Top Banner
11

Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

Apr 19, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

Steel Structures 7 (2007) 227-237 www.kssc.or.kr

Behavior of Steel-framed Buildings in a Fire

Khalifa S. Al-Jabri*

Associate Professor, Department of Civil and Architectural Engineering, College of Engineering,

P.O. Box 33, Sultan Qaboos University, Al-Khod, PC 123, Oman

Abstract

The rules presented in the current codes of practice for the design of steel-framed buildings in fire conditions are based onresults of isolated member tests carried out in a laboratory that followed a prescribed standard fire curve. The laboratory testingconditions were obviously different from the real situation wherein the structure is subjected to natural fires. It has been knownfor many years, from observations of accidental fires, that structural members behave better in fire when they constitute partof a structural arrangement than when they are tested in isolation. These observations have been confirmed by results fromexperimental fire tests conducted on full-scale multi-story steel-framed buildings. It has been demonstrated that members thatform part of the structure can withstand much higher temperatures than those tested singly. This has raised doubts concerningthe conservative design approaches provided by current fire engineering design codes. The attack on the twin towers of theWorld Trade Centre in New York on 11 September 2001 has prompted close examination of the way in which buildings canfail in fires and has brought into the public eye the hazards that fires can pose to major building structures. This paper examinesthe effects of fires on multi-story steel-framed buildings that provide vital knowledge of the behavior of real buildings. Thismay allow for the construction of safer buildings in the future.

Keywords: Steel-framed building, composite floor, fire, Cardington frame, tensile membrane action

1. Introduction

All common building materials lose strength when

heated to high temperatures. Figs. 1 and 2 show the

deterioration of steel and concrete mechanical properties

with increasing temperatures. Although steel does not

melt below 1,500oC, at a temperature of around 600oC, its

yield strength declines to about one-third of its yield

strength at ambient temperature. At 800oC, its yield

strength is reduced to 11%, and at 900oC, to 6%. The

elastic modulus of steel is similarly reduced with increasing

temperatures, but at a higher rate. Due to internal

cracking and chemical changes, concrete also loses

strength and stiffness as temperature increases, as seen in

Fig. 2. Since concrete has much lower thermal conductivity

than steel, a concrete encasement is often used as a fire

protection for steel. The degradation of structural materials’

stiffness and strength at high temperatures may, in some

incidents, cause the structure to collapse under severe fire

conditions. Materials such as steel obviously need to be

designed to withstand the effects of a fire in order to

ensure the safety of people and property. The design must

ensure no additional threat, to either escaping occupants

or fire fighters, of the possible collapse of the steel

structure.

Using applied fire protection remains the most common

way of satisfying structural fire resistance requirements.

Typical types of protection materials consist of boards,

sprays, and intumescent paints, with the choice of material

depending on cost, appearance, durability, and the required

architectural features. The required thickness is based on

the principle of ensuring that the steel remains below

550oC for the specified fire resistance period without the

*Corresponding authorTel: +968-2414-1333; Fax: +968-2441-3416E-mail: [email protected]

Figure 1. Degradation of the mechanical properties ofsteel at elevated temperatures.

Page 2: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

228 Khalifa S. Al-Jabri

necessary consideration of the type of structural steel

element. This assumes that fully stressed steel members

will fail when they reach a temperature of approximately

550oC. This temperature is known as the “limiting

temperature,” according to the British design code

(BS5950: Part 8, 1990). Although the philosophy behind

specifying the protection thickness is adequate, it is not

necessarily the only solution, and it can be extremely

conservative in misestimating the structural capability,

since it ignores the inherent fire resistance of the overall

stiffness of the structure.

From an engineering point of view, however, it is

probably more economical to design a structure so that it

will withstand fire without protection rather than to

design it for normal conditions and then apply protection.

To address this issue, there has been much interest in

understanding the behavior of different structural members

in fire, either in isolation or as part of a more complete

structure. In the beginning, research was conducted on

the actual behavior of isolated bare-steel beams and

columns, which resulted in the first ever fire design code,

BS 5950: Part 8 (1990). The fire occurrence is treated as

an accidental limit state, with its own associated loads

and materials. Despite current design codes providing a

more scientific basis for the provision of fire resistance to

steel-framed structures, they are based on isolated

member tests that did not take into consideration the

interaction between members (BS5950: Part 8, 1990;

EC3: Part 1.2, 1993; and EC4: Part 4, 1994).

Observations from real fires (Newman, 1991; Burgess,

2002; and Burgess, 2005) and recent fire tests on the

Cardington full-scale test frame (Moore and Lennon,

1997) have demonstrated the great importance of structural

continuity and interaction between structural members in

enhancing the performance of steel-framed buildings in

fires. Structural continuity may not have a significant

effect on the behavior of structural members at ambient

conditions, but in a fire, it can play an important role in

enhancing the survival time of the structure. Many

aspects of structural behavior occur due to the interaction

between members, and cannot be predicted or observed

from isolated tests. Some of these features include local

buckling in the vicinity of the joints, deformations due to

adjacent structures’ restraint of thermal expansion, the

redistribution of internal forces, the behavior of the

structures during the cooling phase, and the tensile

membrane action of the composite floor at large

displacements. Furthermore, unlike the standard fire

curve wherein the fire regime follows a pre-defined

temperature-time relationship, a natural fire is characterized

by three phases: a growth phase, a fully developed phase,

and a decay phase (Fig. 3). It is necessary to evaluate not

only the effect on the structural resistance during the

heating phase, but also the high cooling strains in the

beams and adjacent joints induced by the distortional

deformation of the heated elements during the decay

phase (Wald et al., 2004).

In view of the above, it is apparent that current fire

engineering design codes (BS5950: Part 8, 1990; EC3:

Part 1.2, 1993; and EC4: Part 4, 1994) do not address the

actual behavior of buildings in fires, since buildings do

not act as a series of individual members. In reality,

structures have significant reserve strength to withstand

temperatures much higher than those described as

limiting temperatures in the design codes. As a result,

concepts have started to emerge that can lead to the

development of more rational and economical design

methods for steel-framed structures in fires. This may

eventually reduce or even eliminate reliance on

traditional fire protection methods, and may allow

construction of safer, more efficient, and more effective

fire-resistant buildings in the future.

Figure 2. Degradation of the mechanical properties ofconcrete at elevated temperatures.

Figure 3. Natural fire and ISO834 standard fire.

Page 3: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

Behavior of Steel-framed Buildings in a Fire 229

2. Review of Current Fire Engineering Design Methods for Steel-framed Structures

Existing codes (BS5950: Part 8, 1990; EC3: Part 1.2,

1993; and EC4: Part 4, 1994) for the design of steel-

framed structures in fires have introduced design methods

that treat fire as one of the basic limit states. These design

codes were developed, however, mainly from standard

fire tests on isolated columns and beams. In these tests,

the columns were 3.0 m high and the beams were 4.5 m

long, and the failure criteria were governed largely by the

size of the furnace. A beam element was assumed to have

failed when the deflection exceeded span/30, whereas

column failure was deemed to have occurred when the

column could no longer carry the applied load. Based on

the results of such tests, design methods such as the

limiting temperature method and the moment resistance

method were proposed.

2.1. The limiting temperature method

BS 5950: Part 8 (1990) uses the concept of the load

ratio as a measure of the applied load that a member can

resist at the time of a fire. Because the limiting temperature

is virtually independent of the heating time, it depends on

the load that the member carries. The load ratio is defined

as the ratio between the applied load or moment at time

of a fire and the load or moment of resistance at 20ºC.

The load ratio is a useful concept because it allows

different-sized elements to be considered in the same

way. For instance, a 200 mm-deep beam will fail at

approximately the same temperature as a 400 mm-deep

beam if they are both working at the same load ratio. In

practical designs, the load ratio will vary from 0.45 to

0.55. Load ratios much higher than 0.6 are very rare,

although the maximum value can be as high as 0.7 for an

element that carries purely the dead weight of the

structure. Using the concept of the load ratio, it is

possible to design some members that can withstand a

fire without fire protection, and to use a lesser degree of

fire protection for other members. A member that fails at

550oC will require more protection than if the same

member fails at 700oC.

After the load ratio is calculated, the limiting temperatures

(or the maximum allowable temperatures) can be

obtained from Table 5 in BS5950: Part 8. This Table is

reproduced in Table 1. To allow for the use of unprotected

steel, the limiting temperatures are compared with the

design temperatures that are tabulated in BS5950: Part 8

and that correspond to the expected maximum temperature

of the section for a given period of fire resistance (i.e., 30

and 60 minutes). If the limiting temperature is less than

the design temperature, the beam does not require fire

protection. Despite the simplicity of the load ratio

method, it is very conservative, because it does not take

advantage of the full capacity of the member.

2.2. The moment resistance method

The moment resistance method is based on the plastic

design of a structural element, as shown in Fig. 4. If the

temperature distribution across a beam is known, the

moment resistance can be calculated. This is a powerful

but simple tool that until recently was used only by

researchers. The cross-section is divided into groups of

elements of approximately equal temperatures. The

reduced strength of various elements of the cross-section

can be calculated at elevated temperatures using the

strength reduction factors given in Table 1 in BS5950:

Part 8. The plastic neutral axis of the cross-section is

determined and the moment of resistance (or the moment

capacity) of the section is determined by multiplying the

reduced strength of each element by the distance from the

neutral axis and summing up all the elements in the

section. The applied moment is calculated using the fire

limit state load factor. If the moment of resistance

(capacity) is greater than the applied moment, the section

is adequate and protection is not required.

Table 1. Limiting temperatures for the design of members (From BS5950: Part 8)

Description of member Limiting temperature (oC) at a load ratio of:

0.7 0.6 0.5 0.4 0.3 0.2

Members in compression

λ 70 510 540 580 615 655 710

λ > 70 but λ ≤ 180 460 510 545 590 635 635

Members in bending supporting a concrete or composite slab

a) Unprotected beams or beams protected with ductile protection 590 620 650 680 725 780

b) Other protected beams 540 585 625 655 700 745

Members in bending not supporting concrete

a) Unprotected beams or beams protected with ductile protection 520 555 585 620 660 715

b) Other protected beams 460 510 545 590 635 690

Members in tension 460 510 545 590 635 690

λ is the slenderness, i.e., the effective length divided by the radius of gyration.

Page 4: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

230 Khalifa S. Al-Jabri

3. Full-scale Experimental Tests on Steel-framed Structures in a Fire

On the 23rd of June 1990, a fire broke out in the

partially completed 14-story building in the Broadgate

development (Anon, 1991). The fire began at a large

contractors’ hut on the first floor, and smoke spread

undetected throughout the building. The fire detection

and sprinkler system was not yet operational after

working hours. The fire lasted for 4.5 hours, including 2

hours when the fire exceeded 1,000oC. The direct losses

from the fire exceeded £25 million, but only a fraction of

such loss (£2 million) represented the structural frame

and floor damage. The major damage was to the building

fabric as a result of the smoke. Moreover, the structural

repairs after the fire took only 30 days. The structure of

the building was a steel frame with composite steel deck

concrete floors, and it was only partially protected at this

stage of the building’s construction. During and after the

fire, despite significant deflections in the elements exposed

to the fire, the structure behaved well, and none of the

columns, beams, and floors collapsed.

The Broadgate phase 8 fire provided the first opportunity

to examine the influence of a fire on the structural

behavior of a modern fast-track steel-framed building

with composite construction. Prompted by the evidence

from Broadgate, the Building Research Establishment

(BRE) built an 8-story composite steel and lightweight

concrete frame at their large-scale test facility at

Cardington. The frame was subjected to six full-scale fire

tests [2 by BRE and 4 by British Steel (now CORUS)] in

1995 and 1996, which allowed observation and recording

of the behavior of the structure during a fire.

These tests were commonly known as Cardington frame

fire tests. The structure was designed as a composite

building typical of contemporary medium-rise commercial

buildings in the United Kingdom, with roof-mounted

services. The plan dimensions were 45 × 21 m, which

provided a footprint area of 945 m2. The general frame

arrangement is shown in Fig. 5. These tests included: (1)

a restrained beam test, (2) a plane frame test, (3) a first

corner test, (4) a second corner test, (5) a large compartment

test, and (6) a demonstration test. A brief summary of

these tests is given below, whereas and a detailed

description is presented elsewhere (Armer and Moore,

1994; Lennon, 1997; O’Connor and Martin, 1998; and

Bailey et al., 1999).

The restrained beam test comprised a heated 8 m × 3 m

area on the seventh floor. A gas furnace was used to heat

the 9 m 305 × 165UB40 secondary test over the central 8

m of its length.

The plane frame test was conducted using a gas-fired

furnace to heat up a 21 m × 2.5 m area on the fourth floor,

and included three connected spans of primary beams.

The first corner test was carried out in a 10 m × 7 m

area on the second floor. The fire was generated by firing

wooden cribs with densities of 45 kg/m2. The internal

beams were left unprotected whereas the columns and

edge beams were protected. Lightweight concrete block

walls were used in the construction of a compartment.

The second corner test was conducted in a compartment

built from fire-resistant board. In this test, a 9 m × 6 m

area on the third floor was heated by burning wooden

cribs with densities of 45 kg/m2. The columns were fully

protected whereas the internal and edge beams were all

left unprotected.

The large compartment test was carried out in a

Figure 4. Stress distribution analysis at the fire limit state (BS 5950: Part 8, 1990).

Figure 5. General beam framing arrangement of theCardington test frame.

Page 5: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

Behavior of Steel-framed Buildings in a Fire 231

compartment built with a fire-resistant wall across the

21 m width of the building, with fire-resistant board used

to block off the entrance to the passenger hoist. As in the

two previous tests, the fire in the compartment was

generated by burning wooden cribs with densities of 45

kg/m2 in a 21 m × 18 m area on the third floor.

The demonstration office fire test was carried out on a

compartment constructed using concrete blockwork. The

fire load consisted of actual office furniture and documents

that weighed as much as 45 kg/m2 of wood. The test area

was 18 m wide by up to 10 m deep. The columns and

beam-to-column connections were protected whereas the

primary and secondary beams were left unprotected.

Figure 6 shows the floor plan of the Cardington frame

that marks the locations of the fire tests.

4. Behavior of Structural Members in a Fire

4.1. Single beam behavior

Figure 7 compares the beam behavior of the Cardington

frame (Robinson, 1998) in the standard fire test and the

restrained beam test. The beam tested was 356 × 165 UB,

with load ratios of 0.37 and 0.4 in the standard test and

the Cardington test, respectively. In the standard test, the

beam tested was 4.5 m long, with simply supported end-

conditions without axial restraint, whereas in the Cardington

frame test, the unprotected 9 m beam was connected to

the other members in the frame.

The beam behavior in the isolated member test indicated

that the deflection began slowly as the temperature rose

and progressively accelerated up to the termination point

of span/30 at a temperature of approximately 705oC. In

the Cardington frame test, however, the beam behaved

differently, with an almost constant deflection rate

throughout, as shown in Fig. 7. The beam showed no sign

of ‘runaway’ even when the beam temperature was

875oC, at which point the test was terminated because of

a breakdown in the deflection measuring instrumentation.

Had the deflection continued at the same rate, a

temperature of over 1,000oC would have been needed to

achieve the standard test criterion of span/30 deflection

without protection. This comparison raises an obvious

concern on whether or not standard fire tests reflect the

real performance of structural members in a fire.

4.2. Behavior of joints

Two types of joints were used in the Cardington frame:

flexible (partial-depth) end-plates and fin-plates. Flexible

end-plates were used for beam-to-column joints and fin-

plates for beam-to-beam joints. These joints are usually

considered pinned joints, which are assumed to transfer

only shear forces and to have sufficient flexibility to

allow rotation. Observations from Cardington fire tests

(Al-Jabri et al., 1999; Al-Jabri and Hago, 2003; and Wald

et al., 2004) indicate that the joints showed signs of being

subjected to high tensile forces (Fig. 8). In the flexible

end-plate joints, the plates were fractured on one side

whereas the other side remained intact, as shown in Fig.

9(a); and in the fin-plates (the beam-to-beam joints), the

bolts were sheared [Fig. 9(b)]. The fracture resulted from

the high tensile forces that developed during the cooling

of the connected beam under the large rotations

associated with the pinned joints. During the cooling, the

steel contracted, which caused a reduction in the vertical

deflection of the beams, whereas a significant tensile

force was applied to the weld that connected the partial-

depth end-plate to the beam flange (Fig. 8). This tensile

strain on cooling was relieved by the plate fracture or the

bolt shear of the joint (Fig. 9). The shear fracture of the

bolts in the fin-plate joints suggests that flexible end-plate

joints act more reliably in a fire than fin-plate joints. The

behavior of joints during structural cooling has to be

further investigated. The suggested initial solution to this

situation is the design of the joint to have significant

ductile behavior so as to ensure that the shear capacity is

maintained when the what? is subjected to high tensile

forces in a fire (Sarraj et al., 2006).

The temperature of the beam bottom flange was

Figure 6. Floor plan of the Cardington test frame showingthe locations of the fire tests.

Figure 7. Comparison of the deflection behavior of abeam in the standard test and in the Cardington frametest.

Page 6: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

232 Khalifa S. Al-Jabri

considerably higher than the temperature of the joint. At

the maximum temperature, the joint temperature was

almost 200oC lower than the limiting temperature of the

beam. The temperature of the bottom bolts was higher

than the temperature of the top bolt, due to the shielding

by the composite slab, which acted as a heat sink to the

top part of the joint. The end-plate was hotter than the

bolts at the same level because of the ratio of the bolt

diameter to the end-plate thickness.

Also, observations of the joint behavior at Cardington

showed that the bolts and welds did not suffer from

premature failure during the heating phase of the fire.

During the cooling phase, however, a number of bolts

that acted in a single shear in the fin-plate joints suffered

from shear failure due to the tensile forces generated by

thermal contraction, as shown in Fig. 9(b).

4.3. Local buckling of beams

Most of the internal beams showed signs of local

buckling during the heating phase in the lower flange and

in a part of the web in the vicinity of the joints, as shown

in Fig. 10. This behavior was caused by the restraint to

thermal expansion and the negative moment caused by

the rotational restraint from the joint. The restraint to

thermal expansion was provided by the surrounding

cooler structure and the structural continuity of the test

frame. This took place in the beam due to the inability of

the lower flange of the beam to transfer the high axial

forces induced in the beam to the adjacent beam-columns

after the closure of the gap in the lower part of the joints

(Wald et al., 2004). Conservatively, therefore, the joints

should be assumed as ‘pinned’ and the connected beams

as simply supported, allowing larger mid-span deflections

to develop than when the beams are semi-rigidly

connected. Local buckling was found as but a minor

concern in isolated member fire tests (Al-Jabri et al.,

1999).

4.4. Behavior of columns

Observations from the Cardington fire tests show that

the internal and external columns were subjected to high

moments, which caused local squashing of the columns,

as shown in Fig. 11, although no collapse occurred because

the structure had the ability to carry the column load

using an alternative load path (i.e., the tensile membrane

action in the composite floor). These moments were

caused by the expansion of the connecting beams, the

expansion of the heated floor relative to the other floors,

and the induced P-δ effects during a fire. Bailey (2000)

pointed out that if these moments were simply included

within the present member design procedure outlined in

BS5950: Part 8 or EC3: Part 1.2, the calculations would

show that the columns would fail during the fire due to

local plasticity. In steel-framed braced structures, however,

bracing causes the existence of a significant degree of

redundancy. Large localized stresses are caused by

induced column moments that result in the formation of

plastic hinges at the floor level without causing overall

collapse. It is suggested that due to the good inherent fire

resistance of the composite slab, it provides horizontal

restraint to the columns at the floor level during a fire.

Therefore, the influence of the expanding beams on the

overall stability of the column must be checked during a

fire, although it is assumed that the localized plasticity

can be accommodated.

To investigate the stability of columns during a fire in

view of the column behavior observed from the Cardington

Figure 8. Tensile forces induced on the joints during thecooling phase of the structure.

Figure 9. Typical failure modes of joints in the Cardington frame tests.

Page 7: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

Behavior of Steel-framed Buildings in a Fire 233

tests, a simple analytical model was developed by Bailey

(2000). In the model, columns were subjected to both

axial loads and moments caused by the expansion of the

connected beams. Although the effect of floor slabs was

ignored in the analysis, the heated beams were restrained

axially to provide lateral restraint to the columns. The

effects of many parameters that influence the column

behavior, such as the beam-to-column heating rate, the

beam and column sizes, the beam-to-column connection

rigidity, the axial restraint at one end of the heated beams,

the span of the beams, the base rigidity of the heated

column, and the column axial load were studied. The

results of the analysis indicated that instability occurred

in the column, even though it was forced into a double

curvature and restrained at the floor level. This instability

was caused by the P-δ effect in the column, which was

enhanced due to the enforced deflected shape of the

column from the expansion of the connecting beams. The

analysis also showed that parameters such as the beam-

to-column heating rates, the beam cross-section size, the

span of the beams, the end-rigidity of the heated column,

and the column axial load have significant influence on

column instability. The parameters that had a nominal

effect on the behavior of the column, however, included

the column cross-section size, the beam-to-column

connection rigidity, and the horizontal restraint to the

heated beams.

4.5. Behavior of the composite floor

The composite flooring system used in the Cardington

full-scale frame comprised steel downstand beams that

acted compositely with a floor slab and that were constructed

using a trapezoidal steel deck, lightweight concrete, and

an anticrack A142 steel mesh. The overall depth of the

slab was 130 mm, and the mesh was situated 15 mm

above the steel deck. The results of the isolated fire tests

conducted on the beam-to-column composite joints (Al-

Jabri et al., 1998) showed that the composite slab above

the joints caused a 20-30% reduction in the beam’s top

flange temperatures in comparison with the beam’s

bottom flange temperature. This suggested that the concrete

slab acted as an insulation and a heat sink to the top of the

beam, which enhanced the joint performance at an

elevated temperature.

Observations from the Cardington fire tests and other

large building fires have shown that the behavior of the

composite floor slab plays a crucial role in providing

enhanced fire resistance to the structure. Also, it has been

confirmed that the performance of a steel frame with a

composite flooring system is significantly better than that

suggested by current fire design methods (Martin and

Moore, 1997 and Bailey et al., 1999) due to the presence

of tensile membrane action in the composite slab during

the fire at large displacements. At the fire limit state, large

displacements of the structure are acceptable provided

that the fire is contained within the compartment of origin

so that the risk of the fire spreading throughout the

building would be low. During a fire, when significant

numbers of unprotected secondary steel beams are

damaged, the lightly reinforced composite slab acts as a

membrane supported by cold perimeter beams and

protected columns. Due to the failure of unprotected steel

beams to carry any load, the composite slab utilizes its

full bending capacity to span the adjacent cooler members.

With increasing displacement, the composite slab acts as

a tensile membrane that carries the loads in the

reinforcement. In the case of simply supported edges, the

supports will not anchor these tensile forces and a

compressive ring will form around the edge of the slab.

Failure will only occur at large displacements with the

fracture of the reinforcement. Fig. 12 shows the tensile

membrane action of a composite slab with no horizontal

restraint around its perimeter.

Current fire engineering design methods completely

ignore the beneficial effect of the tensile membrane

action at large displacements that occurs in the composite

slab during a fire. To take advantage of this behavior

when designing structures in a fire, a new design method

has been developed (Bailey et al., 2000; Bailey and

Moore, 2000a; Bailey and Moore, 2000b; Bailey, 2001;

and Bailey, 2003) to calculate the performance of steel-

Figure 10. Local buckling of primary and secondarybeams in the vicinity of the joints.

Figure 11. Squashing of the columns due to large moments.

Page 8: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

234 Khalifa S. Al-Jabri

framed buildings with composite flooring systems that

are subject to fire. The method is valid for both square

and rectangular slabs and conforms to the mode of

behavior observed in the Cardington full-scale fire tests.

This method uses a simple energy approach to calculate

the load-carrying capacity of a composite flooring

system. The energy of the lightly reinforced composite

slab is based on the yield-line approach that has been

modified to account for the enhancement caused by in-

plane forces. The developed method has been extended

further to incorporate the membrane action of the slab

and the beam systems that act compositely (Bailey, 2004).

The basic principle of the design method (Bailey and

Moore, 2000a; Bailey and Moore, 2000b; and Burgess,

2005) is summarized below.

(a) The entire floor slab is subdivided into a series of

square or rectangular slab panels (Fig. 13) that incorporate

a number of unprotected composite beams. The size of

the slab panels may be governed by the fire compartment

size.

(b) During a fire, each slab panel area supports the

applied load via the membrane action of the composite

slab and the flexural strength of the grillage of the

composite beam within the panel area. Protected beams

or beams designed to have sufficient resistance to support

an applied load for the duration of a fire are assumed to

support the perimeter of the slab panel. The columns are

either protected or designed to withstand the applied load

for the fire resistance period.

(c) The fire resistance period is defined in BS5950: Part

8 and EC4: Part 1.2, and the total load on the structure at

the fire limit state is calculated.

(d) The maximum allowable vertical displacement for

each slab panel is determined, the beam with the highest

load ratio is identified, and the load carried by the beam

for the specified fire resistance period is calculated.

(e) The flexural strength at the fire limit state of the

composite slab is calculated, considering only the mesh

reinforcement and the concrete components using the

yield-line theory. The slab strength can be enhanced due

to the membrane action based on the maximum allowable

vertical displacement.

(f) The load-carrying capacity of the slab panel is

calculated by summing up the capacity of the composite

slab and the capacity of the grillage of the unprotected

composite beams.

(g) If the load-carrying capacity of the slab panel and

the grillage of the unprotected composite beams are

greater than the applied load, the beams can be left

unprotected.

This method is commonly known as the BRE design

method, which has been developed further (Newman et

al., 2000) into a series of design tables that allow the

designer to leave large numbers of secondary beams

unprotected in buildings requiring 30- and 60-minute fire

resistance, although some compensation features, such as

increased mesh size and density, may be required.

The developed method was validated against the results

of the Cardington fire tests and of the 15 small-scale tests

conducted on horizontally unrestrained slabs that were

subjected to large vertical displacements (Foster, 2004).

A comparison of the results of the two tests shows that

the design method compares well with experimental

results and is generally conservative. It has been

recommended that the method be further developed to

account for the bond characteristics of reinforcement and

the change in the failure mode for some orthotropically

reinforced slabs. A number of numerical studies and

computer simulations have also been performed (Huang

et al., 2003a; Huang et al., 2003b; Izzuddin and Elghazouli,

2004a; Izzuddin and Elghazouli, 2004b; Burgess, 2005;

and Foster et al., 2006) to investigate the influence of the

Figure 12. Tensile membrane action of a floor slab withno horizontal restraint around its perimeter.

Figure 13. Division of the floor plate into slab panels.The floor plate is divided into (a) square panels and (b)rectangular panels.

Page 9: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

Behavior of Steel-framed Buildings in a Fire 235

tensile membrane action in the composite slab on the

behavior of composite buildings in a fire based on the

results of the Cardington fire tests. These studies confirm

that while exposed steel temperatures in composite

buildings remain below 400oC, the much cooler concrete

slab plays only a trivial role in the load-carrying

mechanism, apart from generating the thermal curvature

of the composite beams. For steel temperatures higher

than 500oC, the significance of the slab progressively

increases. At very high temperatures, the floor slab

becomes the main load-bearing element and the floor

loads above the fire compartment are carried largely by

tensile membrane forces developed mainly in the steel

anticracking mesh or reinforcing bars.

5. Overall Frame Behavior during a Fire

The Cardington frame tests showed that the composite

frame suffered from considerable deformation without

any form of collapse (Fig. 14), even though the unprotected

beams reached temperatures of up to 1,100oC. At this

temperature level, BS5950: Part 8 indicates that only 3%

of the member’s design strength remains. The results of

the experiment on the Cardington frame demonstrated

with no doubt the major contribution of the composite

floor to the survival of the frame in a fire. The floor

performed very well in all the tests, which supports the

results of previous small-scale tests (Al-Jabri et al., 1999)

that showed that this type of floor system has good

inherent fire resistance. Results of analytical studies

conducted on the Cardington frame tests (Gillie et al.,

2001; Gillie et al., 2002; Huang et al., 2002; and O’Connor

et al., 2003) confirmed that the effects of thermal

expansion dominated the response of the structure and

that the material degradation and the gravity loading were

of secondary importance. It was also found that at extreme

temperatures, the significant load-carrying mechanism was

the tensile membrane action in the reinforcement mesh

and that gravity loading can influence the magnitude of

the tensile forces produced. The results also suggested

that one method of helping maintain structural integrity in

composite structures during extreme fires is to ensure that

a sufficient degree of ductile reinforcement is present in

the concrete floor slabs. Columns were found to be more

critical than beams and will need protection in multi-story

buildings. The behavior of joints during cooling due to

restraint to thermal expansion has to be further investigated.

The survival of steel-framed structures from the severe

fire conditions experienced in the Cardington project or in

buildings subjected to real fire accidents raises a number

of fundamental issues as to whether or not current design

methods that were based on isolated members tests reflect

the true behavior of structures in a fire. The results of

these tests confirmed that current fire engineering design

methods are too conservative. A new design method has

been developed (Newman et al., 2000) that takes into

account the inherent fire resistance of steel-framed

members in fire conditions. This method incorporates the

beneficial effects of a composite slab and beam systems

on the survival time of steel-framed structures in a fire.

6. Conclusions

This paper examined the effects of fires on the behavior

of multi-story steel-framed buildings. It can be concluded

that, even in the context of the structural fire engineering

approach of modern design codes, predictions of behavior

based on furnace tests or numerical modeling of isolated

members are unreliable. The behavior of the members

within a continuous, compartmented structure is very

different from the behavior of isolated members. Structural

continuity, restraint to thermal expansion provided by the

adjacent members, the beam-to-column joints, and the

tensile membrane action of the composite slab have

demonstrated a significant positive influence on the

bevahior of the entire structure in the event of a fire. Data

from the Cardington fire tests and the subsequent

experimental and analytical studies provide fundamental

information that is very important for researchers studying

the performance of steel-framed buildings in a fire for

many years to come so as to develop new design approaches

that take into account the interaction between structural

members in a fire. The developments that have already

taken place in the past few years have been very

significant in understanding the reality of structural

behavior in a fire, which will undoubtedly lead to the

emergence of new rational fire engineering design

methods that may allow us to construct safer, more

effective, and more efficient fire-resistant buildings in the

future.

References

Al-Jabri, K.S. and Hago, A.W. (2003). Towards a Rational

Approach to the Design of Steel-framed Buildings in a

Fire. Proceedings of the 9th Arab Structural Engineering

Conference, United Arab Emirates, pp. 971-978.

Al-Jabri, K.S.; Burgess, I.W.; Lennon, T.; and Plank, R.J.

Figure 14. Overall frame deformation after a fire in theCardington 8-story building.

Page 10: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

236 Khalifa S. Al-Jabri

(1999). The Performance of Frame Connections in a Fire.

ACTA POLYTECHNICA-EUROSTEEL ‘99, 39(5), pp.

65-75.

Al-Jabri, K.S.; Lennon, T.; Burgess, I.W.; and Plank, R.J.

(1998). Behavior of Steel and Composite Beam-Column

Connections in a Fire. Journal of Constructional Steel

Research, 46(1-3), Paper No. 180.

Anon (1991). Investigation of the Broadgate Phase 8 Fire.

Structural Fire Engineering. Steel Construction Institute,

United Kingdom.

Armer, G.S.T. and Moore, D.B. (1994). Full-scale Testing on

Complete Multi-story Structures. The Structural Engineer,

72(2), pp. 30-31.

Bailey, C.G. (2000). The Influence of the Thermal Expansion

of Beams on the Structural Behavior of Columns in Steel-

framed Structures during a Fire. Engineering Structures,

22, pp. 755-768.

Bailey, C.G. (2001). Membrane Action of Unrestrained

Lightly Reinforced Concrete Slabs at Large

Displacements. Engineering Structures, 23, pp. 470-483.

Bailey, C.G. (2003). Efficient Arrangement for Membrane

Behavior of Composite Floor Slabs in Fire Conditions.

Journal of Constructional Steel Research, 59, pp. 931-

949.

Bailey, C.G. (2004). Membrane Action of Slab/Beam

Composite Floor Systems in a Fire. Engineering

Structures, 26, pp. 1691-1703.

Bailey, C.G. and Moore, D.B. (2000a). The Structural

Behavior of Steel Frames with Composite Floor Slabs

Subject to Fire (Part 1: Theory). The Structural Engineer,

78(11), pp. 19-27.

Bailey, C.G. and Moore, D.B. (2000b). The Structural

Behavior of Steel Frames with Composite Floor Slabs

Subject to Fire (Part 2: Design). The Structural Engineer,

78(11), pp. 28-33.

Bailey, C.G.; Lennon, T.; and Moore, D.B. (1999). The

Behavior of Full-scale Steel-framed Buildings Subjected

to Compartment Fires. The Structural Engineer, 77(8),

pp. 15-21.

Bailey, C.G.; White, D.S.; and Moore, D.B. (2000). The

Tensile Membrane Action of Unrestrained Composite

Slabs Simulated under Fire Conditions. Engineering

Structures, 22, pp. 1583-1595.

BS 5950 (1990). Structural Use of Steelwork in Buildings

(Part 8: Code of Practice for Fire-resistant Design).

British Standard Institution, London, UK.

Burgess, I.W. (2002). Fire Resistance of Framed Buildings.

Physics Education, 37(5), pp. 390-399.

Burgess, I. (2005). Performance and Design of Multi-story

Composite Buildings in a Fire. Proceedings of the KICT

2005 Annual Conference / CUFER Annual Technical

Seminar, Seoul, Korea, pp. 89-109.

EC 3 (1993). Design of Steel Structures (Part 1.2: General

Rules, Structural Fire Design). ENV1993-1-2, European

Committee for Standardization, Brussels.

EC 4 (1994). Design of Composite Steel and Concrete

Structures (Part 1.2: General Rules, Structural Fire

Design). ENV1994-1-2, European Committee for

Standardization, Brussels.

Foster, S.J.; Bailey, C.G.; Burgess, I.W.; and Plank, R.J.

(2004). Experimental Behavior of Concrete Floor Slabs

at Large Displacements. Engineering Structures, 26, pp.

1231-1247.

Foster, S.J.; Burgess, I.W.; and Plank, R.J. (2006). Modeling

the Membrane Action of Model-scale Slabs at Ambient

and Elevated Temperatures. Structures in Fire Workshop,

Aveiro, Portugal, pp. 635-646.

Gillie, M.; Usmani, A.S.; and Rotter, J.M. (2001). A

Structural Analysis of the First Cardington Test. Journal

of Constructional Steel Research, 57, pp. 581-601.

Gillie, M.; Usmani, A.S.; and Rotter, J.M. (2002). A

Structural Analysis of the Cardington British Steel Corner

Test. Journal of Constructional Steel Research, 58, pp.

427-442.

Huang, Z.; Burgess, I.W.; and Plank, R.J. (2002). Modeling

of Six Full-scale Fire Tests on a Composite Building. The

Structural Engineer, 80(19), pp. 30-37.

Huang, Z.; Burgess, I.W.; and Plank, R.J. (2003a). Modeling

the Membrane Action of Concrete Slabs in Composite

Buildings in a Fire (Part I: Theoretical Derivation).

Journal of Structural Engineering, 129(8), pp. 1093-1102.

Huang, Z.; Burgess, I.W.; and Plank, R.J. (2003b). Modeling

the Membrane Action of Concrete Slabs in Composite

Buildings in a Fire (Part II: Validations). Journal of

Structural Engineering, 129(8), pp. 1103-1112.

Izzuddin, B.A. and Elghazouli, M. (2004a). Failure of

Lightly Reinforced Concrete Members under Fire (Part I:

Analytical Modeling). Journal of Structural Engineering,

130, pp. 3-18.

Izzuddin, B.A. and Elghazouli, M. (2004b). Failure of

Lightly Reinforced Concrete Members under a Fire (Part

II: Parametric Studies and Design Considerations).

Journal of Structural Engineering, 130, pp. 19-32.

Lennon, T. (1997). Cardington Fire Tests: Survey of Damage

to an Eight-story Building. BRE Internal Report GD1286/

86, Building Research Establishment, UK.

Martin, D.M. and Moore, D.B. (1997). Introduction and

Background to the Research Program and Major Fire

Tests at BRE Cardington. National Steel Construction

Conference, London, UK, pp. 37-84.

Moore, D.B. and Lennon, T. (1997). Fire Engineering

Design of Steel Structures. Progress in Structural

Engineering and Materials, 1(1), pp. 4-9.

Newman, G.M. (1991). Structural Fire Engineering:

Investigation of the Broadgate Phase 8 Fire. Steel

Construction Institute, UK.

Newman, G.M.; Robinson, J.T.; and Bailey, C.G. (2000).

Fire Safety Design: A New Approach to Multi-story Steel-

framed Buildings. SCI-P288, Steel Construction Institute,

UK.

O’Connor, M.A. and Martin, M.D. (1998). Behavior of a

Multi-story Steel-framed Building Subjected to a Fire

Attack. Journal of Constructional Steel Research, 46(1-3),

Paper No. 169.

O’Connor, M.A.; Kirby, B.R.; and Martin, D.M. (2003).

Behavior of a Multi-story Composite Steel-framed

Building in a Fire. The Structural Engineer, 81(2), pp. 27-

36.

Robinson, J. (1998). Fire--A Technical Challenge and a

Market Opportunity. Journal of Constructional Steel

Research, 46(1-3), Paper No. 179.

Sarraj, M.; Burgess, I.W.; Davison, J.B.; and Plank, R.J.

Page 11: Behavior of Steel-framed Buildings in a FireC).pdf · 2018-01-08 · (BS5950: Part 8, 1990). Although the philosophy behind specifying the protection thickness is adequate, it is

Behavior of Steel-framed Buildings in a Fire 237

(2006). Finite Element Modeling of Fin-plate Steel

Connections in a Fire. Structures in Fire Workshop,

Aveiro, Portugal, pp 315-326.

Wald, F.; Simõnes da Silva, L.; Moore, D.; and Santiago, A.

(2004). Experimental Behavior of Steel Joints in a

Natural Fire. ECCS-AISC Workshop, Amsterdam, The

Netherlands.