You are on page 1of 15

Journal of Constructional Steel Research 136 (2017) 162–176

Contents lists available at ScienceDirect

Journal of Constructional Steel Research


journal homepage: www.elsevier.com/locate/jcsr

Stability and design of continuous steel beams in the strain-hardening range MARK
a,* b
A.S.J. Foster , L. Gardner
a
University College London, London, UK
b
The Imperial College of Science, Technology and Medicine, London, UK

A B S T R A C T

We examine the lateral stability implications of allowing for strain-hardening in the design of continuous steel
beams through a programme of experiments, numerical modelling and parametric studies. Six tests are
performed on continuous beams that are partially restrained against lateral torsional buckling. Restraint
spacings are chosen to give non-dimensional lateral torsional slenderness values of 0.3 and 0.4. Bending
resistances determined by the continuous strength method (CSM), which takes into account strain-hardening, are
shown to be exceeded. We present a numerical model validated against the laboratory test data to conduct
parametric studies that investigate the range of slenderness values for which the CSM is beneficial, and to
examine the interaction between beam segments with unequal loads. Neglecting beneficial interactions between
neighbouring beam segments, to achieve the degree of rotation capacity required for Class 1 sections designed
using the continuous strength method, closer restraint spacing than the minimum specified by EN 1993-1-1
(2005) is required. A basic design approach is presented that incorporates a limiting lateral torsional slenderness
for the CSM of 0.2, and a simple transition function from bending resistances predicted by simple plastic theory
to those predicted by the CSM.

1. Introduction that the elastic and inelastic moment distributions are the same.
However, for statically indeterminate structures, the elastic in-plane
The lateral stability of beams is an important consideration in distribution of moment is not necessarily the same as the final inelastic
structural steel design. The current practice in European design codes in-plane distribution and so is not strictly valid for determining the
(EN 1993-1-1 (2005)), when designing in the inelastic range, is to inelastic buckling load of such a structure [3].
specify a maximum lateral torsional buckling slenderness λLT below Numerous investigations into the critical buckling loads of elastic
which the effects of lateral instability can be ignored and the cross- continuous beams [4–10] and inelastic continuous beams [3,11-13]
section can fully yield; beyond this limit, buckling instabilities can have been carried out, typically considering behaviour up to the plastic
arise. This method assumes an elastic-perfectly plastic (EPP) stress- design moment. This paper examines the lateral stability of continuous
strain curve. In a newly proposed design method, referred to herein as beams in the strain-hardening range. The objective is to establish
the continuous strength method (CSM) [1], strain-hardening is allowed practical limitations on lateral restraint spacing such that the effects of
for in the material model, with the limiting strain being dependent lateral torsional buckling can be ignored and the full indeterminate
upon the local plate slenderness λ p . In this research we examine the strain-hardened cross-section capacity can be achieved.
implications of allowing for strain-hardening in the design of contin- To these ends, we conduct a series of experiments on continuous
uous beams and hence the limiting value of λLT . beams with variations in restraint spacings. We present a numerical
When designing an unbraced statically determinate I-beam in the model for inelastic lateral torsional buckling that reproduces and
inelastic range, the limiting value of non-dimensional slenderness λLT , extends the test data in a parametric study which is then used to
below which the effects of lateral torsional buckling can be ignored, is inform and develop a preliminary design format.
based upon knowledge of the inelastic critical moment. The inelastic
buckling load of statically determinate structures can be determined
analytically using the method proposed by Foster and Gardner [2],
which is based upon modifications of elastic predictions. Key to this
simplified analysis is that there is no plastic redistribution, meaning

*
Corresponding author.
E-mail address: a.foster@ucl.ac.uk (A.S.J. Foster).

http://dx.doi.org/10.1016/j.jcsr.2017.05.006
Received 9 November 2016; Received in revised form 25 April 2017; Accepted 10 May 2017
Available online 26 May 2017
0143-974X/ © 2017 Elsevier Ltd. All rights reserved.
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Table 1
Summary tensile and compressive coupon test data.

Coupon E fy fu ϵf ϵu

designation (N/mm2) (N/mm2) (N/mm2) (%) (%)


305 ×127 ×48-TW 198700 402 528 24.2 19.1
305 ×127 ×48-TF 191700 391 534 30.8 17.1
305 ×165 ×40-TW 201340 459 599 21.6 13.7
305 ×165 ×40-TF 204200 436 585 25.1 15.7
305 ×127 ×48-CW 207600 408 − − −
305 ×127 ×48-CF 213100 408 − − −
305 ×165 ×40-CW 204223 482 − − −
305 ×165 ×40-CF 218700 454 − − −

In this equation Wy is the major axis plastic section modulus for


Class 1 and 2 cross-sections, the elastic section modulus for Class 3
cross-sections and an effective section modulus for Class 4 cross-
sections, fy is the material yield strength and Mcr is the elastic critical
moment for lateral torsional buckling, which is a function of member
length L (EN 1993-1-1, 2005). Where λLT ≤ 0.4 , the design buckling
resistance moment of the member Mb,Rd may be taken as the design
bending resistance Mc,Rd of the cross-section, assuming that γM0 = γM1.
The yielded zones that characterise the formation of plastic hinges
weaken the cross-section and member with regard to lateral torsional
buckling [14]. As such, EN 1993-1-1 stipulates that the length between
lateral restraints must not exceed the stable length Lstable. For uniform
beam segments with I- or H-sections with h/tf < 40ϵ under linear
moment without significant axial compression, the stable length may be
taken as:
⎧ 35ϵiz for 0.625≤ψ ≤1
L stable =⎨
⎩ (60−40ψ )ϵiz for−1≤ψ ≤0.625 (2)
in which ϵ = 235/ fy , ψ is the ratio of end moments in the segment, h is
the overall height of the cross section and tf is the plate thickness of the
Fig. 1. Buckling modes for a two-span continuous beam.
flange. More detailed provisions exist for members with axial force and
this is defined as Lm in Annex BB.3 of EN 1993-1-1.

2.2. Load interaction relationships

Of relevance to this study is the work carried out on continuous


beams [4-8,15] whose response differs from simply supported beams as
a result of the interactions between adjacent spans. Trahair [16–18]
demonstrates that for a two-span continuous beam, the behaviour when
fully loaded lies in between the cases when one span is unloaded. Fig. 1
presents the various buckling modes of a two-span continuous beam
that is prevented from twisting and deflecting at its supports, which for
a given load position, vary according to the relative magnitudes of the
loading intensities QAB and QBC. In Fig. 1b, only segment AB is loaded
and during buckling it is elastically restrained by the unloaded segment
BC, with the opposite being the case in Fig. 1c. If the loads are equal
(Fig. 1c), the two segments are effectively independent of one another
and so buckle as two simply supported beams with effective lengths LAB
and LBC.
Fig. 2. Elastic critical load combinations for a two-span continuous beam. The effect of proportional variations in QBC relative to QAB on the
critical buckling loads for each span on the continuous beam in Fig. 1a
2. Key design aspects is plotted in Fig. 2. In between the extreme cases of QBC = 0, QAB = 0
and QBC = QAB a series of intermediate buckling loads will develop,
2.1. Lateral restraint spacing producing a complete and non-linear interaction relationship for the
critical load combinations on each span. On the basis of this accurate
A non-dimensional slenderness limit λLT = 0.4 , represented graphi- relationship, a simplified linear relationship was proposed by Tra-
cally as a plateau, below which the effects of lateral torsional buckling hair [7], constructed by interpolating between the loads corresponding
can be ignored is defined by EN 1993-1-1 (2005) as: to the critical buckling modes in Fig. 1 b - 1d. For use in routine design,
the critical points corresponding to QBC = 0, QAB = 0 and QBC = QAB
Wy fy required for the interpolations must be determined using tabulations
λLT =
Mcr (1) provided by Trahair [6], but this approach has not gained widespread
acceptance in practice; instead, the conservative method of neglecting

163
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Table 2
Summary of geometric properties for the continuous beam test specimens.

Test designation λp λp e0 ω0 b hw tf tw

(flange) (web) (mm) (mm) (mm) (mm) (mm) (mm)


305 × 127 × 48, λLT = 0.4 , R 0.31 0.30 1.0 0.100 127.51 283 14.33 8.73
305 × 127 × 48, λLT = 0.4 , K1 0.31 0.30 1.0 0.100 127.29 283 14.46 8.68
305 × 127 × 48, λLT = 0.4 , K2 0.31 0.30 1.0 0.100 127.59 282 14.77 9.07
305 × 165 × 40, λLT = 0.4 , R 0.57 0.44 10.5 0.083 165.86 284 9.93 6.54
305 × 165 × 40, λLT = 0.4 ,K1 0.57 0.44 2.5 0.083 166.81 286 10.44 6.40
305 × 165 × 40, λLT = 0.3, K1 0.57 0.44 6.0 0.083 165.49 287 10.05 6.53

Inclinometer Load spreader L.V.D.T


Strain gauge
and roller Load transfer
rods

Load transfer Central support Roller


plate and load cells system

Laboratory floor

Loading
Laboratory basement jack Bolt
Load cell

(a) Illustration of test configuration and instrumentation.


Loading point L1 L2

Test specimen
Lateral restraint

(b) Plan view of test configuration illustrating locations of lateral restraints.


F F
Msup

Mspan
R2 R1 R2

(c) Schematic bending moment diagram.


Fig. 3. Schematic illustrations of continuous beam tests.

continuity between spans is generally preferred [19], or a numerical ϵcsm 0.25


= 3.6
but ≤ 15 for λ p ≤ 0.68
analysis of the full system is conducted. ϵy λp (4)

in which ϵy is the yield strain of the extreme outer fibre of the cross-
2.3. The continuous strength method (CSM) section material and λ p is the local cross-section slenderness, defined in
Eq. (5) as:
The continuous strength method is a design approach for structural
steel elements that considers the influence of strain-hardening. λp = (fy /σcr ) (5)
Presently, cross-section level design equations are available for mem-
bers in bending and compression [20]. Under examination in this In this equation σcr is the lesser of the elastic buckling stress of the
research is the CSM bending resistance function, which we denote as cross-section, or its most slender constituent plate element.
Mcsm,Rd, applicable for λ p ≤ 0.68 and is defined in Eq. (3) as: A key assumption of current plastic design is a perfectly plastic post-
yield response, whereupon attaining the plastic moment capacity, no
Wpl fy ⎛ ⎛ ⎞ ⎛ W ⎞ ⎛ϵ ⎞ −α ⎞
Mcsm,Rd = ⎜ 1 + Esh Wel ⎜ ϵcsm − 1⎟ − ⎜1− el ⎟ ⎜ csm ⎟ ⎟⎟ further increases in capacity occur and infinite rotations can be
γM0 ⎜⎝ E Wpl ⎝ ϵ y ⎠ ⎝ Wpl ⎠ ⎝ ϵ y ⎠ ⎠ (3) achieved. Introducing strain-hardening prevents plastic hinges from
rotating at a constant moment, but significant increases in capacity
In this equation E is the modulus of elasticity, Esh is the strain- beyond Mpl are possible for cross-sections with low λ p values [20]. To
hardening modulus taken as E/100 for carbon steel, α is an exponent reconcile the CSM with the familiar virtual work format of simplified
that relates to the cross-section geometry with a value of 2 for I-sections plastic design, the CSM moment capacities at individual plastic hinge
in bending about the major axis, Wel and Wpl are the elastic and plastic locations can be modified to reflect relative deformation de-
section moduli and ϵcsm/ϵy is the strain ratio, which defines the mands [20,21]. A summary account of this procedure is provided in
maximum permissible strain and is given by Eq. (4): Foster and Gardner [22].

164
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

M/Mpl

1.0

R pl

pl u ur

Fig. 5. Definition of rotation capacity.

Fig. 4. Non-dimensional support moment versus support rotation curves for the tested
continuous beams.

Table 3
Ultimate moments, rotations and load capacities of the tested continuous beams.

Test designation Mu,sup Mu,sup Mu,span Mu,span Fu Fu R


Mpl Mcsm Mpl Mcsm Fcol Fcol,csm

305 × 127 × 48, 1.11 0.99 1.16 1.03 1.17 1.05 2.71
λLT = 0.4 , R
305 × 127 × 48, 1.13 1.01 1.18 1.05 1.25 1.12 3.99
λLT = 0.4 , K1
305 × 127 × 48, 1.16 1.04 1.21 1.08 1.21 1.09 3.11
λLT = 0.4 , K2
305 × 165 × 40, 1.07 1.08 1.11 1.13 1.13 1.14 1.50
λLT = 0.4 , R
305 × 165 × 40, 1.09 1.10 1.14 1.15 1.14 1.15 3.38
λLT = 0.4 , K1
305 × 165 × 40, 1.09 1.10 1.14 1.15 1.13 1.15 2.97
λLT = 0.3, K1
Mean 1.17 1.12 Fig. 6. Normalised load versus displacement curves for the tested specimens.

165
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Fig. 7. Overall computational regime for the inelastic buckling analysis of continuous beams.

3. Experimental programme to achieve the desired values of λLT for each test, with L1 chosen in
relation to L2 to ensure that buckling occurs in the L1 region first.
3.1. Introduction To mitigate premature failure through web crippling, web stiffeners
are provided at the positions of the supports and the loading points for
A testing programme comprising tensile and compressive material each test. Vertical displacements are measured using pull-wire trans-
coupon tests, and tests on beams with discrete lateral restraints is ducers and end rotations are measured using inclinometers. Tests are
carried out at the Building Research Establishment and Imperial College performed under load control, applying force at each loading point
London on hot-rolled grade S355 steel I-sections. Full details of the using two hand operated 1000 kN hydraulic jacks; forces are mon-
material properties tests are reported in Foster and Gardner [22]. A itored with load cells. The progression of strain throughout the test is
summary of the key material properties derived from these tests is monitored using two linear electrical resistance post-yield strain gauges
presented in Table 1. In this table, T denotes a tensile test, C denotes a that are bonded to the extreme tensile and compressive fibres of the
compressive test, W denotes a coupon taken from the web and F specimens at their mid-spans. Simple support conditions are provided
denotes a coupon taken from the flange. Other symbols are defined as by a roller system, permitting longitudinal movement at one end. All
follows: E is the modulus of elasticity, fy is the material yield strength, fu instrumentation signals are recorded at one-second intervals using the
is the ultimate tensile strength, ϵf is the strain at fracture calculated over data acquisition system DATASCAN.
the standard gauge length set out in EN 10002-1 (2001), and ϵu is the Fig. 4 a and 4b presents the non-dimensional moment-rotation
strain at the ultimate tensile stress. For the continuous beam tests, two responses for all of the continuous beam tests, in which θpl is the elastic
cross-section sizes are used: 305×127 ×48 UB, which has a Class 1 component of the rotation when Mpl is reached. Table 3 presents a
flange and web, and 305×148 ×40 UB, which has a Class 2 flange and summary of the ultimate test moments measured at the loading points
a Class 1 web. The basic geometric properties of the tested specimens (Mu,span) at the ultimate load of the system normalised by Mpl and Mcsm,
(web height hw, flange thickness tf, web thickness tw), global imperfec- and the ultimate system load Fu normalised by the plastic collapse load
tion magnitudes (e0), and local flange imperfection magnitudes (ω0) are Fcol and CSM collapse load Fcol,csm. In this table, F denotes the value of
reported in Table 2. Detailed descriptions of the imperfections mea- each of the two point loads. The test results show that the CSM bending
surement procedures are provided in Foster and Gardner [22]. resistance is achievable at both the loading points and at the central
support for all specimens with the exception of the
305 × 127 × 48, λLT = 0.4 R test (see Fig. 4a).
3.2. Tests on continuous beams with discrete lateral restraints The ultimate rotation capacity R, defined as:

We test six continuous beams with two cross-sections θur


R= −1
(305 × 127 × 48 UB and 305 × 165 × 40 UB), equipped with lateral θpl (6)
restraints positioned at intervals corresponding to λLT = 0.4 and
λLT = 0.3. The schematic test configuration for the continuous beam in which θur is defined as the mid-span rotation at which the bending
tests is shown in Fig. 3a. For all tests, lateral restraints are provided at moment is descending and falls below Mpl (see also Fig. 5). Test values
the loading points and at the supports (Fig. 3b). We vary the length L2 are reported in Table 3. It is not possible to determine θur in all tests. In

166
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Fig. 8. Programme structures for in-plane analyses.

such cases, the terminal test rotation is used instead, which is taken as loads as predicted by the CSM for indeterminate structures are
the last recorded value. EN-1993-1-1 requires a minimum rotation achieved. The predictive capacity of the CSM collapse load is presented
capacity of R = 3 for a cross section to be considered Class 1. Where in Table 3, which compares it with the conventional collapse loads
achieved, the results show that this requirement is sufficient to attain derived from a plastic analysis Fcol. An improvement can be seen when
F
Mcsm. We speculate that with closer lateral restraints, higher values of R compared with ultimate test loads, with mean values of F u = 1.12
col,csm
could be attainable. Fig. 6 a and 6b, which presents the non-dimen- Fu
and Fcol
= 1.17 respectively.
sional loading histories in terms of F/Fcol,csm, shows that the collapse

167
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Fig. 9. Schematic moment-curvature behaviour with strain-hardening.

Fig. 12. Programme structure for the elastic buckling analysis.

25

20

15
Error (%)

Fig. 10. Cross-section discretisation for the layered model.


10

0
0 4 8 12 16
Number of elements
Fig. 13. Comparison of finite element results with analytical result for the elastic
buckling moment of a beam subject to uniform moment.

4. Numerical modelling

4.1. Introduction and model overview

For the purposes of this research, the finite element method is


employed, largely for its flexibility and widespread use in the literature.
For its application to the problems considered in this paper, two general
steps are taken. Firstly, a static in-plane pre-buckling analysis is
performed to determine the distribution of bending and axial forces
along the length of the structure for use in the incremental (geometric)
component of a buckling analysis. Secondly, knowing these forces (and
stresses), the points of transition from elastic to yielded to strain-
Fig. 11. Predictor-corrector algorithm for inelastic behaviour.
hardened material properties are then established and introduced into

168
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Fig. 14. Validation of the numerical model against analytical results.

Table 4 (iv) If stresses exceed the yield stress, initiate a materially non-linear
Comparison of the laboratory test results on continuous beams with the numerical model. in-plane analysis with a trial value of the inelastic load factor. This
will be some value that results in the maximum moment being
Test designation Test Mu,span/ FE Mu at test θu
slightly in excess of the yield moment;
305 × 127 × 48, λLT = 0.4 , Rigid 0.95 (v) Extract the distribution of yielded and strain-hardened material
305 × 165 × 40, λLT = 0.4 , Rigid 0.96 from the materially non-linear analysis. Using this information,
perform a linear eigenvalue analysis to obtain an inelastic critical
load and corresponding inelastic critical moment;
the flexural stiffness component of a buckling analysis [23]. For (vi) If the trial and inelastic critical values for the load factor are
statically indeterminate inelastic structures, these steps cannot be sufficiently close, accept the solution, otherwise introduce another
separated as stresses and yield distributions are interrelated by a system trial load.
of non-linear equilibrium equations [24]. In practice, the procedure for
arriving at an inelastic load prediction using the finite element method The relationships between these steps are summarised in Fig. 7.
can be subdivided into the following stages (adapted from Nether-
cot [25] and Yoshida et al. [3]) :
4.2. Basic modelling procedures
(i) Perform an in-plane (pre-buckling) analysis to establish the
4.2.1. In plane analysis
distribution of internal forces due to a nominal distribution of
The initial in-plane analysis in this paper uses two subroutines. The
concentrated and distributed applied loads;
first is a linear elastic mixed-formulation Timoshenko beam model and
(ii) Using the information obtained from the pre-buckling analysis,
the second is a materially non-linear mixed formulation layered
perform an elastic buckling analysis to determine a critical load
Timoshenko beam model. Using the pure-displacement finite element
factor λcr corresponding to the applied loads;
formulation, where displacements are the primary variables:
(iii) Evaluate the subsequent distribution of internal forces with a
second in-plane analysis. If any internal stresses exceed the 1
material yield stress σy, proceed to an inelastic analysis, otherwise
Π(u) =
2
∫V ϵTCϵdV − ∫V uTfdV where ϵ=∂ϵ u
(7)
accept the critical load from the elastic buckling analysis;
in which u is a vector of element displacements, ϵ is a vector of element

169
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

1.4 In order to determine the spread of plasticity along the beam and
through its cross-section, material non-linearity needs to be introduced.
1.2 For a beam, the general expression for element stiffness ke is given by
Eq. (9):
1 Le
ke = ∫0 BT CBdx
(9)
0.8
M/Mpl

in which C is the constitutive matrix and B is the strain-displacement


0.6 matrix. With reference to Fig. 9, in the elastic range C = EI (where EI is
the flexural stiffness) and during plastic deformations, C = EIt (where
0.4 EIt is the tangent flexural stiffness).
For an increment dM in bending moment in the post-yield range, the
Test
0.2 resulting change in curvature dχ is represented as:
Numerical model
dχ = dχe + dχp (10)
0
0 0.01 0.02 0.03 0.04 in which χe and χp are the respective elastic and plastic curvature
Rotation (rad) components. In the strain-hardening range, plastic deformations are
dependent upon the current degree of plastic deformation, the extent to
which is characterised by the strain-hardening parameter H:
dM dM EIt
H= = =
dχp dχ −dχe 1−EIt / EI (11)
1.4
Using Eqs. (10) and (11), the incremental moment curvature
1.2 relationship is given by:

1 EIH ⎛ EI ⎞
dM = dχ = EI ⎜1− ⎟ dχ
EI +H ⎝ EI +H ⎠ (12)
0.8
M/Mpl

such that in the post-yield range C = EI ( EI


. The shear force-
1− EI + H )
shear strain relationship is assumed to remain elastic [24].
0.6
For continuous beams, the degree of moment redistribution is
closely linked to the degree of plastic penetration at different locations
0.4
along the member. In order to account for this, a layered cross-section is
Test used, where the depth of the cross-section is subdivided into k layers
0.2
Numerical model (see Fig. 10).
In the layered formulation, individual layer stresses are used, such
0
that bending moments M and shear forces Q are determined as:
0 0.01 0.02 0.03 0.04
Rotation (rad) M= ∑ bk σx,k zk tk ; Q= ∑ bk τx,k tk
k k (13)

in which (with reference to Fig. 10) b is the layer breadth, t is the layer
thickness, z distance of the mid-surface of layer k from the neutral axis,
Fig. 15. Experimental and numerical moment versus end rotation curves for continuous σ is the uniaxial stress, τ is the shear stress, and x denotes the
beams. longitudinal direction. Along similar lines, flexural rigidity EI is
evaluated as:
strains derived from element displacements, C is the constitutive matrix
and f is the applied loading vector. The two-node Timoshenko beam EI = ∑ Ek bk zk2 tk
k (14)
element, when used with a pure-displacement formulation, will result
in shear locking behaviour, as described by [26]. By invoking a mixed If the stress at layer k exceeds the uniaxial yield stress σy, then the
formulation, shear locking can be avoided if it is assumed that both Young's modulus Ek at that layer is replaced with Ek(1 − Ek/(Ek + H)).
displacements and strains are unknown variables in the total potential The overall computational scheme for the inelastic in-plane analysis
and it is assumed that there is a constant element shear strain γ [26]. is summarised in Fig. 8b. The specific steps that need to be followed in
Adopting a linear displacement representation [27], the resulting the analysis are to update the stresses, update the plastic strains, update
element flexural stiffness matrix k is given by Eq. (8): the yield stress limit, identify the loading and unloading paths, and
finally to satisfy equilibrium for virtual displacements. With reference
⎡ Gh Gh

Gh Gh ⎤
⎢ Le 2 Le 2 ⎥ to Fig. 11, at load step r the stress σir, k in layer k of element i with a strain
⎢ ⎥ increment Δϵri, k is evaluated as σir, k = σi(,rk−1) + Ei, k Δϵri, k . Until the yield
⎢ Gh
k = ⎢ Gh
2 ( GhLe
4
Eh3
+ 12L
e ) − (− Gh
2
GhL e
4
Eh3
− 12L
e ) ⎥
⎥ stress σy is reached, this approach requires no further modification.
⎢− −
Gh Gh

Gh ⎥ When making the transition from point A (in the elastic range) to point
⎢ Le 2 Le 2 ⎥ B (in the strain-hardening range), a predictor-correction approach is
⎢ Gh ⎥
⎢ 2
⎣ ( GhL e
4
Eh3
− 12L
e )− Gh
2 ( GhL e
4
Eh3
+ 12L
e ) ⎥
⎦ (8)
necessary [28]. This has three steps. Firstly, compute the elastic stress
predictor σei,k using Eq. (15) to locate A’ in Fig. 11:
Using the procedure outlined in Fig. 8a, this matrix can be σei, k = σi(,rk−1) + Ei, k Δϵri, k (15)
assembled into a global stiffness matrix and solved for the global
displacements U for any given set of boundary conditions, global Secondly, formulate the correction factor R for the load increment
geometry, material and cross-section properties. using Eq. (16):

170
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Lspan Lspan
P P
Lspan/2 Lspan/2

Lspan Lspan
P P
Lspan/3 Lspan/3

Lspan Lspan
P P
2Lspan/3 2Lspan/3

Fig. 16. Loading configurations for the parametric studies.

σei, k −σy 4.2.2. Buckling analysis


R=
σi(,rk−1)−σei, k (16) The finite element model used in the linear buckling analysis
subroutine is based upon the stiffness and stability matrices derived
which can be derived from similar triangles in Fig. 11. R determines the by Barsoum and Ghallager [29]. These matrices were developed for an
proportion of the strain increment that is plastic. Finally, calculate the arbitrary open section subjected to any combination of distributed,
corrected stress at point B using Eq. (17): concentrated, axial, moment and torsional loading.
σir, k = σi(,rk−1) + [(1−R ) Ei, k +REti, k ]Δϵri, k In a similar manner to regular displacement-based finite element
(17)
formulations, expressions for strain energy and the total potentials for
If the material has previously yielded (i.e. the transition from point the applied loads are formulated. Within the strain energy expression,
C to point D), the same steps are followed, but with R = 1 in Eq. (17). In buckling deformations can be separated from pre-buckling deforma-
the inelastic range, permanent plastic deformations are experienced tions, with the latter being excluded from the buckling analysis and
and so the plastic strains must be computed for each load increment. instead computed during the in-plane analysis. Cubic basis functions
Re-writing Eq. (17): are then substituted for the displacement terms in the out-of-plane
strain energy expression. The resulting expressions are arranged into
σir, k = σy + Eti, k RΔϵri, k ≡ σy + Δσir, k (18) two matrices: the first containing terms for axial, flexural and torsional
stiffness - the element flexural stiffness matrix kfe. The second contain-
in which and Δσir, k RΔϵri, k
are the stress and strain components involved
ing terms that are dependent upon geometric terms and the element
in plastic flow respectively. The plastic strain increment is:
forces determined in the pre-buckling analysis - the element incre-
σir, k mental or geometric stiffness matrix kge. Summing the potential energy
Δϵrpi, k = RΔϵri, k −
Ei, k (19) of the applied loads f and the strain energy terms to form total potential
energy Π, then for a small variation δΠ, stable equilibrium is where:
During plastic straining, the yield limit of the material changes as a
δ Π = 0: f=k ef u+kgeu (21)
result of plastic flow. The updated yield stress σyir , k is calculated using
Eq. (20): For elastic instability problems, the intensity of the element forces is
σyir , k = σy + H Δϵrpi, k unknown. Hence, the geometric stiffness matrix needs to be numeri-
(20)
cally evaluated for an arbitrary load intensity λ for any distribution of
In the linear elastic range, the principle of virtual displacements is element forces [30]. The stability problem reduces to:
satisfied. However, in the inelastic range, stress adjustments produce δ Π = 0: k ef u+λkgeu−f = 0 (22)
out of balance forces that may not satisfy equilibrium. At the element
interface level, forces between neighbouring yielded and elastic ele- Considering the second variation of Π:
ments are initially out of equilibrium. Overall equilibrium must be
δ 2 Π > 0 ⇒ stable equilibrium
satisfied by redistributing the force imbalance to all other elements in
δ 2 Π=0 ⇒ bifurcation of equilibrium
the system. This is achieved using a corrective displacement increment
within the loading step. Equilibrium is then satisfied approximately Hence:
once a given tolerance has been achieved within the desired number of
iterations. [k ef +λkge] u =0 (23)

171
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

1.5 1.5
MLP2 = MLP1
MLP2 = 2MLP1 MLP2 = MLP1
1.25 1.25
MLP2 = 2MLP1
1 1
M/Mcsm, LP2

M/Mcsm, LP2
Load at 1/2 point Load at 1/2 point
0.75 Load at 1/3 point 0.75 Load at 1/3 point
Load at 2/3 point Load at 2/3 point

0.5 MLP2 = 0.5MLP1 0.5

0.25 0.25 MLP2 = 0.5MLP1

0 0
0 0.25 0.5 0.75 1 1.25 1.5 0 0.25 0.5 0.75 1 1.25 1.5
M/Mcsm, LP1 M/Mcsm, LP1

1.5
MLP2 LP1

1.25
MLP2 = 2MLP1
1
M/Mcsm, LP2

Load at 1/2 point


0.75 Load at 1/3 point
Load at 2/3 point

0.5

0.25 MLP2 = 0.5MLP1

0
0 0.25 0.5 0.75 1 1.25 1.5
M/Mcsm, LP1

Fig. 17. Interaction buckling relationships for a two span 305 × 127 × 48 UB continuous beam.

The non-trivial solution requires: λpl,cr and corresponding moment Mpl,cr are evaluated. The trial load
λtrial is then adjusted until the difference between Mtrial,pl and Mpl,cr
⎛e ⎞
det ⎜k+λkge⎟ = 0 satisfies the specified tolerance (1%) for a given number of iterations.
⎝f ⎠ (24)
This procedure is carried out using the bisection method. Defining
which is a linear eigenvalue problem with the critical load being the tolerance as |Mtrial,pl − Mcr,pl|, the method proceeds by evaluating
a b c a b
lowest eigenvalue λ = λcr. The specific expressions for the terms of kfe α tol (λ trial ), α tol (λ trial ), and α tol (λ trial ), where λ trial and λ trial are chosen such
and kge are reported in Barsoum and Ghallager [29]. Element stiffness that the corresponding αtol have opposite signs and
c a b a b
matrices are assembled so that the potential energy for the system is the λ trial = (λ trial + λ trial )/2 . If λ trial and λ trial result in different signs in the
sum of the potential energies of the component elements. The eigen- first iteration, λ trial c
is selected instead of λ trial b
as the new trial load factor
value analysis is then performed on the system, rather than the in the next iteration. This process is allowed to continue for a specified
elemental equations. A description of the implementation of the number of iterations until the error is less than or equal to the specified
buckling analysis is provided in Fig. 12. tolerance.
This model has been systematically validated against known
analytical results by Barsoum and Ghallager [29]. In all cases,
convergence to the analytical solution was achieved with negligible 4.3. Validation and convergence
error using no more than six elements. An example for a beam subject
to uniform moment is provided in Fig 13, in which the error is defined To validate the numerical model, two approaches were taken: firstly
as: a comparison of the numerical model results against two standard
analytical cases and secondly, a comparison to laboratory test data. For
Error(%) = Mcr,numerical − Mcr,analytical (25) the first approach, two analytical expressions for the critical buckling
moment are formulated. The first is the elastic critical moment Mcr with
in which Mcr,numerical is the critical moment as determined from the
all material properties (E and G) assumed to be in the elastic range. The
numerical buckling analysis and Mcr,analytical is the elastic critical
second is the inelastic critical moment Mcr,t which uses effective values
moment of a simply supported beam subjected to a uniform moment.
of material properties (Et and Gt) according to the tangent modulus
The final step in the analysis is to iterate towards the inelastic
theory. All data is plotted against the non-dimensional slenderness
buckling load. Using a trial load factor λtrial, a corresponding maximum
parameter λLT with reference to the elastic section modulus Wel:
trial moment Mtrial,pl is determined from an inelastic analysis. Using the
material properties from this analysis, a critical buckling load factor

172
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

1.5 1.5
MLP2 = MLP1
MLP2 = 2MLP1 MLP2 = MLP1
1.25 1.25
MLP2 = 2MLP1
1 1
M/Mcsm, LP2

M/Mcsm, LP2
Load at 1/2 point Load at 1/2 point
0.75 Load at 1/3 point 0.75 Load at 1/3 point
Load at 2/3 point Load at 2/3 point

0.5 0.5

0.25 MLP2 = 0.5MLP1 0.25 MLP2 = 0.5MLP1

0 0
0 0.25 0.5 0.75 1 1.25 1.5 0 0.25 0.5 0.75 1 1.25 1.5
M/Mcsm, LP1 M/Mcsm, LP1

1.5
MLP2 = MLP1

1.25
MLP2 = 2MLP1
1
M/Mcsm, LP2

Load at 1/2 point


0.75 Load at 1/3 point
Load at 2/3 point
0.5

0.25 MLP2 = 0.5MLP1

0
0 0.25 0.5 0.75 1 1.25 1.5
M/Mcsm, LP1

Fig. 18. Interaction buckling relationships for a two span 305 × 165 × 40 UB continuous beam.

Wel fy numerical data due to the assumption of a bi-linear material model (the
λLT = material model in the numerical analysis employs only the measured
Mcr (26)
yield stress from the tests).
Fig. 14 plots these curves for the two cross-sections used in this The overall buckling model is composed of three distinct numerical
study alongside the data generated by the numerical model. The models, each with separate convergence demands. Through a series of
configuration is a two-span continuous beam with concentrated loads trial studies, overall convergence with significantly less than 1% error
applied at the two mid-spans in equal proportions (zero interaction). between Mtrial,pl and Mcr,pl was achieved using 60 cross-section layers
The results show that the numerical model satisfies two theoretical and 200 elements. For a stable solution in the materially non-linear
expectations: firstly, where λLT = M / Mel = 1 the analytical and numer- model, it was found that 100 load increments were sufficient.
ical expressions coincide; secondly, the numerical model results
approach Mcr,t as full plasticity is able to develop through the cross-
section at low values of λLT . 5. Parametric studies
For the second approach, moment rotation data obtained from the
laboratory tests are compared with data generated by the numerical A series of parametric studies are performed, focusing on variations
model using the measured material and geometric properties of the test in λLT . The aim of these studies is to arrive at a limiting value of λLT for
specimens. These are summarised in Table 4 and show acceptable the CSM that can satisfy a variety of basic structural and loading
agreement between the tests and numerical model. configurations. For these studies, two cross-sections are used
Fig. 15 plots the comparative moment-rotation responses of the tests (305 × 127 × 48 UB and 305 × 165 × 40 UB) and three basic loading
and the numerical model. The initial slope of the numerical model is configurations are chosen: (i) concentrated loads at the mid-spans; (ii)
perfectly straight, whilst the experimental data shows some rounding concentrated loads at the 1/3 spans; and (iii) concentrated loads at the
throughout. This is likely to be caused by elastic deformations of the 2/3 spans. These are illustrated in Fig. 16.
load transfer apparatus and friction between the specimen and the The parametric studies are organised into two groups. The first
lateral restraints. For the numerical model, yielding commences at the focuses on interaction relationships and the second assumes zero
yield moment, which is an intrinsic assumption of the computer model. interactions, instead focusing upon a wide range of λLT values for the
The marginally earlier yielding shown by the test data is likely to be due three structural configurations under consideration. In all cases, a bi-
to residual stresses. For the 305 × 165 × 40 UB specimen, there is linear material model is assumed with E/Esh = 100.
evidence of a small plateau in the test data, which is not apparent in the

173
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

1.6 1.2
Load at 1/2 point
1.4 Load at 1/3 point
1 Load at 2/3 point
1.2
F2/Fcol, csm

M/Mcsm
1
0.8
0.8 Load at 2/3 point
Load at 1/2 point
0.6 Load at 1/3 point 0.6
0.4

0.2 0.4
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2
F1/Fcol, csm LT

1.6 1.2
Load at 1/2 point
1.4 Load at 1/3 point
1 Load at 2/3 point
1.2
F2/Fcol, csm

M/Mcsm

1
0.8
0.8 Load at 1/3 point
Load at 1/2 point
0.6 Load at 2/3 point 0.6
0.4

0.2 0.4
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2
F1/Fcol, csm LT

Fig. 19. Collapse load interaction relationships for continuous beams with λLT = 0.20 Fig. 20. Lateral torsional buckling slenderness relationships for two-span continuous
critical spans. beams.

5.1. Interaction relationships 1.2

For this component of the study, interaction relationships are


constructed by holding the load in one span at a constant value of P
1.0
and allowing the load in the other span to take a different value υP
where 0 ≤ υ ≤ 1. Three variations in lateral torsional buckling
F/Fcol, csm

slenderness λLT are considered for each cross-section and loading


configuration, λLT = 0.25, 0.20, 0.18. 0.8
The parametric study is performed using the computer model
discussed in Section 4. The inelastic critical moments at the loading 305×127×48 (load at 1/3 point)
points non-dimensionalised by the corresponding CSM moment are 305×127×48 (load at 1/2 point)
shown in the interaction diagrams of Figs. 17 and 18. For the low
0.6 305×127×48 (load at 2/3 point)
305×165×40 (load at 1/3 point)
slenderness values used, in most cases sufficient strain-hardening takes
305×165×40 (load at 1/2 point)
place to attain critical buckling moments at or above the in-plane CSM 305×165×40 (load at 2/3 point)
capacities. For spans whose loads are placed at the 2/3 loading point, 0.4
the full CSM moment capacity is generally not attained at the loading 0.0 0.2 0.4 0.6 0.8 1.0 1.2
points, but sufficient yielding still occurs for a full CSM mechanism to
LT
develop.
At high and low ratios of MLP1/MLP2, extensive strain-hardening can Fig. 21. CSM collapse loads versus lateral torsional buckling slenderness for all sections
be observed, and this develops due to extensive pre-bucking redistribu- and load configurations.
tion of bending moments [3]. This redistribution is favourable with

174
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

Table 5 5.2. Restraint spacing


Critical CSM restraint spacings compared with EN-1993-1-1 restraint spacings.
The variations in non-dimensional critical moment with the lateral
Lcsm/Lstable
torsional buckling slenderness parameter λLT are plotted for each
305 × 127 × 48 UB 305 × 165 × 40 UB section type and the corresponding loading configurations in Fig. 20
Load at the 1/3 span 0.81 0.76 a and 20b. In the transition from elastic to plastic behaviour, there are
Load at the 1/2 span 0.80 0.76 some fairly pronounced differences in the critical moments for the
Load at the 2/3 span 1.09 1.08
section types and their various configurations. However, at low
slenderness values, these disparities are greatly reduced as extensive
1.3 strain-hardening is able to occur before the onset of inelastic buckling.
In all cases, CSM levels of bending resistance are attained for both
Mcsm/Mpl = 1.15 section types and all configurations at a slenderness λLT = 0.18, which is
1.2 in line with the analytical and numerical results presented by Foster
M/Mpl or χLT or χLT,csm

Mcsm/Mpl = 1.10 = χLT,csm and Gardner [2] for statically determinate beams.
Mcsm/Mpl = 1.05 Fig. 21 plots the lateral torsional buckling slenderness against the
1.1 critical load, normalised by the CSM collapse load. This clearly shows a
more favourable spacing, with all cross-sections attaining Fcol,csm at
λLT = 0.2 . For loading in close proximity to the support at the 2/3 span,
1 EN 1993-1-1 buckling curve a Fcol,csm is achieved at λLT = 0.3. Table 5 compares the critical lengths at
which the CSM collapse load is attainable Lcsm with the EN 1993-1-1
stable length Lstable (Eq. (2)) and shows that in cases where loading is at
0.9 1/3 of the span, Lstable is approximately 25% too long for the CSM.
EN 1993-1-1 buckling curve b Where loading is concentrated about the central support, there is
approximate agreement.
0.8
0 0.2 0.4 0.6 0.8 1
λLT 6. Design recommendations
Fig. 22. Illustration of the application of the factor χLT,csm for a typical range of values of
Mcsm [2].
Adopting the conservative approach of zero interactions between
loaded segments discussed in Section 2.2, the limiting slenderness and
respect to lateral torsional buckling as it tends to occur at the interior capacity reduction relationships proposed in Foster and Gardner [2] for
simply supported beams may be applied. In their research, a limiting
support where reductions in rigidities have minimal effect upon
member stability. This can be observed with greater clarity in Fig. 19 lateral torsional buckling slenderness of λLT = 0.2 is adopted, with the
a and 19b which plots the CSM collapse load interaction relationships following transition curve from the CSM to plastic moment capacities:
for the three load configurations. For loads positioned at the 1/3 and 1/
(λLT −0.2) ⎛ Mcsm ⎞ M
2 spans there is little deviation from the zero interaction load, whilst for χLT,csm = ⎜1− ⎟ + csm for 0.2 ≤ λLT ≤ 0.4
loading at the 2/3 points, there is considerable deviation. This 0.2 ⎝ Mpl ⎠ Mpl (27)
difference is due to the fact that yielding is concentrated in the support
in the latter case, with the majority of the beam remaining elastic and in which χLT,csm is a factor applied to Mpl to obtain Mcsm in the region
able to provide restraint to the more heavily loaded span. Where 0.2 ≤ λLT ≤ 0.4 . For λLT ≤ 0.2 , the full value of Mcsm may be used.
csm
loading is applied > 0.5Lspan away from the support, two separate Furthermore, if Mcsm < Mpl, then λLT = 0.2 remains, but the transition
regions of yielding develop, reducing the available material for restraint should be from Mel rather than Mpl at λLT = 0.4 . These design expres-
and the capacity for interaction. sions are illustrated in Fig. 22 for a typical range of values of Mcsm [2].
The comparative performance between the design format and the
results of the parametric studies are plotted in Fig. 23.

1.2 1.2

Load at 1/2 point


1 1
CSM transition
M/Mcsm

M/Mcsm

0.8 0.8

Load at 1/2 point


0.6 0.6
CSM transition

0.4 0.4
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
λ LT λ LT

Fig. 23. Mpl to Mcsm transition relationships.

175
A.S.J. Foster, L. Gardner Journal of Constructional Steel Research 136 (2017) 162–176

7. Conclusions [4] M. Salvadori, Lateral bending of beams of rectangular cross-section under bending
and shear, Proc First US Natl. Congress Appl. Mech. 1 (1951) 403–405.
[5] N. Trahair, Interaction buckling of narrow rectangular continuous beams, Civ. Eng.
We investigate the minimum values of lateral torsional buckling Trans. Inst. Eng. Aust. CE10 (1968) 167–172.
slenderness λLT required to achieve the cross-section capacity predicted [6] N. Trahair, Elastic stability of propped cantilevers, Civ. Eng. Trans. Inst. Eng. Aust.
CE10 (1968) 94–100.
by the CSM for statically indeterminate members through a programme [7] N.S. Trahair, Elastic stability of continuous beams, J. Struct. Div. ASCE 95 (1969)
of six tests on continuous beams. Two S355 cross-sections (one Class 1 1295–1312.
and one Class 2) are considered with lateral restraint spacings [8] D. Nethercot, N. Trahair, Lateral buckling approximations for elastic beams,
Struct. Eng. 54 (1976) 197–204.
corresponding to λLT = 0.3 and 0.4. Results show that, for the cases [9] P.F. Dux, S. Kitipornchai, Elastic buckling of laterally continuous I-beams, J.
considered, the CSM-compatible collapse load is attainable, with the Struct. Div. ASCE 108 (1982) 2099–2116.
CSM on average providing enhanced predictions of capacity over [10] N.S. Trahair, Beams and beam columns, inelastic lateral buckling of beams, Applied
Science Publishers, 1983, pp. 35–69.
conventional plastic methods, but closer restraint spacing than is
[11] H. Yoshida, Y. Imoto, Inelastic lateral buckling of restrained beams, J. Struct. Div.
currently required for traditional plastic design would be beneficial ASCE EM2 (1973) 9666–9679.
for member rotation capacity. [12] T. Poowannachaikul, N.S. Trahair, Inelastic buckling of continuous steel I-beams,
A numerical model is validated against the experimental results and Civ. Eng. Trans. Inst. Eng. Aust. CE18 (2) (1976) 134–139.
[13] P.F. Dux, S. Kitipornchai, Buckling approximations for inelastic beams, J. Struct.
is shown to be able to capture the observed physical behaviour. On the Div. ASCE 110 (1984) 559–574.
basis of this, parametric studies are conducted to investigate the [14] J.M. Davies, Strain hardening, local buckling and lateral-torsional buckling in
implications of varying the restraint spacing and the interactions plastic hinges, J. Constr. Steel Res. 62 (2006) 27–34.
[15] O. Pettersson, Combined bending and torsion of I-beams of monosymmetrical cross-
between adjacent unequally loaded spans on the ultimate capacities section, 10 Division of Building Statics and Structural Engineers (Royal Institute of
of the beams. The results from the parametric studies show that, in line Technology Stockholm), 1952, p. 1.
with recent research on statically determinate structures [2], distances [16] N.S. Trahair, Stability of I-beams with elastic end restraints, J. Inst. Eng. Aus. 37
(1965) 157.
between lateral restraints need to be reduced to attain the full CSM [17] N.S. Trahair, Elastic stability of I-beams elements in rigid-jointed frames, J. Inst.
capacity compared with the requirements of EN 1993-1-1 to achieve Eng. Aus. 38 (1966) 171.
Mpl. Further research on a wider range of cross-section properties and [18] N.S. Trahair, Elastic stability of frame structures, Ph.D. thesis, University of Sydney,
1967.
steel grades is recommended.
[19] N.S. Trahair, M.A. Bradford, D.A. Nethercot, L. Gardner, The Behaviour and Design
of Steel Structures to EC3, 3rd edn, Taylor and Francis, Abingdon, 2008.
Acknowledgements [20] L. Gardner, F. Wang, A. Liew, The influence of strain hardening on the behaviour
and design of steel structures, Int. J. Struct. Stab. Dyn. 11 (05) (2011) 855–875.
[21] F. Wang, A deformation based approach to structural steel design, Ph.D. thesis,
This project is funded by the Building Research Establishment Trust, Imperial College of Science, Technology and Medicine, 2011.
whose sponsorship is gratefully acknowledged. The authors would like [22] A.S.J. Foster, L. Gardner, Ultimate behaviour of continuous steel beams with
to thank Dr. Julie Bregulla and Mr. David Brooke from the Building discrete lateral restraints, Thin-Walled Struct. 88 (2015) 58–69.
[23] N.S. Trahair, Flexural Torsional Buckling of Structures, Spon Press, 1993.
Research Establishment, and Mr. Gordon Herbert from Imperial College [24] D.R.J. Owen, E. Hinton, Finite Elements in Plasticity, Pineridge Press, 1980.
London for their assistance in the experimental programme. Dr. [25] D.A. Nethercot, The solution of inelastic lateral stability problems by the finite
Jonathan Gosaye Fida Kaba is duly acknowledged. element method, Proceedings of the 4th Australian Conference on Mechanics of
Structures and Materials, 1 1973, pp. 183–190.
[26] K.J. Bathe, Finite Element Procedures, Prentice Hall, 1996.
References [27] T.J.R. Hughes, R.L. Taylor, S. Kanoknukulchai, A simple and efficient finite
element for bending, Int. J. Numer. Methods Eng. 11 (1977) 1529–1543.
[28] J.N. Reddy, An Introduction to Nonlinear Finite Element Analysis, Oxford
[1] L. Gardner, The continuous strength method, Proc. Inst. Civ. Eng. Struct. Build. 161
University Press, 2004.
(3) (2008) 127–133.
[29] R.S. Barsoum, R.H. Ghallager, Finite element analysis of torsional and torsional-
[2] A.S.J. Foster, L. Gardner, Stability of steel beams using the continuous strength
flexural stability problems, Int. J. Numer. Methods Eng. 2 (1970) 335–352.
method, Thin-Walled Struct. 100 (2016) 1–13.
[30] R.H. Gallagher, Finite Element Analysis: Fundamentals, 1st edn, Prentice-Hall,
[3] H. Yoshida, D.A. Nethercot, N.S. Trahair, Analysis of lateral buckling of continuous
Englewood Cliffs, NJ, 1975.
beams, IABSE Proceedings P-3/77, 1977, pp. 1–14.

176

You might also like