Sei sulla pagina 1di 34

37

Structural Design
for Fire Safety

37.1 Introduction .................................................. 37-1


37.2 Fire Severity................................................... 37-2
Standard Fire Exposure  Natural Fire Exposure
37.3 Introduction to Heat Transfer............................ 37-8
Conduction  Convection  Radiation  Some Simplified
Solutions of Heat Transfer  Section Factors  Thermal
Properties of Materials
37.4 Design of Structural Elements at Elevated
Temperatures ................................................. 37-14
General  Mechanical Properties of Steel and Concrete at
Elevated Temperatures  Design of Steel Elements  Composite
Composite Steel/Concrete Members
37.5 Design for Unprotected Steelwork ...................... 37-27
Integration of Structural Load Bearing and Fire Protection
Functions of Concrete  Utilizing Whole Building Performance
Yong C. Wang in Fire
School of Aerospace,
37.6 Concluding Remarks ....................................... 37-31
Mechanical and Civil Engineering,
The University of Manchester, References ............................................................. 37-32
Manchester, U.K. Appendix .............................................................. 37-33

37.1 Introduction
Structural fire safety is one of the three requirements that have to be fulfilled by a fire resistant
construction, whose function is to ensure that a fire in a building is contained within the compartment
of origin so that occupants in other parts of the building can escape to safety and fire damages do not
become excessive. To achieve this, the load bearing structure of a fire resistant construction should
not collapse in fire. The other two fire resistance requirements are
 Insulation. The unexposed surface of a fire resistant construction should not be heated excessively
and cause further ignition. Clearly, whether any material will be ignited or not will not only depend
on the temperature of the unexposed surface, but also on its nature and its relative position to the
unexposed surface. Nevertheless, at present, regulations worldwide limit the average temperature
rise on the unexposed surface to 140 C and the maximum local temperature rise to 180 C.
 Integrity. Gaps should not develop in fire resistant construction to spread fire.
The practice of dividing a building into a number of compartments bounded by fire resistant
construction is called fire resistant compartmentation. It should be pointed out that fulfillment of the

0-8493-1569-7/05/$0.00+$1.50
# 2005 by CRC Press 37-1

Copyright 2005 by CRC Press


37-2 Handbook of Structural Engineering

above three fire resistance requirements only applies to those elements of construction that are
necessary for the fire resistant compartment to contain fire. Other elements, whose failure to fulfill these
requirements does not lead to a failure of the fire resistant compartment, do not require any fire
safety design consideration. Before the designer commences detailed structural fire safety design
calculations, he should work with the client and the fire service authority to determine the size of the
fire resistant construction. This will depend on factors such as fire regulations on the maximum size
of fire resistant compartmentation, insurance premium, and fire brigade access and is beyond the
scope of this chapter.
Until recently, assessment of fire resistance of a construction is performed experimentally in standard
fire resistance test furnaces and under the standard fire condition. Each country has its own fire
resistance test standard [e.g., ASTM E-119 (ASTM 1985) in the United States, BS 476 (BSI 1987) in
the United Kingdom, and ISO 834 (ISO 1975)], but they are largely similar. The standard fire resistance
test has many shortcomings, for example, high cost, time consuming, limitation of specimen size,
idealized loading condition, idealized support condition, lack of repeatability, and unrealistic fire
exposure. It is now possible to perform some fire resistant design by calculations and the objective of
this chapter is to introduce the reader to structural design calculations to ensure stability of load
bearing members in the event of a fire.
It should be pointed out that the behavior of a complete structure in fire and that of isolated elements
can be different because a complete structure will have characteristics, such as load redistribution,
structural interactions, that will not exist in isolated elements. Also, it is worth noting that current
calculation methods are based on flexural behavior at small deflections. The behavior of structural
elements at large deflections can be vastly different and such a different behavior may be explored to
improve structural fire safety design. Behaviors of elements at large deflections and complete structures
in fire are beyond the scope of this chapter. Interested readers may consult the book by Wang (2002).
In general, design calculations to check structural safety in fire involves three parts:

1. Assessment of the fire severity to which a structural member is exposed. For structural fire safety
design, a fire is usually quantified by a temperature–time relationship of the fire.
2. Evaluation of the temperature field in the structural member under the above fire condition.
3. Calculation of the remaining load carrying capacity of the structural member at elevated
temperatures and comparison with the applied load.

This chapter will introduce the reader to all three aspects of structural fire safety design, while
emphasizing on the third.
It is understandable that the September 11 tragedy has initiated the interest of many engineers in
structural fire safety design. However, it must be mentioned that there had already been great progresses
on this topic well before the September 11 event in many parts of the world, particularly in Europe and
the United Kingdom. At present, there is a systematic and comprehensive coverage of structural fire
safety design methods in Europe, through developments of the so-called Eurocodes. Thus, this chapter
will adopt Eurocodes as the basis of its design guidance. However, it is hoped that there will be sufficient
explanations of the fundamental engineering principles so that the basis of Eurocode design rules can
be similarly adopted in different design environments.

37.2 Fire Severity


Fire severity to structural fire safety design is akin to applied mechanical loads on a structure to
structural design at ambient temperature. Fires can occur anywhere; however, for structural fire safety
design, they are often assumed to take place in a building enclosure. In such a situation, starting from
ignition, a fire can go through a number of stages. In the early stages, combustion is restricted to local
areas near the ignition source and temperatures of the combustion gases are low, the safety of a structure
exposed to fire attack is very rarely threatened. Later, under certain circumstances, a localized fire can

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-3

transform very quickly to involve all the combustible materials in the fire enclosure. The transition from
a localized fire to a fire engulfing the entire enclosure is called flashover and the fire at this stage is called a
postflashover fire. At this stage, the combustion gas temperatures are very high and stability of the
building structure may be threatened, the consequence of which can lead to rapid fire spread and loss of
life and property. It is this stage of the fire behavior that will be described below. Interested readers
should consult some excellent textbooks, such as Drysdale (1999) and Karlsson and Quinterie (2000), to
obtain a deep understanding of enclosure fire behavior.

37.2.1 Standard Fire Exposure


There are two ways of dealing with postflashover fires for structural fire safety design. First, the standard
fire temperature–time relationship may be adopted. The standard fire equation is similar in different
countries of the world. In the European standard (CEN 2000a), the standard fire temperature–time
relationship is given by
Tfi ¼ Ta þ 345 logð8t þ 1Þ ð37:1Þ
where the standard fire exposure time (t) is in minutes and the fire temperature Tfi and ambient
temperature Ta are in degrees celsius.
Equation 37.1 is used for wood based, or cellulosic, fires. For fire resistant design of offshore
structures, the standard hydrocarbon fire curve should be used. This fire has a much faster rate of initial
increase in temperature. The standard hydrocarbon fire temperature–time relationship is given by
(CEN 2000a)
 
Tfi ¼ 1080 1  0:325e0:167t  0:675e2:5t þ Ta ð37:2Þ
Figure 37.1 plots the two standard fire curves. It is obvious that the standard cellulosic fire curve gives
a monotonically increasing temperature–time relationship that cannot be sustained in any real fire. In
order to reflect some reality in the standard fire exposure, a limiting time of fire exposure is specified.
This is the familiar standard fire resistance rating. In standard fire resistance design calculations, spe-
cifications for the required standard fire resistance rating are based on very broad criteria such as the
occupancy type and height of a building. Whilst these criteria give a broad indication of fire load and
consequence of fire exposure, they do not consider other important factors that affect the behavior of an
enclosure fire such as ventilation condition and construction materials.

37.2.2 Natural Fire Exposure


As a regulatory control tool, the standard fire exposure is simple to use. However, fires cannot be
expected to behave according to the standard temperature–time relationship. For more realistic

Hydrocarbon curve, Equation 37.2


1200
1000
Temperature, °C

800
Cellulosic curve, Equation 37.1
600
400
200
0
0 50 100 150 200 250
Fire exposure time, min

FIGURE 37.1 A comparison of standard cellulosic and hydrocarbon fire temperature–time relationships.

Copyright 2005 by CRC Press


37-4 Handbook of Structural Engineering

assessment of performance of structures in fire, it is necessary to quantify realistic fire behavior.


Assuming that a fire enclosure is at the same temperature during the postflashover phase, the fire
temperature–time relationship may be determined by carrying out an energy balance analysis for the fire
enclosure; that is
heat input into the fire (heat released from combustion) = heat losses from the fire
Figure 37.2 depicts a postflashover enclosure fire situation. Heat losses from the fire include
1. Heat lost to the outside by hot gases flowing out of the fire compartment through openings ðQ_ lc Þ:
2. Heat lost to the enclosure lining ðQ_ lw Þ:
3. Heat lost to the outside environment by radiation through the opening ðQ_ lr Þ:
4. Heat required to increase the combustion gas temperature ðQ_ lg Þ.
The rate of heat release is the most important factor. At present, due to the difficulty of numerically
modeling fire spread and random distribution of combustible materials in a fire enclosure, it is not
possible to accurately calculate the rate of heat release of a fire. Nevertheless, from considerations of the
main governing factors of burning, a number of empirical equations have been derived. From basic
hydrodynamics, pffiffiffiffiffiit can be shown that the amount of hot gases flowing out of a fire compartment is
related to Av hv , the so-called ventilation factor, where Av is the opening area and hv is the opening
height. In order to sustain burning, cold air should be supplied into the fire compartment to replenish pffiffiffiffiffi
the lost hot gases. Thus, the amount of cold air entering the fire compartment pffiffiffiffiis
ffi also related to Av hv . It
follows that the amount of fresh oxygen supply to the fire is related to Av hv . If burning is ventilation
controlled, that is, the rate of burning is governed
pffiffiffiffi
ffi by the amount of fresh oxygen available, the rate of
heat release of a fire is a function of Av hv . On the other hand, if the opening is large but the burning
area is small, a fire can become fuel controlled, that is, the rate of burning is governed by the available
surface area of the fuel bed or combustible materials. Also, the amount of available combustible
materials, that is, the fuel load, will determine the duration of burning.
Fire development inside a fire enclosure will also be affected by thermal properties of the fire enclosure
lining materials, that is, the bounding walls and floors. A material that has a low thermal conductivity,
that is, heat is difficult to penetrate the material, will lose a small amount of heat through the material.
A material that has a high thermal capacitance, that is, a large amount of heat is required to raise its
temperature, will absorb a large amount of heat of the burning fire and vice versa. Combining these two
factors, the quantity that is used to describe the thermal properties of fire enclosure lining materials is
krC where k and rC are the thermal conductivity and thermal capacitance of the lining materials,
respectively.
Using the three quantities mentioned above, that is, the ventilation factor, the fire enclosure lining
material property, and the fuel load, a number of approximate temperature–time relationships of

. .
Qlw Qlc
.
Qlw
.
Qlr

. .
Qlg Qlw

FIGURE 37.2 Fully developed enclosure fire, showing various heat losses.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-5

postflashover enclosure fires have been developed. Among them, the so-called parametric temperature–
time curves of Eurocode 1 Part 1.2 (CEN 2000a), based on the results of Pettersson et al. (1976), are
widely accepted. As shown in Figure 37.3, a parametric fire curve has an ascending branch and a
descending branch. The ascending branch is used to describe the temperature–time relationship of a fire
during its growth and steady burning stages, when it is ventilation controlled. The descending branch
describes the decay period of the fire. The ascending branch is expressed by
 
Tfi ¼ 1325 1  0:324e0:2t  0:204e1:7t  0:472e19t ð37:3Þ
where the modified time t (in hours) is related to the real time t (in hours) by
t ¼ tG ð37:4Þ
in which G is a dimensionless parameter, given by
  
O 2 1160 2
G¼ ð37:5Þ
0:04 b
In Equation 37.5, O is the ventilation factor defined as
pffiffiffiffiffi
Av hv
O¼ ð37:6Þ
At
in which At is the total enclosure
pffiffiffiffiffiffiffiffiffi (including openings) area.
In Equation 37.5, b ¼ krC [in J/(m2 s1/2 K)] is the overall thermal property of the fire enclosure
lining material. For a fire enclosure constructed of a combination of different lining materials, com-
plicated equations have recently been introduced in Eurocode 1 Part 1.2 (CEN 2000a) to find an
equivalent value of b.
The ascending branch of the fire temperature–time relationship terminates at time (td ) when the
maximum fire temperature is obtained. This time is a function of the fire load in the fire enclosure and
is given by
qt,d G
td ¼ 0:00013 (in hours) ð37:7Þ
O
In Equation 37.7, qt,d is the fire load density (in MJ/m2) related to the total surface area of the fire
enclosure At. Since fire load density is usually specified with regard to the floor area Af, the fire load per
enclosure area qt,d is related to the fire load per floor area (qf,d) using
qt,d ¼ qf ,d Af =At ð37:8Þ
It can be seen that the ascending branch of the fire temperature–time curve is not dependent on the
fire load. This is because a fire is assumed to be ventilation controlled and the rate of heat release is the

Temperature

Tfi,max

td* Time

FIGURE 37.3 Parametric time–temperature curve of Eurocode 1 Part 1.2.

Copyright 2005 by CRC Press


37-6 Handbook of Structural Engineering

same, depending only on the ventilation condition. The effect of fire load is to change the duration of
burning td according to Equation 37.7.
For simplicity, the descending branch is given by a straight line. Since structural behavior is only
slightly affected by the descending branch of the fire temperature–time relationship, more complicated
equations for the descending branch are not justified. The rate of the descending branch depends on the
fire duration. The fire temperature during cooling is given by
Tfi ¼ Tfi,max  625ðt  td Þ for td
0:5
Tfi ¼ Tfi,max  250ð3  td Þðt  td Þ for 0:5 < td < 2:0 ð37:9Þ
Tfi ¼ Tfi,max  250ðt  td Þ for td 2:0
In Equation 37.9, Tfi,max is the maximum fire temperature, obtained by substituting the time in
Equation 37.7 into Equation 37.3.
In Eurocode 1 Part 1.2 (CEN 2000a), the limit of application of the above fire temperature–time
relationship is for fire compartments up to 100 m2 in floor area with the maximum compartment
height at 4 m. For larger or taller compartments, the effect of nonuniform temperature distribution
in the fire enclosure cannot be ignored. Unfortunately, simple methods are not available yet.
From previous discussions, it is clear that the fire temperature–time relationship depends on the
amount of combustible materials (or fuel load) in a fire enclosure, the ventilation condition, and thermal
properties of the fire enclosure lining material. During a design, the ventilation condition and thermal
properties of the fire enclosure lining material may be estimated from construction details, that is,
the window size and construction materials. Thermal properties of some enclosure lining materials
may be found in Table 37.1.
The design fire load is building specific. However, since the exact type and amount of combustible
materials will not be known during the design stage, it is unlikely that the design fire load can be known
with any certainty. In fire engineering design calculations, it is common to specify a generic fire load for a
type of building, depending on its proposed use. This is similar to specifying a general structural load for
structural design at ambient temperature. Values in Table 37.2 may be used as a guide. More detailed
information on fire load may be obtained from a Conseil International du Batiment (CIB) report (CIB
1986). It is important to point out that there are many uncertainties about the design fire load. When
conducting a fire engineering design, the designer should perform a sensitivity study to investigate the
consequence of adopting a range of possible fire loads.

EXAMPLE 37.1
Natural fire exposure
Figure 37.4 shows the dimensions and other design data of a fire enclosure. Evaluate the postflashover
fire temperature–time curve inside the enclosure.

TABLE 37.1 Thermal Properties of Generic Fire Protection Materials


Thermal Moisture
Generic Density, conductivity, Specific heat, content,
material kg/m3 W/(m K) J/(kg K) (% by wt.)

Sprayed mineral fiber 250–350 0.1 1050 1.0


Vermiculite slabs 300 0.15 1200 7.0
Vermiculite/gypsum slabs 800 0.15 1200 15.0
Gypsum plaster 800 0.2 1700 20.0
Mineral fiber sheets 500 0.25 1500 2.0
Aerated concrete 600 0.3 1200 2.5
Lightweight concrete 600 0.8 1200 2.5
Normal weight concrete 2200 1.7 1200 1.5

Source: Lawson, R.M. and Newman, G.M., 1996, Structural Fire Design to EC3 & EC4, and Comparison With BS 5950,
Technical Report, SCI Publication 159, The Steel Construction Institute.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-7

TABLE 37.2 Design Fuel Load Density


Fuel load Fuel load density, qf,d
density band Type of construction (MJ/m2)
I Assembly, low hazard 350
Open sided car parks
II Assembly, ordinary hazard 500
Residential
Office
Industrial, low hazard
Storage, low hazard
III Assembly, high hazard 750
Shops and commercial
IV Industrial, high hazard 1000
V 1250
VI Storage, high hazard 1500

Source: British Standards Institution (BSI), 2001, Draft BS 9999, Code of Practice for Fire Safety in the
Design, Construction and Use of Buildings (London: British Standards Institution).

4m Other design data


Design fire load
720 MJ/m2 of floor area

Enclosure lining material


lightweight concrete blocks,
density 1600 kg/m3,
1.5 m Opening 9m specific heat 1200 J/(kg K)
thermal conductivity 0.8 W/(m K)

6m

FIGURE 37.4 Enclosure information for Example 37.1.

Calculation results
Floor area: Af ¼ 54 m2
Total enclosure area: At ¼ 228 m2
Design fire load density per m2 enclosure area: qt,d ¼ 720 54/228 ¼ 170.5 MJ/m2
Window area: Av ¼ 9 m2 pffiffiffiffiffiffi
9 1:5
Opening factor: O ¼ ¼ 0:04835 m1=2
228
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Lining property: b ¼ 1600 0:8 1200 ¼ 1239:4 J=ðm2 s1=2 KÞ
   
0:04835 2 1160 2
Equation 37.5 gives G ¼ ¼ 1:28
0:04 1239:4
170:5 1:28
Equation 37.7 gives td ¼ 0:00013 ¼ 0:587 h
0:04835

Equation 37.3 gives Tfi,max ¼ 844 þ 20 ¼ 864 C
The real time at the maximum fire temperature td ¼ 0.587/1.28 ¼ 0.458 h
For the cooling part, Equation 37.9 will be used, giving the time t necessary to reach the ambient
temperature as t ¼ 1.937 h. The real time is 1.937/1.28 ¼ 1.5136 h.
Figure 37.5 plots the complete fire temperature–time relationship.

Copyright 2005 by CRC Press


37-8 Handbook of Structural Engineering

900
Fire, temperature,
800 Example 37.1

700

600 Unprotected steel,


Temperature, °C

Example 37.3
500

400
Protected steel,
300 Example 37.4

200

100

0
0 10 20 30 40 50 60
Fire exposure time, min

FIGURE 37.5 Temperature–time relationships.

37.3 Introduction to Heat Transfer


Having determined the temperature–time relationship of a fire to which a structural member is exposed,
the next step is to calculate the temperatures in the structural member. Building structures are usually
treated as either one-dimensional, such as beams and columns, or two-dimensional, such as slabs and
walls. For a one-dimensional member, it is common to assume that there is no temperature variation
along its length. For a two-dimensional member, temperatures are assumed to change only through its
thickness. Calculating temperatures in a structural member exposed to fire involves using heat transfer
analysis. Only a brief introduction to heat transfer theory will be given in this chapter, with particular
emphasis on applications under fire conditions.
There are three basic mechanisms of heat transfer: conduction, convection, and radiation. In con-
duction, energy or heat is exchanged in solids on a molecular scale without any movement of macro-
scopic portions of matter relative to one another. Convection refers to heat transfer at the interface
between a fluid and a solid surface. Here, the exchange of heat is due to fluid motion. This motion may
be the result of an external force, causing the fluid to flow over the solid surface at speed. This is called
forced convection. Convection can also occur due to buoyancy-induced flow when there is a tem-
perature gradient in the fluid, causing a density gradient. This is called natural convection. Radiation is
the exchange of energy by electromagnetic waves that, like visible light, can be absorbed, transmitted or
reflected at a surface. Unlike conduction and convection, heat transfer by radiation does not require any
intervening medium between the heat source and the receiver. Thus, in the context of structural fire
safety design, heat conduction describes the heat transfer process inside a structural member and heat
convection and radiation describe the thermal boundary condition of the structural member.

37.3.1 Conduction
The basic equation for one-dimensional heat conduction is Fourier’s law of heat conduction. It is
expressed as
dT
Q_ ¼ k ð37:10Þ
dx
where, refering to Figure 37.6, dT is the temperature difference across an infinitesimal thickness dx. Q_ is
the rate of heat transfer (heat flux) across the material thickness. The minus sign in Equation 37.10
indicates that heat flows from the higher temperature side to the lower temperature side.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-9

Unit
area

.
Q = – k dT
dx

x x + dx
(T ) (T + dT )

FIGURE 37.6 Heat conduction in one dimension.

Unit
area

.
Q = – k ∆T
∆x

∆x

x x + ∆x
(T ) (T + ∆T )
T1 T2

FIGURE 37.7 Temperature distribution with constant thermal conductivity.

The constant of proportionality k is the thermal conductivity of the material. In many practical
applications of fire safety engineering, the material thermal conductivity within the relevant temperature
range may be approximated as a constant. Thus, Equation 37.10 may be replaced by its finite difference
equivalent
T2  T1 Dx
Q_ ¼ k or T1  T2 ¼ Q_ ð37:11Þ
Dx k
where, refering to Figure 37.7, T1 and T2 are temperatures at the two sides of a material and Dx is the
material thickness. Dx/k expresses the thermal resistance of the material.

37.3.2 Convection
Heat convection and radiation are considered at the interface between a structural member and the fire
or the ambient temperature air. When applying thermal boundary conditions, it is often assumed that

Copyright 2005 by CRC Press


37-10 Handbook of Structural Engineering

the heat exchange between the fluid and the structural surface is related to the temperature difference at
the interface. Therefore, on the fire side from fire to structural surface
Q_ ¼ hfi ðTfi  Ts Þ ð37:12Þ
On the ambient temperature air side from structural surface to the ambient temperature air
Q_ ¼ ha ðTs  Ta Þ ð37:13Þ
where Tfi and Ta are the fire and air temperatures, respectively and Ts is the structural surface
temperature.
Quantities hfi and ha are the overall heat exchange coefficients on the fire and air side, respectively.
Depending on the relationship between the fire/air and the structural surface, the heat exchange
coefficients (hfi and ha) may only contain the convective component (hc), the radiant component (hr),
or both.
The convective heat transfer coefficient depends on many factors. However, for structural fire
applications, the structural temperatures are relatively insensitive to its exact values. Eurocode 1 Part 1.2
(CEN 2000a) recommends constant convective heat transfer coefficients as follows: on the fire side,
hc ¼ 25 W/m2 and on the air side, hc ¼ 10 W/m2.

37.3.3 Radiation
If a structural surface is in direct contact with fire/air, radiant heat transfer between the structural
surface and the fire/air may be assumed to occur between two very large parallel plates of area A, whose
distance apart is small compared with the size of the plates so that radiation at their edges is negligible.
Under this circumstance and assuming graybody radiant heat transfer, the radiant heat exchange
coefficient is

hr ¼ er sðT22 þ T12 ÞðT2 þ T1 Þ ð37:14Þ

where s is the Stefan–Boltzmann coefficient (¼ 5.876 108 W/(m K4)) and er is often referred to as the
resultant emissivity given by
1 e1 e2
er ¼ ¼ ð37:15Þ
1=e1 þ 1=e2  1 e1 þ e2  e1 e2

in which e1 and e2 are the graybody emissivities of the two surfaces, that is, that of the structural surface
and fire/air, respectively.

37.3.4 Some Simplified Solutions of Heat Transfer


General heat transfer problems are difficult to solve and will usually require the use of numerical heat
transfer procedures. However, for two common cases of unprotected and protected steelworks exposed
to fire attack, simple analytical solutions have been derived to enable their temperatures to be calculated
quickly. These simple analytical solutions have been derived by using the ‘‘lumped mass method,’’ that is,
the entire steel mass is given the same temperature. For unprotected steelwork
h As
DTs ¼ ðTfi  Ts Þ Dt ð37:16Þ
rs Cs V

where V and As are the volume and exposed surface area of the steel element, respectively, rs is the
density of steel, and Cs is the specific heat of steel. The ratio As/V in Equation 37.16 is often referred to
as the section factor of the steel element. Tfi and Ts are the fire and steel temperatures, respectively. h is
the total heat transfer coefficient between the fire and the steel surface, including both the convective
and radiant components. When using Equation 37.16, a step-by-step approach is necessary and the
time increment should be small (Dt < 5 s).

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-11

For protected steelwork


ðTfi  Ts ÞAs =V Cp rp As
DTs ¼ Dt  ðef=10  1ÞDTfi , where f ¼ tp ð37:17Þ
ðtp =kp ÞCs rs ð1 þ ð1=3ÞfÞ Cs rs V
Additional symbols in Equation 37.17 include tp, the fire protection thickness; kp, thermal conductivity
of the fire protection; Cp, specific heat of the fire protection; rp, density of the fire protection; and DTf,
the increment in fire temperature during the time interval Dt. The time increment should not be too
large. When using Equation 37.17, the time increment (Dt) should not exceed 30 s.
Because of the second term in Equation 37.17, it is possible that at the early stage of increasing fire
temperature, the increase in steel temperature (DTs) may be negative. In this case, the steel temperature
increase should be taken as zero.

37.3.5 Section Factors


Equations 37.16 and 37.17 clearly indicate that the temperature rise in a steel element is directly
related to the section factor As/V, that is, the ratio of the heated surface area to the volume of the steel
element. Consider a unit length of a steel element where the end effects are ignored, the section factor
may alternatively be expressed as Hp/A, where Hp is the fire exposed perimeter length of the steel cross-
section and A is the cross-sectional area of the steel element. Section factors for a few common types
of steel sections exposed to fire are given in Table 37.3.

EXAMPLE 37.2
Section factor
Calculate the section factor (Hp/A) for the two cases shown in Figure 37.8.

Calculation results
Case 1, Figure 37.8a
Hp ¼ 2 400 þ 150 3  2 10 ¼ 1230 mm,
A ¼ 2 15 150 þ ð400  15 2Þ 10 ¼ 8200 mm2
Hp =A ¼ 0:15 mm1 ¼ 150 m1
Case 2, Figure 37.8b
Hp ¼ 2pRo ¼ 300p, A ¼ pðRo 2  Ri 2 Þ ¼ 2900p
1
Hp =A ¼ 0:1034 mm ¼ 103:4 m1

37.3.6 Thermal Properties of Materials


In order to use Equations 37.16 and 37.17, it is necessary to have available information on the thermal
properties (thermal conductivity k, density r, and specific heat C) of steel and insulation materials.
37.3.6.1 Steel
The thermal properties of steel are known with reasonable accuracy and the following values are given
in Eurocode 3 Part 1.2 (CEN 2000b):
Density
rs ¼ 7850 kg/m3
Thermal conductivity [W/(m K)]
Ts
ks ¼ 54  for 20 C
Ts
800 C
300
ks ¼ 27:3 for Ts > 800 C

Copyright 2005 by CRC Press


37-12 Handbook of Structural Engineering

TABLE 37.3 Section Factors of a Steel Element


Fire exposure situation Ap/V

2ð2B  tw þ D Þ
Unprotected steel section exposed to fire exposure around all sides
As

2ðB þ D Þ
Fire exposure on all sides of board protection
As

2ðB  tw Þ þ B þ 2D
Fire protection on three sides: profile protection
As

2D þ B
Fire protection on three sides: board protection
As

Note: B ¼ section width, D ¼ steel depth, As ¼ cross-sectional area, tw ¼ web thickness.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-13

(a) (b)

Concrete slab

15 Ri = 140
10
Ro = 150
400
150

Case 1 All units in mm Case 2

FIGURE 37.8 Dimensions of steel cross-sections.

Specific heat [J/(kg K)]

Cs ¼ 425 þ 0:773Ts  0:00169Ts2 þ 2:22 106 Ts3 for 20 C


Ts
600 C
13, 002
Cs ¼ 666  for 600 C < Ts
735 C
Ts  738
17, 820
Cs ¼ 545  for 735 C < Ts
900 C
Ts  731
Cs ¼ 650 for Ts > 900 C

37.3.6.2 Insulation Materials


It is much more difficult to obtain information on the thermal properties of fire protection materials.
This is partly due to the specific nature of fire protection materials, which have complicated and variable
chemical reactions at high temperatures. An important factor contributing to this lack of information is
that most fire protection materials are proprietary systems from different manufacturers and com-
mercial sensitivity prevents publication of this type of information.
Information in Table 37.1 should only be used as a general guide for a few generic types of fire
protection material.
Intumescent coatings offer a number of advantages, for example, architectural appearance and the
possibility of offsite applications and are increasingly being used as fire protection materials to steel
structures. At present, there are no suitable simplified design equations for steel structures protected
with intumescent coatings. The designer has to rely on information provided by the intumescent coating
manufacturer from standard fire resistance tests.

EXAMPLE 37.3
Temperatures in unprotected steelwork
For the cross-section in Example 37.2, case 1, calculate the unprotected steel temperature under the
natural fire condition evaluated in Example 37.1. For steel, assume a constant density of 7850 kg/m3
and a constant specific heat of 650 J/(kg K). Also, assume a constant resultant emissivity of 0.5 and
a convective heat transfer coefficient of 25 W/(m2 K).

Calculation results
The calculations are performed for intervals of 5 s and results of only the first time increment are shown.
Equation 37.15 gives hr ¼ 2.95 W/(m2 K). The total heat transfer coefficient is 27.95 W/(m2 K).

Copyright 2005 by CRC Press


37-14 Handbook of Structural Engineering

After a time increment of Dt ¼ 5 s, Equation 37.3 in Example 37.1 gives a fire temperature of
41.74 C. The steel temperature is 20 C. The section factor from case 1 of Example 37.2 is 150 m1.
Equation 37.17 gives DTs ¼ 0.0616 C. Thus, the steel temperature after 5 s is 20.0616 C.
Figure 37.5 plots the unprotected steel temperature development. The maximum steel temperature is
829.25 C, reached just after 30 min.

EXAMPLE 37.4
Temperatures in protected steelwork
For the cross-section in case 2 of Example 37.2, calculate the protected steel temperature under
the natural fire condition obtained in Example 37.1. Fire protection is by sprayed lightweight concrete
whose thermal properties are given in Table 37.1. Assume the same steel thermal properties as in
Example 37.3.

Calculation results
The calculations are performed for intervals of 5 s and results of only the first time increment are shown.
Equation 37.17 gives f ¼ 1.129 and DTs ¼ 2.54 C. As pointed out in Section 37.3.4, this negative
number should be changed to 0.
Figure 37.5 plots the protected steel temperature development. The maximum steel temperature is
602.58 C, reached at 51.25 min.

37.4 Design of Structural Elements at


Elevated Temperatures
37.4.1 General
The remaining sections of this chapter will introduce the reader to design calculations to assess structural
stability at elevated temperatures. In Europe, the structural fire safety design of steel elements is covered
in ENV 1993-1-2 (CEN 2000b), commonly known as Eurocode 3 Part 1.2 (to be referred to as Eurocode 3
hereafter), composite steel/concrete elements in ENV 1994-1-2 (CEN 2001) or Eurocode 4 Part 1.2
(Eurocode 4), reinforced concrete structures in ENV 1992-1-2 (CEN 1996), or Eurocode 2 Part 1.2,
timber structures in ENV 1995-1-2 (CEN 2000c) or Eurocode 5 Part 1.2 and masonry structures in ENV
1996-1-2 (CEN 1997) or Eurocode 6 Part 1.2. At present, design calculations for steel and composite
steel/concrete structures are advanced. Design calculations for concrete and timber structures are
relatively brief and there is a significant lack of information for fire safety design of masonry structures.
Also, there is very little information available to enable fire safety design calculations for structures made
of more ‘‘specialist’’ construction materials such as glass, fiber reinforced plastics, aluminum. This
chapter will present detailed information to enable fire safety design calculations of steel and composite
steel/concrete structural elements. Structures using other materials will be not be dealt with in this
chapter due to a significant lack of information.
It should be borne in mind that fire is an accidental event and its coincidence with extreme structural
loading is rare. Fire attack is more likely to occur during normal use of a building. Therefore, for
structural fire safety design, the design structural loads should be those present during normal service
of a building, and further reduced to allow for occupant escape and combustible materials burning off.
Investigations into the collapse of the World Trade Center (FEMA 2002) have raised concern over the
reliability of fire protection materials. It is assumed in this chapter that a fire protection material is able to
fulfill its intended functions during the life of the protected structure and also during a fire exposure. This
may be ensured by making sure that the fire protection material can stick to the protected structure or by
limiting deflections of the structure. Also, fire protection materials may get damaged. At present, there are
very few studies of this problem and there is insufficient information to help develop a sensible simple

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-15

design guide to assess the acceptable extent of damage to fire protection materials. In the light of this, the
designer has to ensure that any damage to the fire protection material is repaired.
The structural fire safety design criterion is the same as at ambient temperature, that is, the residual
load carrying capacity of a structural member should not be lower than the applied load under the fire
condition.

37.4.2 Mechanical Properties of Steel and Concrete


at Elevated Temperatures
37.4.2.1 Steel
The stress–strain relationships of steel at elevated temperatures depend on whether steady state or
transient state testing is employed. In steady state testing, the material temperature is held at a constant
value and stress is changed. In transient state testing, stress is applied and the material temperature is
then changed. Transient state testing is preferred since it reflects the realistic situation of a structure in
fire, where structural loads are applied before fire exposure. In transient state testing, the rate of heating
has some influence due to creep strain. But since the steel creep strain is small, mechanical testing of steel
at elevated temperatures is usually carried out by using a typical heating rate of about 10 C/min as found
in realistic steel structures exposed to fire conditions.
In Eurocode 3, the stress–strain curve of steel consists of a straight line for the initial response,
followed by an elliptical relationship and then a plateau. Table 37.4 gives the mathematical descriptions
used in Eurocode 3 and Figure 37.9 provides an illustration of this model and shows various parameters
to be used in the mathematical model. In order to use this model, the reduced strength and stiffness of
steel at elevated temperatures are required as input data and Table 37.5 gives their values, expressed as
ratios of the values at elevated temperatures to that at ambient temperature. These ratios are often
referred to as retention factors.
The thermal expansion strain (eth) of steel is given by
eth ¼ 2:416 104 þ 1:2 105 T þ 0:4 108 T 2 for T
750 C
eth ¼ 0:011 for 750 < T
860 C
eth ¼ 0:0062 þ 2 105 T for T > 860 C

TABLE 37.4 Mathematical Model of the Stress–Strain


Relationship of Steel at Elevated Temperatures
Strain range Stress s

e
ep,T eET
rhffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
b  2 i
ep,T < e < ey,T fp,T cþ a2  ey,T  e
a
ey,T
e
et,T fy,T
fp,T
Parameters ep,T ¼ , ey,T ¼ 0.02, et,T ¼ 0.15
ET
 
  c
Functions a2 ¼ ey,T  ep,T ey,T  ep,T þ
ET
2
  2
b ¼ c ey,T  ep,T ET þ c
 2
fy,T  fp,T
c¼   
ey,T  ep,T ET  2 fy,T  fp,T

Source: European Committee for Standardisation (CEN),


2000b, Draft prEN 1993-1-2, Eurocode 3: Design of Steel Structures,
Part 1.2: General Rules, Structural Fire Design (London: British
Standards Institution).

Copyright 2005 by CRC Press


37-16 Handbook of Structural Engineering

Stress 

fy,T

fp,T

ET

p,T y,T t,T Strain 

FIGURE 37.9 Stress–strain relationship of hot-rolled steel at elevated temperatures.

TABLE 37.5 Retention Factors of Steel at Elevated Temperatures

Effective yield strength Proportional limit Slope of the linear


Steel temperature, (relative to fy at 20 C), (relative to fy at 20 C), elastic range (relative to
T ( C) ky,T ¼ fy,T=fy kp,T ¼ fp,T=fy Ea at 20 C), kE,T ¼ ET/Ea
20 1 1 1
100 1 1 1
200 1 0.807 0.9
300 1 0.613 0.8
400 1 0.42 0.7
500 0.78 0.36 0.6
600 0.47 0.18 0.31
700 0.23 0.075 0.13
800 0.11 0.050 0.09
900 0.06 0.0375 0.0675
1000 0.04 0.025 0.045
1100 0.02 0.0125 0.0225
1200 0 0 0

Note: The effective yield strength is defined at 2% strain.


Source: European Committee for Standardisation (CEN), 2000b, Draft prEN 1993-1-2, Eurocode 3: Design of
Steel Structures, Part 1.2: General Rules, Structural Fire Design (London: British Standards Institution).

In simple calculations, the coefficient of thermal expansion of steel may be assumed to be a constant
so that the incremental thermal expansion strain is given by

eth ¼ 14 105 DT

37.4.2.2 Concrete
37.4.2.2.1 Thermal Strains
The thermal strain of concrete is complex and is influenced by a number of factors. According to
Anderberg and Thelandersson (1976) and Khoury and coworkers (Khoury 1983; Khoury et al. 1986), the
thermal strain of concrete may be divided into thermal expansion strain, creep strain, and a stress
induced transient thermal strain. Interested readers should refer to the above references for more

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-17

Stress
c,T

fc,T A

Equation 37.19

B
cu,T crush,T Strain c,T

FIGURE 37.10 Stress–strain relationship of concrete at elevated temperatures.

detailed information on how to evaluate these different strain components. Eurocode 4 (CEN 2001)
takes a simple approach and gives the coefficient of thermal expansion of concrete as
eth ¼ 1:8 104 þ 9 106 T þ 2:3 1011 T 3 for 20 C
T
700 C;
eth ¼ 0:0014 for T > 700 C:

37.4.2.2.2 Stress–Strain Relationships


The mechanical properties of concrete are more variable than those of steel. Phan and Carino (1998,
2000) recently carried out a survey of mechanical properties of concrete (including high strength
concrete) at elevated temperatures. Khoury (1992) provided an explanation of the variability in concrete
mechanical properties at elevated temperatures. Values in Eurocode 4 (CEN 2001) may be regarded as
the lower bound values of different test results for normal strength concrete.
Figure 37.10 shows the Eurocode 4 model for the stress–strain relationship of concrete and definitions
of various parameters. The stress–strain relationship is divided into two parts: the ascending part and
the descending part.
The Eurocode 4 equation for the ascending part is
(  ,"   #)
ec,T ec,T 3
sc,T ¼ fc,T 3 2þ ð37:18Þ
ecu,T ecu,T

where sc,T, ec,T, fc,T, ecu,T, are, respectively, the stress, strain, peak stress, and strain at peak stress for
concrete at elevated temperature T.
From Equation 37.18, the initial Young’s modulus of concrete may be obtained from
3 fc,T
Ec,T ¼ ð37:19Þ
2 ecu,T
The descending part is a straight line, joining the peak point (A) with the point of concrete crush (B)
in Figure 37.10.
Values of fc,T, ecu,T, and ecrush,T, are required to determine the complete stress–strain relationship
of concrete at elevated temperatures. Table 37.6 gives their values recommended by Eurocode 4
(CEN 2001). This table also gives the retention factors for modulus of elasticity.

37.4.3 Design of Steel Elements


37.4.3.1 Steel Beams
In general, the load carrying capacity of a steel beam depends on the bending moment capacity of its
cross-section and its slenderness if lateral torsional buckling occurs. Unlike structural design at ambient

Copyright 2005 by CRC Press


37-18 Handbook of Structural Engineering

TABLE 37.6 Strength, Strain Limits, and Elastic Modulus of Normal Weight Concrete (NWC) and
Lightweight Concrete (LWT) at Elevated Temperatures
kc,T ¼ fc,T /fc,a kE,T ¼ ET/Ea
3 3
Temperature NWC LWC ecu,T 10 ecrush,T 10 NWC LWC
20 1 1 2.5 20.0 1 1
100 0.95 1 3.5 22.5 0.844 0.889
200 0.9 1 4.5 25.0 0.72 0.8
300 0.85 1 6.0 27.5 0.618 0.727
400 0.75 0.88 7.5 30.0 0.5 0.587
500 0.60 0.76 9.5 32.5 0.369 0.468
600 0.45 0.64 12.5 35.0 0.257 0.366
700 0.30 0.52 14.0 37.5 0.16 0.277
800 0.15 0.4 14.5 40.0 0.075 0.2
900 0.08 0.28 15.0 42.5 0.038 0.132
1000 0.04 0.16 15.0 45.0 0.018 0.071
1100 0.01 0.04 15.0 47.5 0 0.017
1200 0.0 0.0 15.0 50.0 0 0

Source: European Committee for Standardisation (CEN), 2001, prEN 1994-1-2, Eurocode 4: Design of Composite Steel
and Concrete Structures, Part 1.2: Structural Fire Design (London: British Standards Institution).

temperature, lateral torsional buckling of a steel beam is usually not a problem in fire. This is because
steel beams that are required to have fire resistance are floor beams whose compression flanges are
restrained by the floor slabs. Steel beams that should be checked for lateral torsional buckling at ambient
temperature, for example, roof beams, do not require fire resistance. Therefore, the following design
method will only consider the cross-sectional bending resistance of a beam.
Eurocode 3 gives two methods to calculate the plastic bending moment capacity of a steel beam.
The first method is the bending moment capacity method that is generally applicable to cross-sections
with nonuniform temperature distributions. In this method, the steel cross-section is divided into
a number of thin slices of approximately the same temperature. The plastic bending moment capacity
of the cross-section is calculated according to the reduced strengths of steel at the temperatures of
these slices.

EXAMPLE 37.5
Plastic bending moment capacity of a nonuniformly heated steel beam
An example is given in Table 37.7 to illustrate this method. Input information for this example are
shown in Figure 37.11.
As can be seen, the plastic bending moment capacity method requires many calculations. Hence, the
only benefit of using this more elaborate method is to explore possible benefits of nonuniform tem-
perature distribution in the cross-section of a beam, particularly to justify the use of unprotected
steelwork.
Instead of using the plastic bending moment method, Eurocode 3 also gives an alternative method that
is much simpler to use. In this simple method, the plastic bending moment capacity of a beam is given by
Mp,fi ¼ ky,T Mp =k1 ð37:20Þ

where Mp is the plastic bending moment capacity of the cross-section at ambient temperature; ky,T is the
retention factor for the effective yield strength of steel at the maximum temperature in the lower flange.
Thus, ky,TMp gives the reduced plastic bending moment capacity of the cross-section at a uniform
temperature T. The modification factor k1 is used to account for nonuniform temperature distribution
in the cross-section and k1 ¼ 0.7. For the above example, ky,TMp ¼ 137.5 kN m. Using Equation 37.20
gives a value of 196.5 kN m, which is close to that obtained using the more time-consuming plastic
bending moment capacity method.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-19

TABLE 37.7 An Example of Using the Bending Moment Capacity Method


Temperature, Resultant force, Lever arm from top, Bending
Part i Ti ( C) Fi ¼ py(Ti) Ai (kN) di (mm) moment ¼ Fi di (kN m)
Upper flange 650 0.8 ¼ 520 565.81 8 4.53
Web 1 528.13 86.08 39.8 3.43
Web 2 544.38 24.10 70.76  1.71
55.96 94.56 5.29
Web 3 560.63 73.93 135.0 9.98
Web 4 576.88 67.74 182.6 12.37
Web 5 593.13 61.56 230.2 14.17
Web 6 609.38 55.75 277.8 15.49
Web 7 625.63 50.21 325.4 16.34
Web 8 641.88 44.68 373.0 16.66
Lower flange 650.0 266.16 404.8 107.74
Total NA 0 NA 188.39

(a) (b) (c) Compression

16
T
di
PNA
Ti Fi
412.8
9.5

179.5 Tension

FIGURE 37.11 Input data for calculations in Table 37.7: (a) dimensions (mm), (b) temperature distribution, and
(c) cross-section layers.

37.4.3.2 Steel Columns


It is commonly assumed that a steel column is surrounded by fire on all sides so that the steel tem-
perature is uniform. Because the stress–strain relationships of steel at elevated temperatures are non-
linear and cannot be assumed to be elastic–perfectly plastic, calculations of the column resistance in fire
are slightly different from those at ambient temperature. In Eurocode 3, the column axial compressive
resistance is calculated from
Pc,fi ¼ wfi Pu,fi ð37:21Þ

where Pu,fi is the column squash load at elevated temperature T and is calculated from
Pu,fi ¼ As ky,T , fy ð37:22Þ

wfi is the column strength reduction factor to account for the column slenderness effect and is given by
1 1 fi þ l
2

wfi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi , with ffi ¼ 1 þ al fi ð37:23Þ
f þ f2   l2 2
fi fi fi
pffiffiffiffiffiffiffiffiffiffiffiffiffi
where a is an imperfection factor, a ¼ 0:65 fy =235.
The column slenderness in fire  lfi is calculated using
sffiffiffiffiffiffiffiffi
fi ¼ l
l  k y ,T ð37:24Þ
kE,T

Copyright 2005 by CRC Press


37-20 Handbook of Structural Engineering

where 
l is the column slenderness at ambient temperature. 
l is defined as
rffiffiffiffiffiffiffiffi
 ¼ l fy , with l ¼ Le
l ð37:25Þ
p2 E ry

where Le and ry are the column buckling length and radius of gyration of the column cross-section about
the relevant axis of buckling.

37.4.3.3 Steel Connections


The behavior of steel connections in fire is complicated due to complex temperature distributions in
connections and connection interactions with the adjacent structure. Fortunately, since a connection is
much ‘‘bulkier’’ than the connected members, its temperatures are much lower, and, if designed
properly, it is rarely the weak link in the structure. Current design methods require that when fire
protection materials are applied to a structure, the thickness of protection applied to a connection
should be based on the thickness required for whichever of the members jointed by the connection that
has the highest section factor Hp/A.
However, it should be pointed out that the above comments are related to steel structures in flexural
bending under fire conditions, where a connection does not have to resist any tensile force. High tensile
forces can develop in the connected steel beams to fracture connections, which was observed during
the Cardington fire research on the steel framed building (Newman et al. 2000) when the connections
were cooling, or found from investigations of the World Trade Center collapse (FEMA 2002) due to
possible development of catenary action in the floor truesses at very large deflections. If the designer
anticipates tensile forces to develop in the connected beams, it is important that their values are
quantified accurately and connections are designed to resist such tensile forces.

37.4.3.4 Other Types of Steel Structures


37.4.3.4.1 Stainless Steel Structures
Due to architectural demand and superior corrosion resistance, stainless steels are becoming more
widely used. Although their fire resistance is only a minor factor in determining whether to use stainless
steel or not, stainless steel does have superior fire resistance to conventional carbon steels.
Baddoo (1999) gives some information on the strength retention factors of stainless steels at different
elevated temperatures. Whilst conventional carbon steel loses about 50% of its strength at a temperature
of around 600 C, the temperature that gives the same loss in the strength of stainless steel is much
higher, at about 800 C. Also, the surface of stainless steel has a much higher reflectivity, hence the
emmissivity (1  reflectivity) of stainless steel is much lower. Typically, the emmissivity of stainless steel
is about 0.3 to 0.4, compared to about 0.8 for carbon steel. This gives a much lower temperature in a
stainless steel structure in fire. For realistic loading conditions, it is almost certain that stainless steel
structures will be able to be engineered to provide sufficient fire resistance without the need for fire
protection.
With suitable modifications to take into consideration the reduced strength and stiffness of stainless
steel at elevated temperatures, the same design method for carbon steel may be extended to stainless steel
structures (Baddoo and Burgen 1998).

37.4.3.4.2 Portal Frames


Portal frames are usually single storey buildings with a small density of occupants. Escape in the case of
fire attack is relatively easy. Therefore, fire safety of a portal frame only becomes a requirement when the
portal frame is adjacent to another building and it is necessary to prevent fire spread from the portal
frame building to the adjacent building. Fire spread from a portal frame building is usually through
collapsed walls; consequently, the structural safety requirement for a portal frame under fire attack is to
ensure that the portal frame columns in the walls remain upright so that the walls are stable. The portal
frame girders may be allowed to collapse.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-21

The Steel Construction Institute in the United Kingdom has developed a design guide (Newman 1990)
for portal frames in fire. Using this guide, portal frames are usually designed without fire protection.

37.4.3.4.3 Water Cooled Structures


The design of water cooled structures (Bond 1975) relies on the principle that boiling water has a very high
value of convective heat transfer coefficient, some 3 or 4 orders higher than the convective heat transfer
coefficient of air. If boiling water is replenished, heat on a water cooled steel structure is taken away and
steel temperatures remain low. Typically, steel temperatures in a water cooled structure do not exceed
150 C. Therefore, structural stability of a water cooled structure is rarely a design issue. Design is mainly
concerned with hydraulic calculations to ensure sufficient water supply and circulation in case of fire.
Water cooling a steel structure to achieve fire protection is expensive and because of this it is rarely
used solely for the purpose of fire protection. It is usually combined with other functions. An excellent
example (Bressington 1997) of recent application of this technique is in the roof truss of the cargo
handling facility of Hong Kong Air Cargo Terminals Ltd. The steel structural roof truss is made of
circular hollow sections and is used as the water distribution pipe for sprinklers. On operation of the
sprinklers in the event of a fire, internal water flow through the steelwork members also provides
sufficient cooling.

37.4.3.4.4 External Steelwork


In the case of a building enclosure fire, fire exposure on the external steelwork differs from that on the
interior steelwork in two ways:
 The fire temperature to the external steelwork is much lower than that to the interior steelwork.
 The external steelwork may not be directly engulfed in fire.
These two differences ensure that temperatures in the external steelwork are kept lower than their failure
temperatures so that fire protection is not necessary.
Law and O’Brien (1989) developed a design method for external steelwork and the design method in
Eurocode 3 is based on their work.

37.4.4 Composite Steel/Concrete Members


37.4.4.1 Composite Slabs
Composite slabs are constructed from reinforced concrete slabs in composite action with steel decking
underneath. The steel decking acts as support to the concrete during construction and is generally
profiled to maximize structural efficiency. Composite slabs usually form the floor of a fire resistant
compartment. Hence, they should meet all the requirements of fire resistant construction; that is, in
addition to sufficient load bearing resistance, they should also have adequate insulation and maintain
their integrity during fire attack.
Composite floor slabs are noncombustible and will not suffer integrity failure by burning through.
However, the problem of integrity failure may occur at the junctions between a composite slab and other
construction elements. Particular attention should be paid to the slab edges where large cracks may occur
due to large rotations. It is important that reinforcement bars should be made continuous over the
supports.
To check whether a slab can fulfill the insulation requirement, it is necessary to carry out a
heat transfer analysis to determine temperatures on the unexposed surface of the slab. Results of this
temperature analysis can then be used to determine the minimum slab thickness above which the
unexposed surface temperature is unlikely to exceed the allowed values. For most applications where
the required standard fire resistance rating does not exceed 90 min, the required minimum slab
thickness will almost certainly be less than that required by other functions such as control of
deflections. Hence, the insulation requirement for composite slabs is very rarely a problem in fire
resistant design.

Copyright 2005 by CRC Press


37-22 Handbook of Structural Engineering

Uniformly distributed load

M– ,fi M– ,fi
L L L

M+ ,fi M+ ,fi M+ ,fi

FIGURE 37.12 A continuous slab exposed to fire.

37.4.4.1.1 Load Bearing Capacity of One-Way Spanning Composite Slabs


When calculating the load bearing resistance of a composite slab in fire, it is often assumed that it is
one-way spanning, being effective only in the direction of the concrete rib. For a continuous composite
slab, the plastic design method may be used, in which plastic hinges are assumed to form at the supports
and locations of the maximum bending moment in the span of the slab. For example, for the interior
span of a continuous slab under uniformly distributed load, as shown in Figure 37.12,
1
Mþ,fi þ M,fi Mfi, max , i:e:, Mþ,fi þ M,fi wL2 ð37:26Þ
8
where M þ,fi and M ,fi are the sagging and hogging bending moment resistance of the slab, respectively,
and Mfi,max is the maximum free bending moment in the slab under fire conditions.
For the end span of a continuous slab with uniformly distributed load, as shown in Figure 37.12,

1 1 M2
Mþ,fi þ M,fi ¼ wL2 þ 2 ,fi ð37:27Þ
2 8 2L w
or approximately

1
Mþ,fi þ 0:45M,fi wL2 ð37:28Þ
8

In order to determine the slab load carrying capacity in fire, the sagging bending moment capacity
M þ,fi and hogging bending moment capacity M ,fi should be evaluated.
When calculating the sagging bending moment capacity of a slab, the reinforcement near the fire side
is in tension and the compressive concrete is on the unexposed side of the slab. Since the temperature
rise on the unexposed side of the slab is required to be below 140 C to fulfill the insulation requirement,
the concrete in compression can be assumed to be cold and its cold strength may be used when
calculating the sagging bending moment capacity of the slab. Contributions from the steel decking are
usually ignored because the decking will be unprotected and may debond because it is under direct fire
attack. Figure 37.13 shows the calculation procedure. In Figure 37.13, fc is the design strength of concrete
in bending at ambient temperature, Ar, py,r, and ky,r(T ) are the area, design strength at ambient
temperature, and strength retention factor of the reinforcement at temperature T, and ky,r(T ) may be
obtained from Figure 37.14.
Under a hogging bending moment, the compression face of a composite slab is exposed to fire where
there is a very steep temperature gradient. When calculating the hogging bending moment capacity, the
composite slab, including the concrete in the ribs, should be divided into a number of layers each of
approximately constant temperature. The contribution of each layer should be evaluated separately and
then integrated to give the total slab resistance.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-23

fc

dc C = Fr
D M+,fi = Fr × d
d = D – 0.5 × dc = D – 0.5 × Fr /f c

Fr = Ar × ky,r(T) × fy,r

FIGURE 37.13 Calculation method for sagging bending moment capacity.

1
0.9
0.8
Effective yield strength
Retention factor

0.7
0.6
0.5
0.4
0.3 Elastic modulus
0.2
0.1
0
0 200 400 600 800 1000 1200
Temperature, °C

FIGURE 37.14 Retention factors of cold worked reinforcing steel (from European Committee for Standardisation
(CEN), 2001, prEN 1994-1-2, Eurocode 4: Design of Composite steel and Concrete Structures, Part 1.2: Structural Fire
Design (London: British Standards Institution)).

37.4.4.1.2 Load Bearing Capacity of Two-Way Spanning Slabs


The load carrying capacity of a composite slab in one-way spanning is usually sufficient under
fire conditions. However, in some cases, it may be necessary to utilize the slab strength in two-way
spanning. This is particularly the case when it is necessary to justify the elimination of fire protection from
some of the slab-supporting steel beams. The yield line analysis (Johansen 1962) for reinforced concrete
slabs may be used to give a safe estimate of the slab load carrying capacity in two-way spanning.
The real benefit of utilizing the strength of a slab in two-way spanning is the possibility of using
tensile membrane action in the slab, under which the strength of the slab can be many times higher
than that given by yield line analysis. Recently, a design method has been developed to use tensile
membrane action in steel framed buildings to eliminate fire protection to some steel beams.
For detailed design equations and examples, reference should be made to the two papers by Bailey
and Moore (2000a, 2000b) and a publication by the United Kingdom’s Steel Construction Institute
(Newman et al. 2000).

37.4.4.2 Composite Beams


For a conventional composite beam with concrete slabs on top of the steel section, the sagging
bending moment capacity of the composite cross-section may be calculated by the bending moment
capacity method, similar to steel beams (Section 37.4.3.1). The concrete in compression may be assumed
to be cold and the steel temperatures may be calculated using Equations 37.16 or 37.17. If the steel
section is protected, it may be assumed to have a uniform temperature distribution and its section
factor Ap/V is that of the entire section, calculated according to Table 37.3. If the steel section

Copyright 2005 by CRC Press


37-24 Handbook of Structural Engineering

is unprotected, the steel section will have a nonuniform temperature distribution. For temperature
calculations, the steel section may approximately be divided into two parts: the upper flange and the
lower flange plus the web.

EXAMPLE 37.6
Plastic bending moment capacity of a composite cross-section
Figure 37.15 shows a composite cross-section exposed to fire underneath the slab. The steel section is
unprotected and fire exposure is according to the temperature–time relationship in Example 37.1.
At ambient temperature, the design strength of steel is 275 N/mm2 and the design compression strength
of concrete is 20 N/mm2. Calculate the minimum sagging bending moment capacity of the composite
cross-section in fire.

Calculation results
According to the dimensions in Figure 37.15, the section factor of the top flange is 80 m1 and that of
the web/bottom flange is 147 m1.
Following calculations in Example 37.2, the maximum top flange temperature is 802.8 C reached at
32.25 min and the maximum lower flange/web temperature is 847.2 C reached at 28.75 min. It is
interesting to notice that under natural fire exposure, different parts of a structural member will reach
their maximum temperatures at different times. Therefore, in theory, calculations of the plastic bending
moment capacity of the cross-section should be performed as a function of time. For simplicity, in this
example, the maximum temperatures of the top flange and bottom flange/web, attained at different
times, are used.
From Table 37.5, the residual steel strengths and tensile capacity of the steel cross-section are
Top flange: fy ¼ 0.1086 275 ¼ 29.9 N/mm2, As ¼ 2250 mm2, Nuf ¼ 67.3 kN
Bottom flange/web: fy ¼ 0.08639 275 ¼ 23.76 N/mm2
The lower flange tension resistance: Nlf ¼ 53.5 kN
The web area is 3700 mm2, giving the web tension resistance: Nwf ¼ 87.9 kN
The total tensile resistance of the steel cross-section: Ns ¼ Nuf þ Nwf þ Nlf ¼ 208.7
Assume the top of the concrete slab is cold. The depth of concrete in compression is 208.7 1000/
(20 2000) ¼ 5.22 mm.

2000

Concrete slab 130

15

10
400

150 15

Units in mm

FIGURE 37.15 Composite cross-section dimensions for Example 37.6.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-25

Taking moment about the top surface of the slab, the residual sagging bending moment capacity of the
composite cross-section is
Mf ¼ ½208:7 5:22=2 þ 67:3 ð130 þ 15=2Þ þ 87:9 ð130 þ 400=2Þ
þ 53:5 ð130 þ 400  15=2Þ=1000
¼ 65:7 kN m

37.4.4.3 Composite Columns


37.4.4.3.1 Resistance to Axial Load According to Eurocode 4
The general equation for calculating the squash load of a composite column is
X X X
Pu,fi ¼ ðAi fi Þc þ ðAi fi Þs þ ðAi fi Þr ð37:29Þ

where fi is the design strength of the ith layer and subscripts ‘‘c,’’ ‘‘s,’’ and ‘‘r’’ represent concrete,
steel, and reinforcement, respectively. Due to nonuniform temperature distribution, each component
of the composite cross-section is divided into a number of layers of approximately the same temperature.
Similarly, the rigidity (EI ) of the composite cross-section is calculated using the following equation:
X X X
ðEIÞfi ¼ ðEIÞi,s þ 0:8ðEIÞi,c þ ðEIÞi,r ð37:30Þ
where symbols E and I are the initial modulus of elasticity and second moment of area of the
appropriate component material about the relevant axis of buckling of the entire composite cross-
section, respectively.
The composite column compression resistance in fire is given by
w
Pc,fi ¼ fi Pu,fi ð37:31Þ
1:2
where the compression strength reduction factor in fire is calculated by
1 h   2
i
wfi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi with ffi ¼ 0:5 1 þ a lfi  0:2 þ lfi ð37:32Þ
ffi þ f2fi   l2fi

in which the initial imperfection factor a has a value of 0.49.


The column slenderness factor  lfi in fire is defined by
sffiffiffiffiffiffiffiffiffi
 Pu,fi
lfi ¼ ð37:33Þ
Pcr,fi

where the Euler buckling load in fire is calculated using


p2 ðEIÞfi
Pcr,fi ¼ ð37:34Þ
Le2
in which Le is the column effective length.

37.4.4.3.2 Simplified Temperature Calculation Method for Unprotected


Concrete Filled Columns
The temperature calculation method is based on the method of Lawson and Newman (1996) with
modification by Wang (2000). This method assumes that the composite column is unprotected.
In this method, the steel shell temperature is calculated by
Ts ¼ C2 Tfi ð37:35Þ
where Tfi is the standard fire temperature and C2 is a multiplication factor depending on the fire
resistance time. C2 is given by
120  FR
C2 ¼ 1  0:02t , but C2
1:0 ð37:36Þ
120

Copyright 2005 by CRC Press


37-26 Handbook of Structural Engineering

800 For fire rating of:


120 minutes
700 90 minutes
600 60 minutes T
30 minutes x
Temperature, °C

500
400
300
200
100
0
0 20 40 60 80 100
Depth from bottom of the slab x, mm

FIGURE 37.16 Temperatures of a concrete slab exposed to the standard fire from underneath (from European
Committee for Standardisation (CEN), 2001, prEN 1994-1-2, Eurocode 4: Design of Composite steel and Concrete
Structures, Part 1.2: Structural Fire Design (London: British Standards Institution)).

TABLE 37.8 Multiplication Factor C1

Diameter or size of Distance of center of layer from out surface (mm)


square section (mm) 10 30 50 70 >70
200 1.08 1.22 1.41 1.60 1.80
300 1.05 1.14 1.22 1.36 1.50
400 1.03 1.09 1.18 1.25 1.35
500 1.02 1.07 1.12 1.18 1.25

Source: Lawson, R.M. and Newman, G.M., 1996, Structural fire design to EC3 & EC4, and comparison with BS 5950,
Technical Report, SCI Publication 159, The Steel Construction Institute.

where FR is the fire resistance rating in minutes and t is the thickness of the steel shell in mm. The
concrete temperature is calculated by

Tc ¼ C1 C2 Tslab ð37:37Þ

where C1 is a multiplication factor depending on the composite section size and location of the
concrete (and is independent of the standard fire resistance time) and Tslab is the temperature in an
infinitely wide concrete slab exposed to fire on one side (given in Figure 37.16). Values of C1 are given in
Table 37.8.
If reinforcement is used, the reinforcement temperature should be taken as that of the concrete at the
same location. Furthermore, for reinforcement in the corners of a square section, due to heating from
two sides of the composite section, the reinforcement temperature should be calculated from an
equivalent depth of half the concrete cover depth.

EXAMPLE 37.7
Compression resistance of a composite column
Figure 37.17 shows the dimensions of a concrete filled circular steel section. The effective length of the
column is 4 m. Calculate the column compression resistance for a standard fire resistance period of
60 min. At ambient temperature, steel has a design strength of 275 N/mm2 and Young’s modulus
of 205,000 N/mm2. Concrete has a design compression strength of 40 N/mm2 and modulus of elasticity
of 20,000 N/mm2.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-27

Circular hollow section


Outside diameter 300 mm
Wall thickness 20 mm

Concrete
inside

FIGURE 37.17 Column cross-section size for Example 37.7.

TABLE 37.9 Calculation Results for Example 37.7

Radius Tslab Tzone A I fi Ei (fA)I (EI)I


Zone (mm) ( C) ( C) (mm2) (cm4) (N/mm2) (N/mm2) (kN) (kN m2)
Steel 130–150 — 756 17,593 17,329 44.77 22,058 787.6 3322.4
Con. 110–130 421 480 15,080 10,933 25.2 7,875 380 861
Con. 90–110 250 305 12,566 6,346 33.84 12,250 425 777.4
Con. 70–90 150 204 10,053 3,267 35.92 14,310 361 467.5
Con. 0–70 130 195 15,394 1,886 36.1 14,512 556 273.7

Calculation results
Divide the composite cross-section into steel tube, three concrete rings of equal thickness of 20 mm, and
a concrete core of 70 mm radius. Table 37.9 gives, for each zone of the composite cross-section, tem-
perature (in  C, calculated using Equations 37.35–37.37), area (mm2), second moment of area about a
principle axis of the entire cross-section (cm4), reduced design strength (in N/mm2), and Young’s
modulus (in N/mm2, from Table 37.5 for steel and Table 37.6 for concrete) at elevated temperatures,
compression resistance (in kN) and rigidity (EI, in kN m2).
Equation 37.29 gives Pu,fi ¼ 2509.6 kN. Equation 37.34 gives the Euler buckling load in fire
Pcr,fi ¼ 5226.1 kN, giving a column slenderness in fire 
lfi ¼ 0:7885. Equation 37.32 gives wfi ¼ 0.563 and
Equation 37.31 gives the column compression resistance in fire Pc,fi ¼ 1413 kN.
37.4.4.3.3 High Strength Concrete Filled Columns
With the introduction of high strength concrete, the load carrying capacity of a concrete filled column
can be further enhanced. However, the increase in fire resistance is relatively small because high strength
concrete loses its strength at a much lower temperature than normal strength concrete. By adding a small
amount of steel fibre to the concrete, the elevated temperature performance of high strength concrete
can be much improved and the performance of fibre reinforced high strength concrete filled steel
columns is similar to that with normal strength concrete filling (Kodur and Wang 2001). Provided the
strength and stiffness retention factors are available, Equations 37.29 and 37.30 can also be used.

37.5 Design for Unprotected Steelwork


The quest for knowledge is one of the main drivers of research to investigate the behavior of steel and
composite structures under fire conditions. However, it should be recognized that the desire to reduce or

Copyright 2005 by CRC Press


37-28 Handbook of Structural Engineering

eliminate fire protection to steelwork is an equally strong incentive to carry out these studies. Fire
protection to steelwork can represent a significant part of the total steel structural cost and the elim-
ination of fire protection to steelwork represents a significant saving in construction cost to the client.
But more importantly, by reducing the use of fire protection, steel becomes more competitive and the
steel industry can benefit from an increased market share. After the September 11 event, it is also
appropriate to consider using unprotected steelwork in fire situations for safety. Without fire protection,
there would be no problem related to possible unreliable use of fire protection materials.
There are a number of ways of designing for unprotected steelwork, including risk assessment to
reduce the requirement of fire resistance, using the so-called fire resistant steel (Sakumoto 1998) to
increase the strength of steel at elevated temperatures, over-design steel elements at ambient temperature
so as to increase their reserve of strength in fire, integrating the functions of fire protection and structural
load bearing of concrete, and utilizing advanced structural behavior. Wang and Kodur (2000) provide
a summary of these techniques. This section will give a brief introduction to the last two methods
because they can be readily implemented in practice.

37.5.1 Integration of Structural Load Bearing and Fire Protection


Functions of Concrete
It should be appreciated that it is very rare for steelwork to be used alone. Steel is usually used in
combination with other materials, in particular with concrete. Concrete is not only a structural material,
it also has good thermal insulation properties. Therefore, by combining these two functions of concrete,
composite structures may be constructed to give inherently high fire resistance. The systems that will be
described below have made special considerations of fire resistance in their design and construction.
The following paragraphs will give a short description of their main features and the inherent standard
fire resistance that they can achieve. This should enable the designer to determine quickly a possible
structural load bearing system where the main design concern is to use unprotected steelwork. More
detailed information may be found in Bailey and Newman (1998).
37.5.1.1 Beams
Three types of construction may be used:
1. Slim floor/asymmetric beam, shown in Figures 37.18a and b. In the slim floor construction, a
wide plate is welded to the bottom flange of a universal column section and composite floor slabs
are supported on the wide plate. In an asymmetric steel beam, the bottom flange is rolled wider
than the top flange. Both systems use the same principle to achieve unprotected steelwork: the

(a) (b)

(c) (d)

FIGURE 37.18 Steel/composite beams of high fire resistance: (a) slim floor beam, (b) asymmetrical beam, (c) shelf
angle beam, and (d) partially encased steel beam.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-29

web of the steel section is protected by the concrete and provides the majority of the bending
resistance of the steel beam at elevated temperatures. Only the steel section is assumed to have
load carrying capacity, but lateral torsional buckling is prevented by the concrete slabs.
2. Shelf angle beams, shown in Figure 37.18c. In this system, steel angles are welded to the web of a
steel beam and these angles are used to support precast concrete floor units. This system is
mainly used to reduce the structural depth of the floor. Since the angles, the upper flange, and the
upper portion of the web of the steel section are shielded from fire exposure, 60 min of fire
resistance can be achieved using this system without fire protection.
3. Partially encased beams, shown in Figure 37.18d. By casting concrete in between the flanges of a
regular universal beam section, only the downward side of the lower flange will be exposed to
fire. Both the web and the upper flange are shielded from fire exposure and can provide high
structural resistance. Composite floor slabs may be connected to the top of the partially encased
steel beam via shear connectors to obtain composite action. Since concrete is cast between
flanges of the steel section, no temporary formwork is necessary. By using reinforcement,
standard fire resistance of up to 3 h can be obtained without fire protection to the steelwork.
Table 37.10 summarizes the standard fire resistance rating that can be achieved by different types of
unprotected steel beams.
37.5.1.2 Columns
Three types of unprotected columns may be used:
1. Columns with blocked-in webs as shown in Figure 37.19a. In this construction, lightweight
aerated concrete blocks are placed between the flanges of a universal steel section. The aerated
concrete blocks not only provide good insulation to the column web, they also reduce the
average column flange temperature compared to a bare steel column. A standard fire resistance
rating of 30 min can be achieved without additional fire protection.
2. Partially encased steel columns with unreinforced and reinforced concrete as shown in Figure
37.19b. In a column with blocked-in web, the lightweight aerated concrete only provides
insulation to the steel section and the system cannot provide 60 min fire resistance. If normal

TABLE 37.10 Standard Fire Resistance Rating of Unprotected Steel Beams


Standard fire resistance
Type of construction time (min)
Bare universal beams 15
Slim floor/asymmetrical beams 60
Shelf angle beams 60
Partially encased beams >60

Source: Bailey, C.G. and Newman, G.N., 1998, The design of steel framed
building without applied fire protection, The Structural Engineer, 76(5), 77–81.

(a) (b) (c)

FIGURE 37.19 Steel/composite columns of high fire resistance: (a) blocked-in web, (b) partially encased, and
(c) concrete filled.

Copyright 2005 by CRC Press


37-30 Handbook of Structural Engineering

TABLE 37.11 Standard Fire Resistance Rating of Unprotected Steel Columns


Standard fire resistance
Type of column time (min)
Universal column 15
Blocked-in column 30
Partially encased with unreinforced concrete 60
Partially encased with reinforced concrete >60
Concrete filled hollow section without reinforcement 60
Concrete filled hollow section with reinforcement >60

Source: Bailey, C.G. and Newman, G.N., 1998, The design of steel framed building
without applied fire protection, The Structural Engineer, 76(5), 77–81.

strength concrete is used to provide composite action, much higher fire resistance can be
obtained. If unreinforced concrete is used, 60 min of fire resistance can be obtained.
Reinforcement may be used to give much higher fire resistance.
3. Concrete filled hollow steel sections as shown in Figure 37.19c. Concrete filling of hollow steel
sections is a very practical solution to form composite columns. Either unreinforced or
reinforced concrete may be used. This type of column has been described in some detail in
Section 37.4.4.3. To summarize, unreinforced concrete filled columns can achieve 60 min of fire
resistance. If reinforcement is used, much higher fire resistance may be obtained.
To summarize, Table 37.11 gives the standard fire resistance time that can be achieved by different
types of unprotected columns.
The usefulness of Table 37.10 and Table 37.11 is to enable readers to reach a decision quickly on the
form of construction to achieve the required standard fire resistance without fire protection.

37.5.2 Utilizing Whole Building Performance in Fire


It has long been recognized that the behavior of a whole building in fire is much better than that of its
individual members. By achieving a better understanding of whole building behavior, it is possible to
achieve the objective of eliminating fire protection in conventionally designed and constructed steel
framed buildings. Two methods based on the whole building behavior may be considered.

37.5.2.1 Utilization of Structural Redundancy


Up to now, the design of steel structures for fire safety has generally been based on the assessment of
individual structural members, that is, each structural member should achieve the required fire resis-
tance. However, it should be realized that for a building to remain stable under fire conditions, serving
the principal need of containing a fire and preventing its spread, it is not absolutely necessary for every
individual structural member to remain stable. Fire protection may be eliminated for some steel
members if, in the absence of these members, an alternative load path due to structural redundancy can
be developed to retain structural stability.
For example, in a multistory steel framed structure, floor loads are transferred by floor slabs to the
supporting steel beams and the beam reactions are then transmitted to the supporting steel columns and
thence to the foundations. For fire design, it is usually assumed that the load path selected for the
ambient temperature design remains unchanged under fire conditions and that each member in this load
path has to have sufficient fire resistance.
It is now appreciated that the load transmission path in a structure is not a fixed one. If one load path
breaks down, other alternative load paths may exist and safely transfer the applied loads to the foun-
dation. For example, Figure 37.20a shows part of a steel framed structure and the usual load carrying
path adopted in the ambient temperature design. However, if the floor slabs are designed to have higher
load bearing resistance than required at ambient temperature, it is quite possible for the alternative load
carrying sequence in Figure 37.20b to develop in fire.

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-31

(a)

Load path
Slabs ⇒ secondary beams
Secondary beams ⇒ primary beams
Primary beams ⇒ columns

(b)

Some (dotted lines) secondary


beams have failed. Remaining
beams become edge beams

Load path
Slabs (two way) ⇒ edge beams
Edge beams ⇒ columns

FIGURE 37.20 An example of alternative load paths in a structure: (a) load path of ambient temperature and
(b) possible load path in fire.

At ambient temperature (Figure 37.20a), the secondary beams are needed to control excessive slab
deflections. Under fire conditions, applied floor loads are reduced and large slab deflections are
permissible. Thus, failure of some secondary beams is permissible provided a sufficient network of
beams remain available to keep transfer the slab load to the columns. A possible system of this type is
shown in Figure 37.20b. In Figure 37.20b, fire protection for the dotted secondary beams is not required.
Of course, the design load carrying capacity of the slab can be further increased even to bypass some
main steel beams.

37.5.2.2 Developing Better Understanding of Structural Behavior in Fire


Structural fire safety design has been developed based on extending current structural design methods at
ambient temperature to elevated temperatures. At ambient temperature, due to the need to avoid large
deflections, design methods are based on flexural behavior at small deflections. These current design
methods can be very conservative and structural members may have much higher strength if large
deflections are allowed. By exploiting this enhanced load bearing capacity, fire protection may be
eliminated. An example is the utilization of tensile membrane action in reinforced composite slabs in
steel framed buildings (Newman et al. 2000).

37.6 Concluding Remarks


This chapter has given an introduction to different aspects of structural fire safety design, including
evaluation of the severity of fire exposure, calculations of temperatures in structures exposed to fire, and
assessment of load carrying capacities of structures at elevated temperatures. An important emphasis of
this chapter is structural fire safety design from first principles, rather than following the prescriptive
route. Materials presented in this chapter are based on developments in Europe. However, the funda-
mental knowledge should be equally applicable in other situations. A number of examples have been

Copyright 2005 by CRC Press


37-32 Handbook of Structural Engineering

provided to illustrate the applications of various specific recommendations. There have been tre-
mendous advances in fire safety design of steel and composite steel/concrete structures such that these
types of structures are now increasingly being fire engineered to improve design. However, there are still
large scopes for research and development in structural fire safety, especially with regard to other types of
structures, such as concrete, timber, masonry. It is hoped that this chapter will arouse the interest of the
reader in this subject. Structural fire safety design is a relatively new and evolving subject. It is expected
that future editions of this chapter will provide more comprehensive coverage of this important topic of
structural engineering.

References
American Society for Testing and Materials (ASTM), 1985, ASTM E 119-83, Standard Methods of Fire
Tests of Building Construction and Materials, ASTM, Philadelphia.
Anderberg, Y. and Thelandersson, S., 1976, Stress and deformation characteristics of concrete at high
temperatures, 2: Experimental investigation and material behaviour model, Division of Structural
Mechanics and Construction, Lund Institute of Technology.
Baddoo, N. and Burgan, B.A., 1998, Fire resistant design of austenitic structural stainless steel, J. Constr.
Steel Res., 46(1:3), CD-Rom, paper No. 243.
Baddoo, N., 1999, Stainless steel in fire, Struct. Eng., 77(19), 16–17.
Bailey, C.G. and Newman, G.N., 1998, The design of steel framed buildings without applied fire pro-
tection, Struct. Eng., 76(5), 77–81.
Bailey, C.G. and Moore, D.B., 2000a, The structural behaviour of steel frames with composite floor slabs
subject to fire, Part 1: Theory, Struct. Eng., 78(11), 19–27.
Bailey, C.G. and Moore, D.B., 2000b, The structural behaviour of steel frames with composite floor slabs
subject to fire, Part 2: Design, Struct. Eng., 78(11), 28–33.
Bond, G.V.L., 1975, Fire and Steel Construction: Water Cooled Hollow Columns (London: Constrado).
Bressington, P., 1997, Integrated active fire protection: two systems for the price of one, Building Eng.,
June, 10–13.
British Standards Institution (BSI), 1987, British Standard 476, Fire Tests on Building Materials and
Structures, Part 20: Method for Determination of the Fire Resistance of Elements of Construction
(General Principles) (London: British Standards Institution).
British Standards Institution (BSI), 2001, Draft BS 9999, Code of Practice for Fire Safety in the Design,
Construction and Use of Buildings (London: British Standards Institution).
Conseil International du Batiment (CIB), 1986, Design guide — structural fire safety, Fire Safety J., 10,
79–137.
Drysdale, D., 1999, An Introduction to Fire Dynamics (2nd edition) (Chichester: John Wiley & Sons).
European Committee for Standardisation (CEN), 1996, ENV 1992-1-2, Eurocode 2: Design of Concrete
Structures, Part 1.2: General Rules, Structural Fire Design (London: British Standards Institution).
European Committee for Standardisation (CEN), 1997, ENV 1996-1-2, Eurocode 6: Design of Masonry
Structures, Part 1.2: General Rules, Structural Fire Design (London: British Standards Institution).
European Committee for Standardisation (CEN), 2000a, Draft prEN 1991-1-2, Eurocode 1: Basis of
Design and Actions on Structures, Part 1.2: Actions on Structures — Actions on Structures Exposed
to Fire (London: British Standards Institution).
European Committee for Standardisation (CEN), 2000b, Draft prEN 1993-1-2, Eurocode 3: Design
of Steel Structures, Part 1.2: General Rules, Structural Fire Design (London: British Standards
Institution).
European Committee for Standardisation (CEN), 2000c, ENV 1995-1-2, Eurocode 5: Design of Timber
Structures, Part 1.2: General Rules, Structural Fire Design (London: British Standards Institution).
European Committee for Standardisation (CEN), 2001, prEN 1994-1-2, Eurocode 4: Design of Composite
Steel and Concrete Structures, Part 1.2: Structural Fire Design (London: British Standards
Institution).

Copyright 2005 by CRC Press


Structural Design for Fire Safety 37-33

FEMA, 2002, World Trade Center Building Performance Study: Data Collection, Preliminary Observations,
and Recommendations, Federal Emergency Management Agency.
International Standards Organization (ISO), 1975, ISO 834: Fire Resistance Tests, Elements of Building
Construction (Geneva: International Organization for Standardization).
Johansen, K.W., 1962, Yield Line Theory (London: Cement and Concrete Association).
Karlsson, B. and Quantiere, J.G., 2000, Enclosure Fire Dynamics (Boca Raton, FL: CRC Press).
Khoury, G.A., 1983, Transient Thermal Creep of Nuclear Reactor Pressure Vessel Type Concretes,
PhD Thesis, Department of Civil Engineering, Imperial College of Science, Technology and
Medicine.
Khoury, G.A., Grainger, B.N., and Sullivan, P.J.E., 1986, Transient thermal strain of concrete during
first heating cycle to 600 C, Mag. Concrete Res., 37(133), 195–215.
Khoury, G.A., 1992, Compressive strength of concrete at high temperatures: a reassessment, Mag.
Concrete Res., 44(161), 291–309.
Kodur, V.K.R. and Wang, Y.C., 2001, Performance of high strength concrete filled steel columns at
ambient and elevated temperatures, in Tubular Structures IX (ed. Puthli, R. and Herion, S.) (Lisse:
Balkema Publishers).
Law, M. and O’Brien, T., 1989, Fire Safety of Bare External Structural Steel (Ascot: The Steel Construction
Institute).
Lawson, R.M. and Newman, G.M., 1996, Structural Fire Design to EC3 & EC4, and Comparison with BS
5950, Technical Report, SCI Publication 159, The Steel Construction Institute.
Newman, G.M., 1990, Fire and Steel Construction: The Behaviour of Steel Portal Frames in Boundary
Conditions (2nd edition) (Ascot: The Steel Construction Institute).
Newman, G.M., Robinson, J.T., and Bailey, C.G., 2000, Fire Safety Design: A New Approach to
Multi-Storey Steel-Framed Buildings, SCI Publication P288, The Steel Construction Institute.
Pettersson, O., Magnusson, S.E., and Thor, J., 1976, Fire Engineering Design of Steel Structures,
Publication 50, Swedish Institute of Steel Construction.
Phan, L.T. and Carino, N.J., 1998, Review of mechanical properties of HSC at elevated temperatures,
ASCE J. Mater. Civil Eng., 10(1), 58–64.
Phan, L.T. and Carino, N.J., 2000, Fire performance of high strength concrete: research needs,
Proceedings of ASCE Congress 2000, May, Philadelphia, USA.
Sakumoto, Y., 1998, Research on new fire-protection materials and fire-safe design, ASCE J. Struct. Eng.,
125(12), 1415–1422.
Wang, Y.C., 2000, A simple method for calculating the fire resistance of concrete-filled CHS columns,
J. Constr. Steel Res., 54, 365–386.
Wang, Y.C. and Kodur, V.K.R., 2000, Research towards use of unprotected steel structures, ASCE
J. Struct. Eng., 126(12), 1442–1450.
Wang, Y.C., 2002, Steel and Composite Structures, Behaviour and Design for Fire Safety (London: Spon
Press).

Appendix
WWW Sites
Nowadays the internet can provide a rich source of information on everything. There are many web
sites that give information on fire related topics. The following web sites are well-known organizations
that have a interest in the topic of this chapter, structural fire engineering. They are divided into three
groups: government and research organizations whose main functions are legislation, research, and
dissemination; academic institutions whose main functions are research and education; and industrial
companies that are involved in research and development to some extent, but whose main interest is
in the application of fire engineering in practical projects. It is inevitable that this list is biased toward
U.K. organizations.

Copyright 2005 by CRC Press


37-34 Handbook of Structural Engineering

A: Government and Research Organizations


www.bre.co.uk: Building Research Establishment (BRE), U.K.
www.safety.odpm.gov.uk/fire: Office of the Deputy Prime Minister, U.K.
www.nrc.ca: National Research Council of Canada (NRCC), Canada
www.bfnl.nist.org: Building and Fire Research Lab, National Institution of Science and Technology
(NIST), U.S.
www.nfpa.org: National Fire Protection Association (NFPA), U.S.
www.sfpe.org: Society of Fire Protection Engineers (SFPE), U.S.
www.iafss.org: International Association for Fire Safety Science (IAFSS)
www.vtt.fl: VTT Building Technology, Finland
www.sp.se: Swedish National Testing and Research Institute (SP), Sweden
www.sintef.no: Norwegian Fire Research Laboratory (SINTEF), Norway
www.factorymutual.com: Factory Mutual, U.S.
www.cticm.fr: Centre Technique Industriel de la Construction Métallique (CTICM), France
www.tno.bouw.nl: TNO Building and Construction Research, The Netherlands
B: Academic institutions
www.structuralfiresafety.com: One-stop-shop for structural fire safety, set up by Manchester Centre for
Civil and Construction Engineering, University of Manchester Institute of Science and Technology
(UMIST), U.K.
www.steelinfire.org.uk: Steel In Fire Forum (STIFF), a network coordinated by the University of
Sheffield, U.K.
www.ed.ac.uk: University of Edinburgh, U.K.
www.ulst.ac.uk: University of Ulster, U.K.
www.brand.lth.se: Lund University, Sweden
www.wpi.edu: Worcester Polytechnic Institute, U.S.
www.enfp.umd.edu: University of Maryland, U.S.
www.civil.canterbury.ac.nz: University of Canterbury, New Zealand
www.vut.edu.au: Victoria University of Technology, Australia
C: Industrial Companies
www.corusgroup.com: U.K. steel manufacturer
www.arup.com: Ove Arup & Partners, Engineering consultancy, U.K.
www.burohappold.com: Buro Happold Consulting Engineers, Engineering Consultancy, U.K.
www.steel-sci.org: The Steel Construction Institute, U.K.

Copyright 2005 by CRC Press

Potrebbero piacerti anche