Вы находитесь на странице: 1из 34

ABSTRACT

Beam-Columns are the main structural elements of the structure, if fails effects globally, so
design of column in all aspects is an important step to take first while considering the stability
of structure. Current practice of IS codes doesnt consider the some of the effects on beamcolumn such as fire, which causes a thermal gradient in structural elements that causes a
thermal bowing and moment reversal due to shift in center of stiffness. Some of the foreign
codes like Eurocode and American consider fire design by considering the uniform thermal
gradient, but non-uniform thermal gradients are not taken into account which causes a major
distortion in P-M interaction curves (reduction in plastic capacity). Therefore, there is a need
to predict the plastic capacity of beam-column under realistic fire scenario.
Keywords: Steel, Reinforced concrete, Beam-column, Fire, Thermal gradient, Thermal
bowing, Shift in center of axis, P-M interaction curve.

iii

LIST OF FIGURES

FIGURE NO.

CAPTION

PAGE NO.

Structural-fire test furnace

Temperature zones for the direct validation model and


thermocouple locations

Discretized cross-sections of (a) C1-S, (b) C1-W, (c) C2-S, and


(d) C2-W

Combined axial load (P)moment (M) capacity diagram


normalized by the axial yield force (Py) and the plastic moment
capacity (Mp): comparison of diagram with and without thermal
gradients

Critical buckling stress (Fcr) normalized by the yield stress (Fy)


for a uniformly loaded plate at 500C with pin boundary
conditions on both ends. Shown as a function of the nondimensional slenderness ratio c,T

Plot of cross-sectional temperature distribution at t = 30 min for


(a) C2-S and (b) C2-W

Column orientation, boundary conditions, and moment


distribution

Failure patterns of beamcolumns (a) C1-S, (b) C1-W, (c) C2S, and (d) C2-W after fire tests

Failure pattern of beamcolumn C1-S: (a) experimental


specimen and (b) shell model

11

10

Averaging the temperature in plates of steel beam

14

11

Comparison of steel temperature predicted by Eq. (5) with that


obtained from finite element analysis

15

12

Temperature and ultimate stress distributions for strong and


weak axis orientation in a WF section

17

13

Variation of Es and Fy as a function of steel temperature

18

vi

FIGURE NO.

CAPTION

PAGE NO.

14

Characterizing P-M diagram for a WF section with thermal


gradient in the strong direction

19

15

Eccentricity between center of stiffness and center of geometry


of a WF section with thermal gradient

20

16

Linearized P-M diagram for a WF section with thermal gradient


in weak direction

21

17

Factors influencing behavior of RC column under fire

24

18

Discretization of column and flowchart of numerical model

25

19

Independent deformation of column segment

28

LIST OF TABLES

TABLE NO. CAPTION

PAGE NO.

Summary of fire test parameters

Fcr/Fy values (based on proposed predictive equations) for the


W8x48 flange and web plates with varying boundary conditions,
loading conditions, and temperatures

11

Recorded times for column failure in minutes

12

vii

CHAPTER 1
INTRODUCTION

1.1 General
The beam-columns are the main structural elements that are designed for combined axial load and
bending moment. The regular code based design of beam columns is based on P-M interaction
curve which is based on cover and material. The effect of fire on structural elements causes a
thermal expansion and thermal gradient which causes several members to act as beam-columns.
Thermal gradient in a member through depth is caused when member subjected to fire less than
four sides. Perimeter columns and roof beams are more prone to thermal gradient, whose top
surface is shielded by slab. As these members if restrained against thermal expansion and thermal
gradient results in thermal bowing, and shift in center of stiffness, in which when load acted
through geometrical centroid causes additional moments that causes reduction in plastic capacity
of a section. Thus the thermal gradient causes the distortion in P-M interaction curves that may be
conservative or non-conservative when compared to regular code of practice.

CHAPTER 2
STEEL BEAM-COLUMNS SUBJECTED TO FIRE
Few studies have addressed the performance of steel beam-columns subjected to thermal gradient
through their depth due to non-uniform heating.

2.1 Quiel SE. et al. (2011)


Quiel SE. et al. (2011) examined that during a fire, columns on the perimeter of a building will be
subjected to moments induced by both a thermal gradient and the restraint of axial expansion by
adjacent heated beams, which themselves develop axial load. These members thus act as beamcolumns because they were then subject to a combination of axial load plus moment caused by a
combination of gravity plus thermal loading. A two-pronged procedure was adopted to predict the
behavior of the perimeter column as a beam-column, considering both the individual member
response (including thermal gradients) and the global response (including the interactions of
adjacent members).

2.2 Quiel S.E. et al. (2014)


Quiel S.E. et al. (2014) examined the mechanics and capacity of wide flanged steel beam-columns
with non-uniform temperature and this study evaluated the adequacy of different methodologies
to predict the capacity and response caused by non-uniform thermal gradients through the depth
of beam-columns. Three models with different levels of complexity were used to evaluate the fire
response of beamcolumns under non-uniform temperature gradients: (1.) code based equations
(European and American based codes, assumes uniform thermal gradient); (2.) a fiber-beam
element model; and (3.) a shell element model that discretizes the full cross section and length and
is capable of capturing local (i.e. plate) instability.
To generate the data to validate the computational models experimental fire tests that are
performed by Dwaikata M.M.S. et al. (2010) were adopted. The experimental fire tests were used
to investigate the mechanics and capacity of steel beam-columns that develop a thermal gradient
through their depth when exposed to fire. Wide-flanged specimens were loaded axially and tested
vertically in a furnace shown in Fig.1. The placement of insulation simulated a realistic three-sided
heating scenario such as that experienced by a column on the perimeter of a building frame.
2

Table.1-Summary of fire test parameters.


Specimen
Fire exposure
Axis orientation

SFRM
Initial P/Py
thickness(mm)
C1-S
ASTM E119
Strong
44
0.25
C1-W
ASTM E119
Weak
44
0.25
C2-S
Design fire
Strong
38
0.45
C2-W
Design fire
Weak
38
0.45
All specimens are a W8x48 cross-section with A992 Grade 50 steel (yield stress,Fy = 345 MPa)
and were fabricated to a length of 3.3 m. where, S-strong axis; W-weak axis

Fig.1. Structural-fire test furnace

Four beam-columns designating C1-S, C1-W, C2-S and C2-W with test parameters shown in
Table.1 were subjected to fire for generating the test data in validating the computer models.
Structural fire testing furnace shown in Fig.1 which can monitor both standard and design fire
scenarios, was used to measure the fire resistance of steel beam-columns. All the specimens were
insulated on the sides exposed to fire and was set free on top and bottom ends which were not
exposed to fire.
All the thermal insulations, thickness of the insulation, discretization of cross section and
insulations that are removed in specific locations to induce the thermal gradient, as shown in
Fig.2&3.

Fig.2. Temperature zones for the direct validation model and thermocouple locations.

Fig.3. Discretized cross-sections of (a) C1-S, (b) C1-W, (c) C2-S, and (d) C2-W

2.2.1 Predictive models


The three models considered in the study were: the code based model, the fiber-beam element
model and the shell model. The data extracted from the fire tests was used to evaluate the results
of computational modeling of experimental specimens. A detailed description of these three
models follows.

I. Code-based model
The code based models were based on European and American standards.
Current practice for the design of steel beam-columns subject to fire was to calculate the capacity
of the members assuming a uniform temperature distribution through the depth of the section. This
assumption may be acceptable for some members, but there were cases where the member will be
heated non-uniformly, thereby developing a thermal gradient through its depth.
The European code (EN 1993-1-2) (2001) and the American code (AISC. Steel construction
manual. 13th ed.2005) neglects the thermal gradients, which permits the one-dimensional heat
transfer (e.g., a lumped mass approach) to calculate the thermal response of beam-columns and
due to this, codes doesnt consider the change in interaction of axial load and moment.
If thermal gradients are neglected the shape of the capacity envelope approximates a diamond
shape as shown in Fig.4 in which the section has failed in reality when thermal gradients are
present. This shows that design based on both codes assumes diamond shape of P-M interaction
curve. In addition to this, the approximation of codes cant provide the design for moments that
are induced by thermal gradients and shift in center of stiffness, where that should not be neglected,
which were important considerations in calculating moment demand, which should also consider
the effects of the adjacent structures that can also impose moments.
Full section yielding under combined P and M is one limit state (reaching of capacity) that leads
to failure but there can be other limit states that lead to failure beforehand such as overall buckling
and local buckling.
Takagi and Deierlein (2007) evaluated both the Eurocode and American codes for overall
buckling under uniform temperature and found that the American code significantly overestimates
the strength of sections under combined P and M, while the European code is more accurate.
5

Fig.4. Combined axial load (P)moment (M) capacity diagram normalized by the axial yield
force (Py) and the plastic moment capacity (Mp): comparison of diagram with and without
thermal gradients.

Fig.5. Critical buckling stress (Fcr) normalized by the yield stress (Fy) for a uniformly loaded
plate at 500C with pin boundary conditions on both ends. Shown as a function of the nondimensional slenderness ratio c,T.
6

Where, c ,T =

( t )F k

12(1 2 ) b

y ,T

k Ek E ,T
2

ky,T is reduction factor for yield strength; kE,T is reduction factor for youngs modulus; k is based
on boundary conditions; b is width of plate and t is thickness of plate
Quiel and Garlock (2010) have recently evaluated both codes for local plate (flange and web)
buckling based on finite element simulations and came to the same conclusion: the American code
severely overestimates the capacity of plates at elevated temperature and the Eurocode provides a
good approximation but only for larger values of the non-dimensional slenderness ratio, which
was modified for temperature T (c,T).
An example of this trend was shown in Fig.5, which predicts the critical buckling stress (Fcr)
normalized by the yield stress (Fy) for a stiffened plate (pinned on both ends) uniformly loaded at
a temperature of 500C. Fig.5 shows the normalized buckling stress trend based on both codes ,
finite element solution and a predictive equation proposed by Quiel and Garlock (2010). The
same predictive equations were used to estimate in the shell model section if local buckling was
expected.

II. Fiber-beam element model


Two types of fiber-beam elements were used: (1) the forensic fiber model which was experimental
based; (2) a priori model which was presumptive based. SAFIR, a software specially designed for
the analysis of structures exposed to fire, was used to perform uncouple thermal and structural
analysis.
In the fiber-beam element models used for this study, the 3.3-m length of the column is divided
into 64 elements, resulting in a 50-mm discretization. This discretization is based on an
optimization study where larger discretization did not yield different results. Residual stresses are
not considered in any computational model because relaxation of initial residual stresses is likely
to occur in fire-exposed steel members due to an increase in steel temperature.
Global geometric imperfections were also not included in the computational models. The beam
columns used for the experiments on which this computational study was based were purposefully
selected to experience plastic failure and therefore had little if any initial out-of-straightness.
Neglecting initial global imperfections in this case is therefore justified.

For the strong axis specimens (i.e. the specimens with thermal gradients parallel to the web),
each flange is assumed to have constant temperature through its thickness and across its width.
These temperatures are equal to those recorded by the thermocouples in the experiments. The
temperature of the web was obtained by linearly interpolating between the flange temperatures and
the temperature recorded at the middle of the web.
For the weak axis specimens (i.e. the specimens with gradients parallel to the flanges), linear
interpolation was used to obtain temperatures between thermocouples. Extrapolation using the
same linear relationships was used to project the temperature profile in the flange tips beyond the
range of the thermocouples. Examples of these piecewise linear thermal distributions were shown
in Fig.6 a and b for a strong axis section and weak axis section.

Fig.6. Plot of cross-sectional temperature distribution at t = 30 min for (a) C2-S and (b) C2-W.

Fig.7. Column orientation, boundary conditions, and moment distribution.


8

The boundary conditions, loading, and gradient orientation used for the fiber models are shown in
Fig.7. The bottom end of each specimen is modeled as fixed to approximate the bolted angle
connection used in the experiment. Vertical and rotational linear springs represent the resistance
to thermal expansion and thermal bowing provided by the piston-to-plate connection.
When compared with the experimental model with fiber beam models failure patterns (see Fig.8)
has shown buckling of the flange and web plates and this buckling was likely squashing of the
plates after full plate section yield and not a local instability. But the experimental and
computational results has shown good agreement on the location of failure along the length of the
column.

Fig.8. Failure patterns of beamcolumns (a) C1-S, (b) C1-W, (c) C2-S, and (d) C2-W after fire
tests.
9

III. Shell element model


The advantage to a shell model was that it can capture local buckling whereas the fiber models
cannot.
Quiel and Garlock (2010) performed an extensive computational thermal and structural analysis
using SAFIR (a software package specifically designed for the analysis of structures exposed to
fire) to predict the critical buckling stress of plates with various boundary conditions, loading and
temperature. They proposed their own predictive equations, since they found the American and
European codes to inadequately predict plate buckling, (see Fig.5).
Based on their proposed equations, Table.2 lists the values of the critical buckling stress (Fcr)
normalized by the yield stress (Fy) for seven different boundary conditions and loading cases for
the flanges and web of the W8x48 specimen heated to four different temperatures.
The real boundary conditions of the plates are somewhere between pinned (e.g., cases 1 and 4)
and fixed (e.g., cases 2 and 5). Values of Fcr/Fy equal to one means that the plate (flange or web)
will yield before it buckles.
The Table.2 indicates that except for Cases 1 and 4, plate yielding controls over plate buckling for
temperatures less than 500C. It was assumed that local buckling does not control the failure limit
state for temperatures less than 500C. At temperatures higher than 500C, the Fcr/Fy values start
to become less than one, which means that if the temperature in the plates exceed this value, it is
possible that local buckling will control the failure limit state.
To examine if local buckling would develop during the experiments, a shell element model of each
specimen was developed and analyzed. The shell element model can be directly compared to the
forensic fiber model since both use the same boundary conditions, loading, and thermal inputs.
This model used four-noded shell elements using square discretization per shell based on an
optimization study in SAFIR.
Initial local geometric imperfections were included in both the web and flanges in order to
realistically model the potential for local buckling. These imperfections were modeled as
sinusoidal with the same wavelength as the lowest buckling mode obtained from linear analysis of
the plates at ambient temperature. The maximum magnitude of the imperfection was modeled as
the plate width (d for the web and bf/2 for the flange) divided by 200.

10

Table.2-Fcr/Fy values (based on proposed predictive equations) for the W8x48 flange and web
plates with varying boundary conditions, loading conditions, and temperatures.
Case No.

Loading and
boundary
cinditions

T = 300C

T = 500C

T = 600C

T=700C

1.00

0.93

0.78

0.79

1.00

1.00

0.85

0.87

1.00

1.00

0.85

0.87

1.00

0.89

0.75

0.76

1.00

0.99

0.84

0.85

1.00

1.00

0.92

0.93

1.00

1.00

0.95

0.97

Stiffened
(Web)

Unstiffened
(Flange)

Fig.9. Failure pattern of beamcolumn C1-S: (a) experimental specimen and (b) shell model.
11

Shell element model was compared with fire test data reported by Dwaikata M.M.S. et al. (2010)
to examine the thermal, structural and plastic response.
Experimental specimen (C1-S) and shell element model has shown the same failure pattern with
significant flange deformation as shown in Fig.9. Recorded time of column failures for different
models was shown in Table.3

Table.3-Recorded times for column failure in minutes.


Specimen

Experiment

C1-S

Fiber-beam models

Shell model

Forensic

Priori

220

219

198

210

C1-W

215

212

197

196

C2-S

90

89

85

85

C2-W

93

87

80

82

2.3 Dwaikat Mahmud and Kodur Venkatesh (2010)


Dwaikat Mahmud and Kodur Venkatesh (2010) proposed a simplified approach for evaluating
plastic axial and moment capacity curves for beam columns with non-uniform thermal gradients
Restrained steel members, when exposed to fire develop significant forces and this transforms the
behavior of a beam (or column) into that of a beam-column. The load carrying capacity of such
beam-columns is determined through axial and moment capacity curves (P-M curves).
Codes and standards (European and American) recommend the use of uniform average
temperature for establishing the P-M curves at elevated temperatures. This assumption, though
adequate for cases where temperature in steel is uniform, such as a column exposed to fire from
four sides, may not be valid for columns or beams exposed to fire from 1, 2, or 3 sides since
significant thermal gradients develop across the section. These thermal gradients can cause severe
distortion in the P-M curves and render the capacity curves based on uniform temperature
inadequate for evaluating the strength of such beam-columns.
A simplified approach was proposed for adjusting the uniform temperature plastic P-M curves to
account for the shape distortion resulting from fire-induced thermal gradients. The proposed
method employs a two-step process in which the cross-sectional steel temperatures are calculated
12

first, and then the distorted P-M diagram is computed by adjusting the P-M diagrams based on a
uniform ``averaged'' temperature. The magnitude of P &M influences the load carrying capacity
of beam-columns which is generally expressed using the following relation at ambient conditions.

P
M
+b
1
Pn Mn

(1)

Where a and b are interaction coefficients, Mn and Pn are the moment and axial factored
capacities, respectively. At room temperature, the applied forces (P and M) and their
corresponding resistive capacities (Mn and Pn) are assumed to remain constant. But, under fire
conditions, the fire-induced restraint forces and the corresponding capacities vary as a function of
fire exposure characteristics and properties of the beam (or column).

2.3.1 Thermal analysis


The proposed approach is valid for standard fire exposure and is developed based on results from
detailed parametric study and nonlinear regression analysis. Steel temperature was obtained by
carrying out thermal analysis using ANSYS, finite element program. The results of the finite
element analysis are used in a regression analysis so as to arrive at a simple equation for predicting
temperature rise in steel as a function of fire exposure time.
The temperatures obtained via the finite element analysis were averaged for every time step of fire
exposure as follows: The average temperatures of each steel plate (top-cool and bottom-hot flanges
and web) are computed as the arithmetic mean of temperatures of three points per plate, as shown
in Fig.10. The computed mean temperatures of the cool (top) and hot (bottom) flanges and web
(TCF; THF, and Tw) are then used to compute the overall average sectional temperature (Ts) through
weighted average method as follows:

Ts =

AF TCF + Aw Tw + AF THF
2 AF + Aw

(2)

Where AF and Aw are the areas of the flange and web, respectively.
The averaged steel temperature Ts is then used in intensive nonlinear regression analysis so as to
arrive at a simple equation for predicting steel temperature under fire exposure.

13

Fig.10. Averaging the temperature in plates of steel beam.

2.3.2 Nonlinear regression analysis


The finite element model validated above is utilized to undertake a parametric study to generate
data for temperature profile developed in a steel section upon fire exposure. The factors that were
varied in this parametric study included sectional geometric and insulation properties.
The main assumption used in the development of the proposed expression was that steel
temperature (Ts) can be expressed as a product of a correlation function (Ct) and fire temperature
(Tf):
Ts = Ct Tf

(3)

Where Ct is dependent on fire exposure time t, section geometry and insulation properties. Section
geometry is expressed in terms of section factor (Ap/V), where Ap and V are the heated perimeter
and the volume of unit length of the section, respectively. The insulation characteristics are
expressed in terms of the (tp/kp) ratio, where tp and kp are the thickness and the thermal conductivity
of the applied insulation, respectively.
The shape of the correlation function (Ct) emerged from linearization of the logarithmic ratio (ln(1-Ts/Tf)). Therefore, it is assumed that the logarithmic ratio (-ln(1-Ts /Tf )) is a linear function
of a regression coefficient (s), namely: -ln(1-Ts/Tf )=s t. By rearranging, the following form of the
correlation coefficient can be obtained as Ct = (1-e-st). The temperature of steel should converge to
the temperature of fire after long time of exposure, therefore, Ct 1 as t

. Nonlinear

regression is then carried out to obtain an expression for the coefficient (s). The value of (s) is
14

computed for different combinations of (Ap/V) and (tp/kp) generated from parametric study. The
variation of (s) is a function of (tp/kp) for the range of (Ap/V) is in a rational form:

1
s=

Ap
V
Ap
V

+ 2
+ 2

tp
kp
tp
kp

+ 3
(4)
+ 3

The values of the regression coefficients (1; 2; 3; 1; 2; 3) are rounded off so that they fit in
a simple formula for protected and unprotected steel as follows:
-

For protected steel sections:


Ts = (1-e-st )Tf ,
1 =0.000013; 2=0; 3=0; 1=0; 2=1; 3=0.033

(5)

For unprotected steel sections:


Ts = (1-e-st )Tf,
1 =0.00033; 2=0; 3=0; 1=0; 2=0; 3=1

(6)

Fig.11. Comparison of steel temperature predicted by Eq. (5) with that obtained from finite
element analysis.
Fig.11 shows a comparison between finite element predictions of steel temperature and those
predicted using Eq. (5). The parameters used in this comparison were generated by multiplying
the values obtained from parametric study with random numbers ranging from 0.75 to 1.25. This
procedure was essential to ensure the independency between the finite element analysis results
used in the regression analysis and those used in the comparison shown in Fig.11. The comparison
shows that Eq. (5) can reasonably predict steel temperatures below 750C; while steel temperatures
15

are overestimated by Eq. (5) for Ts > 750C. This overestimation was because of the phase change
that occurs in steel at temperatures between 750C and 800C, which is not captured by Eq. (5).
Any attempt to make adjustments to Eq. (5) in the range Ts > 750C will result in a complex
equation for predicting steel temperature. Thus Eq. (5) was deemed to be acceptable for generating
steel temperature conservatively below 750C and this is the range of temperatures encountered in
the failure of beam-columns in most practical situations.
The developed simple equation for predicting temperature distribution in steel sections exposed to
fire, is used to modify the P-M diagrams to account for thermal gradients induced by fire exposure.
The P-M diagrams will be established for sections with linear thermal gradients, and then
adjustment will be made so as to be able to use the P-M diagrams based on uniform average steel
temperature.

2.3.3 P-M Interaction development


In the generation of the P-M diagrams, the main assumption was that the beam-columns are braced
against lateral or torsional buckling. Further, the influence of local buckling is neglected since the
beam-column was assumed to be made of a compact section, as required in a design situation.
Stability factors (e.g. KL for columns) can be achieved independently from the P-M diagrams.
This can be done through adjusting both axial and bending capacities (using Pcr and Mcr instead of
plastic capacities FyAs, and FyZx, respectively), and by amplifying applied moment due to lateral
bowing of the member. Pcr and Mcr, under the influence of thermal gradient, can be computed
based on Eurocode recommendations, or the upcoming AISC specifications manual. In terms of
global stability, Eurocode recommends an effective length of 0.5Lo for columns in intermediate
storey and 0.7Lo for columns in top storey for fire resistance design. To generate the plastic beamcolumn capacity for a section with thermal gradient, each fiber of steel section is subjected to full
ultimate stress that corresponds to the temperature in that fiber, i.e. Fu ky(Ti). Fig.12 shows a wide
flanged (WF) section with thermal gradient and the corresponding ultimate stress Fu ky(Ti) at each
fiber of temperature Ti. As shown in Fig.12, the nonlinear thermal gradient is approximated to a
linear gradient.
In order to generate capacity P-M envelops for WF sections with thermal gradients, few
assumptions were made in order to simplify the resulting expressions for computing the capacity

16

envelopes of beam-columns. First, the beam was assumed to be exposed to fire from three sides,
and the resulting thermal gradient was approximated as a linear gradient.

Fig.12. Temperature and ultimate stress distributions for strong and weak axis orientation in
a WF section.
The temperature in fire exposed WF section is calculated using the proposed simplified equation
(Eq. (5)). In the case of standard fire exposure, the average temperatures of steel section and in
cool (top) flange (Ts;Ave, and Ts;CF ) are computed using the sectional factors for the section (Ap/V)
and for the top flange (Ap/V) TF , respectively:

Ts;Ave = -(1-e-st)Tf,

0.013 A p

1000 V
s=
tp

+ 0.033

kp

Ts;CF = -(1-e-sTFt)Tf, sTF

(7)

0.013 Ap


1000 V CF Ap
2 t F + BF t w
=
,
=

tp

t F BF
V CF
+ 0.033
kp

(8)

For any case of steel temperature distribution, the ultimate axial force and bending moment are
computed by direct numerical integration over the cross-sectional depth (d) using the following
algorithms:
P = Fu dA = sign ( yi YNA ) Fu k y (Ti ) beff yi

M GC = yFy dA = sign( yi YNA ) ( yi YGC ) Fu k y (Ti ) beff yi


17

(9)
(10)

Where the function ``sign(x)'' is the function that gives the sign of a real number (+1 for x > 0 and
-1 for x < 0) and the positive sign indicates compression. The terms in Eqs. (8) and (9) are
illustrated in Fig.12. The bending moment MGC is calculated around the center of geometry (YGC)
of the section, and the positive sign indicates tension at the top (cooler) fiber. The center of
geometry is defined as the geometrical centroid of the section, while the neutral axis (YNA) is
defined as the axis at which the strain is zero. To obtain the entire P-M diagram for the same
temperature profile of the section, the location of the neutral axis (YNA) is varied along the depth
of the section. For each location of the neutral axis, a pair of P and M values is computed using
Eqs. (8) and (9), respectively. The location of the neutral axis is moved from zero to d in
increments of yi= 0:001d. Fig.13 shows the degradation in strength and stiffness of steel as a
function of temperature that was adopted in the calculation of P-M diagrams. These strengthtemperature relationships were adopted from the ASCE manual of practice. The distortion in the
shape of the normalized P-M diagrams does not depend on the relative sizes of the sections;
however, the shape critically depends on the axis orientation (i.e. on the direction of thermal
gradient). The distortion in the P-M diagrams is more severe for the case of thermal gradient in
strong direction as compared to the case of weak direction. Therefore, it was proposed here to
adjust the P-M diagrams for the case of severe distortion due to thermal gradient in the strong
direction, while the shape of P-M diagrams will be maintained for the case of weak direction. The
following subsections will present treatment of P-M diagrams for both strong and weak axis
orientation.

Fig.13. Variation of Es and Fy as a function of steel temperature.


18

2.3.4 Strong axis orientation


The basic features of the distorted plastic P-M diagrams for a WF section with thermal gradient in
the strong direction are compared to the case of a uniform temperature in Fig.14. The figure shows
that the value of moment capacity under peak axial capacity (point A in Fig.14) moves back and
forth (to point A1 in Fig.14) depending on the eccentricity (e) between center of stiffness ``CS''
and center of geometry ``GC'' that is caused by the thermal gradient in a WF section.

Fig.14. Characterizing P-M diagram for a WF section with thermal gradient in the strong
direction.
The following assumptions were made in order to obtain a simple method for deriving the distorted
1. The P-M diagram represents a parallelogram with its sides being parallel to the P-M segments
derived based on the average temperature of the section. Specifically, it was assumed that the
segments AB and AC were parallel to the segments A1B1and A1C1 respectively, as shown in
Fig.14.
2. The magnitude of the shift (MTG) in the P-M capacity envelope (Fig.14) is assumed to be
numerically equal to the ultimate axial capacity (Pu;Tave) of the section multiplied by the
eccentricity (e) between the center of geometry (GC) and of the center of stiffness (CS) of the
section as shown in Fig.15. The ultimate capacity is computed based on the average temperature
of the section, i.e.:

MTG = e Pu,Tave = e ky (Ts, Ave ) Fy As


19

(11)

Fig.15. Eccentricity between center of stiffness and center of geometry of a WF section with
thermal gradient.
To compute the eccentricity (e) between the YCS and YGC, the reduction in the elastic modulus of
steel was assumed to vary linearly across the depth of the section as shown in Fig.15. Each plate
of the section was assumed to have a constant rate of reduction (kE) in the elastic modulus
depending on its average temperature. The reduction in elastic modulus of the steel in the web was
assumed to equal the average of the reductions of both the top (cool) and bottom (hot) flanges.
With these assumptions, the eccentricity (e) between YGC and YCS can then be calculated as
follows:
e = YCS YGC =

A k (T ) y
A k (T )
i

d
2

d 2 B t k (T ) + t w dk E (Ts , Ave )
e = F F E s ,CF
1

2
k E (Ts , Ave )(2 BF t F + t w d )

(12)

Where BF ; tF ; tw; d are dimensions of the section as shown in Fig.15, and kE(T) is the reduction
factor for elastic modulus at steel temperature T .
The modified equations of the plastic P-M diagrams for wide flange section with linearized thermal
gradient in the strong direction was be written as:

M + M TG
Zx Fu,Ts,ave
M M TG
Zx Fu,Ts,ave

P
1.0
As Fu,Ts,ave

1.0
As Fu,Ts,ave

20

and

(13)

2.3.5 Weak axis orientation


In the case of thermal gradient along the weak axis, the distortion in plastic P-M diagrams is
smaller as compared to the case of thermal gradient along strong axis. As shown in Fig.16, the PM diagram for WF sections with thermal gradients in the weak direction constructed by assuming
the moment capacity ratio (M/Mu) to increase linearly as the axial capacity ratio (P/Pu) decreases

Fig.16. Linearized P-M diagram for a WF section with thermal gradient in weak direction.

until the moment capacity ratio (M/Mu) reaches its maximum value of 1.0 at P/Pu=0.5, after which
the moment capacity remains constant. Based on this assumption, the P-M diagrams for WF
sections subjected to thermal gradients in their weak direction can be expressed as:

If

P
1

A s Fu,Ts,ave
2
1 M + M TG
P
+
1.0
2 Z y Fu,Ts,ave
A s Fu,Ts,ave

(14)

If

P
1
,
A s Fu,Ts,ave 2
M + M TG
1.0
Z y Fu,Ts,ave
21

CHAPTER 3
REINFORCED CONCRETE BEAM-COLUMNS SUBJECTED TO
FIRE
So far, chapter.2 is discussed about the steel beam-columns subjected to fire, but in India most of
the constructions are reinforced concrete structures. Concrete is specified in buildings and civil
engineering projects for several reasons, sometimes cost, and sometimes speed of construction or
architectural appearance, but one of concretes major inherent benefits is its performance in fire,
which may be overlooked in the race to consider all the factors affecting design decisions.
Concrete usually performs well in building fires. The low thermal conductivity of the concrete
associated with its great capacity of thermal insulation of the steel bars is responsible for this good
behavior. However, when its subjected to prolonged fire exposure or unusually high temperatures,
concrete can suffer significant distress. Because concretes pre-fire compressive strength often
exceeds design requirements, a modest strength reduction can be tolerated. But large temperatures
can reduce the compressive strength of concrete so much that the material retains no useful
structural strength.
A phenomenon such as the concrete spalling may compromise the fire behavior of the elements.
Spalling of concrete leads to a decrease in the cross section area of the concrete column and thereby
decrease the resistances to axial loads and moment carrying capacities, as well as the reinforcement
steel bars become exposed directly to high temperatures, this makes the structural element to act
as beam-column.
A study of RC columns is important because these are primary load bearing members, and a
column could be crucial for the stability of the entire structure.

3.1 Raut Nikhil and Kodur Venkatesh (2011)


Raut Nikhil and Kodur Venkatesh (2011) given an approach for modeling the fire response of
RC columns under biaxial bending. This approach accounts for high-temperature material
properties, geometric and material nonlinearity, fire-induced spalling, and restraint effects and can
be applied under realistic fire and loading scenarios. The validity of the approach is established by
comparing the predictions from the model with results from full-scale fire resistance tests.
Reinforced concrete (RC) columns, when exposed to fire, are often subjected to biaxial bending
22

arising from eccentricity in loading or one-, two-, and three-sided exposure or due to non- uniform
spalling.
The buildup of such biaxial bending effects in an RC column was illustrated in Fig.17, where a
structural frame in a building was exposed to fire. Fig.17 (b) through (d) illustrates the
development of thermal gradients in center (four-sided), peripheral (three-sided), and corner
column (two adjacent side) fire exposure, respectively, under fire conditions. It can be seen that
the occurrence of uniaxial bending (peripheral column) and biaxial bending (corner column) can
be a common occurrence in most practical scenarios in buildings.
The current fire resistance provisions in codes (i.e., Eurocode & American code) and standards are
mostly based on standard fire tests with all four sides exposed to fire; hence, they may not be fully
applicable to one-, two-, or three-sided fire exposure.
The presence of uniaxial bending induces an eccentricity in an originally axially loaded column
due to the shift in the neutral axis, resulting from a degradation of the strength and stiffness
properties of concrete and reinforcing steel. Thus, the column is subjected to an additional moment
along with the axial load; this leads to a reduction in the fire resistance of the column.
In the case of corner columns (biaxial bending), the neutral axis shifts and also rotates, thus
inducing eccentricity along both the axes. This causes the column to experience additional
moments along both the axes. The induced eccentricity in both cases also increases lateral and
axial deformations due to P- effects.
In addition to fire exposure, uneven fire-induced spalling can also produce significant uniaxial or
biaxial bending on the column. Such fire-induced spalling may be more prevalent in columns made
of high-strength concrete (HSC). Spalling is a phenomenon in which chunks of concrete fall off
from the surface of a concrete structure when exposed to high and rapidly rising temperatures. The
spalling in concrete is primarily dependent on the rate of temperature rise (fire scenario),
permeability (typically related to strength and presence of fines such as silica fume), fire exposure,
and other conditions (such as geometry and aggregate type). The development of differential
thermal stresses can also significantly influence spalling in concrete. Spalling occurs when the
pore pressure in a concrete member, due to evaporated moisture, exceeds the tensile strength of
concrete. This phenomenon is schematically illustrated in Fig.17 (e), where the pore pressure
development and loss of tensile strength is plotted as a function of temperature.

23

Fig.17. Factors influencing behavior of RC column under fire.

The effect of such biaxial bending was taken into consideration in evaluating the fire resistance of
RC columns in this paper, the model was applied to undertake a set of parametric studies and the
observations shown that the biaxial bending effect arising from one-sided, two adjacent-sided, and
three-sided fire exposure significantly decreases the fire resistance of RC columns.

3.1.1 Numerical model for predicting fire response


Numerical model is based on a macroscopic finite element approach has been developed for tracing
the fire response of RC columns. Numerical model has been extended to include the effect of
biaxial bending resulting from the thermal gradients, exposure conditions, loading, or spalling.
The macroscopic finite element model (FEM) uses a series of moment-curvature relationships for
tracing the response of the column in the entire range of behavior from a linear elastic stage to the
24

collapse stage under any given fire and loading scenario. The model incorporates a hydrothermal
spalling sub-model to predict the occurrence of fire-induced spalling in HSC columns.
In the macroscopic FEM, an RC structural member was divided into a number of segments (Fig.18
(b)) along its length and the midsection of the segment was assumed to represent the behavior of
the whole segment. This midsection was divided into elements, as shown in Fig.18 (d). The fire
resistance analysis was carried out by incrementing time steps (refer to Fig.18 (e)). At each time
step, the model performs analysis through three main steps, namely:
1. Establishing fire temperature due to fire exposure;
2. Carrying out heat transfer analysis to determine temperature distribution across each segment.
The temperature data are used to evaluate the pore pressure in concrete, which is then used to
determine the extent of spalling within the cross section; and

Fig.18. Discretization of column and flowchart of numerical model


25

3. Performing the strength analysis through the following three sub-steps: (a) calculating the fireinduced axial restraint force in the RC structural member; (b) generating the M- relationship
(using the axial force computed previously) for each segment; and (c) performing structural
analysis of the overall member to compute lateral and axial deformations and internal forces.
The fire temperatures were calculated by assuming that the column was exposed to a fire whose
temperature follows that of the standard fire exposure or any other design fire scenario. For design
fires, the time-temperature relationship specified in Eurocode-1 (2002) for typical fuel load and
ventilation factors was built into the model. The thermal analysis was carried out for each segment
in the column. The temperature was assumed to be uniform along the length of the segment and
thus the calculations were performed for a unit length of each segment. The governing equation
and boundary condition for heat transfer analysis was taken as
k 2T + Q = c

T
t

T
T
k
ny +
nz = ht ( T T )
y
y

(15)
(16)

Where k is thermal conductivity; cis heat capacity; T is temperature; t is time; Q is heat source
defined as the amount of heat produced for a unit volume of the material; ht is the heat transfer
coefficient; T is fire or ambient temperature depending on the type of exposure; and ny and nz are
components of the vector normal to the boundary in the plane of the cross section.

M T + K T = F (t )

(17)

Where,

N N T
N N T
T
K = k
+k
dA + N N ds
x x
y y
A

M = cNN T dA
A

F = NQdA + N T ds

Where N is the vector of shape functions; and is the heat transfer coefficient depending on the
boundary.
The finite element analysis is applied to solve Eq. (17) and compute the temperature distribution
within the cross section of each segment. As part of strength analysis, M- relationships are
26

established as a function of time for all the segments and they are, in turn, used to trace the response
of the column under fire conditions.
Using generated M- relationships, the flexural rigidity of the segment at a particular time step is
evaluated using the segmental stiffness approach and was used to calculate strength, axial, and
lateral deformations of the column through the stiffness approach. As part of this segmental
stiffness matrix, loading and displacement vectors were set up for each longitudinal segment.
These segmental matrixes (vector) were assembled in the form of a nonlinear global stiffness
equation as

( K

+ K geo { } = {Pf } ;

K g { } = { Pf } + {Ps } ;
{Ps } = K geo { }

(18)

Where Kg is the global stiffness matrix (computed from the M- relationship); Kgeo is the geometric
stiffness matrix; is the nodal displacements; Pf is the equivalent nodal load vector due to applied
loading; and Ps is the equivalent nodal load vector due to the P- effect.
At every time step, each segment of the column was checked for failure under thermal and strength
failure limit states. In the strength limit state, the failure of an RC column was said to occur when
the applied axial load (or moments) exceeds the load- (or moment-) carrying capacity of the
column. The column was said to have failed in crushing if the curvature of the column exceeds the
maximum curvature in the moment-curvature relationship (i.e., corresponding to the ultimate
strain in concrete) at that time step, or if the global stiffness matrix (elastic stiffness matrixgeometric stiffness matrix) becomes negative or singular, the column is said to have failed in
buckling.

3.1.2 Interpretation of biaxial bending


The aforementioned procedure was applied for fire resistance analysis under axial loading
conditions. The effect of one-, two-, or three-sided fire exposure is incorporated through proper
thermal boundary conditions in Eq. (18); thus, the relevant temperature profile across the column
cross section was generated.

27

Fig.19. Independent deformation of column segment.

To account for biaxial bending arising from uneven spalling or loading or one-, two-, or threesided fire exposure, the 2-D stiffness matrix in Eq. (19) was replaced with a 3-D stiffness matrix.
The 2-D stiffness matrix Kg in Eq. (20) accounts for bending and lateral deformation in one plane
only, but in the case of biaxial bending the column bends in both directions, therefore the bending
and lateral deformations along the other direction also need to be considered. Thus, the 6 x 6
stiffness matrix was replaced by a 10 x 10 stiffness matrix (two additional degrees of freedom per
node). The five different degrees of freedom to be considered in a beam-column segment at each
node are shown in Fig.19. The segmental stiffness matrix [K(1010)] is computed considering
axial and bending deformations separately. The sectional properties were assumed to be constant
within the segment at a given time step. Thus, elastic analysis can be performed. The segmental
matrix was derived by separately solving the axial and flexural components using the forcedisplacement relations. The flexural components in the y- and z-directions and the three force
components in the x-, y-, and z-directions were given by

P = EAux' ; my = EI yu"z ; mz = EI z u"y

(19)

Where P is the axial force acting on the column; my and mz are the moments about the y- and zaxis; EA is the axial rigidity of the segment; and EIy and EIz are the flexural rigidity of the segment
about the y- and z-axes.

28

The final stiffness matrix for each segment can be written as


(5 5)

[ K (10 10) ] = K AA (5 5)

BA

K AB (5 5)

K BB (5 5)

(20)

In Eq. (19) the stiffness matrices KAA, KBB, and KAB are expressed as
EI y ky3S y

2 2Cy ky LS y

0
[ KAA ] = [ KBB ] =

2
EI y ky (1 Cy )
2 2C k LS
y
y
y

EI y k y3 S y

2 2Cy k y LS y

0
[ K AB ] = [ KBA ] =

2
EI y k y (1 C y )
2 2C k LS
y
y
y

EI z kz3Sz
2 2Cz kz LSz

EA
L

EI z kz2 (1 Cz )
2 2Cz kz LSz

EI z kz (Sz kz LCz )
2 2Cz kz LSz

EI z k z3 Sz
2 2Cz kz LS z

EI z kz3 (1 Cz )
2 2Cz kz LSz

EA

EI z k z2 (1 Cz )
2 2Cz k z LSz

EI z k z (S z kz LCz )
2 2Cz kz LSz

EI z kz3 (1 Cz )
2 2Cz kz LSz

EI y ky2 (1 Cy )

2 2Cy ky LS y

EI y ky (S y K y LCy )
2 2Cy ky LS y

EI y k y2 (1 C y )

2 2C y k y LS y

EI y k y (S y K y LC y )
2 2Cy k y LS y

(21)

(22)

Where Cy=cos(kyL); Sy=sin(kyL); Cz=cos(kzL) and Sz=sin(kzL)


ky =

P
; kz =
EI y

P
EI z

3.1.3 Validation of model


The macroscopic FEM models was validated by comparing predictions from the analysis with two
columns, normal strength concrete column (NSC) and HSC column.
The response of the columns was evaluated using the previously described model and fire
resistance was computed based on the strength failure criterion. The columns were discretized, as
shown in Fig.18 (b) and (d) and loaded eccentrically.
29

It was observed that model predictions were close to measured axial deformations throughout the
fire exposure time for both NSC and HSC columns, and The fire resistance of the HSC column is
higher than that of the NSC column, where the HSC column is of a larger cross-sectional area and
is made of carbonate aggregate, which helped in achieving higher fire resistance. Furthermore, the
bent ties (135 degrees) in the columns helped to minimize the spalling beyond the concrete core.
The results from the analysis were applied to determine the fire resistance of the columns based
on the strength failure criteria.
The fire resistance of NSC column from the macroscopic FEM based on strength failure criteria
was 208 minutes, which is close to the measured fire resistance in the test (214 minutes). The
predicted failure was through buckling and this is in agreement with the observed failure pattern.
Similarly, for HSC column, the predicted and measured fire resistance was 123 and 118 minutes,
respectively, based on the strength criteria. The model predicted HSC column HSC to fail through
buckling and the test reported a combined crushing and buckling failure.
So, it should be noted that the model cannot predict combined failure modes because whichever
failure limit state is reached first is taken as failure.
Comparing the behavior of the predicted axial deformation for HSC column to the model
predictions for the same column but with a concentric load, the fire resistance increased from 208
to 360 minutes. Similarly for HSC column under a concentric load, the fire resistance increased to
346 minutes from 118 minutes under eccentricity. Thus, eccentricity has a significant influence on
fire resistance and needs to be accounted for in the analysis.

30

REFERENCES
1. AISC. Steel construction manual. 13th ed. Chicago: American Institute of Steel
Construction; 2005.
2. CEN. Eurocode 3: design of steel structures, Part 1.2: general rules structural fire design
(ENV 1993-1-2). Brussels: European Committee for Standardization (CEN); 2001.
3. Dwaikat Mahmud; Kodur Venkatesh. A simplified approach for evaluating plastic axial
and moment capacity curves for beam-columns with nonuniform thermal gradients. Eng
Struct. 2010;32(5):1423-1436.
4. Dwaikat MMS, Kodur VKR, Quiel SE, Garlock MEM. Experimental behavior of steel
beamcolumns subjected to fire-induced thermal gradients. J Constr Steel Res
2010;67(1):308.
5. EN 1991-1-2, Eurocode 1: Actions on Structures. Part 1-2: General ActionsActions on
Structures Exposed to Fire, European Committee for Standardization, Brussels, Belgium,
2002, pp. 1-59.
6. Quiel SE, Garlock MEM. Calculating the buckling strength of steel plates exposed to fire.
Thin-Wall Struct 2010;48:68495.
7. Quiel SE; Garlock M.E.M.; Dwaikat M.M.S.; Kodur V.K.R.. Predicting the demand and
plastic capacity of axially loaded steel beam-columns with thermal gradients. Eng
Struct.2014;58:49-62.
8. Quiel SE, Garlock MEM, Paya-Zaforteza. Closed-form procedure for predicting the
capacity and demand of steel beamcolumns under fire. J Struct Eng ASCE
2011;137(9):96776.
9. Raut Nikhil; Kodur Venkatesh. Response of reinforced concrete columns under fireinduced biaxial bending ACI Structural Journal. 2011;108(5):610-619.
10. Takagi J, Deierlein GG. Strength design criteria for steel members at elevated
temperatures. J Constr Steel Res 2007;63:103650.

31

Вам также может понравиться