You are on page 1of 12

International Journal of Steel Structures Online ISSN 2093-6311

https://doi.org/10.1007/s13296-020-00388-4 Print ISSN 1598-2351

Effect of Temperature‑Induced Moment‑Shear Interaction on Fire


Resistance of Steel Beams
Mohannad Zeyad Naser1   · Venkatesh Kodur2

Received: 8 March 2019 / Accepted: 29 July 2020


© Korean Society of Steel Construction 2020

Abstract
The interaction between bending and shear effects in steel beams can be amplified under fire conditions due to rapid deg-
radation in strength and stiffness properties of steel, together with temperature-induced local instability effects. This paper
presents temperature-induced moment-shear (M-V) interaction phenomenon in compact (Class 1) steel beams. Results
generated from numerical studies are utilized to quantify the effects of temperature-induced critical parameters influencing
moment-shear interaction, shear and flexural sectional capacity, as well as instability in steel beams under fire conditions.
The major findings of this work are two folds: (1) occurrence of temperature-induced instability adversely reduces shear
capacity, as compared to flexural capacity, and (2) this rapid degradation in shear capacity trigger moment–shear interaction
phenomenon at elevated temperatures. Eventually, this shifts failure mode in steel beams towards a shear dominant failure
mechanism on the interaction envelope.

Keyword  Fire · Moment–shear interaction · Steel beams · Instability

Abbreviations ν Poisson’s ratio


Mu Applied bending moment b, t Width and thickness of plates subjected to bending
Vu Applied shear force h, tw Width and thickness of plates subjected to shear
Mp Plastic flexural capacity k Plate buckling coefficient
Vp Plastic shear capacity fy Yield strength of steel section
Myf Yield moment considering the flanges only Zx Plastic section modulus
Ag Gross-sectional area fy,T Yield strength of steel section at temperature
bf Flange width τyw Shear yield strength of the steel web
hw Clear web depth d Overall depth for hot-rolled beams
r Radius of fillet at the web-flange joint Cv Web shear coefficient that depends on slenderness
tw Web thicknesses of web
tf Flange thicknesses
σcr Critical stress for pure bending
τcr Critical stress for pure shear 1 Introduction
E Elastic modulus
In most building applications, steel beams are primarily
subjected to load effects arising from bending moment, and
* Mohannad Zeyad Naser thus main consideration in design of beams is to satisfy
mznaser@clemson.edu
http://www.mznaser.com bending requirements under both ambient and fire condi-
tions (Zentz 2002; Crisan and Dubina 2016). However, in
Venkatesh Kodur
kodur@egr.msu.edu certain scenarios, shear effects can dominate response of
steel beams especially when subjected to certain loading
1
Glenn Department of Civil Engineering, Clemson configurations such as high concentrated (point) loads act-
University, Clemson, SC 29634, USA ing on beams, as in the case of transfer girders and beams
2
Civil and Environmental Engineering, Michigan State connecting to offset columns in buildings (Hall 1954).
University, 3546 Engineering Building, 428 S. Shaw Lane, Also, shear effects can be significant in beams with slender
East Lansing, MI 48824‑1226, USA

13
Vol.:(0123456789)
International Journal of Steel Structures

webs, such as built-up sections and deep beams. Recent 2 Response of Steel Beams Under Fire
studies have highlighted the susceptibility of steel beams Conditions
and structures to shear and interaction-based failure, espe-
cially under fire conditions (Naser and Kodur 2016, 2017; In case of beams in buildings, level of bending moment
Aziz et al. 2015; Reis et al. 2016a, b, 2017a). (as compared to moment capacity) is usually much higher
In a general sense, it is possible that a beam can be than that of shear force (when compared to shear capac-
subjected to bending moment alone, but not to shear alone ity). Hence, in many cases shear effects are negligible
since shear is the derivative (i.e. rate of change) of bending and bending dominates behavior and failure of beams.
moment. As a result, beams are often subjected to combi- Figure 1a illustrates response and failure mechanism of a
nation of bending and shear. In critical regions of a beam, typical steel beam subjected to dominant bending moment
closer to location of maximum shear force and bending and exposed to fire. It can be seen that failure of this beam
moment, interaction of bending moment and shear effects occurs once bending moment exceeds reduced level of
can accelerate plastification in the beam (and eventually moment capacity resulting from temperature effects. In a
failure). Despite this “natural” vulnerability to pre-mature similar manner, failure of a beam loaded with high shear
failure, provisions in most design standards such as AISC forces (and minor bending moment) occurs once the shear
(AISC, Steel Construction Manual 2011), AASHTO capacity falls below the level of applied shear force (see
(AASHTO LRFD Bridge Design Specifications 2017) and Fig. 1b).
AS 2327.1 (Standards Australia International 2003) do not There could also be a third scenario in which failure
specifically account for combined effects of moment and occurs due to interaction of bending and shear stresses
shear under fire conditions. located at transition zones. A transition zone is defined as
In fact, current AISC and AASHTO design specifi- a section (in a beam) where resistance mechanism changes
cations neglect moment–shear interaction in hot-rolled from flexure-based into shear-based dominant loading
W-shaped steel sections whether at ambient or fire condi- mechanism. This situation, wherein interaction between
tions. Only in the case of Eurocodes 3 (2005), dealing with moment and shear can dominate response of beams, is
steel structures, moment–shear interaction in W-shaped illustrated by tracing response of a typical steel beam from
beams at room temperature is to be considered under loading stage to failure under fire exposure as shown in
these two conditions: (1) when applied shear force does Fig. 1c. This figure shows a simply supported beam sub-
not induce web buckling, and (2) when shear force exceeds jected to a uniformly distributed load and two point loads
half the plastic shear resistance (Eurocode 3 2005). When and these point loads produce varying bending moment
moment–shear interaction is deemed to occur, Eurocode 3 and shear force across the span of the beam.
applies a reduction to moment capacity to allow the web to As can be seen in Fig. 1c, there are three different criti-
be fully utilized in resisting shear force (Eurocode 3 2005). cal sections. The first critical section, located at mid-span
The same Eurocode provisions allow extending room of the beam, is where peak bending moment occurs and
temperature design expressions to fire conditions if appro- the failure can occur at this section once temperature-
priate material property reduction factors are applied to induced degradation in moment capacity reaches below
account for temperature-induced degradation in strength the level of bending moment due to applied loading. The
and stiffness. Eurocode 3 provides recommendations on second critical section, located close to end supports, is
degradation of steel properties as a function of temperature mainly dominated by shear effects, and failure at this sec-
and assumes degradation in modulus of elasticity and yield tion could occur when temperature-induced degradation
strength of steel to start at 150 and 400 °C, respectively. in shear capacity reaches below the level of shear force.
As a result, temperature-induced web buckling, which In between these two critical sections lays a transition
is influenced by degradation in modulus of elasticity, zone where bending and shear effects are quite high (but
could occur at 150 °C, prior to any degradation in yield do not reach maximum values), and the combined effects
strength of steel (which starts at 400 °C). This contradicts can lead to failure in this region. The effect of combined
key design aspects that promote a ductile failure through forces, together with temperature-induced degradation
web yielding over shear buckling. Hence, extendibility of in sectional moment and shear capacity, combined with
Eurocode 3 moment–shear interaction design expressions temperature-induced instability, significantly complicates
to fire conditions can be debatable. stress distribution, load path, and failure mechanisms.
In order to explore the underlying mechanics of such A number of previous studies focused on moment–shear
phenomenon, this paper investigates effect of load ratio interaction effects at ambient conditions (Zentz 2002;
and temperature-induced instability on the development of Crisan and Dubina 2016). Results from these studies have
moment–shear interaction in fire exposed W-shaped hot- shown that interaction between bending moment and
rolled compact (Class 1) steel beams.

13
International Journal of Steel Structures

Fig. 1  Variation of bending UDL


moment and shear force in
beams under fire conditions

+ve
(a) Bending failure (shear effects are minor)

Point load Point load

+ve

- ve
(b) Shear failure (bending effects are minor)

Point load UDL Point load

+ve

-ve

Moment-shear Moment-shear
dominant region dominant region

(c) Failure due to moment-shear interaction

13
International Journal of Steel Structures

shear force is quite weak at ambient conditions due to two in beams loaded with high-moment and high-shear loading
main reasons, (1) bending effects govern failure of simply could lead to unconservative designs specifically at the junc-
supported beams, i.e. beam plastifies upon reaching its ture of web and flange where normal and shear stresses can be
flexural capacity which occurs much earlier than reaching substantial. Despite these findings, current editions of AISC
shear capacity, and (2) compact webs tend to yield before and AASHTO continue to neglect moment–shear interaction
they buckle when subjected to combination of moment effects.
and shear loading. While these observations hold true at Unlike North American codes, Eurocode 3 still requires
ambient conditions, the same observations may not be checking combined effects of flexural and shear interac-
consistent with results from recent studies of fire exposed tion in the design of steel beams. Eurocode 3 accounts for
steel beams (Aziz et al. 2015; Naser and Kodur 2017). To moment–shear interaction in cases where, (1) shear force
this date, very little research has been carried out on the does not induce shear buckling of the web plate, and (2)
interaction of bending and shear effects in beams under when the shear force exceeds half the plastic shear resistance
fire conditions (Naser 2016; Kodur and Naser 2018; Reis of that particular section. When moment–shear interaction
et al. 2017b). occurs, Eurocode simply reduces the moment capacity to
some degree using a shear interaction coefficient, ρ, such
that:

3 Moment–Shear Interaction at Elevated Mp = 1 − 𝜌2 fy Z (3)


( )

Temperatures
where, ρ =  V u − 1 = − 1 ≥ 0.
2V 2Vu

p Av fy ∕ 3
In order to quantify moment–shear interaction phenomenon, In Eq. 3, the plastic shear capacity, Vp, is valid for bf
Basler (1961) developed the following interaction equation tf/tw hw ≥ 0.6 as long as the average shear stress in the
mainly derived for slender webs at ambient temperature. section does not cause web yielding, and Av = Ag – 2bF
)2 )2 tf + (tw + 2r)tf ≥ hwtw, where Ag is the gross-sectional area, bf
Mu − Myf
( (
Vu is the flange width, hw is the clear web depth, r is the radius
+ =1 (1)
Mp − Myf Vp of fillet at the web–flange joint and tw and tf are web and
flange thicknesses, respectively.
where, Mu and Vu are the applied bending moment and
Equation 3 has been revised based on results from tests
shear force, respectively, Mp and Vp are the plastic flexural
on steel beams that showed web to be able to carry a con-
and shear capacity, respectively and Myf is the yield moment
siderable shear even when steel sections reach the plastic
considering the flanges only.
moment. Thus, Eurocode 3 provisions now combine Eqs. 1
Equation 1 has gone through number of revisions and an
and 3 into Eq. 4 and these equations are intended for plastic
improved equation (Eq. 2) was finally adopted by the AISC
failure analysis:
(2011) and AASHTO (2017) specifications.
)2
Mf
( )(
Mu V Mu 2Vu
+ 0.625 u = 1.375 + 1− −1 =1 (4)
Mp Vp (2) Mp Mp Vp

However, the above expression, along with moment–shear All of the above presented equations were derived from
interaction phenomenon, was discarded in recent editions of classical analysis on flat plates with different aspect ratios
AISC manual as a result of a comprehensive survey com- and boundary conditions. The fundamental form of these
missioned by American Iron and Steel Institute (AISI), the equations is given by Eq. 5 and simply represents a “load-
Federal Highway Administration (FHWA), and the Ameri- to-capacity” ratio:
can Society of Civil Engineers (ASCE) and conducted by ( )2 ( )2
𝜎 𝜏
White et al. (2008). + =1 (5)
𝜎cr 𝜏cr
White et al. (2008) statistically evaluated moment–shear
interaction effects reported in numerous experiments. Their where σ and τ are the applied bending and shear stress,
analysis concluded that moment–shear interaction either does respectively, σcr is the critical stress for pure bending and
not occur or its effect is minor and can be neglected. White τcr is the critical stress for pure shear. Both of these critical
et al. (2008) recommendations were founded based on the stresses are calculated as:
fact that it is possible to independently consider the effect of
bending moment and shear force rather than a combination of k𝜋 2 E
these effects. However, a follow up study carried out by Lee
𝜎cr = (6)
12 1 − 𝜈 2 (b∕t)2
( )
et al. (2013) revealed that ignoring moment–shear interaction

13
International Journal of Steel Structures

k𝜋 2 E fire exposure zone, and (2) the presence of a concrete slab


𝜏cr = which, due to insulating properties of concrete, act as a heat
)( / )2 (7)
12 1 − 𝜈 2 h t
(
w sink that attracts much of the temperature in the top flange.
Temperature-induced degradation in modulus of steel can
where E is elastic modulus, ν is Poisson’s ratio, b, t, h, tw be calculated as the product of room temperature modulus
are width and thickness of plates subjected to bending and (E) by a reduction factor (β) applied to represent degradation
shear and k is a plate buckling coefficient which is a function in modulus at that temperature (as given in fire design stand-
of both plate aspect ratio and wavelength parameter related ards i.e. Eurocode 3). Hence, when temperature in the top
to restraint conditions along the longitudinal boundaries (i.e. flange is 250 °C, the degradation of elastic modulus in the
fixed or simply supported etc.). The plate buckling coef- flange (and top portion of the web) equals to E250°C = β × E = 
ficient of unstiffened web in bending and shear equals to 0.85 × 210 = 178.5 GPa (given that β = 0.85 at 250 °C). This
23.9 and 5.34, respectively (Lee et al. 2013; Ziemian 2010). stiffness can be 2.7 times higher than that of the bottom por-
Earlier studies assumed fixed–fixed or simply supported tion of the web assuming it has a temperature similar to that
boundary conditions when evaluating plate buckling coef- of the bottom flange i.e. 600 °C (E600°C = 0.31 × 210 = 65.1
ficient and critical buckling stress in steel beams (Lee et al. GPa) (Eurocode 3 2005). This large variation in modu-
2013; Ziemian 2010; Kusuda and Thurlimann 1958). These lus causes reduction in overall plate stiffness and, most
studies have shown that following such assumptions lead to importantly, affect web restraint conditions (i.e. rigidity of
good correlation with experimental data at ambient condi- flanges).
tions. Unfortunately, the validity of these assumptions under When temperature-induced degradation in stiffness in
fire conditions has not been fully investigated or verified the bottom flange reaches this low level of initial stiffness,
(Timoshenko and Woinowsky-Krieger 1959; Selamet and this flange may not be able to provide the same level of
Garlock 2012). A closer look into Eqs. 6 and 7 reveals that restraint as that at ambient conditions (or even as that of
critical stress is a function of geometric features of web the top flange). In this case, the bottom portion of web has
plates as well as modulus of elasticity. Since modulus of more flexibility to laterally move due to the weaker restraint
elasticity is a temperature-dependent property, the critical provided by bottom flange. As a result, the assumption of
stress can then be significantly influenced by temperature having plate boundary conditions of fixed–fixed as pro-
rise and associated degradation in modulus of elasticity. posed by Chern and Ostapenko (1969) or simply-supported
In order to illustrate the effect of temperature rise on as proposed by Porter et al. (1975) and Lee and Yoo (1998)
bending and shear critical buckling stresses (i.e. corre- is questionable and may not hold true under fire exposure
spondingly moment–shear interaction), the response of conditions. In fact, the actual restraint conditions is of more
W-shaped fire-exposed steel beams is examined herein. complex nature (Sharp and Clark 1971).
When a W-shaped steel beam is exposed to fire (say standard In order to verify this observation, a simple steel plate
fire conditions), thermal gradients develop along the depth was modeled using a finite element model developed in
of beam cross section (comprising of top flange, web, and ANSYS. The plate is discretized using “shell 181” elements
bottom flange) due to the fact that the beam is often heated (a four-noded element with six degrees of freedom: transla-
from the bottom side (Kodur and Naser 2013). Due to the tions in the x, y, and z directions, and rotations about the x,
small thickness of the flanges, it can then be safe to assume y, and z axes). “Shell 181” was supplemented with elastic
that both bottom and top flange have uniform temperature material properties, namely modulus of elasticity and Pois-
rise (in which the bottom flange being much hotter than the son’s ratio. The plate had an aspect ratio of 4, thickness of
top flange). The web, on the other hand, can experience large 10 mm and was subjected to shear loading on all sides (and
thermal gradients due to its high depth and being located without any in-plane force or moment). A scale factor of
between a hot (bottom) flange and a cooler (top) flange. Depth/10,000 was selected to induce initial imperfection
Since stiffness properties of steel degrade with rise in to this plate (Garlock and Glassman 2014). Three loading
temperature, the stiffness of bottom flange will be much scenarios were studied. In the first scenario, the plate was
smaller than that in the top flange. Observations from recent analyzed at ambient conditions where steel modulus of elas-
fire tests on beams have shown that when temperature at the ticity equals to 210 GPa. The analysis shows that the critical
bottom flange reaches 600 °C, temperature at the top flange buckling stress for this plate equals to 428 MPa which is
(and top portion of the web) can be lower by 200–300 °C within 5% from that obtained using Eq. 7. Figure 2a shows
(Aziz et al. 2015; Naser and Kodur 2017). Thus, when tem- von-Mises stress distribution in this plate.
perature in the bottom flange reaches 600 °C, the tempera- In the second scenario, and in order to investigate effect
ture on the top flange can be conservatively assumed to be of temperature rise on stress distribution, the plate was sub-
250 °C. The lower temperature of the top flange is attrib- jected to uniform temperature of 600 °C. Hence, in this case,
uted to two factors, (1) farther distance from top flange to both top and bottom boundaries have similar stiffness. The

13
International Journal of Steel Structures

Fig. 2  Von-Mises stress distri-


bution in steel plate (web) under
various conditions

(a) Stress distribution of steel plate with modulus of (b) Deformed shape
210 GPa

Major out-of-plane
deformation

(c) Stress distribution of metal plate with modulus (d) Deformed shape
of 65.1 GPa

100°C
Minor out-of-plane
deformation

600°C

(e) Stress distribution of steel plate where section (f) Deformed shape
has varying modulus of 210, 157.5 and 65.1 GPa

stress distribution shown in Fig. 2b reveals that the plate has in stiffness (Aziz et al. 2015; Naser and Kodur 2017). More
similar stress distribution (but with lesser magnitude) to that importantly, results of this analysis infer that this variation in
in the first scenario. This due to the fact that the only differ- shear stress, once combined with bending effects and applied
ence between the two cases is the use of a reduced modulus to Eq. 6, can trigger changes in the moment–shear interac-
of elasticity of 65.1 GPa. tion behavior in fire-exposed steel beams that may trigger
The effect of thermal gradients was also studied as part premature failure.
of the third scenario in which the plate was assumed to
have three temperature fields of 100, 350 and 600 °C (in
top flange, web, and bottom flange, as shown in Fig. 2e). 4 Evaluation of Flexural and Shear Capacity
Results from the analysis shows that stress distribution (see Under Ambient and Fire Conditions
Fig. 2c, e) as well as out-of-plane deformations (see Fig. 2d,
f) are much different than that observed in the two scenarios As discussed earlier, the main objective of this paper is to
due to the variation in stiffness as a function of temperature quantify the effect of temperature-induced strength degra-
rise across the depth of the plate. These observations do not dation and instability effects on moment–shear interaction
rule out the possibility of tension field action developing behavior of steel beams exposed to fire conditions. For these
at elevated temperatures, especially when the strength-to- objectives to come through, the degrading flexural and shear
slenderness ratio of web becomes too small. These results capacity in steel beams need to be quantified first.
also agree with observations from fire tests and clearly show The current provisions for evaluating flexural capacity
that stress distribution and buckling behavior of a steel plate of beams under fire conditions are through extending room
changes as a function of temperature-induced degradation temperature design expressions available in AISC (2011)

13
International Journal of Steel Structures

and Eurocode 3 (2005). In these provisions, flexural fail- development of above equation can be found elsewhere
ure can occur when a beam becomes fully plastic (reaching (Kodur and Naser 2018).
plastic moment capacity) and once flexural capacity falls A general procedure for evaluating interaction effects
below the bending moment resulting from applied loading. of shear force and moment in fire exposed steel beams is
As such, flexural capacity of a W-shaped section is given as: proposed through steps outlined in the flow chart shown in
Fig. 3. The proposed procedure takes into account both flex-
Mp = fy Zx (8) ural and shear limit states to evaluate failure in steel beams.
where fy is the yield strength of steel section, and Zx is the The computed flexural and shear capacity (of a steel beam)
plastic section modulus. is compared against effects of applied bending moment and
For evaluating flexural capacity under fire conditions, shear force at various target temperatures (100, 200, 300 °C
codal provisions extend room temperature design proce- etc.). Once flexural and/or shear capacity drops below level
dure with due consideration to temperature-induced reduc- of bending moment and/or shear force, failure is said to
tion in yield strength of steel. All that is needed is to take occur. These steps outlined in the flow chart can be program-
into account appropriate reduction in yield strength of steel mable into a simple and robust computer code.
at a specified temperature (i.e. 100, 200, 300 °C etc.) and
this is given as:
5 Effect of Temperature‑Induced Capacity
Mfire = fy,T Zx (9) Degradation
where fy,T is the yield strength of steel section at tem- It can be inferred from above discussion that one of the
perature, T. major differences in behavior of steel beams under ambi-
This approach simplifies flexural capacity calculations at ent and fire conditions is the fact that strength and stiffness
elevated temperature and can be repeated at various tempera- properties of steel degrades with rise in temperature. Since
tures, to derive the moment capacity-temperature response flexural and shear capacity reduces relative to degradation
history. in strength properties, the reductions in properties can also
In the case of shear response, shear capacity at room tem- change “load-to-capacity” ratio (often of constant magnitude
perature can be evaluated as: at ambient conditions) leading to developing moment–shear
interaction under fire conditions. In order to quantify such
Vp = 𝜏yw dtw Cv (10)
ratio, this section investigates the effect of temperature-
where τyw is the shear yield strength of the steel web induced property degradation on moment–shear interaction
(τyw = 0.6 fyw ), tw is the thickness of the web, d is the overall of steel beams.
depth for hot-rolled beams, Cv is the web shear coefficient Discussion in Sect. 5 presented number of design expres-
that depends on slenderness of web. sions that can be applied to evaluate both flexural and shear
AISC provisions do not specifically state an extension capacities. Once evaluated, these capacities are compared to
of Eq. 10 to evaluate shear capacity-temperature response existing bending moment and shear force actions, mainly to
history. This is in contrast to the case of flexural capacity check if the “load-to-capacity” ratio satisfies interaction con-
evaluation, where AISC allows Eq. 8 to be extended to Eq. 9. ditions. If one (or more) of these conditions is not satisfied,
Naser and Kodur (2018) have looked into the extendibility of then moment–shear interaction is to be neglected. However,
Eq. 10 to fire conditions and found that this equation overes- it is possible for a beam not to experience interaction effects
timates shear strength in fire-exposed hot-rolled beams as it at ambient conditions, but to develop such effects under high
does not specifically account for temperature-induced insta- temperature conditions due to degradation in strength prop-
bility effects. As a result, they concluded that direct exten- erties (and corresponding degradation in moment and shear
sion of Eq. 10 cannot be applied to account for evaluation for capacity).
shear capacity under elevated temperature. They proposed Temperature raise in steel and associated degradation in
modified equation to account for yield strength degradation strength increases the ratio of bending moment-to-flexural
as well as temperature-induced instability effects (Kodur and capacity and shear force-to-shear capacity (such that fy,550 °C 
Naser 2018). <  < fy,25 °C → M550°C <  < M25 °C → Mu/M550 °C >  > Mu/M25 °C).
{ Since most codal provisions disregard moment–shear inter-
Vfire =
0.6fy Aw Cv , T < 150◦ C
(11) action in beams, and codes which account for such inter-
0.6fy,T 𝛽tw2 Cv,T , T ≥ 150◦ C action does not account for this phenomenon under fire
conditions, steel beams can be vulnerable and severely
where β is the critical slenderness parameter, Cv,T is
under-designed if subjected to moderate-to-high levels of
temperature-dependent web shear coefficient. Details on
bending moment and shear loading under fire conditions.

13
International Journal of Steel Structures

Fig. 3  Evaluation of flexural and shear capacity in fire exposed steel beams

In order to demonstrate the above observation, a sim- leads to significant degradation in strength properties in
ple example is carried out herein. In this example, a the range of 62% (fy,550 °C= β × fy,25 °C = 0.62 × 345 = 214 M
W18 × 40 made of Grade 345 MPa possessing flexural and Pa). As a result, moment and shear capacity of this beam
shear capacity of 443 kN.m and 752.4 kN is selected for will also reduce, such that:
analysis (AISC, Steel Construction Manual 2011). This
Mp,550◦ C = 0.62 × 443 = 274.7 kN m
beam is subjected to bending moment and shear force
equivalent to 40% of its moment and shear capacity (i.e.
Mu = 0.4 × Mp = 0.4 × 443 = 177 kN m and Vu = 0.4 × Vp = 0 Vp,550◦ C = 0.62 × 752.4 = 466.5 kN
.4 × 752.4 = 301 kN). Inferring to the discussion presented
in Sec. 4.0, both AISC and AASHTO provisions do not Since the magnitude of applied loading (Mu = 177 kN m
account for moment–shear interaction effects and hence and Vu = 301 kN) is often maintained under fire conditions,
according to these provisions, moment–shear interaction in the ratio of bending moment-to-flexural capacity and shear
the selected W18 × 40 beam is deemed minor and could be force-to-shear capacity at 550 °C increases (from 40%) to
neglected. In the case of Eurocode provisions, the applied about 64 and 65%, respectively:
shear force level equals to 40% and does not exceeds half Mu 177 Vu 301
the plastic shear resistance. As a result, the moment–shear = = 0.64, = = 0.65
Mp,550◦ C 274.7 Vp,550◦ C 466.5
interaction effects in this beam are also assumed to be minor
and could be neglected. While AISC and AASHTO provisions neglect interac-
In the case this beam is exposed to fire and temperature tion effects, revisiting Eurocode provisions to check for
in beam (steel) reaches 550 °C. This rise in temperature

13
International Journal of Steel Structures

moment–shear interaction shows that shear force exceeds in this beam is minor and lays in the low moment-low
half the plastic shear resistance at 550  °C, and hence shear region. However, with the rise in temperature, the
moment–shear interaction can occur. This simple example level of moment–shear interaction increases until failure
clearly shows that moment–shear interaction can develop occurs.
under fire conditions due to temperature-degradation Figure  4c shows the effect of increasing “load-to-
in strength properties. Figure 4a, b further trace “load- capacity” ratio for bending (or shear), with levels rang-
to-capacity” ratio in the selected beam as a function of ing between 25, 40 and 80%, as a function of temperature
elevated temperature. It can be seen from trends plotted rise on failure of steel beams. These results agree with
in these figures that as temperature rises, higher property that shown in Fig. 4a, and further demonstrate the rapid
degradation occurs in material properties and the applied rise in “load-to-capacity” ratio at temperatures higher than
“load-to-capacity” ratio increases which accelerates devel- 400 °C, occurring due to the rapid degradation in yield
opment of moment–shear interaction effects. In other strength of steel at temperatures exceeding 400 °C. Fig-
words, at ambient condition, the moment–shear interaction ure 4c also shows that failure (i.e. load/capacity ratio = 1)

Fig. 4  Effect of temperature- 152 mm


induced property reduction on mm

moment–shear interaction 152 mm


454 mm

13 mm

(a) Typical loading conditions


1.5

1.25

1
High shear-low High moment-
moment region high shear region
0.75

0.5 25°C
500°C
600°C
0.25 700°C
Low moment- High moment-low Simple failure envelope
low shear region shear region EC failure envelope
0
0 0.25 0.5 0.75 1 1.25 1.5

(b) Development of moment-shear interaction under fire conditions for load-to-ratio level
of 40%
1.5

1.25

0.75

0.5
Failure
25%
0.25 40%
80%
0
0 100 200 300 400 500 600 700 800 900

(c) Development of moment-shear interaction for various “load-to-capacity” ratios under


fire conditions

13
International Journal of Steel Structures

under a high load/capacity ratio can occur at lower temper- 6 Effect of Temperature‑Induced Instability
atures. For example, failure occurs at 635, 550 and 500 °C
in case of applied loading with magnitude of 25, 40 and Number of studies have shown that temperature-induced
80%, respectively. instability can start in steel beams at temperature as low as
In order to further investigate moment–shear interac- 150 °C (Naser 2016; Kodur and Naser 2018, 2017). This
tion behavior of beams, fire response of three W18 × 40 is due to degradation in modulus of steel at temperatures
beams, referred to as “Beam 1”, “Beam 2” and “Beam beyond 150 °C. Once modulus degrades, buckling capac-
3”, is traced when subjected to different patterns of load- ity of plates (web or flanges) reduces and causes additional
ing. These beams are subjected to high shear-low moment losses to shear (or flexural) capacity. More specifically,
(Beam 1), medium shear-medium moment (Beam 2), and recent fire tests and numerical studies have pointed out vul-
low shear-high moment (Beam 3), respectively. The high nerability of steel web to temperature-induced instability
shear-high moment and low shear-low moment loadings (Kodur and Naser 2018; Reis et al. 2017b). This vulner-
were calculated as 75 and 25%, respectively, of shear and ability arises from the fact that web is usually thinner and
flexural capacity of W18 × 40 section while the medium more slender than flanges. Further, steel webs are exposed
loading was equivalent to 40% of shear and flexural capac- to higher thermal (fire) loading; since web has larger surface
ity of the same section. area and exposed to the fire from two sides. Hence, strength
Figure 5 traces developed moment–shear interaction (and modulus) properties of steel in web can degrade at a
behavior of these beams at temperatures of 25, 500, 600 rapid rate than that in flanges. As a result, shear capacity
and 700 °C. It can be seen from plotted data that “Beam of steel beams can degrade at a much higher rate than flex-
1” and “Beam 3” reaches failure envelope at a lower tem- ural capacity since small area of web is main contributor to
perature as compared to “Beam 2” which is loaded with shear capacity (when compared to the area of two flanges). It
medium levels of shear and moment. This is due to the should be noted that previous studies have shown that losses
fact higher levels of applied loading develop larger level due to temperature-induced web instability effects can be as
of stresses within beam cross section, specifically in web high as 20–25% (Aziz et al. 2015; Kodur and Naser 2018).
(in case of high shear forces i.e. “Beam 1”) or in flanges The effect of temperature-induced instability on
(in case of high bending moment, i.e. “Beam 3”). At tem- moment–shear interaction of a typical steel beam (similar to
perature of 600 °C, these large stresses can reach reduced “Beam 1” shown in previous sections) can be seen in Fig. 6a.
yield strength of steel which leads to beam plastification. This figure shows that effect of temperature-induced instabil-
In order for the same magnitude of stresses to develop ity starts to be apparent at temperatures exceeding 150 °C
in “Beam 2”, further degradation to strength properties i.e. compact steel section transforms to a non-compact (and
need to take place which occurs at higher temperature than possibly into slender) section with further rise in temperature
that reached in “Beams 1 and 3”. Thus, “Beam 2” fails at (Naser and Kodur 2016). The difference in moment–shear
slightly higher temperature than that of “Beams 1 and 3”. interaction response by including instability effects and that

Fig. 5  Moment–shear interac-
tion in Beams 1, 2 and 3 1.75

1.5

1.25
Failure envelope
1 25°C
High shear-low 500°C
High moment-
moment region high shear region 600°C
0.75 700°C

0.5

0.25
Low moment- High moment-low
low shear region shear region
0
0 0.25 0.5 0.75 1 1.25 1.5 1.75

13
International Journal of Steel Structures

understanding on development of moment–shear interac-


tion under fire conditions, there is scope for further research
to extend this study towards tracing different states of
moment–shear interaction and exploring the use of fiber-
reinforced concrete (FRC) in concrete slabs to better stabi-
lize composite beams subjected to combined loading effects.
Finally, the following key conclusions can be drawn:

1. While moment–shear interaction in steel beams is minor


and can be neglected at ambient conditions, steel beams
can experience significant moment–shear interaction
(a) effects that can get amplified under severe fire condi-
tions due to the rapid degradation in shear capacity trig-
ger moment–shear interaction phenomenon at elevated
temperatures
2. Development of temperature-induced instability effects
can reduce moment capacity and shear capacity of steel
beams under fire conditions by about 20–25%.
3. Temperature-induced instability adversely reduces shear
capacity, as compared to flexural capacity; thus, effec-
tively shifting failure in steel beams towards a shear
dominant failure mechanism on the interaction envelope.
4. The adverse effects of temperature-induced sectional
instability are more apparent in the case of shear capac-
(b) ity, as oppose to flexural capacity. The occurrence of
such sectional instability can shift failure of compact
Fig. 6  High-temperature moment–shear interaction in beams with steel beams towards the shear dominant region on the
and without temperature-induced instability effects (PS. the shear moment–shear interaction envelope.
dominant region is highlighted)

of when instability is excluded is about 5% at temperature Acknowledgements  This material is based upon the work supported by
the National Science Foundation under Grant number CMMI-1068621
range between 150–500 °C. This variation then rapidly grows
to Michigan State University. Any opinions, findings, and conclusions
to 10 and 17% at temperatures of 600 and 700 °C, respectively. or recommendations expressed in this paper are those of the authors
Incorporating temperature-induced instability effects cause the and do not necessarily reflect the views of the sponsors.
beam to fail at slightly lower temperature of 603 °C (as com-
pared to 635 °C when the instability effect is neglected, see
Fig. 4b).
Data plotted in Fig. 6b also shows that occurrence of insta- References
bility further reduces shear capacity, as compared to flexural
capacity. In other words, temperature-induced instability shift AASHTO LRFD Bridge Design Specifications. ( 2017). American
Association of State Highway and Transportation Officials, Wash-
Vu/Vp ratio towards shear dominant failure on the interaction
ington, DC.
failure envelope (Vu/Vp = 1). As a result of this instability, the AISC, Steel Construction Manual. (2011). 14th Edition. American
beam fails in shear mode and not in simultaneous flexural/ Institute of Steel Construction, Chicago, Illinois.
shear mode (see highlighted portion of Fig. 6). Figure 6 also Aziz, E., Kodur, V. K. R., Glassman, J., & Garlock, M. (2015). Behav-
ior of steel bridge girders under fire conditions. Journal of Con-
shows that magnitude of instability effects can be substantial
structional Steel Research, 106, 11–22. https​://doi.org/10.1016/j.
at higher temperatures (and higher level of loading). jcsr.2014.12.001.
Basler, K. (1961). Strength of plate girders under combined bending
and shear. In Proceeding of ASCE, 87, (ST7), Reprint No. 187,
Fritz Laboratory Reports Paper 71.
7 Conclusions Cern, C., & Ostapenko, A. (1969). Ultimate strength of plate girder
under shear. Fritz Eng. Lab. Rep. No. 328.7, Lehigh University,
This paper presents temperature-induced capacity degra- Bethlehem, PA.
dation under combined effect of moment and shear in fire Crisan, A., & Dubina, D. (2016). Bending–shear interaction in
short coupling steel beams with reduced beam section. Journal
exposed steel. While this study presented a fundamental

13
International Journal of Steel Structures

of Constructional Steel Research, 122, 190–197. https​://doi. Porter, D. M., Rocky, K. C., & Evans, H. R. (1975). The collapse
org/10.1016/j.jcsr.2016.03.020. behaviors of plate girders loaded in shear. Structural Engineer-
Eurocode 3. (2005). Design of steel structures, Part 1–2: General rules- ing, 53, 313–325.
structural fire design. Document CEN, European Committee for Reis, A., Lopes, N., & Real, P. (2016a). Numerical study of steel plate
Standardization, UK girders under shear loading at elevated temperatures. Journal of
Garlock, M. E., & Glassman, J. D. (2014). Elevated temperature Constructional Steel Research, 117, 1–12.
evaluation of an existing steel web shear buckling analytical Reis, A., Lopes, N., & Real, P. (2016b). Shear–bending interaction in
model. Journal of Constructional Steel Research. https​://doi. steel plate girders subjected to elevated temperatures. Thin-Walled
org/10.1016/j.jcsr.2014.05.021. Structures, 104, 34–43.
Hall, W. J. (1954). Shear deflection of wide flange steel beams in the Reis, A., Lopes, N., & Real, P. (2017a). Design of steel plate gird-
plastic range. Urbana, IL: Wright Air Development Center, Uni- ers subjected to shear buckling at ambient and elevated tempera-
versity of Illinois. tures: Contribution from the flanges. Engineering Structures, 152,
Kodur, V. K. R., & Naser, M. Z. (2013). Effect of shear on fire response 437–451.
of steel beams. Journal of Constructional Steel Research, 97, Reis, A., Lopes, N., & Real, P. V. (2017b). Design of steel plate girders
48–58. https​://doi.org/10.1016/j.jcsr.2015.03.015. subjected to shear buckling at ambient and elevated temperatures:
Kodur, V. K. R., & Naser, M. Z. (2017). Effect of local instability on Contribution from the flanges. Engineering Structures, 152, 437–
fire response of steel beams. PSU Research Review. https​://doi. 451. https​://doi.org/10.1016/j.engst​ruct.2017.09.020.
org/10.1108/PRR-05-2017-0025. Selamet, S., & Garlock, M. (2012). Predicting the maximum com-
Kodur, V. K. R., & Naser, M. Z. (2018). Approach for shear capac- pressive beam axial force during fire considering local buckling.
ity evaluation of fire exposed steel and composite beams. Jour- Journal of Constructional Steel Research, 71, 189–201. https​://
nal of Constructional Steel Research, 141, 91–103. https​://doi. doi.org/10.1016/j.jcsr.2011.09.014.
org/10.1016/j.jcsr.2017.11.011. Sharp, M. L., & Clark, J. W. (1971). Thin aluminum shear webs. Jour-
Kusuda, T., & Thurlimann, B. (1958). Strength of wide-flange beams nal of the Structural Division.
under combined influence of moment, shear and axial force, Fritz Standards Australia, Composite structures. Part 1: Simply supported
Laboratory Reports Paper 1658. Bethlehem: Lehigh University. beams AS2327.1, Standards Australia International, Sydney,
Lee, S. C., Lee, D. S., & Yoo, C. H. (2013). Flexure and shear interac- 2003.
tion in steel I-girders. Journal of Structural Engineering, 139, Timoshenko, S. P., & Woinowsky-Krieger, S. (1959). Theory of plates
1882–1894. https:​ //doi.org/10.1061/(ASCE)ST.1943-541X.00007​ and shells. New York: McGraw-Hill.
46. White, D. W., Barke, M. G., & Azizinamini, A. (2008). Shear Strength
Lee, S. C., & Yoo, C. H. (1998). Strength of plate girder web panels and moment–shear interaction in transversely stiffened steel
under pure shear. Journal of Structural Engineering, 124, 184– I-girders. Journal of Structural Engineering, 134, 1437–1449.
194. https:​ //doi.org/10.1061/(ASCE)0733-9445(1998)124:2(184). https​://doi.org/10.1061/(ASCE)0733-9445(2008)134:9(1437).
Naser, M. Z. (2016). Response of steel and composite beams subjected Zentz, A. L. (2002). Experimental moment–shear interaction and TFA
to combined shear and fire loading. East Lansing: Michigan State behavior in hybrid plate girders. Doctoral dissertation, University
University. of Missouri-Columbia, USA.
Naser, M. Z., & Kodur, V. K. R. (2016). Factors governing onset Ziemian, R. D. (2010). Guide to stability design criteria for metal
of local instabilities in fire exposed steel beams. Journal of structures. Hoboken: Wiley.
Thin-Walled Structures, 98, 48–57. https​://doi.org/10.1016/j.
tws.2015.04.005. Publisher’s Note Springer Nature remains neutral with regard to
Naser, M. Z., & Kodur, V. K. R. (2017). Comparative fire behavior jurisdictional claims in published maps and institutional affiliations.
of composite girders under flexural and shear loading. Journal
of Thin-Walled Structures, 116, 82–90. https​://doi.org/10.1016/j.
tws.2017.03.003.

13

You might also like