prior to 650 1C, these temperature–time curves were
reasonably close to the pre-defined heating curve shown in Fig. 5.
However, beyond 650 1C, the temperature increase rate could not
be maintained at 10 1C/min. As a result, the maximum tempera-
ture of 800 1C could only be achieved after 195 min, instead of
180 min as initially planned. Fig. 8(b) depicts temperatures mea-
sured at thermocouples TC-1 and TC-5, which were at a 70 mm
distance from the internal wall of heating panel (Fig. 4). It can be
seen that the temperatures obtained were identical. However,
temperatures at these points were lower than those measured at
heat elements due to heat flux. Temperatures measured at the
column surfaces are shown in Fig. 8(c). These temperatures were
lower than those at the heating elements as well as inside the
furnace. This is due to a greater distance from the specimens to Fig. 8. Temperature measurement. (a) Built-in temperature sensors, (b) thermo-
heating elements compared to those of TC-1 and TC-5. Besides, couples inside the furnace and (c) thermocouples located at specimen surfaces.
since there were steel meshes covering around the column
to protect the heat elements from damage due to concrete spalled section of specimen C1-3-25. Points 1, 8, and 4 were respectively
debris (Fig. 4), the heat flux to the surfaces of the specimen was at a corner main bar, stirrups, and concrete at the centre of the
also impeded. cross-section B–B in Fig. 1(d). Due to their relative positions from
Since the geometric properties and the test conditions of all the column surfaces, point 1 had the highest temperature profile
specimens were identical, the test results of a specimen will be whereas point 4 at the centre of section had the lowest among the
discussed in detail. In Fig. 9, the upper continuous and centered three points. Good agreement between the recorded and the
lines, respectively represent the fire curves measured at the predicted temperatures could be observed during the first
furnace heat elements and specimen surfaces. It can be noted that 210 min of the test. Beyond this time, the actual temperatures
these lines are ideally simplified from the measured curves shown were higher than the FEA predictions. At the end of the heating
in Fig. 8(c). This is to form an identical thermal boundary condition process, the temperatures measured at points 1, 8, and 4 were 386,
in numerical analysis as a comparison basis for all test specimens. 301, and 126 1C, respectively. The corresponding predicted tem-
The lower curves in Fig. 9 show the measured and the predicted peratures were at 371, 292, and 115 1C, respectively. For other
temperatures at various points within the mid-height cross- column specimens, the measured temperatures were also higher
K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13 7

Fig. 9. Temperature distribution at mid-height cross-sections (C1-3-25).

Fig. 10. Experimental stiffness of axial restraint. (a) Restraint steel beam and (b) high-stiffness test rig.

Table 4
Experimental axial restraint stiffness.

Series Column Axial restraint

2 2 2 2
Ec kN/mm ) Ac (mm ) Es (kN/mm ) As (mm ) Kc (kN/mm) KA (kN/mm) KB (kN/mm) Kr (kN/mm) αr ¼ Kr/Kc(%)

S1 30.42 88115 196.15 1885 861.59 32.31 1218.4 31.14 3.6

than those obtained from numerical analysis at the end of the 5.2. Restraint stiffness
tests. This is because in the tests, concrete spalling caused inner
parts of the column cross-section to be directly exposed to fire, The actual axial restraint stiffness KA that the transverse beam
resulting in a greater heat transfer rate. Meanwhile, concrete exerted onto the heated column specimen could be experimen-
spalling was not modelled in the numerical simulation, resulting tally determined based on the relationship between the beam
in an underestimation in numerical results. In order to adjust the mid-span deflection and the restraint force induced in the column.
numerical predictions to be closer to the experimental tempera- As explained in Section 3.5 and Fig. 6(a), the restraint force in
ture measurements, the surface heating curve had to be modified Fig. 10(a) is obtained from load cells LC1, LC2 whereas the mid-
as shown in Fig. 9. Beyond 210 min of fire exposure, an increase span deflection is calculated from the readings of LVDTs as L1l-0.5
rate of 10 1C/min was applied so that the fire curve reached 800 1C (L9+L10). A value of 32.3 kN/mm of KA, which is the slope of the
at 236 min. This adjustment was conducted on a trial-and-error curve, was obtained. This experimental result is of 1.9% different
basis to fit the numerical and measured temperature curves after from the design flexural stiffness of the steel beam shown in
210 min. Hence, the time-dependent numerical temperature dis- Table 3.
tribution could be used for the subsequent structural analysis on During the tests, there were small relative movements between
the numerical model presented in Section 4 and Fig. 7(c). the bulkheads (Fig. 2) due to the high stiffness of the self-reaction
8 K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13

Fig. 11. Column deflections. (a) Axial deformation-time relationship (u-curves) and (b) mid-height lateral deflection-time relationship (v-curves).

test rig. Hence, an additional restraint with the stiffness, namely, FEA u-curves during the first 190 and 170 min of the tests,
KB, was experimentally determined to be 1218.4 kN/mm based on respectively. Beyond 170 min, the test and the FEA u-curves of
the gradient of the linear graph between the bulkhead relative C1-2-00 started to deviate. However, the test and the FEA u-curves
deformations measured by L7 and L14 and the total measured of C1-1-00 stopped at 288 and 310 min, while those u-curves of
axial force Nt ¼ N0+ΔNaT (Fig. 10(b)). For numerical modelling, an C1-2-00 respectively stopped at 251 and 270 min, which were
idealised elastic spring was adopted at the column end A with an quite close to each other. For C1-1-00, there were some spikes at
equivalent stiffness Kr given as 1/Kr ¼1/KA+1/KB. The column axial both the test u- and v-curves at 216, 243, 269, and 288 min. This
stiffness is Kc ¼(EcAc+EsAs)/Lc where Ec and Es are the respective was because the LVDT readings were affected by explosive con-
measured elastic moduli of concrete and reinforcing steel; Ac and crete spalling, as presented in Section 5.6. For the remaining
As are cross-sectional areas of the concrete and reinforcing bars, specimens, the experimental results agreed reasonably well with
respectively; and Lc is the effective length of the test column. the FEA predictions during the first 160 min. After that, although a
Values of Kr, Kc, and the corresponding axial restraint ratio αr ¼Kr/ similar trend could be observed from the test and the FEA v-curves
Kc are included in Table 4. The value of 0.036 obtained for αr was of C1-3-25 and those u-curves of C1-4-40, deviations between the
within a practical range of 0.01 to 0.35 as mentioned in the test and the FEA results occurred in both axial deformation and
Broadgate Fire Report 1991 [21] for steel columns, whereas there is mid-height lateral deflection of C1-5-60 and C1-6-80. At 159 min
no reported range of restraint ratio for concrete columns. of the C1-5-60 test, spikes on both the test u- and v-curves were
resulted from the occurrences of concrete spalling. For C1-6-80,
5.3. Deflections at elevated temperatures from 199 min onwards the reading of L8 reached its limit so that
its test v-curve remained unchanged at a value of 43.3 mm,
Fig. 11(a) and (b) shows the profiles of axial deformations and although the column continued to deform. The divergence in the
mid-height lateral deflections of all the test specimens, which will test and the numerical results observed from Fig. 11(a) will be
be hereafter respectively denoted as u- and v-curves. The test and discussed in Section 5.4 of thermal-induced axial forces, which
the FEA curves are plotted in continuous and hidden lines, have a close relationship with axial deformation.
respectively. The FEA curves were obtained from the numerical Fig. 11(b) shows that at elevated temperatures, lateral deflec-
model introduced in Section 4, incorporating the modified thermal tion of C1-2-00, which had an initial load level of 0.55, was higher
analysis and the actual axial restraint stiffness obtained from than that of C1-100 loaded with an initial level of 0.45. It is shown
experiment (Section 5.2). that both the test and the FEA v-curves of C1-2-00, C1-3-25, C1-4-
Relatively good agreement between the test and the FEA curves 40, C1-5-60, and C1-6-80, which all were under the same initial
of C1-1-00 and C1-2-00 could be observed. The test u-curves of load level of 0.55, were in an increasing manner with the extent of
C1-1-00 and C1-2-00 were nearly parallel to the corresponding eccentricity. This observation confirms that at elevated
K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13 9

Fig. 12. Thermal-induced restraint forces (ΔN-curves).

Fig. 14. Loading-heating processes in restrained columns.

and the FEA ΔN-curves of C1-6-80 with the largest eccentricity of


80 mm, experienced the steepest gradient compared to all the other
specimens. The corresponding curves of C1-2-00 with zero eccentri-
city had the lowest gradient. Both the test and the FEA ΔN-curves of
Fig. 13. Maximum normalised restraint forces. other specimens, namely, C1-3-25, C1-4-40, and C1-5-60 were
arranged in a sequential manner proportional to the magnitude of
eccentricity. Fig. 7 shows that unlike columns under pure compression
temperatures, a column subjected to a higher initial load level or a with the maximum axial resistance of Nud, axial axial NRd of an
larger eccentricity has a higher lateral deflection. This is because eccentrically-loaded column is eccentricity-dependent, such that if
when either initial load level or eccentricity was increased, the e′ 4 e then N′R d oN R d. As a result, with the same reduction factor μfi,
second-order effect associated with deterioration of material when first-order eccentricity e increases, the initial applied load N0
stiffness at elevated temperatures was also more significant, decreases, resulting in an increase in normalised restraint forces.
leading to an increase in the development of lateral deflection. Third, the numerical predictions were always higher than the test
results. The maximum FEA normalised restraint forces of C1-1-00 and
5.4. Thermal-induced restraint forces C1-2-00 were 12.2 and 7.8%, respectively. In the C1-1-00 and C1-2-00
tests, these maximum values were 3.8 and 2.4%, respectively.
Fig. 12 shows the development of the ratios between the The divergence in the test and the numerical results of axial
thermal-induced restraint forces ΔNaT and the initial applied load deformation and thermal-induced restraint forces can be
N 0 , which will be hereafter abbreviated as normalised restraint explained based on the phenomena of concrete spalling and
forces, or ΔN-curves. The test and the FEA results of all the column transient strain.
specimens, which are respectively plotted in continuous and During fire tests, when the surface concrete layers spalled off,
hidden lines, had similar trends as illustrated by the maximum the interior concrete parts and reinforcing steel were directly
values of normalised restraint forces shown in Fig. 13. exposed to fire. Thus, the axial force applied on the column was
First, the higher the load level μfi, the lower the normalised sustained by the smaller-in-area but hotter and weaker parts of
restraint forces were attained. C1-1-00 and C1-2-00 were identical the cross-section and the column experienced greater mechanical
specimens subjected to different initial applied loads with reduction deteriorations. As a result, although the thermal boundary condi-
factors of 0.45 and 0.55, respectively. The test and the FEA ΔN-curves tion for numerical analysis was modified to fit with the measured
of C1-2-00 were both lower than the corresponding curves of C1-1- temperature distribution, the numerical model could not reflect
00. This is because under a higher level of axial force, the column the fact that due to concrete spalling the column had a decreased
experiences a higher axial deformation at the initial loading stage at axial stiffness and was weaker to response to thermal action since
ambient condition and then has less ability to response to thermal thermal elongation was offset by mechanical deformation. Hence,
action, resulting in a lower development of restraint forces. the actual axial deformation and thermal-induced restraint forces
Second, at the same initial load level, columns under larger were both different from those predicted by numerical simulation
eccentricities experienced higher normalised restraint forces. The test that ignores the physical effect of concrete spalling.
10 K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13

On the other hand, the numerical analysis used in this study shown in Fig. 15(b), in which the outermost concrete cover spalled off
adopted the Eurocde model at elevated temperatures for concrete. from the corners of the column cross-section.
In this constitutive material model, total strain εtot is a combina-
tion of thermal strain εth and mechanical strain εm, which 5.6. Failure modes
implicitly includes stress-related strain εs, transient strain εtr,
and creep strain εcr. Hence, the Eurocode model can be considered The failure modes of an RC column subjected to uniaxial bending at
as an implicit model. Transient strain develops in concrete under elevated temperatures are illustrated in Fig. 16. The column of
unchanged compressive stress when temperature increases [22]. eccentricity-dependent resistances (NRd, MRd) at room temperature is
That means this type of concrete strain only changes when subjected to an initial load N 0 ¼ μf iN R d when heating occurs. This
temperature increases at a constant level of stress whereas it initial condition is represented by point I. As temperature rises,
remains unchanged with increasing stress at a constant tempera- material failure may occur at point M where the non-linear load–
ture. Different from unrestrained columns, when an axially- moment curve coincides with the regressing M–N interaction diagram
restrained RC column is heated, it experiences the combined resulted from solely strength degradations of materials. Point S at the
actions of increasing compressive stress and elevated temperature. peak of the load–moment curve represents the column failure in an
As can be seen in Fig. 14, these simultaneous processes of loading instability manner, which develops run-away lateral deflection result-
and heating can be divided into two separate independent phases: ing in a rapid increase of secondary bending moment whereas the
(i) heating with constant loading (A1B, A2C, and A3D); and (ii) axial force reaches the maximum value and starts to reduce, so-called
loading with constant temperature (BA2, CA3, and DA4). In numer- stability failure. After passing through point S, if the descending load–
ical analyses adopting implicit model, transient strain of concrete moment curve coincides with the continuously-regressing M–N inter-
is considered in both phases. However, in reality transient strain action diagram at point C, the column would have failed in a
only exists in the heating phase by definition. This fact results in combination manner of stability and material failures, so-called
the divergence between the results obtained from test and combined failure mode. The time intervals, namely, tM, tS, and tC, for
numerical analysis using implicit models. However, since there the load–moment curve to develop from initial point I to failure points
have been limited analytical and numerical models incorporating M, S, and C can be respectively referred to as the material, stability, and
reliable elevated-temperature material properties to predict con- combined failure times.
crete spalling and to consider transient strain explicitly [23], the Images of the failed specimens are shown in Fig. 17. All six
aspects should be quantitatively studied in future. specimens had single curvature bending with significant mid-
height flexural deflections, indicating stability failure. Combined
failure mode can be observed in specimens C1-1-00, C1-2-00,
5.5. Concrete spalling
C1-3-25, C1-4-40, and C1-5-60, which had localised damages due
to concrete crushing as evidences of material failure.
It was observed from the tests that severe concrete spalling
occurred in most of the test specimens. Due to the steel meshes
wrapping around the specimens to protect the heating elements from
concrete debris, it was impossible to capture concrete spalling images
within the enclosed furnace. Hence, concrete spalling could only be
recorded when there was either a gentle or an explosive sound heard
from the furnace. Fig. 15(a) shows the recorded times at which
concrete spalling occurred in C1-1-00 test. The first gentle noise was
recorded at 161 min, followed by other cracking noises at 216, 243,
269, and 288 min. Temperature of concrete surface when the first
concrete spalling occurred was 328 1C. The column failed at 310 min
with a loud noise of explosive spalling. For C1-2-00, C1-3-25, C1-4-40,
and C1-6-80, the first concrete spalling was recorded at 149, 154, 167,
and 160 min, with the corresponding surface temperatures of 276, 297,
354, and 320 1C, respectively. In the C1-5-60 test, a loud noise of
explosive spalling was recorded at 159 min followed by premature
failure. Images of spalled concrete debris captured after testing are Fig. 16. Failure modes of columns at elevated temperatures.

Fig. 15. Concrete spalling observed in fire tests. (a) Specimen C1-1-00 and (b) debris of spalled concrete.
K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13 11

Fig. 17. Images of failed specimens. (a) C1-1-00, (b) C1-2-00, (c) C1-3-25, (d) C1-4-40, (e) C1-5-60 and (f) C1-6-80.

Material failure at the mid-height cross-section of C1-1-00 is


shown in Fig. 18. Since the column resistance was reached at failure, Cracking at tension zone

concrete in compression zone crushed and broke off into pieces. Due
Crushing at compression zone
to elevated-temperature effect, concrete spalled off as well as
debonded from longitudinal bars and stirrups. Being subjected to high Failure of stirrups
compression forces and exposed to high temperatures, the long-
itudinal bars in compression zone buckled between two stirrups. Buckling of compression bar
The stirrups themselves failed as the 901 hooks were bent outwards.

5.7. Failure times

Fig. 19 depicts in a four-axis combined chart the typical structural


responses obtained from experimental and numerical analyses. The
left and the right vertical axes are for the normalised restraint forces
Concrete spalling Debonding
(ΔN-curve) and the mid-height lateral deflection (v-curve), respec-
tively. The lower and the upper horizontal axes represent the fire
exposure time and the modified time-dependent temperature at the
column surfaces, respectively. The test and the FEA curves are plotted Fig. 18. Failure at mid-height cross-section of C1-1-00.
in continuous and hidden lines, respectively. At the initial points
(IN1 and IN2), there was no restraint force and the measured lateral
deflection (Iv2) was close to that predicted by FEA (Iv1). Prior to about stability failure occurred in C1-6-80 since the ΔN-curve dropped at
160 min of fire exposure (point A), reasonable agreement between the its peak. Second, the remaining five specimens had combined
test and the FEA curves could be observed. As explained in Section 5.5, failure since there were both peaks and end points on their test
in this stage of agreement (IA), there was no concrete spalling. After ΔN-curves. Failure modes and failure times obtained from the tests
concrete spalling had occurred at point A, the increase rate of the test and predicted by FEA are compared in Table 5. It is shown that the
ΔN-curve gradually reduced until the curve reached its peak (point test failure times in stability and combined modes of C1-1-00
SN2). As explained in Fig. 16, since the axial force attained the agreed well with the FEA predictions with, respective ratios of 1.10
maximum value at point SN2, this point represents stability failure. and 0.93. Meanwhile, C1-5-60 had large discrepancies between
During the AS stage, the test v-curve gradually increased to a run- the test and the FEA stability failure times as well as those of
away deflection, indicating that the column was about to buckle. After combined failure times with ratios of 1.61 and 1.59, respectively.
passing through point SN2, the test ΔN-curve descended slightly and The premature combined failure, which is denoted as PC in
dropped at end point CN2, which represents the combined failure Table 5, was due to an explosive spalling occurred at 159 min of
mode illustrated in Fig. 16. It is noteworthy that the peak and the end the test (Section 5.5). With the respective mean ratios of 1.27 and
point of the test ΔN-curve (SN2 and CN2) all had shorter time co- 1.13, the stability failure times and the combined failure times of
ordinates compared to those of the corresponding points of the FEA all test specimens were generally overpredicted by numerical
ΔN-curve (S′N1 and C′N1). This means that the numerical analyses analyses which ignored the effect of concrete spalling.
overpredicted both stability failure times and combined failure times
of all the test specimens. In Fig. 19, the co-ordinates of points SN2 and
CN2 in the lower horizontal axis can be respectively referred to as the 6. Conclusions
measured stability failure time tSTest and combined failure time tCTest .
Similarly, the corresponding co-ordinates of points S′N1 and C′N1 are This paper presents the fire tests conducted on six full-scale
the FEA-predicted stability and combined failure times, which are specimens to enhance the understanding of structural responses
denoted as tSFEA and tCFEA , respectively. of reinforced concrete columns subjected to the combined effects
Some observations can be obtained from the test and the FEA of uniaxial bending and axial restraint at elevated temperatures.
ΔN-curves of all the column specimens shown in Fig. 12. First, The design of specimens, test set-up, instrumentation, and
12 K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13

Fig. 19. Failure modes and failure times.

Table 5 Acknowledgements
Test and FEA failure times.
The authors would like to acknowledge the financial contribu-
Specimens TEST results FEA predictions Comparison
tions of MINDEF-NTU joint applied R&D co-operation program
tSTest tCTest Failure tSFEA tCFEA Failure t S FEA t C FEA with research fund MINDEF-NTU/JPP/FY05/14, as well as the
(min) (min)
mode (min) (min)
mode tSTest tCTest research project AnSTAR M47030016. The authors are also grateful
to Mr. Yao Yao for his contribution in the early stage of the
C1-1-00 244 310 C 268 288 C 1.10 0.93 experimental programme.
C1-2-00 195 251 C 256 270 C 1.31 1.06
C1-3-25 215 252 C 256 274 C 1.19 1.03
C1-4-40 209 221 C 256 280 C 1.23 1.27 References
C1-5-60 159 180 PC 256 286 C 1.61 1.59
C1-6-80 214 – S 254 294 C 1.19 –
[1] T.T. Lie, T.D. Lin, Influence of restraint of fire performance of reinforced
Mean 1.27 1.13 concrete columns, in: First International Symposium, International Association
COV 0.16 0.32 for Fire Safety Science, 1985.
[2] Y.C. Wang, D.B. Moore, Effect of thermal restraint on column behaviour in a
Notes: S: Stability—C: Combined—PC: premature combined failure mode. frame, in: The Fourth International Symposium of Fire safety science, Ottawa,
1994, pp. 1055–1066.
[3] P.G. Shelpherd, I.W. Burgess, On the buckling of axially restrained columns in
procedure of the experimental programme conduced are fire, Eng. Struct. 33 (2011) 2832–2838.
[4] J.P.C. Rodrigues, I.C. Neves, J.C. Valente, Experimental research on the critical
described in detail. The novelty of the test programme is that it
temperature of compressed steel elements with restrained thermal elonga-
is capable of simultaneously simulating the effects of uniaxial tion, Fire Saf. J. 35 (2000) 77–98.
eccentricity and thermal-induced restraint so that axially- [5] Z.F. Huang, K.-H. Tan, Analytical fire resistance of axially-restrained steel
restrained RC columns can be tested under various levels of columns, J. Struct. Eng. -ASCE 129 (11) (2003) 1531–1537.
[6] F. Ali, A. Nadjai, S. Silcock, A. Abu-Tair, Outcomes of a major research on fire
uniaxial eccentricity and initial applied load at elevated tempera- resistance of concrete columns, Fire Saf. J. 39 (2004) 433–445.
tures. The data acquired, which are clearly discussed and com- [7] A. Benmarce, M. Guenfoud, Experimental behaviour of high-strength concrete
pared with numerical predications using SAFIR program, include columns in fire, Mag. Concr. Res. 57 (5) (2005) 283–287.
[8] K.-H. Tan, W.S. Toh, G.H. Phng, Z.F. Huang, Structural responses of restrained
structural responses of temperature-dependent axial and lateral steel columns at elevated temperatures. Part 1: Experiments, Eng. Struct. 29
deformations, thermal-induced restraint forces, failure modes and (2007) 1641–1652.
failure times. The research outcomes are as follows: (1) Column [9] Z.F. Huang, K.-H. Tan, Structural response of restrained steel columns at
elevated temperatures. Part 2: FE simulation with focus on experimental
lateral deflection at elevated temperatures is adversely affected by secondary effects, Eng. Struct. 29 (2007) 2036–2047.
both eccentricity and initial load level; (2) the development of [10] Z.F. Huang, K.-H. Tan, G.H. Phng, Axial restraint effects on the fire resistance of
thermal-induced restraint forces increases with eccentricity but composite columns encasing I-section steel, J. Constr. Steel Res. 63 (2007) 437–447.
[11] B. Wu, Y-H. Li, Experimental study on fire performance of axially-restrained
decreases with initial load level; and (3) numerical analyses that NSC and HSC columns, Struct. Eng. Mech. 32 (5) (2009) 635–648.
ignore concrete spalling and implicitly accounts for concrete [12] A.M.B. Martins, J.P.C. Rodrigues. Behavior of concrete columns subjected to
transient strain, overestimate normalised thermal-induced fire, in: Proceedings of the Sixth International Conference Structures in Fire,
USA, 2010, pp. 181–188.
restraint forces. It can be also seen that the effects of concrete
[13] European Committee for Standardization (CEN), Design of Concrete Structures—
spalling and transient strain should be quantitatively studied in Part 1–2: General Rules, Structural Fire Design, EC2-1.2, EN 1992-1-2:2004, 2004.
future by using advance finite element software. Besides, the [14] ACI/TMS Committee 216. Code Requirements for Determining Fire Resistance
combined effect of biaxial bending and restraint should also be of Concrete and Masonry Construction Assemblies (ACI 2.16.1-07/TMS-216-07).
American Concrete Institute, 2007.
investigated for RC columns under standard as well as real fire [15] J.M. Frassen, SAFIR. A thermal/structural program modelling structures under
conditions. fire, Eng. J., A.I.S.C 42 (3) (2005) 143–158.
K.-H. Tan, T.-T. Nguyen / Fire Safety Journal 60 (2013) 1–13 13

[16] European Committee for Standardization (CEN), Design of Concrete Structures— [20] J.C. Dotreppe, J.M Franssen, Y. Vanderzeypen, Calculation method for
Part 1–1: General Rules and Rules for Buildings, EC2-1.1, EN 1992-1-1:2004, 2004. design of reinforced concrete columns under fire conditions, ACI Struct. J. 96
[17] C.G. Bailey, The influence of the thermal expansion of beams on the structural (1) 9–18.
behaviour of columns in steel-framed structures during a fire, Eng. Struct. 22 [21] Steel Construction Institute. Structural Fire Engineering Investigation of
(2000) 755–768. Broadgate Phase 8 Fire. Steel Construction Industry Forum. Fire Engineering
[18] Q. Zhenhai, Shear Behaviour of Steel Members and Beam-to-column Joints Group. June, 1991.
Under Elevated Temperatures, Ph.D. Thesis, Nanyang Technological University, [22] Y. Andeberg, Fire-exposed Hyperstatic Concrete Structures—An Experimental
Singapore, 2007, pp. 113. and Theoretical Study. Division of Building Fire Safety and Concrete Structures,
[19] J.M. Franssen, J.B. Sheleich, L.G. Cajot, W. Azpiazu, A simple model for fire Lund Institute of Technology, 1976.
resistance of axially loaded members—comparison with experimental results, [23] V.K.R. Kodur, N.K. Raut, Fire resistance of reinforced concrete columns—state-
Constr. Steel 37 (1996) 175–204. of-the-art and research needs, 255, ACI Special Publication97–124.

You might also like