Sunteți pe pagina 1din 42

Lecture 7.9.

1: Unrestrained Beams I
OBJECTIVE/SCOPE
To develop an understanding of the phenomenon of lateral-torsional instability; to identify
the controlling parameters and to show how theory, experiment and judgement are combined
to produce a practical design method. The design procedure given in Eurocode 3 [1] is used
as an illustration of such a method.
PREREQUISITES
Lecture 6.1: Concepts of Stable and Unstable Elastic Equilibrium
Lectures 6.6: Buckling of Real Structural Elements
Lectures 7.8: Restrained Beams
RELATED LECTURES:
Lecture 7.9.2: Unrestrained Beams II
Lectures 7.10: Beam Columns
SUMMARY
This lecture begins with a non-mathematical introduction to the phenomenon of lateral
torsional buckling. It presents a simple analogy between the behaviour of the compression
flange and the flexural buckling of a strut. It summarises the principal factors influencing
lateral stability and briefly describes the role of bracing in improving this.
A brief explanation is given of the reasons why the elastic theory, discussed in Lecture 7.9.2,
requires modification before being used as a basis for the design rules for unrestrained beams.
A summary of the background to Eurocode 3 [1] is also presented.
NOTATION
C coefficient to account for type of loading
d overall depth
EI
z
flexural rigidity about the minor axis
f
d
design strength of material
f
y
material yield strength
i
z
minor axis radius of gyration
k coefficient to account for conditions of lateral support
L span
M
b.Rd
buckling resistance moment
M
cr
elastic critical buckling moment
M
pl
plastic moment of cross-section
M
Rd
moment resistance of cross-section
t
f
flange thickness
u lateral deflection
o
LT
parameter in design formula, see Equation (2)
_
LT
reduction factor for lateral-torsional buckling
LT
beam slenderness

LT
basic slenderness

1
parameter used to determine
LT
, see Equation (4)
| twist
|
LT
parameter used to determine _
LT
, see Equation (2)
moment ratio, see Equation (5)
1. STRUCTURAL PROPERTIES OF SECTIONS USED
AS BEAMS
When designing a steel beam it is usual to think first of the need to provide adequate strength
and stiffness against vertical bending. This leads naturally to a cross-sectional shape in which
the stiffness in the vertical plane is much greater than that in the horizontal plane. Sections
normally used as beams have the majority of their material concentrated in the flanges, which
are relatively narrow so as to prevent local buckling. The need to connect beams to adjacent
members with ease normally suggests the use of an open section, for which the torsional
stiffness will be comparatively low. Figure 1, which compares section properties for four
different shapes of equal area, shows that the high vertical bending stiffness of typical beam
sections is obtained at the expense of both horizontal bending and torsional stiffness.

2. RESPONSE OF SLENDER BEAMS TO VERTICAL
LOADING
It is known from our understanding of the behaviour of struts that, whenever a slender
structural element is loaded in its stiff plane (axially in the case of the strut), there exists a
tendency for it to fail by buckling in a more flexible plane (by deflecting sideways in the case
of the strut). Figure 2 illustrates the response of a slender cantilever beam to a vertical end
load; this phenomenon is termed lateral-torsional buckling. Although it involves both a lateral
deflection (u) and twisting about a vertical axis through the web (|), as shown in Figure 3,
this type of instability is quite similar to the simpler flexural buckling of an axially loaded
strut. Loading the beam in its stiffer plane (the plane of the web) has induced a failure by
buckling in a less stiff-direction (by deflecting sideways and twisting).


Of course, many types of construction effectively prevent this form of buckling, thereby
enabling the beam to be designed with greater efficiency as fully restrained (see Lecture
7.8.1). In this context it is important to realise that during erection of the structure certain
beams may well receive far less lateral support than will be the case when floors, decks,
bracings, etc., are present, so that stability checks, at this stage, are also necessary.
Lateral-torsional instability influences the design of laterally unrestrained beams in much the
same way that flexural buckling influences the design of columns. Thus the bending strength
will now be a function of the beam's slenderness, as indicated in Figure 4, requiring the use in
design of an iterative procedure similar to the use of column curves in strut design. However,
because of the type of structural actions involved, the analysis of lateral-torsional buckling is
considerably more complex. This is reflected in a design approach which requires a rather
greater degree of calculation.



3. SIMPLE PHYSICAL MODEL
Before considering the analysis of the problem, it is useful to attempt to gain an insight into
the physical behaviour by considering a simplified model. Since bending of an I-section beam
is resisted principally by the tensile and compressive forces developed in two flanges, as
shown in Figure 5, the compression flange may be regarded as a strut. Compression members
exhibit a tendency to buckle and in this case the weaker direction would be for the flange to
buckle downwards. However, this is prevented by the presence of the web. Therefore the
flange is forced to buckle sideways, which will induce some degree of twisting in the section
as the web too is required to deform. Whilst this approach neglects the real influence of
torsion and the role of the tension flange, it does, nevertheless, approximate the behaviour of
very deep girders with very thin webs or of trusses or open web joists. Indeed, early attempts
at analysing lateral-torsional buckling started with this approach.




4. FACTORS INFLUENCING LATERAL STABILITY
The compression flange/strut analogy, discussed in the previous section, is also helpful in
understanding the following:
1. The buckling load of the beam is likely to be dependent on its unbraced span, i.e.the
distance between points at which lateral deflection is prevented, and on its lateral
bending stiffness (EL
z
) because strut resistance EL
z
/L
2
.
2. The shape of the cross-section may be expected to have some influence, with the web
and the tension flange being more important for relatively shallow sections, than for
deep slender sections. In the former case the proximity of the stable tension flange to
the unstable compression flange increases stability and also produces a greater
twisting of the cross-section. Thus torsional behaviour becomes more important.
3. For beams under non-uniform moment, the force in the compression flange will no
longer be constant, as shown in Figure 6. Therefore such members might reasonably
be expected to be more stable than similar members under a more uniform pattern of
moment.
4. End restraint which inhibits development of the buckled shape, shown in Figure 3, is
likely to increase the stability of the beam. Consideration of the buckling
deformations (u and |) should make it clear that this refers to rotational restraint in
plan, i.e.about the z-axis (refer back to Figure 5 and 3). Rotational restraint in the
vertical plane affects the pattern of moments in the beam (and may thus also lead to
increased stability) but does not directly alter the buckled shape, as shown in Figure 7.
















A more rigorous analysis, substantiating the above four points, is presented in Lecture 7.9.2.
This lecture also deals with warping of the cross-section and the influence of level of
application of load on stability; factors not illustrated by the simplified model presented here.
5. BRACING AS A MEANS OF IMPROVING
PERFORMANCE
Bracing may be used to improve the strength of a beam that is liable to lateral-torsional
instability. Two requirements are necessary:
1. The bracing must be sufficiently stiff to hold the braced point effectively against
lateral movement (this can normally be achieved without difficulty).
2. The bracing must be sufficiently strong to withstand the forces transmitted to it by the
main member (these forces are normally a percentage of the force in the compression
flange of the braced member).
Providing these two conditions are satisfied, then the full in-plane strength of a beam may be
developed through braces at sufficiently close spacing. Figure 8, which illustrates buckled
shapes for beams with intermediate braces, shows how this buckling involves the whole
beam. In theory, bracing should prevent either lateral or torsional displacement from
occurring. In practice, consideration of the buckled shape of the beam cross-section shown in
Figure 3 suggests that bracing is potentially most effective when used to resist the largest
components of deformation, i.e. a lateral brace attached to the top flange is likely to be more
effective than a similar brace attached to the bottom flange.

6. DESIGN APPLICATION
Direct use of the theory of lateral-torsional instability for design is inappropriate because:
- The formulae are too complex for routine use, e.g. Equation (17) of Lecture 7.9.2.
- Significant differences exist between the assumptions which form the basis of the
theory and the characteristics of real beams. Since the theory assumes elastic
behaviour, it provides an upper bound on the true strength (this point is discussed in
general terms in Lecture 6.6.2).
Figure 9 compares a typical set of lateral-torsional buckling test data obtained using actual
hot-rolled sections with the theoretical elastic critical moments given by Equation (17) of
Lecture 7.9.2.

In Figure 9a only one set of data for a narrow flanged beam section is shown. The use of the
LT
non-dimensional format in Figure 9b has the advantage of permitting results from
different test series (using different cross-sections and different material strengths) to be
compared directly. In both figures three distinct regions of behaviour can be observed:
- Stocky beams which are able to attain M
pl
, with values of
LT
below about 0,4 in
Figure 9b.
- Slender beams which fail at moments close to M
cr
, with values of
LT
above 1,2 in
Figure 9b.
- Beams of intermediate slenderness which fail to reach either M
pl
or M
cr
, with 0,4 <
LT
< 1,2 in Figure 9b.
Only in the case of beams in region 1 does lateral stability not influence design; such beams
can be designed using the methods of Lecture 7.8.1. For beams in region 2, which covers
much of the practical range of beams without lateral restraint, design must be based on
considerations of inelastic buckling suitably modified to allow for geometrical imperfections,
residual stresses, etc., (see Lecture 6.1). Thus both theory and tests must play a part, with the
inherent complexity of the problem being such that the final design rules are likely to involve
some degree of empiricism.
Section 7 outlines the provisions of Eurocode 3 [1] with regard to beam design, assuming
typical sections as shown in Figure 10a and 10b. It should be noted that sections of the type
illustrated in Figure 10b, with one axis of symmetry, e.g.channels, may only be included if
the section is bent about the axis of symmetry, i.e. loads are applied through the shear centre
parallel to the web of the channel. Singly-symmetrical sections bent in the other plane, e.g. an
unequal flanged I-section bent about its major-axis as shown in Figure 10c, may only be
treated by an extended version of the theory of Lecture 7.9.2, principally because the section's
shear centre no longer lies on the neutral axis.

7. METHOD OF EUROCODE 3
The buckling resistance moment [1] is given by:
M
bRd
= _
LT
M
Rd
(1)
where M
Rd
is the moment resistance of the cross-section
_
LT
is the reduction factor for lateral-torsional buckling
In determining M
Rd
the section classification should, of course, be noted and the appropriate
section modulus used in conjunction with the material design strength f
d
. The value of _
LT

depends on the beam's slenderness
LT
and is given by:
_
LT
= 1/ {|
LT
+ [|
LT
2
- LT
2
]
1/2
} (2)
where |
LT
= 0,5 [1 + o
LT
(
LT
- 0,20) +
LT
2
]
and o
LT
= 0,21 for rolled sections
o
LT
= 0,49 for welded beams
Figure 11 illustrates the relationship between _
LT
and
LT
, showing how it follows the
pattern of behaviour exhibited by the test data of Figure 9. When
LT
s 0,4, the value of _
LT

is sufficiently close to unity that design may be based on the full resistance moment M
Rd
.

The slenderness
LT
, which is a measure of the extent to which lateral-torsional buckling
reduces a beam's load carrying resistance, is a function of M
Rd
and M
cr
. M
cr
is the elastic
critical buckling moment, a quantity similar in concept to the Euler load for a strut since it is
derived from a theory (see Lecture 7.9.2) that assumes "perfect" behaviour, i.e. an initially
straight member, elastic response, no misalignment of the loading, etc..
Thus
LT
is taken as:
LT
= (3)
For calculation purposes Equation (3) may be rewritten as:
LT
= [
LT
/

1
] (4)
where
1
= t[E/f
y
]
1/2

=
and
LT
=
This expression represents a conservative approximation for any uniform plain I or H shape
with equal flanges - see Annex F2 of EC3.
The above expression for
LT
is valid for loading giving uniform moment over a span whose
ends are prevented from deflecting laterally and from twisting about a vertical axis passing
through the web. This is the basic case for lateral stability (Figure 12) for which a full
theoretical treatment is provided in Lecture 7.9.2. Variations in the conditions of loading
and/or lateral support may be allowed for by introducing modifying factors into the
expressions for
LT
or M
cr
.

For example, if there is a moment gradient between points of lateral restraint,
LT
is
calculated as follows:

LT
= (5)
where C
1
= 1,75 - 1,05 + 0,3
2
s 2,35
and is the end moment ratio defined in Figure 13.

Taking as an example the end span of a continuous beam for which = 0 gives C
1
=1,75 and
thus
LT
will be reduced to 0,76 (= 1/\ 1,75) of the value for uniform moment, leading to an
increase in _
LT
and thus in M
bRd
.
Variations in the conditions of lateral restraint may be treated by introducing k-coefficients to
modify the geometrical length L into kL when determining M
cr
. For conditions with more
restraint, values of k < 1,0 are appropriate, leading to an increase in M
cr
and thus, via a
reduction in
LT
, to increases in _
LT
and M
bRd
.
Similarly additional C-coefficients may be used directly in the determination of M
cr
to
provide modified values of
LT
appropriate for a wide range of load types. In particular, this
method should be used to calculate the reduced M
cr
appropriate for destabilising loads. These
are loads that act above the level of the beam's shear centre and are free to move sideways
with the beam as it buckles, as shown in Figure 14.

For cross-sections of the type illustrated in Figure 9c, for which the shear centre and centroid
do not lie on the same horizontal axis, evaluation of M
cr
becomes more complex and is
covered by Annex F of Eurocode 3 [1].
8. CONCLUDING SUMMARY
- Beams that are not restrained along their length and are bent about their strong axis
are subject to lateral torsional buckling.
- Unbraced span, lateral slenderness (L/i
z
), cross-sectional shape (torsional and warping
rigidities), moment distribution and end restraint are the primary influences on
buckling resistance.
- Bracing of sufficient stiffness and strength, that restrains either torsional or lateral
deformations, may be used to increase buckling resistance.
- Although elastic critical load theory provides a background for understanding the
behaviour of laterally unrestrained beams, it requires both simplifications and
empirical modification if it is to form a suitable basis for a design approach.
- In order to check the lateral buckling resistance of a trial section, its effective
slenderness
LT
must first be obtained.
- Variation in either lateral support conditions or the form of the applied loading may
be accommodated in the design process by means of coefficients k and C, used to
modify either the basic slenderness
LT
or the basic elastic critical moment M
cr
.
9. REFERENCES
[1] Eurocode 3: "Design of Steel Structures": ENV 1993-1-1: Part 1.1: General rules and
rules for buildings, CEN, 1992.
10. ADDITIONAL READING
1. Narayanan, R., Editor, "Beams and Beam Columns: Stability and Strength", Applied
Science Publishers 1983.
Chapters 1 - 3 deal with various aspects of behaviour and design of laterally
unrestrained beams.
2. Chen, W. F. and Atsuta, T. "Theory of Beam Columns Volume 2, Space Behaviour
and Design", McGraw Hill 1977.
Chapter 3 deals with laterally unrestrained beams.
3. Timoshenko, S. P. and Gere, J. M., "Theory of Elastic Stability" Second Edition,
McGraw Hill 1961.
Basic derivations for the elastic critical moment for a variety of beam problems are
provided in Chapter 6.
4. Bleich, F., "Buckling Strength of Metal Structures", McGraw Hill 1952.
Chapter 4 presents the basic theory of lateral buckling of beams.
5. Galambos, T. V., "Structural Members and Frames", Prentiss Hall 1968.
Chapter 2 deals with the fundamentals of elastic behaviour, whilst Chapter 3 covers
elastic and inelastic behaviour and design of laterally unrestrained beams.
6. Trahair, N. S. and Bradford, M. A., "The Behaviour and Design of Steel Structures",
Chapman and Hall, Second Edition, 1988.
Laterally unrestrained beams are dealt with in Chapter 6.





Lecture 7.9.2: Unrestrained Beams II
OBJECTIVE/SCOPE
To derive the basic theory of elastic lateral-torsional buckling and to discuss the physical
significance of the resulting expressions.
PREREQUISITES
Simple bending theory
Simple torsion theory
Lecture 6.4: General Methods for Assessing Critical Loads
Lectures 7.5: Columns
Lecture 7.9.1: Unrestrained Beams I
RELATED LECTURES
Lectures 7.10: Beam Columns
RELATED WORKED EXAMPLES
Worked Example 7.9: Laterally Unrestrained Beams
SUMMARY
This lecture presents the basic elastic theory for lateral-torsional buckling of beams,
commencing with the case of a simply supported beam under uniform moment. Variations in
load pattern, load level and degree of end restraint are discussed. The theoretical derivations
are separated into two Appendices, the main text being limited to a discussion of the
underlying assumptions and the physical significance of the derived expressions.
NOTATION
b flange width
C
1
coefficient to allow for type of loading
h
f
distance between flange centroids
EI
w
warping rigidity
EI
z
minor axis flexural rigidity
E Young's modulus
F applied load
F
cr
elastic critical buckling load
GI
t
torsional rigidity
d overall depth of section
L span
M moment
M
cr
applied elastic critical buckling moment
t
f
flange thickness
t
w
web thickness
1. INTRODUCTION
The basic model used to illustrate the theory of lateral-torsional buckling is shown in Figure
1. It assumes the following:
- beam is initially straight
- elastic behaviour
- uniform equal flanged I-section
- ends simply supported in the lateral plane (twist and lateral deflection prevented, no
rotational restraint in plan)
- loaded by equal and opposite end moments in the plane of the web.


This problem may be regarded as being analogous to the basic pin-ended Euler strut.
The beam is placed in its buckled position, as in Figure 2, and the magnitude of the applied
load necessary to hold it there determined by equating the disturbing effect of the end
moments, acting through the buckling deformations, to the internal (bending and torsional)
resistance of the section.

The derivation and solution to the equations leading to the critical value of applied end
moments (M
cr
) at which the beam of Figure 1 just becomes unstable is provided in Appendix
1. The physical significance of the solution and its application in cases where the assumptions
listed above do not apply are discussed in Sections 2 and 3 that follow.
2. PHYSICAL SIGNIFICANCE OF THE SOLUTION
The buckled shape of the beam, Figure 2, is now compared with the expression for the elastic
critical moment of Equation (17) in Appendix 1, i.e.
M
cr
= (17)
The presence of the flexural (EI
z
) and torsional (GI
t
and EI
w
) stiffnesses of the member in the
equation is a direct consequence of the lateral and torsional components of the buckling
deformations. The relative importance of the two mechanisms for resisting twisting is
reflected in the second square root term. Length is also important, entering both directly and
indirectly via the t
2
EI
w
/L
2
GI
t
term. It is not possible to simplify Equation (17) by omitting
terms without imposing limits on the range of application of the resulting approximate
solution. Figure 6 shows quantitatively the application of Equation (17) to the different types
of beam sections defined in the earlier Lecture 7.8.1. The region of the curves for both I-
sections of low length/depth ratios corresponds to the situation in which the value of the
second square root term in Equation (17) adopts a value significantly in excess of unity. Since
warping effects (see Appendix 1) will be most important for deep sections composed of thin
plates, it follows that the t
2
EI
w
/L
2
GI
t
term will, in general, tend to be large for short deep
girders and small for long shallow beams.

Figure 7 gives some quantitative indication of the effect of shape of cross-section for
structural steel I-beams, by comparing values of M
cr
for a beam (I) and a column (H) having
approximately equal in-plane plastic moment capacities. Clearly, lateral-torsional buckling is
a potentially more significant design consideration for the beam section which is much less
stiff laterally.

3. EXTENSION TO OTHER CASES
3.1 Load Pattern
The equivalent of Equation (8) (Appendix 1) may be set up and solved for a variety of other
load cases. Since the applied moment at any point within the span will now be a function of
x, the mathematics will be more complex. As an example, consider the beam subjected to a
central load acting at the level of the centroidal axis shown in Figure 8, for which the analysis
is outlined in Appendix 2.

The solution for this example may conveniently be compared with the basic case in terms of
the critical moments for each, i.e. maximum moment when the beam is on the point of
buckling.
Basic case: M
cr
= (t/L) \(EI
x
GI
t
) \[1+ (t
2
EI
w
/L
2
GI
t
)] (17)
Central load: M
cr
= (4,24/L) \(EI
x
GI
t
) \[1+ (t
2
EI
w
/L
2
GI
t
)] (21)
The ratio of the two constants t/4,24=0,74 is the reciprocal of the coefficient C
1
introduced in
Lecture 7.9.1. Its value is a direct measure of the severity of a particular pattern of moments
relative to the basic case.
Figure 9, which gives C
1
factors for various loading patterns, shows how lateral stability
generally increases as the moment pattern becomes less uniform.

3.2 Level of Application of Load
For transverse loads free to move sideways with the beam as it buckles, the level of
application of load (relative to the centroid) is important. The solution for a point load applied
at any level relative to the beam's centroidal axis may conveniently be obtained using an
energy approach, as outlined in Appendix 2. When the load is applied to either the top flange
or the bottom flange, e.g. by a crane trolley, the solution of Equation (21) may still be used,
providing the numerical constant is replaced by a variable, the value of which depends upon
the ratio L
2
GI
t
/EI
w
as shown in Figure 10. The reason why top flange loading and bottom
flange loading are respectively more or less severe than centroidal loading may be
appreciated from the sketches in Figure 10, which show the destabilising and stabilising
effects. Clearly this would be expected to become more significant as the depth of the section
increases and/or the span reduced, i.e. as L
2
GI
t
/EI
w
becomes smaller.

3.3 Conditions of Lateral Support
It has already been suggested in Lecture 7.9.1, that lateral support arrangements which inhibit
the growth of the buckling deformations will improve a beam's lateral stability. Equally, less
effective conditions will reduce stability. Providing the appropriate boundary conditions can
be incorporated into the analysis methods of Appendices1 and 2, any arrangement can be
dealt with.
A convenient way of including the effect of different support conditions is to redefine L in
Equation (17) as an effective length l, with the exact value of l/L depending upon the degree
of lateral bending and/or warping restraint provided. In Eurocode 3 [1] this approach is split
into the use of two factors:
k referring to end rotation on plan.
k
w
referring to end warping.
It is recommended that k
w
be taken as unity unless special provision for warping fixity is
made; k may vary from 0,5 for full fixity, through 0,7 for one end fixed and one end free, to
1,0 for no rotational fixity.
One case of particular practical interest is the cantilever, for which some results are presented
in Figure 11.

These show that:
1. Cantilevers under end moment are less stable than similar, simply supported, beams.
2. Concentrating the moment adjacent to the support, as happens when the applied
loading changes from pure moment to an end load or to a distributed load, improves
lateral stability.
3. The effect of load height is even more significant for cantilevers than for simply
supported beams.
3.4 Continuous Beams
Continuity may be present in two different forms:
1. In a beam that has a single span vertically but is subdivided, by intermediate lateral
supports, so that it exhibits horizontal continuity between adjacent segments, see
Figure 12a.
2. In the vertical plane as illustrated in Figure 12b.

For the first case a safe design will result if the most critical segment, treated in isolation, is
used as the basis for designing the whole beam. For the second case account should be taken
of the actual moment diagram within each span, produced by the continuity, by using the C
1

factor. If the top flange can be considered as laterally restrained because of attachment to a
concrete slab, particular attention should be paid to the regions in which the lower flange is in
compression, e.g. the support regions or regions where uplift loads can occur.
3.5 Beams Other than Doubly-Symmetrical I-sections
The basic theoretical solution of Equation (17) is valid for members that are symmetrical
about their horizontal axis, e.g. a channel with the web vertical, providing the moments act
through the shear centre (which will not now coincide with the centroid). However, sections
symmetrical only about the vertical axis, e.g. an unequal flanged I, require some modification
so as to allow for the so-called Wagner effect. This arises as a direct result of the vertical
separation of the shear centre and the centroid and leads to either an increase or a decrease in
the section's torsional rigidity. Thus lateral stability will be improved when the larger flange
is in compression and reduced when the smaller flange is in compression as compared with
equal flange sections having comparable properties.
Sections with no axis of symmetry will not actually buckle but will deform by bending about
both principal axes and by twisting from the onset of loading. They should therefore be
treated in the same way as symmetrical sections under biaxial bending.
3.6 Restrained Beams
The elastic critical moment for a doubly symmetrical I-beam provided with continuous elastic
torsional restraint, of stiffness equal to K| , is:
M
cr
=
Rearranging this shows that the beam behaves as if its torsional rigidity GI
t
were increased to
(GI
t
+ K| L
2
/t
2
), thereby permitting a ready assessment of the effectiveness of the restraint.
An important practical example of such a restraint would be that provided by the bending
stiffness of profiled steel sheeting (used typically in roof construction) spanning at right
angles to the beam.
4. CONCLUDING SUMMARY
- The elastic critical moment which causes lateral-torsional buckling of a slender beam
may be determined from an analysis which has close similarities to that used to study
column buckling.
- Examination of the expression for the elastic critical moment for the basic problem
enables the influence of cross-sectional shape, as it affects the beam's resistance to
lateral bending (EI
z
), torsion (I
t
) and warping (I
w
), to be identified; it also
demonstrates the importance of unbraced span length.
- Extensions to the basic theory permit the effects of load pattern, end restraint and
level of application of destabilising loads to be quantified.
- Load patterns which produce non-uniform moment may be compared with the basic,
uniform moment case using the coefficient C
1
; since most of these other cases will be
less severe, C
1
values greater than 1,0 are the norm.
5. REFERENCES
[1] Eurocode 3: "Design of Steel Structures": ENV 1993-1-1: Part 1.1: General rules and
rules for buildings, CEN, 1992.
6. ADDITIONAL READING
1. Narayanan, R., Editor, "Beams and Beam Columns: Stability and Strength", Applied
Science Publishers 1983.
Chapters 1 - 3 deal with various aspects of behaviour and design of laterally
unrestrained beams.
2. Chen, W. F. and Atsuta, T. "Theory of Beam Columns Volume 2, Space Behaviour
and Design", McGraw Hill 1977.
Chapter 3 deals with laterally unrestrained beams.
3. Timoshenko, S. P. and Gere, J. M., "Theory of Elastic Stability" Second Edition,
McGraw Hill 1961.
Basic derivations for the elastic critical moment for a variety of beam problems are
provided in Chapter 6.
4. Bleich, F., "Buckling Strength of Metal Structures", McGraw Hill 1952.
Chapter 4 presents the basic theory of lateral buckling of beams.
5. Galambos, T. V., "Structural Members and Frames", Prentice Hall 1968.
Chapter 2 deals with the fundamentals of elastic behaviour, whilst Chapter 3 covers
elastic and inelastic behaviour and design of laterally unrestrained beams.
6. Trahair, N. S. and Bradford, M. A., "The Behaviour and Design of Steel Structures",
Chapman and Hall, Second Edition, 1988.
Laterally unrestrained beams are dealt with in Chapter 6.
APPENDIX 1: ANALYSIS OF LATERAL-TORSIONAL
BUCKLING
Derivation of Governing Equations
The deformed state of the beam is shown in Figure 3, which identifies the deflections (u and
v) and the twist (|). A new co-ordinate system q , , which deflects with the beam, is also
illustrated.

Bending in the , and q , planes and twisting about the , axis are governed by:
EI
y
(1)
EI
z
(2)
GI
t
(3)
In Equations (1) and (2) the flexural rigidities and curvatures in the , and the q , planes
have been replaced by the values for the yx and zx planes, on the basis that | is a small angle.
Equation (3) includes both mechanisms available in a thin-walled section to resist twist; the
first term corresponds to that part of the applied torque which is resisted by the development
of shear stresses, whilst the second term allows for the influence of restrained warping. This
latter phenomenon arises as a direct result of the type of axial flange deformation, illustrated
in Figure 4a, that occurs in an I-section subject to equal and opposite end torques. The two
flanges tend to bend in opposite senses about a vertical axis through the web, with the result
that originally plane sections do not remain plane. On the other hand, for the cantilever of
Figure 4b, it is clear that warping deformations must be at least partly inhibited elsewhere
along the span, since they cannot occur at the fixed end. This induces additional axial stresses
in the flanges; the pair of couples, or bimoment, due to this additional stress system provides
part of the section's resistance to twist. In the case of lateral instability, restraint against
warping arises as a result of adjacent cross-sections wanting to warp by different amounts.

For an I-section, the relative magnitudes of the warping constant I
w
and the torsion constant I
t

are:
I
w
= I
z
h
f
2
/4 and I
t
=
They will be affected principally by the thickness of the component plates and by the depth of
the section. For compact column-type sections the first term in Equation (3) will tend to
provide most of the twisting resistance, whilst the second term will tend to become dominant
for deeper beam shapes.
Consideration of the buckled shape using Figures 2, 3 and 5 enables the components of the
applied moment in the , and q , planes and about the , axis to be obtained as:
M = Mcos| , Mq = Msin| , M, = Msino (4)

Since | is small, cos| ~ 1 and sin| ~ |, whilst Figure 5 shows that sino may be approximated
by - . Thus Equations (1) - (3) may be written as:
EI
y
= M (5)
EI
z
= M| (6)
GI
t
(7)
Since Equation (5) contains only the vertical deflection (v), it is independent of the other two;
it controls the in-plane response of the beam described in Lecture 7.5.1. Equations (6) and (7)
are coupled in u and | , the buckling deformations; their solution gives the value of elastic
critical moment (M
cr
) at which the beam becomes unstable. Combining them gives:
EI
w
(8)
Solution
The solution of Equation (8) is made far simpler if the warping stiffness (I
w
) is assumed to be
zero. The results obtained are then directly applicable to beams of narrow rectangular cross-
section but are conservative for the normal range of I-sections. Equation (8) therefore reduces
to:
(9)
Putting
2
= enables the solution to be written as:
| = Acos x = Bsin x (10)
Noting the boundary conditions at both ends gives
When x = 0, | = 0; then A = 0 (11)
When x = L, | = 0; then Bsin L = 0
and either B = 0, or (12)
sin L = 0 (13)
The first possibility gives the unbuckled position whereas the second gives:
L = 0, t, 2t (14)
and the first non-trivial solution is:
L = t (15)
which gives:
M
cr
= (t/L) \(EI
x
GI
t
) (16)
Since the form of Equation (9) is identical to the form of the basic Euler strut equation all of
the same arguments about its solution apply.
Returning to the original Equation (8), this may be solved to give:
M
cr
= (t/L) \(EI
x
GI
t
) \[1+ (t
2
EI
w
/L
2
GI
t
)] (17)
The inclusion of warping effects therefore enhances the value of M
cr
by an amount which is
dependent on the relative values of EI
w
and GI
t
.
APPENDIX 2: BUCKLING OF A CENTRALLY
LOADED BEAM
Using the approach of Appendix 1 and noting from Figure 7 that the vertical load will
produce a moment about the x-axis of W(u
o
- u)/2 when the beam is in its buckled position,
enables Equations (4) to be re-written as:
M =
Mq = (18)
M, =
Replacing Equations (5) - (7) by their revised forms and eliminating u from the second and
third of these gives:
EI
w
(19)
which may be solved for W
cr
to yield approximately:
W
cr
= 5,4 (20)
The moment at midspan is then:
M
cr
= (21)
The alternative means of obtaining elastic critical loads uses the energy method, in which the
work done by the applied load during buckling is equated to the additional strain energy
stored as a result of the buckling deformations. Considering an element of the longitudinal
axis of the beam of length dx located at C, bending in the , plane causes the end B of the
beam to rotate in the , plane by:
(22)
The vertical component of this is:
(23)
Summing these for all elements between x= 0 and x = L/2 gives the lowering of the load W
from which the work is:
(24)
The strain energy stored as a result of lateral bending, twisting and warping is:
(25)
Assuming a buckled shape of the form:
(26)
and equating Equations (24) and (25) enables the critical value of W to be obtained.
Use of this technique permits examination of the case in which the load is applied at a level
other than the centroidal axis. Assuming W to act at a vertical distance (a) above the centroid,
the additional work will be:
Wa (1 - cos |
o
) = Wa |
o
2
/2
in which |
o
is the value of | at the load point. This must be added to Equation (24).

S-ar putea să vă placă și