Fire Performance of Thin-Walls and Steel Structures
Fire Performance of Thin-Walls and Steel Structures
of Thin-Walled
Steel Structures
Fire Performance
of Thin-Walled
Steel Structures
Yong Wang
Mahen Mahendran
Ashkan Shahbazian
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reasonable
efforts have been made to publish reliable data and information, but the author and publisher
cannot assume responsibility for the validity of all materials or the consequences of their use. The
authors and publishers have attempted to trace the copyright holders of all material reproduced in
this publication and apologize to copyright holders if permission to publish in this form has not
been obtained. If any copyright material has not been acknowledged please write and let us know
so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information
storage or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.
copyright.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC),
222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that
provides licenses and registration for a variety of users. For organizations that have been granted a
photocopy license by the CCC, a separate system of payment has been arranged.
Preface ix
Authors xi
List of notations xiii
v
vi Contents
References 89
Index 95
Preface
ix
x Preface
Yong Wang,
Mahen Mahendran,
Ashkan Shahbazian
Authors
xi
List of notations
LOWER CASE
b width
d depth of steel cross-section
e eccentricity/moisture content
f a constant
f y yield strength of steel
h convective heat transfer coefficient
k thermal conductivity
t time
x,y,z coordinates
UPPER CASE
A area
Cp specific heat
E Young’s modulus
L length/span/height
N compression resistance
Pn design resistance under compression
∙
Q heat
R thermal resistance
T temperature
Tf fire/furnace temperature
Ts surface/steel temperature
W width of panel
xiii
xiv List of notations
SUBSCRIPT
0 ambient condition
b buckling
cap related to capacitance
cond conduction
conv convection
cr critical
d distortional
e effective/Euler buckling
Ed design effect
f flange/fire
l local
rad radiation
Rd design resistance
ref reference
s surface
y yield
GREEK LETTER
α coefficient of thermal expansion
ε emissivity
ρ density
σ Stefan-Boltzmann constant
λ slenderness
Δ difference
Σ sum
Introduction
Fire Safety
Requirements and
1
Implications for
Thin-Walled Steel
Construction
This chapter presents the general requirements of fire safety and, in particular,
fire resistance, the implications for thin-walled steel construction, methods of
evaluating fire resistance of thin-walled steel structures and the contents of
this book.
1
2 Fire Performance of Thin-Walled Steel Structures
solve the problem of lack of qualified skills, its market share is still
relatively small. User confidence in the quality and flexibility of this
type of construction appears to be an issue.
Whatever the method of construction, the final product is a panel with enclosed
thin-walled steel members, with or without interior insulation. Figure 1.2 shows
examples of thin-walled steel structural floors and wall panels. It is the fire
performance of the complete panel assembly that will be the topic of this book.
However, as far as thin-walled steel structures are concerned, only a small num-
ber of these requirements affect their design and construction, as explained next.
wall is exposed to an external fire, as in the case of external fire spread, the
external fire temperature will be much lower than the internal fire temperature.
Therefore, the external fire exposure will in general be less detrimental to the
structure than interior fire exposure.
Therefore, the general scope and assumptions of this book are:
2.1 INTRODUCTION
Thin-walled steel structural components are made using a cold-forming pro-
cess (roll-forming or press-braking) at ambient temperature. Unlike hot-rolled
and welded heavy steel sections, which are only available in limited shapes
and sizes, thin cold-formed steel sections can be custom designed and made
to enhance structural and cost efficiencies and to suit specific applications.
Unlipped and lipped channel sections (Figure 2.1) are still commonly used
in the construction of wall, floor and roof systems. Z-sections are also used
in applications such as purlins and joists. However, complex sections shown in
Figure 2.2 can be easily roll-formed using low- and high-strength cold-rolled
steel coils with thicknesses in the range of 0.55–6.35 mm since automated
advanced manufacturing technologies are available now.
11
12 Fire Performance of Thin-Walled Steel Structures
Supacee Supazed
There is great potential to use similar HFS sections and optimised C- and
Z-sections in a range of applications including wall, floor and truss systems.
Recently, an HFS similar to HFC (Figure 2.6) was developed and used in
modular units in Korea due to their excellent performance and cost-effi-
ciency (Ha et al. 2016). Using simultaneous cold-formed and electric weld-
ing process, the so-called MCO s ections with 100–120 mm flange widths are
being produced with depths in the range of 200–390 mm and thicknesses in
the range of 4.5–10 mm. A smaller clinched HFS is also used as truss mem-
bers in Australia.
2.2.2 Prefabricated Structural
and Modular Units
The availability of new automated roll-forming facilities means that complex
and optimised thin-walled steel sections can be easily cold-formed to produce
them with tight tolerances and custom cut lengths. The new facilities can inte-
grate CAD design to produce pre-cut, punched and sized steel members, which
can then be readily used to make the basic structural wall and floor panels and
roof trusses inside a factory in a controlled environment. Figure 2.7 shows typi-
cal prefabricated wall and floor panels with suitable studs and joists at 600 mm
centres (also 300 and 400 mm). They can all be then transported to the site
(Figure 2.7) and assembled using, in most cases, the efficient self-drilling screw
fasteners in constructing low- and mid-rise buildings (Figures 2.3 and 2.4).
This off-site manufacturing of light steel framing (wall and floor panels and
roof trusses) will considerably reduce labour cost and waste in the factory and
2 • Applications of Thin-Walled Steel Structures 17
FIGURE 2.8 A modular building. (From Steel Construction Institute (SCI), Fire
resistance of light steel framing, SCI Publication P424, Steel Construction Institute,
Ascot, UK, 2019. With permission.)
storeys, and then both stud size and thickness (0.75–3 mm) can be reduced for
the higher storeys. Light steel buildings can be easily modified and extended if
required, disassembled and reused at another site.
FIGURE 2.11 Improved board joints. (With kind permission from Springer
Science+Business Media: Fire Technol., A review of parameters influencing
the fire performance of light gauge steel-framed walls, 2017, Kesawan, S. and
Mahendran, M.)
22 Fire Performance of Thin-Walled Steel Structures
2.3.2 Insulation
Insulation (glass fibre, cellulosic fibre or rock fibre) is commonly used in
light steel wall and floor systems to enhance their thermal and acoustic
performance at ambient temperature. The use of cavity insulation has been
shown to increase the insulation-based FRL for non-load-bearing walls,
but it was found to be detrimental to the FRL of load-bearing walls (Kodur
and Sultan 2006, Gunalan et al. 2013). With cavity insulation, the consider-
able rise in temperatures on the fireside plasterboards and stud hot flanges,
together with a larger temperature gradient, led to premature stud failures.
To overcome this problem, Kolarkar and Mahendran (2008) proposed exter-
nally insulated light gauge steel-framed (LSF) walls, where an insulation
layer was sandwiched between two plasterboard layers (Figure 2.9). Fire
tests demonstrated 30% improvement in FRL when external insulation was
used. The use of similar external insulation systems will enhance the fire
resistance of both non-load-bearing and load-bearing walls. If composite
panels consisting of boards and insulation can be prefabricated and installed,
the installation cost will remain the same. Enhanced insulation materials
with improved thermal properties can also be developed by varying their
compositions.
different profiles have the same FRL for a given load ratio if their depth and
flange widths are the same. Increase in stud thickness is likely to improve the
FRL of LSF walls. Effects of web depth on the FRL of walls are dependent on
the type of failure mode of studs (section yielding or local buckling or major
axis buckling) and thermal bowing deflections. In summary, for a chosen wall/
floor depth, FRL cannot be improved by stud/joist section profile. However,
elevated temperature mechanical property reduction factors vary depending
on the type/grade of steel used. Ariyanayagam and Mahendran (2018) showed
that the use of steels of varying strength grades significantly influenced the
FRL of load-bearing walls. Hence developing and using cold-formed steels
with higher elevated temperature mechanical property reduction factors, espe-
cially in the range of 400°C–700°C, can provide higher FRLs for thin-walled
steel wall and floor systems.
2.4 SUMMARY
The last section has briefly explained the effects of key parameters on the fire
resistance of thin-walled steel construction systems and suggested methods to
enhance fire resistance. The next chapter will provide further details of fire
performance characteristics of different types of thin-walled construction sys-
tems based on fire resistance tests.
Fire Resistance
Tests 3
Because of complexity of detailing of thin-walled steel structural systems,
most fire design standards worldwide use the so-called standard fire tests to
determine their fire resistance level (FRL). Standard fire resistance tests are
also used by researchers to determine and investigate the FRL of the con-
structed system as a function of key influential parameters. This chapter first
presents the standard fire resistance test in general, and then describes the
fire resistance characteristics of different types of thin-walled steel wall and
floor systems.
25
26 Fire Performance of Thin-Walled Steel Structures
where:
T f is the average furnace temperature (°C) at time t
t is the time elapsed (min)
3 • Fire Resistance Tests 27
FIGURE 3.2 Standard fire time-temperature curve. (From ISO 834-1, Fire resis-
tance tests – elements of building construction, Part 1: General requirements.
International Organization for Standardization, Geneva, Switzerland, 1999.)
Depending on the intended function of the tested system, the fire resistance
level (FRL) of the tested system is the time when one of the following three
fire resistance criteria is reached:
Figure 3.3 shows a typical fire test set-up used to assess a cold-formed LSF
wall system where pre-determined loads are applied to the studs via the bot-
tom track. Fire tests of LSF floor systems are conducted in a horizontal fur-
nace with fire exposure from underneath after applying the pre-determined
vertical loads to the floor boards.
Fire testing using the standard fire time-temperature curve gives good
comparative results for building components tested under identical condi-
tions, and also valuable basic test data. They are commonly used by product
28 Fire Performance of Thin-Walled Steel Structures
Grade 300 steel (2.85 m height) and 1.0 mm Grade 450 steel (3.6 m height) and
used at 600 mm spacing with a central row of noggings without any interior
insulation. The tests were conducted on horizontally placed panels by loading
via the tracks. Major axis flexural buckling associated with local buckling of
mid-height cold flanges occurred in two tests while in the third test flexural
torsional buckling occurred due to the loss of lateral restraint to the compres-
sion flange by the thinner (9.5 mm only) ambient side plasterboards. With the
use of 12.5 and 9.5 mm boards, 32 min fire resistance was reached while using
16 mm boards gave 72 min fire resistance.
Kodur and Sultan’s (2006) experimental study included 14 LSF wall pan-
els with 0.84 and 0.912 mm thick studs spaced at 406 mm (610 mm in one test)
and lined with one or two 12.7 mm gypsum boards on both sides (15.9 mm
boards in two tests). Glass fibre, rock fibre and cellulose fibre cavity insula-
tions were used. Local buckling of studs was the dominant structural failure
mode except in two tests where overall buckling was observed. Fire resistance
was reduced when cavity insulation was used, with cellulose fibre performing
better than others.
Alfawakhiri (2001) conducted six standard fire resistance tests of load-
bearing LSF walls made of 0.912 mm thick 90 mm studs (low yield strength of
228 MPa). Major axis flexural buckling of studs occurred towards the furnace
with compressive failure of the cold flange near mid-height. When cavity insu-
lation was used, stud hot flange temperatures were increased while cold flange
temperatures were reduced, and the wall moved away from the furnace and
failed by hot flange compression failure. Using cavity insulation reduced fire
resistance. The use of noggings and ambient side cross-bracing was considered
to have eliminated flexural-torsional and minor axis flexural buckling of the
studs. Local compressive failures of studs occurred at one of the four holes
located along the length. The use of resilient channels slightly reduced the fire
resistance times.
Feng and Wang (2005) conducted six standard fire tests of load-bearing
LSF wall panels of 2.2 × 2.0 m lined with 12.5 mm gypsum board lining and
cavity insulation. Lipped channel studs of 100 × 54 × 15 mm made of 1.2 and
2.0 mm thick S350 steels were used with two holes near the ends. Increasing
the load ratio (0.2, 0.4 and 0.7) decreased the stud failure time significantly.
Except for one test (local buckling at the hole), the studs failed by flexural
torsional buckling. This was considered to be due to the inability of fireside
plasterboard to restrain the studs although the ambient side plasterboard was
able to prevent minor axis flexural buckling. Their study noted that plaster-
board fall-off could have been triggered by stud failure, instead of plasterboard
fall-off causing stud failure.
Zhao et al. (2005) conducted 29 tests of 1.2 × 2.8 m LSF walls with two studs
spaced at 600 mm with gypsum plasterboards on one or both sides, and with
30 Fire Performance of Thin-Walled Steel Structures
or without cavity insulation. Studs of varying sizes were used and loaded con-
centrically or eccentrically for load ratios from 0.2 to 0.6. Test results showed
that stud failure is essentially governed by the maximum temperature in fire.
Steel studs supported by plasterboards on both sides failed by major axis flex-
ural buckling with local buckling of the cold flange. This failure mode was
observed with eccentric loading towards the fireside. In some other tests, the
same failure mode occurred but with movement away from the fire and com-
pressive failure of the hot flange. These observations show the presence of
complex stud failure modes due to combined thermal and structural effects
and are similar to those observed by other researchers as discussed above.
Gunalan et al. (2013) conducted 11 full-scale fire tests of load-bearing
walls made of 1.15 mm high-strength G500 cold-formed steels. Single or dou-
ble gypsum plasterboard lining of 16 mm thickness was used with or without
rock fibre, glass fibre and cellulosic fibre cavity insulation. Typical plaster-
board and stud time-temperature curves from their standard fire tests of double
plasterboard–lined walls are shown in Figure 3.4. They are similar to those
observed by earlier researchers and show that gypsum plasterboards are able to
keep the cavity temperatures low and delay steel stud temperature rise for a
long time. The figure also shows presence of non-uniform temperature distri-
butions (temperature gradient) across the wall/stud depth. Many researchers
(Kaitila 2002, Feng et al. 2003a, Gunalan et al. 2013) have confirmed that it is
acceptable to assume uniform temperatures in the flanges and lips and a linear
web temperature distribution as shown in Figure 3.5.
FIGURE 3.4 Time-temperature curves from a standard fire test of LSF wall.
(Reprinted from Thin-Walled Struct., 65, Gunalan, S. et al., Experimental study of
cold-formed steel wall systems under fire conditions, 72–92, Copyright 2013, with
permission from Elsevier.)
3 • Fire Resistance Tests 31
Hot Flange
Linear Web
variation
Cold Flange
FIGURE 3.7 LSF wall panels after failure. (Reprinted from Thin-Walled Struct., 65,
Gunalan, S. et al., Experimental study of load bearing cold-formed steel wall sys-
tems under fire conditions, 72–92, Copyright 2013, with permission from Elsevier.)
Gunalan et al.’s (2013) fire tests demonstrated integrity of the inner
p lasterboard until failure (Figure 3.7) and showed that there was sufficient
lateral restraint to the studs from the plasterboards until failure. In fact, in
all the fire tests of load-bearing walls with load ratios of 0.2 and above, steel
stud failure occurred before integrity or insulation failure. This is despite
plasterboard fall-off that started to occur when the inner plasterboard sur-
face temperature reached about 900°C. Other researchers (Alfawakhiri
2001, Feng et al. 2003b, 2003c, Zhao et al. 2005) also suggested that it would
be appropriate to assume sufficient plasterboard lateral restraints to prevent
minor axis flexural buckling and flexural-torsional buckling failures of the
steel studs.
From the above standard fire tests of load-bearing LSF walls conducted in
many different countries, the following useful general observations are made:
FIGURE 3.9 Composite panel: (a) with external insulation and (b) stud time-
temperature curves (HF-hot flange, CF-cold flange).
36 Fire Performance of Thin-Walled Steel Structures
The fire performance of the new LSF wall system was investigated using
small-scale and full-scale fire tests and numerical studies by Kolarkar and
Mahendran (2012), Gunalan et al. (2013), Kesawan and Mahendran (2015) and
Chen and Ye (2014). The use of external insulation resulted in reduced temper-
atures of the inner plasterboard layer, and so its calcination and deterioration
were delayed. It also allowed direct heat transfer from the fireside plaster-
board to the ambient side plasterboard. Hence in contrast to cavity-insulated
walls (Figure 3.8a), the stud hot flange temperatures and associated tempera-
ture gradients were significantly reduced as shown in Figure 3.9b. Fire tests
showed that the FRLs of load-bearing lipped channel stud walls increased by
about 25%, with rock fibre insulation providing the best outcome. Kesawan
and Mahendran (2015) showed that similar improvements can be obtained
for LSF walls made of other stud sections such as hollow flange channels.
In terms of insulation performance, the corresponding FRLs are likely to be
the same for both cavity and externally insulated panels based on both test and
numerical results.
Furthermore, Gunalan et al.’s (2013) test results showed that the vertical
plasterboard joints on the inner boards were protected by the external insulation
layer even after the outer layer fall-off, thus preventing localised temperature
rise in the stud hot flange. Therefore, plasterboard joints in LSF walls with
external insulation are not as detrimental as those in cavity-insulated or
uninsulated walls.
FIGURE 3.10 HFC stud sections. (Reprinted from Thin-Walled Struct., 98(A),
Kesawan, S. and Mahendran, M., Predicting the performance of LSF walls made
of hollow flange sections in fire, 111–126, Copyright 2019, with permission from
Elsevier.)
FIGURE 3.11 Cracking and opening-up of joints and their effects on stud time-
temperature curves. (Reprinted from Ariyanayagam, A.D. et al., Thin-Walled
Struct., 107, Detrimental effects of plasterboard joints on the fire resistance of
light gauge steel frame walls, 597–611, Copyright 2016, with permission from
Elsevier.)
were not directly exposed to fire through the opened-up joints. They also
showed that since the joint effect is localized by only affecting the stud hot
flange and fireside plasterboard temperatures, its effect on the insulation-based
FRL of non-load-bearing walls is negligible. Innovative plasterboard joint
arrangements can also be used along the studs instead of back-blocking to
reduce construction time/cost.
3 • Fire Resistance Tests 39
FIGURE 3.14 Mass loss of MgO boards. (Reprinted from Fire Saf. J., 90, Rusthi,
M. et al., Fire tests of magnesium oxide board lined light gauge steel frame wall
systems, 15–27, Copyright 2017, with permission from Elsevier.)
3 • Fire Resistance Tests 41
at 600 mm spacing while the third one had glass fibre cavity insulation.
The cavity-insulated wall panel exhibited higher hot flange temperatures and
associated high temperature gradients and larger lateral deflections of the wall
as observed for gypsum plasterboard–lined walls. All three wall panels suf-
fered severe board cracking on the ambient side (Figure 3.15) and failed by
integrity criterion, giving only a 30 min FRL. The high mass loss caused board
cracking in all three tests.
The use of noggings led to excessive board cracking as it restricted ther-
mal expansion and bowing of the boards. When noggings were removed, joint
opening and cracking due to excessive shrinkage and bowing of the boards
caused panel failure (Figure 3.15). Fire rated sealants between boards and
studs and in the joints could not withstand the excessive movement caused by
board bowing and shrinkage. Strengthening the board joints with high tem-
perature mortar also did not prevent joint cracking. Fibre mesh appeared to
have been removed along the recessed board edges, which also affected the
board strength in fire. This study has shown that mass loss characteristics of
boards, joint detailing and stud spacing are important factors affecting the
FRLs of LSF walls.
Hanna et al.’s (2015) fire tests of load-bearing walls also showed similar
observations. In their tests, ambient side cracks were seen at 40 and 50 min,
and they were wide open allowing the passage of hot gases. To eliminate the
problems with MgO board, Chen et al. (2013a) suggested the use of MgO board
as the inner layer and gypsum plasterboard as the outer layer. These studies
seem to indicate that MgO boards are not an effective solution in improving
the fire resistance of LSF walls.
FIGURE 3.15 Fire tests of MgO board-lined LSF walls. (Reprinted from Fire
Saf. J., 90, Rusthi, M. et al., Fire tests of magnesium oxide board lined light
gauge steel frame wall systems, 15–27, Copyright 2017, with permission from
Elsevier.)
42 Fire Performance of Thin-Walled Steel Structures
New types of fire protective boards are being regularly introduced into
the building sector. It is essential that full-scale fire tests are conducted to
demonstrate that they can provide thin-walled steel structures with ade-
quate FRLs. Furthermore, to facilitate advanced analyses, it is important
that manufacturers and suppliers provide temperature-dependent thermal
properties of their products such as specific heat, thermal conductivity and
mass loss.
Source: Reprinted from Thin-Walled Struct., 137, Dias, Y. et al., Full-scale fire tests of steel- and plasterboard-sheathed web-stiffened stud
walls, 81–93, Copyright 2019, with permission from Elsevier.
3 • Fire Resistance Tests 43
44 Fire Performance of Thin-Walled Steel Structures
FIGURE 3.16 Failures of steel-sheathed LSF walls. (a) LSF wall with internal steel
sheathing, (b) LSF wall with external steel sheathing. (Reprinted from Thin-Walled
Struct., 137, Dias, Y. et al., Full-scale fire tests of steel and plasterboard sheathed
web-stiffened stud walls, 81–93, Copyright 2019, with permission from Elsevier.)
FIGURE 3.17 Failures of LSF floor panels. (From Alfawakhiri, F. and Sultan,
M.A., Loadbearing capacity of cold-formed steel joists subjected to severe
heating, Proceedings of the 9th International Conference Proceedings, Interflam
2001, Edinburgh, UK, Vol. 1, pp. 431–442, 2001. With permission from National
Research Council of Canada.)
reasons, the FRL of load-bearing floors was significantly reduced by the use of
cavity insulation. Integrity and insulation failures were not observed, and the
floor panel collapsed due to section failure of the joists (Figure 3.17).
Sakumoto et al. (2003) conducted six fire tests of 4.26 × 2.95 m LSF
floors made of back-to-back lipped channel joists lined with composite boards
including plywood on the top and one or two layers of gypsum plasterboards
of thicknesses 9.5, 12.5 and 15 mm. Their tests also showed that the FRL
increased significantly with increasing thickness and number of boards, but
cavity insulation did not reduce the FRL as expected. Plasterboard fall-off was
identified as the key factor influencing the FRL.
Zhao et al. (2005) conducted full-scale fire tests of two floor systems of
5.5 × 2.99 m made of 250 × 2.5/2.0 mm lipped channel joists at 600 mm spac-
ing and lined with two layers of gypsum board and cavity insulation. With
fire on one side, non-uniform temperature distribution was developed, which
caused thermal bowing deformations in addition to deflections caused by the
applied load. With increasing temperatures on the fireside and plasterboard
fall-off, the joists failed in bending due to reduced mechanical properties in
fire. Localised web crippling failures were also observed at the joist to track
supports.
Baleshan and Mahendran (2016) conducted three fire tests of LSF floors
made of 180 × 1.15 mm lipped channel joists with two layers of 16 mm plaster-
board on the fireside under a load ratio of 0.4. Rock fibre insulation was used in
46 Fire Performance of Thin-Walled Steel Structures
the cavity in the second test while it was used between the two plasterboards as
external insulation in the third test, following the recommendation of Kolarkar
and Mahendran (2008, 2012). In all three tests, the joists moved towards the
fireside and failed locally by web crippling near the supports. The use of cavity
insulation reduced the failure time from 107 to 99 min due to higher hot flange
temperatures and larger lateral deflections caused by greater temperature gra-
dients across the floor depth. On the other hand, the use of external insulation
increased the failure time to 139 min. These observations are similar to those
observed for LSF walls as discussed in Sections 3.2.2 and 3.2.3.
From the aforementioned standard fire tests of LSF floors, the following
useful observations are made:
FIGURE 3.19 Uniform elevated temperature tests and failure modes of short
thin-walled steel columns: (a) electrical furnace, (b) distortional buckling, (c) local
buckling of SHS, (d) local web buckling of lipped channel and (e) local flange
buckling of unlipped channel. (Reprinted from J. Constr. Steel Res., 65, Ranawaka,
T. and Mahendran, M., Distortional buckling tests of cold-formed steel com-
pression members at elevated temperatures, 249–259, Copyright 2009, with
permission from Elsevier.)
3 • Fire Resistance Tests 49
51
52 Fire Performance of Thin-Walled Steel Structures
(1) (3)
Growth phase Decay phase
(2)
Temperature
Ignition Time
Table 4.1 lists the most commonly used nominal fire conditions and their
parameters.
The nominal fire temperature–time curves are:
T f 660 1 0.687 e 0.32 t 0.313 e 3.8 t 20 (4.2)
T f 1080 1 0.325 e 0.167 t 0.675 e 2.5 t 20 (4.3)
where:
T f is the gas/fire temperature in °C
t is time in min
1200
1000
Temperature (θ, ºC)
800
600
400
200
θmax
Temperature
t*max Time
FIGURE 4.4 Travelling fire. (From Stern-Gottfried, J. and Rein, G., Fire Saf. J., 54,
96–112, 2012. With permission.)
enclosure fire behaviour requires special expertise and detailed fire dynamics
knowledge. They are rarely used by structural engineers.
Enclosure fires may be modelled using zone models, where the enclosure
is divided into a very small (typically 2) number of zones, each having uniform
properties or computational fluid dynamics (CFD) models. Popular zone and
CFD models for modelling enclosure fires include the following (Wang et al.
2012):
dT
Q cond kA (4.4)
dx
T T T T
k k k Cp (4.5)
x x y y z z t
where:
t is time [sec]
ρ is the density of the material [kg/m3]
Cp is the specific heat of the material [J/kg K]
where:
Q conv is the rate of convective heat transfer (W)
hconv is the convective heat transfer coefficient in W/m2°C
As is the surface area through which convective heat transfer takes place
Ts is the surface temperature [°C]
T f is the temperature of the gas sufficiently far from the surface [°C] (fire
temperature)
4
Q rad As T f 273.15 Ts 273.15
4
(4.7)
where:
Q rad is the rate of radiant heat transfer (W)
ε is the emissivity of the surface. The property emissivity, whose value is
in the range between 0 and 1, is a measure of how closely a surface
approximates a blackbody surface, which has an emissivity value
of 1 ( 1)
σ is the Stefan–Boltzmann constant ( 5.67 10 8 W / m 2 K 4 )
As is the surface area through which radiant heat transfer takes place
Ts is the surface temperature
T f is the surrounding air (fire) temperature
Figure 4.6 illustrates different modes of heat transfer for a thin-walled steel
structural panel with fire exposure from one side (Shahbazian and Wang
2014b).
perform heat transfer simulations. There are also a few specialist heat transfer
programs that have been specifically developed for performing heat transfer
calculations for structures under fire conditions, although some were developed
a long time ago and are not maintained. They include FIRES-T3 (Iding et al.
1977), TASEF (Gerlich et al. 1996), WALL2D (Takeda and Mehaffey 1999)
and SAFIR (Franssen and Gerney 2017).
Over a small time interval, the heat transfer between any two layers can be
generally written as:
T
Q (4.9)
R
where:
∆T is the temperature difference between these two layers
ΣR is the total thermal resistance in the heat transfer path
62 Fire Performance of Thin-Walled Steel Structures
R
dT
Q cap cap (4.10)
dt
where:
t is time
ΣRcap is the total heat capacitance (mass times specific heat) of the layer
4 • Numerical Modelling of Fire Resistance of Thin-Walled 63
In this simplified method, the thermal resistance is calculated using the
weighted average of the materials within the heat transfer path. Details are as
follows:
• For heat transfer between the fire and the slice of gypsum board on
the fire exposed side (point 1), the thermal resistance is the total of
the thermal boundary layer and ½ of the gypsum slice.
• For heat conduction between two adjacent gypsum slices (points
2–3, 3–4, 4–5, 10–11, 11–12 and 12–13), the thermal resistance is
the total of the two halves of each slice.
• For heat transfer between the slice of gypsum board on the air side
with the air layer, the thermal resistance is the total of the air layer
and ½ of the gypsum slice (point 14).
• For heat transfer between a gypsum slice and a steel slice or between
two adjacent steel slices, calculation of the total thermal resistance
should include different materials that form the slices. For each
slice, the thermal resistances are in parallel.
• To calculate the heat capacitance of each slice, all the materials
within that slice should be included.
We 45 0.85b f in mm (4.11)
4.4 SUMMARY
The distribution of temperatures in thin-walled steel sections as part of a
panel exposed to fire on one side can be complex. However, for evaluations
of the load carrying capacity of LSF structures, the temperature profile can
be simplified. Furthermore, the two-dimensional heat transfer problem can
be approximately reduced to a simple one-dimensional problem by using the
weighted average of thermal properties of the different constituent parts of the
thin-walled steel structural panel.
Elevated
Temperature
Properties of
5
Materials
5.1 INTRODUCTION
In fire engineering calculations of structural resistance in fire, it is necessary
to have accurate and reliable information on thermal and mechanical prop-
erties of materials of the structure. Compared to hot-rolled steel structures,
there are more issues for the material properties of cold-formed thin-walled
steel structures. Cold-formed steel structures are made by cold working of
thin-walled cold-rolled steel strips into structural shapes. This introduces
strain hardening around corners, and the amount of strain hardening may be
affected by the thickness of the structure. Lightweight board and insulation
materials, such as gypsum plasterboard and mineral fibres, are used in thin-
walled steel construction systems. The thermal properties, in particular ther-
mal conductivity, of these materials are not constant. Since temperatures
attained in steel sections of thin-walled construction are critically depen-
dent on insulation p roperties, thermal properties of insulation materials and
gypsum plasterboards should be provided. This chapter will review avail-
able data from major sources of relevant information in literature and make
recommendations.
65
66 Fire Performance of Thin-Walled Steel Structures
5.3.1 General
In cold-formed thin-walled steel panel construction for walls and floors,
lightweight fire protection and interior/exterior insulation materials are used
to provide sufficient fire resistance and thermal/sound insulation. Accurate
data of thermal properties of these materials are critical for safe and reliable
calculation of fire resistance of these systems. In fact, since the mechanical
properties of cold-formed steel decrease sharply at elevated temperatures
in the region of interest, small errors in the calculation of steel temperature
results can result in very large errors of calculating structural resistance.
For example, when the steel temperature is increased from 500°C to 600°C
(20% increase), according to the EN 1993-1-2 recommendation of reduc-
tion factors of steel yield strength and modulus of elasticity in Table 5.1, the
yield strength and modulus of elasticity of steel decrease by 43% and 48%,
respectively.
The thermal properties of a material include density (ρ), specific heat (Cp)
and thermal conductivity (k). The product of density and specific heat (ρ Cp) is
thermal capacitance of the material, measuring the amount of heat (energy)
required to raise the temperature of per unit volume of material by 1 degree.
Since fire protection and insulation materials used in cold-formed thin-walled
steel structures are lightweight, the accuracy of calculating steel temperatures
is insensitive to even large changes in thermal capacitance of fire protection/
insulation materials. On the other hand, fire protection and insulation materi-
als have low thermal conductivities, therefore, temperatures attained in the
steel structure vary almost linearly with variations in thermal conductivity of
protection/insulation materials.
The thermal conductivities of fire protection and insulation materials at
ambient temperature can be expected to be provided by the material manu-
facturers because they have to provide such information for evaluating the
thermal insulation performance of the panels. However, the ambient tempera-
ture values should not be used in fire engineering calculations. This is because
the thermal conductivity of lightweight materials increases with temperature.
70 Fire Performance of Thin-Walled Steel Structures
k k0 Const T 3 (5.1)
where:
k and k0 are thermal conductivity of the material at temperature T and
ambient temperature, respectively
T is the material temperature
5.3.2 Gypsum
Gypsum plasterboards are commonly used in cold-formed thin-walled steel
panels as fire protection. Due to the presence of a large amount of free and
chemically bound water, their thermal properties are affected by water move-
ment and evaporation.
The total amount of water is about 24% by weight, consisting of about 21%
of chemically bound water and 3% of free water. Water evaporation occurs in
two stages, the first stage starting at about 100°C, during which about 75% of
the chemically bound water and free water are lost. The temperature range
over which the second stage of water evaporation occurs is still subject to
5 • Elevated Temperature Properties of Materials 71
TABLE 5.2 Thermal property models for some common types of insulation
material
BASE VALUE OF
DENSITY SPECIFIC HEAT THERMAL CONDUCTIVITY
MATERIAL ρ (kg/m3) (J/kg·K) (W/m·K)
3
T
Rock fibre 155–180 900 k 0.022 0.1475
1000
3
T
Mineral wool 165 840 k 0.03 0.2438
1000
3
T
Calcium silicate Various 900 k k0 Const
1000
k0 0.23
1000
2540
Const 0.08
2540
3
T
Vermiculite Various 900 k k0 Const
1000
k0 0.27
1000
1000
Const 0.18
1000
Source: With kind permission from Taylor & Francis: Performance-Based Fire Engineering of
Structures, 2012, Wang, Y.C. et al.
debate, with some stating immediately following completion of the first stage
at about 200°C and others saying as late as 600°C. Because the amount of
water involved in the second stage is relatively low, it is acceptable to assume
that water evaporation is complete in the temperature range between 100°C
and 200°C.
Based on the above assumption of water evaporation, the thermal properties
of gypsum plasterboard can be quantified as follows.
5.3.2.1 Density
The density of gypsum plasterboard is 100% of its ambient density until 100°C,
then linearly changing to 76% (100%–24%) at 200°C and maintaining at this
value at higher temperatures.
72 Fire Performance of Thin-Walled Steel Structures
FIGURE 5.1 Behaviour of gypsum and interior insulation after a fire test.
(Reprinted from Fire Saf. J., 40, Feng, M. and Wang, Y.C., An experimental study
of loaded full-scale cold-formed thin-walled steel structural panels under fire
conditions, 43–63, Copyright 2005, with permission from Elsevier.)
74 Fire Performance of Thin-Walled Steel Structures
5.5 SUMMARY
The accuracy of any analysis depends on the accuracy of input of the material
properties. For the evaluation of fire resistance of structures, accurate data
of both thermal and mechanical properties of the materials are necessary.
For thin-walled steel structures, this chapter has recommended temperature-
dependent thermal and mechanical property values for some protective mate-
rials and steel, respectively. However, significant gaps in knowledge on the
material properties of thin-walled steel structure still exist and further research
studies are necessary.
Performance-
Based Design
Methods of
6
Thin-Walled
Steel Members
at Elevated
Temperatures
Thin-walled steel members usually form part of a panel system, either as part
of walls or floors. When designing for fire resistance, these panels are assumed
to be exposed to fire from one side. Therefore, temperature distributions in
thin-walled steel members are non-uniform, with steep gradients. Thin-walled
steel members are also prone to different modes of buckling, including local
buckling, distortional buckling and global buckling. Therefore, the behaviour
of thin-walled steel members in fire is very complicated. It is not surprising that
so far there is no widely accepted design calculation method for t hin-walled
steel members in fire.
This chapter will present a number of different methods that are either in
design codes and standards or were a result of extensive research, and explain
their scopes of application. The authors hope that they become the basis of
further assessment by interested readers who may wish to develop a design
method for their applications.
75
76 Fire Performance of Thin-Walled Steel Structures
where:
NRd, f is the load carrying capacity of the thin-walled steel member at the
fire limit state
Nb, Rd is the load carrying capacity of the member in normal conditions
Tref is the reference temperature of the thin-walled steel section, which is taken
as the mean temperature of the section. This may be taken as the average of
the flange temperatures.
SRF(Tref ) is the strength reduction factor for cold-formed steel at Tref
according to BS EN 1993-1-2 (2005). Because the extension method is based
on having availability of a valid fire test, the steel temperatures are from the
fire test. In fire resistance tests on thin-walled steel panels, the thinnest steel
section is usually used. Therefore, the test temperatures can be applied to
thicker steel sections, which will have lower temperatures.
The multiplication factor of ‘0.6’ is to allow for the effects of ther-
mal bowing due to non-uniform temperature distribution in the steel
cross-section.
This method allows the fire test result to be used for steel members that
are slightly longer (+10% in height), thicker steel members that resist higher
loads at the same fire resistance time, or to allow steel members to resist
higher loads at lower fire resistance times for which temperature data would
be available. However, this method is not applicable to steel members with
reduced load level to achieve longer fire resistance times for which experi-
mental data of temperatures would not be available. Due to gross approxima-
tions, this method should only be extended (extrapolated) over a small range,
e.g. about 10%.
The results of Feng et al. (2003d) indicate that it is possible to use the effective
width method to calculate the resistances of thin-walled steel members in wall
construction with a reasonable level of accuracy. Using the first yield condition
to calculate the cross-section resistance would give lower resistances than using
the partial plasticity condition, but overall differences are small.
80 Fire Performance of Thin-Walled Steel Structures
L2 T
m (6.2)
8d
where:
α is the coefficient of thermal expansion of steel
L is the column length (panel height)
ΔT is the temperature difference between the exposed and unexposed
sides
d is depth of the cross-section
• Global buckling:
for λ e ≤ 1.5:
2
Pne 0.495 e Py (6.3)
for λ e > 1.5:
0.462
Pne 2 Py (6.4)
e
• Local buckling:
for λ l ≤ 0.776:
Pnl = Pne (6.5)
for λ l > 0.776:
Pcrl Pcrl
0.75 0.75
Pnl 1 0.22 Pne (6.6)
Pne Pne
• Distortional buckling:
for λ d ≤ 0.561:
Pnd = Py (6.7)
for λ d > 0.561:
P P
0.7 0.7
In the above formulas Py is the equivalent squash load of the cross-section; Pcre
is the critical elastic global buckling load; Pcrl is the critical elastic local buck-
ling load; Pcrd is the critical elastic distortional buckling load; λ e, λ l and λ d
are the slenderness for global, local and distortional buckling modes, respec-
tively; Pne, Pnl and Pnd are the column axial strength for global, local and distor-
tional buckling modes, respectively; Pn is the final design strength under axial
compression.
The DSM equations above have been demonstrated to be applicable for
thin-walled members in wall panels under both the standard and parametric
design fire curves of EN 1991-1-2 (2005).
Pcre = 68.04 kN
Pcrl = 388.85 kN
Pcrd = 317.7 kN
6.6 SUMMARY
This chapter has presented an outline of a number of methods of calculating
the load carrying capacity of thin-walled steel structures at elevated tempera-
tures. Thin-walled steel structures have non-uniform temperature distributions
and complex modes of failure. Therefore, there is still no universally accepted
design calculation method. Nevertheless, as have been demonstrated in this
chapter, it is possible to use the existing ambient temperature design methods
as the basis for further development.
References
Alfawakhiri, F. (2001). Behaviour of cold-formed steel framed walls and floors in stan-
dard fire resistance tests. PhD thesis, Carleton University, Ottawa, Canada.
Alfawakhiri, F. and Sultan, M.A. (2001). Loadbearing capacity of cold-formed
steel joists subjected to severe heating. Proceedings of the 9th International
Conference Proceedings, Interflam 2001, Edinburgh, UK, Vol. 1, pp. 431–442.
Alfawakhiri, F., Sultan, M.A. and MacKinnon, D.H. (1999). Fire resistance of
load bearing steel-stud walls protected with gypsum board. Journal of Fire
Technology, Vol. 35, pp. 308–335.
American Iron and Steel Institute (AISI) (2016). S100-16: North American specifica-
tion for the design of cold-formed steel structural members, 2016 edition, AISI,
Washington, DC.
Anapayan, T. and Mahendran, M. (2012). Improved design rules for hollow flange sec-
tions subject to lateral distortional buckling. Thin-Walled Structures, Vol. 50,
pp. 128–140.
Ang, C.N. and Wang, Y.C. (2009). Effect of moisture transfer on specific heat of gyp-
sum plasterboard at high temperatures. Construction and Building Materials,
Vol. 23, pp. 675–686.
Ariyanayagam, A. and Mahendran, M. (2018). Fire performance of load bearing
LSF walls made of low strength steel studs. Thin-Walled Structures, Vol. 130,
pp. 487–504.
Ariyanayagam, A. and Mahendran, M. (2019). Influence of cavity insulation on the
fire resistance of light gauge steel framed walls. Construction and Building
Materials, Vol. 203, pp. 687–710.
Ariyanayagam, A.D. and Mahendran, M. (2017). Fire tests of non-load bearing light
gauge steel frame walls lined with calcium silicate boards and gypsum plaster-
boards. Thin-Walled Structures, Vol. 115, pp. 86–99.
Ariyanayagam, A.D., Kesawan, S. and Mahendran, M. (2016). Detrimental effects of
plasterboard joints on the fire resistance of light gauge steel frame walls. Thin-
Walled Structures, Vol. 107, pp. 597–611.
Avery, P., Mahendran, M. and Nasir, A. (2000). Flexural capacity of hollow flange
beams. Journal of Constructional Steel Research, Vol. 53, pp. 201–223.
Babrauskas, V. (1979). COMPF2, a program for calculating post-flashover fire tem-
peratures (Vol. 991). Department of Commerce, National Bureau of Standards.
Baleshan, B. and Mahendran, M. (2016). Experimental study of LSF floor
systems under fire conditions. Advances in Structural Engineering.
doi:10.1177/1369433216653508.
Bandula Heva, Y. and Mahendran, M. (2013). Flexural-torsional buckling tests of
cold-formed steel compression members at elevated temperatures. Steel and
Composite Structures, Vol. 14, No. 3, pp. 205–227.
89
90 References
Dias, Y., Mahendran, M. and Keerthan, P. (2019b). Full-scale fire tests of steel and plas-
terboard sheathed web-stiffened stud walls. Thin-Walled Structures, Vol. 137,
pp. 81–93.
European Committee for Standardization (CEN) (2002). Eurocode 1: Actions on
structures – Part 1–2: General actions – Actions on structures exposed to fire.
European Committee for Standardization, Brussels, Belgium.
European Committee for Standardization (CEN) (2005). EN 1993-1-2, Eurocode 3:
Design of steel structures – Part 1.2: General rules – Structural fire design.
European Committee for Standardization, Brussels, Belgium.
European Committee for Standardization (CEN) (2006a). EN 1993-1-3 2006, Eurocode
3: Design of steel structures – Part 1–3: General rules – Supplementary rules for
cold-formed members and sheeting. European Committee for Standardization,
Brussels, Belgium.
European Committee for Standardization (CEN) (2006b). EN 1993-1-5: Eurocode
3: Design of steel structures – Part 1–5: Plated structural elements. British
Standards Institution, London, UK.
Feng, M. (2003). Numerical and experimental studies of cold-formed thin-walled steel
studs in fire. PhD thesis, University of Manchester.
Feng, M. and Wang, Y.C. (2005). An experimental study of loaded full-scale cold-
formed thin-walled steel structural panels under fire conditions. Fire Safety
Journal, Vol. 40, pp. 43–63.
Feng, M., Wang, Y.C. and Davies, J.M. (2003a). Structural behaviour of cold-formed
thin-walled short steel channel columns at elevated temperatures, part 1:
Experiments. Thin-Walled Structures, Vol. 41, No. 6, pp. 543–570.
Feng, M., Wang, Y.C. and Davies, J.M. (2003b). Structural behaviour of cold-formed
thin-walled short steel channel columns at elevated temperatures, part 2: Design
calculations and numerical analysis. Thin-Walled Structures, Vol. 41, No. 6,
pp. 571–594.
Feng, M., Wang, Y.C. and Davies, J.M. (2003c). Thermal performance of cold-formed
thin-walled steel panel systems in fire. Fire Safety Journal, Vol. 38, No. 4,
pp. 365–394.
Feng, M., Wang, Y.C. and Davies, J.M. (2003d). Axial strength of cold-formed thin-
walled steel channels under non-uniform temperatures in fire. Fire Safety
Journal, Vol. 38, pp. 679–707.
Franssen, J.M. and Gernay, T. (2017). Modeling structures in fire with SAFIR®:
Theoretical background and capabilities. Journal of Structural Fire Engineering,
Vol. 8, No. 3, pp. 300–323.
Gerlich, J.T., Collier, P.C.R. and Buchanan, A.H. (1996). Design of light steel‐framed
walls for fire resistance. Fire and Materials, Vol. 20, No. 2, pp. 79–96.
Ghazi Wakili, K., Koebel, M., Glaettli, T. and Hofer, M. (2015). Thermal conductivity
of gypsum boards beyond dehydration temperature. Fire and Materials, Vol. 39,
pp. 85–94.
Gunalan, S., Bandula Heva, Y. and Mahendran, M. (2014). Flexural-torsional buck-
ling behaviour and design of cold-formed steel compression members at elevated
temperatures. Engineering Structures, Vol. 79, pp. 149–168.
Gunalan, S., Bandula Heva, Y. and Mahendran, M. (2015). Local buckling studies of
cold-formed steel compression members at elevated temperatures. Journal of
Constructional Steel Research, Vol. 108, pp. 31–45.
92 References
Gunalan, S., Kolarkar, P.N. and Mahendran, M. (2013). Experimental study of load
bearing cold-formed steel wall systems under fire conditions. Thin-Walled
Structures, Vol. 65, pp. 72–92.
Ha, T.H., Cho, B.H., Kim, H. and Kim, D.J. (2016). Development of an efficient
steel beam section for modular construction based on six-sigma. Advances
in Material Science and Engineering, Vol. 2016, Article ID 9687078.
doi:10.1155/2016/9687078.
Hanna, M.T., Abreu, J.C.B., Schafer, B.W. and Abu-Hamd, M. (2015). Post-fire buck-
ling strength of CFS walls sheathed with magnesium oxide or ferrocement
boards. Proceedings of the Annual Stability Conference, Structural Stability
Research Council, Nashville, TN, March 24–27, 2015.
Iding, R., Bresler, B. and Nizamuddin, Z. (1977). FIRES-T3, A computer program for
the fire response of structures-thermal. Report No. UCB FRG 77, 15.
ISO 834-1 (1999). Fire resistance tests – Elements of building construction, Part 1:
General requirements. International Organization for Standardization, Geneva,
Switzerland.
Jatheeshan, V. and Mahendran, M. (2016a). Experimental study of cold-formed steel
floors made of hollow flange channel section joists under fire conditions. ASCE
Journal of Structural Engineering, Vol. 142, No. 2, p. 04015134.
Jatheeshan, V. and Mahendran, M. (2016b). Fire resistance of LSF floors made of hol-
low flange channels. Fire Safety Journal, Vol. 84, pp. 8–24.
Kaitila, O. (2002). Imperfection sensitivity analysis of lipped channel columns at
high temperatures. Journal of Constructional Steel Research, Vol. 58, No. 3,
pp. 333–351.
Kankanamge, N.D. and Mahendran, M. (2011). Mechanical properties of cold-formed
steels at elevated temperatures. Thin-Walled Structures, Vol. 49, pp. 26–44.
Keerthan, P. and Mahendran, M. (2011). New design rules for the shear strength
of LiteSteel beams. Journal of Constructional Steel Research, Vol. 67,
pp. 1050–1063.
Keerthan, P., Mahendran, M. and Frost, R.L. (2013). Fire safety of steel wall systems
using enhanced plasterboards. Proceedings of the 19th International CIB World
Building Congress 2013, Brisbane, Australia, pp. 1–12.
Kesawan, S. and Mahendran, M. (2015). Predicting the performance of LSF walls made
of hollow flange sections in fire. Thin-Walled Structures, Vol. 98(A), pp. 111–126.
Kesawan, S. and Mahendran, M. (2016). Thermal performance of load‐bearing walls
made of cold‐formed hollow flange channel sections in fire. Fire and Materials,
Vol. 40, No. 5, pp. 704–730.
Kesawan, S. and Mahendran, M. (2017a). Fire performance of LSF walls made of hol-
low flange channel studs. Journal of Structural Fire Engineering, Vol. 8, No. 2,
pp. 149–180.
Kesawan, S. and Mahendran, M. (2017b). A review of parameters influencing the fire
performance of light gauge steel framed walls. Fire Technology. doi:10.1007/
s10694-017-0669-8.
Kodur, V.K.R. and Sultan, M.A. (2006). Factors influencing fire resistance of load-
bearing steel stud walls. Fire Technology, Vol. 42, pp. 5–26.
Kolarkar, P.N. and Mahendran, M. (2008). Thermal performance of plasterboard lined
steel stud walls. Proceedings of the 19th International Specialty Conference on
Cold Formed Steel Structures 2008, St. Louis, MO, pp. 517–530.
References 93
95
96 Index
L scope, 9
uniform temperature distribution, 76
limiting temperature method, 76–77 prefabricated structural units, 16–18
M S
methods of building construction, 2 section profiles
modular construction, 16 effects on fire resistance, 22, 36, 46
modular units, 16–18 hollow flange sections, 16
innovation, 14
N mechanical properties, 66–68
thermal properties, 65
non-uniform temperature thin-walled steel sections, 12
design method, 76 steel
simplification, 30 grade, 1
numerical modelling joists, 22–23
heat transfer, 59 mechanical properties, elevated
structural behaviour, 64 temperatures, 66–68
sheathing, 23, 42
P studs, 22–23
thickness range, 1
panel construction, 3
performance based design methods T
general, 75
non-uniform temperature distribution, thermal bowing, 82
76–78 thickness range, 1