Sei sulla pagina 1di 295

Thin-Walled Structures

v

Contents

CHAPTER 1

Airplane and ship structures

1

Structures and Engineering

1

Principal structural units

2

Design

4

Loads

6

Function of ight vehicle structural members

6

Ships structures

9

Key words and concepts from Chapter 1

16

References

16

CHAPTER 2

Bars Subjected to Axial Loads

17

Axially loaded bar

17

The tensile test

19

Effect of temperature on strain

23

Bar reference axis

24

Linear elastic response

26

EXAMPLE 2.1 Axial bar with a specified uniform distributed load and specified end
displacements 27
EXAMPLE 2.2 A bar with fixed ends and subjected to an axial point force. 28

Work and energy methods

30

Concept of virtual displacement

30

Principle virtual work

32

EXAMPLE 2.3 Approximating the response of a bar using PVW. 34

Contents

vi

Thin-Walled Structures

Strain energy density

36

EXAMPLE 2.4 Strain energy of a bar with fixed ends and subjected to an axial point
force. 38
EXAMPLE 2.5 An elastic bar subjected to two forces and a thermal load 40

Castiglianos rst theorem

41

EXAMPLE 2.6 Response of a stepped bar by Castiglianos first theorem 43

Complementary virtual work

45

Complementary strain energy

47

Relationship between the complementary strain energy and the strain energy
densities

49

EXAMPLE 2.7 Application of complementary virtual work to an elastic bar 50

Generalized form of Castiglianos second theorem

52

EXAMPLE 2.8 Stepped bar response by Castiglianos second theorem 54
EXAMPLE 2.9 A suspended bar subjected to self weight 55

Trusses

56

EXAMPLE 2.10 Three bar planar truss 60
EXAMPLE 2.11 Three bar truss with lack of fit 61

References

62

Problems

63

CHAPTER 3

Axial Normal Stress in Pure Bending and Extension

67

Pure bending

67

Geometry of deformation

68

Bending normal stress exure formula

75

EXAMPLE 3.1 Bending normal stress distribution in a cantilever beam with a thin-
walled zee section. 78
EXAMPLE 3.2 Lateral displacements of the zee section beam 80

Moments of areas

83

EXAMPLE 3.3 Thin-walled zee section properties by the composite body technique 86
EXAMPLE 3.4 Semicircular section with two stringers 88

Extension, pure bending, and thermal effects for multi-material beams

89

EXAMPLE 3.5 A multi-material beam with a symmetric cross section 92

Problems

96

CHAPTER 4

Axial Force, Shear Force and Bending Moment
Diagrams

99

Method of sections

99

Differential equation method

101

EXAMPLE 4.1 Cantilever wing with tip tank 105
EXAMPLE 4.2 The air load acting on a wing given as discrete data. 109

Semi-graphical method

116

EXAMPLE 4.3 Uniform barge with symmetric load 116

Buoyancy Force Distribution on Ships

118

Thin-Walled Structures

vii

Contents

References

121

Problems

121

CHAPTER 5

Bending of Beams under Transverse Loads

125

Approximations for slender beams

125

Beam displacements

126

EXAMPLE 5.1 A statically indeterminate beam with an unsymmetrical cross section.
131
EXAMPLE 5.2 Contact between two cantilever beams. 135

Complementary virtual work and complementary energy

139

Complementary virtual work

140

Complementary strain energy

142

Beam displacement by Castiglianos second theorem

145

EXAMPLE 5.3 End rotations of a simply supported beam subject to an end moment 145
EXAMPLE 5.4 Tip displacement of a cantilever wing spar a under distributed load 147
EXAMPLE 5.5 Strut-braced spar 149

Problems

153

CHAPTER 6

Shear Flow due to Shear Forces

155

Shear ows and shear stresses due to bending in a rectangular section beam

155

Shear ows due to transverse shear forces in open section beams

158

EXAMPLE 6.1 Shear flow distribution in a tee beam 160

Shear center of a thin-walled open section

163

EXAMPLE 6.2 Shear center location in an unsymmetrical section 167

Skin-stringer idealization

170

EXAMPLE 6.3 Shear flows in a stringer-stiffened C-section 173

Inuence of transverse shear deformations on bending

177

Transverse shear strains, forces, and complementary energy density

178

Complementary energy density obtained from a two-dimensional element of the
wall

180

Determination of the transverse shear compliances

183

EXAMPLE 6.4 Shear compliances of a stiffened blade section 184
EXAMPLE 6.5 Deflection of a cantilevered beam due to bending and shear
deformation. 186

Problems

188

CHAPTER 7

Bars Subjected to Torsional Loads

191

Uniform torsion of a circular tube

191

Uniform torsion of an open section

197

EXAMPLE 7.1 Torsional response of a thin-walled open section and an equivalent
closed section 201

Non-uniform torsion; governing boundary value problem

202

Contents

viii

Thin-Walled Structures

EXAMPLE 7.2 A uniform distributed torque acting on a bar with fixed ends. 203
EXAMPLE 7.3 A point torque acting on a bar with fixed ends. 204

Virtual work and strain energy

205

Strain energy density

207

Complementary virtual work and energy

208

Complementary strain energy

209

Unit twist of a single cell beam due to shear ow

210

Hookes law

211

Shear strain-displacement relation

212

Tangential displacement of a typical point on the contour

212

Relation between the shear ow and unit twist

214

Uniform torsion of a thin-walled closed section with a contour of arbitrary
shape

215

EXAMPLE 7.4 Torsion of a circular, bi-material section 218

Shear center of a closed section

219

EXAMPLE 7.5 Shear center location of a single-cell, closed section having one axis of
symmetry. 219

Uniform torsion of multi-cell closed sections

222

EXAMPLE 7.6 Uniform torsion of a two-cell section 225

Resultant of a constant shear ow in a curved branch

226

EXAMPLE 7.7 Torsion of a five cell closed section; circuit shear flow 229

Torsion of hybrid sections

231

References

232

Problems

233

CHAPTER 8

Criteria for Initial Yielding

237

Ductile and brittle behavior

237

Criteria for initial yielding of ductile materials

238

Stress transformation equations for generalized plane stress

241

Principal stresses and maximum shear stress

243

EXAMPLE 8.1 Maximum shear stress in tensile test 246
EXAMPLE 8.2 Mohrs circle for hydrostatic stress state 246
EXAMPLE 8.3 Principal stresses and maximum shear stress 247

Octahedral shear stress

249

Mises criterion for initiation of yielding

252

Maximum shear-stress criterion

254
EXAMPLE 8.4 Factor of safety against initial yielding 256
EXAMPLE 8.5 Stress responses of a stringer-stiffened, single cell beam. 257
EXAMPLE 8.6 Minimum weight design of the beam in Example 8.5 subject to a
constraint on initial yielding 261
References 264
Thin-Walled Structures ix
Contents
CHAPTER 9 Buckling 265
One-degree of freedom model 265
Static equilibrium 266
Stability analysis 267
Perfect Columns 269
EXAMPLE 9.1 Critical load for clamped-free boundary conditions (B) 272
Imperfect columns 275
Eccentric load 275
Geometric imperfection 277
Column Design Curve 279
Inelastic buckling 280
Bending of thin plates 283
Contents
x Thin-Walled Structures

Thin-Walled Structures

1

CHAPTER 1

Airplane and ship
structures

The objective of Thin-Walled Structures is to provide an understanding of the basic concepts of stress and defor-
mation of stiffened-shell structures with applications to aerospace and ocean vehicles. In this chapter we discuss
why a structure carries load, the types of loads acting on airplane and ship structures, the denitons and functions
of principal structural units, and the relation of structural design to the overall design.

1.1 Structures and Engineering

A structure may be dened as any assemblage of materials which is intended to sustain loads. Structures function
to protect people and things, and are so common and familiar to us that when we are informed of their use and
form we conceive of ourselves as knowledgeable. After all, it does not take a structural engineering degree to
build an ordinary shed. However, when asked to build an airplane or a ship we would probably be more hesitant,
since if an airplane or a ship breaks many people are likely to be killed. The crux of the issue in the design of
complex structures, like aircraft, is to not only know the use and form of the structure, but also to know

why

a
structure is able to carry load.
Although Galileo (1564-1642) began an inquiry into the strength of materials, it is Robert Hooke (1635-
1702) who is credited with providing the answer as to why structures carry load. A historical account of Hookes
discoveries is discussed in chapter two of an informative and readable book on structures by Gordon (1978). Gor-
dons text is the source for what is written here. By Newtons law of action and reaction we know that isolated
forces do not exist in nature. A force acting on an inanimate solid is reacted by a force produced by the solid. But
how does the solid produce such a reaction force? Hooke in 1676 recognized that every kind of solid changes its
shape when subjected to a mechanical force and it is the change in shape which enables the solid to provide the
reaction force. This shape change extends to the very ne scale of molecules where there is a large resistance to
stretching and compressing of chemical bonds. Hookes measurements also showed that many solid materials
recovered their original shape when they are unloaded; i.e, they are elastic. The science of elasticity is about the
interactions between forces and deections in materials and structures. The formulation of the mathematical the-
ory of elasticity for a solid continuum is well established; e.g., see Sokolnikoff (1956). Exact elasticity solutions
are known for a small, but important, number of structural problems. However, the mathematical solutions to
elasticity problems are usually challenging. Approximations to elasticity theory which exploit the geometry and

Airplane and ship structures

2

Thin-Walled Structures

material construction of practical structures is called structural mechanics. For example, the structural mechanics
of beams, plates, and shells approximate elasticity by exploiting the fact that one or two dimensions of the solid
body are very small with respect to the remaining dimensions. Structural mechanics reduces the mathematical
complexity somewhat relative to three-dimensional elasticity, and it is the theory most often used by the practic-
ing engineer.
Two important properties common to the analysis of any structure are

stiffness

and

strength

. Stiffness is a
measure of the force required to produce a given deection, and strength refers to the force, or force intensity,
necessary to cause failure. A criterion for failure is required in order to determine the strength of a structure, and
this depends upon the particular application. For example, failure can be dened when a stress (internal force
intensity) exceeds the yield stress of the material, or failure can mean excessive displacements which occur dur-
ing buckling. The stiffness and strength of a structure depend on its geometrical conguration, connections, and
the stiffness and strength of the materials from which it is made.
It is important to recognize that structures are made from materials, and that the history of structures follows
the development of materials and the development of tools to fabricate the materials. The evolution of the air-
frame, for example, is tied closely to the introduction of materials and cost-effective means for their fabrication.
For example, early aircraft were constructed of wire-braced wood frames with fabric covers. Currently, advanced
composite materials are very attractive for weight-sensitive structures, like aircraft, because of their high stiff-
ness-to-weight and strength-to-weight ratios.
The distinction between structures and materials is not always clear. It may be said that the forward-swept
wing on Grumman's X-29 demonstrator airplane is a structure, and the material it is made from is an advanced
composite. However, advanced ber-reinforced composites are made from stiff, strong, continuous bers embed-
ded in a pliant matrix. The complex constitution of an advanced composite, therefore, may be considered either
as materials or structures.

1.2 Principal structural units

The principal structural units of xed-wing airplanes are the fuselage, wings, stabilizers, control surfaces, land-
ing gear, nacelle and engine mounts. Light airplanes are shown in Fig. 1.1 and a heavier airplane (Douglas DC-3)
is shown in Fig. 1.2. A cargo ship is depicted in Fig. 1.3, and its principal structural units will be discussed in
more detail in Section 1.6. After some study of the structures shown in these gures, it is reasonable to suggest
that some principal structural units such as the fuselage, wings, and ship hull have the features commonly attrib-
utable to a beam. That is, two dimensions of the overall component, or the cross-sectional dimensions, are small
with respect to the third, or longitudinal, dimension. Indeed, to simplify the analysis of such complex structures,
we can approximate them as slender, built-up beams!
Hence, the basic conceptualizing we make for complex vehicle structures such as aircraft and ships is that a
fuselage, a wing, or a hull is a thin-walled beam. That is, a vehicle structure as a whole is assumed to be a one-
dimensional structural element in the mathematical sense that its response under load can be described by ordi-
nary differential equations. Aero-hydrodynamic and other loads that act on the structure cause extensional, bend-
ing, and torsional deformations of the structure. The cross section of the structure is built from many actual
structural elements such as spars, frames, and panels. This beam assumption is particularly suited for the analysis
required in preliminary design. Of course not all principal structural units can be modeled as a beam. In constrast
to a high aspect ratio wing, a delta wing whose span and chord are of comparable value (low aspect ratio) is not
modeled very well by using the beam assumption.

Thin-Walled Structures

3

Principal structural units
Fig. 1.1 Principal structural units of light airplanes (from Aircraft Basic Science, 1948)
(b) Taylorcraft airplane
(a) Global Swift

Airplane and ship structures

4

Thin-Walled Structures

1.3 Design

An aircraft or ship may be considered a system consisting of several subsystems; a structural subsystem, a con-
trol subsystem, a propulsion subsystem, cargo handling subsystem, etc. Vehicle design must consider the total, or
integrated, system to achieve optimal performance.

An important contribution to the overall vehicle performance
is to minimize weight in the structural subsystem design.

A minimum weight vehicle structure can carry the same
payload with less fuel consumption. In addition, a lighter structure can reduce operating costs through less main-
tenance, and also may reduce initial cost by requiring less labor for fabrication. Modern engineering design has
been revolutionized by the development of high-speed computers combined with optimization theory. As a math-
ematical problem in optimization, structural design may be considered as the development of a computational
algorithm for choosing member spacing and dimensions such that weight is minimized (objective) subject to
constraints on strength and stiffness. The role of structural mechanics in this design process is to provide a
description of the state of stress and deformation throughout the structure for a given structural conguration,
such that the constraints can be evaluated. Structural mechanics provides the theory for the analysis of the struc-
tural response (state of stress and deformation).
Fig. 1.2 Principal structural units of a Douglas airplane (from Aircraft Basic Science, 1948)

Thin-Walled Structures

5

Design

The overall dimensions of a vehicle structure are usually determined by more general requirements rather
than for structural considerations. Thus, the structural design begins with the overall dimensions given, and two
levels of design may be distinguished. The rst level is called

preliminary design

, and at this level the locations
and dimensions of the principal structural members are determined. The second level is called

detail design

, and
Fig. 1.3 Levels of structural analysis for a large ship (from Hughes, 1983)

Airplane and ship structures

6

Thin-Walled Structures

at this level the geometry and dimensions of the local structure like joints, cutouts, reinforcements, etc. are deter-
mined.

1.4 Loads

The rst step in preliminary design is to determine the external loads acting on the structure. Maneuvering ight
vehicles are subjected to gravity, aerodynamic forces, and inertial loads. In addition, the landing loads, and wind
gust loads during turbulent weather must be considered. Ships are subjected to gravity, buoyancy forces, and
inertial loads. Also, wave loads and other hydrodynamic loads such as slamming, sloshing of liquid cargos, etc.,
must be considered in ship design. The calculation of aerodynamic and hydrodynamic forces are sufciently
complex such that their determination is done by specialists rather than by designers.
Loads on a vehicle structure may be classied as static or dynamic, and either deterministic or probabilistic.
Gust loading conditions for aircraft and wave loading conditions for ships are not known with absolute certainty,
so that these load magnitudes are estimated on a statistical basis using probability theory together with past expe-
rience. The type of loading has a direct inuence on the type of structural response analysis required. For exam-
ple, dynamic loading requires a structural dynamics analysis.
In traditional structural design most of these loads are not affected by the structural conguration or dimen-
sions of the members. They are a function of the wing shape, or hull shape, and other nonstructural factors.
Hence, the determination of the loads, a very crucial aspect of the design process, is essentially a separate task
typically performed by an aerodynamicist or hydrodynamicists. In modern exible vehicles the loads greatly
inuence the shape and so aeroelastic or hydroelastic load analysis must be performed. That is, the structural and
aero/hydrodynamic analysis must be combined to obtain the correct loads. This interaction between the structure
and the shape is so important for exible vehicles made from composite materials that it is expected that in the
future the shape and structural design will be combined.

1.5 Function of ight vehicle structural members

The following description of the functions of ight vehicle structures is excerpted from Rivello (1969, Section 7-
6).

The structure of a ight vehicle usually has a dual function: it transmits and resists the
forces which are applied to the vehicle, and it acts a cover which provides the aerodynamics
shape and protects the contents of the vehicle from the environment. This combination of roles
is fortunate since, from the standpoint of structural weight, the most efcient location for the
structural material is at the outer surface of the vehicle. As a result, the structures of most ight
vehicles are essentially thin shells. If these shells are not supported by stiffening members,
they are referred to as

monocoque

. When the cross-sectional dimensions are large, the wall of
a monocoque structure must be relatively thick to resist bending, compressive, and torsional
loads without buckling. In such cases a more efcient type of construction is one which con-
tains stiffening members that permit a thinner covering shell. Stiffening members may also be
required to diffuse concentrated loads into the cover. Constructions of this type are called

semimonocoque

. Typical examples of semimonocoque body structures are shown in Fig. 7-5.
While at rst glance these structures appear to differ considerably, functionally there are simi-
larities. Both have thin-sheet coverings, longitudinal stiffening members, and transverse sup-

Thin-Walled Structures

7

Function of ight vehicle structural members

porting elements which play similar structural roles.
In semimonocoque structures the cover, or skin, has the following functions:
1. It transmits aerodynamic forces to the longitudinal and transverse supporting members by
plate and membrane action (Chap. 13).
2. It develops shearing stresses which react the applied torsional moments (Chap. 8) and
shear forces (Chap. 9).
3. It acts with the longitudinal members in resisting the applied bending and axial loads
(Chaps. 7, 15, and 16).
4. It acts with the longitudinals in resisting the axial load and with the transverse members in
reacting the hoop, or circumferential, load when the structure is pressurized.
In addition to these structural functions, it provides an aerodynamic surface and cover for
the content of the vehicle.

Spar webs

(Fig. 7-5b) play a role that is similar to function 2 of the
skin.
The longitudinal members are known as

longitudinals

,

stringers

, or

stiffeners

. Longitudi-
nals which have large cross-sectional areas are referred to as

longerons

. These members
serve the following purposes:
1. They resist bending and axial loads along with the skin (Chap. 7).
2. They divide the skin into small panels and thereby increase its buckling and failing stresses
(Chaps. 15 and 16).
3. They act with the skin in resisting axial loads caused by pressurization.
Longitudinal
stringers
Transverse
frames
Cover Skin
Cover skin
Spar web
Spar cap
Transverse
rib
Longitudinal
stringers
(a)
(b)
Fig. 7-5 Typical semimonocoque construction. (a) Body
structures: (b) aerodynamic surface structures.

Airplane and ship structures

8

Thin-Walled Structures

The spar caps in aerodynamic surface perform functions 1 and 2.
The transverse members in body structures are called

frames

,

rings

, or if they cover all or
most of the cross-sectional area,

bulkheads

. In aerodynamic surfaces they are referred to as

ribs

. These members are used to:
1. Maintain the cross-sectional shape.
2. Distribute concentrated loads into the structure and redistribute stresses around structural
discontinuities (Chap. 9).
3. Establish the column length and provide end restraint for the longitudinals to increase their
column buckling stress (Chap. 14).
4. Provide edge restraint for the skin panels and thereby increase the plate buckling stress of
these elements (Chap. 16).
5. Act with the skin in resisting the circumferential loads due to pressurization.
The behavior of these structural elements is often idealized to simplify the analysis of the
assembled component. The following assumptions are usually made:
1. The longitudinals carry only axial stress.
2. The webs (skin and spar webs) carry only shearing stresses.
3. The axial stress is constant over the cross section of each of the longitudinals, and the
shearing stress is uniform through the thickness of the webs.
4. The transverse frames and ribs are rigid within their own planes, so that the cross section is
maintained unchanged during loading. However, they are assumed to possess no rigid-
ity normal to their plane, so that they offer no restraint to warping deformations out of
their plane.
When the cross-sectional dimensions of the longitudinals are very small compare to the
cross-sectional dimensions of the assembly, assumptions 1 and 3 result in little error. The
webs in an actual structure carry signicant axial stresses as well as shearing stresses, and it
is therefore necessary to use an analytical model of the structure which includes this load-car-
rying ability. This is done by combining the effective areas of the webs adjacent to a longitudi-
nal with the area of the longitudinal into a

total effective area

of material which is capable of
resisting bending moments and axial forces. A method for determining this effective area is
given in Sec. 15-7. In the illustrative examples and problems on stiffened shells in this and suc-
ceeding chapters it may be assumed that his idealization has already been made and that
areas given for the longitudinals are the total effective areas. The fact that the cross-sectional
dimensions of most longitudinals are small when compared with those of the stiffened-shell
cross section makes it possible to assume without serious error that the area of the effective
longitudinal is concentrated at a point on the midline of the skin where it joins the longitudinal.
The locations of these idealized longitudinals will be indicated by small circles, as shown in
Fig. 7-6b. In thin aerodynamic surfaces the depth of the longitudinals may not be small com-
pared to the thickness of the cross section of the assembly, and more elaborate idealized
model of the structure may be required.
The fewer the number of longitudinals, the simpler the analysis, and in some cases several
longitudinal may be lumped into a single effective longitudinal to shorten computations. (Fig. 7-
6). On the other hand, it is sometimes convenient to idealize a monocoque shell into an ideal-
ized stiffened shell by lumping the shell wall into idealized longitudinals, as shown in Fig.7-
7,and assuming that the skin between these longitudinals carries only shearing stresses.. The
simplication of an actual structure into an analytical model represents a compromise, since
elaborate models which nearly simulate the actual structure are usually difcult to analyze. A
more complete discussion of the idealization of shell structures will be found in Ref. 4
Once the idealization is made, the stresses in the longitudinals due to bending moments,

Thin-Walled Structures

9

Ships structures

axial load, and thermal gradients can be computed from the equations of this chapter if the
structure is long compared to its cross-sectional dimensions and if there are no signicant
structural or loading discontinuities in the region where the stresses are computed. In many
ight structures the cross section tapers; the effects of this taper upon the stresses are dis-
cussed in Chap. 9. When discontinuities or other conditions arise which violate the analytical
assumptions made in the Bernoulli-Euler theory, it is necessary to analyze the stiffened shell
as an indeterminate structure (Chaps. 11 and 12).

1.6 Ships structures

The following description of the distortion and functions of ship structures is excerpted from Muckle (1967).

The Distortion of the ships structure
The study of the static forces on the ship has shown that the ship can bend in a longitudi-
nal vertical plane like a beam. This is one of the most important types of distortion to which the
ship is subjected, and is one in which the entire structure of the ship takes part. While consid-
ering this longitudinal bending of the structure it should be mentioned that it is also possible for
the ship to bend in horizontal plane. Consider a ship moving diagonally across a regular wave
system as in Fig. 4. The crests are not perpendicular to the centre line of the ship and Fig. 5
Actual skin and web
carries axial and
shear stresses
Effective longitudinals
(axial stress only)
Idealized webs
(shear stress only)
(a)
(b)
Fig. 7-6 Idealization of semimonocoque structure. (a) Actual
structure; (b) idealized structure
Wall carries axial
and shear stresses
(a)
Effective longitudinals
(axial stress only)
Idealized web
(shear stress only)
(b)
Fig. 7-7 Idealization of a monocoque shell. (a) Mono-
coque shell; (b) idealization

Airplane and ship structures

10

Thin-Walled Structures

shows that the slope of the waves at various points in the length of the ship varies, being
sometimes positive and sometimes negative. This means that there are sideways forces acting
on the ship which will not only cause swaying, but also bending in the horizontal plane. This
bending has in the past been neglected and it is safe to say that the forces and moments gen-
erated are likely to be of small amount.
Referring again to Fig. 5, it will be evident that, because of the variation in the wave slope
at different sections in the length, not only will sideways forces be generated but there will also
be moments applied at the various sections. As these may change sign along the length of the
ship, twisting is possible with the consequent generation of torsional stresses. Once again it is
perhaps doubtful whether this type of distortion is important from the point of view of the
strength of the structure. The problem has been, partially investigated in the past, and at the
present there appears to be some interest in it in view of the tendency to increase the size of
hatch openings, thus reducing the torsional rigidity of the structure.
Consider now a transverse section of a ship as shown in Fig. 6. This section is subject rst
of all to static pressure due to the surrounding water. It will also be subjected to internal load-
ing due to the weight of the structure itself and the weight of the cargo etc. which is carried.
The effect of these static forces is to cause transverse distortion of the section, as shown by
dotted lines in Fig. 6. It is worthy of note that this type of distortion would take place regardless
of whether there was bending in the longitudinal direction. It is possible therefore to recognise
an entirely independent study dealing with the transverse deformation of the ships structure.
W
a
v
e

c
r
e
s
t
W
a
v
e

c
r
e
s
t
W
a
v
e

c
r
e
s
t
W
a
v
e

c
r
e
s
t
Fig. 4 Ship moving diagonally across waves
3/4 length
F.P.
A.P.
1/4 length
Amidships
Fig. 5 Wave surface at various
positions in length

Thin-Walled Structures

11

Ships structures

Not only do the water pressure and the local internal loads cause transverse bending but it
is possible to have local deformation of the structure due to these forces. A typical example of
this is the bottom plating of a ship between oors or longitudinals. Fig. 7 shows a strip of such
plating between two oors or longitudinals. The tendency is for the plating to bend as a beam
in between these members. Other parts of the structure which could be deformed under local
loads are tank top plating, bulkheads, girders under heavy loads such as machinery etc. In this
way it will be seen that there is another aspect of the strength of the structure which may be
dened as local deformation.
Summarising this section, it is clear that the overall problem of the strength of the ships
structure may be conveniently divided into three sections:
(1) Longitudinal strength (2) Transverse strength (3) Local strength
Since any given part of the structure of the ship may be subjected to one or more of the
modes of distortion discussed, it will be seen that the resultant state of stress in that part could
Cargo load
Cargo load
Water pressure load
Fig. 6 Distortion of transverse section due to static loading
Floors
Outer bottom
Inner bottom
Water pressure
Fig. 7 Distortion of bottom plating due to water pressure

Airplane and ship structures

12

Thin-Walled Structures

be very complex. It is for this reason that, in a rst study at least of the strength of the ships
structure, longitudinal bending, transverse bending and local bending are treated entirely inde-
pendent, so that each of the three divisions of the subject of strength of ships quoted above
can be investigated separately. This is the only realistic way of tackling the problem.
Function of the ships structure
It has been shown that the ship is capable of bending in a longitudinal vertical plane and it
follows therefore that there must be material in the ships structure which will resist this bend-
ing; or in other words there must be material distributed in the fore and after direction to full
this purpose. It follows that any material distributed over a considerable portion of the length of
the ship will contribute to the longitudinal strength. Items which come into this category are the
side and bottom shell plating, inner bottom plating and any decks which there may be. Fig. 8 is
an outline section showing these items. As far as decks are concerned, it is usual to consider
only the material abreast the line of openings, such as hatches and engine casings.
It will be clear that this longitudinal material forms a box girder of very large dimensions in
relation to its thickness. Consequently, unless the plating was stiffened in some way it would
be incapable of with standing compressive loads. For this reason therefore it becomes neces-
sary to t transverse rings of material spaced from 2 ft. to 3 ft. apart throughout the length of
the ship. This is the procedure which is adopted in what is usually called a

transversely framed


ship. The transverse stiffening consists of three parts; in the bottom between the outer and
inner bottoms there are several vertical plates called oors which have lightening and access
holes cut in them as shown in Fig. 9; in the sides of the ship rolled sections called

side frames

,
are welded to the plating (see Fig. 9); the decks are also supported by rolled sections welded
to the plating, called

beams

. The oors, side frames and beams at the various decks are con-
nected by means of brackets so that a continuous transverse ring of material is provided. As
stated earlier, the spacing of these transverse rings, usually called the

frame spacing

, is
between 2 ft. and 3 ft. and depends upon the length of the ship. It will be seen that the effect of
Upper deck
plating
2nd deck
plating
Side
shell
Inner bottom
plating
Centre
girder
Margin
plate
Bottom shell
Fig. 8 Section through ship showing material resisting longitudinal bending

Thin-Walled Structures

13

Ships structures

supporting the plating in this way is to reduce the unsupported span and hence to raise the
buckling strength of the plating, to enable it to carry compressive loads.
Another function of these transverse rings is to prevent transverse distortion of the struc-
ture, so that the oors, side frames and beams are the main items contributing to the trans-
verse strength of the structure of the ship. The main force involved here is that due to water
pressure and, as this will be greatest on the bottom of the ship, the bottom structure should be
very heavy. This is in fact so, a very heavy girder being provide by the oor plate in conjunction
with its associated inner and outer bottom plating. The side of the ship is also subjected to
water pressure of rather lesser magnitude, and in this case adequate stiffening is provided by
the girder consisting of the side frame welded to the side shell plating. As far as decks are con-
cerned, here again the beam with its associated deck plating forms an effective built-up girder.
The main factor determining the sizes of the beams is the load which they have to carry. This
load may be a cargo load, a load due to passengers or, in the case of a weather deck some
weather load.
Other items of the structure which contribute to transverse strength are watertight bulk-
heads. Their primary object is, of course, to divide the ship into a series of watertight compart-
ments, but since they consist of transverse sheets of plating they have very considerable
transverse rigidity and hence contribute greatly to the prevention of transverse deformation of
the structure.
The structure shown in Fig. 9 is typical of a transversely framed ship. It is common practice
nowadays to adopt a different form of construction in which the sides of the ship are stiffened
transversely whilst the decks and bottom are stiffened by means of longitudinals. This type of
construction is shown in Fig. 10. As will be shown later, the effect of stiffening the deck and
bottom by longitudinal members instead of transverse members is to increase very greatly the
buckling strength of the plating, and it is largely for this reason that this method of construction
has been adopted.Since these longitudinals are effectively attached to the plating they contrib-
ute also to the general longitudinal strength of the structures. The longitudinals have to carry
cargo and water pressure loads and so, in order to reduce their scantlings, they must be sup-
ported at positions other than at bulkheads. This is achieved by introducing deep transverse
beams in the decks spaced some 6 to 12 ft. apart and by having transverse plate oors in the
Upper deck
beam
Tween deck
pillar
Tween deck
frame
Second deck
beam
Hold
pillar
Floor
plate
Tank side
bracket
Side girder
Centre
girder
Fig. 9 Section through ship showing transverse structure

Airplane and ship structures

14

Thin-Walled Structures

bottom at the same spacing. These widely spaced transverse members, in conjunction with
closely spaced side framing, then provide the transverse strength of the structure.
The longitudinal system of framing has often also been extended to the sides of the ship
as well as the decks and bottom. In fact when initially developed for use in oil tankers this was
the method which was adopted. This was called the Isherwood System. At a later stage in the
development of the tanker the combined system of longitudinals in the bottom and deck with
transverse side framing was employed. In many of the larger oil tankers of the present day,
however, the complete longitudinal framing system has been used. Figure 11 shows the mid-
ship section of such a tanker.
Where transverse beams are employed in the decks of ships it would be impracticable to
Tween deck
frame
Upper deck
longitudinals
2nd deck
longitudinals
Side frame
Deep transverses
spaced 6 to 12 ft.
apart
Inner bottom
longitudinals
Outer bottom
longitudinals
Fig. 10 Section through ship with longitudinally stiffened decks and bottom
Flat bar deck
longitudinals
Side
girder
Centre girder
Wing bulkhead
Flat bar longitudinals
Side girder
Center girder
Flat bar bottom
longitudinals
Transverses spaced
about 10 ft. apart
Flat bar side longitudinals
Fig. 11 Section through large modern oil tanker

Thin-Walled Structures

15

Ships structures

run these from side to side of the ship without intermediate support. It is therefore necessary
to introduce pillars to support the beams. In the early development of the iron and steel ship
these pillars were closely spaced, generally being on alternate beams with a longitudinal angle
runner tted under the beams to spread the load to those beams not supported by pillars. This
meant that access to the sides of cargo holds could only be made between two pillars, so that
the available space was only about 5 ft. The later development was to support the deck beams
by one or more heavy longitudinal girders and to support these girders by means of wide-
spaced pillars. With this arrangement there would be probably two girders in the breadth of the
ship each supported by two pillars in the length of the hold. Such an arrangement is shown in
Fig. 12. By supporting the deck beams with lines of pillars or heavy longitudinals, the scant-
lings of the beams are greatly reduced and, further, by the use of wide-spaced pillars access
to the holds is made easy. When longitudinal stiffening of decks is used, the system of con-
struction just described can be imagined to have been turned around, The longitudinals
replace the beams and the deep transverse beams replace the longitudinal deck girders in the
transversely framed ship.
In addition to their functions in resisting longitudinal and transverse bending, many of the
parts of the structure referred to in this section have also to support local loads. Thus beams
and girders will often be subjected to loads due to machinery and loads produced by lifting
Hatch coaming
Hatch coaming
Pillars
Girder
Girder
B
u
l
k
h
e
a
d
B
u
l
k
h
e
a
d
Pillars
Inner bottom
Outer bottom
Elevation through hold
Girder
Girder
Deck
beams
Pillars
Pillars
Plan at deck
Fig. 12 Wide-spaced pillar and girder arrangement in transversely framed ship

Airplane and ship structures

16

Thin-Walled Structures

equipment such as derricks and the like. The outside plating of the ship has also to withstand
water pressure, and this could produce local bending of the plating between the stiffening
members such as oors and frames. In general it could be said that nearly every structural
member in the ship is a local strength member.
The foregoing discussion has shown briey the functions which the various parts of the
ships structure have to perform. It can be seen that particular part of the structure may have to
perform several functions at the same time. In succeeding chapters methods for determining
the stresses in the various parts will be dealt with in detail.

1.7 Key words and concepts from Chapter 1

structure
elasticity
structural mechanics
stiffness
strength
preliminary design
types of loads
monocoque & semimonocoque
spar, spar caps, spar web
bulkheads, ribs, rings
structural functions of the skin, longitudinals, and frames
idealization of semimonocoque structure
stresses due to bending and torsion
longitudinal, transverse, and local strength of ship structures
transversely framed, longitudinally framed, and Isherwood system of framing of ships
girder, pillar, beam, oor plate
hatch, hatch coaming
buckling strength
scantlings

1.8 References

Anon.,

Aircraft Basic Science

, 1948, First Edition, Northrop Aeronautical Institute, Charles E. Chapel, Chief
Editor, McGraw-Hill Book Company, Inc, p. 59 & 60.
Gordon, J.E., 1978,

Structures: or, Why things dont fall down

, (A Da Capo paperback) Reprint. Originally
published by Harmondsworth: Penguin Books, pp. 33-44.
Hughes, O.F., 1983,

Ship Structural Design

, John Wiley and Sons, New York, N.Y., p. 8.
Muckle, W., 1967,

Strength of Ships' Structures

, E. Arnold Inc., pp. 5-12.
Rivello, R. M., 1969,

Theory and Analysis of Flight Structures

, McGraw-Hill, pp. 143-147.
Sokolnikoff, I.S., 1956,

Mathematical Theory of Elasticity

, Second Edition, McGraw-Hill Book Company,
New York.

Thin-Walled Structures

17

CHAPTER 2

Bars Subjected to Axial
Loads

A bar is a structural member that is relatively long along one axis and relatively compact in cross section in
planes perpendicular to the axis. Bars can be straight or curved. Bars are among the most widely use structural
elements. In this chapter only straight bars are considered that are subjected to loads directed along the reference
axis of the bar. The reference axis is parallel to the long axis of the bar and will be dened in Section 2.4. Axial
loads applied along the reference axis of a straight bar cause extensional and/or compressive deformations. A
slender bar in compression is likely to buckle and in that case the bar is called a column. Buckling results in a
combination of bending and compressive deformations of the column. Loads applied perpendicular to the refer-
ence axis cause the bar to bend, and in that case the bar is called a beam. Beams are the subject of the next chap-
ter.
The three basic steps to analyzed the static response of any structure are discussed for a bar in Section 2.1 to
Section 2.5. These three fundamental steps of static structural mechanics are


equilibrium conditions,


strain-displacement conditions, or conditions of geometric t, and


a material law, or constitutive behavior.
Work and energy methods are presented in Section 2.6 to Section 2.11, which includes the topics of virtual work,
strain energy, complementary virtual work, complementary strain energy, and Castiglianos theorems. Applica-
tions of the energy method to trusses is presented in Section 2.12.

2.1 Axially loaded bar

Consider a straight bar of length

L

, whose cross section is uniform along its length with its cross-sectional area
denoted by

A

. The bar is referred to a Cartesian coordinate system

x

,

y

, and

z

with the

z

-axis parallel to the length
and the

x

and

y

axes in a plane parallel to the cross section. The origin of the

z

-axis is taken at the left end of bar,
so . The bar is subjected to the following loads: a distributed force per unit length of intensity ,
either an axial force or axial displacement at the left end, and to either an axial force or axial dis-
0 z L p
z
z ( )
Q
1
q
1
Q
2

Bars Subjected to Axial Loads

18

Thin-Walled Structures

placement at the right end. The distributed force intensity , forces

Q

1

and

Q

2

, and the corresponding
displacements

q

1

and

q

2

, respectively, are dened positive if they act in the positive in the positive

z

-direction.
See Fig. 2.1. Under the imposed loads, the bar is in tension and/or compression.

Equilibrium

The internal normal force, or axial force, acting in the

z

-direction is denoted by function ,
and

N

is positive if tensile and negative if compressive. See Fig. 2.2. If we consider an interior element of the bar
as shown in the center sketch of Fig. 2.2 and a positive normal force is dened to act in the positive

z

-direction
on a positive

z

-face, then the action-reaction law requires a positive normal force acting on the negative z-face to
act in the negative z-direction. A positive

z

-face of this interior element is the face whose normal pointing away
from the material inside the element is in the positive

z

-direction. Conversely, a negative

z

-face of this interior
element is the face whose normal pointing away from the material inside the element is in the negative

z

-direc-
tion. Force equilibrium in the

z

-direction of the differential interior element of the bar shown in the gure, in the
limit as , gives the following differential equation of equilibrium.

(2.1)

(In Fig. 2.2 note that coordinate

z

* where the distributed load intensity is evaluated on the differential element
approaches

z

in the limit as the element length goes to zero; i.e., in the limit.)
Let the axial displacement function be denoted by

w(z)

. The function

w(z)

is the displacement in the

z

-direc-
tion of a particle located at

z

in the undeformed bar due to the imposed loads as is shown in Fig. 2.3. The axial
q
2
p
z
z ( )
x
y
z w ,
y
L
Q
2
q
2
, Q
1
q
1
,
Fig. 2.1 Axially loaded bar
p
z
z ( )
Cross section
N z ( )
z
L
Q
2
Q
1
dz
N dN +
p
z
z
*
( )dz
N ( )

N
z
N L ( )

z z
*
z dz + < <
0
0
Fig. 2.2 Free body diagrams for equilibrium of the bar
left end right end
dz 0
dN
dz
------- p
z
z ( ) + 0 = 0 z L < <
z
*
z

Thin-Walled Structures

19

The tensile test

displacement of all material points in the cross section is assumed to be the same, but the displacement changes
from cross section to cross section. Hence, the axial displacement is a function of

z

and not of

x

and

y.

The
boundary conditions at are

(2.2)

and the boundary conditions at are

(2.3)

Note that we cannot specify both the displacement and its corresponding force at a point on the body. If we do
not specify the displacement then equilibrium applied separately at the boundaries of the bar (at either z = 0 or z
= L) relates the external force to the internal bar force at the boundary point. See the free body diagrams of the
left and right ends of the bar in Fig. 2.2.

Strain-displacement

Extensional deformation of an element of the bar is shown in Fig. 2.3. The axial normal
strain due to extension is given by

(2.4)

The axial normal strain


z

is a function of

z,

and it is dimensionless it is the change in length divided by the
original length. If the axial displacement function

w

is a constant value for 0


L, then the bar displaces as a
rigid body and the normal strain is zero, as is evident from eq. (2.4).
To complete the analysis for the response of the bar, we need to relate the internal axial force

N

to the axial
normal strain


z

, and this requires a material law. The material law is obtained from material characterization
tests of carefully designed specimens as discussed in the following section.

2.2 The tensile test

Material characterization involves devising experiments for a very simple loading situation from which a mate-
rial law can be developed to predict behavior in general loading situations. The tensile test is an important mate-
rial characterization test in which a slender member is pulled parallel to its axis. Metallic tensile test specimens
are usually circular section bars that are designed to achieve as nearly as possible a uniform state of axial normal
stress and strain in portion of the specimen called the gage length. The test is usually conducted in a universal
load frame and the with an attached load cell to measure the axial force applied to the specimen and an exten-
someter, or electrical resistance strain gage, to measure the elongation/strain of the gage length. The load frame
z 0 =
either w 0 ( ) q
1
or N 0 ( ) Q
1
but not both = =
z L =
either w L ( ) q
2
or N L ( ) Q
2
but not both = =
z
dz
w z ( )
w dw +
dz dw +
z - axis
y
z dz + z w z ( ) + z dz w dw + + +
Fig. 2.3 Extensional deformation of an element of the bar

z
Limit
dz dw + ( ) dz
dz
------------------------------------
dw
dz
------- = =
dz 0
Bars Subjected to Axial Loads
20 Thin-Walled Structures
may be servo-hydraulically controlled using a feedback system to maintain specied load magnitude or elonga-
tion as a function of time. The loading rate is slow in static testing so that inertia effects are negligible, and usu-
ally the load is increased monotonically in time until fracture of the specimen. In addition, servo-hydraulic load
frames can apply cyclic loading at frequencies of around10Hz, which is used for fatigue testing. Electrical sig-
nals from the load cell and strain transducers are conditioned and converted to digital form. The digital response
data can be stored and then plotted using personal computers and laboratory software. The American Society of
Testing Methods (ASTM) publishes standards for testing materials. The standard governing the tensile test of
ductile metals is ASTM E8 Standard Test Methods for Tension Testing of Metallic Materials.
The engineering stress in the gage section is dened as the axial force N divided by the original cross-
sectional area A of the specimen. The engineering normal strain is dened as the elongation of the mate-
rial in the gage length divided by the original gage length L, and for uniform extension in the gage section it is the
same as the point-wise denition given in eq. (2.4). Typical engineering stress-strain plots for a low carbon steel
and an aluminum alloy are shown in Fig. 2.4. For most engineering materials there is a linear relationship
between the axial normal stress and normal strain near the origin of the plot as is shown in the gure. The
slope of the linear portion is a material property called the modulus of elasticity, and is denoted by E. The value
of the stress where the stress-strain plot deviates from a straight line is called the proportional limit, and is
denoted by . The proportional limit is difcult to measure since under test conditions the deviation from a
straight line is subject to some judgement. The deformation of the material in the linear region is usually elastic.
Elastic deformation is dened as the deformation that disappears on removal of the load. The largest stress for
which elastic deformation occurs is called the elastic limit. The elastic limit of a material is also difcult to mea-
sure precisely since the specimen must be unloaded and reloaded to determine it. Just after the linear region of
the stress-strain plot for some low carbon steels, there is a relative maximum stress followed by a relative mini-
mum, and the deformation, or strain, begins to increase rapidly for small changes in the load. (To observe this
behavior, the elongation or displacement of the specimen is controlled and the load is measured.) The relative

z
L

yu

z
0

z
fracture
E
1

z
0

z
fracture

0.2
0.002 =
(a) some steels (b) aluminum alloy
L 1
z
+ ( )
N
N
L
A

z
N
A
---- =

yl
f

f
, ( )

f

f
, ( )
Fig. 2.4 Typical stress-strain curves for steel and aluminum alloy from tensile tests

z

z

p
Thin-Walled Structures 21
The tensile test
maximum value of the stress is called the upper yield point , and the stress at the subsequent relative mini-
mum value is called the lower yield point . Values of the upper yield point for metals are sensitive to the load-
ing rate and accidental bending stresses. Yielding is associated with plastic deformation of the material. Plastic
deformation is dened as deformation which is independent of time and remains on release of the load. The prin-
cipal physical mechanism causing plastic deformation in metals is slippage between planes of atoms in the crys-
tal grains of the material (Dowling, 1993).
Loading, unloading, and reloading of a metallic specimen beyond its yield stress is depicted in Fig. 2.5. The
unloading slope is nearly the same as the slope of the linear elastic portion, or E. The total strain at load is the
sum of the plastic strain (permanent portion upon removal of the load) and the elastic strain (recoverable por-
tion). Since plastic deformation of the material results in a change in size and shape of the structural component,
it is undesirable in design. Hence, yielding of the material is an important phenomena to quantify.
Under service loads, it is required that a structural component not be stressed beyond the yield stress. In
design, the condition of no yielding under service loads is called a limit state.
Aluminum alloys do not exhibit the abrupt yield point of low carbon
steels; i.e., there is no stress just after the linear elastic portion where the
stress-strain curve has a zero slope. See Fig. 2.4b. Instead, following the lin-
ear elastic region, the slope of the stress-strain curve continuously decreases
until a relative maximum engineering stress occurs deep into the response
regime where plastic deformation is dominate. For such material behavior we
dene an offset yield stress. A straight line is drawn parallel to the linear elas-
tic portion of the stress-strain curve starting from a strain
on the strain axis. The stress at the intersection of this straight line with the
stress-strain curve is dened to be the yield strength of the material.
Note that the strain of 0.002, or 0.2% (percent strain is dened as ), is
plastic strain, since unloading the specimen from the point on
the stress strain-curve would follow the straight dashed line in Fig. 2.6 and the strain of 0.002 would not be
recovered. However, a permanent strain of 0.2% is not considered detrimental for most structural components,
and the 0.2% offset yield strength has the advantage of being a precisely dened quantity. The offset yield stress
is generally the most satisfactory means of dening the yielding event for engineering materials.

yu

yl

z 0
A B

0B
total strain =

0A
plastic strain =

AB
elastic strain =

0B

0A

AB
+ =
Fig. 2.5 Loading and unloading a tensile test specimen beyond the yield stress
E
1
E
1

yield
0
0.002

yield

z
Fig. 2.6 0.2% offset
yield strength

0.2
0.002 = =

yield
100

yield

yield
, ( )
Bars Subjected to Axial Loads
22 Thin-Walled Structures
After yielding the load may have to increase to cause further plastic defor-
mation of the specimen. (Also, the elastic deformation increases.) The increase
in load required for further plastic deformation after yielding is called strain
hardening. The maximum tensile load carried by the specimen occurs during
strain hardening, and this maximum load divided by the original cross-sectional
area is called the ultimate tensile strength, or just the tensile strength, of the
material and is denoted by . At the maximum load the deformation of most
metal specimens becomes localized in the form of an abrupt reduction in cross section along a small length in the
gage section. Prior to the maximum load the deformation is spatially uniform in the gage section. Plastic defor-
mation becomes concentrated in the reduced cross section after the maximum load. The non-uniform deforma-
tion is called necking, and its location in the gage section depends on imperfections in the particular specimen
which are random in nature. The load decreases after necking commences. The tensile test reaches its conclusion
when a small crack develops at the center of the neck and spreads outward to complete the fracture. The engi-
neering stress at fracture is denoted by and the engineering strain at fracture is denoted by . Note that the
engineering stress is dened with respect to the original cross-sectional area.
In addition to the axial elongation of the bar in the tensile test, most engi-
neering materials exhibit a lateral contraction of the cross section. In axial
compression, the cross section expands. In the linear elastic range of mate-
rial behavior, the ratio of the lateral normal strain to the axial normal strain
is found to be a constant called Poissons ratio. Poissons ratio is denoted
by , and it is a dimensionless quantity. See Fig. 2.7. It has the same value
in tension and compression. The x and y axes of the Cartesian system
(x,y,z) lie in the cross section, so the lateral normal strains are denoted by
and . That is, measurements of the diameter of the test specimen in
the tensile test reveal that the normal strain in the x- and y-directions are
The negative sign is introduced to make the Poissons ratio positive for lateral contraction under uniaxial exten-
sion. It is implicit in the above expression that the Poissons ratio is the same along the x- and y-directions. If a
material is isotropic, then the Poissons ratio is necessarily the same in the x- and y-directions. An isotropic mate-
rial is one for which the material properties are independent of direction. In general, a material in which the prop-
erties vary with direction is called anisotropic. Materials whose properties vary from point to point are said to be
heterogeneous, and a material whose properties are uniform from point to point are said to be homogeneous.
For the linear elastic range of material behavior, the material law relating the axial stress to the three nor-
mal strains is
(2.5)
N N
necking deformation

f

f

x
or
y

1
0
Fig. 2.7 Lateral normal
strains in the tension test

z
E
----- =

x

y

x

y

z
= =

x

z
E =
y

z
E =
z

z
E = (uniaxial loading only)
Thin-Walled Structures 23
Effect of temperature on strain
Typical values of the modulus of elasticity and Poissons ratio are listed in the table below for selected metals.
2.3 Effect of temperature on strain
In the elastic region of material behavior changes in temperature can cause two effects:
The elastic constants (E and ) of the material can change with temperature.
The temperature change causes the material to strain in the absence of stress.
For many structural materials a change in temperature of a few hundred degrees Fahrenheit does not result in
much change in the elastic constants. We will neglect the effect of temperature on the elastic constants in this text
and consider only the second effect. The strain caused by a temperature change in the absence of stress is called
thermal strain, and thermal strain is denoted by . For an isotropic material, symmetry arguments show that a
rectangular parallelepiped of material deforms into rectangular parallelepiped of larger dimensions as the tem-
perature increases. That is, the thermal strain must be a pure expansion or contraction of the material with no dis-
tortion or shear. For the axially loaded bar, the important thermal strain component is the axial component .
Although the thermal strain is not exactly a linear function of the temperature change, for temperature changes of
a few hundred degrees Fahrenheit the actual thermal strain is nearly linear with the change in temperature. If the
temperature of the material is changed from to , then the thermal strain is approximated by
(2.6)
in which is called the linear coefcient of thermal expansion. The coefcient of thermal expansion is
abbreviated as CTE. The dimensional units of are or , since strain is a dimensionless quantity. Val-
ues of the linear coefcient of thermal expansion for selected metals and metal alloys are listed in the table
below. The total strain in the bar is the sum of the strains due to mechanical stress and thermal strain. That is,
(2.7)
Room temperature modulus of elasticity and Poissons ratios for selected metals and metal
alloys (from Callister, 1997)
Material
Modulus of elasticity
Poissons ratio,
dimensionless 10
6
psi GPa
Aluminum and aluminum alloys 10 69 0.33
Copper 16 110 0.34
Steels, plain carbon 30 207 0.30
Titanium and titanium alloys 15-16.8 104-116 0.34

z
t
T
0
T

z
t
T T
0
( ) =
1 F
0
1 C
0

z
E
----- T T
0
( ) + =
{
||||
mechanical thermal
Bars Subjected to Axial Loads
24 Thin-Walled Structures
2.4 Bar reference axis
Consider the bar to be made from a homogeneous, isotropic material. Under the assumption of uniform exten-
sional deformation of the bar over its cross section, the distribution of the axial normal stress is uniform over
the cross section.With respect to an arbitrary coordinate system , in the cross section, this uniform normal
stress distribution is statically equivalent to an axial force and a moment with components and as
shown in Fig. 2.8.
The relationships of static equivalence are
(2.8)
Room temperature linear coefcients of thermal expansion (from Callister, 1997)
Material
Coefcients of thermal expansion
Aluminum 13.1 23.6
Aluminum alloys (cast) 10.0-13.6 18.0-24.5
Aluminum alloys (wrought) 10.8-13.4 19.5-24.2
Copper 9.4 17.0
Steel (low alloy) 6.2-7.1 11.1-12.8
Steel (plain carbon) 6.1-6.7 11.0-12.0
Titanium and titanium alloys 4.2-5.4 7.6-9.8
10
6
F
0
10
6
C
0

z
x y
N M
x
M
y
x
y
x
y

z
dA
M
x
M
y
N
N
x
c
x
y
c
y

C
statically equivalent
Fig. 2.8 Statical equivalence of the normal stress to the bar resultants.
N
z
A d
A

= M
x
y
z
A d
A

= M
y
x
z
A d
A

=
Thin-Walled Structures 25
Bar reference axis
Substitute the material law for the axial normal stress from eq. (2.7) into these conditions of statical equiva-
lence to get
Since the bar is homogenous, the material properties are independent of the cross-sectional coordinates. Also, the
extensional strain is independent of the cross-sectional coordinates, and the change in temperature is uniform
over the cross section, in pure extensional deformation. Hence, the previous equations become
(2.9)
where denotes the rst moment of the cross-sectional area about the -axis and denotes the rst area
moment about the -axis. These rst area moments are dened by the integrals
(2.10)
For pure extension of the bar with no bending, the distribution of the axial normal stress must be equivalent to an
axial force and no bending moments. The bending moments can be made to vanish by moving the resultants to a
special point in the cross section. The coordinates of the point in the cross section where the moments
due to a uniform distribution of axial normal strain vanish is called the centroid. Static equivalence yields the
resultants at the centroid as
(2.11)
Solving the last two of eqs. (2.11) for the coordinates of the centroid, and using eqs. (2.9) we get
(2.12)
The location of the centroid depends on distribution of the area of the section about the - coordinate system.
The origin of the parallel Cartesian system x, y is at the centroid of the cross section as is shown in Fig. 2.8. The
two coordinate systems are related by
(2.13)
The rst area moments have dimensional units of L
3
. Equations (2.12) establish the location of where the z-
axis passes through the cross section. That is, the reference axis of the bar, or the z-axis, passes through the cen-
troid of each cross section. If the bar is uniform along its length, then the reference axis is a straight line. Also,
eqs. (2.12) imply the rst moments of the cross-sectional area about the x and y axis system vanish; i.e.,
N E
z
T T
0
( ) ( ) [ ] A d
A

=
M
x
y E
z
T T
0
( ) ( ) [ ] A d
A

= M
y
x E
z
T T
0
( ) ( ) [ ] A d
A

=

z
N EA
z
T T
0
( ) [ ] =
M
x
Q
x
E
z
T T
0
( ) ( ) [ ] = M
y
Q
y
E
z
T T
0
( ) ( ) [ ] =
Q
x
x Q
y
y
Q
x
y A d
A

= Q
y
x A d
A

=
x
c
y
c
, ( )
N N =
M
x
M
x
y
c
N 0 = =
M
y
M
y
x
c
N 0 = =
x
c
Q
y
A = y
c
Q
x
A =
x y
x x
c
x + = y y
c
y + =
Q
x
y A d
A

0 = = Q
y
x A d
A

0 = =
Bars Subjected to Axial Loads
26 Thin-Walled Structures
From the rst of eqs. (2.9) and the rst of eqs. (2.11), the material law for the axial force is given by
(2.14)
where the product the modulus of elasticity and the cross-sectional are (EA) is called the axial stiffness , or exten-
sinal stiffness, of the bar, and denotes the so-called thermal force. The thermal axial force is dened by
(2.15)
Note that eq. (2.14) shows that the magnitude of the thermal force is equal to the magnitude of the actual internal
bar force only if the strain vanishes. For example, if the bar is constrained from changing its length
( ), the distributed load intensity vanishes ( ), and there is a change in temperature from
the stress free state, then .
2.5 Linear elastic response
As stated at the beginning of this chapter, there are three basic steps to determine the static response of a structure
to specied loads: equilibrium conditions, strain-displacement relations, and material laws. For the uniaxially
loaded bar problem, we assume the it is made of a linear elastic, isotropic, homogeneous material, and summa-
rize these fundamental equations as follows.
equilibrium, eq. (2.1):
strain-displacement, eq. (2.4):
material law, eq. (2.14):
Further lets assume the temperature is independent of the axial coordinate z as well as being uniform over
the cross section; i.e., the bar is subjected to a spatially uniform change in temperature. Substitute eq. (2.4) for
the strain in eq. (2.14), and in turn substitute the resulting equation into the equilibrium eq. (2.1) for the axial
force to get
(2.16)
where we used for a spatially uniform change in temperature. If the axial stiffness EA is indepen-
dent of z, that is, the bar is uniform along its length, then eq. (2.16) becomes
(2.17)
Equation (2.16) is the governing ordinary differential equation for the displacement function w(z), 0 z L,
and is of second order. For a second order differential equation we need two boundary conditions to determine
the two constants that arise in the general solution of eq. (2.16). These boundary conditions are given in eqs. (2.2)
and (2.3). Using the material law and the strain displacement equations these boundary conditions become
N EA
z

z
t
( ) EA
z
N
t
= =
N
t
N
t
EA
z
t
EA T T
0
( ) = =
q
1
q
2
0 = = p
z
z ( ) 0 =
N N
t
=
dN
dz
------- p
z
z ( ) + 0 = 0 z L < <

z
dw
dz
------- =
N EA
z
N
t
=
z d
d
EA
z d
dw
\ )
[
p
z
z ( ) = 0 z L < <
dN
t
dz 0 =
EA
z
2
2
d
d w
p
z
z ( ) = 0 z L < <
Thin-Walled Structures 27
Linear elastic response
(2.18)
(2.19)
EXAMPLE 2.1 Axial bar with a specied uniform distributed load and specied end displacements
Consider a linear elastic, uniform bar subjected to a uniformly dis-
tributed load with intensity , specied end displacements q
1

and q
2
, and a uniform change in temperature T - T
0
from the stress free
state as shown in Fig. 2.9.Determine the axial displacement, normal
strain, and normal force of the bar.
Solution The governing differential equation (2.17) for the displace-
ment reduces to
Integrate this equation twice with respect to z to get
where c
1
and c
2
are constants obtained in the indenite integration. Substitute this expression into the boundary
condition at the left end and right end to get
Solve these last two equation for constants c
1
and c
2
to nd
Substitute these results for the constants of integration in the solution for the axial displacement to get
(2.20)
The axial strain in the bar is dened by eq. (2.4), and substituting eq. (2.20) for w(z) we get
(2.21)
Finally, substitute the strain given by eq. (2.21) into eq. (2.14) to nd the distribution of the axial force as
(2.22)
either w 0 ( ) q
1
or EA
dw
dz
-------
z 0 =
N
t
Q
1
but not both = =
either w L ( ) q
2
or EA
dw
dz
-------
z L =
N
t
Q
2
but not both = =
p
0
z w ,
L
w 0 ( ) q
1
= w L ( ) q
2
=
T T
0
0
Fig. 2.9 Bar of Example 2.1
p
z
p
0
=
z
2
2
d
d w
p
0
EA
------- = 0 z L < <
w z ( )
p
0
EA
-------
z
2
2
----
\ )
[
c
1
z c
2
+ + =
p
0
EA
-------
0
2
2
-----
\ )
[
c
1
0 c
2
+ + q
1
=
p
0
EA
-------
L
2
2
-----
\ )
[
c
1
L c
2
+ + q
2
=
c
1
q
2
q
1

L
----------------
p
0
L
2EA
----------- + = c
2
q
1
=
w z ( ) q
1
1
z
L
---
\ )
[
q
2
z
L
---
p
0
2EA
----------- z
2
Lz ( ) + =

z
q
2
q
1
( )
L
---------------------
p
0
2EA
----------- 2z L ( ) =
N z ( )
EA
L
------- q
2
q
1
( ) p
0
z
L
2
---
\ )
[
N
t
=
Bars Subjected to Axial Loads
28 Thin-Walled Structures
where . Equations (2.20) to (2.22) constitute the complete solution for the linear thermoelas-
tic response of the uniform bar subject to a uniform distributed load intensity p
0
, uniform temperature change
, and end displacement q
1
and q
2
.
We can use eq. (2.22) and the relations and (refer to eqs. (2.2) and (2.3)), to get
(2.23)
(2.24)
Addition of these results gives
which is the condition of overall axial force balance for the bar.
EXAMPLE 2.2 A bar with xed ends and subjected to an axial point force.
Determine the axial displacement and normal force for the uniform bar
subjected to a point force F as shown in Fig. 2.10. Each end of the bar is
xed to a rigid support. There is no change in temperature from the stress
free state, and the specied data are, the axial stiffness EA, dimensions a, b,
and L, and F.
Solution This is a statically indeterminate problem so all three steps (equi-
librium, strain-displacement, and a material law) are needed to solve for the
response of the bar. The governing differential equation (2.17) is valid in
the open intervals 0 < z < a and a < z < L, but is not valid at the location of the point force since a point force is
mathematically equivalent to an innite value of the distributed load intensity p
z
acting over a zero length.
Hence, we will need to solve the governing differential equation separately in each open interval. That is
(2.25)
Two constants of integration will occur in the solution of this differential equation in each open interval for a total
of four unknown constants. Thus, we need four boundary conditions. Since the ends of the bar are xed, we have
(2.26)
The displacement of the bar must be continuous at z = a, so
(2.27)
where implies the limit of for values of , and implies the limit of for values of .
The last boundary condition comes from axial force equilibrium at the point of application of the force.
(2.28)
N
t
EA T T
0
( ) =
T T
0

N 0 ( ) Q
1
= N L ( ) Q
2
=
Q
1
EA
L
------- q
2
q
1
( )
p
0
L
2
--------- N
t
+ =
Q
2
EA
L
------- q
2
q
1
( )
p
0
L
2
--------- N
t
=
Q
1
Q
2
p
0
L + + 0 =
z
a b
L
F
Fig. 2.10 Point force acting on
a clamped bar
EA
z
2
2
d
d w
0 = 0 z a < < and a z L < <
w 0 ( ) 0 = w L ( ) 0 =
w a

( ) w a
+
( ) =
a

z a z a < a
+
z a z a >
N a

( ) N a
+
( ) F + + 0 = or EA
dw
dz
-------
z a

EA
dw
dz
-------
z a
+

F + + 0 =
Thin-Walled Structures 29
Linear elastic response
The solution to eqs. (2.25) is
where c
1
, c
2
, c
3
, and c
4
are constants to be determined from the four boundary conditions. From the two bound-
ary conditions given by eqs. (2.26), we nd and . Using these results in the continuity condi-
tion given by eq. (2.27) we get
and in the force balance given by eq. (2.28) we get
These last two equations can be solved for the constants c
1
and c
3
, and the solution is
, where we used b = L - a. Hence the displacement function is
(2.29)
The normal force in each interval is obtained from , so we have
(2.30)
The displacement and normal force are plotted with respect to the axial coordinate in Fig. 2.11. Note that the dis-
placement is continuous in z, and that the normal force has a jump discontinuity at z = a. The value of the discon-
w z ( ) c
1
z c
2
+ = 0 z a < <
w z ( ) c
3
z c
4
+ = a z L < <
c
2
0 = c
4
Lc
3
=
c
1
a c
3
a L ( ) =
EAc
1
EAc
3
F + + 0 =
c
1
bF ( ) EAL ( ) =
c
3
aF ( ) EAL ( ) =
w z ( )
bF
EAL
-----------z 0 z a
aF
EAL
----------- L z ( ) a z L
|
|

|
|
=
N EA dw dz ( ) =
N z ( )
b
L
---F 0 z a
a
L
---F a z L
|
|

|
|
=
w
z
z
0
0
N
a
a
L
L
abF
EAL
-----------
b
L
---F
a
L
---F
Fig. 2.11 Displacement and normal force
distributions for the clamped bar
subjected to a point force.
Bars Subjected to Axial Loads
30 Thin-Walled Structures
tinuity in the normal force is , which satises the rst of eqs. (2.28).
2.6 Work and energy methods
In particle mechanics the incremental work of a force acting on a particle is the product of the force and the com-
ponent of the incremental displacement of the particle in the direction of the force. If we have a system of parti-
cles, then the incremental work of the forces acting on their respective particles can be dened as the sum the
incremental work of the forces acting on their respective particles. For the problems of solid mechanics in which
we assume the body is made of a continuum of particles, each innitesimal particle, or material point, is identi-
ed by its coordinates in the undeformed or reference state of the body. The displacement components of the par-
ticle are continuous, single valued functions of its coordinates in the reference state so that the body in the
deformed conguration remains a continuum; i.e., the deformation is such that no holes or overlapping of mate-
rial points occur in the deformed conguration.
Hence, to dene incremental work in continuum structures we need to account for small variations in the
displacement variables which are themselves continuous functions of the coordinates. In the mathematical sense,
a bar is called a one-dimensional structure, since the dependent variables describing the displacement, strain, and
the internal normal force are functions of a single independent variable z, 0 z L, where L is the length of the
bar. The coordinate z locates a material point in the undeformed bar.
2.6.1 Concept of virtual displacement
Assume the bar is in equilibrium under external loads , and a temperature change . The
displacement , the strain , and the axial force are those of this equilibrium state. Now give the
end displacements an innitesimal change in magnitude, denoted by . These innitesimal changes in
the displacements are called virtual displacements. The word virtual means in essence and not in fact. That is, the
virtual displacements are possible displacements, but do not necessarily coincide with the actual displacements
of the bar. Since the bar is assumed to be a continuum, the displacements of the material points along the bar will
also change, and the change in axial displacement is a function denoted by . The virtual displacement
function must be a continuous, single valued function of z for the bar not to fracture, and its rst deriva-
tive, or virtual strain , must be piecewise continuous so that its integral is computable. Also, the virtual
displacement function must satisfy the end conditions and .
A displacement function is said to be kinematically admissible if it satises continuity conditions and speci-
ed boundary conditions on the displacement, if any. A kinematically admissible function may not satisfy the
governing boundary value problem for the response of the bar. However, within the set of all kinematically
admissible displacement functions one will satisfy the governing differential equation of equilibrium, eq. (2.16),
and the specied boundary conditions on the axial force, if any, which are given in eqs. (2.2) and (2.3). That is,
the exact displacement function of the equilibrium state is in the set of kinematically admissible displacement
functions.
The virtual displacement function is a new innitesimal function, and to emphasize this fact it is also repre-
sented as
N a
+
( ) N a

( )
a
L
---F
\ )
[
b
L
---F
\ )
[
F = =
Q
1
Q
2
p
z
z ( ) , , T T
0
( )
w z ( )
z
z ( ) N z ( )
q
1
and q
2
w z ( )
w z ( )

z
z ( )
w 0 ( ) q
1
= w L ( ) q
2
=
Thin-Walled Structures 31
Work and energy methods
(2.31)
where is an innitesimal scalar parameter and the function is an arbitrary, but kinematically admissible,
function. The varied displacement function is dened as
(2.32)
and the varied displacement equals the displacement
of the equilibrium state of the deformed bar as
. This displacement situation is depicted in
Fig. 2.12. The particle at position P having the coor-
dinate z in the undeformed bar is displaced to posi-
tion having the coordinate in the
equilibrium conguration of the deformed bar.
Imposition of the innitesimal virtual displacement
from the equilibrium state of the deformed
bar causes the particle to move to the position
having the coordinate . The point of all this is that the virtual displacement is a change in the displace-
ment following a particle, and the particle is identied by its position z in the undeformed bar. That is, from the
mathematical viewpoint, the virtual displacement is an innitesimal change in the displacement holding the inde-
pendent variable z xed in value. In mechanics, incremental work is dened as product of the force in the direc-
tion of the displacement times an innitesimal change in the displacement of the particle on which the force acts.
Hence, for a continuum, like our bar example, we have to introduce the concept of an innitesimal change in dis-
placement holding the independent variable xed in value in order to compute the incremental work of a force
acting on a particle.
Although the virtual displacement is a new displacement function, the use of the prex greek letter
is meant to convey a change in the function in analogy to the use of letter d used in calculus. However,
the greek letter "" means an innitesimal change in the displacement for a xed value of z, where the symbol
"d" in calculus means the change in a function with respect to a change in its independent variable. The differ-
ence between the variation of a function and the differential of a function is depicted in Fig. 2.13. Since in the
ordinary calculus the differential of the function w(z) is , the physical interpretation of
dw is the difference in the displacements of two particles, one particle originally at z + dz and the second one
w z ( ) z ( ) =
z ( )
w

z ( ) w z ( ) z ( ) + =
y
0
z
w z ( )
w

z ( )
w z ( )
P P
*
P

Fig. 2.12 Displacements of a particle originally a


point P.
0
P
*
z w z ( ) +
w z ( )
P

z w

z ( ) +
w z ( )
w z ( )
w

z ( )
w z ( )
w z ( )
dw
z z dz +
dz
q
1
0 L
q
2
q
1
q
1
+
q
2
q
2
+
z
Fig. 2.13 The variation of a function and the differential of a function
dw w z dz + ( ) w z ( ) =
Bars Subjected to Axial Loads
32 Thin-Walled Structures
originally at z in the undeformed body. The physical interpretation of the innitesimal function w(z) is change in
the displacement of a single particle originally a z in the undeformed body. The "" is considered an operator in a
sense similar to the differential operator "d" in calculus. The "" operator means the variation in a function and
not the differential of a function with respect to a change in its independent variable. The distinction between the
variation of a function and the differential of a function is essential in formulating the concept of work in a con-
tinuous system of particles, and the mathematics dealing with the variation of functions is called the calculus of
variations. That is, the calculus of variations is a mathematical tool by which work and energy methods are
applied to systems with innitely many degrees of freedom. A more complete description of the calculus of vari-
ations applied to structural problems is presented in the text by Langhaar (1962), and Courant and Hilbert (1953)
present the essential mathematical framework. However, the reader may be interested in the recent text book by
Dym (1997) which presents energy methods for continnum models with limited use of the calculus of variations.
We will encounter the variation of the derivative of a function and the variation of the denite integral of a
function in the discussions that follow. These are dened as (refer to Fig. 2.13)
(2.33)
Substitute the for the total displacement the relation into these denitions to get
(2.34)
These relations show that the variational operator and the derivative, and the variational operator and the integral
are interchangeable.
2.6.2 Principle virtual work
The incremental work of the external forces acting through the virtual displacements is denoted by , and
this quantity is called the external virtual work. Internal and external forces are held constant during the (inni-
tesimal) virtual displacements, so that from denition of incremental work in mechanics, we write the external
virtual work as
(2.35)
Since it is assumed that the bar is in an equilibrium state prior to the application of the virtual displacements, we
have from the boundary conditions on the axial force, see eqs. (2.2) and (2.3), that
Hence, eq. (2.35) can be written as
Differentiating the rst term in the integral on the right hand side of this equation, and using the interchange of

dw
dz
-------
\ )
[
dw

dz
-------
dw
dz
------- w z d
0
L

\ )
| j
[
w

z d
0
L

w z d
0
L

z ( ) w

z ( ) w z ( ) w z ( ) + =

dw
dz
-------
\ )
[
d
dz
----- w ( ) = w z d
0
L

\ )
| j
[
w z d
0
L

=
W
ext
W
ext
Q
1
q
1
Q
2
q
2
p
z
z ( )w z ( ) z d
0
L

+ + =
Q
1
q
1
Q
2
q
2
+ N 0 ( )w 0 ( ) N L ( )w L ( ) +
z d
d
Nw ( ) z d
0
L

= =
W
ext
z d
d
Nw ( ) p
z
z ( )w z ( ) + z d
0
L

=
Thin-Walled Structures 33
Work and energy methods
the derivative and variation, as shown by the rst of eqs. (2.34), we get
The rst term in the integral on the right-hand side of this equation vanishes via the equilibrium condition given
in eq. (2.1), and the second term contains the variation in the strain as dened in eq. (2.4) Hence,
(2.36)
Since the integral on the right-hand side of this equation is determined by the internal bar force and the incremen-
tal strain, it is dened as the internal virtual work . The internal virtual work is dened as
(2.37)
What the manipulations from eq. (2.35) to (2.37) show is that for a bar in equilibrium, the external virtual work is
equal to the internal virtual work for a kinematically admissible variation in the displacement function .
This proves the necessary condition that if the body is in equilibrium then the external virtual work equals the
internal virtual work for a kinematically admissible variation in the displacement. In problem solving we assume
that equating the external to the internal virtual work for every kinematically admissible displacement is suf-
cient for equilibrium of the body. We state this principle as follows.
For the axially loaded bar, the principle of virtual work states that the bar is in equilibrium if
(2.38)
where the external virtual work is given by eq. (2.35) and the internal virtual work is given by eq. (2.37). The
principle of virtual work is an integral form of the equations of equilibrium for the bar. The mathematical condi-
tions of kinematic admissibility of the virtual displacement function w(z) are that it is continuous, its rst deriv-
ative is piecewise continuous, and that it vanishes at points where the displacement w(z) is prescribed. That is, the
virtual displacement, or variation in the displacement, must be a possible displacement of the bar.
If no restrictions are placed on the virtual displacement function w(z) other than kinematic admissibility,
then the PVW leads to the exact differential equation of equilibrium for the bar, eq. (2.1), and the associated
boundary conditions, eqs. (2.2) and (2.3). Moreover, it will be shown in Example 2.5 that the PVW for a linear
elastic bar gives the exact solution if the assumed displacement function has the exibility to represent the exact
displacement. Hence, it may seem that the PVW yields nothing new. However, in complex structures, an exact,
closed form, mathematical solution is rarely known. We must resort to approximate solutions, and the PVW is a
powerful technique to develop approximate solutions based on assumed displacement functions containing
unknown parameters. In fact, the PVW is the basis of many nite element procedures used in structural mechan-
ics. The use of the PVW to obtain an approximate solution is illustrated in the following example.
Principle of Virtual Work (PVW)
If the external virtual work (W
ext
) equals the internal virtual work (W
int
) for every
kinematically admissible variation in the displacement, then the body is in equilibrium.
W
ext
z d
dN
p
z
+
\ )
[
w N
z d
dw
\ )
[
+ z d
0
L

=
W
ext
N
z
z d
0
L

=
W
int
W
int
N
z
z d
0
L

w z ( )
W
ext
W
int
= for every kinematically admissible w z ( )
Bars Subjected to Axial Loads
34 Thin-Walled Structures
EXAMPLE 2.3 Approximating the response of a bar using PVW.
Consider the bar of Example 2.2 again, and take length a equal to b so that a = b = L/2; that is, the point force F
acts at the center of the bar. Assume the axial displacement of the bar of the form , in which
the unknown parameter q
1
represents the axial displacement at z = L/2.
a) Show that the assumed displacement function is kinematically admissible.
b) Use the principle of virtual work to determine the equilibrium equation associated with the assumed dis-
placement function.
c) Assume the material is linear elastic and solve for the unknown parameter q
1
.
d) Calculate the percentage error in the maximum displacement and maximum stress with respect to the
exact solution.
Solution (a) To be kinematically admissible the displacement function must be continuous in the interval 0 z
L, and satisfy prescribed boundary conditions. The sine function is continuous in the interval, and its derivative
with respect to z, or the strain, is also continuous. The prescribed boundary conditions on the axial displacement
are w(0) = 0 and w(L) = 0. Since the function is zero at z = 0 and z = L, the prescribed displacement
boundary conditions are satised. Thus, the assumed displacement is kinematically admissible.
(b) The variation in the assumed displacement, or virtual displacement, is , where
q
1
is the variation in the unknown parameter q
1
. The only external force acting on the bar that performs work is
the point force F, so the external virtual work is
The internal virtual work is given by eq. (2.37), which requires the virtual strain. The virtual strain is merely the
derivative with respect to z of the virtual displacement function; i.e., . Hence
the internal virtual work is
Now equate the external virtual work to the internal virtual work, and rearrange the result to get
The PVW gives the equilibrium equation as
(2.39)
Note that at this point we cannot carry out the integration in this last equation since we do not know how the axial
force depends on the displacement. The relationship that is missing is the material law; i.e., the PVW gives the
w z ( ) q
1

z
L
---
\ )
[
sin =
z L ( ) sin
w z ( ) q
1
z L ( ) sin =
W
ext
Fw L 2 ( ) Fq
1
= =

z
d
dz
----- w ( ) q
1

L
---
z
L
---
\ )
[
cos = =
W
int
N
z
z d
0
L

q
1

L
--- N
z
L
---
\ )
[
cos z d
0
L

= =

L
--- N
z
L
---
\ )
[
cos z d
0
L

F q
1
0 = q
1

L
--- N
z
L
---
\ )
[
cos z d
0
L

F 0 =
Thin-Walled Structures 35
Work and energy methods
equilibrium condition related to the kinematically admissible displacement independent of the material law.
(c) For a linear elastic material and no change in temperature from the stress free state, the material law is
and the strain-displacement relation is . Substitute the approximate displacement into
the strain-displacement relation, and then substitute this result into the material law to get
Substitute this expression for the axial force into the equilibrium condition, eq. (2.39), and rearrange it to get
Performing the integration on the left-hand side of this expression we have
Evaluating the limits in this equation we nd
We have determined the unknown parameter q
1
in the assumed displacement function terms of the applied force,
length of the bar, and the extensional stiffness of the bar. Finally, the displacement approximation is
(2.40)
The normal force is obtained from , so it is approximated as
(2.41)
(d) For a = b = L/2, the maximum displacement of the exact solution, eq. (2.29), occurs at the center of the
bar and is . The maximum displacement in the approximate solution, eq. (2.40), also
occurs at the center of the bar and is . The percentage error in the approximate solu-
tion is given by
So the approximate displacement at the center of the bar is 18.9% less than the exact value. The fact that the
approximate displacement is less than the exact displacement means our assumed displacement function is too
restrictive and results in a stiffer response than what the exact solution gives.
N EA
z
=
z
dw dz =
N EAq
1

L
---
z
L
---
\ )
[
cos =
EAq
1

L
---
\ )
[
2

z
L
---
\ )
[
cos
2
z d
0
L

F =
EAq
1

L
---
\ )
[
2
z
2
---
L
4
------ 2
z
L
---
\ )
[
sin +
0
L
F =
EAq
1

2
2L
------ F = solving for q
1
q
1
2

2
-----
FL
EA
-------
\ )
[
=
w z ( )
2

2
-----
FL
EA
-------
\ )
[

z
L
---
\ )
[
sin = 0 z L
N EA dw dz ( ) =
N
2

---F
z
L
---
\ )
[
cos =
w
exact
FL ( ) 4EA ( ) =
w
approx
2FL ( )
2
EA ( ) =
w
exact
w
approx

w
exact
----------------------------------- 100
1 4 2
2
( )
1 4
--------------------------------- 100 18.9% = =
Bars Subjected to Axial Loads
36 Thin-Walled Structures
The maximum value of the normal force in the exact solution, eq. (2.30), occurs in the left half of the bar and
is . The maximum value of the normal force in the approximate solution, eq. (2.41), occurs at z =
0 and is . The percentage error in the approximate solution for the normal force is
The maximum normal force in the approximate solution exceeds the exact value by 27.3%.
We can improve our approximate solution in Example 2.3 by assuming
(2.42)
in which parameters q
1
and q
3
are unknown and are independent. Note that this approximate displacement func-
tion is continuous and satises the prescribed displacement boundary conditions w(0) = 0 and w(L) = 0; that is,
eq. (2.42) is a kinematically admissible displacement function for the bar with xed ends. In fact, each of the
individual sine functions in this equation are kinematically admissible, so there sum is kinematically admissible.
Also note that axial displacement at the center of the bar is w(L/2) = q
1
- q
3
, and the axial displacements at w(L/
3) = w(2L/3) = q
1
. We are approximating the axial displacement by the superposition of two sine functions of dif-
ferent frequencies as is depicted in Fig. 2.14. The principle of virtual work will lead to two equilibrium condi-
tions: one condition determined by the independent virtual displacement q
1
and the second condition
determined from the independent virtual displacement q
3
. Explicit determination of these equilibrium condi-
tions is one of the problems in Section 2.13.

2.6.3 Strain energy density
Lets denote the integrand of the internal virtual work expression in eq. (2.37) as the incremental quantity .
That is,
N
exact
F 2 =
N
approx
2F ( ) =
N
exact
N
approx

N
exact
----------------------------------- - 100
1 2 2
1 2
-------------------------- 100 27.3% = =
w z ( ) q
1

z
L
---
\ )
[
sin q
3
3
z
L
---
\ )
[
sin + = 0 z L
0.2 0.4 0.6 0.8 1
zL
-1
-0.5
0.5
1
0.2 0.4 0.6 0.8 1
zL
-1
-0.75
-0.5
-0.25
0.25
0.5
0.75
1
Fig. 2.14 Sine functions of a two term
approximation to the axial
displacement

z
L
---
\ )
[
sin
3
z
L
---
\ )
[
sin
U
0
Thin-Walled Structures 37
Work and energy methods
(2.43)
For a linear elastic material, we can eliminate the axial force in this equation using the material law in eq. (2.14)
to get
(2.44)
where we used eq. (2.4) to eliminate the strain in terms of the displacement derivative. The form of this incre-
mental quantity suggests there is a function of the derivative of the displacement, or strain, such that its variation
is given by eq. (2.44). If such a function exists, it is actually a function of a function (the displacement deriva-
tive), and it is denoted by , where the prime means ordinary derivative; i.e., . Functions
that depend on a functions are called functionals. But what is meant by the variation of the functional
? To answer this question we consider evaluating the functional for the varied function . Since
the varied function contains the scalar parameter , and as (refer to eq. (2.32)), the change
in with respect to the function can be examined by considering as an ordinary function of
the parameter . Thus, the variation of the functional is dened as
(2.45)
Regard the quantity as a simple variable in the functional , and by using the chain rule for differentiation,
we write the previous equation as
since as , and where . But eq. (2.31) gives . So the variation of func-
tional is
(2.46)
Equate this denition to eq. (2.44) and we have
which must old for every kinematically admissible . Consequently,
Integrate this relation this relation with respect to , treating as a simple variable as in the ordinary calculus,
to nd
(2.47)
U
0
N
z
N
dw
dz
-------
\ )
[
= =
U
0
EA
dw
dz
-------
z
t

dw
dz
-------
\ )
[
=
U
0
w z ( ) [ ] w dw dz =
U
0
w z ( ) [ ] w

z ( )
w

z ( ) w z ( ) 0
U
0
w z ( ) U
0
w

z ( ) [ ]

U
0
U
0
w + [ ] U
o
w [ ]

--------------------------------------------------------
0
lim
| |

| |

d
d
U
0
w

[ ]
0 = | |

| |
=
w

U
0
U
0
w


U
0

( )
0 =

w
U
0
= =
w

w 0 w

= w =
U
0
U
0
w
U
0
w =
U
0
w
U
0
w EA w
z
t
[ ]w = =
w
w
U
0
EA w
z
t
( ) =
w w
U
0
w z ( ) [ ]
1
2
---EA w ( )
2
EA
z
t
w =
Bars Subjected to Axial Loads
38 Thin-Walled Structures
in which the constant of integration is zero since is taken be zero if the strain is zero. We have also assumed
in this integration that the thermal strain is independent of the strain. The functional or , given
by eq. (2.47) is called the strain energy density. The dimensional units of the strain energy density in this case are
(F-L)/L. Notice that the partial derivative of the strain energy density with respect to the strain gives the axial
force; i.e.,
(2.48)
The fact that is an important property of the strain energy density. In fact, an elastic material can
be dened as one for which a strain energy density function exists.
Return to the internal virtual work expression given in eq. (2.37) and identify the integrand on the right-hand
side as the variation in the strain energy density as given in eq. (2.47). We write the internal virtual work as
(2.49)
where we interchanged the variational operator and the integral per the second of eqs. (2.34), and then dened
the strain energy as
(2.50)
The strain energy represents the energy stored in the bar due to elastic deformation. It has the dimensional units
of F-L. Note that the temperature, which is contained in the thermal strain, enters the strain energy as a parame-
ter.
EXAMPLE 2.4 Strain energy of a bar with xed ends and subjected to an axial point force.
Consider the bar with xed ends subjected to the axial point force as discussed in Example 2.2 on page 28 .
(a) Determine the strain energy in the bar.
(b) Determine the strain energy in the approximate solution of the bar given in Example 2.3.
(c) Determine the percentage error in the approximate strain energy with respect to its exact value.
Solution (a) The displacement function of the exact solution is given by eq. (2.29). Differentiating this expres-
sion with respect to z determines the strain, and this strain is
Substitute this result into the expression for the strain energy given by eq. (2.50), and note that in this example
the thermal strain vanishes, to nd
U
0
U
0
w z ( ) [ ] U
0

z
[ ]
N

U
0
EA
z

z
t
( ) = =
N U
0

z
=
W
int
U
0
z d
0
L

U = =
U U
0
z d
0
L

EA
1
2
---
z
2

z
t

\ )
[
z d
0
L

= =

z
bF
EAL
----------- 0 z a
aF
EAL
----------- a z L
|
|

|
|
=
Thin-Walled Structures 39
Work and energy methods
Since the bar is uniform, the extensional stiffness EA is uniform in z. So, after integration, the strain energy
becomes
where we used the fact that . For the case of a = b = L/2, the strain energy in the exact solution
reduces to
(b) The approximate solution for the displacement is given by eq. (2.40), which is repeated below.
Differentiating this expression to nd the strain we get
Substitute this result into the expression for the strain energy given by eq. (2.50), and again note there is no ther-
mal strain, to nd
Evaluating this result we get
(c) The exact strain energy is determined in part (a) as , and the strain energy in the approxi-
mate solution is determined in part (b) as . The percentage error in the approximate solution
with respect to the exact solution is given by

The next example illustrates that the PVW as applied to a linear elastic bar gives the exact solution, if the
assumed displacement function can represent the exact displacement. Of course, in solving for the response of
more complex structures the exact form of the displacement function is not known. It is reassuring, however, that
U
1
2
--- EA
bF
EAL
-----------
\ )
[
2
z d
0
a

1
2
--- EA
aF
EAL
-----------
\ )
[
2
z d
a
L

+ =
U
ab
2
F
2
2EAL
2
-----------------
a
2
bF
2
2EAL
2
----------------- +
abF
2
2EAL
-------------- = =
b L a =
U
LF
2
8EA
----------- = a b L 2 = = ( )
w z ( )
2

2
-----
FL
EA
-------
\ )
[

z
L
---
\ )
[
sin = 0 z L

z
2

---
F
EA
-------
\ )
[

z
L
---
\ )
[
cos = 0 z L
U
1
2
--- EA
2

---
F
EA
-------
\ )
[

z
L
---
\ )
[
cos
2
z d
0
L

=
U
2F
2

2
EA
-------------
z
L
---
\ )
[
cos
2
z d
0
L

2F
2

2
EA
-------------
z
2
---
L
4
------ 2
z
L
---
\ )
[
sin +
0
L
1

2
-----
F
2
L
EA
----------
\ )
[
= = =
U
exact
LF
2
8EA
----------- =
U
approx
1

2
-----
F
2
L
EA
----------
\ )
[
=
U
exact
U
approx

U
exact
------------------------------------ 100
1 8 1
2

1 8
---------------------------- 100 18.9% = =
Bars Subjected to Axial Loads
40 Thin-Walled Structures
if the exact displacement is substituted into the PVW, the exact equilibrium equations are obtained from the prin-
ciple.
EXAMPLE 2.5 An elastic bar subjected to two forces and a thermal load
Show that virtual work eq. (2.38) leads to equilibrium equations for the displacements of the linear elastic bar of
Example 2.1 on page 27. For simplicity, take the distributed load for all z. The bar is subjected to
specied end forces and , and a spatially uniform temperature change . The end displacements
and are unknown, since the end forces are prescribed. Take the kinematically admissible displacement to
be
(2.51)
Parameters q
1
and q
2
are unknown and independent. Parameters q
1
and q
2
physically represent the end displac-
ments of the bar since and .
Solution Since end displacements are not known the variation in the displacement function, or virtual displace-
ment, is
and the variation in the derivative of the displacement, or virtual strain, is
(2.52)
From eq. (2.49) the internal virtual work is given by the variation of the strain energy. That is,
From eq. (2.47) the strain energy density is
,
and its variation, where the variation of a functional is dened by eq. (2.46), is
Obtain the derivative of the displacement from eq. (2.51) and substitute it, and eq. (2.52) for the variation of the
derivative, into the previous equation to get
Substitute this variation of the strain energy density into the internal virtual work to get
p
z
z ( ) 0 =
Q
1
Q
2
T T
0
( )
q
1
q
2
w z ( ) q
1
1
z
L
---
\ )
[
q
2
z
L
--- + =
w 0 ( ) q
1
= w L ( ) q
2
=
w z ( ) q
1
1
z
L
---
\ )
[
q
2
z
L
---
\ )
[
+ =
w
q
2
q
1
( )
L
---------------------------- =
W
int
U
0
z d
0
L

=
U
0
w z ( ) [ ]
1
2
---EA w ( )
2
EA
z
t
w =
U
0
w
U
0
w EA w
z
t
( )w = =
U
0
EA
q
2
q
1
( )
L
---------------------
z
t

q
2
q
1
( )
L
---------------------------- =
Thin-Walled Structures 41
Castiglianos rst theorem
The external virtual work for no distributed loading, see eq. (2.35), becomes
The principle of virtual work, eq. (2.38), is
Rearrange this result into the form
(2.53)
where the thermal force . Since the virtual displacements are independent, we can take and
, so that eq. (2.53) leads to
which in turn gives
(2.54)
Now take and and then eq. (2.53) leads to
which in turn gives
(2.55)
Equations (2.54) and (2.55) agree with the exact result as derived in Section 2.5 and specied by eqs. (2.23) and
(2.24) if the distributed load intensity is set to zero. We get the exact result because the given displacement
function, eq. (2.51), is the exact displacement for this case, as can be veried from eq. (2.20).
2.7 Castiglianos rst theorem
Example 2.5 showed that the exact relationship between the end forces, the corresponding end displacements,
and the thermal force, can be determine from virtual work for a uniform bar with no distributed load acting on it
if the displacement function specied is exact. That is, if the displacement function specied corresponds to the
displacement of the equilibrium state of the elastic bar, then virtual work gives the exact result for the point
forces acting on the bar. In the example, the displacement function given by eq. (2.51) is continuous in z, can rep-
resent a state of uniform strain, and the end displacements appear as parameters in the function. Of course, the
displacement function corresponding to the equilibrium state is not known without solving the governing differ-
W
int
EA
q
2
q
1
( )
L
---------------------
z
t
q
2
q
1
( ) =
W
ext
Q
1
q
1
Q
2
q
2
+ =
Q
1
q
1
Q
2
q
2
+ EA
q
2
q
1
( )
L
---------------------
z
t
q
2
q
1
( ) = q
1
and q
2

Q
1
EA
L
------- q
2
q
1
( ) N
t
+ q
1
Q
2
EA
L
------- q
2
q
1
( ) N
t
+ q
2
+ 0 = q
1
and q
2

N
t
EA
z
t
= q
1
0
q
2
0 =
Q
1
EA
L
------- q
2
q
1
( ) N
t
+ q
1
0 = q
1
0
Q
1
EA
L
------- q
2
q
1
( ) N
t
+ =
q
1
0 = q
2
0
Q
2
EA
L
------- q
2
q
1
( ) N
t
+ q
2
0 = q
2
0
Q
2
EA
L
------- q
2
q
1
( ) N
t
=
p
0
Bars Subjected to Axial Loads
42 Thin-Walled Structures
ential equation (2.16) subject to boundary conditions (2.2) and (2.3). But the procedure followed in this example
can be extended to structures built-up of bars subjected to end forces and a uniform temperature change from the
stress free state, since the displacement of each bar is of the form of eq. (2.51). These observations motivate Cas-
tiglianos rst theorem.
The virtual work eq. (2.38) is extended to an elastic structural system subjected to many external point forces
, , but no distributed loads. The displacements at the points of application of the forces, in the
direction of the forces, are denoted , , and are assumed to be independent. Forces and dis-
placements are said to be corresponding forces and displacements, since they are dened to act at the same
point on the bodys surface and in the same direction. We consider the external point forces as specied, and the
corresponding displacements as unknown. The virtual work eq. (2.38) becomes
(2.56)
In principle, it is possible to write the displacement functions for the structural system such that the independent
s appear as parameters in the functions. Then the strains in terms of the s are determined from the strain-
displacement equations similar to eq. (2.4). Finally, the strains are substituted into the strain energy of the system
and integration over the volume of the system is carried out explicitly. Hence, the strain energy of the structural
system becomes a function of the s; i.e.,
Hence, the variation in the strain energy is
(2.57)
and the virtual work, eq. (2.56), becomes
(2.58)
Since the virtual displacements are independent, we conclude from eq. (2.58) that
(2.59)
This is the equation of Castiglianos rst theorem, which is also called Castiglianos theorem part I. Note the sim-
ilarity of eq. (2.59) to the property of the strain energy density that the internal bar force is equal to the derivative
of the strain energy density with respect to the strain; or, . This theorem is stated for the general
case as follows.
Q
n
n 1 2 N , , , =
q
n
n 1 2 N , , , = Q
n
q
n
Q
n
q
n
n 1 =
N

U =
q
n
q
n
q
n
U U q
1
q
2
q
N
, , , ( ) =
U
q
n

U
q
n
n 1 =
N

=
Q
n
q
n

\ )
[
q
n
n 1 =
N

0 = q
n
n 1 2 N , , , =
Q
n
U
q
n
-------- = n 1 2 N , , , =
N U
0

z
=
Thin-Walled Structures 43
Castiglianos rst theorem
To illustrate the application of Castiglianos rst theorem consider the two-force member in Example 2.5 on
page 40 again. Substitute the displacement function given in eq. (2.51) into the strain-displacement relation of eq.
(2.4) to determine the strain as
Substitute this expression for the strain into the strain energy given by eq. (2.50) to get
Perform the integral over the length of the bar to nd
where the thermal force is dened as . From eq. (2.59) we nd
which coincide with the exact solution given in Example 2.1 on page 27 in eqs. (2.23) and (2.24) when the dis-
tributed load intensity set to zero.
EXAMPLE 2.6 Response of a stepped bar by Castiglianos rst theorem
Determine the relationship for between the external forces and the corresponding displacements for equilibrium
of the axially loaded bar containing a step change in cross section as shown below. There is no change in temper-
ature from the stress-free state. Use Castiglianos rst theorem.
Castiglianos rst theorem
If the strain energy of an elastic structure is expressed in terms of the independent
displacement components , , in the direction of the prescribed point
forces , and there are no distributed loads, then the rst partial derivative
of the strain energy with respect to the displacement is equal to the corresponding
force or
q
n
n 1 2 N , , , =
Q
1
Q
2
Q
N
, ,
q
n
Q
n
Q
n
U
q
n
-------- = n 1 2 N , , , =

z
q
2
q
1
( )
L
--------------------- =
U EA
1
2
---
q
2
q
1

L
----------------
\ )
[
2

z
t
q
2
q
1

L
----------------
\ )
[
z d
0
L

=
U
1
2
---
EA
L
------- q
2
q
1
( )
2
N
t
q
2
q
1
( ) =
N
t
EA
z
t
=
Q
1
q
1

U EA
L
------- q
2
q
1
( ) 1 ( ) N
t
+ = =
Q
2
q
2

U EA
L
------- q
2
q
1
( ) N
t
= =
Bars Subjected to Axial Loads
44 Thin-Walled Structures
Solution The deformation of the bar is a state of uniform strain in each section. Using the displacement function
corresponding to uniform strain given by eq. (2.51) in Example 2.5 on page 40 as a guide, we write the axial dis-
placement functions for each section of the bar as
Note that these displacement functions are continuous, w(0) = 0, w(L
1
) = q
1
for both functions, and w(L
1
+ L
2
) =
q
2
. From the strain-displacement relation, eq. (2.4), these displacements yield the strains in each section as
The displacement is continuous at z = L
1
, but the strain is discontinuous at unless the external forces are
applied in such a manner that . The strain energy is the sum of the strain energies in
each section as determined from eq. (2.50) with the thermal strain set equal to zero; i.e.,
Substituting for the strains in terms of the displacements we get
Castiglianos theorem gives
Writing these results in matrix form, we have
EA ( )
1
EA ( )
2
q
1
Q
1
,
q
2
Q
2
,
L
1
L
2
z w ,
Fig. 2.15 Stepped bar
w z ( ) q
1
z
L
1
----- = 0 z L
1

w z ( ) q
1
1
z L
1

L
2
--------------
\ )
[
q
2
z L
1

L
2
--------------
\ )
[
+ = L
1
z L
1
L
2
+

z
( )
1
q
1
L
1
----- =
z
( )
2
q
2
q
1

L
2
---------------- =
z L
1
=
L
1
L
2
+ ( )q
1
L
2
q
2
0 =
U
1
2
--- EA ( )
1

z
( )
1
2
z d
0
L
1

1
2
--- EA ( )
2

z
( )
2
2
z d
L
1
L
1
L
2
+ ( )

+ =
U
1
2
---
EA
L
-------
\ )
[
1
q
1
2
1
2
---
EA
L
-------
\ )
[
2
q
2
q
1
( )
2
+ =
Q
1
q
1

U EA
L
-------
\ )
[
1
q
1
EA
L
-------
\ )
[
2
q
2
q
1
( ) 1 ( ) + = =
Q
2
q
2

U
0
EA
L
-------
\ )
[
2
q
2
q
1
( ) + = =
Thin-Walled Structures 45
Trusses
The 2 x 2 matrix
(2.60)
is called the stiffness matrix of the structure. Note that the stiffness matrix is symmetric. A symmetric stiffness
matrix results for any linear elastic structural system undergoing small displacements.
2.8 Trusses
Trusses are idealized as an assemblage of two-force members connected by smooth ball-and-socket joints in
three-dimensional trusses, or by smooth hinge joints in a planar truss. External forces are assumed to act only at
the joints. The line connecting the joints at the end of the bar is assumed to coincide with the reference axis of the
bar. Hence, the axial force and strain in each bar is uniform along its length, and the bar is either in tension or
compression. A planar truss consisting of fteen bars and eight joints is shown in Fig. 2.16. Each joint in a pla-
nar truss has two degrees of freedom, one horizontal and the other vertical. Hence, there are sixteen degrees of
freedom for this truss. At joint i, which is often called node i, i = 1, 2,..., 8, the horizontal displacement is denoted
by and the vertical displacement is denoted by . The positive directions for the displacements and cor-
responding forces in the fteen bar truss are dened as shown in Fig. 2.16. The original coordinates of the joints
and the sixteen displacements completely dene the conguration of the truss in the deformed state. Cas-
tiglianos rst theorem is particularly useful in the analysis of the static structural response of trusses. The dis-
placements and the corresponding forces , , used in the formulation of Castiglianos
theorem are the displacements and forces at the joints, or nodes.
To use Castiglianos rst theorem, we need the strain energy of the truss in terms of the nodal displacements.
Begin with a generic bar, say, the i-th bar whose original length is and elongation in the deformed state is .
Since the strain is uniform along the bar the axial strain of the i-th bar is
Q
1
Q
2
k
1
k
2
+ k
2

k
2
k
2
q
1
q
2
= where k
1
EA
L
-------
\ )
[
1
= k
2
EA
L
-------
\ )
[
2
=
k
1
k
2
+ ( ) k
2

k
2
k
2
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
1
2 4
3 5
6
7
8
Fig. 2.16 A 15 bar truss
joint numbering degree of freedom numbering
q
2i 1
q
2i
q
n
Q
n
n 1 2 16 , , , =
L
i

i
Bars Subjected to Axial Loads
46 Thin-Walled Structures
(2.61)
The bar may also be subjected to a uniform temperature change from the stress free state, and the thermal strain
due to temperature for the i-th bar is denoted as . From eq. (2.50) the strain energy in the i-th bar is
(2.62)
The integrand in this equation is independent of the z-coordinate along the reference axis of the bar, and the
result of the integration is simply the length of the bar times the integrand. Substitute eq. (2.61) for the strain, and
denote the thermal force in the i-th bar as , so that the strain energy becomes
(2.63)
The strain energy of the assemblage is simply the sum of the strain energies in each bar; i.e.,
(2.64)
We have assumed the temperature change is spatially uniform in each bar, but it can be different from bar to bar.
The elongation of a truss bar depends on the nodal displacements at its end points. Hence, Castiglianos rst the-
orem for the truss shown in Fig. 2.16 gives
(2.65)
Since the directions of the nodal displacements at the end of a typical bar do not coincide with the reference
axis of the bar, it is necessary to express the elongation of the bar in terms of the nodal displacements in order to
use Castiglianos theorem.
Geometry of deformation To relate the elongation of a typical bar to its nodal displacements, requires a study
of the geometry of the deformation of the bar. The end nodes, or joints, of the bar in the undeformed state are
labeled by integers i and j as shown in Fig. 2.17. In the x-y plane of the truss, the coordinates of the beginning
node i are , and the coordinates of the end node j are . The square of the length of the undeformed
bar is determined by the pythagorean theorem as
(2.66)
and the direction cosines of the undeformed bar are
(2.67)
As is depicted in Fig. 2.17, the bar displaces and rotates to a new position in the plane. The coordinates of
the beginning node in the deformed state are , and the coordinates of the end node in the
deformed state are . The counterclockwise rotation of the bar is denoted by the angle .

i
L
i
----- =

i
t
U
i
EA ( )
i
1
2
---
i
2

i
t

i
z d
0
L
i

=
N
i
t
EA ( )
i

i
t
=
U
i
1
2
---
EA
L
-------
\ )
[
i

i
2
N
i
t

i
=
U
1
2
---
EA
L
-------
\ )
[
i

i
2
N
i
t

all bars

=
Q
n
EA
L
-------
\ )
[
i

i
N
i
t

q
n

i
\ )
| j
[
i 1 =
15

= n 1 2 16 , , , =
x
i
y
i
, ( ) x
j
y
j
, ( )
L
2
x
j
x
i
( )
2
y
j
y
i
( )
2
+ =
cos
x
j
x
i

L
--------------- = 90 ( ) cos sin
y
j
y
i

L
-------------- = =
x
i
q
2i 1
+ y
i
q
2i
+ , ( )
x
j
q
2 j 1
+ y
j
q
2 j
+ , ( )
Thin-Walled Structures 47
Trusses
Let denote the length of the bar in the deformed state. The square of the length of the bar in the deformed
conguration is
(2.68)
and the direction cosines of the bar in the deformed conguration are
(2.69)
(2.70)
Dene the relative elongation of the bar in the x-direction as
(2.71)
and the relative elongation of the bar in the y-direction as
(2.72)
Rearranging eq. (2.68) and using the above denitions we get
(2.73)
The left-hand sides of eqs. (2.69) and (2.70) are expanded by the trigonometric identity for the sum of angles,
and the denitions given by eqs. (2.71) and (2.72) are used in the right-hand sides of eqs. (2.69) and (2.70) to get
(2.74)
(2.75)
x
i
y
i
x
j
y
j
i
j
L
L
*

q
2i 1
q
2i
q
2 j 1
q
2 j
i
*
j
*
x
y
0
+
Fig. 2.17 Displacement, rotation, and elongation of a truss bar
L
*
L
*
( )
2
x
j
q
2 j 1
+ ( ) x
i
q
2i 1
+ ( ) [ ]
2
y
j
q
2 j
+ ( ) y
i
q
2i
+ ( ) [ ]
2
+ =
+ ( ) cos
x
j
q
2 j 1
+ ( ) x
i
q
2i 1
+ ( )
L
*
---------------------------------------------------------------- =
+ ( ) sin
y
j
q
2 j
+ ( ) y
i
q
2i
+ ( )
L
*
--------------------------------------------------- =

x
q
2 j 1
q
2i 1
( )
L
-------------------------------------

y
q
2 j
q
2i
( )
L
------------------------- =
L
*
( )
2
L
2
cos
x
+ ( )
2
sin
y
+ ( )
2
+ [ ] =
cos cos sin sin
L
L
*
------ cos
x
+ ( ) =
cos sin sin cos +
L
L
*
----- sin
y
+ ( ) =
Bars Subjected to Axial Loads
48 Thin-Walled Structures
If denotes the elongation of the bar, then . The latter expression can be written as
where the strain of the bar is denoted by . Expand the right-hand side of eq. (2.73), use
the trigonometric identity , and let to get
(2.76)
Expand the left-hand side of this result, subtract one from each side, and the divide each side by two to get
(2.77)
To nd the sine of the rotation angle, multiply eq. (2.74) by , multiply eq. (2.75) , and then add the
resulting equations. We obtain
(2.78)
Given the relative x-direction displacements and the relative y-direction displacements of the nodes, the strain of
the bar is determined from eq. (2.77) and the rotation of the bar is determined by eq. (2.78). For most trusses, the
strains and rotation of the bars are very small. For innitesimal deformations, the following quantities are very
small
Hence, eq. (2.77) yields the following approximation for the bar strain
(2.79)
and eq. (2.78) yields the following approximation for the rotation
(2.80)
Substitute eqs. (2.71) and (2.72) for the relative elongations into eq. (2.79) to get
Hence, the elongation is
(2.81)
For small strain and small rotations of the bar, eq. (2.81) shows that the elonga-
tion is the sum of the projections of the relative displacements onto the refer-
ence axis of the undeformed bar, as is shown in Fig. 2.18.
EXAMPLE 2.7 Three bar planar truss
The planar truss shown in Fig. 2.19 consists of three bars and one movable joint. Take the thermal strains to be
zero. Determine the 2 x 2 stiffness matrix using Castiglianos rst theorem.
Solution The elongation of each bar as determined from eq. (2.81) is
(2.82)
L
*
L + =
L
*
L 1 + ( ) = L =
cos
2
sin
2
+ 1 = L
*
L 1 + ( ) =
1 + ( )
2
1 2 cos ( )
x
2 sin ( )
y

x
2

y
2
+ + + + =
1
1
2
--- +
\ )
[
cos ( )
x
sin ( )
y
1
2
---
x
2

y
2
+ ( ) + + =
sin cos
sin
1
1 + ( )
---------------- cos ( )
y
sin ( )
x
[ ] =
0 1 0
x
1 0
y
1
cos ( )
x
sin ( )
y
+
cos ( )
y
sin ( )
x

L
---
q
2 j 1
q
2i 1

L
--------------------------------
\ )
[
cos
q
2 j
q
2i

L
--------------------
\ )
[
sin +

q
2 j 1
q
2i 1
( )
q
2 j
q
2i
( )
i
j
Fig. 2.18
cos ( ) q
2 j 1
q
2i 1
( ) sin ( ) q
2 j
q
2i
( ) + =

i

i
( ) cos q
1

i
( ) sin q
2
+ = i 1 2 3 , , =
Thin-Walled Structures 49
Trusses
Equation (2.65) results from the application of Castiglianos theorem, and applied to this example gives
These results are written in the matrix form
(2.83)
where the elements of the stiffness matrix are
(2.84)
Note that this example is statically indeterminate, since we have only two equilibrium equations at the movable
joint for three unknown bar forces. Given the applied forces , matrix eq. (2.83) is solved for the nodal
displacements . From eq. (2.82) the elongation of each bar is then computed, and from these elonga-
tions the bar forces are determined from
(2.85)
EXAMPLE 2.8 Three bar truss with lack of t
q
1
Q
1
,
q
2
Q
2
,

1

2

3
EA
L
-------
\ )
[
1
EA
L
-------
\ )
[
2
EA
L
-------
\ )
[
3
Fig. 2.19 Three bar truss
Q
1
EA
L
-------
\ )
[
i

i
( ) cos q
1

i
( ) sin q
2
+ [ ]
i
( ) cos [ ]
i 1 =
3

=
Q
2
EA
L
-------
\ )
[
i

i
( ) cos q
1

i
( ) sin q
2
+ [ ]
i
( ) sin [ ]
i 1 =
3

=
Q
1
Q
2
k
11
k
12
k
21
k
22
q
1
q
2
=
k
11
EA
L
-------
\ )
[
i

i
( ) cos
2
i 1 =
3

= k
22
EA
L
-------
\ )
[
i

i
( ) sin
2
i 1 =
3

=
k
12
k
21
EA
L
-------
\ )
[
i

i
( ) cos
i
( ) sin
i 1 =
3

= =
Q
1
and Q
2
q
1
and q
2
N
i
EA
L
-------
\ )
[
i

i
= i 1 2 3 , , =
Bars Subjected to Axial Loads
50 Thin-Walled Structures
Consider the same three bar truss of Example 2.7, but now assume that bar 1 was too short and had to be
stretched an amount in order to connect to the joint. This is a case of lack of t, and lack of t is common in
the fabrication of structures. That is, before the external loads are applied ( ), the truss bars experi-
ence initial forces due to the lack of t of bar 1. Determine the initial forces in the bars using Castiglianos rst
theorem.
Solution The strain energy for bar 1 is determined by the total elongation of the bar, which is the sum of initial
elongation due to lack of t plus the elongation experienced by bar 1 after the connection to the joint is made.
That is, the total elongation of bar 1 is , where denotes the additional elongation of bar 1 after the
connection of the bar to the joint is made. Hence, the total strain energy for the system in this case is
The elongation of each bar after the connection is made is
Use Castiglianos rst theorem to get
or
Substitute into this equation the relation between the nodal displacements and the elongation of each bar, eq.
(2.82), after the connection is made to get
where the elements of the stiffness matrix are the same as given in Example 2.7. Setting and ,
since no external forces are applied to the joint just after assembly, we can solve the above matrix equation for
the joint displacements to get
From this solution we can calculate the elongation of each bar after assembly from
and nally calculate the initial bar forces from

1
Q
1
Q
2
0 = =

1

1
+
1
U
1
2
---
EA
L
-------
\ )
[
1

1

1
+ ( )
2
1
2
---
EA
L
-------
\ )
[
i

i
2
i 2 =
3

+ =

i

i
( ) cos q
1

i
( ) sin q
2
+ = i 1 2 3 , , =
Q
n
EA
L
-------
\ )
[
1

1

1
+ ( )
q
n

1 EA
L
-------
\ )
[
i

i
q
n

i
i 2 =
3

+ = n 1 2 , =
Q
n
EA
L
-------
\ )
[
i

i
q
n

i
i 1 =
3

EA
L
-------
\ )
[
1
q
n

1
\ )
| j
[

1
+ = n 1 2 , =
Q
1
Q
2
k
11
k
12
k
12
k
22
q
1
q
2
EA
L
-------
\ )
[
1

1
( ) cos

1
( ) sin
+ =
Q
1
0 = Q
2
0 =
q
1
q
2
1
k
11
k
22
k
12
2
( )
---------------------------------
k
22
k
12

k
12
k
11

1
( ) cos

1
( ) sin
EA
L
-------
\ )
[
1

1
=

i

i
( ) cos q
1

i
( ) sin q
2
+ = i 1 2 3 , , =
Thin-Walled Structures 51
Complementary virtual work

2.9 Complementary virtual work
For the bar shown in Fig. 2.1 on page 18 consider the loading situation where end displacements and and
the distributed load intensity are prescribed. Also, we assume that the axial displacement function is
continuous and single valued with and ; i.e., kinematically admissible. The external
forces and are interpreted as reactions to the prescribed loading. Now consider an arbitrary, innitesimal
change in the external forces and from this deformation state with the distributed load intensity function
unchanged. These innitesimal changes in the forces are called a virtual forces, and are denoted as and
. The virtual forces cause a change in the internal normal force in the bar. The change in the internal normal
force is denoted by the function . Take the internal virtual force to satisfy the equilibrium conditions for
this case; i.e.,
(2.86)
(2.87)
Note that the distributed load does not enter into the equilibrium differential equation (2.86) since it is not part of
the virtual force system (Q
1
, Q
2
, N(z)). The virtual force system is required to satisfy equilibrium conditions
separately from the actual force system. If the eq. (2.86) is integrated over the length of the bar we get
Substitute into this last expression the boundary conditions from eqs. (2.87) to get
(2.88)
The last condition, eq. (2.88), is necessary for overall equilibrium of the bar under the application of the external
virtual forces. The solution to eqs. (2.86) to (2.88) is that the internal virtual force is uniform in z with
(2.89)
Note that eqs. (2.89) are satised by the overall equilibrium condition (2.88) of the bar. Only one of the external
virtual forces, either or , is independent and permitted to be varied arbitrarily.
Internal actions, like function above, that satisfy the equilibrium differential equations and those
boundary conditions where the forces are specied are called statically admissible functions. A statically admis-
sible function may not satisfy the governing boundary value problem for the response of the bar. However, within
the set of all statically admissible functions one will lead to strains determined from the material law and the dis-
placements determined from the strain-displacement relations such that the specied boundary conditions on the
displacements are satised.
N
1
EA
L
-------
\ )
[
1

1

1
+ ( ) = N
2
EA
L
-------
\ )
[
2

2
= N
3
EA
L
-------
\ )
[
3

3
=
q
1
q
2
p
z
z ( ) w z ( )
w 0 ( ) q
1
= w L ( ) q
2
=
Q
1
Q
2
Q
1
Q
2
Q
1
Q
2
N z ( )
d
dz
----- N ( ) 0 = 0 z L < <
N z = 0 ( ) Q
1
= N z = L ( ) Q
2
=
z d
d
N ( ) z d
0
L

0 = or N L ( ) N 0 ( ) 0 =
Q
1
Q
2
+ 0 =
N z ( )
N Q
1
Q
2
= =
Q
1
Q
2
N z ( )
Bars Subjected to Axial Loads
52 Thin-Walled Structures
Dene the product of the prescribed displacements and with there corresponding virtual forces
and , respectively, as the external complementary virtual work ; i.e.,
(2.90)
From the solution for the virtual force given in eq. (2.89), the complementary external virtual work can be written
as
(2.91)
Now we relate the strain to the end displacements by integrating the strain-displacement relation given by eq.
(2.4) over the length of the bar to get
(2.92)
Equation (2.92) is a condition of compatibility
1
; it relates the strain in the deformed state of the bar to the speci-
ed end displacements. Substitute eq. (2.92) for in eq. (2.91) to get
The virtual force is spatially uniform via equilibrium eq. (2.86), so that this last equation can be written as
(2.93)
The integrand of this equation contains the strain and the virtual force, and these quantities are dened internal to
the bar. Therefore, we dene the right-hand side of eq. (2.93) as the internal complementary virtual work ;
i.e.,
(2.94)
The internal complementary virtual work is the integral over the length of the bar of the product of the strain in
the deformed state with the virtual force.
We have shown in the process of going from eq. (2.90) to (2.94) that, for compatible displacements and
strains satisfying eq. (2.92), and for virtual forces satisfying the equilibrium conditions given in eqs. (2.86) to
(2.88), the external complementary virtual work is equal to the internal complementary virtual work. That is, we
have proved the necessary condition that if the displacements and strains are compatible, and the virtual force
system is statically admissible, then the external complementary virtual work equals the internal complementary
virtual work. Compatibility means that the displacements and strains satisfy the strain-displacement equations
1. Compatibility for a deformed body means that the displacements are continuous and single-valued functions such that
strains are computable and that no gaps or discontinuities occur in the deforming body. More generally, compatibility
means that the strains have to satisfy certain mathematical conditions such that, when integrated, the strain functions lead
to continuous, single-valued displacements.
q
1
q
2
Q
1
Q
2
W
ext
*
W
ext
*
q
1
Q
1
q
2
Q
2
+ =
W
ext
*
q
2
q
1
( )N =

z
z d
0
L

z d
dw
z d
0
L

w L ( ) w 0 ( ) q
2
q
1
= = =
q
2
q
1

W
ext
*

z
z d
0
L

\ )
| j
[
N =
W
ext
*

z
N z d
0
L

=
W
int
*
W
int
*

z
N z d
0
L

Thin-Walled Structures 53
Complementary virtual work
and prescribed displacement boundary conditions. In problem solving we assume that equating external comple-
mentary virtual to the internal complementary virtual work for every statically admissible system of virtual
forces is sufcient for the displacements and strains of the body to be compatible. We state this principle as fol-
lows.
Note that the phrase variation in the forces means the virtual forces. For the axially loaded bar, the PCVW
means the displacement function w(z) and strain function
z
(z) are compatible if
(2.95)
where the external complementary virtual work is given by eq. (2.90) and the internal complementary virtual
work is given by eq. (2.94). The principle of complementary virtual work is independent of the material behavior.
Apply the principle of complementary virtual work to our bar example. Substitute eq. (2.90) for the external
virtual work and eq. (2.94) for the internal virtual work to get
Requiring the virtual forces to satisfy equilibrium conditions given by eqs. (2.86) to (2.88) results in the solution
given by eq. (2.89). Hence, we can eliminate two of the virtual forces in terms of the third in the third. Say we
take and , so that the principle becomes
This last equation is satised if and only if
which coincides with the compatibility condition, eq. (2.92).
2.9.1 Complementary strain energy
Now assume the material of the bar is elastic and linear. Solve the material law given by eq. (2.14) for the strain
and then substitute it into the internal complementary virtual work of eq. (2.94) to get
(2.96)
Let the integrand of this equation be denoted by the incremental quantity . So
Principle of Complementary Virtual Work (PCVW)
If the external complementary virtual work ( ) equals the internal complementary virtual work
( ) for every statically admissible variation in the forces, then the displacements and strains of the
body are compatible.
W
ext
*
W
int
*
W
ext
*
W
int
*
= for every statically admissible virtual force N z ( )
q
1
Q
1
q
2
Q
2
+
z
N z d
0
L

=
N Q
2
= Q
1
Q
2
=
q
2
q
1

z
z d
0
L

Q
2
0 = Q
2

q
2
q
1

z
z d
0
L

0 =
W
int
*
1
EA
------- N N
t
+ ( )N z d
0
L

=
U
0
*
Bars Subjected to Axial Loads
54 Thin-Walled Structures
(2.97)
This form of the incremental quantity suggests there is a functional of the axial force whose variation is given by
eq. (2.97). This functional is called the complementary strain energy density, and is denoted by . Fol-
lowing the developments we used to nd the strain energy density functional (refer to eqs. (2.45) to (2.47)) we
would nd in a similar manner that the variation of the complementary strain energy density is given by
(2.98)
Equating eq. (2.97) and eq. (2.98) for every statically admissible virtual force gives
(2.99)
Integrating this expression with respect to N, and taking the constant of integration to be zero when N = 0, we get
the complementary strain energy density as
(2.100)
The thermal force appears as a parameter in the complementary strain energy density, since it assumed indepen-
dent of the normal force N. The important property of the complementary strain energy density is
(2.101)
which can be seen from eqs. (2.99) and (2.14). A comparison of eq. (2.101) with eq. (2.48) for the strain energy
density shows the dual attributes that these energies possess: the derivative of the complementary strain energy
density with respect to the axial force gives the strain, and the derivative of the strain energy density with respect
to the strain gives the axial force.
Return to the internal complementary virtual work given by eq. (2.96), and use eq. (2.97) to write it as
Now interchange the variational operator and the integral operator in accordance with the second of eqs. (2.34),
and write this equation as
where the complementary strain energy for the bar is dened as
(2.102)
U
0
*
1
EA
------- N N
t
+ ( )N =
U
0
*
N z ( ) [ ]
U
0
*
N

U
0
*
( )N =
N

U
0
*
( )
1
EA
------- N N
t
+ ( ) =
U
0
*
1
EA
-------
1
2
---N
2
NN
t
+
\ )
[
=

z
N

U
0
*
( ) =
W
int
*
U
0
*
z d
0
L

=
W
int
*
U
*
=
U
*
U
0
*
z d
0
L

1
2EA
----------- N
2
2N
t
N + ( ) z d
0
L

= =
Thin-Walled Structures 55
Relationship between the complementary strain energy and the strain energy densities
2.10 Relationship between the complementary strain energy and the
strain energy densities
To get a clearer understanding of the complementary strain energy density, consider a plot of the axial force ver-
sus axial strain for the uniaxially loaded bar with no change in temperature from the stress free state as is shown
in Fig. 2.20. For a linear elastic material under these isothermal conditions, the material law for the bar is
which plots as a straight line with slope of EA to one (Fig. 2.20a). The area between the straight line
and the strain axis is . From eq. (2.47) the strain energy density is given as .
Hence, the area between the force-strain curve and the strain axis is the strain energy density. For isothermal
deformation, the complementary strain energy density from eq. (2.100) is . The area between
the force-strain curve and the force axis is . Hence, the area between the force-strain curve and the
force axis is the complementary strain energy density. For the linear elastic material under isothermal conditions,
the strain energy density and complementary strain energy density are equal in value, . However, it is
preferred to express the strain energy density as a functional of the strain and the complementary strain energy as
a functional of the axial force, since the former was obtained by considering virtual displacements and the latter
was obtained by considering virtual forces.
Examination of Fig. 2.20a shows that the relationship between the strain energy density and complementary
strain energy density can also be expressed as
(2.103)
Actually, this relationship is more general than for a linear elastic material under isothermal deformation. Equa-
tion (2.103) is valid for nonlinear elastic material behavior as is depicted in Fig. 2.20b, and is valid if the defor-
mations is caused in part by a temperature change from the stress free state. For a nonlinear elastic material law
and/or in the non-isothermal deformation the complementary strain energy density is not equal in value to the
strain energy density ( ), but eq. (2.103) remains valid.
U
0
U
0
*
0

z
N
EA
1
U
0
U
0
*
0

z
N
(a) linear elastic (b) nonlinear elastic
Fig. 2.20 Strain energy density and complementary strain energy density
under isothermal conditions for a uniaxially loaded bar.
N EA
z
=
1
2
---
z
( ) EA
z
( ) U
0
EA
z
2
( ) 2 =
U
o
*
N
2
2EA ( ) =
1
2
--- N ( )
N
EA
-------
\ )
[
U
0
U
0
*
=
U
0
*
U
0
N
z
+ =
U
0
*
U
0

Bars Subjected to Axial Loads


56 Thin-Walled Structures
We derive the complementary strain energy density for a linear elastic material under non-isothermal defor-
mation using eq. (2.103). Start with the strain energy density as given by eq. (2.47), which is
Use the material law, eq. (2.14), to eliminate the strain in this equation to express the strain energy in terms of the
axial force, and get
which upon simplication becomes
(2.104)
Now substitute this result for the strain energy density in eq. (2.103), and substitute the strain-force form of the
material law for the strain in eq. (2.103), to get
After simplication this equation reduces to
(2.105)
Compare this expression to the complementary strain energy density given by eq. (2.100) and we see that the
expressions differ by the thermal force term . However, this term is irrelevant since it does not con-
tribute to the property of the complementary strain energy density that
(2.106)
That is, both eqs. (2.100) and (2.105) will yield the same resulting form of the strain-force form of the material
law if substituted into eq. (2.106). Hence, we can drop the term in eq. (2.105) without loss of gen-
erality since it is independent of the axial force.
Comparing the complementary strain energy density given by eq. (2.100) to the strain energy density written
in terms of the axial force, eq. (2.104), it is clear that because of the temperature effect.
EXAMPLE 2.9 Application of complementary virtual work to an elastic bar
Consider the uniform bar xed against rigid body displacement by a support at z = 0, as is shown in the gure
below. That is, the displacement . Take the end displacement to be specied, but not necessarily zero.
End forces and are regarded as reactions to the prescribed displacements. In addition, the bar is subject to
a uniformly distributed load of intensity , and a uniform change in temperature from the stress free
state. Assume the bar is in a state of equilibrium with the displacements and strains being compatible under the
U
0
1
2
---EA
z
2
EA
z
t

z
=
U
0
1
2
---EA
N
EA
-------
N
t
EA
------- +
\ )
[
2
N
t
N
EA
-------
N
t
EA
------- +
\ )
[
=
U
0
1
2
---
N
2
EA
-------
1
2
---
N
t
( )
2
EA
------------- =
U
0
*
1
2
---
N
2
EA
-------
1
2
---
N
t
( )
2
EA
-------------
\ )
[
N
N
EA
-------
N
t
EA
------- +
\ )
[
+ =
U
0
*
1
2EA
----------- N
2
2NN
t
N
t
( )
2
+ + [ ] =
N
t
( )
2
2EA ( )

z
N

U
0
*
( ) =
N
t
( )
2
2EA ( )
U
0
*
U
0

q
1
0 = q
2
Q
1
Q
2
p
0
T T
0
( )
Thin-Walled Structures 57
Relationship between the complementary strain energy and the strain energy densities
imposed loads. Apply the principle of complementary virtual work and use, in addition, that the material of the
bar is linear elastic to relate the end displacement to the applied loads and the reaction .
Solution From eq. (2.90) the external complementary virtual work is , since . The internal virtual
work is the variation of complementary strain energy, since the bar material is elastic; i.e., . The
complementary strain energy is given by eq. (2.102) for a linear elastic material. Hence, the principle of comple-
mentary virtual work for this linear elastic case is
Interchange the variational operator and the integral operator on the right-hand side of this equation in accor-
dance with the second of eqs. (2.34), then determine the variation of the integrand per the denition of the varia-
tion of a functional given by eq. (2.46). Mathematically these operations are
Perform the differentiation in the integrand to get
(2.107)
Following the principle of complementary virtual work, the virtual forces are to satisfy equilibrium conditions
given in eqs. (2.86) to (2.88), which have the solution given in eq. (2.89). Hence, on the basis of virtual force
equilibrium we can take , and then eq. (2.107) becomes
This equation is satised if and only if
(2.108)
Equation (2.108) can be recognized as the condition of compatibility for the bar made of a linear elastic material
by substituting the strain from the material law given in eq. (2.14) into the integrated form of the strain-displace-
ment relation given in eq. (2.92).
q
2
Q
2
q
2
Q
2
,
z
p
0
L
q
1
0 =
EA constant =
q
2
Q
2
q
1
0 =
W
int
*
U
*
=
q
2
Q
2

1
2
---
N
2
EA
------- 2
N
t
N
EA
---------- + z d
0
L

\ )
| j
[
=

1
2
---
N
2
EA
------- 2
N
t
N
EA
---------- + z d
0
L

\ )
| j
[
N
N
2
2EA
-----------
N
t
N
EA
---------- +
| |

| |
N z d
0
L

=
q
2
Q
2
N
EA
-------
N
t
EA
------- + N z d
0
L

=
N Q
2
=
q
2
N
EA
-------
N
t
EA
------- + z d
0
L

\ )
| j
[
Q
2
0 = Q
2

q
2
N
EA
-------
N
t
EA
------- + z d
0
L

=
Bars Subjected to Axial Loads
58 Thin-Walled Structures
The bar was assumed to be in equilibrium prior to application of the
virtual force . Hence, the axial force must satisfy the equilib-
rium eq. (2.1) with . Alternatively we can impose equilibrium
of the free body diagram of the right portion of the bar cut at z as shown in
Fig. 2.21. Equilibrium gives
(2.109)
Note that the force is unknown at this point since the displacement was specied instead of . Substi-
tute eq. (2.109) for the axial force in eq. (2.108) and perform the integration to get
(2.110)
The exact solution was obtained Section 2.5 as eq. (2.24). Solving eq. (2.24) for displacement and setting
we arrive at eq. (2.110). Equation (2.110) can be used to nd the force in terms of the prescribed
quantities , and .
In summary, we have related the displacement to the corresponding point force by rst using the
principle of complementary virtual work to establish compatibility between the displacement and the strain
; second, we used the linear elastic material law to relate the strain to the axial force , and nally equilibrium
was used to determine the axial force such that the unknown force appeared as parameter.
2.11 Generalized form of Castiglianos second theorem
Based on the results for the elastic bar subjected only to a variation in the point force given in Example 2.9,
the complementary virtual work eq. (2.95) is extended to an elastic structural system restrained against rigid
body displacements, and subjected to many external point forces , . Distributed loads and
thermal strains can act on the system as well. The point forces may be independently varied since the variations
in the reactions at the support points compensate to keep the structure in equilibrium. The corresponding dis-
placements at the points of application of the forces, in the direction of the forces, are denoted ,
. The complementary virtual work eq. (2.95) becomes
(2.111)
From the differential equilibrium equations and the prescribed boundary conditions on the forces, it is possible,
in principle, to write the internal actions (axial force, bending moments, torque, etc.) of the structure in terms of
the independent external point forces . Since the complementary strain energy of the structure is determined
from integrals of the internal actions over the domain of the structure, it becomes, after integration, a function of
N z ( )
p
0
Q
2
z

L
Fig. 2.21
Q
2
N z ( )
p
z
z ( ) p
0
=
N z ( ) Q
2
p
0
d
z
L

+ Q
2
p
0
L z ( ) + = =
Q
2
q
2
Q
2
q
2
L
EA
-------Q
2
L
EA
-------N
t
p
0
L
2
2EA
----------- + + =
q
2
q
1
0 = Q
2
q
2
p
0
N
t
q
2
Q
2
q
2

z
N
N
Q
2
Q
n
n 1 2 N , , , =
q
n
n 1 2 N , , , =
q
n
Q
n
n 1 =
N

U
*
=
Q
n
Thin-Walled Structures 59
Generalized form of Castiglianos second theorem
the s; i.e.,
Hence, the variation in the complementary strain energy is
(2.112)
and the complementary virtual work eq. (2.111) becomes
(2.113)
Since the virtual forces are independent, we conclude from eq. (2.113) that
(2.114)
Equation (2.114) is the general form of Castiglianos second theorem. Note the similarity of eq. (2.114) to the
property of the complementary strain energy density that the strain is equal to the derivative of the complemen-
tary strain energy density with respect to the internal bar force; or, . This theorem is stated for
the general case as (Langhaar, 1962):
Although this theorem is derived from complementary virtual work where we regarded the displacement as pre-
scribed and the corresponding point forces as unknown, it is usually used in practice to determine the displace-
ment corresponding to a point force applied to the structure. The example problems to follow will illustrate this
point.
If the material of the bar is linear elastic and there are no thermal strains, we showed in Section 2.10 that the
strain energy density and the complementary strain energy density have the same value. The strain energy and
complementary strain energy have the same value as well since they are dened as integrals over the domain of
the structure of their respective densities. Hence, in this restricted case we can replace by in eq. (2.114) to
get
(2.115)
Equation (2.115) is Castiglianos second theorem, which is also called Castiglianos theorem part II. Note that
eq. (2.115) implies that the strain energy is written in terms of the internal actions of the structure rather than in
Generalized form of Castiglianos second theorem
If an elastic structure is mounted such that rigid body displacements are impossible
and certain point forces act on the structure, in addition to distributed
loads and thermal strains, the displacement component , of the
point of application of force in the direction of is determined by the equation
Q
n
U
*
U
*
Q
1
Q
2
Q
N
, , , ( ) =
U
*
Q
n

U
*
( )Q
n
n 1 =
N

=
q
n
Q
n

U
*
( )
\ )
[
Q
n
n 1 =
N

0 = Q
n
n 1 2 N , , , =
q
n
U
*
Q
n
---------- = n 1 2 N , , , =

z
U
*
0
N =
Q
1
Q
2
Q
N
, , ,
q
n
n 1 2 N , , , =
Q
n
Q
n
q
n
U
*
Q
n
---------- =
U
*
U
q
n
Q
n

U
= (linear elastic, no thermal strain)
Bars Subjected to Axial Loads
60 Thin-Walled Structures
its natural form, which is in terms of the strains. Since we prefer to write the strain energy in terms of the strains
and the complementary strain energy in terms of the internal forces or stresses, the implementation of Cas-
tiglianos second theorem in this text will use the complementary strain energy on the right-hand side of eq.
(2.115) instead of the strain energy.
EXAMPLE 2.10 Stepped bar response by Castiglianos second theorem
Determine the relationship between the displacements and the corresponding forces of the stepped bar in Exam-
ple 2.6 on page 43 using Castiglianos second theorem. There is no change in temperature from the stress free
state.
Solution From the free body diagrams shown in Fig. 2.22,
equilibrium givesthe axial force as
The complementary strain energy stored in the structure due
to elastic deformation is the sum of the energies from each
segment. (Note: the complementary strain energy and strain
energy are the same in this case.) That is,
From Castiglianos second theorem
Substituting for the axial forces determined from equilibrium we get
Performing the integrations gives
N z ( )
N z ( )
Q
1
Q
2
Q
2
z
z
L
1
L
1
L
2
+
Fig. 2.22
N z ( ) Q
1
Q
2
+ = 0 z L
1
< <
N z ( ) Q
2
= L
1
z L
1
L
2
+ < <
U
*
1
2
---
N
2
EA ( )
1
-------------- z d
0
L
1

1
2
---
N
2
EA ( )
2
-------------- z d
L
1
L
1
L
2
+ ( )

+ =
q
1
Q
1

U
*
( )
N
EA ( )
1
--------------
Q
1

N
z d
0
L
1

N
EA ( )
2
--------------
Q
1

N
z d
L
1
L
1
L
2
+ ( )

+ = =
q
2
Q
2

U
*
( )
N
EA ( )
1
--------------
Q
2

N
z d
0
L
1

N
EA ( )
2
--------------
Q
2

N
z d
L
1
L
1
L
2
+ ( )

+ = =
q
1
Q
1
Q
2
+ ( )
EA ( )
1
------------------------ 1 ( ) z d
0
L
1

Q
2
EA ( )
2
-------------- 0 ( ) z d
L
1
L
1
L
2
+ ( )

+ =
q
2
Q
1
Q
2
+ ( )
EA ( )
1
------------------------ 1 ( ) z d
0
L
1

Q
2
EA ( )
2
-------------- 1 ( ) z d
L
1
L
1
L
2
+ ( )

+ =
q
1
L
EA
-------
\ )
[
1
Q
1
Q
2
+ ( ) =
Thin-Walled Structures 61
Generalized form of Castiglianos second theorem
Finally, write these results in matrix form as
The 2 x 2 matrix
(2.116)
is called the exibility matrix. It is the inverse of the stiffness matrix obtained in Example 2.6, eq. (2.60); i.e.,

EXAMPLE 2.11 A suspended bar subjected to self weight
A a uniform bar is suspended from the ceiling and is subjected to self weight and a tip force as shown in Fig.
2.23. There is no change in temperature from the stress free state, and the material is linear elastic. The material
has a specic weight denoted by which has dimensional units of . Determine the displacement at the tip
of the bar using Castiglianos second theorem.
q
2
L
EA
-------
\ )
[
1
Q
1
Q
2
+ ( )
L
EA
-------
\ )
[
2
Q
2
+ =
q
1
q
2
1
k
1
-----
1
k
1
-----
1
k
1
-----
1
k
1
-----
1
k
2
----- +
Q
1
Q
2
= k
1
EA
L
-------
\ )
[
1
= k
2
EA
L
-------
\ )
[
2
=
1
k
1
-----
1
k
1
-----
1
k
1
-----
1
k
1
-----
1
k
2
----- +
k
1
k
2
+ k
2

k
2
k
2
1
k
1
-----
1
k
1
-----
1
k
1
-----
1
k
1
-----
1
k
2
----- +
k
1
k
2
+
k
1
----------------
k
2
k
1
-----
k
1
k
2
+
k
1
---------------- k
2
1
k
1
-----
1
k
2
----- +
\ )
[

0
k
2
k
1
----- k
2
1
k
1
-----
1
k
2
----- +
\ )
[
+
1 0
0 1
= =
Q
2
F L
3

dz
z
N
L
Adz
N dN +
EA
g
q
2
Q
2
,
Fig. 2.23 Suspended bar under self
weight and a tip force
Bars Subjected to Axial Loads
62 Thin-Walled Structures
Solution The bar is subjected to a distributed load intensity caused by self weight. From the free body diagram
of an element of the bar shown in Fig. 2.23, the equilibrium differential equation is
The force prescribed at the tip gives the boundary condition for the axial force as
Integrating the differential equation with respect to z from z to L gives
and from the boundary condition prescribed on the axial force, we nd the axial force is
The complementary strain energy for no thermal force, eq. (2.102), is
The tip displacement is determined from Castiglianos second theorem as
Now substitute for the axial force this equation to get
Thus, after integration, we nd the tip displacement is
This result for the displacement can be evaluated for , of course, which suggests a method to compute
the displacement at a point on the structure where no point force acts. Merely, put a force at the point in question,
apply Castiglianos theorem, then set the point force to zero.
2.12 References
Callister, W.D., 1997, Materials Science and Engineering, Fourth Edition, John Wiley & Sons, Inc., New York,
pp. 777-779, 788, 789.
Courant, R. and Hilbert, D., 1953, Methods of Mathematical Physics, Volume I, Interscience Publishers, Inc.,
z d
dN
A + 0 = 0 z L < <
N L ( ) Q
2
=
z d
dN
z d
z
L

N L ( ) N z ( ) A z d
z
L

= =
N z ( ) Q
2
A L z ( ) + =
U
*
1
2
---
N
2
EA
------- z d
0
L

=
q
2
Q
2

U
*
( )
1
2
---
Q
2

N
2
EA
-------
\ )
[
z d
0
L

N
EA
-------
Q
2

N
\ )
[
z d
0
L

= = =
q
2
Q
2
A L z ( ) + ( )
EA
----------------------------------------- 1 ( ) z d
0
L

=
q
2
L
EA
-------Q
2
L
2
2E
--------- + =
Q
2
0 =
Thin-Walled Structures 63
Problems
New York, pp. 164-274.
Dym, C.L.,1997, Structural Modeling and Analysis, Cambridge University Press, Cambridge, United King-
dom, pp. 73-95.
Dowling, N.E., 1993, Mechanical Behavior of Materials, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, p.
110.
Langhaar, H.L., 1962, Energy Methods in Applied Mechanics, John Wiley and Sons, Inc., New York, pp. 75-
103, 133-136.
Oden, J.T., and Ripperger, E.A., 1981, Mechanics of Elastic Structures, Second Edition, Hemisphere Publish-
ing Corporation, New York, pp. 241-257.
2.13 Problems
1. Take the bar of Example 2.1 to have a solid circular cross section with a diameter of 0.50 inches, and take its
length L = 20 inches. The material is aluminum alloy with a modulus of elasticity , yield
strength in tension and compression of magnitude , and thermal coefcient of expansion
.
a) For the bar subjected to the loads , , , and
, determine the maximum axial displacement, the maximum tensile normal stress,
and the maximum magnitude of the compressive normal stress. (Partial answer: maximum tensile nor-
mal stress is 3,494 psi.)
b) Assume proportional loading such that , , , and
, where is a dimensionless load factor. Determine the largest positive value of
such that normal stress is between the limits and the maximum displacement
is less than 0.0180 inches.
2. A bolt is threaded through a tubular sleeve, and the nut is turned up just tight by hand as shown. Using
wrenches, the nut is then turned further, the bolt being put in tension and the sleeve in compression. If the bolt
has 16 threads per inch, and the nut is given an extra quarter turn (90
0
) by the wrenches, estimate the tensile force
in the bolt if the bolt and sleeve are of steel and the cross-sectional areas are: Bolt area = 1.00 in
2
, Sleeve area =
0.60 in
2
.
E 10
6
10 psi =

yield
47
3
10 psi =
13
6
10 /F =
p
0
265 lb/in. = q
1
0 in. = q
2
0.0060 in. =
T T
0
100 F =
p
0
265 lb/in. ( ) = q
1
0 in. = q
2
0.0060 in. ( ) =
T T
0
100 F ( ) =

z

yield

z

yield

w
max
6 in.
Bars Subjected to Axial Loads
64 Thin-Walled Structures
3. Consider the two term approximation for the displacement of the bar in Example 2.3 on page 34 as given by
eq. (2.42), which is repeated below.
This is a kinematically admissible displacement function with parameters q
1
and q
3
unknown. (Refer to Fig. 2.14
on page 36 and the discussion proceeding it.)
a) Use the principle of virtual work to determine the two equilibrium equations associated with the virtual
displacements q
1
and q
3
. Do not assume a material law has been specied.
b) Now assume the material is linear elastic and solve for the unknown parameters q
1
and q
3
. Note: the fol-
lowing denite integral, with m and n integers, is encountered in the solution.
c) Calculate the percentage error in the strain energy, maximum displacement, and maximum normal force
with respect to the exact solution. The exact solution is given in Example 2.2 on page 28.
4. A uniform bar with axial stiffness EA and length L is xed at each end, and it is subjected to an axial distrib-
uted load , . We seek an approximate solution using the principle of virtual work.
Assume the kinematically admissible axial displacement function
where the unknown parameter q represents the displacement at and at . Use the principle
of virtual work, and assume the material of the bar is linear elastic, to determine
a) The value of parameter q in terms of the given data.
b) The strain energy stored in the bar and the percentage difference with respect to the strain energy for the
exact solution. The strain energy in the exact solution is .
c) The end forces and and the percentage difference with respect to the exact solution. The end
forces in the exact solution are .
w z ( ) q
1

z
L
---
\ )
[
sin q
3
3
z
L
---
\ )
[
sin + = 0 z L
m
z
L
---
\ )
[
cos n
z
L
---
\ )
[
cos z d
0
L

L
2
--- m n =
0 m n
|
|

|
|
=
p
z
z ( ) p
0
2z
L
---------
\ )
[
sin = 0 z L
z w ,
L
p
z
z ( )
EA
Q
1
Q
2
w z ( )
32
3
------q
z
L
---
\ )
[
1
z
L
---
\ )
[
1 2
z
L
---
\ )
[
=
z L 4 = z 3L 4 =
U p
0
2
L
3
( ) 16
2
EA ( ) =
Q
1
Q
2
Q
1
Q
2
p
o
L ( ) 2 ( ) = =
Thin-Walled Structures 65
Problems
5. Each bar in the truss shown has a cross-sectional
area of 1.0 in.
2
, and a modulus of elasticity of 10
7
psi.
There is no change in temperature. Use Castiglianos
rst theorem to nd
a) the horizontal and vertical displacements of
node 1,
b) the stress in psi in each bar, and
c) the horizontal and vertical support reactions
at node 5.
6. The bars in the truss shown have the following
cross-sectional areas: ,
, ,
. The modulus of elasticity of
each bar is 10
7
psi. Compute the vertical displacement
of the right-hand joint using Castiglianos second theo-
rem. Note this truss is statically determinate and all bar
forces can be determined in terms of external load Q.
7. Use Castiglianos second theorem to compute the
horizontal displacement of the right-hand joint of the
previous problem.
8. The plane truss shown represents a single bay of a
wing spar truss. For all bars: and . The cross-sectional areas of the bars are:
2580 mm
2
for the horizontal bars, 387 mm
2
for the vertical bars, and 2690 mm
2
for the diagonal bars. The upper
horizontal bar is heated to above the zero stress temperature, and all other bars remain at the zero stress
temperature. Two 45 kN lift forces act at joints 1 and 2.
Use Castiglianos rst theorem to nd
a) stiffness matrix in kN/mm,
b) displacement of all joints in mm,
1 2
3
4
5
100 in.
10,000 lb.
45
60
45
30
x
y
30 in. 30 in.
40 in.
1
2 3 4
5 6
Q 30 000 lb. , =
A
1
1.0 in.
2
=
A
2
A
4
2.0 in.
2
= = A
3
1 2 in.
2
=
A
5
A
6
3 2 in.
2
= =
E 75 GPa = 23.0
6
10 C =
250C
810 mm
1080 mm
250C
45kN
45kN
1 3
2 4
Bars Subjected to Axial Loads
66 Thin-Walled Structures
c) all boundary reactions in kN, and
d) the stresses in MPa in each bar.

Thin-Walled Structures

67

CHAPTER 3

Axial Normal Stress in
Pure Bending and
Extension

In the rst four sections the exure formula for straight, prismatic beams made from a single homogenous, iso-
tropic, and linear elastic material is developed. The cross section can be asymmetric or symmetric, and the exam-
ples emphasize thin-walled approximations. The exure formula (eq. 3.27) relates the axial normal stress acting
on the beam's cross section to the components of the bending moment. In the last section we consider combined
extension, pure bending, and thermal effects for multi-material beams. The axial normal stress in this latter case
is given by eqs. (3.66-3.68).

3.1 Pure bending

The beam is assumed to be initially straight with uniform cross section and uniform material properties along its
length. The

z

-axis is taken along the length of the beam and the

x

- and

y

-axes are in the cross section. Each end of
the beam is subjected to equal and opposite couples, which are equivalent to a moment denoted by

M

. These cou-
ples act in the plane, called the plane of loading, which is inclined to the

y-z

plane by the angle


as shown Fig.
3.1. The moment of each end couple is directed perpendicular to the plane of loading. Equilibrium dictates that
each cross section is subjected to the applied moment

M

. This loading condition is called pure bending because
the shear force vanishes along the entire length. The bending moment

M

projects on the

x

- and

y

-axes as
and . The sign convention for positive bending moments

M

x

and

M

y

is such that
they tend to cause elongation of the bers parallel to the

z

-axis that have positive

x

and

y

coordinates (rst quad-
rant of the

x-y

plane in the cross section). Thus

M

x

is positive if directed along the positive

x

-axis by the right-
hand screw rule, and

M

y

is positive if directed along the negative

y

-axis by the right-hand screw rule.
M
x
M cos = M
y
M sin =

Axial Normal Stress in Pure Bending and Extension

68

Thin-Walled Structures

3.2 Geometry of deformation

Under the special conditions of pure bending described above, the assumptions of the deformed geometry of the
beam are listed in the table below.

Plane of bending

The plane in which the beam bends is denoted the



plane where the -axis is inclined
by the angle


from the

y

-axis as shown in Fig. 3.2. Note that the plane of bending is not assumed to coincide
with the plane of loading; i.e.,


. Coordinate axes and are rotated through the angle


relative to the

x


and

y

coordinate axes as shown in the gure. A point in the cross section can be located by

x

and

y

coordinates
and by and coordinates. Consider a typical point

A

with coordinates

x

and

y

. The and coordinates of
point

A

are determined by the geometry shown in Fig. 3.3. From Fig. 3.3, the relationship between these coordi-
nates is

(3.1)

The cosines of the angles between the positive coordinate axes, or direction cosines, can be written in tabular
form as shown in Fig. 3.2. For example, the direction cosine of the angle between the -axis and the y-axis is
. Note that the sum of the squares of the direction cosines in a row or column equals
unity. Equations (3.1) show that the rows of the direction cosine table are associated with the - coordinates.
The inverse of eqs. (3.1) is



Geometry of deformation in pure bending

1. The beam deforms into a plane circular arc, and
2. That plane sections originally perpendicular to the

z

-axis remain plane and perpendicular
to the

z

-axis in the deformed beam.
A
A
M sin
z
y
M cos
M sin
M cos
x
y
M
x
M
y
M
plane of loading

Section A-A
Fig. 3.1 A uniform beam subjected to equal and opposite couples at each end.
y z y
x y
x y x y
x x cos y sin =
y x sin y cos + =
x
90 + ( ) cos ( ) sin =
x y

Thin-Walled Structures

69

Geometry of deformation

Hence, the columns of the direction cosine table are associated with the

x -y

coordinates

Axial normal strain

The particles on the

z

-axis in the undeformed beam displace to a circular arc of radius



in
the plane of bending, which we call the

z

-curve. See Fig. 3.4. If the beam is considered to be made up of longitu-
dinal bers, then bers parallel to the

z

-axis in the undeformed beam deform into concentric circular arcs. Some
of these parallel bers are elongated and some are shortened. Consequently, the bers on some surface between
the top and bottom surfaces of the beam, called the neutral surface, will remain the same length. Although we do
not know the location of the neutral surface at this point, we take the

z

-axis to lie in the neutral surface of the
x
y
M
plane of loading

y
x
plane of bending
Fig. 3.2 Projection of the plane of loading and the plane of bending onto
the cross section of the beam.
cos
cos sin
sin
x
y
x y
direction cosines
x cos
y sin
x

x
y
x sin
y cos
y
A
90
90
90
90
Fig. 3.3 Point A located in two Cartesian
coordinate systems, where one system is
rotated through angle relative to the other.
x x cos y sin + =
y x sin y cos + =
Axial Normal Stress in Pure Bending and Extension
70 Thin-Walled Structures
undeformed beam. We will show later that the neutral surface in the undeformed beam coincides with the -z
plane.
An element in the undeformed beam of length z bounded by two parallel line segments and is
shown in Fig. 3.5. Line is part of the z-axis and line is characterized by a constant -coordinate value.
In the deformed geometry these line segments map to concentric circular arcs and . Let denote the
rotation of a line element originally parallel to the -direction in the undeformed beam, positive clockwise.
Since cross sections remain perpendicular to the neutral surface in the deformed beam, the angle is also equal
to the rotation of the line element coinciding with z-axis. In Fig. 3.5, the line is shown to rotate counter-
clockwise which corresponds to a , and line is shown to rotate clockwise through angle .
Then, the included angle between sections and is . The radius of the arc is as described
in the previous paragraph. It is assumed that the distortion of the cross section in the plane of bending is negligi-
x

z-curve in the neutral surface.


z
y
Fig. 3.4 Beam deformed into a circular arc in the plane of bending.
PQ
)
RS
)
y
y
z
z
z
P
Q
R S

+

P
*
Q
*
R
*
S
*
y
Fig. 3.5 Deformation of an element in the plane of bending.

PQ
)
RS
)
y
P
*
Q
*
)
R
*
S
*
)
y
P
*
R
*
)
Q
*
S
*
)
+
P
*
R
*
)
Q
*
S
*
)
P
*
Q
*
)
Thin-Walled Structures 71
Geometry of deformation
ble, so that the distance between the two circular arcs and remains . Hence, the radius of arc
is . The strain of line element is dened as
(3.2)
From Fig. 3.5
(3.3)
and since line element is in the neutral surface . Thus, eq. (3.2) becomes
(3.4)
Also, we have , since the line element is in the neutral surface. From this relation we
have in the limit as and , the following equation for the curvature of the z-curve in the deformed
beam
(3.5)
where the prime denotes a derivative with respect to z. Equation (3.5) is consistent with the denition of curva-
ture as the change in the slope angle of the curve with respect to its arc length. Hence,
(3.6)
Equation (3.6) shows that the normal strain of a line element parallel to the z-axis is proportional to the distance
from the z-axis times the curvature. Fibers above the z-axis ( ) are stretched ( ) if the beam is bent
convex side up ( ). Fibers below the z-axis are shortened ( ) if the beam is bent convex side up (Fig.
3.4).
Projection of the z-curve onto the x-z and y-z planes Now we need to express the normal strain given by eq.
(3.6) in terms of the x-y coordinate system. This is accomplished by projecting the z-curve in the deformed beam,
which lies in the plane, onto the coordinate planes x-z and y-z. The z-curve is represented in the x-y-z coor-
dinate system by introducing displacement functions that represent the x-, y-, and z-direc-
tion displacements, respectively, of a particle at point P on the z-axis. Under the deformation the particle
originally at P with coordinates displaces to the point P* with coordinates . See
Fig. 3.6.
Also the displacements of the particle at P can be represented in the - -z coordinate system. Let
denote the displacements in the - and -directions, respectively, of the particle at point at P. These dis-
placement components are related to components and by the considering the position of the projec-
tion of P* onto the x-y plane. The projection of P* onto the x-y plane is labeled point in Fig. 3.6 and Fig.
P
*
Q
*
)
R
*
S
*
)
y
R
*
S
*
)
y + RS
)

z
R
*
S
*
RS
RS
------------------------- =
))
)
R
*
S
*
y + ( ) =
)
P
*
Q
*
)
P
*
Q
*
PQ R S = = =
)))

z
R
*
S
*
P
*
Q
*

P
*
Q
*
--------------------------------
y + ( )

-----------------------------------------
y

--- = = =
))
)
P
*
Q
*
z = =
)
P
*
Q
*
)
z 0 0
1

---
d
dz
------ = =

z
y =
y 0 >
z
0 >
1 0 >
z
0 <
y z
u z ( ) v z ( ) and w z ( ) , ,
0 0 z , , ( ) u z ( ) v z ( ) z w z ( ) + , , ( )
x y u z ( )
v z ( ) x y
u z ( ) v z ( )
P
3
*
Axial Normal Stress in Pure Bending and Extension
72 Thin-Walled Structures
3.7. From Fig. 3.7, the displacement components in the - coordinates are related to those in the x-y coordi-
nates by
(3.7)
which is of the same form as the coordinate transformations given by eq. (3.1) with the coordinates replaced by
their respective displacements. Since the z-curve lies in the plane, displacement for all z. That is,
point is in the -z plane. Also, the angle is independent of the z-coordinate, since the beam is assumed to
deform into a plane circular arc. Next we relate the rotation, or slope, of an element of the z-curve in the plane of
bending to the rotations its projections onto the x-z and y-z planes.
x
y
z
u z ( )
v z ( )
w z ( )
P: 0 0 z , , ( )
P
*
P
3
*
P
2
*
P
1
*

y
Q
*
Fig. 3.6 Displacements of the particle from P to P*, and the rotations of the
projections of a line element P*Q*.
x
y

y
x
plane of bending
u
u
v
v
P
P
3
*
Fig. 3.7 Lateral displacement components of the particle at point P.
x y
u u cos v sin =
v u sin v cos + =
y z u 0 =
P
3
*
y
Thin-Walled Structures 73
Geometry of deformation
The displacement and rotation of an innitesimal line element on the z-axis to on the z-curve is
shown in Fig. 3.8. Since the line element does not change length we have by the Pythagorean theorem in the
deformed state that . Hence, in the limit as , the derivative of the axial dis-
placement is related to the derivative of the lateral displacement by
(3.8)
As is shown in Fig. 3.8, the rotation, or slope, of the z-curve is related to displacement by
(3.9)
where we assume the angle is small so that the sine of it can be replaced by the angle itself in radians. Differ-
entiating eqs. (3.7) with respect to z, and noting that and the angle is constant for all z, we get
(3.10)
The derivatives of the displacement components in the x - and y -coordinate directions are directly related to
the rotations of the projections of the z-curve onto the x-z and y-z planes. The projection of point P* on the onto
the y-z plane is labeled , and its projection onto the x-z plane is labeled , as is shown in Fig. 3.6. Points
and and the projections of the differential line element along the z-curve onto the coordinate planes shown in
Fig. 3.6 are also shown in Fig. 3.9. From the differential geometry of Fig. 3.9, we have in the limit as
that
(3.11)
Since is assumed small with respect to unity, eq. (3.8) yields that is also small with respect to unity. Thus,
we neglect with respect to one in the denominators of eqs. (3.11), and approximate the rotations of the projec-
tions as
(3.12)
PQ
)
P
*
Q
*
)
z
y
z
*
v
w
v dv +
w dw +

dz
z z dz +
P Q
P
*
Q
*
dv
z
*
dz
*
+
dz dw +
Fig. 3.8 Displacement and rotation of an innitesimal line element on the z-axis
in the plane of bending.
dz ( )
2
dz dw + ( )
2
dv ( )
2
+ = dz 0
1 w + ( )
2
v ( )
2
+ 1 =
v
sin
dv
dz
------ v = =

u 0 =
0 u cos v sin =
u sin v cos + =
P
1
*
P
2
*
P
1
*
P
2
*
dz 0

x
tan
v
1 w +
--------------- =
y
tan
u
1 w +
--------------- =
v w
w

x
v =
y
u =
Axial Normal Stress in Pure Bending and Extension
74 Thin-Walled Structures
where the tangent of a small angle is approximated by the angle itself in radians. Thus, eqs. (3.10) become
(3.13)
which are the relations between the rotation of the z-curve to the rotations of its projections onto the coordinate
planes. Solving these equations for the rotations in the x-y coordinate system, we get
(3.14)
Equations (3.14) show that for small rotations, the rotations of the projected curves are simply the components
of the rotation of the z-curve .
Normal strain in the x-y-z system Finally, the strain given by eq. (3.6) is written in terms of the x-y system by
substituting the second of eqs. (3.1) for , and the derivative of the second of eqs. (3.13) for , to get
Rearrange this equation to
Differentiating the rst of eqs. (3.13), recalling that angle is independent of z, results in the relation
Now use this result in the previous equation to get the strain expressed as
(3.15)
where the curvatures of the projected curves are determined from eqs. (3.12) to be
(3.16)

x
v
v dv +
z
*
dz
*
+ z
*
dz dw +
dv
y
projection in y-z plane
z
P
1
*

y
u
z
x
projection in x-z plane
dz dw + P
2
*
du
u du +
z
*
z
*
dz
*
+
Fig. 3.9 Rotations of the projections of the z-curve onto the cartesian planes.
0
y
cos
x
sin + =

y
sin
x
cos + =

y
sin =
x
cos =
y

z
x sin y cos + ( )
y
sin
x
cos + ( ) =

z
x
y
sin
2

x
sin cos + ( ) y
y
cos sin
x
cos
2
+ ( ) + =

y
cos
x
sin =

z
x
y
y
x
+ =

y
u =
x
v =
Thin-Walled Structures 75
Bending normal stress exure formula
3.3 Bending normal stress exure formula
The strain is related to the normal stress acting on the cross section by the
material law, and we consider the beam to be made from a linear elastic, isotropic
material. For uniaxial loading only, this material law is given by eq. (2.5) on p. 22 .
Under multi-axial loading the material law is obtained by superpostion of the normal
strains for line elements parallel to the x-, y-, and z-directions due to
the combined action of normal stresses acting on faces of an innites-
imal element normal to to the x-, y-, and z-directions. Superposition of normal strains using eq. (2.5) on p. 22
yields
(3.17)
where E is the modulus of elasticity and is Possions ratio. The material law for a linear elastic, isotropic mate-
rial is called Hooke's law. Material properties E and which vary from point to point in the cross section (x-y
plane) are permissible, but the material properties are assumed uniform along the z-axis. Thus, a laminated beam
of more than one material may be considered, which is addressed in Section 3.5. However, in this section we
limit consideration to a beam made of a single homogeneous material, so that E and are independent of the spa-
tial coordinates x, y and z. Now we make the third basic assumption of beam theory:
Neglecting stress components and , Hooke's law becomes
(3.18)
Eliminating the strain in eq. (3.18) via eq. (3.15) we get
(3.19)
The distribution of normal stresses given by eq. (3.19) over the cross section is, in general, statically equiva-
lent to the resultants N, M
x
, and M
y
as shown in Fig. 3.10. The axial load N, and bending moments M
x
and M
y

are related to the normal stress by
(3.20)
where dA = dxdy is an area element in the cross-sectional area denoted by A. Substituting eq. (3.19) into eq.
(3.20) and integrating over the cross section we obtain
Assumption regarding the material law
3. Lateral normal stresses and are assumed small with respect to the axial normal
stress , and hence are neglected in Hookes law.

z
x
y
z

z

z

x

y
and
z
, ,

x

y
and
z
, ,

z
1
E
---

E
---

E
---

E
---
1
E
---

E
---

E
---

E
---
1
E
---

z
=

x

y

y

y

z
E
z
=

z
E x
y
y
x
+ ( ) =
N
z
A d
A

= M
x
y
z
A d
A

= M
y
x
z
A d
A

=
Axial Normal Stress in Pure Bending and Extension
76 Thin-Walled Structures
(3.21)
in which we dened
(3.22)
and
(3.23)
In obtaining the results shown in eqs. (3.21), the modulus of elasticity E was brought outside the area inte-
grals since it is independent of x and y for a homogeneous material cross section. The integrals in eqs. (3.22) and
(3.23) are purely geometric quantities. They reect how the area is distributed with respect to the x- and y-coordi-
nate axes. The quantities Q
x
and Q
y
in eqs. (3.22) are called rst area moments since the power of the coordi-
nate in the integrand is unity, while the I
xx
and I
yy
in eqs. (3.23) are called second area moments since the power
of the coordinates in the integrand is two. The quantity I
xy
is called the product area moment.
For the case of pure bending, only moments M
x
and M
y
are non-zero, so the axial force N must vanish. Set-
ting N = 0 in the rst of eqs. (3.21) implies that
(3.24)
The rst area moments vanish if the x-axis and the y-axis pass through the centroid of the cross-sectional area.
This locates the z-axis; i.e., the z-axis of the beam passes through the centroid of each cross section. For centroi-
dal coordinates in the cross section, the moment-curvature relationship of eq. (3.21) can be written in the matrix
form

z
dA
x
y
z
N
M
x
M
y
z
x
y
Fig. 3.10 Statical equivalence of the normal stress to the beam resultants.

N EQ
x

x
EQ
y

y
+ =
M
x
EI
xx

x
EI
xy

y
+ =
M
y
EI
xy

x
EI
yy

y
+ =
Q
x
y A d
A

= Q
y
x A d
A

=
I
xx
y
2
A d
A

= I
yy
x
2
A d
A

= I
xy
xy A d
A

=
N 0 = ( ) Q
x
Q
y
0 = =
Thin-Walled Structures 77
Bending normal stress exure formula
(3.25)
where the second are moments are computed for centroidal coordinates in the cross section. The curvature-
moment relationship is the inverse of this, or
(3.26)
Equations (3.25) and (3.26) are regarded as the material law for beam bending, and these equations are valid only
if the x- and y-axes pass through the centroid of the cross section.
The formula for the bending normal stress
z
in eq. (3.19) can be written in terms of the bending moments
M
x
and M
y
by using the curvature-moment relation in eq. (3.26). Substituting the curvatures from eq. (3.26) into
eq. (3.19) we get
(3.27)
or we can re-write this as
(3.28)
Equation (3.27), or its equivalent eq. (3.28), is called the exure formula, since it determines the normal stress at
cross-sectional coordinates x and y due to bending. Notice that the bending normal stress is zero at the origin,
which coincides with the centroid of the cross section, and that the bending normal stress varies linearly in x and
y over the cross section. Thus, the extreme values of the bending normal stress occur somewhere on the boundary
of the cross section where coordinates x and y can attain their largest values.
The plane of bending of the beam can be determined in terms of the bending moment components M
x
and
M
y
by rst differentiating eqs. (3.14) with respect to z and recognizing that the z-curve lies in the plane of bend-
ing (angle is independent of z for pure bending). Second, solve the resulting equations for to get
(3.29)
Now use eq. (3.26) to eliminate the curvatures in eq. (3.29) to nd
(3.30)
Thus, knowing the bending moment components and the second area moments of the cross section, the angle ,
and hence the plane of bending, is determined by eq. (3.30). The axis normal to the plane of bending through the
centroid is labeled as the axis in Fig. 3.2. On the axis , so that from strain-curvature relation given
by eq. (3.6) the normal strain on the axis. Also, from Hookes law this means that the bending normal
M
x
M
y
E
I
xx
I
xy
I
xy
I
yy

y
=

y
1
E I
xx
I
yy
I
xy
2
( )
-----------------------------------
I
yy
I
xy

I
xy
I
xx
M
x
M
y
=

z
I
xy
M
x
I
xx
M
y
+
I
xx
I
yy
I
xy
2

--------------------------------------
\ )
[
x
I
yy
M
x
I
xy
M
y

I
xx
I
yy
I
xy
2

-----------------------------------
\ )
[
y + =

z
I
yy
y I
xy
x
I
xx
I
yy
I
xy
2

--------------------------
\ )
[
M
x
I
xx
x I
xy
y
I
xx
I
yy
I
xy
2

--------------------------
\ )
[
M
y
+ =
tan

y

x
------- =
tan
I
xy
M
x
I
xx
M
y
+
I
yy
M
x
I
xy
M
y

---------------------------------------- =
x x y 0 =

z
0 = x
Axial Normal Stress in Pure Bending and Extension
78 Thin-Walled Structures
stress on the axis. Since on the axis, the axis is called the neutral axis . The particles in
-z plane map to the neutral surface in the deformed beam. The equation for the neutral axis in the cross section
is determined from the condition that in the rst of eqs. (3.1). We get
(3.31)
where the tangent of is determined by eq. (3.30).
The neutral axis passes through the centroid of the cross section, since the centroid was used as the origin of
the x-y system. If the axial force in addition to non-zero bending moments, then the axis of zero stress, or
the neutral axis, will no longer pass through the centroid, but will be parallel to the line obtained from eq. (3.31).
From Fig. 3.1, recall that the plane of loading is located by
Using this result in eq. (3.30), we get
(3.32)
Equation (3.32) shows that, in general, the plane of loading and the plane of bending do not concide ( ). The
plane of loading and the plane of bending coincide if , or , or
.
EXAMPLE 3.1 Bending normal stress distribution in a cantilever beam with a thin-walled zee section.
The cantilever beam shown in Fig. 3.11 is subjected to a bending moment M at its tip. The cross-sectional
dimensions a and t are considered known with 0 < t/a << 1; i.e., this is a thin-walled section. Note that only the
center line of the wall is drawn in the sketch of the cross section and not the thickness, since the thickness is
small. The wall center line is called the contour of the cross section. Given that the second area moments about
the centroidal axes x and y are
(3.33)
determine the neutral axis in the cross section and the distribution of the bending normal stress.
Solution First, note that the components of the bending moment are and . Thus, the plane
of the couple whose moment is M is the y-z plane, or the angle for the plane of loading shown in Fig.
3.1. From eq. (3.30)
and from eq. (3.31) the equation of the neutral axis in the cross section is .

z
0 = x
z
0 = x x
x
y 0 =
0 x sin y cos + ( )
NA
= or y
NA
x
NA
tan =
N 0
tan
M
y
M
x
------- =
tan
I
xy
I
xx
tan +
I
yy
I
xy
tan
------------------------------------ =

I
xy
0 and M
x
0 = = I
xy
0 and M
y
0 = =
I
xy
0 and I
xx
I
yy
= =
I
xx
8
3
---a
3
t = I
yy
2
3
---a
3
t = I
xy
a
3
t =
M
x
M = M
y
0 =
=
tan
I
xy

I
yy
---------
a
3
t ( )
2
3
---a
3
t
------------------
3
2
--- = = = so 56.31 =
y
NA
3
2
--- x
NA
=
Thin-Walled Structures 79
Bending normal stress exure formula
The exure formula, eq. (3.28), for this example becomes
We plot the bending normal stress on the contour only. Coordinates x and y are related on the contour. That is,
The neutral axis and the bending normal stress distribution are shown in Fig. 3.12.
x
y
a
t, typical
a
a
a
centroid
C
z
y
M
Fig. 3.11 Pure bending of a cantilevered beam with a zee cross section
Cross Section

z
2
3
---a
3
t
\ )
[
y a
3
t ( )x
8
3
---a
3
t
\ )
[
2
3
---a
3
t
\ )
[
a
3
t ( )
2

--------------------------------------------------------- M ( ) 6y 9x + ( )
M
7a
3
t
---------- = =
top flange: a x 0 y a =
web: x 0 = a y a
bottom flange: 0 x a y a =
Axial Normal Stress in Pure Bending and Extension
80 Thin-Walled Structures
EXAMPLE 3.2 Lateral displacements of the zee section beam
Determine the lateral displacement functions u(z) and v(z), 0 z L, for the zee section beam of Example 3.1.
Solution First note that this a statically determinate problem. So the equilibrium equations are satised for
and for all , where M denotes the specied end moment. To nd the lateral dis-
placements and , , we begin with the curvature-moment relations given by eqs. (3.26). These
equations are repeated below.
From equilibrium we have and for . From the given data for the second area
moments we know , , and . Thus,
(3.34)
Substitute these data into curvature-moment relations above to get
(3.35)
Integrate these equations with respect to z once to get
x
y
y
NA
3
2
--- x
NA
=
3
7
---
M
a
2
t
-------
3
7
---
M
a
2
t
-------
6
7
---
M
a
2
t
-------
6
7
---
M
a
2
t
-------

z
M
Fig. 3.12 Bending normal stress
distribution along the contour of the
zee section
M
x
M = M
y
0 = z 0 L , ( )
u z ( ) v z ( ) 0 z L

y
1
E I
xx
I
yy
I
xy
2
( )
-----------------------------------
I
yy
I
xy

I
xy
I
xx
M
x
M
y
=
M
x
M = M
y
0 = 0 z L
I
xx
8
3
---a
3
t = I
yy
2
3
---a
3
t = I
xy
a
3
t =
det I ( ) I
xx
I
yy
I
xy
2

8
3
---
2
3
--- 1 ( )
2
a
6
t
2
7
9
---a
6
t
2
= = =

x
6
7
---
M
Ea
3
t
----------- =
y
9
7
---
M
Ea
3
t
----------- =
Thin-Walled Structures 81
Bending normal stress exure formula
(3.36)
The constants due to indenite integration are determined from the boundary conditions. At z = 0 the beam is
clamped to the rigid wall. Hence the cross section at z = 0 is prevented from rotation, and this means
and . Substituting these boundary conditions into eqs. (3.36), we nd and . Hence,
the rotations are
(3.37)
The rotation-displacement relations are given in eqs. (3.12), and are repeated below.
Substitute these relations into eqs. (3.37) to get
(3.38)
Integrate these equations with respect to z to get
(3.39)
The constants of indenite integration are evaluated by the boundary conditions. Since the beam is xed to the
rigid wall at z = 0, we have and . Substitute these boundary conditions into eqs. (3.39) to
nd that and . Hence the lateral displacements of the beam are
(3.40)
A plot of the z-axis in the deformed beam, or z-curve, is shown in Fig. 3.13. In the plot the scaled lateral dis-
placements are dened by

x
6
7
---
M
Ea
3
t
-----------
\ )
[
z c
1
+ =
y
9
7
---
M
Ea
3
t
-----------
\ )
[
z c
2
+ =

x
0 ( ) 0 =

y
0 ( ) 0 = c
1
0 = c
2
0 =

x
6
7
---
M
Ea
3
t
-----------
\ )
[
z =
y
9
7
---
M
Ea
3
t
-----------
\ )
[
z = 0 z L

x
v =
y
u =
v
6
7
---
M
Ea
3
t
-----------
\ )
[
z = u
9
7
---
M
Ea
3
t
-----------
\ )
[
z =
v
3
7
---
M
Ea
3
t
-----------
\ )
[
z
2
c
3
+ = u
9
14
------
M
Ea
3
t
-----------
\ )
[
z
2
c
4
+ =
v 0 ( ) 0 = u 0 ( ) 0 =
c
3
0 = c
4
0 =
v
3
7
---
M
Ea
3
t
-----------
\ )
[
z
2
= u
9
14
------
M
Ea
3
t
-----------
\ )
[
z
2
= 0 z L
Axial Normal Stress in Pure Bending and Extension
82 Thin-Walled Structures
The view of the lateral displacements at the end of the beam are shown in Fig. 3.14.
v v
Ea
3
t
ML
2
-----------
\ )
[
= u u
Ea
3
t
ML
2
-----------
\ )
[
=
0
0.25
0.5
0.75
1
zL
0
0.2
0.4
0.6
u

0
0.1
0.2
0.3
0.4
v

0
0.25
0.5
0.75
1
zL
z-curve
Fig. 3.13 The z-axis in the deformed, zee section beam .
x
y
x
y
56.3 =
u L ( )
v L ( )
M
Fig. 3.14 Displacement of the tip of the cantilevered, zee section beam.
Thin-Walled Structures 83
Moments of areas
3.4 Moments of areas
First and second area moments of the cross-sectional area need to be determined before using the exure for-
mula. The most frequently used methods in the computation of the moments of areas are the parallel axis theo-
rem and the composite body technique. Also, the geometric approximations appropriate for thin-walled sections
are introduced and discussed in the examples of this section.
Parallel Axis Theorem Consider two parallel axes systems
in the cross section. The origin of the cartesian axes x and y
coincide with the centroid of the cross-sectional area, which
is labeled C in Fig. 3.15. The second cartesian system and
has its origin at an arbitrary point O, the axis is parallel
to the x axis, and the axis is parallel to the y axis. The
and system should not be confused the rotated system
introduced earlier in the discussion of the plane of bending
(Fig. 3.2). The and system in this section is not rotated
relative to the x and y system. The location of the centroid in
the and system is denoted by coordinate values
. Usually the and system is selected as some-
thing convenient to start with, and the rst and second area
moments with respect to the and system are computed or
looked-up in tables. Then the coordinates of the centroid are computed and the parallel axis theorem is
used to nd the second area moments in the x and y system.
In the and system, the rst area moments are dened as
(3.41)
in which the area element . The relationship between the two parallel coordinate systems is deter-
mined from the location of a generic point in the plane in each system. That is
(3.42)
The relationship between the area elements is where , since the values are
xed. If eqs. (3.42) are substituted into eqs. (3.41), we get
(3.43)
where
(3.44)
Since the origin of the x and y system is at the centroid, the rst moments Q
x
and Q
y
are zero by denition. Set-
ting and in eqs. (3.43), we can solve to nd the location of the centroid as
x
y
y
x
C
O
y
c
x
c
y y
c
y + =
x x
c
x + =
Fig. 3.15 Parallel cartesian axes systems .
x
y x
y x
y
x y
x y
x
c
y
c
, ( ) x y
x y
x
c
y
c
, ( )
x y
Q
x
y A d
A

= Q
y
x A d
A

=
dA dxdy =
x x
c
x + = y y
c
y + =
dA dA = dA dxdy = x
c
y
c
, ( )
Q
x
y
c
A Q
x
+ = Q
y
x
c
A Q
y
+ =
A A d
A

= Q
x
y A d
A

= Q
y
x A d
A

=
Q
x
0 = Q
y
0 =
Axial Normal Stress in Pure Bending and Extension
84 Thin-Walled Structures
(3.45)
In the and coordinate system the second area moments are
(3.46)
Second area moments are often called moments of inertia in analogy to moments of inertia of mass elements
used in rigid body dynamics. The fact that eqs. (3.46) are second moments of area elements and not mass ele-
ments should be kept in mind even if the terminology moments of inertia is used in the context of beam bend-
ing. Now substitute eqs. (3.42) for the and coordinates into eqs. (3.46) to get
(3.47)
where the second area moments in the x and y system are dened as
(3.48)
Since the x and y coordinates are centroidal, Q
x
= Q
y
= 0, and eqs. (3.47) reduce to
(3.49)
Equations (3.43) and (3.47) are the generalized parallel axis theorem, but in problem solving we use eqs. (3.43)
to nd the centroid and then the parallel axis theorem reduces to the use of eqs. (3.49). Note that eqs. (3.48) show
that the I
xx
and I
yy
are always positive in value with dimensional units of L
4
. The product area moment I
xy
can be
positive, zero, or negative in value. The product area moment I
xy
is zero if either the x axis or y axis is a axis of
symmetry of the cross section.
Radius of gyration It is common with respect to the second area moments I
xx
and I
yy
to dene radii of gyration
by the denitions
(3.50)
The radii of gyration r
x
and r
y
have dimensional units of length. However the radii of gyration do not locate a
physically signicant point in the cross section. For example, , where it the radius of gyration
with respect to the -axis. (Using the parallel axis theorem, the relation between the radius of gyration about the
-axis to the x-axis is .)
Approximations in thin-walled sections For thin-walled sections, the thickness t of the wall is much smaller
than the largest dimension in the cross section. This geometry permits simplications with respect to computing
the rst and second area moments without loss of signicant accuracy with respect to an exact computation. As
an example we model the thin-walled zee section with thin rectangular elements for the web and anges as is
x
c
Q
y
A
------ = y
c
Q
x
A
------ =
x y
I
xx
y
2
A d
A

= I
yy
x
2
A d
A

= I
xy
xy A d
A

=
x y
I
xx
y
c
2
A 2y
c
Q
x
I
xx
+ + =
I
yy
x
c
2
A 2x
c
Q
y
I
yy
+ + =
I
xy
x
c
y
c
A x
c
Q
x
y
c
Q
y
I
xy
+ + + =
I
xx
y
2
A d
A

= I
yy
x
2
A d
A

= I
xy
xy A d
A

=
I
xx
y
c
2
A I
xx
+ = I
yy
x
c
2
A I
yy
+ = I
xy
x
c
y
c
A I
xy
+ =
r
x
I
xx
A = r
y
I
yy
A =
r
x
y
c
r
x
+ r
x
x
x r
x
2
y
c
2
r
x
2
+ =
Thin-Walled Structures 85
Moments of areas
shown in Fig. 3.16 The gap and overlap at the corners are ignored. Instead of an actual drawing of the zee sec-
tion, or a sketch of the mathematical representation as shown in Fig. 3.16, we merely draw the contour of the
section to represent it. The contour consists of the locus of points in the middle of the wall for each branch of the
cross section. The contour for the zee section is illustrated in Fig. 3.11. The term branch is used to mean either
a ange or web of the zee section. The slope of the contour is continuous within the branch. At the junctions
between branches, like the web-ange junctions of the zee section, the contour usually exhibits a discontinuous
slope. The contour may be a curve in the x-y plane for more complex sections.
For branches that can be represented by a thin-walled
rectangular area, we can obtain simple formulas for the
second area moments. Consider a thin rectangular area,
where 0 < t << l, represented by a straight line contour
inclined at a angle as is shown in Fig. 3.17. The con-
tour coordinate is denoted by s, and the area element is
. The x and y coordinates of the point s on
the contour are given by and .
Hence, the second area moments are computed from
(3.51)
t
h
Fig. 3.16 Modeling a thin-walled zee section with rectangular elements for the web and anges
0
t
h
--- 1 <
x
y

l
t
ds
s
l/2
Fig. 3.17 A thin rectangular area inclined with
respect to the centroidal coordinates.
dA tds =
x s cos = y s sin =
I
xx
s sin ( )
2
t s d
l 2
l 2

l
3
t
12
------ sin
2
= =
I
yy
s cos ( )
2
t s d
l 2
l 2

l
3
t
12
------ cos
2
= =
I
xy
s sin ( ) s cos ( )t s d
l 2
l 2

l
3
t
12
------ sin cos = =
Axial Normal Stress in Pure Bending and Extension
86 Thin-Walled Structures
Composite area technique The composite body technique for
computing the centroid and second area moments is a method
applicable to cross-sectional areas that can be subdivided into
simple geometric shapes whose properties are known. For our
example of the thin-walled zee section, we can subdivide it into
three rectangular areas and use the formulas in eqs. (3.51) for
the second area moments about the local centroidal axis system
in each rectangle. With the centroid and second area moments
known for each sub-area, the parallel axis theorem is used to
transfer their properties to a parallel reference system. Then a
summation of the rst and second moments for each sub-area
about the reference axis system determines the total properties
for the cross section. The pertinent equations for an area subdi-
vided into N sub-areas labeled (see Fig. 3.18)
are
(3.52)
(3.53)
(3.54)
(3.55)
where are the known properties of the i-th sub-area. The overall centroid of the area
is computed from the last two formulas in eqs. (3.52). The second area moments for the reference system deter-
mined from eqs. (3.53) to (3.55) are then used with the coordinates of the overall centroid to compute the second
area moments about the centroidal x and y axes by the parallel axis theorem. This method is best illustrated by
example.
EXAMPLE 3.3 Thin-walled zee section properties by the composite body technique
Determine the centroid and the second area moments for the thin-walled zee section of Example 3.1. The section
is sub-divided into three rectangular branches as shown in Fig. 3.19. One branch corresponds to the web and two
branches correspond to the anges.
Solution First we nd the centroid. Equations (3.52) are represented in the table shown below. Summation of the
appropriate columns gives
x
i
x
i
y
i
y
i
C
i
1
2
N
x
y
i-th sub-area
Fig. 3.18 Composite body technique
i 1 2 N , , , =
A A
i
i 1 2 , , =
N

= Ay
c
y
i
A
i
i 1 2 , , =
N

= Ax
c
x
i
A
i
i 1 2 , , =
N

=
I
xx
I
xx
y
2
A + ( )
i
i 1 2 , , =
N

=
I
yy
I
yy
x
2
A + ( )
i
i 1 2 , , =
N

=
I
xy
I
xy
xyA + ( )
i
i 1 2 , , =
N

=
A
i
x
i
y
i
I
xx
( )
i
I
yy
( )
i
I
xy
( )
i
, , , , ,
A 4at = x
c
A 0 = y
c
A 4a
2
t =
Thin-Walled Structures 87
Moments of areas
so that the centroid has coordinates and .
The second area moments are computed for the reference coordinate system using the second table
shown below. Note that for the local centroidal coordinate systems in each branch we can identify the angle in
eqs. (3.51) as , , and . These values of the angle in each branch are used to com-
pute the local second area moments in each rectangular branch via eqs. (3.51). From the summation of the col-
umns, the second area moments in the system via eqs. (3.53) to (3.55) are
1 at a/2 0
a
2
t/2
0
2 2at 0 a 0
2a
2
t
3 at a/2 2a
a
2
t/2 2a
2
t
Sum 4at 0
4a
2
t
1 0 0
a
3
t/4 a
3
t/12
0 0
2
(a
2
)2at (2a)
3
t/12
0 0 0 0
3
(2a)
2
at
0
a
3
t/4 a
3
t/12 a
3
t
0
Sum
6a
3
t 2a
3
t/3 a
3
t/2 a
3
t/6 a
3
t
0
x
1
y
1
x
2
y
2
x
3
y
3
a 2
a
a
a
a 2
a
x
y
Fig. 3.19 Zee section approximated by three rectangular areas.
x
y y ,
x
y
c
a =
C
O
O
t = thickness of all branches
x
c
0 = y
c
a =
x y , ( )
i
A
i
x
i
y
i
x
i
A
i
y
i
A
i

1
0 =
2
90 =
3
0 =
i
y
i
2
A
i
I
xx
( )
i
x
i
2
A
i
I
yy
( )
i
x
i
y
i
A
i
I
xy
( )
i
x y , ( )
Axial Normal Stress in Pure Bending and Extension
88 Thin-Walled Structures
Now we use the parallel axis theorem to transfer these moments to the centroidal system. Equations (3.49) give
These are the second area moments given the problem statement of Example 3.1.
EXAMPLE 3.4 Semicircular section with two stringers
The beam cross section consists of a thin-walled semicircular web of radius a and thickness t, and two stringers
each with area . The section and the reference coordinate system are shown Fig. 3.20. Determine
the location of the centroid and the second area moments.
Solution Since this cross section is symmetric with respect to the axis, the centroid is located on this axis, and
the product area moment is zero. In thin-walled construction the stringers cross-sectional dimensions are
small with respect to the largest dimension of the cross section. Hence, the stringer is modeled by its area A
s
con-
centrated at the stringers centroid. Also, the second area moments of the stringer area are neglected with respect
to the transfer terms in the parallel axis theorem. It is convenient to use the polar angle in the moment compu-
tations for the web, where , as is shown in the gure. The differential area of the web is
, and the coordinates to this differential area are and . The
cross-sectional area is
I
xx
6a
3
t 2a
3
t 3 + 20a
3
t ( ) 3 = =
I
yy
a
3
t ( ) 2 a
3
t ( ) 6 + 2a
3
t ( ) 3 = =
I
xy
a
3
t =
I
xx
I
xx
y
c
2
A
20
3
------a
3
t a
2
4at ( )
8
3
---a
3
t = = =
I
yy
I
yy
x
c
2
A
2
3
---a
3
t 0 ( ) 4at ( )
2
3
---a
3
t = = =
I
xy
I
xy
x
c
y
c
A a
3
t 0 ( ) a ( ) 4at ( ) a
3
t = = =
A
s
at ( ) 2 =
x x ,
y
A
s
A
s
O
C
y
a
t
ds
x
y

s
A
s
A
s
O
Fig. 3.20 Stringer-stiffened, semicircular section.
x
I
xy
2 2
dA tds tad = = x acos = y asin =
Thin-Walled Structures 89
Extension, pure bending, and thermal effects for multi-material beams
From the second of eqs. (3.41), the rst area moment about the axis is
and from the second of eqs. (3.43) with by the denition of the centroid we nd . Thus,
the centroid is located at
The second area moment about the axis is (see the rst of eqs. (3.46))
and the second area moment about the axis is
Now we use the parallel axis theorem, eqs. (3.49), to get the second area moments about the centroidal system.
To summarize, the second area moments about the centroidal system are

3.5 Extension, pure bending, and thermal effects for multi-material
beams
Strain-displacement Extensional deformation of an element of the beam is shown in Fig. 3.21. The axial nor-
mal strain due to extension is given by
A ta d

2
---

2
---

2A
s
+ at 2

2
---at
\ )
[
+ 2at = = =
y
Q
y
acos ( )ta d
2
2

2a
2
t = =
Q
y
0 = x
c
A 2a
2
t =
x
c
2a
2
t
2at
--------------
a

--- = = y
c
0 =
x
I
xx
asin ( )
2
ta d
2
2

a
2
A
s
a ( )
2
A
s
+ + a
3
t

2
---
1
4
--- 2 sin
\ )
[
2
2
a
3
t +
3
2
---a
3
t = = =
y
I
yy
acos ( )
2
ta d
2
2

2 0 ( )
2
A
s
+ a
3
t

2
---
1
4
--- 2 sin +
\ )
[
2
2

2
---a
3
t = = =
I
xx
I
xx
y
c
2
A
3
2
---a
3
t 0 ( )
2
A
3
2
---a
3
t = = =
I
yy
I
yy
x
c
2
A

2
---a
3
t
a

---
\ )
[
2
2at a
3
t

2
---
2

---
\ )
[
= = =
I
xx
4.712a
3
t = I
yy
0.934a
3
t = I
xy
0 =
Axial Normal Stress in Pure Bending and Extension
90 Thin-Walled Structures
For combined bending and extension, we superpose the extensional strain and the bending strain, eq. (3.15), to
obtained the total strain. Hence,
(3.56)
Material law Consider Hookes law for the case of thermal strain. A change in temperature causes an expan-
sion of an unrestrained beam element without an associated mechanical stress. If mechanical stresses are present
then the total strain is composed of the sum of the mechanical strain and the thermal strain. That is,
(3.57)
in which denotes the temperature, the stress-free temperature, the coefcient of linear thermal expan-
sion (CTE) of the material, and denotes the modulus of elasticity of the material. If the temperature is given in
degrees fahrenheit ( ), then the CTE has dimensional units . In the material law for the beam given by
eq. (3.57), we have assumed that the lateral stresses are negligible with respect to the axial normal
stress and that the thermal strains in the x- and y-directions can be neglected. Solving eq. (3.57) for the axial
normal stress we get
(3.58)
where
(3.59)
In general the material properties can be a function of the temperature as well. However, we assume
that over moderate temperature changes these material properties do not vary signicantly from an average value.
In addition to the thermal effects, we assume that the beam cross section is hetrogeneous; i.e., the material
properties and are functions of coordinates x and y. In the case of heterogeneity, we dene a reference mod-
ulus of elasticity , which is usually selected as some convenient positive value in problem solving. The refer-
ence modulus is independent of x and y. We write eq. (3.58) as
z
dz
w z ( )
w dw +
dz dw +
z
y
Fig. 3.21 Extensional deformation of a element of the beam

z
dz dw + ( ) dz
dz
------------------------------------ = which in the limit as dz 0 gives
z
w =

z
w x
y
y
x
+ + =
{
||||
extension bending

z

z
E T T
0
( ) + =
||||||
mechanical thermal
T T
0

E
F 1 F

x
and
y

z
E
z
ET =
T T T
0
=
E and
E
E
0
Thin-Walled Structures 91
Extension, pure bending, and thermal effects for multi-material beams
(3.60)
Resultants We substitute the strain-displacement relation,eq. (3.56), into eq. (3.60) for the axial normal strain to
get
(3.61)
Recall that the beam resultants, rst given by eqs. (3.20), are
Now substitute eq. (3.61) for the axial normal stress in these beam resultants to get the material law for the beam.
We write the nal expression of the material law for the resultants in the matrix form
(3.62)
where the modulus weighted area is
(3.63)
the modulus weighted rst area moments are
(3.64)
the modulus weighted second area moments are
(3.65)
the modulus weighted product area moment is
(3.66)
the thermal axial force is
(3.67)
and the thermal bending moments are
(3.68)

z
E
0
E
E
0
------
\ )
[

z
ET =

z
E
0
E
E
0
------
\ )
[
w x
y
y
x
+ + ( ) ET =
N
z
A d
A

= M
x
y
z
A d
A

= M
y
x
z
A d
A

=
N
M
x
M
y
E
0
A
*
Q
x
*
Q
y
*
Q
x
*
I
xx
*
I
xy
*
Q
y
*
I
xy
*
I
yy
*
w

y
N
t
M
x
t
M
y
t
=
A
*
E
E
0
------ A d

=
Q
x
*
E
E
0
------ y A d

= Q
y
*
E
E
0
------ x A d

=
I
xx
*
E
E
0
------ y
2
A d

= I
yy
*
E
E
0
------ x
2
A d

=
I
xy
*
E
E
0
------ xy A d

=
N
t
ET A d

=
M
x
t
y ET ( ) A d

= M
y
t
x ET ( ) A d

=
Axial Normal Stress in Pure Bending and Extension
92 Thin-Walled Structures
We take the origin of the x-y coordinate system at the modulus-weighted centroid. The modulus-weighted
centroid is characterized by the fact that the modulus-weighted rst area moments vanish; i.e.,
(3.69)
We will show how to use this property to nd the modulus-weighted centroid in the next example. With the ori-
gin of the x-y coordinates at the modulus-weighted centroid, the material law for the resultants, eq. (3.62),
reduces to
(3.70)
By locating the origin at the modulus-weighted centroid we de-couple the extensional response of the beam from
its bending response.
Lastly we write the normal stress given by eq. (3.61) in terms of the resultants. Invert eq. (3.70)) and
solve for the extensional strain , and curvatures . to get
(3.71)
(3.72)
Now substitute eq. (3.71) for and eqs. (3.72) for into eq. (3.61) to eliminate these terms in the
equation for . The equation for the normal stress in the form
(3.73)
where is the mechanical portion of normal stress and is the thermal portion of the normal stress. The for-
mulas for these components are
(3.74)
(3.75)
In eqs. (3.74) and (3.75) the modulus E and coefcient of thermal expansion are functions of x and y in gen-
eral.
EXAMPLE 3.5 A multi-material beam with a symmetric cross section
A beam of three materials is subjected to a positive bending moment only. The maximum tensile normal
stress is 35 ksi. Find the maximum compressive normal stress.
Q
x
*
Q
y
*
0 = =
N
M
x
M
y
E
0
A
*
0 0
0 I
xx
*
I
xy
*
0 I
xy
*
I
yy
*
w

y
N
t
M
x
t
M
y
t
=

z
w
x
and
y
w
1
E
0
A
*
------------ N N
t
+ ( ) =

y
1
E
0
I
xx
*
I
yy
*
I
xy
* 2
( )
----------------------------------------
I
yy
*
I
xy
*

I
xy
*
I
xx
*
M
x
M
x
t
+
M
y
M
y
t
+
=
w
x
and
y

z

z
m

z
t
+ =

z
m

z
t

z
m
E
E
0
------ N
I
xx
*
M
y
I
xy
*
M
x
( )x
I
xx
*
I
yy
*
I
xy
*2
( )
--------------------------------------------
I
yy
*
M
x
I
xy
*
M
y
( )y
I
xx
*
I
yy
*
I
xy
*2
( )
------------------------------------------- + + =

z
t
E
E
0
------ N
t
I
xx
*
M
y
t
I
xy
*
M
x
t
( )x
I
xx
*
I
yy
*
I
xy
*2
( )
--------------------------------------------
I
yy
*
M
x
t
I
xy
*
M
y
t
( )y
I
xx
*
I
yy
*
I
xy
*2
( )
-------------------------------------------- + + ET =

M
x
Thin-Walled Structures 93
Extension, pure bending, and thermal effects for multi-material beams
Solution In this example we take the reference modulus , so that , , and
. The modulus weighted area from eq. (3.63) is
Note that the modulus weighted area is twice the geometric area. We choose the - system centered at the bot-
tom of the beam as shown in Fig. 3.22. Since the -axis is an axis of symmetry, both in geometry and in mate-
rial properties, the modulus weighted centroid lies on this axis. We need only to nd the location on the -axis of
the modulus weighted centroid; i.e., . Begin by calculating the modulus weighted rst area moment about the
-axis. See eqs. (3.64). That is,
so
We now use the parallel axis theorem for the rst area moment as given by eqs. (3.43), but generalized to the
multi-material case to write
Hence
Note that the modulus weighted centroid is located at the intersection of materials 1 and 2, which does not coin-
1
2
3
x
y
x
y
2in
2in
2in
1in
y
c
*
M
x
E
1
35msi =
E
2
20msi =
E
3
5msi =
1msi 10
6
psi =
Fig. 3.22 Bending of a multi-material beam with a symmtric cross section
E
0
10Msi = E
1
E
0
3.5 = E
2
E
0
2 =
E
3
E
0
0.5 =
A
*
E
E
0
------ A d

E
i
E
0
------ A
i
i 1 =
3

3.5 2in
2
2 2in
2
0.5 2in
2
+ + 12in
2
= = = =
x y
y
y
y
c
*
x
Q
x
*
E
E
0
------ y A d

E
i
E
0
------ y
i
A
i
i 1 =
3

3.5 5in 2in


2
2 3in 2in
2
0.5 1in 2in
2
+ + = = =
Q
x
*
48in
3
=
Q
x
*
Q
x
*
y
c
*
A
*
0 for the modulus weighted centroid = =
y
c
*
Q
x
*
A
*
------
48in
3
12in
2
------------- 4in = = =
Axial Normal Stress in Pure Bending and Extension
94 Thin-Walled Structures
cide with the geometric centroid of the cross-sectional area.
Now that the modulus weighted centroidal coordinates x-y are located, the modulus weighted second area
moments are calculated. Again, the y-axis is an axis of symmetry, both in geometry and in material properties, so
that the modulus weighted product area moment about the x-y axes vanishes; i.e., . Since
, the formula for the axial normal stress, eqs. (3.73) and (3.74), reduces to
Thus, we only need to compute . This computation proceeds as follows
In achieving this result we used the fact that the (geometric) second area moment of the i-th area element about
the x-axis is , which is based on the composite body technique and parallel axis theorem.
See eq. (3.53). Performing the computations we have
Only material 1 is in tension for . So the axial normal stress is a maximum in material 1 at
;i.e., Thus,
Hence . The maximum compressive normal stress in material 2 is
The maximum compressive normal stress in material 3 is
Therefore, the magnitude of the maximum compressive normal stress is 20ksi.
As a function of y, the bending normal stress is given by
I
xy
*
0 =
N M
y
I
xy
*
0 = = =

z
E
E
0
------
M
x
I
xx
*
------- y = 2in y 4in
I
xx
*
I
xx
*
E
E
0
------ y
2
A d

E
i
E
o
------ y
2
A d

A
i

i 1 =
3

E
i
E
o
------ y
2
A d

A
i

i 1 =
3

E
i
E
o
------ I
xx
y
2
A + ( )
i
i 1 =
3

= = = =
y
2
A d

A
i

I
xxi
y
i
2
A
i
+ =
I
xx
*
3.5
1
12
------ 1 2
3
1 ( )
2
2 +
\ )
[
2
1
12
------ 1 2
3
1 ( )
2
2 +
\ )
[
+ + =
0.5
1
12
------ 1 2
3
3 ( )
2
2 +
\ )
[
24in
4
=
M
x
0 >
y 2in =
z
35ksi =
35ksi 3.5
M
x
24in
4
------------- 2in =
M
x
120
3
10 lb-in =

z
E
2
E
0
------
M
x
I
xx
*
------- 2in ( ) 2 5000lb/in
3
2in ( ) 20ksi = = =

z
E
3
E
0
------
M
x
I
xx
*
------- 4in ( ) 0.5 5000 4 ( ) 10ksi = = =
Thin-Walled Structures 95
Extension, pure bending, and thermal effects for multi-material beams
and this is plotted in Fig. 3.23. Note that the bending normal stress jumps at the material interfaces since the
modulus is discontinuous, but is linear in y within a material layer. The axial normal strain in terms of the kine-
matic quantities , , and is given by eq. (3.56). For this problem , so the from eq.
(3.71); i.e., there is no axial extensional component to the strain. Also, , which
gives from eq. (3.72) that and . From eq. (3.56), the axial strain distribution in y -
coordinate is
The maximum tensile strain is at y = 2 in, and the maximum compressive strain is
at y = -4 in. (The quantity is read as microstrain.) The strain is continuous in
the coordinate y as is shown in Fig. 3.23.

z
17 500 lb/in
3
, y 0 y 2in <
10 000 lb/in
3
, y 2in y 0 < <
2 500 lb/in
3
, y 4in y 2in <
\
|
|
[
=
w
x

y
N N
t
0 = = w 0 =
M
y
M
y
t
M
x
t
I
xy
*
0 = = = =

x
M
x
E
0
I
xx
*
( ) =
y
0 =

z
y
x
y
M
x
E
0
I
xx
*
------------- y
5000 lb/in
3
10
6
10 lb/in
2
------------------------------- 500 10
6
in
1
( )y = = = = 4in y 2in
1000 10
6
in/in 1000 =
2000 10
6
in/in 2000 =
4
2
2
y in ,

z
,
0
2000
1000
4
2
2
y in ,
20 10 5
35

z
ksi ,
1
2
3
x
y
2in
2in
2in
1in
M
x
0 >
Fig. 3.23 Distribution of normal stress and strain through the thickness for the multi-material beam
Axial Normal Stress in Pure Bending and Extension
96 Thin-Walled Structures
3.6 Problems
1.For the thin-walled Y-section shown at right, the
branch thickness t is much smaller that the branch
length l.
a)Determine the location of the centroid.
b)Determine the second area moment of the
cross section about the x axis.
c)If the bending moment components M
x
> 0 and M
y

= 0, determine the magnitudes of the largest tensile
and compressive normal stress due to bending. State
the coordinates where they occur.
2. For the thin-walled S-section shown at right the
radius a = 5mm, the wall thickness t = 0.64mm, and
the bending moment components are M
x
= 3500
Nmm, M
y
= 0. Determine the angle of the neutral
axis in the cross section, and the magnitude of the
maximum bending normal stress.
3.Half of the cross section of a ship is shown on the
next page. Only the material that is effective in the
longitudinal bending is illustrated in the gure.
Determine the area A, location of the centroid ,
and the second area moment about the x-axis (
)for the full section. Use the tabular format for
the computations as given in Example 3.3. All plat-
ing has a thickness t = 14 mm unless other wise noted. The descriptions of the numbered structural elements
shown in the gure are given in the table.
item # Description
1 outer bottom
2 Inner bottom
3 Center girder
4 & 5 Side girders
6 Bilge (curved portion)
7 Side plating
8 Second deck plating
9 & 11
Hatch side girders
10 Strength deck plating
x
y
c
l l
x
y y ,
45 45
C
O
Each of the five branches: length l and thickness t
y
c
I
xx
4a
a
t
a
t

x
y
C
neutral axis in the cross
section
y
c
I
xx
L500 400 25
Thin-Walled Structures 97
Problems
4 m
9 m
5.5 m
R = 1 m
3.5 m
6.5 m
1
2
3 4 5
6
7
8
9
10
11
x
x
C
y
c
t/2
t
C
x
R
2R
I
xx

2
---R
3
t
1
2
---
4

2
-----
\ )
[
=
A

2
---Rt =
C
500mm
400mm
139mm
L500 400 25
x
A 0.0225m
2
=
I
xx
6.077
4
10 m
4
=
Axial Normal Stress in Pure Bending and Extension
98 Thin-Walled Structures

Thin-Walled Structures

99

CHAPTER 4

Axial Force, Shear Force
and Bending Moment
Diagrams

The following three methods to construct equilibrium shear force and bending moment diagrams for beams are
described


the method of sections


the differential equation method, and


the semi-graphical approach.
In actual practice the distinction between these methods becomes blurred, because they may be used simul-
taneously by the analyst. In the last section of this chapter the buoyancy force distribution on ships is described
so that the longitudinal bending response can be computed.

4.1 Method of sections

If an aerospace or ocean vehicle structure is assumed to behave as a slender bar built-up from many members,
then in order to determine stresses in the members we follow the approach in mechanics of materials and rst
determine the distribution of the internal forces and moments along the length. The objective of this section is to
review how to determine these internal actions and plot them along the length of the structure. The relationship
between the internal forces and moments to the stresses will be discussed in subsequent chapters. The review is
limited to statically determinate problems such that internal forces and moments may be obtained by the equa-
tions of statics alone.
Consider a straight slender bar with uniform cross section, which has a length 3

c

, and is simply supported at
two locations as shown in Fig. 4.1. In a right-handed Cartesian coordinate system (

x,y,z

), take the

z

-axis is paral-
lel to the length, the

y

-axis in the plane of the gure, and the

x

-axis perpendicular to the plane of the gure. The
origin is at the left end of the bar; . The bar is in equilibrium subject to the external loads

F

and
which act in the

z-y

plane. Neglect the weight of the bar relative to the loads

F

and , as is frequently done
in structures where the applied loads are much greater than the weight of the structure. The load

F

is a point
force acting at the left end of the bar, and is replaced by its horizontal and vertical components

F

z

and

F

y

. The
0 z 3c p
y
z ( )
p
y
z ( )

Axial Force, Shear Force and Bending Moment Diagrams

100

Thin-Walled Structures

load is a distributed load and has units of force per unit length. It is assumed positive if it acts vertically
upward (positive y-direction), and negative if it acts downward. At a typical value for

z

we want to nd the inter-
nal axial force

N(z)

, shear force

V

y

(

z

), and bending moment

M

x

(

z

). We write the internal actions as

N(z)

,

V

y

(

z

),
and

M

x

(

z

), since they are mathematically functions of the coordinate

z

. The basis for their determination is equi-
librium.
Free-body diagrams of the bar removed from its supports and by imagining the bar is cut at some value of

z


are shown in Fig. 4.2. The internal actions

N

,

V

y

, and

M

x

are shown as well as the unknown support reactions

A

y

,

B

z

, and

B

y

. Let us assume the distributed load vanishes in this discussion. Internal actions

N

,

V

y

, and

M

x


are shown in their assumed positive senses. A positive

z

-face has its outward normal in the positive

z

-direction,
and a negative

z

-face has its outward normal in the negative

z

-direction. That is, on a positive

z

-face a positive
axial force

N

acts in a positive

z

-direction, a positive shear force

V

y

acts in the positive

y

-direction, and the posi-
tive moment

M

x

acts clockwise, or as a vector in the positive

x

-direction by the right-hand screw rule. By New-
ton's third law on action/reaction, positive values of

N

,

V

y

, and

M

x

acting on a negative

z

-face have a sense
opposite to their positive values on the positive

z

-face. The sign convention for the internal forces and moments
needs be carefully followed.
The usual procedure to determine

N(z)

,

V

y

(

z

), and

M

x

(

z

), is to rst draw an overall free-body diagram of the
bar removed from its supports and nd the unknown support reactions, and then section the bar at various

z

-loca-
tions to nd

N

,

V

y

, and

M

x

. The reader should verify that the support reactions in Fig. 4.2 are ,
, and , where

F

y

and

F

z

are the known applied loads. Note that the force

A

y

has a sense
opposite to what was originally assumed.
The overall equilibrium free-body diagram of the bar is shown at the top of Fig. 4.3. The diagrams shown
from top to bottom directly below the overall free-body diagram are the axial force diagram, the shear force dia-
gram, and the bending moment diagram, respectively. The axial force diagram is obtained by sectioning the bar
at any value of

z

, 0 <

z

< 3c, placing a positive internal force

N

on the cut faces, and then summing forces in the

z

-direction for one of the two free-body diagrams. Thus,

N

(

z

) is a constant, for all

z

, 0 <

z

< 3c. In a similar man-
ner one obtains

V

y

(

z

) and

M

x

(

z

). Note that the shear force has a discontinuity at

z

= c since a point force acts
there. It is necessary to consider free-body diagrams in two separate ranges of

z

to draw the shear force diagram;
0 <

z

< c, and c <

z

< 3c. The magnitude of the jump in the shear force is equal to the magnitude of the point
z
y
F
F
y
F
z
c 2c
Fig. 4.1 A slender bar simply supported at two locations and subjected to force F and
distributed load . p
y
z ( )
p
y
z ( )
p
y
z ( )
p
y
z ( )
A
y
3 2 ( )F
y
=
B
z
F
z
= B
y
F
y
2 =
Thin-Walled Structures 101
Differential equation method
force: . The shear force is a piecewise constant for this problem. The bending
moment is piecewise linear; i.e., M
x
(z) = z F
y
, 0 < z < c, and M
x
(z) = (3c - z)(F
y
/2), c < z < 3c. The bending
moment is continuous at z = c, but has a discontinuity in slope , where
. As illustrated in Fig. 4.2, there are two free-body diagrams for each cut. Either one may be
used to nd N, V
y
, and M
x
. If the left free-body diagram is used to nd N, V
y
, and M
x
, then equilibrium of the
right free-body diagram will give the same values of N, V
y
, and M
x
. If it does not, there is either a math error, the
sign convention is violated, or overall equilibrium is in error. Sketching the axial force, shear force, and bending
moment diagrams by the method illustrated in this section is called the method of sections.
4.2 Differential equation method
The distributed load intensity , the shear force V
y
(z), and bending moment M
x
(z) are related by simple dif-
ferential equations at z, if no point forces or concentrated couples act at z. These differential equations are useful
in the construction of the shear force and bending moment diagrams.
F
y
F
z
N
V
y
M
x
z
V
y
M
x
z
c
Ay
2c
By
Bz
Left and right FBDs for cuts in the range 0 < z < c
F
y
F
z
A
y
B
y
B
z
c 2c
Fig. 4.2 Free body diagrams (FBDs) of the slender bar with = 0. p
y
z ( )
overall FBD
V
y
c
+
( ) V
y
c
-
( ) 3 2 ( )F
y
=
M'
x
c
+
( ) M'
x
c
-
( ) 3 2 ( )F
y
=
M'
x
dM
x
dz =
p
y
z ( )
Axial Force, Shear Force and Bending Moment Diagrams
102 Thin-Walled Structures
Consider a portion of a straight beam subjected to distributed load whose intensity is as shown in Fig.
4.4. By convention is positive upwards and negative downwards. A free-body diagram of a portion z-
long of the beam is shown in Fig. 4.5. The shear force and bending moment change with z, and so their values at
z + z are different from their values at z. The distributed load acting on the segment z is replaced by single
force of magnitude acting long a line of action given by z = z*, where z < z* < z + z. Mathematically this
is permissible by the mean value theorem (for integrals) for continuous functions of a single variable, and from
this theorem we have
F
z
F
y
3/2 F
y
F
y
/2
z
z
z
z
0
N
F
z
V
y
0
- F
y
F
y
/2
M
x
0
- cF
y
c 2c
+ N
V
y
M
x
N
V
y
M
x
Fig. 4.3 Axial force N, shear force V
y
, and bending moment M
x
diagrams
for the slender bar with = 0, p
y
z ( )
p
y
z ( )
p
y
z ( )
p
y
z
Thin-Walled Structures 103
Differential equation method
Note that in the limit as , , and .
In the free-body diagram of Fig. 4.6 we sum forces vertically, divide by z, and take the limit as to
get
(4.1)
Often it is more convenient to use an integrated form of eq. (4.1). Integrating it from z
1
to z we obtain
p
y
1
z
------ p
y
( ) d
z
z z + ( )

=
z 0 z
*
z p
y
p
y

z
y
Fig. 4.4 Distributed load intensity shown acting in the positive sense
p
y
z ( )
M
x
+ M
x
M
x
V
y
+ V
y
V
y
z
z + z
Fig. 4.5
p
y
z ( )
M
x
+ M
x
V
y
+ V
y
V
y
M
x
p
y
z
*
( )z
z + z
z*
z
Fig. 4.6
z 0
z d
dV
y
p
y
=
Axial Force, Shear Force and Bending Moment Diagrams
104 Thin-Walled Structures
(4.2)
The integrated form is valid only if there are no point loads between z
1
and z. If we sum moments at z in the free-
body diagram, divide by z, and take the limit as we get
(4.3)
The integrated form of eq. (4.3) is
(4.4)
which is valid if no point couples act between z
1
and z. Equations (4.1) and (4.3) are the differential equations of
equilibrium for the beam. Equation (4.1) is valid at z if no point force acts there, and eq. (4.3) is valid at z if no
point couple acts there. Considering separately the equilibrium of a point force F
0
and a point couple with
moment magnitude C
0
acting at z = z
0
, as shown in Fig. 4.7, we obtain
(4.5)
Distributed loads may be replaced by a resultant force acting at the center of pressure. This procedure is con-
venient in many situations. For example, the air load distribution on a wing may be replaced by lift and drag
forces acting at the center of pressure. A segment of a beam from z = z
1
to a typical value of z, z > z
1
, which has a
distributed load with intensity acting on it is shown in Fig. 4.8. Using as a dummy variable to measure
the axial position, the resultant force is
(4.6)
and the center of pressure is given by
V
y
z ( ) V
y
z
1
( ) p
y
( ) d
z
1
z

=
z 0
z d
dM
x
V
y
=
M
x
z ( ) M
x
z
1
( ) V
y
( ) d
z
1
z

+ =
V
y
z
0
+
( )
M
x
z
0
+
( )
V
y
z
0
-
( )
M
x
z
0
-
( )

z
y
z
0
F
0
C
0
Fig. 4.7 Point force and couple acting at z
0
on an inintesimal beam element
V
y
z
0
+
( ) V
y
z
0
-
( ) F
0
= M
x
z
0
+
( ) M
x
z
0
-
( ) C
0
=
p
y
z ( )
F
y
z ( )
F
y
z ( ) p
y
( ) d
z
1
z

=
z
p
z ( )
Thin-Walled Structures 105
Differential equation method
(4.7)
Both the resultant force and center of pressure are functions of z. Equations (4.6) and (4.7) are con-
ditions of statical equivalence for the distributed load .
EXAMPLE 4.1 Cantilever wing with tip tank
Consider the cantilever wing with tip tank as shown in Fig. 4.9. Given the weight of the tip tank and its contents
W, the distance e of the weight W from the wing tip, the wing span L, and the value of the distributed load inten-
sity at the wing root, determine the shear force and bending moment along the span. The solution to this
problem is given by a Mathematica 4.0 program listed below.
V
y
z
1
( )
M
x
z
1
( )
V
y
z ( )
M
x
z ( )
z
1
z
p
z ( )
z
F
y
z ( )
V
y
z
1
( )
M
x
z
1
( )
V
y
z ( )
M
x
z ( )
z
1

z
p
y
z ( )
Fig. 4.8 Resultant force at the center of pressure statically equivalent to a distributed load.
z
p
z ( )F
y
p
y
( ) d
z
1
z

=
F
y
z ( ) z
p
z ( )
p
y
z ( )
p
y
z ( ) p
0
z
L
--- =
z
L
e
W
Fig. 4.9 Cantilever wing with tip tank.
p
0
Axial Force, Shear Force and Bending Moment Diagrams
106 Thin-Walled Structures

1 2 3 4 5 6
Example 4.1
Shear force and bending moment diagrams for a cantilever wing with tip tank.
Input the distributed load function.
p
y
= p
0
z

L
;
The shear force and bending moment distributions are determined from eqs. (4.2) and
(4.4) with z
1
=0. The shear force and bending moment at the wing tip (z = 0) are
denoted by V
y0
and M
x0
, respectively.
V
y
= V
y0
-

p
y
z
-
z
2
p
0

2 L
+V
y0
M
x
= M
x0
+

V
y
z
M
x0
-
z
3
p
0

6 L
+z V
y0
Boundary conditions at the wing tip, obtained from equilibrium of the tip tank, deter-
mined the shear force V
y0
and moment M
x0
.
bc
1
= HV
y
. z 0L - W
bc
2
= HM
x
. z 0L - e W
- W +V
y0
- e W +M
x0
slv1 = Solve@bc
1
== 0, V
y0
D
88V
y0
W<<
V
y0
= V
y0
. slv1@@1DD
W
slv2 = Solve@bc
2
== 0, M
x0
D
88M
x0
e W<<
M
x0
= M
x0
. slv2@@1DD
e W
Thin-Walled Structures 107
Differential equation method
Print@ "Shear force V
y
HzL = ", V
y
D
Print@"Bending moment M
x
HzL = " M
x
D
Shear force V
y
HzL = W -
z
2
p
0

2 L
Bending moment M
x
HzL =
i
k
j
je W +W z -
z
3
p
0

6 L
y
{
z
z
Plot the shear force and bending moment diagrams for the following parameter
values: L = 144 in, p
0
= 70 lb/in, W = 500 lbs, and e = 6 in. (Plots labled p1 and
p2 have been suppressed using the DisplayFunction option.)
p1 =
Plot@HV
y
. 8 L 144, p
0
70, W 500, e 6<L,
8z, 0, 144<,
PlotRange 81000, - 5000<,
GridLines Automatic,
AxesLabel 8"z,inches", "V
y
, lbs"<,
PlotLabel
StyleForm@"Shear force diagram",
"Section"D,
DisplayFunction - > IdentityD
Graphics
p2 =
Plot@HM
x
. 8 L 144, p
0
70, W 500, e 6<L,
8z, 0, 144<,
GridLines Automatic,
AxesLabel 8"z,inches", " M
x
, lb- in"<,
PlotLabel
StyleForm@"Bending moment diagram",
"Section"D,
DisplayFunction - > IdentityD
Graphics
Show@GraphicsArray@88p1<, 8p2<<DD
1 2 3 4 5 6
The magnitude of the bending moment is largest at the wing root, and this vlaue is
important for wing structural design. Its value in lb-in is
M
x
. 8 L 144, p
0
70, W 500, e 6, z 144<
- 166920
(The plots are are shown on the next page.)
Axial Force, Shear Force and Bending Moment Diagrams
108 Thin-Walled Structures
1 2 3 4 5 6
20 40 60 80 100 120 140
z,inches
-50000
-40000
-30000
-20000
-10000
10000
M
x
, lb- in
Bending moment diagram
20 40 60 80 100 120 140
z,inches
-5000
-4000
-3000
-2000
-1000
1000
V
y
, lbs
Shear force diagram
Shear force and bending moment diagrams for the cantilever wing with tip tank
Thin-Walled Structures 109
Differential equation method
The shear force and bending moment at the wing tip are determined from equilibrium of the tip tank, which
gives V
y
(0) = W and M
x
(0) = eW. Note that eq. (4.1) shows that the slope on the shear diagram is equal to the
distributed load intensity. For example at z = 0, , so that the slope of the shear diagram is zero at z = 0.
Similarly, the slope on the moment diagram is equal to the shear force as given by eq. (4.3). In particu-
lar, at z = 45.36 inches the shear force is zero. Thus, in the bending moment diagram the moment is stationary at
z = 45.36 in (i.e., it has a horizontal slope), and M
x
may be either a local maximum, minimum, or a value corre-
sponding to horizontal inection point. It is important to compute the largest magnitude of the bending moment,
and this is accomplished by checking the bending moments where V
y
= 0, at the end points of the beam, and the
locations where the bending moment is discontinuous. Note that the maximum bending moment magnitude
occurs at the wing root, where . The bending moment changes sign, and hence vanishes, at
z = 81.4 inches. In a plot of the beam deection versus z, which is not shown above, the location z = 81.4 inches
is called an inection point because the curvature is changing from concave down (positive M
x
) to concave up
(negative M
x
) as z increases through z = 81.4 inches.
EXAMPLE 4.2 The air load acting on a wing given as discrete data.
The problem statement and data for this example is taken from the aircraft structures text by Peery (1950). How-
ever, the notation is changed to that of the this text, and the solution is given in terms of a Mathematica 3.0 pro-
gram. Since the air load on the wing is given at discrete spanwise locations and not as a mathematical function, it
is useful to use Mathematicas list manipulation capabilities to effect the solution. Before giving the problem
statement, we will discuss some aspects of list manipulations.
Lists provide a mechanism for representing arrays, vectors, matrices, and for grouping together objects such
as data, variables, or expressions. A list is a collection of objects whose symbols are enclosed in braces, {}, and
separated by commas, as in . It is usually more efcient to do operations on
lists rather than to do operations on individual items in the list. A function is applied separately to each element in
the list if it has the attribute Listable. For example, addition, multiplication, and the logarithm have the attribute
Listable, so that

That is, Listable functions in Mathematica are automatically distributed or threaded over lists that appear as its
arguments. The number of elements in a list is given by the built-in function Length [ ]; e.g.,
. A summary of Mathematicas built-in functions used for list manipulation is given in
the table below.
Summary of list manipulation functions in Mathematica (taken from Blachman, 1992)
Function Description
Range [min, max, step] Generates the list {min, .., max} using step (arithmetic progression)
Table [expr, {imax}] Generates a list of imax copies of expr (more general)
Array [s, dim] Generates a list of length dim with elements s[i]
Sort [list] Sorts elements of list into canonical order
p
y
0 ( ) 0 =
dM
x
dz
M
x
166920 lb-in =
item
1
item
2
item
3
item
n
, , , , { }
5 8 11 , , { } 2 3 6 , , { } + 7 5 5 , , { } =
5 8 11 , , { }* 2 3 6 , , { } 10 24 66 , , { } =
Log 5 8 11 , , { } [ ] Log 5 [ ] Log 8] [ ] Log 11] [ ] , , { } =
Length 5 8 11 , , { } [ ] 3 =
Axial Force, Shear Force and Bending Moment Diagrams
110 Thin-Walled Structures
Problem statement: The aerodynamic loads on an airplane wing cannot be represented by a simple equa-
tion. The load per inch of span, , of the airplane wing shown in Fig. 4.10is tabulated in column
two of the Table printed at line Out[14] in the Mathematica program below. Find the shear force and bending
moment diagrams for the wing.
Solution: The values of the shear force and bending moment at various points along the wing are calculated
in the Table (see Out[14] in the code). The points are called stations and are designated by their distances from
Reverse [list] Reverses elements in list
RotateLeft [ list, n] Cycles the elements n positions to the left
RotateRight [ list,n] Cycles the elements n positions to the right
Permutations [list] Generates a list of all possible permutations of the elements of list
Drop [list, n] Drops the rst n elements from list
Take [list, n] Takes the rst n elements from list
First [list] Give the rst element of list
Last [list] Gives the last element of list
list [[n]] or Part [list, n] Gives the nth element
Rest [list] Returns all but the rst element of list
Select [list, crit] Picks out elements in list which meet the criterion of crit
Append [list, elem] Returns a list with elem appended to the end of list
AppendTo [list, elem] Changes list by appending elem to the end
Prepend [list, elem] Returns a list with elem added to the from of list
PrependTo [list, elem] Changes list by adding elem to the front
Insert [list, elem, n] Inserts elem at position n in list
Length [list] Gives the number of elements in list
Dimensions [list] Gives the dimensions of a list or expression
Complement [ list
1
, list
2
, ... ] Gives the complement, i.e., those elements in list
1
but not in list
2
, ...
Intersection [ list
1
, list
2
, ... ] Gives a sorted list of all the elements common to all list
1
, list
2
, ...
Union [list
1
, list
2
, ... ] Gives a sorted list of the distinct elements
Join [list
1
, list
2
, ... ] Joins or concatenates lists together
Partition [list, n] Partition list into sublists of length n
Flatten [list] Flattens out nested lists, i.e., eliminates nested lists
Transpose [list] Transpose
Apply [f, list] Replaces the head of list with f
Map [f,list] Applies f to each element in list
Listable An attribute, if set, automatically maps a functions onto a list
ColumnForm[list] Prints list as a column
MatrixForm[ list] Prints elements in list in a regular array
Summary of list manipulation functions in Mathematica (taken from Blachman, 1992)
Function Description
p
y
z ( ) p z ( ) =
Thin-Walled Structures 111
Differential equation method
the centerline of the airplane, as shown in Fig. 4.10. These distances are measured along the wing rather than
horizontally, since the air loads are perpendicular to the wing. The distances between stations, , are computed
as a list in the code. The value of the shear at any point is obtained as the area under the load curve from that
point out to the wing tip. The load curve is assumed to be a series of straight lines between the known points, and
the area is obtained as the sum of the areas of the trapezoids. The area of the trapezoids are obtained as the prod-
uct of the average height and the base . The change in the shear between two stations is equal to
the area of the load curve between the stations. The shear is then obtained by summation of the -values.
The change in the bending moment between two stations is equal to the area under the shear curve. This
area is also assumed trapezoidal and is obtained by multiplying the sum of the shears at the adjacent stations by
one-half the distance between the stations. The bending moments are obtained by a summation of the -val-
ues. Plots of the air load, shear force, and bending moment distributions are shown at the end of the Mathematica
program
z
Fig. 4.10
z
p
ave
z V
y
V
y
V
y
M
x
M
x
Example 4.2: Numerical quadrature for the shear force and bending moment in a wing
(Peery, 1950, pp. 107-109)
In[1]:= Off@General::spell1D
Input the airload intensity at each z-station and each z-station coordinate as two separate lists. Dimensional
units: z-list, inches; p-list, lb/in.
In[2]:= z = 80, 20, 40, 60, 80, 100, 120, 140, 160, 180,
200, 220, 225<;
p = 8125, 123, 120, 116, 111, 105, 98, 89, 80, 71,
58, 35, 0<;
Compute distances between stations.
In[3]:= Dz = Take@HRotateLeft@zD - zL, Length@zD - 1D
Out[3]= 820, 20, 20, 20, 20, 20, 20, 20, 20, 20, 20, 5<
Axial Force, Shear Force and Bending Moment Diagrams
112 Thin-Walled Structures
.
1 2 3 4 5 6 7
Compute average airload intensity in each interval between stations.
In[4]:= p
ave
= Take@N@HRotateLeft@pD + pL 2D, Length@pD - 1D
Out[4]= 8124., 121.5, 118., 113.5, 108.,
101.5, 93.5, 84.5, 75.5, 64.5, 46.5, 17.5<
The trapezoidal rule of numerical integration is used to compute the change in the shear force over each
interval from eq. (4.2). A list of DV
y
- values is computed by a direct multiplication of lists Dz and p
ave
.
In[5]:= DV
y
= - Dz * p
ave
Out[5]= 8- 2480., - 2430., - 2360., - 2270., - 2160.,
- 2030., - 1870., - 1690., - 1510., - 1290., - 930., - 87.5<
Since V
y
HtipL - V
y
HrootL =

root
tip
HdV
y
dzL z, and V
y
HtipL = 0, the shear force at the root is the negative of
the sum the elements in the DV
y
-list. A simple way to sum elements in a list is to change the Head of the list
from "List" to "Plus" by using the Apply function.
In[6]:= Head@DV
y
D
Out[6]= List
In[7]:= V
y0
=- Apply@Plus, DV
y
D
Out[7]= 21107.5
The shear force at station i is the the partial sum of the of the DV
y
- values over the intevals from 1 to i; i.e.,
V
y
HiL =
j=1
i
DV
y
H jL +V
y0
. We use the Table function to generate a list of the shear force values at each
station.
In[8]:= V
y
= Table@HSum@DV
y
@@jDD, 8j, 1, i<D +V
y0
L, 8i, 1, Length@DV
y
D<D
Out[8]= 818627.5, 16197.5, 13837.5, 11567.5, 9407.5,
7377.5, 5507.5, 3817.5, 2307.5, 1017.5, 87.5, 0.<
Add the value of the shear force at the root to the beginning of this list in order to have the shear force at each
z
i
- station, including the root.
In[9]:= V
y
= PrependTo@V
y
, V
y0
D
Out[9]= 821107.5, 18627.5, 16197.5, 13837.5, 11567.5, 9407.5,
7377.5, 5507.5, 3817.5, 2307.5, 1017.5, 87.5, 0.<
Thin-Walled Structures 113
Differential equation method

1 2 3 4 5 6
Compute the average force in each interval from the list of shear force values.
In[10]:= V
ave
= Take@N@HRotateLeft@V
y
D +V
y
L 2D, Length@V
y
D - 1D
Out[10]= 819867.5, 17412.5, 15017.5, 12702.5, 10487.5,
8392.5, 6442.5, 4662.5, 3062.5, 1662.5, 552.5, 43.75<
The change in the bending moment is computed from eq. (4.4) using the trapezoidal rule of numerical
integration.
In[11]:= DM
x
= Dz * V
ave
Out[11]= 8397350., 348250., 300350., 254050., 209750.,
167850., 128850., 93250., 61250., 33250., 11050., 218.75<
Since M
x
HtipL - M
x
HrootL =

root
tip
HdM
x
dzL z, and M
x
HtipL = 0, the bending moment at the root is the
negative of the sum the elements in the DM
x
-list.
In[12]:= M
x0
= - Apply@Plus, DM
x
D
Out[12]= - 2.00547 10
6
Compute the bending moment at each station.
In[13]:= M
x
= Table@HSum@DM
x
@@jDD, 8j, 1, i<D + M
x0
L,
8i, 1, Length@DM
x
D<D;
M
x
= PrependTo@M
x
, M
x0
D
Out[13]= 9- 2.00547 10
6
, - 1.60812 10
6
, - 1.25987 10
6
,
- 959519., - 705469., - 495719., - 327869., - 199019.,
- 105769., - 44518.7, - 11268.8, - 218.75, 0.=
In[14]:= TableForm@Transpose@8z, p, V
y
, M
x
<D,
TableHeadings
8None, 8"z,in", "p,lbin",
"V
y
,lb", "M
x
,lb- in"<<D
Axial Force, Shear Force and Bending Moment Diagrams
114 Thin-Walled Structures
1 2 3 4 5 6 7
Out[14]//TableForm=
z,in p,lbin V
y
,lb M
x
,lb- in
0 125 21107.5 - 2.00547 10
6
20 123 18627.5 - 1.60812 10
6
40 120 16197.5 - 1.25987 10
6
60 116 13837.5 - 959519.
80 111 11567.5 - 705469.
100 105 9407.5 - 495719.
120 98 7377.5 - 327869.
140 89 5507.5 - 199019.
160 80 3817.5 - 105769.
180 71 2307.5 - 44518.7
200 58 1017.5 - 11268.8
220 35 87.5 - 218.75
225 0 0. 0.
Plots of the airload, shear force, and bending moment distributions. (Intermediate plots have been suppressed.)
In[15]:= p11 = ListPlot@Transpose@8z, p<D,
PlotStyle 8PointSize@0.02D<,
DisplayFunction - > IdentityD
p12 = ListPlot@Transpose@8z, p<D,
PlotJoined - >True,
DisplayFunction - > IdentityD
p1 = Show@p11, p12,
AxesLabel 8"z, in.", "lbin"<,
PlotLabel "Airload distribution",
DisplayFunction - > IdentityD
p21 = ListPlot@Transpose@8z, V
y
<D,
PlotStyle - >8PointSize@0.02D<,
DisplayFunction - > IdentityD
p22 = ListPlot@Transpose@8z, V
y
<D,
PlotJoined - >True,
DisplayFunction - > IdentityD
p2 = Show@p21, p22,
AxesLabel 8"z, in.", "lb"<,
PlotLabel "Shear force",
DisplayFunction - > IdentityD
p31 = ListPlot@Transpose@8z, M
x
<D,
PlotStyle - >8PointSize@0.02D<,
DisplayFunction - > IdentityD
Thin-Walled Structures 115
Differential equation method

In[24]:= Show@GraphicsArray@88p1<, 8p2<, 8p3<<DD


50 100 150 200
z,in
-210
6
-1.510
6
-110
6
-500000
lb- in
Bending Moment
50 100 150 200
z, in.
5000
10000
15000
20000
lb
Shear force
50 100 150 200
z, in.
20
40
60
80
100
120
lbin
Airload distribution
Axial Force, Shear Force and Bending Moment Diagrams
116 Thin-Walled Structures
4.3 Semi-graphical method
The semi-graphical method to draw the shear force and bending moment diagrams is best illustrated by doing an
example.
EXAMPLE 4.3 Uniform barge with symmetric load
Consider a barge at rest in still water with a uniform immersed cross section, and subjected to the symmetrical
loads shown in Fig. 4.11. This is an example of a structure with no boundary supports, and is typical of aero-
space and ocean vehicle structures. We wish to sketch the shear force and bending moment diagrams for the
barge. In this example there is a distributed load acting on the barge due to buoyancy forces produced by displac-
ing the water. Let represent the distributed load intensity due to buoyancy, and is a constant along the
barge because the immersed cross section is uniform and the water is still.
Solution: A semi-graphical method is used to sketch the shear force and bending moment diagrams. In this
approach we rst sketch the distributive load , then the shear force V
y
(z), and nally the bending moment
M
x
(z). Equations (4.1) and (4.3) are used to note that the slope of shear diagram at z is the negative of the distrib-
uted load intensity at z, and the slope of the moment diagram at z is the shear force at z. In addition, eqs. (4.2) and
(4.4) set at z = z
2
give
(4.8)
(4.9)
Equation (4.8) is interpreted in a graphical sense to mean that the difference in the shear force between z
2

and z
1
is the area under the distributed loading diagram from z
1
to z
2
. This is not geometrical area. The area
between the curve and the z-axis has units of force, and may be positive, zero, or negative. Similarly eq.
(4.9) is interpreted to mean the difference in the bending moment is the area under the shear force diagram.
Vertical equilibrium of the entire barge requires the buoyant upthrust equals 40kN, so that .
The total distributed load intensity is the difference between and the magnitude of the downward acting
15kN (total) 15kN (total)
10kN
5m 5m
5m 5m
Fig. 4.11Uniform section barge in still water with symmetric load.
p
b
p
b
p
y
z ( )
V
y
z
2
( ) V
y
z
1
( ) p
y
z ( ) z d
z
1
z
2

=
M
x
z
2
( ) M
x
z
1
( ) V
y
z ( ) z d
z
1
z
2

=
p
y
z ( )
p
b
2kN ( ) m =
p
b
Thin-Walled Structures 117
Semi-graphical method
applied loading intensity. The distributed loading intensity diagram is constructed in this manner as shown in Fig.
4.13.
5 10 15 20
z, m
Shear force diagram
-4
-2
2
4
kN
Fig. 4.12Shear force and bending moment diagrams for the barge in still water
0
2
-1
p
y
z, m
5
10
15
20
kN/m
10 kN
5 10 15 20
z, m
Bending moment diagram
2.5
5
7.5
10
12.5
15
17.5
kN-m
Axial Force, Shear Force and Bending Moment Diagrams
118 Thin-Walled Structures
The point force of 10kN acting at z = 10m is shown schematically in the -diagram as a downward
pointing arrow. Actually, as , because a point force is a nite load acting over zero length.
Point forces are idealizations to actual loads and introduce discontinuities in the mathematical descriptions of
some of the dependent variables. The reader should verify the distributive loading intensity diagram of Fig.
4.13.The shear force diagram is drawn below the loading intensity diagram in Fig. 4.13. Equilibrium at z = 0
requires V
y
(0) = 0, and the slope dV
y
/dz at z = 0 is equal to 1kN/m. The slope is constant between ,
thus V
y
(z) is a straight line in this range of z. The difference in the shear force between z = 5m and z = 0 is equal
to the negative of the area under the curve which is 5kN. Thus V
y
(5) = 5kN since V
y
(0) = 0. At z = 5
+
m the
loading intensity jumps to +2kN/m. The slope of the shear force jumps from 1kN/m to -2kN/m at z = 5m, but the
shear force is itself continuous. The difference V
y
(10) - V
y
(5) is equal to the negative of the area between the
-curve and the z-axis between z = 5m and z = 10m. Thus V
y
(10) - V
y
(5) = -10kN, so V
y
(10) = -5kN. Note
the shear force is zero at z = 7.5m. At z = 10m the point force of 10kN acts. According to the rst of eqs. (4.5)
V
y
(10
+
) - V
y
(10
-
) = 10kN, so that V
y
(10
+
) = 5kN. The slope of the shear at z = 10m is +2kN/m, and remains con-
stant until z = 15m. The difference V
y
(15) - V
y
(10
+
) = -10kN, so that V
y
(15) = -5kN. Finally, the slope changes to
-1kN/m at z = 15
+
m and remains constant in the range 15 < z < 20. The difference V
y
(20) - V
y
(15) = 5kN, so that
V
y
(20) = 0. Checking vertical equilibrium at z = 20m veries that V
y
(20) should be zero.
Moment equilibrium at z = 0 shows M
x
(0) = 0. The slope of M
x
at z = 0 is equal to the shear force at z= 0.
Hence at z = 0, as shown in Fig. 4.13. The slope on the moment diagram increases linearly
from zero at z = 0 to 5kN at z = 5m. Thus M
x
(z) is parabolic from z= 0 to z = 5. The difference M
x
(5) - M
x
(0) is
equal to the area under shear diagram from z = 0 to z = 5m. Hence, M
x
(5) - M
x
(0) = 12.5 kNm, and M
x
(5) = 12.5
kNm since M
x
(0) = 0. From z = 5 to z = 7.5 the slope of the moment decreases from 5kN to zero. At z = 7.5, M
x

is a local maximum with a magnitude of 18.75 kNm. The slope of M
x
(z) for 7.5 < z < 10 is negative, decreasing
linearly from zero to -5kN. The difference M
x
(10) - M
x
(7.5) = -6.25 kNm, so that M
x
(10) = 12.50 kNm. The
slope of M
x
(z) at z = 10m jumps from a -5kN to a +5kN as shown in Fig. 4.13, but the moment itself is continu-
ous. The bending moment diagram in the range 10 < z < 20 is completed in a manner similar to the description of
its construction in the range 0 < z < 10.
In this example the shear force diagram is antisymmetric about z = 10m and the bending moment is symmet-
ric about z = 10m. This follows from the symmetrical loading on the barge and equilibrium eqs. (4.1) and (4.3).
4.4 Buoyancy Force Distribution on Ships
The simple uniform buoyancy distribution acting on the barge in Example 4.3 is an exception to the buoyancy
distributions found in practice. It is true that equilibrium requires the total buoyant upthrust to equal the weight of
the ship and its contents. However, the distribution of the buoyancy and weight along the length of the ship is not
necessarily the same. The difference in the magnitudes of the buoyancy and weight distribution intensities is the
applied load intensity . In ship design three conditions are recognized to compute for the same ship.
These conditions are called
the still water condition,
sagging condition, and
the hogging condition.
p
y
z ( )
p
y
z 10m
0 z 5m < <
p
y
z ( )
p
y
z ( )
dM
x
( ) dz ( ) 0 =
p
y
z ( ) p
y
z ( )
Thin-Walled Structures 119
Buoyancy Force Distribution on Ships
A more detailed account of these conditions on the longitudinal bending of the ship is given by Muckle
(1967) and Zubaly (1996), and here we only summarize the basic ideas.
A ship in still water is shown in Fig. 4.13, and a section between z and z + dz is shown in Fig. 4.14.
Archimedes' principle asserts that the buoyant upthrust is equal to the weight of the uid displaced. Let A(z)
denote the submerged cross section at z, and let denote the specic weight (force per volume) of the uid. The
differential buoyancy force dF
b
acting on the ship over a differential length dz is
(4.10)
Consequently, the buoyant upthrust per unit ship length, which we designate , is equal to A(z); i.e.,
(4.11)
A curve of for a ship as well as the weight per unit length is shown in Fig. 4.15. Overall equilibrium requires
the area under these curves to have the same magnitude. If the submerged cross section is uniform in z, as is the
z
Fig. 4.13
A z ( )dz
dF
b
Fig. 4.14
dF
b
A z ( )dz =
p
b
p
b
dF
b
dz
--------- A z ( ) = =
p
b
weight/length
buoyancy/length
Fig. 4.15
Axial Force, Shear Force and Bending Moment Diagrams
120 Thin-Walled Structures
case for the barge in Example 4.3, the distribution of the buoyancy per unit length is a constant.
At sea a ship is subjected to waves, and this alters the buoyancy distribution. For longitudinal bending of the
ship two extreme static conditions are assumed: sagging and hogging. In each condition, the length of the wave is
assumed to be the length of the ship. This is an accepted assumption for the worst buoyancy distribution caus-
ing the most severe bending of the ship.
The sagging condition is shown in Fig. 4.16. (Also see Fig. 1.3 on page 5.) The wave crests are at the bow
and stern, and the wave trough is amidships. A schematic of the buoyancy per unit length is shown below the ship
in Fig. 4.16. The immersed cross section is the largest at or near the wave crests, and is least near the trough. The
intensity of the buoyancy distribution reects this. In this condition the deck sags and is in compression while the
bottom is in tension. The worst location to concentrate the cargo in the ship is amidships, as this will result in the
largest bending moment.
The hogging condition is depicted in Fig. 4.17. Here the wave troughs are at bow and stern, and the crest is
amidships. The immersed cross section is greatest near amidships and is least near bow and stern. The distribu-
tion of the buoyancy per unit length , shown in Fig. 4.17, reects this situation. In hogging the deck is in ten-
sion and the bottom is in compression. The worst possible locations to concentrate cargo is fore and aft, as this
will produce the greatest bending moment in the ship.
p
b
p
b
Fig. 4.16Sagging
concentrated weight
p
b
p
b
Fig. 4.17Hogging
concentrated weights
Thin-Walled Structures 121
References
4.5 References
Blachman, N.R., 1992, Mathematica: A Practical Approach, Prentice Hall, Englewood Cliffs, New Jersey,
p. 133.
Muckle, W., 1967, Strength of Ships Structures, Edward Arnold Ltd., London, pp. 27-69.
Peery, D.J., 1950, Aircraft Structures, McGraw-Hill, New York, pp. 107-108.
Zubaly, R.M., 1996, Applied Naval Architecture, The Society of Naval Architects and Marine Engineers,
Cornell Maritime Press, Inc., Centreville, Maryland, pp. 195-237.
4.6 Problems
1. The cantilever wing is subjected to a distributed air load , where the total lift (2
wings) at cruise, wing length ft., and . Also, the wing supports an
engine weighing 1000 lbs. Plot the loading diagram, shear force diagram , and bending moment diagram
as functions of z for ft. Partial answer: lb. and lb-ft.
2. A proposed solar airplane called Centurion is being designed to achieve semi-perpetual ight (Aviation
Week & Space Technology, May 4, 1998, p.54). Centurion is a ying wing with a span of 206 ft., an 8-ft. chord,
and no taper or sweep. The wing has ve sections, one center, two mid-span, and two tips. It is supported by four
landing pods. The tip sections have a dihedral to assist in turning and washout twist to prevent tip stall. The
empty weight is predicted to be 1,105 lb., comprising 630 lb. for structure, 160 lb. for engines and propellers, 150
lb. for avionics, 75 lbs for batteries, 20 lb. of miscellaneous and 70 lb. for 7% growth. The aircraft should be able
to take a 100 lb payload to 100,000 ft. It is powered by 14 electric motors producing a maximum of 2 hp. each.
Assume the following: The span-wise airload distribution acting on the wing is as given in problem 1. Each
engine is modeled as a concentrated weight acting its location on the wing. The payload, avionics, batteries, etc.,
lumped are together as a concentrated load at the center, and that the structural weight is uniformly distributed
p
y
z ( )
2L
z
max
------------- 1 z ( )
2
=
L 20 000lbs , = z
max
32.5 = z z z
max
=
V
y
z ( )
M
x
z ( ) 0 z 32.5 V
y
0 ( ) 9 000 , = M
x
0 ( ) 131934 =
y
z
p
y
z ( )
M
x
M
x
V
y
p
y
+
V
y
6 ft.
32.5 ft.
1000 lb engine
fuselage
Axial Force, Shear Force and Bending Moment Diagrams
122 Thin-Walled Structures
along the span. For steady level ight, determine the shear force and bending moment diagrams from the center-
line of the wing to its tip, and show them in a sketch. Label signicant points. The front view of half of the Cen-
turion is shown below.
3. The barge shown below has a uniform cross section along its length and is subjected to a uniformly distrib-
uted load of intensity , in which P has dimensional units of force/length. Also it is subjected to
buoyancy for the extreme hogging condition . Draw the shear force,
, and bending moment, , diagrams. Label signicant points. Note that .
4. A barge has a plan view as shown. All waterplanes are identical. Cargo is loaded evenly in the four rectangu-
lar holds as shown. Neglecting the weight of the barge itself, construct curves of weight, buoyancy, load, shear,
and bending moment for the loaded barge in still sea water. Label the values of each curve at each bulkhead, and
identify the maximum shear and bending moment. (Zubaly, 1996)
z
1
z
2
z
3
z
4
z
5
z
6
z
7
z
8
z
9
z
10
z
z
1
4.0 ft.
z
2
12.0 ft.
z
3
19.2 ft.
z
4
32.7 ft.
z
5
46.3 ft.
z
6
60.0 ft.
z
7
67.1 ft.
z
8
80.6 ft.
z
9
93.4 ft.
z
10
103.0 ft.
Centerline
p
y
z ( ) P =
p
y
z ( )
buoyancy
P 1
z
L
---


cos =
V
y
z ( ) M
x
z ( ) M
x
max
2P L ( )
2
=
z
L
L
P
waterline
Thin-Walled Structures 123
Problems
5. A barge of uniform rectangular construction has a length of 30m, breadth of 10m, depth of 5m, oats at an
even keel in fresh water at a draft of 2m when unloaded. The barge is transversely divided into three equal com-
partments. These compartments are uniformly loaded as follows:
No. 1 hold, 200 tonne; No. 2 hold, 155 tonne; No. 3 hold, 245 tonne
(Note: one metric ton, or tonne, is equal to 1000 kg, and the mass density of fresh water is 1 tonne/m
3
)
You will plot the loading intensity diagram, shear force diagram, and the bending moment diagram for the
loaded barge in a column format. Do not neglect the weight of the barge itself.
a) Since the moments of the weight about amidships are not equal for the loaded barge, the barge trims.
Assume the trim angle is small. Show from overall equilibrium that the draft at z = 0 is , and
that the draft at z = L = 30m is
b) Plot the loading intensity diagram, , where the loading intensity is in N/m, Newton/meter.
(The specic weight in N/m
3
is g times the mass density in kg/m
3
. If for simplicity g is taken as 10
(instead of 9.8) then specic weight of fresh water is .)
c) Determine the shear force, and plot it directly below the loading intensity diagram. Note; the dimen-
sional unit of the shear force is N, or Newtons.
d) Determine the bending moment, and plot it directly below the shear diagram. Note; the dimensional
units of the bending moment are Nm, or Newton-meters.
40 ft 40 ft 40 ft 40 ft 40 ft 40 ft
30 ft
400
tons
950
tons
400
tons
950
tons
empty empty
d
0
3.7 m =
d
L
4.3m =
p
y
z ( )
10m/sec
2
1000kg/m
3
10 000N/m
3
, =
d
0 d
L
200t
155t
245t
y
z
trim angle
2m
unloaded loaded
5m
30m
Axial Force, Shear Force and Bending Moment Diagrams
124 Thin-Walled Structures

Thin-Walled Structures

125

CHAPTER 5

Bending Deections of
Beams under Transverse
Loads

5.1 Approximations for slender beams

In CHAPTER 3 we discussed the special case of pure bending of a beam caused equal and opposite couples
applied to the ends of the beam. In pure bending, the bending moment interior to the beam is equal to the moment
of the applied end couples and the shear force in the beam vanishes. In this chapter, forces and/or distributed
loads act perpendicular to the reference axis causing transverse shear forces to develop interior to the beam. For
example, if a beam is subjected to a transverse distributed load of intensity , then the shear force

V

y

(z

) is
non-zero by equilibrium eq. (4.1) on p. 103 . Thus, moment equilibrium, eq. (4.3) on p. 104 , requires the bend-
ing moment to change with

z

along the beam. The presence of the transverse shear force causes plane cross sec-
tions not to remain plane in the deformed beam. Cross section of the beam will warp out of plane due to the
action of transverse shear. This invalidates the assumptions used for the deformation analysis of the beam under
pure bending (no shear) in Section 3.2. The analysis of the deformations associated with the presence of both
bending and transverse shear is best approached from the theory of elasticity. However, the elasticity analysis is
too complex for routine calculations, and in structural mechanics a simpler theory is used. The simpler theory to
compute stresses in beams under transverse loading is called engineering theory, or technical theory, or Ber-
noulli-Euler theory.
The engineering theory is known from actual practice, and from the few exact elasticity solutions available,
to provide reasonable estimates for long slender beams. As a general guide, the length of the beam should be
greater than ten times the largest cross-sectional dimension for the engineering theory to be applicable.
In the 5.2 we discuss the governing boundary value problem for the displacements of a beam subject to
transverse loads. Following this, complementary virtual work and complementary strain energy equations for



Assumption in the engineering theory of beams

In the engineering theory of beams it is assumed that the distribution of normal stress


z


given by the exure formula, eq. (3.27) on p. 77 , or equivalently eq. (3.28), is essentially
correct even in the presence of a nonuniform bending moment: i.e., when the shear force is
non-zero. This implies the deformations associated with shear are small, and consequently
neglected, with respect to deformations associated with bending.
p
y
z ( )

Bending Deections of Beams under Transverse Loads

126

Thin-Walled Structures

beam bending are derived. The complementary energy method leads to the generalized form of Castiglianos sec-
ond theorem for beams. Application of Castiglianos second theorem to compute beam displacements and rota-
tions at discrete points is illustrated in several examples.

5.2 Beam displacements

Neglecting the displacements due to transverse shear, the formulation of the pure bending problem discussed in
Section 3.1 and Section (3.2) is used to estimate beam displacements. Computing displacements is necessary in
design situations where there are limits on the maximum displacements. Also, in statically indeterminate beam
problems it is necessary to consider all three steps of the structural analysis to solve the problem:

equilibrium

,

strain-displacement

(geometry of deformation), and the

material law

. For beam bending, derivatives of the dis-
placements of the material points on the

z

-axis determine the curvature components of the

z

-curve in the
deformed beam. The curvatures of the

z

-curve, in turn, determine the normal strain distribution of the bers orig-
inally parallel to the

z

-axis in the undeformed beam.

Equilibrium

Let and , denote the lateral load intensities (

F/L

) in the

x

-direction and

y

-direction,
respectively, acting on the beam. These external load intensities are dened positive if they act in their respective
positive coordinate directions as shown in Fig. 5.1. Also shown in the gure, are the positive shear forces,
and , and the positive bending moments, and , acting on the positive

z

-face of the beam. Consider a
free body diagram of a differential element of the beam

dz-

long



obtained by cutting the beam parallel to the

x-y


plane at axial locations

z

and

z + dz.

Equilibrium conditions for bending in the

y-z

plane involve the sum of
forces in the

y

-direction and sum of moments about the

x

-axis of the differential element. In the limit as ,
the sum of forces leads to

(5.1)

and the sum of the moments leads to

(5.2)
p
x
z ( ) p
y
z ( )
p
x
p
y
V
x
V
y
M
x
M
y
x u ,
y v ,
z
Fig. 5.1 External distributed loads and internal actions
V
x
V
y
M
x
M
y
dz 0
dV
y
dz
--------- p
y
+ 0 =
dM
x
dz
----------- V
y
0 =

Thin-Walled Structures

127

Beam displacements

Equations (5.1) and (5.2) were derived in Chapter 2 as eqs. (2.1) and (2.3), respectively. Equilibrium conditions
for bending in the

x-z

plane involves the sum of forces in the

x

-direction and sum of moments about the

y

-axis of
the differential element. In the limit as , the sum of forces in the

x

-direction leads to

(5.3)

and the sum of moments about the negative

y

-direction leads to

(5.4)

Differentiating eq. (5.2) with respect to

z

, and then substituting eq. (5.1) into this result for the derivative of the
shear force gives

(5.5)

Similarly, differentiating eq. (5.4) with respect to

z

, and then substituting eq. (5.3) into this result for the deriva-
tive of the shear force gives

(5.6)

Strain-displacement

The normal strain of line elements parallel to the

z

-axis in the undeformed beam is deter-
mined in Section 3.2, and is given by eq. (3.15) on p. 74 . Repeating this result we have

(5.7)

in which the rotation of the projection of the

z

-curve into the

y-z

plane (Fig. 3.9) is

(5.8)

and the rotation of the projection of the

z

-curve into the

x-z

plane is

(5.9)

These rotations are depicted in Fig. 3.9 on page 74. In eq. (5.8) the displacement component of the material
points on the

z

-axis in the

y

-direction is denoted by function

v(z)

, and in eq. (5.9) the displacement in the

x

-direc-
tion is denoted by function

u(z)

. The derivatives of the rotations in eq. (5.7) are the curvature components of the
projections of the

z

-curve onto the

x-z

and

y-z

coordinate planes. Combining eqs. (5.7) to (5.9), we get

(5.10)

Material law

For a linear elastic, isotropic material the normal stress-strain relation is , where

E

is the
modulus of elasticity of the material. The normal strain distribution given in eq. (5.7) can be substituted into this
material law to get the normal stress distribution over the cross section due to bending. For a beam made of a
dz 0
dV
x
dz
--------- p
x
+ 0 =
dM
y
dz
----------- V
x
0 =
z
2
2
d
d M
x
p
y
z ( ) =
z
2
2
d
d M
y
p
x
z ( ) =

z
x
d
y
dz
--------
\ )
[
y
d
x
dz
--------
\ )
[
+ =

x
dv
dz
------ =

y
du
dz
------ =

z
x
z
2
2
d
d u

\ )
| j
[
y
z
2
2
d
d v

\ )
| j
[
+ =

z
E
z
=
Bending Deections of Beams under Transverse Loads
128 Thin-Walled Structures
homogenous material, it was shown in eq. (3.25) on p. 77 that statical equivalence of the normal stress distribu-
tion to the bending moments leads to
(5.11)
where I
xx
, I
yy
, and I
xy
are the second area moments about the centroidal x-y coordinates in the cross section, and
where the prime denotes an ordinary derivative with respect to z. The inverse of eq. (5.11) is
(5.12)
Now substitute the rotation-displacement relations, eqs. (5.8) and (5.9), into the material law, eq. (5.11), and
in turn substitute the moments from this result into equilibrium eqs. (5.5) and (5.6) to get
(5.13)
(5.14)
Equations (5.13) and (5.14) are the governing ordinary differential equations for the displacement functions u(z)
and v(z) in the open interval , where L is the length of the beam. These equations are coupled in the
sense that u(z) and v(z) appear in both equations if the product area moment I
xy
is non-zero. If the product area
moment is zero, then eq. (5.13) governs function v(z) alone and eq. (5.14) governs function u(z) alone. If the
bending stiffness terms , and are independent of coordinate z, then the governing equations
reduce to
(5.15)
(5.16)
The governing ordinary differential equations (5.15) and (5.16)are fourth order in each displacement func-
tion, which means that their solution will contain four arbitrary constants for each displacement function. These
constants are determined from the boundary conditions at z = 0 and z = L. Boundary conditions specify how the
beam is supported at each end. For bending in the y-z, plane specify at z = 0 and z = L
a) either displacement or shear force , but not both
b) either rotation or bending moment , but not both
The particular choices for the so-called standard boundary conditions for bending in the y-z plane are given in
Fig. 5.2. Other, more complex, boundary conditions exist in practice.
For bending in the x-z plane, specify at z = 0 and z = L
M
x
M
y
E
I
xx
I
xy
I
xy
I
yy

y
=

y
1
E I
xx
I
yy
I
xy
2
( )
-----------------------------------
I
yy
I
xy

I
xy
I
xx
M
x
M
y
=
z
2
2
d
d
EI
xx
z
2
2
d
d v
\ )
| j
[
z
2
2
d
d
EI
xy
z
2
2
d
d u
\ )
| j
[
+ p
y
z ( ) = 0 z L < <
z
2
2
d
d
EI
xy
z
2
2
d
d v
\ )
| j
[
z
2
2
d
d
EI
yy
z
2
2
d
d u
\ )
| j
[
+ p
x
z ( ) = 0 z L < <
0 z L < <
EI
xx
EI
yy
, EI
xy
EI
xx
z
4
4
d
d v
EI
xy
z
4
4
d
d u
+ p
y
= 0 z L < <
EI
xy
z
4
4
d
d v
EI
yy
z
4
4
d
d u
+ p
x
= 0 z L < <
v V
y

x
M
x
Thin-Walled Structures 129
Beam displacements
a) either displacement or shear force , but not both
b) either rotation or bending moment , but not both
The so-called standard boundary conditions for bending in the x-z plane are shown in Fig. 5.3.
v 0 = M
x
EI
xx
z
2
2
d
d v

\ )
| j
[
EI
xy
z
2
2
d
d u

\ )
| j
[
+ 0 = =
z

x
M
x
,
y v V
y
, ,
(1) simple support
z v 0 =
x
v 0 = =
(2) clamped
z
V
y
d
dz
----- EI
xx
z
2
2
d
d v

\ )
| j
[
EI
xy
z
2
2
d
d u

\ )
| j
[
+
\ )
| j
[
0 = =
(3) free
z
V
y
d
dz
----- EI
xx
z
2
2
d
d v

\ )
| j
[
EI
xy
z
2
2
d
d u

\ )
| j
[
+
\ )
| j
[
0 = =
x
0 =
(4) free in y
M
x
EI
xx
z
2
2
d
d v

\ )
| j
[
EI
xy
z
2
2
d
d u

\ )
| j
[
+ 0 = =
clamped about x
Fig. 5.2 Standard boundary conditions for bending in the y-z plane.
u V
x

y
M
y
Bending Deections of Beams under Transverse Loads
130 Thin-Walled Structures

It is possible to de-couple the displacements in the governing differential equations, but the problem may not
totally de-couple because the boundary conditions can involve both displacements. To get the de-coupling of the
differential equations, consider the case were the bending stiffnesses are uniform in the z-coordinate. Substitute
the rotation-displacement relations, eqs. (5.8) and (5.9), into the inverse form of the material law, eq. (5.12), to
get
(5.17)
where the second area moment ratios are dened by
(5.18)
u 0 = M
y
EI
xy
z
2
2
d
d v

\ )
| j
[
EI
yy
z
2
2
d
d u

\ )
| j
[
+ 0 = = z

y
M
y
,
x u V
x
, ,
(1) simple support
z u 0 =
y
u 0 = =
(2) clamped
z
V
x
d
dz
----- EI
xy
z
2
2
d
d v

\ )
| j
[
EI
yy
z
2
2
d
d u

\ )
| j
[
+
\ )
| j
[
0 = =
(3) free
z
V
x
d
dz
----- EI
xy
z
2
2
d
d v

\ )
| j
[
EI
yy
z
2
2
d
d u

\ )
| j
[
+
\ )
| j
[
0 = =
y
0 =
(4) free in x
M
y
EI
xy
z
2
2
d
d v

\ )
| j
[
EI
yy
z
2
2
d
d u

\ )
| j
[
+ 0 = =
clamped about y
x
x
x
Fig. 5.3 Standard boundary conditions for bending in the x-z plane.
v
u
1
ER
yy
------------
1
ER
xy
------------
1
ER
xy
------------
1
ER
xx
------------
M
x
M
y
=
R
xx
I
xx
I
yy
I
xy
2

I
xx
-------------------------- = R
yy
I
xx
I
yy
I
xy
2

I
yy
-------------------------- = R
xy
I
xx
I
yy
I
xy
2

I
xy
-------------------------- =
Thin-Walled Structures 131
Beam displacements
Now differentiate eqs.(5.17) twice with respect to z, and substitute eqs. (5.5) and (5.6) into this result for the sec-
ond derivatives of the moments to get
(5.19)
(5.20)
EXAMPLE 5.1 A statically indeterminate beam with an unsymmetrical cross section.
The uniform beam with a zee cross section is subjected to a uniformly distributed load of intensity in the pos-
itive y-coordinate direction as shown in Fig. 5.4. The beam is clamped in both the y-z plane and the x-z plane at z
= 0. At z = L the beam is simply supported in the y-z plane and clamped in the x-z plane. (The mathematical con-
ditions of support at z = L approximate a door butt hinge with the hinge axis parallel to the x-direction.) The sec-
ond area moments for the cross section are , , and , where dimension a is as
shown in the gure and t denotes the wall thickness. Determine the displacements and , and the bend-
ing moments M
x
(z) and M
y
(z).
Solution For this cross section the second area moment ratios dened by eqs. (5.18) are
The governing equation for bending in the y-z plane is obtained from eq. (5.19) as
(5.21)
and the boundary conditions are
(5.22)
z
4
4
d
d v
p
y
z ( )
ER
yy
-------------
p
x
z ( )
ER
xy
------------- = 0 z L < <
z
4
4
d
d u
p
y
z ( )
ER
xy
-------------
p
x
z ( )
ER
xx
------------- + = 0 z L < <
p
0
I
xx
8
3
---a
3
t = I
yy
2
3
---a
3
t = I
xy
a
3
t =
u z ( ) v z ( )
p
0
z
y,v
L
z
x,u
L
a
a
a
a
x
y
Fig. 5.4 Statically indeterminate beam with a zee cross section.
R
xx
7
24
------a
3
t = R
yy
7
6
---a
3
t = R
xy
7
9
---
\ )
[
a
3
t =
z
4
4
d
d v
p
0
ER
yy
------------ = 0 z L < <
v 0 ( ) 0 =
Bending Deections of Beams under Transverse Loads
132 Thin-Walled Structures
(5.23)
(5.24)
(5.25)
Note that boundary condition (5.25) requires that the displacement response u(z) be known. The governing equa-
tion for bending in the x-z plane is obtained from eq. (5.20) as
(5.26)
and the boundary conditions are
(5.27)
Note the boundary value problem for u(z) given by eqs. (5.26) and (5.27) is independent of the displacement
response v(z). Thus, we can solve eqs. (5.26) and (5.27) independent of the solution of eqs. (5.21) to (5.25). The
general solution to eq. (5.26) is
where c
1
,..., c
4
are arbitrary constants. The boundary conditions, eq. (5.27) at z = 0 require that constants c
3
= c
4

= 0. The boundary conditions at z = L lead to
These simultaneous equations are solved for constants c
1
and c
2
to get
Hence, the solution for u(z), after substitution for R
xy
, is
(5.28)
The general solution of eq. (5.21) for v(z) is
where c
5
,..., c
8
are arbitrary constants. Satisfaction of boundary conditions (5.22) and (5.23) requires c
7
= c
8
= 0.
Boundary condition (5.25)) can be re-written as
Substituting for v(z) and u(z), eq. (5.28), this boundary condition leads to

x
0 ( ) v 0 ( ) 0 = =
v L ( ) 0 =
M
x
L ( ) EI
xx
v ( ) EI
xy
u ( ) +
z L =
0 = =
z
4
4
d
d u
p
0
ER
xy
------------ = 0 z L < <
u u 0 = = at z = 0 and z = L
u
p
0
ER
xy
------------
z
4
24
------ c
1
z
3
6
---- c
2
z
2
2
---- c
3
z c
4
+ + + + =
u L ( ) 0 = ( )
L
3
6
-----c
1
L
2
2
-----c
2
+
p
0
L
4
24ER
xy
------------------ =
u L ( ) 0 = ( )
L
2
2
-----c
1
Lc
2
+
p
o
L
3
6ER
xy
--------------- =
c
1
p
0
L
2ER
xy
--------------- = c
2
p
0
L
2
12ER
xy
------------------ =
u z ( )
3p
0
56Ea
3
t
------------------z
2
L z ( )
2
=
v
p
0
ER
yy
------------
z
4
24
------ c
5
z
3
6
---- c
6
z
2
2
---- c
7
z c
8
+ + + + =
v L ( )
R
xx
R
xy
--------u L ( ) + 0 =
Thin-Walled Structures 133
Beam displacements
Boundary condition (5.24) leads to
Solving these last two equations for c
5
and c
6
, we get
Substituting for the second area moment ratios gives
(5.29)
Plots of the displacement distributions, eqs. (5.28) and (5.29), are shown in Fig. 5.5. The response of the beam in
the plane of loading (y-z plane) is represented by the lateral displacement . The out-of-plane lateral displace-
ment is non zero in this problem since the product area moment . If , then
. See eqs. (5.18) and (5.26). That is, the beam would not bend out of the plane of loading if
. However, as can be seen in Fig. 5.5, the displacement in the plane of loading is much larger than
the out of plane displacement in this example.
The bending moments are determined from the moment-curvature relations given by eqs. eq. (5.11). Substi-
tuting the solutions from eqs. (5.28) and (5.29) into these moment-curvature relations gives
The bending moment diagrams for M
x
and M
y
are shown in Fig. 5.6. Note that the moment for bending in the y-
z plane, or M
x
, is larger than the moment for bending in the x-z plane, or M
y
. Moment M
y
would vanish if
.
Lc
5
c
6
+
p
0
L
2
R
xx
12ER
xy
2
--------------------
p
0
L
2
2ER
yy
--------------- =
L
3
6
-----c
5
L
2
2
-----c
6
+
p
0
L
4
24ER
yy
------------------ =
c
5
L 5R
xy
2
R
xx
R
yy
+ ( ) p
0
8ER
xy
2
R
yy
------------------------------------------------------ = c
6
L
2
3R
xy
2
R
xx
R
yy
( ) p
0
24ER
xy
2
R
yy
--------------------------------------------------- =
v z ( )
p
0
896Ea
3
t
--------------------- z
2
39L
2
71Lz 32z
2
+ ( ) =
v z ( )
u z ( ) I
xy
0 I
xy
0 = u z ( ) 0 =
z 0 L , ( )
I
xy
0 = v z ( )
u z ( )
M
x
p
0
8
----- L
2
5Lz 4z
2
+ ( ) =
M
y
p
0
L
64
--------- L 3z ( ) =
I
xy
0 =
Bending Deections of Beams under Transverse Loads
134 Thin-Walled Structures

0.2 0.4 0.6 0.8 1
zL
0.0005
0.001
0.0015
0.002
0.0025
0.003
Ea
3
t up
0
L
4
0.2 0.4 0.6 0.8 1
zL
0.5
1
1.5
2
2.5
Ea
3
t vp
0
L
4
Fig. 5.5 Lateral displacements of the unsymmetrical section beam
Thin-Walled Structures 135
Beam displacements
EXAMPLE 5.2 Contact between two cantilever beams.
Consider two symmetrical section beams, one resting on top of the other, both clamped to a wall at z = 0.
The top beam is twice as long as the bottom beam, and the top beam is subjected to a downward force P at its tip.
Both beams have the same bending stiffness EI
xx
, which we abbreviate as EI here. See Fig. 5.7. Determine the
displacements due to bending for both beams assuming that they contact only at the tip of the shorter beam. Also
determine the bending moments and shear forces in each beam.
Solution Since the product area moment for each beam is zero, and neither beam is subjected to loads in the x-
direction, the response for each is bending in the y-z plane. Also, the distributed load intensity is zero for
each beam, because we have assumed that the contact between the two beams occurs only at one point. The gov-
erning equation for the displacement of the top beam is
Fi 5 6 B di t di f l 4 1
Fig. 5.6 Bending moment diagrams for example 4.1
0.2 0.4 0.6 0.8 1
zL
-0.03
-0.02
-0.01
0.01
M
y
p
0
L
2
0.2 0.4 0.6 0.8 1
zL
-0.1
-0.05
0.05
M
x
p
0
L
2
Fig. 5.6 Bending moment diagrams for the unsymmetrical section beam.
p
y
z ( )
Bending Deections of Beams under Transverse Loads
136 Thin-Walled Structures
Consequently, the displacement function in the top beam is
(5.30)
(5.31)
The boundary conditions are
(5.32)
and the conditions of continuity of displacement, rotation, and bending moment are
(5.33)
There is a jump in the shear force at z = L/2 due to contact with the lower beam given by
(5.34)
For the lower beam, the governing equation is
and the general solution of this equation is
(5.35)
The boundary conditions for the lower beam are
y, v
z
L/2 L/2
P
P
R
Fig. 5.7 Two cantilever beams in contact with
each other.
z
4
4
d
d v
0 = 0 z
L
2
--- < <
L
2
--- z L < <
v z ( ) v
11
z ( ) c
1
z
3
6
---- c
2
z
2
2
---- c
3
z c
4
+ + + = = 0 z
L
2
---
v z ( ) v
12
z ( ) c
5
z
3
6
---- c
6
z
2
2
---- c
7
z c
8
+ + + = =
L
2
--- z L
v 0 ( ) v 0 ( ) 0 = = M
x
L ( ) EIv L ( ) 0 = = V
y
L ( ) EIv L ( ) P = =
v
11
v
12
= v
11
v
12
= v
11
v
12
= at z
L
2
--- =
V
y
+
L
2
---
\ )
[
V
y
-
L
2
---
\ )
[
R + 0 =
\ )
[
EI v
12
L
2
---
\ )
[
EI v
11
L
2
---
\ )
[
R + + 0 =
z
4
4
d
d v
0 = 0 z
L
2
--- < <
v z ( ) v
2
z ( ) c
9
z
3
6
---- c
10
z
2
2
---- c
11
z c
12
+ + + = = 0 z
L
2
---
Thin-Walled Structures 137
Beam displacements
(5.36)
The condition that the beams are in contact at z = L/2 is
(5.37)
The unknown reaction force R can be eliminated between eq. (5.34) and the third of eqs. (5.36) to get
(5.38)
where we have divided by the common factor EI.
For displacement v
11
, eq. (5.30), to satisfy the rst two clamped boundary conditions of eqs. (5.32) constants
c
4
= c
3
= 0. For displacement v
12
, eq. (5.31), to satisfy the last two boundary conditions of eqs.(5.32) constants
c
5
= P/EI and c
6
= L P/EI. For the displacement of the lower beam v
2
, eq. (5.35), to satisfy the rst two clamped
end conditions of eqs. (5.36) constants c
11
= c
12
=0. For v
2
to satisfy the zero moment condition, third of eqs.
(5.36), constant c
10
= - c
9
L/2. Thus, to this point the displacements are
(5.39)
(5.40)
Substitute both of eqs. (5.39) into continuity of moments, second of eqs. (5.33), to get
Substitute the rst of eqs. (5.39) and eq. (5.40) into the contact condition, eq. (5.37), to get
Substitute the displacements eqs. (5.39) and (5.40) into the contact force continuity condition, eq. (5.38), to get
These last three equations are solved for c
1
, c
2
, and c
9
. We get , , and
.
Only c
7
and c
8
remain unknown. The conditions to determine these constants are the continuity conditions
of displacement and rotation, the rst and second of eqs. eq. (5.33). Using the results for c
1
, c
2
, and c
9
these con-
tinuity conditions become
v 0 ( ) v 0 ( ) 0 = = M
x
L
2
---
\ )
[
EIv
L
2
---
\ )
[
0 = = V
y
L
2
---
\ )
[
EIv
L
2
---
\ )
[
R = =
v
11
L
2
---
\ )
[
v
2
L
2
---
\ )
[
=
v
11
v
12
v
2
+
z
L
2
--- =
0 =
v
11
c
1
z
3
6
---- c
2
z
2
2
---- + = v
12
P
6EI
---------z
3
LP
2EI
---------z
2
c
7
z c
8
+ + =
v
2
c
9
z
3
6
----
Lz
2
4
--------
\ )
[
=
c
1
L
2
--- c
2
LP
2EI
--------- + + 0 =
L
3
48
------c
1
L
2
8
-----c
2
L
3
24
------c
9
+ + 0 =
c
1
c
9
P
EI
------ + 0 =
c
1
P 4EI ( ) = c
2
3LP ( ) 8EI ( ) =
c
9
5P ( ) 4EI ( ) =
L
2
---c
7
c
8
+
5L
3
P
96EI
------------- = c
7

5L
2
P
32EI
------------- + 0 =
Bending Deections of Beams under Transverse Loads
138 Thin-Walled Structures
Thus, we nd that and .
The solution is complete since all of the constants appearing in eqs. (5.30), (5.31), and (5.35) have been
determined. The displacement of the upper beam is
(5.41)
and the displacement of the lower beam is
(5.42)
The assumption that the lower beam contacts the upper beam only at its tip needs to be veried. The condi-
tion of no penetration of the contact between the upper and lower beam is
Substitute eqs. (4.41) and (4.42) into this inequality to get
Multiply this equation by the positive factor EI/P, and after rearrangement we nd
Thus, the assumption on the contact condition between the two beams is correct. A plot of the displacements of
the two beams is shown in Fig. 5.8. The bending moment in each beam is determined by the formula
. Using eqs. (5.41), we nd for the upper beam
and using eq. (5.42) we nd for the lower beam
c
7
5L
2
P ( ) 3EI ( ) = c
8
5L
3
P ( ) 192EI ( ) =
v
1
z ( )
3LP
16EI
------------z
2

P
24EI
------------z
3
0 z
L
2
---
5L
3
P
192EI
---------------
5L
2
P
32EI
-------------z
LP
2EI
---------z
2

P
6EI
---------z
3
+ +
L
2
--- z L
\
|
|
|
[
=
v
2
z ( )
5LP
16EI
------------z
2

5P
24EI
------------z
3
+ =
v
2
v
1
0 z
L
2
---
5LP
16EI
------------z
2

5P
24EI
------------z
3
+
3LP
16EI
------------z
2

P
24EI
------------z
3

z
2
4
----
L
2
--- z +
\ )
[
0 which is valid for 0 z L 2
M
x
z ( ) EIv z ( ) =
M
x1
P
8
--- 3L 2z + ( ) 0 z
L
2
---
P L z ( )
L
2
--- z L
|
|

|
|
=
M
x2
5P
8
------- L 2z ( ) = 0 z L 2
Thin-Walled Structures 139
Complementary virtual work and complementary energy
The bending moment distribution in each beam is shown in Fig. 5.9.
5.3 Complementary virtual work and complementary energy
Beam displacements can also be determined using complementary energy. In particular, the generalized form of
Castiglianos second theorem will be used in the examples in the following section. However, we rst establish
the complementary virtual work principle for the beam. Then for a beam made of a linear elastic material the
complementary strain energy expression is obtained.
0.2 0.4 0.6 0.8 1
z/L
-0.2
-0.15
-0.1
-0.05
v EI/(P L^3)
Fig. 5.8 The displacement functions of the upper beam (solid line) and the lower beam (dashed line)
for the two contacting cantilever beams.
0.2 0.4 0.6 0.8 1
z/L
0.1
0.2
0.3
0.4
0.5
0.6
M/(P L)
Fig. 5.9 The bending moment distributions in the upper beam (solid line) and the lower beam
(dashed line) for the two contacting cantilever beams.
Bending Deections of Beams under Transverse Loads
140 Thin-Walled Structures
5.3.1 Complementary virtual work
Consider combined bending of the beam in both the x-z plane as shown in Fig. 5.10, and in the y-z plane as
shown in Fig. 5.11. In these gures the end displacements and rotations are designated by , .
The set of end displacements and rotations are called generalized displacements. The word generalized in gen-
eralized displacements is used to represent collectively both displacements and rotations. Associated with the
generalized displacements are the corresponding generalized forces , . Generalized forces rep-
resent a set of point forces and moments. The generalized displacement is measured at the same point on the
beam as the corresponding generalized force, and both are measured positive in the same direction. The product
has the dimensional units of work, i.e., F-L. So if is a displacement, then is point force. If is a
rotation in radians, then is a point moment with dimensional units of F-L. The generalize displacements are
related to the beams displacements and , and beam rotations and as shown in the gures. Also, the
generalized forces are related to the beam shear forces and , and beam moments and , as shown in
the gures. From the possible boundary conditions for the beam, we either prescribe the generalized displace-
ment or the corresponding generalized force , but not both.
q
n
n 1 2 8 , , , =
Q
n
n 1 2 8 , , , =
q
n
q
n
Q
n
q
n
Q
n
q
n
Q
n
u v
x

y
q
1
Q
1
,
q
2
Q
2
,
z
x
L
V
x
M
y
q
5
Q
5
,
p
x
q
6
Q
6
,
u 0 ( ) q
1
=

y
0 ( ) q
2
=
V
x
0 ( ) Q
1
=
M
y
0 ( ) Q
2
=
u L ( ) q
5
=

y
L ( ) q
6
=
V
x
L ( ) Q
5
=
M
y
L ( ) Q
6
=
Fig. 5.10 Generalized displacements and forces for bending in the x-z plane.
V
x
V
y
M
x
M
y
q
3
Q
3
,
q
4
Q
4
,
z
y v ,
L
V
y
M
x
q
7
Q
7
,
p
y
q
8
Q
8
,
v 0 ( ) q
3
=

x
0 ( ) q
4
=
V
y
0 ( ) Q
3
=
M
x
0 ( ) Q
4
=
v L ( ) q
7
=

x
L ( ) q
8
=
V
y
L ( ) Q
7
=
M
x
L ( ) Q
8
=
Fig. 5.11 Generalized displacements and forces for bending in the y-z plane.
q
n
Q
n
Thin-Walled Structures 141
Complementary virtual work and complementary energy
To derive the complementary virtual work, we consider the generalized displacements to be prescribed.
Assume the beam displacements and , and rotations and , to be continuous, satisfying
compatibility eqs.(5.8) and (5.9), and to be consistent with the prescribed generalized displacements at z = 0 and
z = L. Known distributed load intensities and can act on the beam. Consider a variation in the gen-
eralized forces with the distributed loads unchanged. Virtual forces cause a variation in the internal
actions , , , and of the beam. Take these variations to satisfy the differential equations of
equilibrium for bending in the x-z plane given by eqs. (5.3) and (5.4); i.e.,
(5.43)
with the boundary conditions on the virtual forces for bending in the x-z plane given by
(5.44)
Similarly, from eqs. (5.1) and (5.2) the differential equilibrium conditions for bending in the y-z plane are
(5.45)
and the boundary conditions for bending in the y-z plane are
(5.46)
The external complementary virtual work performed by displacements acting through the virtual forces is
(5.47)
Substitute for the generalized displacements in eq. (5.47) the relationships to the beam displacements listed in
Fig. 5.10 and Fig. 5.11. Also substitute for the virtual forces in eq. (5.47) the boundary conditions given by eqs.
(5.44) and (5.46). After these substitutions we obtain
Write this last expression as
,
which is mathematically identical to
In the integrand of this equation, distribute the derivative of the sum to the sum of the derivatives of each term,
and then differentiate the product composing each term to get
u z ( ) v z ( )
x
z ( )
y
z ( )
p
x
z ( ) p
y
z ( )
Q
n
Q
n
V
x
M
y
V
y
M
x
z d
d
V
x
( ) 0 =
z d
d
M
y
( ) V
x
0 = 0 z L < <
V
x
0 ( ) Q
1
= V
x
L ( ) Q
5
=
M
y
0 ( ) Q
2
= M
y
L ( ) Q
6
=
z d
d
V
y
( ) 0 =
z d
d
M
x
( ) V
y
0 = 0 z L < <
V
y
0 ( ) Q
3
= V
y
L ( ) Q
7
=
M
x
0 ( ) Q
4
= M
x
L ( ) Q
8
=
W
ext
*
q
n
Q
n
n 1 =
8

=
W
ext
*
u 0 ( )V
x
0 ( )
y
0 ( )M
y
0 ( ) v 0 ( )V
y
0 ( )
x
0 ( )M
x
0 ( ) + =
u L ( )V
x
L ( )
y
L ( )M
y
L ( ) v L ( )V
y
L ( )
x
L ( )M
x
L ( ) + + +
W
ext
*
uV
x

y
M
y
vV
y

x
M
x
+ + + [ ]
0
L
=
W
ext
*
z d
d
uV
x

y
M
y
vV
y

x
M
x
+ + + [ ] z d
0
L

=
Bending Deections of Beams under Transverse Loads
142 Thin-Walled Structures
(5.48)
From the differential equations of equilibrium for the virtual forces, eqs. (5.43) and (5.45), eq. (5.48) reduces to
(5.49)
It was assumed that the displacements and rotations satised the compatibility conditions given by eqs. (5.8) and
(5.9), so the terms multiplying the virtual shear forces in eq. (5.49) vanish for all z. Hence, we get
(5.50)
The right-hand side of this equation contains quantities dened internal to the beam, so we dene the integral on
the right-hand-sid of eq. (5.50) as the internal complementary virtual work. Thus,
(5.51)
in which the prime denotes the derivative with respect to z.
The development from eq. (5.43) to (5.51) shows the external complementary virtual work equals the inter-
nal complementary virtual work for the beam having compatible displacements and rotations if the virtual com-
plementary forces are statically admissible. The principle of complementary virtual work asserts that the
displacements and rotations of the beam are compatible if the external complementary virtual work equals the
internal complementary virtual work for every statically admissible variation of the forces. If we were to apply
the principle of complementary virtual work to the beam problem, then we would obtain the compatibility condi-
tions for the displacements and rotations given by eqs. (5.8) and (5.9).
5.3.2 Complementary strain energy
The principle of complementary virtual work is independent of the material behavior. Now assume the material
of the beam is linear elastic and isotropic. Dene the integrand of the internal complementary virtual work, eq.
(5.51), as the incremental quantity ; i.e.,
(5.52)
For the linear elastic beam, the rotation gradients, or curvatures, were related to the bending moments by eq.
(3.72) on p. 92 . This equation is repeated below as eq. (5.53).
(5.53)
In eq. (5.53) , , and denote modulus-weighted second area moments, is a reference modulus of
W
ext
*
z d
du
V
x
z d
d
y
M
y
z d
dv
V
y
z d
d
x
M
x
+ + + +
0
L

=
u
z d
d
V
x
( )
y
z d
d
M
y
( ) v
z d
d
V
y
( )
x
z d
d
M
x
( ) + + + dz
W
ext
*
z d
du

y
+
\ )
[
V
x
z d
d
y
M
y
z d
dv

x
+
\ )
[
V
y
z d
d
x
M
x
+ + + z d
0
L

=
W
ext
*
z d
d
y
M
y
z d
d
x
M
x
+ z d
0
L

=
W
int
*

x
M
x

y
M
y
+ [ ] z d
0
L

U
0
*
U
0
*

x
M
x

y
M
y
+

y
1
E
0
I
xx
*
I
yy
*
I
xy
* 2
( )
----------------------------------------
I
yy
*
I
xy
*

I
xy
*
I
xx
*
M
x
M
x
t
+
M
y
M
y
t
+
=
I
xx
*
I
yy
*
I
xy
*
E
0
Thin-Walled Structures 143
Complementary virtual work and complementary energy
elasticity, and and are thermal moments due to a linearly varying thermal strain distribution over the
cross section. If the beam is made of a homogeneous material and the reference modulus is taken as the modulus
of the homogeneous material, then the modulus-weighted second area moments reduce to the geometric second
area moments of the cross-sectional area. If the change in temperature from the stress free state is spatially uni-
form over the cross section for the homogeneous material beam, then the thermal moments vanish, but the ther-
mal axial force is non-zero. Substitute eq. (5.53) for the rotation gradients in eq. (5.52) to get
(5.54)
in which
(5.55)
is the determinate of the matrix of modulus-weighted second area moments. (This determinate is positive for real
areas and materials.) The form of eq. (5.54) suggests there is a functional whose variation is pro-
vided by eq. (5.54). Assuming this functional exists, it is called the complementary strain energy density of the
beam having dimensional units of F-L/L. The rst variation of the functional is dened by
(5.56)
Equating eqs. (5.54) and (5.56), and recognizing equality holds for every admissible choice of virtual moments
and , we conclude that
(5.57)
and
(5.58)
From eqs. (5.57) and (5.58), the functional should be quadratic in its arguments and .
Assume that is of the form
(5.59)
in which are coefcients to be determined. Taking partial derivative of eq. (5.59), we get
(5.60)
(5.61)
Compare eq. (5.57) to eq. (5.60) to identify the coefcients
M
x
t
M
y
t
N
t
U
0
*
I
yy
*
M
x
M
x
t
+ ( ) I
xy
*
M
y
M
y
t
+ ( )
E
0
det I
*
( )
-------------------------------------------------------------------------- M
x
+ =
I
xy
*
M
x
M
x
t
+ ( ) I
xx
*
M
y
M
y
t
+ ( ) +
E
0
det I
*
( )
----------------------------------------------------------------------------- M
y
det I
*
( ) I
xx
*
I
yy
*
I
xy
*
( )
2
=
U
0
*
M
x
M
y
, [ ]
U
0
*
M
x
M
y
, [ ]
U
0
*
M
x

U
0
*
( ) M
x
M
y

U
0
*
( ) M
y
+ =
M
x
M
y
M
x

U
0
*
( )
I
yy
*
M
x
M
x
t
+ ( ) I
xy
*
M
y
M
y
t
+ ( )
E
0
det I
*
( )
-------------------------------------------------------------------------- =
M
y

U
0
*
( )
I
xy
*
M
x
M
x
t
+ ( ) I
xx
*
M
y
M
y
t
+ ( ) +
E
0
det I
*
( )
----------------------------------------------------------------------------- =
U
0
*
M
x
M
y
, [ ] M
x
M
y
U
0
*
M
x
M
y
, [ ]
U
o
*
A
11
M
x
2
A
12
M
x
M
y
A
22
M
y
2
A
1
M
x
A
2
M
y
+ + + + =
A
11
A
2
, ,
M
x

U
0
*
( ) 2A
11
M
x
A
12
M
y
A
1
+ + =
M
y

U
0
*
( ) A
12
M
x
2A
22
M
y
A
2
+ + =
Bending Deections of Beams under Transverse Loads
144 Thin-Walled Structures
(5.62)
Compare eq. (5.58) to eq. (5.61) to identify the coefcients
(5.63)
Note that coefcient is determined to be the same value from each comparison. If we have a function of the
form of eq. (5.59), then coefcient can be identied by the mixed partial derivatives in two ways; i.e.,
(5.64)
The order of differentiation is immaterial if all derivatives of the function concerned are continu-
ous. Since we started with the partial derivatives given by eqs. (5.57) and (5.58), a necessary check that the func-
tion exits, is that the mixed partial derivatives obtained from eqs. (5.57) and (5.58) have to be
equal. If the mixed partial derivatives obtained from eqs. (5.57) and (5.58) were not equal, then the complemen-
tary strain energy would not exist. Finally, substitute the coefcients given by eqs. (5.62) and (5.63) into eq.
(5.59) to get
(5.65)
Equation eq. (5.65) is the complementary strain energy density in the most complex beam we consider. If the
beam is made of a homogeneous material with modulus of elasticity E, the complementary strain energy density
is
(5.66)
If, in addition to homogeneity, the beam has a symmetric cross section so that the product area moment vanishes,
then the complementary strain energy is
(5.67)
Finally, if the beam is homogenous, has symmetric cross section, and there are no thermal strains, the comple-
mentary strain energy density is
(5.68)
From eqs. (5.51) and (5.52), the internal virtual work can be written as
Assuming a linear elastic material, the complementary strain energy density functional was shown to exist and is
given by eq. (5.65). Now interchange the integral operator and variational operator in the previous equation to get
A
11
I
yy
*
2E
0
det I
*
( )
--------------------------- = A
12
I
xy
*
E
0
det I
*
( )
------------------------ = A
1
I
yy
*
M
x
t
I
xy
*
M
y
t

E
0
det I
*
( )
------------------------------------ =
A
12
I
xy
*
E
0
det I
*
( )
------------------------ = A
22
I
xx
*
2E
0
det I
*
( )
--------------------------- = A
2
I
xy
*
M
x
t
I
xx
*
M
y
t
+
E
0
det I
*
( )
----------------------------------------- =
A
12
A
12
A
12
M
x

M
y

U
0
*
\ )
[
= or A
12
M
y

M
x

U
0
*
\ )
[
=
U
0
*
M
x
M
y
, [ ]
U
0
*
M
x
M
y
, [ ]
U
o
*
1
E
0
det I
*
( )
-------------------------
1
2
--- I
yy
*
M
x
2
2I
xy
*
M
x
M
y
I
xx
*
M
y
2
+ ( ) I
yy
*
M
x
t
I
xy
*
M
y
t
( )M
x
I
xy
*
M
x
t
I
xx
*
M
y
t
+ ( )M
y
+ + =
U
o
*
1
Edet I

( )
---------------------
1
2
--- I
yy
M
x
2
2I
xy
M
x
M
y
I
xx
M
y
2
+ ( ) I
yy
M
x
t
I
xy
M
y
t
( )M
x
I
xy
M
x
t
I
xx
M
y
t
+ ( )M
y
+ + =
U
o
*
1
2E
-------
M
x
2
I
xx
-------
M
y
2
I
yy
------- +
\ )
[
M
x
t
EI
xx
----------M
x
M
y
t
EI
yy
----------M
y
+ + =
U
o
*
1
2E
-------
M
x
2
I
xx
-------
M
y
2
I
yy
------- +
\ )
[
=
W
int
*
U
0
*
z d
0
L

=
Thin-Walled Structures 145
Beam displacement by Castiglianos second theorem
(5.69)
where the complementary strain energy of the beam is given by
(5.70)
5.4 Beam displacement by Castiglianos second theorem
Consider an elastic beam restrained against rigid body displacements and subjected to generalized forces ,
. Distributed loads and thermal strains can act on the beam as well. The generalized forces may
be independently varied since the variations in the reactions at the support points compensate to keep the struc-
ture in equilibrium. The generalized displacements corresponding to the generalized forces are denoted by ,
. The principle of complementary virtual work is
(5.71)
where the complementary energy for the beam is the integral over the length of the beam of the complemen-
tary strain energy density as expressed by eq. (5.70). The complementary strain energy density is written in terms
of the bending moments and as given by eq. (5.65). Following the development in Section 2.10
for the system of axially loaded bars, we use equilibrium conditions to determine the bending moments as func-
tions of z, , and in terms of the generalized forces . Hence, from eq. (5.65) the complementary strain
energy density is determined as a function of z with the generalized forces appearing as parameters in the func-
tion. Integration of the complementary strain energy density over the length of the beam, refer to eq. (5.70), gives
the complementary energy as a function of the generalized forces; i.e., . The variation
of the complementary strain energy becomes
(5.72)
Substituting eq. (5.72) into eq. (5.71), recognizing that the virtual generalized forces are independent, yields the
equation of the generalized form of Castiglianos second theorem; i.e.,
(5.73)
A statement of the theorem is given in Section 2.10.
EXAMPLE 5.3 End rotations of a simply supported beam subject to an end moment
A simply supported, uniform beam of length L is subjected to a moment at its left end as is shown in Fig.
5.12. The material is homogeneous and linear elastic, the cross section is symmetric ( ), and there are no
W
int
*
U
*
= elastic material only ( )
U
*
U
0
*
z d
0
L

=
Q
n
n 1 2 N , , , =
q
n
n 1 2 N , , , =
q
n
Q
n
n 1 =
N

U
*
=
U
*
M
x
z ( ) M
y
z ( )
0 z L Q
n
U
*
U
*
Q
1
Q
2
Q
N
, , , ( ) =
U
*
Q
n

U
*
\ )
[
Q
n
n 1 =
N

=
q
n
Q
n

U
*
=
Q
1
I
xy
0 =
Bending Deections of Beams under Transverse Loads
146 Thin-Walled Structures
thermal strains. The bending stiffness is EI. Use Castiglianos second theorem to determine the rotation at (a) the
left end, and (b) the right end.
Solution This beam is statically determinate. Moment equilibrium about
point z of the free body diagram of the left section of the beam gives
(5.74)
where is the reaction force at the support.
Part (a) Moment equilibrium of the free body diagram of the entire beam
gives the support reaction to be
Hence the moment in eq. (5.74) becomes
It is important to note that we have determined a moment distribution from equilibrium that contains only the
applied moment . That is, the support reaction is not an independent force, it is a function of . The com-
plementary strain energy is obtained from eqs. (5.68) and (5.70) with one of the two moments in eq. (5.68) iden-
tically zero. Castiglianos second theorem in the case is
Substitute for the bending moment in this equation the function determined from equilibrium. We nd
Performing the denite integration on the right-hand side of the last equation we get
z
Q
1
q
1
,
q
2
L
EI
Fig. 5.12 Simply supported beam subject to an end moment
z
Q
1 M
R
V
Free body diagram of left section
M Q
1
zR + =
R
L
Q
1
R R
(a) Free body diagram of entire beam
R
Q
1
L
------ =
M Q
1
1
z
L
---
\ )
[
=
Q
1
R Q
1
q
1
Q
1

M
2
2EI
--------- z d
0
L

M
EI
------
Q
1

M
\ )
[
z d
0
L

= =
q
1
1
EI
------ Q
1
1
z
L
---
\ )
[
1
z
L
---
\ )
[
z d
0
L

Q
1
EI
------ 1
z
L
---
\ )
[
2
z d
0
L

= =
q
1
Q
1
L
3EI
---------- = q
1
0 clockwise >
Thin-Walled Structures 147
Beam displacement by Castiglianos second theorem
Part (b) To nd the rotation at the right end by Castiglianos second theorem when no moment acts at that point,
we assume an external moment acts at the right end, apply the theorem to compute the corresponding rota-
tion , then set . That is,
(5.75)
Equation (5.74) for the moment is still valid in this case with two
moments applied to the beam. However, the support reaction is different
than in part (a). Moment equilibrium of the free body diagram for the entire
beam gives the support reaction as
Substitute this result for the support reaction into eq. (5.74) to nd the moment as
Again, it is important to note that we have determined a bending moment distribution from equilibrium that only
contains the applied moment and the ctitious external moment . Substitute this bending moment into eq.
(5.75) to get
Now we can set and obtain
Performing the denite integration on the right-hand side of this result we nd
Note that a clockwise moment applied at the left end of the beam results in a counter clockwise rotation
at the right end .
EXAMPLE 5.4 Tip displacement of a cantilever wing spar a under distributed load
Determine the vertical tip displacement of a cantilever wing spar subjected to the distributed lift load of inten-
sity , where is the intensity at the root and L is the span. See Fig. 5.13. The spar mate-
Q
2
q
2
Q
2
0 =
q
2
M
EI
------
Q
2

M
\ )
[
z d
0
L

Q
2
0 =
=
L
Q
1
R R
Q
2
(b) Free body diagram of entire beam
R
R
Q
1
Q
2
+ ( )
L
------------------------ =
M Q
1
Q
1
Q
2
+ ( )
z
L
--- + =
Q
1
Q
2
q
2
1
EI
------ Q
1
Q
1
Q
2
+ ( )
z
L
--- +
z
L
---
\ )
[
z d
0
L

Q
2
0 =
=
Q
2
0 =
q
2
1
EI
------ Q
1
1
z
L
---
\ )
[
z
L
---
\ )
[
z d
0
L

=
q
2
Q
1
L
6EI
----------
\ )
[
= q
2
0 clockwise >
Q
1
0 >
q
2
0 <
q
1
p
y
z ( ) p
0
1 z L ( ) = p
0
Bending Deections of Beams under Transverse Loads
148 Thin-Walled Structures
rial is linear elastic and homogeneous, and there are no thermal gradients. The product area moment of the cross
section is zero, and the bending stiffness is uniform along the span.
Solution Although there is no tip force acting on the beam, to use Castiglianos second theorem we imagine a
ctitious force corresponding to displacement to act on the spar. We determine the tip displacement via
the theorem, and then set . The complementary energy for this case is given by eq. (5.68) with
and eq. (5.70). In mathematical terms the theorem results in
(5.76)
where . The bending moment distribution is obtained by solving the differential equations of equilib-
rium, eqs. (5.1) and (5.2), subject to the boundary conditions on the shear force and bending moment at .
Substitute the distributed load intensity into eq. (5.1) to get
Indenite integration of this equation gives
The boundary condition on the shear force is , which allows for the determination of the constant of
integration; i.e.,
Hence, the shear force is
(5.77)
Substitute the shear force from eq. (5.77) into the equilibrium differential equation (5.2) for the moment to get
z
y v ,
p
y
z ( ) p
o
1
z
L
---
\ )
[
=
q
1
Q
1
,
L
EI
Fig. 5.13 Cantilever wing spar under a distributed lift load
EI
xx
EI =
Q
1
q
1
Q
1
0 = M
y
0 =
q
1
M
x
EI
-------
Q
1

M
x
\ )
| j
[
z d
0
L

Q
1
0 =
=
EI EI
xx
=
z L =
dV
y
dz
--------- p
0
1
z
L
---
\ )
[
=
V
y
p
0
z
z
2
2L
------
\ )
[
c
1
+ =
V
y
L ( ) Q
1
=
c
1
Q
1
p
0
L
2
--------- + =
V
y
z ( )
p
0
2L
------ z L ( )
2
Q
1
+ =
Thin-Walled Structures 149
Beam displacement by Castiglianos second theorem
Introduce a new independent variable by the denition
(5.78)
so that the differential equation for the bending moment becomes
From eq. (5.78) . Indenite integration of the above equation gives
The boundary condition for the bending moment at is . The value corresponds to .
This boundary condition determines that the constant of integration . Thus, the bending moment distribu-
tion in the equilibrium conguration under the distributed load and the ctitious force is
(5.79)
Substitute eq. (5.79) for the bending moment in eq. (5.76), recognize that the bending stiffness is indepen-
dent of z, to get
Now set , and change independent variables according to eq. (5.78), to write the expression for the tip
displacement as
Finally we get the result that
(5.80)
EXAMPLE 5.5 Strut-braced spar
Consider Example 5.4 again, but with a strut attached between the fuselage (simulated by a wall) and tip of the
spar as shown in Fig. 5.14. The strut is a two-force member, or truss bar, aiding to resist the vertical tip displace-
ment of the spar and to reduce the root bending moment in the spar. Determine the vertical displacement of the
tip of the spar using Castiglianos second theorem.
dM
x
dz
-----------
p
0
2L
------ z L ( )
2
Q
1
+ =
z L
dM
x
d
-----------
d
dz
------
p
0
2L
------
2
Q
1
+ =
d dz 1 =
M
x
p
0
6L
------
3
Q
1
c
2
+ + =
z L = M
x
0 = z L = 0 =
c
2
0 =
Q
1
M
x
p
0
6L
------ z L ( )
3
Q
1
z L ( ) + =
q
1
1
EI
------
p
0
6L
------ z L ( )
3
Q
1
z L ( ) + z L [ ] z d
0
L

Q
1
0 =
=
Q
1
0 =
q
1
p
0
6EIL
-------------
4
d
L
0

p
0
6EIL
-------------

5
5
-----
\ )
[
L
0
= =
q
1
p
0
L
4
30EI
------------ =
Bending Deections of Beams under Transverse Loads
150 Thin-Walled Structures
Solution The structural assembly is statically indeterminate. For
example, the overall free body diagram of the assembly has three
unknown forces of support and a moment as
shown in the adjacent sketch. For a coplanar force system there
are three independent equations of equilibrium. Hence, we have
one more unknown force than independent equations of equilib-
rium. This structural system is said to have one redundant force.
To proceed in using Castiglianos second theorem, remove the
strut from the fuselage support and replace the action of the sup-
port on the strut with force . Force is selected as the redun-
dant in this approach. In addition to the applied ctitious force
, we regard the redundant force as an applied force as well.
In this manner we have converted our structural assembly to be
statically determinate. Next we need to determine the complementary energy of the assembly in terms of the
independent forces and . (Note: support reactions are not independent, since the three
equilibrium equations obtained from the free body diagram of the assembly relate them to forces and .)
Then, we apply Castiglianos second theorem to determine the corresponding displacements and . How-
ever, the displacement at the end of the strut connected to the fuselage must be zero to correctly model the origi-
nal assembly. Also, the ctitious force is set to zero after performing the differentiations in Castiglianos
theorem. Hence, the mathematical statements of Castiglianos method to this structural assembly are
(5.81)
(5.82)
Complementary energy The complementary energy consists of the sum of the complementary energies stored
in each member of the assembly. The action of the distributed load on the spar is to bend it, so energy is stored in
the spar due to bending deformation. But bending of the spar upward causes its tip to displace upward and stretch
z
y v ,
p
y
z ( ) p
o
1
z
L
---
\ )
[
=
q
1
Q
1
,
L
EA EI ,

q
2
Q
2
,
EA ( )
b
Fig. 5.14 Strut-braced spar
z
y
p
y
z ( ) p
o
1
z
L
---
\ )
[
=
Q
2
R
z
R
y
C
x
Assembly free body diagram
R
y
R
z
Q
2
, , C
x
Q
2
Q
2
Q
1
Q
2
U
*
Q
1
Q
2
R
y
R
z
and C
x
, ,
Q
1
Q
2
q
1
q
2
Q
1
q
1
U
*
Q
1
----------
Q
1
0 =
=
q
2
U
*
Q
2
----------
Q
1
0 =
0 = =
Thin-Walled Structures 151
Beam displacement by Castiglianos second theorem
the strut. Hence, energy is stored in the strut due to extensional deformation. An axial tensile force in the strut,
results in a compressive force in the spar as a free body diagram of the joint at the spar to strut will show. Thus,
the spar is subject to compressive deformation as well as bending deformation. The energy stored in compression
in the spar must be added to the energy stored in bending. Let denote the axial force in the spar and denote
the axial force in the brace strut, both positive in tension. Then the complementary strain energy is
(5.83)
Equilibrium Free body diagrams of the joints between the strut and
fuselage and strut and spar tip are shown in the adjacent gure. Equilib-
rium at the joint between the strut and fuselage yields
(5.84)
Equilibrium at the joint between the spar and strut results in the follow-
ing boundary conditions at z = L in the spar.
(5.85)
(5.86)
(5.87)
The differential equations of equilibrium for the spar are
(5.88)
(5.89)
(5.90)
The solution to eq. (5.88) subject to boundary condition (5.85) is
(5.91)
in which we used eq. (5.84) to write the axial force in the strut in terms of . The solution to eqs. (5.89) and
(5.90) subject to boundary conditions (5.86) and (5.87) is the same procedure explained in Example 5.4. Omit-
ting the details, the solution for the bending moment in the spar is (refer to eq. (5.79))
(5.92)
Equations (5.84), (5.91), and (5.92) represent the equilibrium solutions in terms of the known distributed load
and the independent forces and . This step is crucial to using Castiglianos theorem.
N N
b
U
*
N
2
2EA
-----------
M
x
2
2EI
--------- +
\ )
[
z d
0
L

N
b
2
2 EA ( )
b
------------------
L
cos
------------
\ )
[
+ =
||||||||
||||||
spar strut
z
L
N
b
V
y
M
x
N

Q
1
Q
2
Free body diagrams of the joints
N
b
N
b
Q
2
=
N L ( ) N
b
cos 0 =
V
y
L ( ) Q
1
N
b
sin + 0 =
M
y
L ( ) 0 =
dN
dz
------- 0 = 0 z L < <
dV
y
dz
--------- p
0
1
z
L
---
\ )
[
+ 0 = 0 z L < <
dM
y
dz
----------- V
y
0 = 0 z L < <
N z ( ) Q
2
cos =
Q
2
M
x
z ( )
p
0
6L
------ z L ( )
3
Q
1
Q
2
sin ( ) z L ( ) + =
Q
1
Q
2
Bending Deections of Beams under Transverse Loads
152 Thin-Walled Structures
Evaluation of displacement equations Substitute the complementary strain energy, eq. (5.83), into eq. (5.81)
to get the following expression for displacement .
From eq. (5.91) , from eq. (5.92) , and from eq. (5.84) . Thus,
the equation for displacement reduces to
in which we used eq. (5.92) again to eliminate . Performing the denite integral on the right-hand-side
of the previous equation we get
(5.93)
The rst term on the left-hand side of this equation is the tip displacement obtained in Example 5.4 when there
was no strut. The second term on the left-hand-side of eq. (5.93) represents the reduction the tip displacement
due to the presence of the strut.
Next substitute the complementary strain energy, eq. (5.83), into eq. (5.82) to get the following expression
for displacement .
From eq. (5.91) , from eq. (5.92) , and from eq. (5.84)
. Use these same equations, (5.91), (5.92), and (5.84), evaluated for , to replace , ,
and , respectively, in the equation for displacement . The resulting expression for is
Evaluating some of the terms in this expression we get
q
1
q
1
N
EA
-------
N
Q
1
---------
\ )
[
M
x
EI
-------
M
x
Q
1
----------
\ )
[
+ z d
0
L

N
b
EA ( )
b
--------------
L
cos
------------
\ )
[
N
b
Q
1
----------
\ )
[
+
Q
1
0 =
=
N Q
1
0 = M
x
Q
1
z L = N
b
Q
1
0 =
q
1
q
1
1
EI
------
p
0
6L
------ z L ( )
3
Q
2
sin ( ) z L ( ) + z L [ ] z d
0
L

=
M
x
Q
1
0 =
q
1
p
0
L
4
30EI
------------
Q
2
sin ( )L
3
3EI
----------------------------- =
q
2
q
2
N
EA
-------
N
Q
2
---------
\ )
[
M
x
EI
-------
M
x
Q
2
----------
\ )
[
+ z d
0
L

N
b
EA ( )
b
--------------
L
cos
------------
\ )
[
N
b
Q
2
----------
\ )
[
+
Q
1
0 =
=
N Q
2
cos = M
x
Q
2
z L ( ) sin =
N
b
Q
2
1 = Q
1
0 = N M
x
N
b
q
2
q
2
q
2
Q
2
cos ( )
EA
--------------------------- cos ( )
1
EI
------
p
0
6L
------ z L ( )
3
Q
2
sin ( ) z L ( ) + z L ( ) sin [ ] +
| |

| |
z d
0
L

+ =
Q
2
EA ( )
b
--------------
L
cos
------------
\ )
[
1 ( )
q
2
Q
2
L cos
2
EA
-------------------------
p
o
sin
6EIL
----------------- z L ( )
4
z d
0
L

Q
2
sin
2
EI
-------------------- z L ( )
2
z d
0
L

Q
2
L
EA ( )
b
cos
--------------------------- - + + =
Thin-Walled Structures 153
Problems
Perform the denite integrations in this equation, rearrange the terms, and recall that to get
(5.94)
We can solve eq. (5.94) for the redundant force in terms of the distributed load intensity . Then, in
turn substitute the result for into eq. (5.93) to nd displacement .
5.5 Problems
1. The uniform cantilever beam shown below is subjected to a uniformly distributed load intensity and is also
restrained by a linear elastic spring at midspan. Neglect the weight of the beam, and determine
a) The displacement function . Plot the normalized displacement versus dimen-
sionless coordinate for .
b) .Plot the normalized bending moment for .
c) Plot the normalized shear force for .
2. A uniform beam with a rectangular cross section rests on a knife edge at its left end, while the right end is
clamped in rigid disk. The bending stiffness , the distance between the knife edge and the beams
connection to the disk is L and the radius of the disk is R. This disk rotates about a xed smooth pin through its
center under the action of applied moment M
a
as shown. Determine the relation between the applied moment M
a

and rotation angle of the disk under the assumption that the angle of rotation is small.
q
2
0 =
Q
2
L cos
2
EA
------------------
L
3
sin
2
3EI
--------------------
L
EA ( )
b
cos
--------------------------- - + +
p
0
L
4
30EI
------------ sin 0 =
Q
2
p
0
Q
2
q
1
v z ( ) v z ( )
v z ( )EI
xx
p
0
z
max
4
( )
--------------------- =
z z z
max
= 0 z 1
M
x
z ( )
M
x
p
0
z
max
2
---------------- = 0 z 1
V
y
z ( )
V
y
p
0
z
max
---------------- = 0 z 1
K
64EI
xx
z
max
3
----------------- =
z
p
y
z ( ) p
0
constant = =
z
max
2
z
max
y v ,
EI
xx
EI =
Bending Deections of Beams under Transverse Loads
154 Thin-Walled Structures

3. A coplanar bar is subjected to an end force as
shown. The bar is uniform with axial stiffness EA and
bending stiffness EI. Use Castiglianos second theorem
to nd
a) the end rotation , and
b) the vertical displacement at the joint.
M
a
,
L
R
z
y v ,
EI
xx
EI =

x
z
1
z
2
L
L
Q
1
q
2
q
3
Q
1
q
2
q
3

Thin-Walled Structures

155

CHAPTER 6

Shear Flow due to Shear
Forces

6.1 Shear ows and shear stresses due to bending in a rectangular
section beam

Consider bending under transverse loads of a rectangular section beam in the

y-z

plane as shown in Fig. 6.1. For
a slender beam, it is assumed that the normal stress is determined from the exure formula, eq. (3.27) on p. 77
, with sufcient accuracy when the beam is subjected to shear. The shear stress component exists if the shear
force

V

y

is present. Our purpose is to determine the shear stress in terms of the shear force.
The rst subscript on the symbol for the shear stress refers to the direction of the outward normal to the
face on which it acts, in this case the

z

-face, and the second subscript refers to the direction of the shear stress, in
this case the

y

-direction. A positive value for shear stress is dened as acting in the positive

y

-direction on a
positive

z

-face. By the action/reaction law, a positive value for shear stress means it acts the negative

y

-direc-
tion on a negative

z

-face. See the stress element in Fig. 6.1. Moment equilibrium of the innitesimal stress ele-

zy

yz

z
z
y
V
y
V
y
M
x
M
x
h/2
h/2
V
y
M
x
x
y
z
t
Fig. 6.1 Shear and normal stresses in a rectangular section beam under general bending.

zy

zy

zy

zy

Shear Flow due to Shear Forces

156

Thin-Walled Structures

ment about the

x

-direction requires the shear stress component acting on the

y

-face in the

z

-direction, or , be
non-zero and equal to ; i.e., . Shear stress component is sometimes called conjugate of . The
sign convention for shear stress is dened in the same manner as for shear stress . Shear stress compo-
nents and are shown acting in their positive senses on the stress element in Fig. 6.1. If the value of the
shear stress is negative, then the senses of the four arrows for the shear stresses in Fig. 6.1 are reversed.
In classical beam theory the shear stress component is determined by axial force equilibrium of free
body diagrams obtained from selected portions of the beam. Imagine a section of the beam obtained by cutting
the beam with planes perpendicular to the

z

-axis at

z

and

z

+


z

,


z

> 0, and a plane perpendicular to the

y

-axis
at a generic value of

y

, where . The free body diagram of the beam below the

y = constant

plane is
shown in Fig. 6.2. The bending normal stress


z

acts over a portion of the cross sectional area at

z

and

z

+


z

.
Because of the presence of the shear force, the bending normal stresses at

z

and

z

+


z

are different. The axial
force due to the stress distribution


z

acting over a portion of the cross-sectional area is

(6.1)

where we substituted the exure formula for the bending normal stress


z

. The integral in the last of these equa-
tions for

F

is the rst area moment of the lower portion of the cross-sectional area; i.e.,

(6.2)

Thus, eq. (6.1) becomes

(6.3)

The action of the top portion of the beam on the bottom portion is represented by the force . Notice that the
surface at is stress free, so that no force or shear stress acts on this surface.; i.e., at
. As , , since internal forces are only transmitted over nite areas. This force is

yz

zy

yz

zy
=
yz

zy

yz

zy

zy

yz

yz
h
2
--- y
h
2
---
z
y
h/2

z
z z + ( )

z
z ( )
F
yz
z z
F F +
F
Fig. 6.2 Free body diagram for a selected portion of the beam.
F
z
x d y d
t 2
t 2

h 2
y

M
x
I
xx
------- y x d y d
t 2
t 2

h 2
y

M
x
I
xx
------- yt y d
h 2
y

= = =
Q
x
y ( ) yt y d
h 2
y

t
2
--- y
2
h
2
---
\ )
[
2
= =
F
M
x
I
xx
-------Q
x
y ( ) =
F
yz
y h 2 =
yz
0 =
y h 2 = z 0 F
yz
0
Thin-Walled Structures 157
Shear ows and shear stresses due to bending in a rectangular section beam
given by
The integral of the shear stress across the width is dened as the shear ow and is denoted by q; i.e.,
(6.4)
Thus, the shear force F
yz
is
(6.5)
Divide this last equation by z and take the limit as to get
(6.6)
which shows that the shear ow is also interpreted as the shear force per unit length.
The free body diagram of the beam segment shown in Fig. 6.2 is correct with respect to equilibrium in the
axial direction. Equating the forces in the z-direction to zero we get
and if we divide this by z and take the limit as we get
Substitute eq. (6.3) for F and (6.6) for the second term in this equation we get
Moment equilibrium of the beam requires , so that the shear ow is given by
(6.7)
If we assume that the shear stresses are uniform across the width of the beam, then , and we get
(6.8)
Since , eq. (6.8) is also the formula for the shear stress in the cross section at y. It is important to note
that is the rst area moment of a portion of the cross-sectional area about the centroidal x-axis. For the
rectangular section, and is given by eq. (6.2), so that we nally obtain
F
yz

yz
x d
t 2
t 2

z d
z
z z +

=
q
yz
x d
t 2
t 2

=
F
yz
q z d
z
z z +

=
z 0
dF
yz
dz
----------- q =
F F F F
yz
+ + + 0 =
z 0
dF
dz
-------
dF
yz
dz
----------- + 0 =
dM
x
dz
-----------
Q
x
y ( )
I
xx
-------------- q + 0 =
V
y
dM
x
dz =
q
V
y
I
xx
------Q
x
y ( ) =
q
yz
t =

yz
V
y
Q
x
y ( )
I
xx
t
--------------------- =

zy

yz
=
Q
x
y ( )
I
xx
h
3
t ( ) 12 = Q
x
y ( )
Shear Flow due to Shear Forces
158 Thin-Walled Structures
The shear stress distribution is parabolic in the cross section, with a maximum value of 1.5 times the average
shear stress, where , and the shear stress is zero at the bottom and top of the beam
since at . This distribution is sketched in Fig. 6.3. The shear stress distribution is statically
equivalent to the shear force. That is,
6.2 Shear ows due to transverse shear forces in open section beams
Consider the thin-walled open cross section shown in Fig. 6.4. Let s denote the contour coordinate, which is
dened as the arc length of the center line of the wall. The s-direction is tangent to the contour. Let n denote the
thickness coordinate perpendicular to the contour. The thickness coordinate has the range , where
t denotes the thickness of the wall. Shear forces V
x
and V
y
are the resultants of the distribution of the shear stress

zy
3
2
---
V
y
ht
------
\ )
[
1
2y
h
------
\ )
[
2
=
h
2
--- y
h
2
---

ave
V
y
A V
y
ht ( ) = =

yz
0 = y h 2 +

zy

ave
---------
1.5
h/2
h/2
y
Fig. 6.3 Parabolic shear stress distribution in the rectangular section beam.
V
y

zy
x d
t 2
t 2

y d
h 2
h 2

3
2
---
V
y
ht
------
\ )
[
1
2y
h
------
\ )
[
2

| |

| |
x d
t 2
t 2

y d
h 2
h 2

V
y
= = =
Fig. 6.4 Shear forces, shear stresses, and shear ow in a general thin-walled open cross section
n
s
t
x
y
s
C
t
V
y
V
x

zn

zs
q
q
zs
n d
t
2
---
t
2
---

= 0
zn

zs

Assume
t 2 n t 2
Thin-Walled Structures 159
Shear ows due to transverse shear forces in open section beams
components and , where is acts tangent to the contour and acts in the direction normal to the con-
tour, that is, in the thickness direction. The shear stress component conjugate to is , and on the stress-free
lateral surfaces where component . Since , at as well. In thin
wall bar theory we neglect the thickness direction component with respect to the tangential component ,
because at and the wall is thin so that it appears that the thickness shear stress component is
not much different from zero. The shear ow is dened as the denite integral across the thickness of the shear
stress component tangent to the contour; i.e.,
(6.9)
The shear ow acts tangent to the contour at s, is positive if it acts in the positive s-direction, and has dimensional
units of F/L. It is a function of the contour coordinate s.
The shear ow at s, or q(s), can be related to the shear ow at s = 0, or q(0), by axial equilibrium. A free
body diagram of a beam segment dened by the portion of the contour from s = 0 to s, and the cross sections at z
and z + z is shown in Fig. 6.5. The axial force due to the bending normal stress
z
is
where the bending normal stress is evaluated on the contour (n = 0) rather than at a generic point through the
thickness of the wall in the thin-walled beam approximation. That is, we assume the distribution of the normal
stress through the thickness of the wall is small with respect to its value at the contour. Substituting the general
exure formula for
z
, eq. (3.27) on p. 77 , into the above equation we get
(6.10)
where the rst area moments with respect to the centroidal axes x and y are
(6.11)
Since the contour is a curve in the x-y plane, the x and y coordinates of the contour are related to the coordinate s;
i.e., and on the contour. These coordinate functions x(s) and y(s) are required to compute

zs

zn

zs

zn

zn

nz
n t 2 =
nz
0 =
zn

nz
=
zn
0 = n t 2 =

zn

zs

zn
0 = n t 2 =
q
zs
n d
t 2
t 2

=
F F +
F
z
s 0 =
q 0 ( )
q 0 ( )
s
q s ( )
q s ( )
Fig. 6.5 Free body diagram of a segment of a branch in the beam.
F
z
t s d
0
s

=
F
I
xx
M
y
I
xy
M
x

I
xx
I
yy
I
xy
2

-----------------------------------Q
y
s ( )
I
yy
M
x
I
xy
M
y

I
xx
I
yy
I
xy
2

-----------------------------------Q
x
s ( ) + =
Q
x
s ( ) y s ( )t s d
0
s

= Q
y
s ( ) x s ( )t s d
0
s

=
x x s ( ) = y y s ( ) =
Shear Flow due to Shear Forces
160 Thin-Walled Structures
the rst area moments dened by eqs. (6.11). Note that the thickness of the wall can be, in general, a function of
contour coordinates as well, as long as t(s) remains small with respect to the overall cross-sectional dimensions.
On the s-faces of the free body diagram in Fig. 6.5, the axial forces are approximated by the shear ow
times the small length z. The shear ows at s = 0 and at s are not the same. Note that the shear ow acting on the
positive z-face is assumed positive in the positive s-direction, so that the positive shear ow on the positive s-face
is in the positive z-direction and a positive shear ow on a negative s-face is in the negative z-direction. Setting
the sum of forces in the axial direction equal to zero in the free body diagram shown in Fig. 6.5 leads to
Divide this equation by z and then take the limit as to get
(6.12)
The derivative of eq. (6.10) is
where we used moment equilibrium of the beam about the x-axis, or , and about the y-axis, or
. Substituting this last result into eq. (6.12), we get the formula for the shear ow due to trans-
verse shear forces as
(6.13)
EXAMPLE 6.1 Shear ow distribution in a tee beam
A symmetric tee-section of dimensions h and t with h >> t is shown in Fig. 6.6. The section is subjected to a
shear forces . The second area moment and the product area moment .
Determine the shear ow distribution and the maximum shear stress magnitude.
F F F q s ( )z q 0 ( )z + + 0 =
z 0
dF
dz
------- q s ( ) q 0 ( ) + 0 =
dF
dz
-------
I
xx
V
x
I
xy
V
y

I
xx
I
yy
I
xy
2

--------------------------------Q
y
s ( )
I
yy
V
y
I
xy
V
x

I
xx
I
yy
I
xy
2

--------------------------------Q
x
s ( ) + =
V
y
dM
x
dz =
V
x
dM
y
dz =
q s ( ) q 0 ( )
I
xx
V
x
I
xy
V
y

I
xx
I
yy
I
xy
2

--------------------------------Q
y
s ( )
I
yy
V
y
I
xy
V
x

I
xx
I
yy
I
xy
2

--------------------------------Q
x
s ( ) =
V
x
0 V
y
0 > , = I
xx
5
24
------h
3
t = I
xy
0 =
s,q s,q
s,q
C
x
y
h
3
4
---h
h/2 h/2
t, typical
Fig. 6.6 Symmetric tee-section beam.
Thin-Walled Structures 161
Shear ows due to transverse shear forces in open section beams
Solution A separate contour coordinate is selected in the web, left ange, and right ange. Choosing a separate
contour coordinate in each branch, rather than using one contour coordinate for the entire cross section, is often a
convenience for subsequent computations. The direction of positive contour coordinate s and, consequently, the
sense of positive shear ow is as shown in Fig. 6.6. Note that we selected the origin of each branch contour coor-
dinate at a free edge, so that q(0) = 0 in each branch. For this particular section eq. (6.13) reduces to
In the web, the cartesian coordinates of the contour are
The rst area moment about the x-axis for the web, refer to the rst of eqs. (6.11), is
Substitute this into the shear ow formula to get
(6.14)
The web shear ow is quadratic in the branch contour coordinate s, is zero at s = 0, and . We set the
derivative of the shear ow with respect to s equal to zero to nd the location where the shear ow may have a
maximum magnitude in the web. Hence
This is the location of the centroid for the entire section, and .
The x and y coordinates of the contour in the left ange are

The rst area moment about the x-axis is
and hence the shear ow in the left ange is
(6.15)
q s ( )
V
y
I
xx
------Q
x
s ( ) =
x s ( ) 0 = y s ( )
3
4
---h s + = 0 s h
Q
x
s ( ) y s ( )t s d
0
s

3
4
---h s +
\ )
[
t s d
0
s

t
3
4
---hs
s
2
2
---- +
\ )
[
= = =
q s ( )
V
y
I
xx
------Q
x
s ( )
V
y
5
24
------h
3
t
--------------
3
4
---h
s
2
--- +
\ )
[
st
18
5
------
V
y
h
------ 1
2
3
---
s
h
---
\ )
[
s
h
--- = = = 0 s h
q h ( )
6
5
---
V
y
h
------ =
dq
ds
------ 0 =
\ )
[
2
3h
------
\ )
[
s
h
--- 1
2
3
---
s
h
---
\ )
[
1
h
--- + 0 = or s ,
3
4
---h =
q
3
4
---h
\ )
[
27
20
------
V
y
h
------ =
x s ( )
h
2
--- s + = y s ( )
h
4
--- = 0 s
h
2
---
Q
x
s ( ) y s ( )t s d
0
s

h
4
---
\ )
[
t s d
0
s

1
4
---hts = = =
q s ( )
V
y
5
24
------h
3
t
--------------
h
4
---ts
6
5
---
V
y
h
------
s
h
---
\ )
[
= = 0 s
h
2
---
Shear Flow due to Shear Forces
162 Thin-Walled Structures
Thus, the shear ow in the left ange is linear in the contour coordinate, and at the junction with the web
.
The x and y coordinates of the contour in the right ange are

The rst area moment about the x-axis is
and hence the shear ow in the right ange is
(6.16)
Thus, the shear ow in the right ange is linear in the contour coordinate and at the junction with the web
.
The shear ow distribution in the cross section is given by eqs. (6.14), (6.15), and (6.16). This distribution is
shown schematically in Fig. 6.7. The shear ows in the two anges are given by the same equation, and are neg-
ative for V
y
> 0. In Fig. 6.7 positive values of the ange shear ows are shown, which means that the arrow indi-
cating the sense of the shear ow in the anges is drawn in the opposite direction to what was originally assumed
positive. Note that the shear ow into the junction from the web is equal to the sum of the shear ows going out
of the junction into the anges. The physical basis of this ow relation at the junction is axial equilibrium at the
junction.
q
h
2
---
\ )
[
3
5
---
V
y
h
------ =
x s ( )
h
2
--- s = y s ( )
h
4
--- = 0 s
h
2
---
Q
x
s ( ) y s ( )t s d
0
s

h
4
---
\ )
[
t s d
0
s

1
4
---hts = = =
q s ( )
V
y
5
24
------h
3
t
--------------
h
4
---ts
6
5
---
V
y
h
------
s
h
---
\ )
[
= = 0 s
h
2
---
q
h
2
---
\ )
[
3
5
---
V
y
h
------ =
3
5
---
V
y
h
------
27
20
------
V
y
h
------
6
5
---
V
y
h
------
3
4
---h
q
q
q
Fig. 6.7 Shear ow distribution in the tee-section.
6
5
---
V
y
h
------
3
5
---
V
y
h
------
3
5
---
V
y
h
------
Shear flows at the junction
of the web and flanges
V
y
Thin-Walled Structures 163
Shear center of a thin-walled open section
The maximum shear ow magnitude is equal to , and occurs in the web at the centroid of the cross
section. Since each branch of the cross section has the same thickness, the maximum shear stress magnitude also
occurs at the same location as the maximum shear ow. Thus,
where the average shear stress is dened as the shear force V
y
divided by the cross-sectional area 2ht. The maxi-
mum shear stress is 2.7 times larger than the average shear stress value, or 170% larger with respect to the aver-
age!
6.3 Shear center of a thin-walled open section
Consider a thin-walled channel section with a web height denoted by h, with anges of equal width denoted by b,
and with the web and anges having the same thickness denoted by t. The section is subjected to a vertical shear
force V
y
> 0 and no horizontal shear, or V
x
= 0. See Fig. 6.8. The objective is to determine the shear ow distri-
bution in the cross section, and then use static equivalence to determine the resultant of the shear ow distribu-
tion.
The x-axis as shown in Fig. 6.8 is an axis of symmetry for the section, so that the product area moment I
xy
=
0. The second area moment about the x-axis is
The positive contour coordinate directions and shear ows in each branch of the cross section are dened as
shown in Fig. 6.8. To determine the shear ows for this problem, eq. (6.13) reduces to
27
20
------
V
y
h
------

max
27
20
------
V
y
ht
------
27
20
------2
V
y
2ht
--------
\ )
[
2.7
ave
= = =
x
y
C
s,q
s,q
s,q
x
y
C
b
b
h/2
h/2
t, typical
V
y
Fig. 6.8 Thin-walled channel section subjected to vertical shear.
I
xx
1
12
------h
3
t 2 0
h
2
---
\ )
[
2
bt + +
1
12
------h
3
t
1
2
---h
2
bt + = =
q s ( ) q 0 ( )
V
y
I
xx
------Q
x
s ( ) = where Q
x
s ( ) y s ( )t s d
0
s

=
Shear Flow due to Shear Forces
164 Thin-Walled Structures
In the lower ange, s = 0 at a free edge so that q(0) = 0, and for . Thus, the rst area
moment of the lower ange segment is , and the shear ow is
(6.17)
The web shear ow at the junction with the lower ange, q(0), is equal to the shear ow of the lower ange
at s = b. This, again, is a consequence of axial equilibrium at the junction. Hence, for the web we have:
from eq. (6.17), for , and
Combining these results, the shear ow in the web is
(6.18)
The upper ange shear ow at the junction with the web, q
web
(0), is equal to the web shear ow at s = h.
Hence, from eq. (6.18) we get . For the upper ange: for
, so . The shear ow in the upper ange is
(6.19)
Note that the upper ange shear ow at s = b is zero. This is as it should be, since the upper ange at s = b coin-
cides with a free edge. If this were not the case, then an error would have been made in the shear ow computa-
tions, and we would have to go back at this point to nd it. The shear ow distributions given by eqs. (6.17) to
(6.19) are shown schematically in Fig. 6.9(a).
Next we determine the resultant of the shear ows in several steps using static equivalence. First, for
straight branches we can integrate the shear ow along the contour to get the branch force. The line of action of
the branch force is the contour line. For the lower ange, the branch force is dened by . Substi-
tute eq. (6.17) for the shear ow to get
(6.20)
For the web, the branch force is dened by . Substitute eq. (6.18) for the shear ow to get
but , so that
y s ( ) h 2 = 0 s b
Q
x
s ( ) hts ( ) 2 =
q s ( )
V
y
I
xx
------
ht
2
-----s = 0 s b
q 0 ( ) V
y
hbt ( ) 2I
xx
( ) = y s ( ) h 2 s + = 0 s h
Q
x
s ( )
h
2
--- s +
\ )
[
t s d
0
s

t
2
--- hs s
2
+ ( ) = =
q s ( )
V
y
t
2I
xx
---------- hb hs s
2
+ ( ) = 0 s h
q
web
0 ( ) q
flange
b ( ) V
y
hbt ( ) 2I
xx
( ) = = y s ( ) h 2 =
0 s b Q
x
s ( ) hts ( ) 2 =
q s ( )
V
y
t
2I
xx
---------- hb hs ( ) = 0 s b
F
1
q s ( ) s d
0
b

=
F
1
V
y
I
xx
------
ht
2
-----s
\ )
[
s d
0
b

V
y
hb
2
t
4I
xx
----------------- = =
F
2
q s ( ) s d
0
h

=
F
2
V
y
t
2I
xx
---------- hb hs s
2
+ ( )
\ )
[
s d
0
h

V
y
t
2I
xx
---------- h
2
b
h
3
2
-----
h
3
3
----- +
V
y
I
xx
------
1
2
---h
2
bt
h
3
t
12
------- + = = =
I
xx
1
12
------h
3
t
1
2
---h
2
bt + =
Thin-Walled Structures 165
Shear center of a thin-walled open section
(6.21)
For the upper ange, the branch force is dened by . Substitute eq. (6.19) for the shear ow
to get
(6.22)
Note that the branch forces in each ange have the same magnitude, eqs. (6.20) and (6.22), have parallel lines of
action, but they have opposite senses. The branch forces are shown in Fig. 6.9(b), and these branch forces are
statically equivalent to the shear ows. Since the ange branch forces are the same magnitude, we let F
1
= F
3
=
F
ange
, and use to nd
(6.23)
The second step in the process of static equivalence is to reduce the coplanar force system consisting of
branch forces to a force and a couple at some convenient point in the cross section. Each time a force is moved to
a parallel line of action a couple is created to maintain static equivalence, as is shown below. If we choose the
center of the web to resolve the force system, we get
x
y
C
x
y
C
V
y
hbt
2I
xx
---------------
V
y
hbt
2I
xx
---------------
V
y
hbt
2I
xx
--------------- 1
h
4b
------ +
\ )
[
a) Shear flow distribution
F
2
V
y
=
F
3
V
y
hb
2
t
4I
xx
----------------- =
F
1
V
y
hb
2
t
4I
xx
----------------- =
b) Branch forces

Fig. 6.9 Shear ow distribution due to bending and the statically equivalent branch forces for the
channel section.
static equivalence
F
2
V
y
=
F
3
q s ( ) s d
0
b

=
F
3
V
y
t
2I
xx
---------- hb hs ( )
\ )
[
s d
0
b

V
y
hb
2
t
4I
xx
----------------- = =
I
xx
1
12
------h
3
t
1
2
---h
2
bt + =
F
flange
V
y
4
------
hb
2
t
1
12
------h
3
t
1
2
---h
2
bt +
-----------------------------------
\ )
| j
| j
[
V
y
3b
2
h
2
6hb +
---------------------
\ )
[
= =
Shear Flow due to Shear Forces
166 Thin-Walled Structures
(6.24)
where R
x
and R
y
are the x- and y-direction components of the resultant force, and C
z
is the moment of the couple
about the z-axis due to the force system. The resolution of the coplanar force system at the center of the web is
shown in Fig. 6.10(a). The system in Fig. 6.10(a) is statically equivalent to the planar force system in Fig.
6.9(b).
The third step in the process of static equivalence is to move the force V
y
to a parallel position to eliminate
the couple and maintain force equivalence. Force V
y
is moved such that the moment of V
y
in its new position
about the center of the web is equal to C
z
. This single force along a specic line of action is the resultant of the
coplanar force system. That is,
where e denotes the perpendicular distance between the lines of action of V
y
in its original position and its nal
position. Note that the force V
y
must be move to the left of its original position. The result of this nal reduction
is shown in Fig. 6.10(b). The resultant is a single force of magnitude V
y
in the positive y-direction, whose line of
action is parallel to the web at a distance e to the left of the web. The intersection of the line action of the result-
ant with the x-axis is called the shear center, which is abbreviated as S.C. in Fig. 6.10(b).
a
a
b
b
d
a
a
b
b
d
a
a
b
b
d
F
F
F
F
F
d F ( )
R
x
F
flange
F
flange
0 = = R
y
V
y
= C
z
hF
flange
=
eV
y
hF
flange
=
x
y
C
V
y
x
y
C
V
y
e
S.C.

hF
flange
a) Force and couple at the center of the web b) Resultant of the planar force system
static equivalence
e
hF
flange
V
y
--------------------
3b
2
h 6b +
--------------- = =
Fig. 6.10 Reduction of branch forces to a single resultant acting at the shear center
Thin-Walled Structures 167
Shear center of a thin-walled open section
The location of the shear center (S.C.) in the cross section is determined by the shear ows due to bending.
The computations given above for the channel section illustrate the general conclusion that the S.C. location
depends on the pattern of the shear ow distribution, and not on the magnitude of the shear force. Shear forces
act in the plane of loading to equilibrate the applied loads. (Recall that and .) The
line of action of the lateral loads and must pass through the shear center if the beam is to bend with-
out twisting about the z-axis. If the line of action of the transverse loads is not through the shear center, then the
loads will cause torsion in addition to bending. For open cross sections with straight branches and one junction,
the shear center is at the junction where all branch forces intersect; i.e., the moment of the branch forces is zero at
the common junction. Three simple open sections illustrating this point are shown in Fig. 6.11.
The location of the shear center for a closed section beam cannot be determined using only the steps outlined
above for the open section. In a closed section the shear ow at the origin of the contour in eq. (6.13) is an addi-
tional unknown. To determine the shear ow at the origin of the contour requires consideration of twist of the
section. The procedure to determine the shear center of a closed section is discussed in Shear center of a closed
section on page 219 .
EXAMPLE 6.2 Shear center location in an unsymmetrical section
For the thin-walled section shown in Fig. 6.12, determine the location of the shear center. The location of
the centroid is shown in Fig. 6.12, and the second area moments about the centroidal axes are
dV
y
dz
--------- p
y
z ( ) =
dV
x
dz
--------- p
x
z ( ) =
p
y
z ( ) p
x
z ( )
S.C.
S.C.
S.C.
Fig. 6.11 Shear center locations for open sections with straight branches and one junction
2a
a
2t
y
x
C
2a
t
2t
3a/8
a
y
x
C
s,q
A
free edge
Fig. 6.12 Unsymmetrical thin-walled channel section
Shear Flow due to Shear Forces
168 Thin-Walled Structures
Solution To nd the vertical location of the shear center we can consider shear forces and .
We do not need to compute the shear ows in each branch to nd the shear center in this example. A good choice
for the point in the cross section where we would resolve the branch forces can simplify the computations. It is
convenient to use the junction of the lower ange and web, labeled point A, as the point about which to sum
moments of the branch forces, since the only the branch force in the upper ange contributes to the moment
about point A. Then static equivalence of the moment of the branch force and V
x
about point A will give the loca-
tion of the line of action for shear force V
x
.
First the shear ow in the upper ange is determined using the positive senses for q and the contour coordi-
nate s as shown in Fig. 6.12. For V
y
= 0 and q(0) = 0, eq. (6.13) reduces to
(6.25)
The cartesian coordinates of the contour coordinate s in the upper ange are
(6.26)
The rst area moments of the portion of the contour from s = 0 to s of the upper ange are
(6.27)
(6.28)
The factor in the denominator of eq. (6.25) is . Substituting these results into eq.
(6.25) we get
(6.29)
The branch force in the upper ange is the integral of this shear ow over the upper ange; i.e.,
(6.30)
Static equivalence means that the moment of the shear force V
x
about point A must equal the moment of the
branch force F about point A, as is illustrated in the sketch below. Let e
y
denote the moment arm for the shear
force. Static equivalence for the moment about point A gives
(6.31)
Although the branch forces for lower ange and web have not been computed, they do not contribute to the
moment about point A. Substitute eq. (6.30) for F into this moment equivalence relation and solve for e
y
to get
I
xx
16
3
------a
3
t = I
yy
53
24
------a
3
t = I
xy
a
3
t =
V
x
0 > V
y
0 =
q s ( )
I
xx
Q
y
I
xy
Q
x
+ ( )V
x
I
xx
I
yy
I
xy
2

------------------------------------------------- =
x s ( )
5
8
---a s = y s ( ) a = 0 s a
Q
x
a ( )2t s d
0
s

2ast = =
Q
y
5
8
---a s
\ )
[
2t s d
0
s

2t
5
8
---as
s
2
2
----
\ )
[
= =
I
xx
I
yy
I
xy
2
97a
6
t
2
( ) 9 =
q s ( )
9
97
------
V
x
a
6
------ 2a
4
s
32
3
------a
3
5
8
---as
s
2
2
----
\ )
[
= 0 s a
F q s ( ) s d
0
a

23V
x
97
------------ = =
2aF e
y
V
x
=
Thin-Walled Structures 169
Shear center of a thin-walled open section
(6.32)
This result for e
y
locates the line of action of the shear force V
x
for beam bending. The shear center is on this par-
ticular line of action of V
x
, and we can locate the exact point on this line of action by considering the separate
problem of and .
Now consider and . Equation eq. (6.13) becomes
(6.33)
Substitute eq. (6.27) for Q
x
and eq. (6.28) for Q
y
into this equation to get the shear ow in the upper ange as
(6.34)
The branch force in the upper ange is
(6.35)
Moment equivalence about point A with the moment arm for shear force V
y
denoted as e
x
, gives
(6.36)
See the sketch below. Thus, the location of the line of action of the shear force V
y
is
A
A
2a
F
V
x
e
y
e
y
46
97
------a =
V
x
0 = V
y
0 >
V
x
0 = V
y
0 >
q s ( )
I
xy
Q
y
I
yy
Q
x
( )V
y
I
xx
I
yy
I
xy
2

-------------------------------------------- =
q s ( )
9
97
------
V
y
a
6
------
53
12
------a
4
s 2a
3
5
8
---as
s
2
2
----
\ )
[
= 0 s a
F q s ( ) s d
0
a

45
194
---------V
y
= =
2aF e
x
V
y
=
A
A
2a
F
V
y
e
x
Shear Flow due to Shear Forces
170 Thin-Walled Structures
(6.37)
The location of the shear center is shown in the sketch below.
6.4 Skin-stringer idealization
The purpose of the numerical example to be presented here is to justify the skin-stringer approximation, or ideal-
ization, to classical engineering beam theory. This approximation is especially well suited to semi-monocoque
(stiffened shell) construction; i.e., reinforced thin-walled structures typically found in vehicles where minimum
weight is important. The cross section of the beam is shown in Fig. 6.13, and we take l = 10 d, d = 2 inches, M
x

= 15,000 in-lb, and V
y
= 10,000 lbs for numerical evaluation. To study what happens when the web thickness t
varies we set t = d.
e
x
45
97
------a =
A
45
97
------a
46
97
------a
S.C.
d
l/2
d
l/2
d
t
x
y
centroid, C
Geometry of cross section Beam resultants
s
q(s)
x-axis
Web shear flow q(s)
Area A(s)
V
y
M
x
Fig. 6.13 Two-anged beam with thin web
Thin-Walled Structures 171
Skin-stringer idealization
The exure formula for the axial normal stress is , in which the second area moment of the cross
section about the x-axis is
(6.38)
The average normal stress in the ange is
(6.39)
The portion of the total moment carried by the anges is
or
(6.40)
The shear ow in the web is
(6.41)
The portion of the total shear force carried by the web is
or
(6.42)

z
M
x
I
xx
------- y =
I
xx
y
2
dA
A

tl
3
12
------ 2
d
4
12
------
l
2
---
d
2
--- +
\ )
[
2
d
2
+ + 83.33 60.67 + [ ]d
4
= = =

ave
M
x
I
xx
-------
l d + ( )
2
----------------
M
x
I
xx
------- 5.5d ( ) = =
M
x
( )
fl
2 y
z
dA ( )
A
fl

2 y
M
x
I
xx
------- y
\ )
[
d dy
l
2
---
l
2
--- d +
\ )
[

2
M
x
I
xx
-------d y
2
dy
l
2
---
l
2
--- d +
\ )
[

= = =
M
x
( )
fl
M
x
---------------
2
3
---
d
I
xx
------
l
2
--- d +
\ )
[
3
l
2
---
\ )
[
3
60.67 83.33 60.67 + ( )
1
= =
q s ( )
V
y
I
xx
------ ydA
A s ( )

V
y
I
xx
------
l
2
---
d
2
--- +
\ )
[
d
2
1
2
---
l
2
--- s +
\ )
[
l
2
--- s
\ )
[
t + = =
q s ( )
V
y
I
xx
------
1
2
--- l d + ( )d
2
t
2
---
l
2
---
\ )
[
2
s
2

\ )
[
+
V
y
I
xx
------ 5.5

2
--- 25
s
d
---
\ )
[
2

\ )
[
+ d
3
= =
V
y
( )
web

zy
tds
l
2
---
l
2
---

q s ( )ds
l
2
---
l
2
---

2 q s ( )ds
0
l
2
---

V
y
I
xx
------ l d + ( )d
2
t
l
2
---
\ )
[
2
s
2

\ )
[
+
| |

| |
ds
0
l
2
---

= = = =
V ( )
web
V
y
----------------
1
I
xx
------
1
2
--- l d + ( )ld
2
2
3
---t
l
2
---
\ )
[
3
+
55 83.33 +
60.67 83.33 +
------------------------------------ = =
Shear Flow due to Shear Forces
172 Thin-Walled Structures
These results are tabulated below for various values of .
A plot of the shear ow in the web for each in the table above is shown in Fig. 6.14.

The assumptions of the skin-stringer approximation are
The anges carry all of the bending moment, but no shear force. Note: a ange is considered a stringer.
Normal stress is uniform over each ange area. In other words, each ange area is assumed to be concentrated
at its centroid for calculation of total section centroid, I
xx
,
z
, and the rst area moment.
= t/d
I
xx
, in
4
eq. (6.38)
Flange

ave
, psi
eq. (6.39)
Flange
(M
x
)

/M
x
eq. (6.40)
Web shear ow q, lb/in, eq. (6.41)
(V
y
)
web
/V
y
eq. (6.42) s = 0 s = 5 in s = 10, in
1/4 1304.0 126.5 0.744 529.1 481.2 337.4 0.930
1/10 1104.0 149.5 0.879 489.1 466.5 398.6 0.918
1/50 997.33 165.4 0.973 461.2 456.2 441.2 0.909
0 970.67 170.0 1 453.3 453.3 453.3 0.907
0.7 0.8 0.9 1
qH500 lbinL
-1
-0.5
0.5
1
sH10 inL
Fig. 6.14 Web shear ow distributions as a function for decreasing web thickness to ange
width ratio.
1 4 =
1 10 =
1 50 =
0 =
Thin-Walled Structures 173
Skin-stringer idealization
The web carries all of the shear force, but no bending moment.
EXAMPLE 6.3 Shear ows in a stringer-stiffened C-section
The C-section shown Fig. 6.15 is stiffened by four stringers and subjected to a shear force with components
and . The stringers are assumed to carry only axial forces and each has the same cross-sec-
tional area . Each branch has a thickness t, and dimensions and .
Determine the shear ow distributions in each branch for and and plot them. Also
determine the maximum shear stress for each thickness.
A
f
= d
2
(l + d)/2
y
x
A
f
(l + d)/2
Only the transfer term in the parallel axis theorem is needed
to compute the second area moment about the x- axis.
The axial normal stress is uniformly distributed over the
stringer area, and the shear ow is uniformly distributed along
the web.
I
xx
2
l d +
2
-----------
\ )
[
2
A
f
968in
4
= =

z
M
x
I
xx
-------
l d +
2
-----------
\ )
[
170.45psi = =
q
V
y
I
xx
------
l d +
2
-----------
\ )
[
A
f
454.55 lb/in = =
V
x
0 = V
y
40 kN =
A
s
150 mm
2
= h 80 mm = b 20 mm =
t 5 mm = t 0.5 mm =
x
y
C
s
1
q
1
,
s
2
q
2
,
s
3
q
3
,
x
y
C
b
b
h
2
---
h
2
---
V
y
S.C.
t typical ,
A
s
A
s
A
s
A
s
Fig. 6.15 C-section stiffened by four stringers
branch shear flows

Shear Flow due to Shear Forces

174

Thin-Walled Structures

Solution

From eq. (6.13), the shear ow formula applicable to this problem is

(6.43)

To use this formula, we should rst compute the second area moment of the cross section. Using the thin wall
approximations and the composite body technique, the second area moment about the centroidal

x

-axis is
Note that the stringers are represented by their cross-sectional areas concen-
trated at their centroids, so that only the transfer term in the parallel axis theo-
rem affects the second area moment computation for the entire section. Begin
with determining the shear ow in the rst branch shown in Fig. 6.15, or the
lower ange. The shear ow at is not zero because of the presence of
the stringer. Axial equilibrium of the free body diagram of the lower right
stringer shown in the adjacent sketch gives

(6.44)

But the axial force in the stringer is , and the derivative of this force is .
From the exure formula , and the derivative . Thus,

(6.45)

where is the rst area moment of the lower right stringer about the

x

-axis given by .
Hence, the shear ow at is . Substitute this result for into eq. (6.43) to
get

(6.46)

The rst area moment of the portion of the branch 1 from to is , and the y-
coordinate to a generic point on the contour is . Hence,
Combining these results, the shear ow in branch 1 becomes

(6.47)
q s ( ) q 0 ( )
V
y
I
xx
------Q
x
s ( ) =
I
xx
I
xx
2
h
2
---
\ )
[
2
bt
h
3
t
12
------- 4
h
2
---
\ )
[
2
A
s
+ + =
dz
q
1
( )dz
0
dN
1
lower right stringer
s
1
0 =
q
1
0 ( )
dN
1
dz
---------- =
N
1

z1
A
s
= dN
1
dz d
z1
dz ( ) A
s
=

z1
M
x
h
2
---
\ )
[
I
xx
= d
z1
dz V
y
h
2
---
\ )
[
I
xx
=
dN
1
dz
----------
V
y
I
xx
------Q
xs1
=
Q
xs1
Q
xs1
h 2 ( ) A
s
=
s
1
0 = q
1
0 ( ) V
y
I
xx
( )Q
xs1
= q
1
0 ( )
q
1
s
1
( )
V
y
I
xx
------ Q
xs1
Q
x1
s
1
( ) + [ ] =
s
1
0 = s
1
Q
x1
s
1
( ) y
1
s
1
( )t s
1
d
0
s
1

=
y
1
s
1
( ) h 2 =
Q
x1
s
1
( )
h
2
--- ts
1
( ) = 0 s
1
b
q
1
s
1
( )
V
y
I
xx
------
h
2
---
\ )
[
A
s
ts
1
+ [ ] = 0 s
1
b
Thin-Walled Structures 175
Skin-stringer idealization
The shear ow in branch 2, or the web, is determined from the same gen-
eral formula given in eq. (6.43), and the shear ow at is determined
separately from the free body diagram of the lower left stringer shown in the
adjacent sketch. Axial equilibrium of the lower left stringer gives
(6.48)
The derivative of the axial force in the stringer is obtained following the same
procedure used for lower right stringer discussed above, which resulted in eq.
(6.45). The result for the lower left stringer is
(6.49)
where the rst area moment about the x-axis of stringer 2 is . Calculating the shear ow at the
end of branch 1 from eq. (6.46), using eq. (6.49) for the derivative the axial force in stringer 2, the shear ow at
the beginning of branch 2 from eq. (6.48) is
(6.50)
The term in brackets on the right-hand-side of eq. (6.50) represents the rst area moment about the centroidal x-
axis of the lower right stringer, the lower ange or branch 1, and the lower left stringer. For the web, or branch 2,
the y-coordinate to the generic point is , so the rst area moment of branch 2 is
It follows from this result, eqs. (6.43) and (6.50), that the shear ow in branch 2 is
(6.51)
For the upper ange, or branch 3, the general shear ow formula is again given by eq. (6.43). It follows from
the procedure used to obtain the beginning shear ow in branch 2 that
(6.52)
The y-coordinate to a generic point on branch 3 is . Hence, the rst area moment of the portion of
branch 3 is
It follows from this result and eqs. (6.43) and (6.52), that the shear ow in branch 3 is
dN
2
q
2
0 ( )dz
q
1
b ( )dz
lower left stringer
s
2
0 =
q
2
0 ( ) q
1
b ( )
dN
2
dz
---------- =
dN
2
dz
----------
V
y
I
xx
------Q
xs2
=
Q
xs2
h 2 ( ) A
s
=
q
2
0 ( )
V
y
I
xx
------ Q
xs1
Q
x1
b ( ) Q
xs2
+ + [ ] =
s
2
y
2
s
2
( ) h 2 s
2
+ =
Q
x2
s
2
( ) y
2
s
2
( )t s
2
d
0
s
2

t
2
--- hs
2
s
2
2
+ ( ) = =
q
2
s
2
( )
V
y
I
xx
------ hA
s
h
2
---tb
t
2
--- hs
2
s
2
2
+ ( ) + = 0 s
2
h
q
3
0 ( )
V
y
I
xx
------ Q
xs3
Q
x2
h ( ) Q
xs2
Q
x1
b ( ) Q
xs1
+ + + + [ ] =
y
3
s
3
( ) h 2 =
Q
x3
s
3
( ) y
3
s
3
( )t s
3
d
0
s
3

h
2
---ts
3
= =
Shear Flow due to Shear Forces
176 Thin-Walled Structures
(6.53)
Note that the shear ow at the end of branch 3 is . We can
check that this is correct by drawing the free body diagram of the upper right
stringer as shown in the adjacent sketch. Axial equilibrium per unit z-coordinate
gives
Similar to the results for the derivatives of the stringer forces given in eqs. (6.45) and (6.49), we have
. Axial equilibrium of the upper right stringer is
which, of course, is satised. If axial equilibrium of the upper right stringer were not satised, we would have to
go back over our previous calculations to search for an error.
Numerical evaluation and plots Given the data , , , and
, the shear ows from eqs. (6.47), (6.51), and (6.53) evaluate as
(6.54)
(6.55)
and
(6.56)
The dimensional units of the shear ows are N/mm for the thickness t and branch contour coordinates speci-
ed in mm. These shear ows are plotted versus the global section contour coordinate s in Fig. 6.16 for t = 5 mm
(solid lines) and for t = 0.5mm (dashed lines). The global contour coordinate is dened by

Note that the shear ow distributions exhibit jumps at the stringer locations, and that the shear ow distribution is
q
3
s
3
( )
V
y
I
xx
------
h
2
---
\ )
[
A
s
tb ts
3
+ [ ] = 0 s
3
b
dN
3
q
3
b ( )dz
dz
upper right stringer
q
3
b ( )
V
y
I
xx
------
h
2
--- A
s
\ )
[
=
q
3
b ( )
dN
3
dz
---------- + 0 =
dN
3
dz
----------
V
y
I
xx
------Q
xs3
=
V
y
I
xx
------
h
2
--- A
s

V
y
I
xx
------
h
2
--- A
s
+ 0 =
h 80 mm = b 20 mm = A
s
150 mm
2
=
V
y
40
3
10 N =
q
1
s
1
( )
40
3
10 ( ) 6000 4ts
1
+ ( )
960000
320000
3
------------------t +
-------------------------------------------------------- = 0 s
1
20 mm <
q
2
s
2
( )
40
3
10 ( ) 12000 800t 40ts
2
t
2
---s
2
2
+ +
\ )
[
960000
320000
3
------------------t +
----------------------------------------------------------------------------------------------- = 0 s
2
80 mm
q
3
s
3
( )
40
3
10 ( ) 6000 800t 40ts
3
+ ( )
960000
320000
3
------------------t +
---------------------------------------------------------------------------- = 0 s
3
20 mm
s s
1
= 0 s
1
20 mm
s s
2
20 mm + = 0 s
2
80 mm
s s
3
100 mm + = 0 s
3
20 mm
Thin-Walled Structures 177
Inuence of transverse shear deformations on bending
more uniform in each branch for the thinner section. The tendency to a spatially uniform shear ow in the thin-
walled branches between stringers corroborates with the earlier results for the thin web as shown in Fig. 6.14.
From inspection of the Fig. 6.16, the maximum shear ow occurs at the center of the web. Thus,
The maximum shear ow is reduced in the thinner section with respect to the thicker section. The maximum
shear stress is estimated from . Hence,
The maximum shear stress occurs for the thinner section. N.B. A shear stress magnitude of 1011 MPa is very
large. For example, aluminum alloy 2024-T4 has an ultimate shear stress of about 280 MPa. Clearly, an alumi-
num section of this alloy with t = 0.5 mm would fail.
6.5 Inuence of transverse shear deformations on bending
Classical beam theory neglects the effects of transverse shear stresses on the bending deformation of slender
beams. That is, classical theory assumes that plane cross sections before deformation remain plane and perpen-
dicular to the neutral axis (z-curve) in the deformed beam. For thin-walled beams the effect of transverse shear-
ing deformation can be signicant even for slender beams. An approximate method to account for transverse
shearing deformations is to assumed that cross sections remain plane but not necessarily perpendicular to the
neutral axis in the deformed beam. Again, we will assume the exure formula for the bending normal stress
obtained from pure bending is sufciently accurate in the presence of transverse shear, and consequently the
20 40 60 80 100 120
s, mm
100
200
300
400
500
q, Nmm
Fig. 6.16 Shear ows in the stiffened C-section for t = 5 mm (solid lines) and t =
0.5 mm (dashed lines)
q
max
q
2
40 ( )
535.7 N/mm t 5 mm =
505.3 N/mm t 0.5 mm =
|

|
= =

max
q
max
t =

max
107 MPa t 5 mm =
1011 MPa t 0.5 mm =
|

|
=
Shear Flow due to Shear Forces
178 Thin-Walled Structures
shear ow formula derived from equilibrium conditions using the exure formula is also sufciently accurate.
The beam theory that is developed in this section is sometimes called Timoshenko beam theory to distinguish it
from classical beam theory.
6.5.1 Transverse shear strains, forces, and complementary energy density
To motivate how we will account for transverse shearing deformations, consider the expression we obtained
for the internal complementary virtual work in Section 5.3. The internal complementary virtual work is given by
the right-hand-side of eq. (5.49) on p. 142 . Repeating this result we have
(6.57)
The factors multiplying the virtual shear forces in this equation, which are zero in the classical theory, represent
beam shear strains. Let denote the beam shear strain in the x-z plane and let denote beam shear strain in
the y-z plane. That is, we dene
(6.58)
These transverse shear strains and represent the reduction in the right angles between lines elements in
the deformed beam that were originally parallel to the x-axis and z-axis, and the y-axis and z-axis, respectively, in
the undeformed beam. The displacement gradients and , and rotations and are assumed to
be very small in magnitude with respect to unity. The beam shears are depicted in Fig. 6.17. Classical beam the-
ory is characterized by and for all values of z.
For an elastic material the integrand of eq. (6.57) is identied as the variation of the complementary strain
energy per unit length of the beam, or the complementary strain energy density . Using the denitions of
beam shear strains, the variation in the complementary strain energy density is written as
(6.59)
W
int
*
z d
du

y
+
\ )
[
V
x
z d
d
y
M
y
z d
dv

x
+
\ )
[
V
y
z d
d
x
M
x
+ + + z d
0
L

x

y

x
z d
du

y
+
y
z d
dv

x
+

x

y
du dz dv dz
y

x

x
0 =
y
0 =

2
---
y

dv
dz
------

x
w
v
z
y
(b) y-z plane

2
---
x

du
dz
------

y
w
u
z
x
(a) x-z plane
Fig. 6.17 Transverse shear strains in (a) the x-z plane, and (b) the y-z plane
U
0
*
U
0
*

x
V
x
z d
d
y
M
y

y
V
y
z d
d
x
M
x
+ + + =
Thin-Walled Structures 179
Inuence of transverse shear deformations on bending
We can identify the terms in this expression for the complementary strain energy density as originating from
transverse shear deformation or bending deformation. Decompose the complementary strain energy density as
(6.60)
where is the complementary strain energy density due to transverse shear deformation and is the com-
plementary strain energy density due to bending deformation. Compare the variation of eq. (6.60) to eq. (6.59) to
get
(6.61)
and
(6.62)
The derivation of the complementary strain energy density due to bending for a linear elastic, homogeneous
material was obtained in Section 5.3 as eq. (5.66) on p. 144 . We do not need to repeat this derivation here, since
this beam theory, as does classical beam theory, assumes that the bending normal stresses are given by exure
formula even in the presence of transverse shear deformations. Neglecting the thermal strain terms, the comple-
mentary strain energy density due to bending is obtained as
(6.63)
where and E is the modulus of elasticity.
To obtain an expression for the complementary strain energy density due to shear for a linear elastic mate-
rial, the form of eq. (6.61) suggests that there is a functional of the shear forces, , whose variation by
denition is
(6.64)
Comparing eqs. (6.61) and (6.64), the beam shear strains can be identied as
(6.65)
For linear material behavior the beam shear strains are linearly related to the shear forces by expressions of the
form
(6.66)
where the coefcients , , , and are shear compliances of the beams cross section having dimen-
sional units of 1/F. From eqs. (6.65) and (6.66), the complementary strain energy density due to shear is deter-
mined to be
(6.67)
U
0
*
U
0
*
s
U
0
*
b
+ =
U
0
*
s
U
0
*
b
U
0
*
s

x
V
x

y
V
y
+ =
U
0
*
b
z d
d
y
M
y
z d
d
x
M
x
+ =
U
0
*
b
1
2Edet I

( )
------------------------ I
yy
M
x
2
2I
xy
M
x
M
y
I
xx
M
y
2
+ ( ) =
det I ( ) I
xx
I
yy
I
xy
2
=
U
0
*
s
V
x
V
y
, [ ]
U
0
*
s
V
x
V
y
, [ ]
U
0
*
s
V
x
------------V
x
U
0
*
s
V
y
------------V
y
+ =

x
U
0
*
s
V
x
------------ =
y
U
0
*
s
V
y
------------ =

x
c
xx
V
x
c
xy
V
y
+ =

y
c
yx
V
x
c
yy
V
y
+ =
c
xx
c
xy
c
yx
c
yy
U
0
*
s
1
2
--- c
xx
V
x
2
2c
xy
V
x
V
y
c
yy
V
y
2
+ + ( ) =
Shear Flow due to Shear Forces
180 Thin-Walled Structures
where if eq. (6.66) is to be derived from eq. (6.67) using the relations in eq. (6.65). The question now
is how do we obtain expression for the shear compliances , , and ? To derive these expressions we
return to the free body diagram of an element of the wall of the beam as was used to derived the shear ow for-
mula in Section 6.2 given as eq. (6.13) on p. 160 .
6.5.2 Complementary energy density obtained from a two-dimensional element of the wall
The complementary strain energy per unit length of the beam is derived again in this subsection, but we start
from a differential element cut from the wall of the beam of dimensions , where t is the wall thick-
ness. Let denote the normal stress resultant dened as the integral of the normal stress over the thickness of
the wall; i.e.,
(6.68)
where n is the coordinate normal to the contour. If the wall is thin with respect to overall cross-sectional dimen-
sions of the beam, the distribution of the normal stress over the thickness of the wall can be ignored and the nor-
mal stress is represented by a uniform distribution through the thickness equal to its value on the contour. For
thin walls then, the normal stress resultant is approximated by where is evaluated on the contour at s.
A free body diagram of an element of the wall of the beam is shown in Fig. 6.18. The stress resultant acting
normal to the s-face of the element is assumed to be negligible with respect to the axial resultant in beam the-
ory, and so it is not shown in the free body diagram. Sum the forces acting in the z-direction to zero yields
In the limit as the element shrinks to zero dimensions we obtain from this equation the differential equation of
equilibrium
(6.69)
c
xy
c
yx
=
c
xx
c
xy
c
yy
ds dz t
n
z
n
z

z
n d
t 2
t 2

=
n
z

z
t
z
ds
dz
q
q
s
------ds +
\ )
[
dz
qdz
q
q
z
------dz +
\ )
[
ds
qds
n
z
n
z
z
--------dz +
\ )
[
ds
n
z
ds
s v
t
,
n
x
y
z w ,
s v
t
,
n

Fig. 6.18 Free body diagram of an innitesimal element of the beam wall.
n
s
n
z
n
z
n
z
z
--------dz +
\ )
[
ds n
z
ds q
q
s
------ds +
\ )
[
dz qdz + 0 =
n
z
z
--------
q
s
------ + 0 =
Thin-Walled Structures 181
Inuence of transverse shear deformations on bending
Summing forces to zero in the s-direction yields
In the limit as the element shrinks to zero dimensions, we obtain from s-direction equilibrium that
(6.70)
The displacement of the element in the z-direction is denoted and the displacement in the s-direc-
tion, or in the direction tangent to the contour is denoted by .The element strains of interest are the axial
normal strain
(6.71)
and the shear strain
. (6.72)
The shear strain is depicted in the adjacent sketch. These displacements must be
continuous and single valued so that no gaps or overlaps of material occur in the
deformed conguration. Continuous, single-valued displacements and strains
satisfying the strain-displacement relations (6.71) and (6.72) are said to be com-
patible.
Now assume the beam element has displacements and rotations that are
compatible and satisfy eqs. (6.71) and (6.72). From this compatible deformation
state consider a virtual change in the normal stress resultant and a virtual
change in the shear ow , that satisfy equilibrium conditions (6.69) and
(6.70); i.e.,
(6.73)
and
(6.74)
Equation (6.74) implies that the virtual shear ow is spatially uniform in the z-coordinate. A shear ow that is
uniform along the length of the beam occurs in the case of a cantilevered beam subjected to transverse forces at it
tip with no distributed loads acting on the beam. Hence, we regard the virtual force system acting on the beam to
consist of only virtual transverse forces acting at the tip of a cantilevered beam, so that the virtual shear forces
and are independent of z. Although this may seem like a limiting situation in which to derive the shear
compliances, in the approximate transverse shear theory being developed here the shear compliances obtained
for this case are assumed applicable to other beam loading situations as well. The complementary virtual work
per unit area of the element is denoted by , and is determined by the displacements acting through the vir-
tual stress resultants. That is, the complementary virtual work of the stress resultants acting on the innitesimal
element is written as (refer to Fig. 6.18)
q
q
z
------dz +
\ )
[
ds qds 0 =
q
z
------ 0 =
w z s , ( )
v
t
z s , ( )

z
w z =

zs
w s v
t
z + =
z w ,
s v
t
,
v
t
z
-------
w
s
-------
2
zs

wall element shear strain


n
z
q
n
z
( )
z
----------------
q ( )
s
-------------- + 0 =
q ( )
z
-------------- 0 =
q
V
x
V
y
U

0
*
Shear Flow due to Shear Forces
182 Thin-Walled Structures
(6.75)
Divide this equation by the differential area , and take the limit as and to get
Distribute the derivative in this last expression and write it as
(6.76)
Since the virtual stress resultants satisfy equilibrium conditions (6.73) and (6.74), the rst two terms on the right-
hand side of eq. (6.76) vanish. Hence,
(6.77)
in which the axial normal strain and the shear strain are identied from eqs. (6.71) and (6.72). For a Hookean
material the strain-stress relations are
(6.78)
where E is the modulus of elasticity, t the wall thickness, and G is the shear modulus. Using this material law in
eq. (6.77) we get that the complementary strain energy per unit area is
(6.79)
Equation (6.79) is integrated over the contour of the cross section from to , where S is the arc
length of the contour, to obtain the complementary strain energy per unit length of the beam. That is, the varia-
tion of the complementary strain per unit length is determined from the variation of the complementary strain
energy per unit area by the denite integral
(6.80)
Substitute eq. (6.79) into this result to get
(6.81)
The two terms in this complementary strain energy density expression can be identied with bending and shear-
ing deformations. The bending deformation effect is represented by the term
(6.82)
Recall that in the thin wall approximation, and that the bending normal stress is given by the ex-
ure formula, eq. (3.28) on p. 77 . If we substituted the exure formula for and its variation into eq. (6.82) and
carried out the details to obtain the complementary strain energy density due to bending, then we would obtain
U

0
*
dsdz w n
z
ds ( ) [ ]
z
z dz +
v
t
qds ( ) [ ]
z
z dz +
w qdz ( ) [ ]
s
s ds +
+ + =
dsdz ds 0 dz 0
U

0
*
z

wn
z
v
t
q + [ ]
s

wq [ ] + =
U

0
*
w
n
z
( )
z
----------------
q ( )
s
-------------- +
\ )
[
v
t
q ( )
z
--------------
w
z
-------n
z
v
t
z
-------
w
s
------- +
\ )
[
q + + + =
U

0
*

z
n
z

zs
q + =

z
1
Et
-----n
z
=
zs
q
Gt
------ =
U

0
*
1
Et
-----n
z
n
z
1
Gt
------qq + =
s 0 = s S =
U
0
*
U

0
*
s d
0
S

=
U
0
*
1
Et
-----n
z
n
z
1
Gt
------qq + s d
0
S

=
U
0
*
b
n
z
Et
-----n
z
s d
0
S

=
n
z

z
t =
z

z
Thin-Walled Structures 183
Inuence of transverse shear deformations on bending
eq. (6.63). A result we have already derived. It is the second term in eq. (6.81) that is of interest here. This second
term is the variation of the complementary strain energy density due to transverse shear deformation given by
(6.83)
6.5.3 Determination of the transverse shear compliances
Two expressions for the variation of the complementary strain energy density due to transverse shear deforma-
tion have been derived; eqs. (6.61) and (6.83). Equating these results we have
(6.84)
From axial equilibrium of an innitesimal element of the wall of the beam, and using the exure formula for the
bending normal stress, we derived the shear ow formula which is given in eq. (6.13). We write this formula in a
slightly different form from what is given in eq. (6.13) as
(6.85)
For an open section having a free edge, the origin of the contour coordinate s can be taken at the free edge so that
. If a stringer is located at the contour origin of a branch, then the shear ow is obtained in terms
of shear forces and/or as is shown in Example 6.3. Consequently, we can write eq. (6.85) in the form
(6.86)
where
(6.87)
and
(6.88)
The variation of the shear ow, or virtual shear ow, is obtained from eq. (6.86) as
(6.89)
Substituting eqs. (6.86) and (6.89) into the right-hand side of eq. (6.84) yields
(6.90)
Equation eq. (6.90) must hold for every virtual force and . Recall that we are considering a cantilever
beam subject to virtual forces equal to and applied at its tip. These virtual forces are independent.
Hence, we conclude from eq. (6.90) that
U
0
*
s
q
Gt
------q s d
0
S

x
V
x

y
V
y
+
q
Gt
------q s d
0
S

=
q s ( ) q 0 ( )
I
xx
Q
y
s ( ) I
xy
Q
x
s ( )
det I ( )
------------------------------------------------V
x

I
yy
Q
x
s ( ) I
xy
Q
y
s ( )
det I ( )
------------------------------------------------V
y
=
q 0 ( ) 0 = q 0 ( )
V
x
V
y
q s ( ) K
x
s ( )V
x
K
y
s ( )V
y
=
K
x
s ( )
I
xx
Q
y
s ( ) I
xy
Q
x
s ( )
det I ( )
------------------------------------------------ =
K
y
s ( )
I
yy
Q
x
s ( ) I
xy
Q
y
s ( )
det I ( )
------------------------------------------------ =
q K
x
s ( )V
x
K
y
s ( )V
y
=

x
V
x

y
V
y
+
K
x
s ( )V
x
K
y
s ( )V
y
( )
Gt
-------------------------------------------------------- K
x
s ( )V
x
K
y
s ( )V
y
[ ] s d
0
S

=
V
x
V
y
V
x
V
y
Shear Flow due to Shear Forces
184 Thin-Walled Structures
(6.91)
and
(6.92)
Compare these expressions for the beam shear strains to those given in eq. (6.66), to nally get the expressions
for the shear compliances. These compliance formulas are
(6.93)
and
(6.94)
The shear compliances given in the above expressions depend on the shear modulus of the material, and the
geometry of the cross section. Note that the dimensional units of the cross-sectional functions and
are 1/L. The dimensional units of the shear modulus are F/L
2
, so that the dimensional units of the shear compli-
ances are 1/F.
EXAMPLE 6.4 Shear compliances of a stiffened blade section
Determine the shear compliances for the blade section stiffened by two
stringers as shown in Fig. 6.19. The section is made of an isotropic, homo-
geneous material with a shear modulus G. Also, determine the inuence of
increasing the stringer area relative to the web area on the shear
stiffness.
Solution The second area moments for this thin-walled section are
(6.95)
Since the second area moment about the y-axis is zero in the thin wall
approximation, this section can only resist bending in the y-z plane. Hence,
we consider beam loading such that and . Let the origin of
the contour coordinate s in the web be located at the lower stringer and take the positive direction of s in the pos-
itive y-direction. The cartesian coordinates of point s are
The rst area moments of the portion of the cross section below s are

x
K
x
2
s ( )V
x
K
x
s ( )K
y
s ( )V
y
+ ( )
Gt
------------------------------------------------------------------- s d
0
S

y
K
y
s ( )K
x
s ( )V
x
K
y
2
s ( )V
y
+ ( )
Gt
------------------------------------------------------------------- s d
0
S

=
c
xx
K
x
2
s ( )
Gt
-------------- s d
0
S

= c
xy
K
x
s ( )K
y
s ( )
Gt
---------------------------- s d
0
S

=
c
yx
K
y
s ( )K
x
s ( )
Gt
---------------------------- s d
0
S

= c
yy
K
y
2
s ( )
Gt
-------------- s d
0
S

=
K
x
s ( ) K
y
s ( )
h
2
---
h
2
--- t
A
s
A
s
x
y
h t 0 >
Fig. 6.19
C
2A
s
ht
I
xx
h
3
t
12
------- 2
h
2
---
\ )
[
2
A
s
+ = I
yy
0 I
xy
0 =
V
y
0 V
x
0 =
x s ( ) 0 = y s ( )
h
2
--- s + = 0 s h
Thin-Walled Structures 185
Inuence of transverse shear deformations on bending
and
From eq. (6.87) the function , and from eq. (6.88) the function . Hence, the for-
mulas for the shear compliances given in eqs. (6.93) and (6.94) become
and
The factor Gt is factored out of the integral in the equation for compliance , since the section is homogeneous
and the thickness of the web is uniform. Substitute for the function in the expression for the compliance
to get
Performing the denite integral in this equation results in
(6.96)
where the second area moment about the x-axis through the centroid given by the rst of eqs. (6.95). Equa-
tion (6.96) is the result sought. The only non-zero transverse shear compliance is , so the material law for
shear deformation of the beam in the y-z plane is . We write the inverse of this material law as
, where is the transverse shear stiffness.
To study the inuence of the stringers on the transverse shear stiffness , we dene the ratio of the cross-
sectional area of the stringers to the total cross-sectional area A by
where . From the denition of , the portion of the cross-sectional area in the web is
For the stringers are not present ( ), and the web resists both bending normal stress and shear
stress. In the extreme case of the web thickness , and the stringers carry all the bending load with
Q
x
s ( )
h
2
--- A
s
y s ( )t s d
0
s

+
h
2
--- A
s

ht
2
-----s
t
2
---s
2
+ = =
Q
y
s ( ) x s ( )t s d
0
s

0 = =
K
x
s ( ) 0 = K
y
s ( ) Q
x
s ( ) I
xx
=
c
xx
c
xy
c
yx
0 = = =
c
yy
1
Gt
------
Q
x
s ( )
I
xx
--------------
2
s d
0
h

=
c
yy
Q
x
s ( )
c
yy
c
yy
1
Gt I
xx
2
-------------
h
2
--- A
s

ht
2
-----s
t
2
---s
2
+
2
s d
0
h

=
c
yy
1
Gt I
xx
2
-------------
h
3
A
s
2
4
------------
h
4
t A
s
12
-------------
h
5
t
2
120
--------- + +
\ )
[
=
I
xx
c
yy

y
c
yy
V
y
=
V
y
k
yy

y
= k
yy
1 c
yy
=
k
yy
2A
s
( ) A =
A ht 2A
s
+ =
ht ( ) A 1 =
0 = A
s
0 =
1 = t 0 =
Shear Flow due to Shear Forces
186 Thin-Walled Structures
the web carrying only shear force . Introducing the denition of the stringer area and web area in terms of
and the total area A into the expressions for and we get after some algebra
(6.97)
and
(6.98)
When there are no stringers ( ), the shear stiffness of the blade section is . The shear stiffness of
the section decreases, or the section becomes more exible in shear, as the stringer area increases as a larger por-
tion of the total area ( increasing from zero, but remaining less than one). A plot of the transverse shear stiff-
ness versus is shown in Fig. 6.20.
EXAMPLE 6.5 Deection of a cantilevered beam due to bending and shear deformation.
Consider a cantilever beam of length subjected to a vertical tip force as shown in Fig. 6.22. The
cross section of the beam is the same as in Example 6.5, and the cross-sectional dimensions shown in Fig. 6.19
are , , and . The moduli of the material are and
. Use Castiglianos second theorem to determine the tip portion of the tip displacement due to
shear deformation of the web.
Solution Castiglianos theorem applied to this problem is
The complementary strain energy is due to bending in the y-z plane and transverse shear in the same plane.
This complementary energy is
V
y

I
xx
k
yy
I
xx
h
2
A
12
--------- 1 2 + ( ) =
k
yy
5
6
---GA
1 ( ) 1 2 + ( )
2
1 3 3.5
2
+ +
----------------------------------------- =
0 = 5GA 6

0.2 0.4 0.6 0.8 1


a
0.2
0.4
0.6
0.8
k
yy
HGAL
Fig. 6.20 Transverse shear stiffness of the blade section as the area of the stringers increases relative
to the web area.

2A
s
A
--------- =
ht
A
----- 1 =
L 50 in. = F
h 10 in. = t 0.02 in. = A
s
0.5 in.
2
= E 10 Msi =
G 3 Msi =

U
*
F
---------- =
U
*
Thin-Walled Structures 187
Inuence of transverse shear deformations on bending
in which the second area moment is given by eq. (6.97), and the transverse shear stiffness by eq. (6.98),
in Example 6.5. Castiglianos theorem yields
Equilibrium determines the bending moment in terms of the tip force by the equation , and
the shear force in terms of the tip force by . Substitute these results into Castiglianos theorem to get
Performing the integrals in the above expression we obtain
Let denote the portion of the tip displacement due to bending, and let denote the portion due to shear. The
ratio of the displacement due to shear to the displacement due to bending is
Substitute into this expression eq. (6.97) for and eq. (6.98) for , and perform some algebra to get
For the specied data the factor is
x
y y
z
L
F ,
Fig. 6.22 A cantilever beam subjected to a tip force
U
*
M
x
2
2EI
xx
-------------- z d
0
L

V
y
2
2k
yy
---------- z d
0
L

+ =
||||||||
bending shear
I
xx
k
yy

M
x
EI
xx
----------
M
x
F
---------- z d
0
L

V
y
k
yy
-------
V
y
F
--------- z d
0
L

+ =
M
x
L z ( )F =
V
y
F =

1
EI
xx
----------
\ )
[
L z ( )F [ ] L z ( ) [ ] z d
0
L

1
k
yy
-------
\ )
[
F [ ] 1 [ ] z d
0
L

+ =

FL
3
3EI
xx
--------------
FL
k
yy
------- + =
||||
bending shear

b

s

b
------
3EI
xx
FL
3
--------------
\ )
[
FL
k
yy
------- =
I
xx
k
yy

b
------
3
10
------
E
G
----
\ )
[
h
L
---
\ )
[
2
1 3 3.5
2
+ + ( )
1 ( ) 1 2 + ( )
----------------------------------------- =

Shear Flow due to Shear Forces


188 Thin-Walled Structures

The ratio of tip displacements evaluates as
(6.99)
The percentage of the tip displacement due to shear relative to the total tip displacement is given by the expres-
sion
Hence, almost 35% of the tip displacement is due to shearing deformation of the web.
6.6 Problems
1. The thickness of each branch in the thin-walled
cross section shown below is 3mm and
. The shear force .
a)Determine the shear ow distribution and sketch
it on the cross section. Indicate on the sketch the
positive sense along the branch.
b)Estimate the shear stress due to transverse bend-
ing at point A.
c)Estimate the maximum shear stress due to trans-
verse bending.
2. Consider the thin-walled, Y-section beam of Problem 1 Section 3.6, in
which we found . Also, . Determine the mag-
nitude, direction, and sense of the shear ow at the centroid if the shear
force components are and .

2A
s
ht 2A
s
+ ( )
-------------------------
2 0.5
10 0.02 2 0.5 + ( )
-------------------------------------------------- 0.8333 = = =

b
------
3
10
------
10Msi
3Msi
---------------
\ )
[
10in.
50in.
------------
\ )
[
2
1 3 0.8333 ( ) 3.5 0.8333 ( )
2
+ +
1 0.8333 ( ) 1 2 0.8333 ( ) + ( )
------------------------------------------------------------------------
\ )
[
0.5337 = =

s

b
+
----------------- 100

s

b

s

b
1 +
------------------------ 100 34.8% = =
50 50 30 30
10
40
C
x
y
A
Note: all dimensions in mm
I
xx
10
5
mm
4
= V
y
5 kN =
l l
x
y
45 45
C
Note: the length of each branch is l
and the thickness of each branch is t
0.6414l
I
xx
2.0238l
3
t = I
yy
l
3
t =
V
x
0 = V
y
0 >
Thin-Walled Structures 189
Problems
3. Determine the location of the shear center e from point A for
the thin-walled, open section shown.
4. Determine the coordinates (e
x
,e
y
), with respect to point O, of the shear center (S.C.) for the thin-walled open
section shown below. The second area moments with respect to the centroidal axes x and y are
a
2

t
e
S.C.

2
axis of symmetry A
I
xx
0.660155 in.
4
= I
yy
0.130305 in.
4
= I
xy
0.127197 in.
4
=
e
x
e
y
O
S.C.
x
y
1.5
2
x
c
y
c
, ( ) 1.02786 0.529989 , ( ) =
x
x
y
y
O
C
t 0.030 typ. =
All dimensions in inches
Shear Flow due to Shear Forces
190 Thin-Walled Structures

Thin-Walled Structures

191

CHAPTER 7

Bars Subjected to
Torsional Loads

7.1 Uniform torsion of a circular tube

Consider uniform tube, or cylindrical shell, of length

L

subjected to a torque

T

at each end as shown in Fig. 7.1.
The inside radius of the tube is denoted by

a

and the outside radius by

b

. Equilibrium requires the internal torque
to be equal to applied torque

T

at each cross section

z

= constant, where . Using the right-hand screw
rule, the sign convention for a positive internal torque is that a positive torque acts in the positive

z

- direction on
a positive

z

-face, and by the action/reaction law a positive torque acts in the negative

z

-direction on a negative

z

-
face. The tube is assumed to made of a homogeneous, isotropic, linear elastic material.
The geometry of deformation of the tube is based on symmetry of the undeformed conguration, symmetry
of the loading, and on isotropic material behavior. Three symmetry arguments concerning the geometry of the
deformation are

1.

The pattern of deformation will not vary along the length of the tube.
Any element cut perpendicular to the

z

-axis of the tube, say of length


z

, will have the same original geome-
try and same loading. Thus, we expect each element to have the same deformation.
x
y
a
b
T
z
,
T z
y
L
T
Fig. 7.1 Circular tube subjected to uniform torsion
0 z L < <

Bars Subjected to Torsional Loads

192

Thin-Walled Structures

2.

Each element has a midplane of symmetry in the sense that the deformed geometry at each end must be the
same.
Consider the element shown in Fig. 7.2(a). If we rotate it and the torques 180 degrees about the

x

-axis we
get the element shown in Fig. 7.2(b). The elements (a) and (b) are identical in original shapes and loadings.
Thus, the deformed geometry of the left end of (b) must be the same as the left end of (a). But the deformed
geometry of the left end of (b) must be identical to the deformed geometry of the right end of (a), since dia-
gram (b) is a simple rotation of (a).

3.

All radial lines in the undeformed cross section must deform into identical curves. This is a result of

axial
symmetry

(circular cross section symmetric with respect to the

z

-axis), and isotropy. Note that tube with a
rectangular cross section would not be axially symmetric.
First, we argue that plane and normal sections in the undeformed tube remain plane and normal to the

z

-axis
in the deformed tube.
On the basis of symmetry argument (3) end surfaces of elements must dish-in or bulge-out as a surface of
revolution, or possibly remain plane as a result of the applied torque. Also symmetry argument (2) implies
both ends do the same. See Fig. 7.3. Further (1) implies each element has the same deformation. But ele-
ments which bulge-out or dish-in cannot t together in the deformed geometry to form a continuous tube.
Thus, plane sections remain plane.
Second, we argue that each cross section rotates without distortion in its plane.
Consider the element of the tube shown in Fig. 7.4(a). It is assumed a diameter deforms in the left-end plane
A
B
C
D
A
B
C
D
180
x
(a) (b)
z
z
Fig. 7.2 Symmetry about the midplane of an element of the torsion tube.
bulge-out
z
dish-in
z
Fig. 7.3 Potential axial symmetric end surface deformations.

Thin-Walled Structures

193

Uniform torsion of a circular tube

as shown. Each diameter would have the same shape by symmetry argument (3). By symmetry argument (2)
diameters on the right-end plane should have the same shape. Rotate the element in Fig. 7.4(a) 180 degrees
about the

x

-axis to get Fig. 7.4(b). The elements in Fig. 7.4(a) and Fig. 7.4(b) have the same original geom-
etry and loading, so the assumed shape of the deformed element should be the same for both. But it is not. If
the family of diameters in the deformed geometry were straight lines, there would be no conict with sym-
metry. Hence, diameters in the undeformed tube remains straight in the deformed geometry, but they may
rotate in their plane. Due to symmetry argument (3) each diameter in the cross section rotates through the
same angle.
Symmetry of deformation has not ruled out symmetrical
expansion or contraction of the circular cross section, or a
lengthening or shortening of the tube. It does not seem plau-
sible that such

dilational

deformations would be an important
part of the deformation. Thus, extensional strains of line seg-
ments in the axial direction (

z

-axis), radial direction (

r

), and
the circumferential direction ( ) are assumed zero. See Fig.
7.5. That is extensional strains, , , and
.
The only motion of a circular cross section is, then, a
rotation about the

z

-axis. Thus, the right angle between line
elements originally parallel to the

r

- and -directions does
not change, which means the shear strain . Also, for small strains the right angle between the line ele-
ments originally parallel to the

r

- and

z

-directions does not change very much, so shear strain . In cylin-
drical coordinates the only non-zero strain is . To compute this shear strain consider an element of the tube of
length


z

and radius

r

, where . According to the symmetry arguments presented above, a straight line
on the periphery of the tube parallel to the

z

-axis will deform into a helix. Line AB is a portion of such a straight
line, and when loads are applied this line moves to become AB as shown in Fig. 7.6. For very small


z

the seg-
ment AB can be considered a straight line segment forming the angle with line AC taken parallel to line
AB. This angle is the change of the right angle between originally perpendicular elements


z

and , and is the
180
x
(a) (b)
z
z
T
T
T
T
Fig. 7.4 A diameter in the end plane of the undeformed body cannot distort from a straight line in
the deformed body due to problem symmetry.
z
r
r

r
x
y
z
Fig. 7.5 Cylindrical coordinates and
differential line elements.

z
0 =
r
0 =

0 =

r
0 =

rz
0 =

z
a r b

z
r

Bars Subjected to Torsional Loads

194

Thin-Walled Structures

shear strain. From the diagram in Fig. 7.6, we have the relationship between the shear strain and twist of the tube
as
which in the limit as gives

(7.1)

This result shows that the shear strain varies in direct proportion to the radius. Since each element of the tube of
length


z

deforms in the same way as any other, we conclude



is a constant along a uniform section of
the tube. The quantity is the twist per unit length or simply the unit twist.
We use Hookes law to relate the shear stress to the shear strain, since we have assumed the material is linear
elastic and isotropic. Hookes law is

(7.2)

where

G

denotes the shear modulus of the material. For an isotropic material recall that

where the modulus of elasticity is denoted by

E

and Poissons ratio is denoted by


. Eliminating the shear strain
by eq. (7.1), we get

(7.3)

The stress distribution given by eq. (7.3) leaves the outside and inside cylindrical surfaces of the tube stress
free, as it should. Internally, each differential element of the tube is in equilibrium because does not change

z
z

r
A
A
B
B
C
x
y
z
Fig. 7.6 Reduction in the right angle between line elements originally parallel to the
axial and circumferential directions due to twist of the body.
r
z

z
z =
z 0

z
r
d
dz
------ =
d
z
dz
d
z
dz

z
G
z
=
G
E
2 1 + ( )
-------------------- =

z
Gr
d
z
dz
-------- =

z
Thin-Walled Structures 195
Uniform torsion of a circular tube
in the -direction (axial symmetry) nor does it change in the z-direction (because of uniformity of deformation
pattern along the length of the tube). The shearing stress is the same on each z- and -face. Hence the differential
element is in equilibrium. See Fig. 7.7.
On the end surface of the tube the resultant of the stress distribution must be equal to the applied torque T.
Because of rotational symmetry of the stress distribution, the resultant force due to the shear stresses must be
zero. The only resultant of the stress distribution is the torque
(7.4)
where dA is the element of area and the integral is taken over the cross-sectional area A of the tube. Substituting
eq. (7.3) into eq. (7.4) we get
(7.5)
where the torsion constant J is dened as
(7.6)
The torsion constant J is the same as the polar area moment for a circular annulus, and has dimensional units of
L
4
. From eq. (7.5) we nd the twist per unit length is
(7.7)
For tube of length L, the total angle of twist between the ends is found by integrating eq. (7.7); i.e.,
(7.8)
Now substitute eq. (7.7) into eq. (7.3) to nd
(7.9)

z
z
r

x
y
z
T

z

z
=

Fig. 7.7 Shear stresses due to torque


T r
z
( ) A d
A

=
T GJ
d
z
dz
-------- =
J r
2
A d
A

r
3
d ( ) r d
0
2

a
b


2
--- b
4
a
4
( ) = = =
d
z
dz
--------
T
GJ
------- =

z
L ( )
z
0 ( )
T
GJ
------- z d
0
L

TL
GJ
------- = =

z
Tr
J
------ = a r b
Bars Subjected to Torsional Loads
196 Thin-Walled Structures
Thin-walled approximations Let the mean radius be denoted by and the wall thickness be denoted by t. The
relationships between the inside and outside radii to the mean radius and wall thickness are
(7.10)
A thin-walled tube is quantied by ; e.g. . The torsion constant, eq. (7.6), becomes
Since the quantity , we can neglect this factor squared with respect to one in the above equation to
get
(7.11)
The variation of the shear stress, eq. (7.9), over the annular section is small for a thin wall. That is, we can
write eq. (7.9) as
which shows that the range of is small for a thin-walled section and can be neglected. Thus, the shear stress
is approximated as a uniform distribution across the wall equal to its value at the mean radius.
(7.12)
Since the shear stress is assumed to be uniform across the wall thickness, we dene the shear ow due torsion q
as . Substituting eq. (7.12) for the shear stress and eq. (7.11) for the torsion constant, the shear ow is
written as
(7.13)
where the area enclosed by the mean radius is . Equation eq. (7.13) is more general than as presented in
this section for the circular tube. It is applicable to the torsion of thin-walled sections other than the circular tube.
Equation (7.13) is called the Bredt-Batho formula. These torsion results for the thin-walled circular tube are
r
a r
t
2
--- =
b r
t
2
--- + =
or the inverse
r
1
2
--- a b + ( ) =
t b a =
0
t
r
-- 1 <
t
r
--
1
20
------
J

2
--- r
t
2
--- +


4
r
t
2
---


4

r
4
2
-------- 1
t
2r
----- +


4
1
t
2r
-----


4
= =
J
r
4
2
-------- 1 4
t
2r
-----


6
t
2r
-----


2
4
t
2r
-----


3
t
2r
-----


4
+ + + +


=
1 4
t
2r
-----


6
t
2r
-----


2
+ 4
t
2r
-----


3
t
2r
-----


4
+


J

2
---r
4
8
t
2r
-----


8
t
2r
-----


3
+ 2r
3
t 1
t
2r
-----


2
+ = =
0
t
2r
-----


1 <
J 2r
3
t = thin-walled circular tube

z
Tr
J
------
r
r
--


= 1
t
2r
-----

r
r
-- 1
t
2r
-----


+
r r

z
Tr
J
------ = thin-walled circular tube
q
z
t =
q
Trt
2r
3
t
--------------
T
2
-------- = =
r
2
=
Thin-Walled Structures 197
Uniform torsion of an open section
depicted in Fig. 7.8., in which we used the contour coordinate s rather than the angle to write the shear stress
as
7.2 Uniform torsion of an open section
Consider uniform torsion of the thin-walled rectangular section shown in Fig. 7.9. The correct solution to the
problem of uniform torsion of bars with non-circular cross sections was rst obtained by Saint-Venant (1855).
The procedure used by Saint-Venant, called the semi-inverse method, was to assume the form of the displace-
ments for the non-circular bar based on the known displacements for the circular section bar. Elasticity theory is
required to fully understand the details of the solution, but elasticity theory is beyond the scope of this text.
Hence, we will only summarize the important aspects of this solution here. The reader is referred to the texts by
Timoshenko and Goodier (1970) and Oden and Ripperger (1981) for in-depth discussions.
The deformed bar is shown in Fig. 7.10. As for the circular section under positive torque, the cross section a
z + dz rotates counter-clockwise relative to the section at z (Fig. 7.10a), but unlike the circular section bar the
cross section does not remain plane. For non-circular cross sections, the cross section displaces out of its plane as
well as rotates about the z-axis. This out-of-plane displacement is a result of the loss of axial symmetry of the
non-circular section relative to the circular section bar. (Refer to the third symmetry argument of Section 7.1.)

z

zs
=
t
r
x
y

s r =
t

zs
q t =
T
q
T
2
-------- =

Fig. 7.8 Shear ow and shear stress in a thin-walled circular tube subjected to a torque.
b
t
L
x
y
z
T
T
Fig. 7.9 Uniform torsion of a bar with a rectangular cross section.
0
t
b
--- 1 <
Bars Subjected to Torsional Loads
198 Thin-Walled Structures
The out-of-plane displacement w(x,y) is called warping of the cross section (Fig. 7.10b). For the thin-walled
rectangular section this warping displacement is approximated by
(7.14)
It is assumed in this torsion solution that the cross sections are free to warp as expressed by eq. (7.14). If the bar
cross section is not free to warp as described, then additional normal stresses arise as result of the constraint on
the warping displacement. Constrained warping complicates the state of stress, and is not considered further here
since constrained warping stresses are known to be boundary layer effects in most structural applications. Under
uniform torsion, the twist per unit length is a constant, and from the elasticity solution for the thin-walled section
it is given by
(7.15)
Comparing this result to the standard torsion formula , we nd the torsion constant is
(7.16)
Shear stress components due torsion are
zx
and
zy
, as shown in the adjacent sketch. The dom-
inate stress under free warping of the narrow rectangular section is the shear stress tangent to
the long side,
zx
, and it is approximated by
(7.17)
The shear stress component
zy
, is negligible except near the ends at , where component
zx
must vanish
because the lateral surfaces of the bar at are stress free. The shear stress distribution through the thick-
ness of the wall is linear, as is shown by eq. (7.17), with a value of zero along the contour line y = 0, and attaining
x
y
w
T
T
w x y , ( )
T
GJ
------- xy =
(b) Warping of the cross section (a) Rotation and warping of the cross sections
x
y
z
Fig. 7.10 Deformation of a thin-walled rectangular section bar under uniform torsion
w x y , ( ) xy
d
z
dz
-------- =
d
z
dz
--------
T
G
1
3
---bt
3


--------------------- =
d
z
dz
--------
T
GJ
------- =
J
1
3
---bt
3
=

zx

zy
t

zx
2G
d
z
dz
--------


y =
t
2
--- y
t
2
---
x
b
2
--- =
x
b
2
--- =
Thin-Walled Structures 199
Uniform torsion of an open section
a maximum magnitude at . Combining eqs. (7.15) and (7.17), the magnitude of the maximum shear
stress is
(7.18)
Equations (7.15) to (7.18) are thin-walled approximations. The elasticity solution provides more general for-
mulas based on the width-to-thickness ratio b/t. These relations are
(7.19)
where the dimensionless parameters k
1
and k
2
are functions of the ratio of b/t. These parameters are listed for var-
ious b/t ratios in table on the next page. From results listed in the table, eqs. (7.15) and (7.18) are seen to be good
approximations for b/t >10.
If the shear ow is calculated from the shear stress distribution given by eq. (7.17) we get
That is, the shear ow due to a linear shear stress distribution through the thickness of the wall is zero. The shear
stresses are not zero, but the shear ow is zero for an open section. This is an important difference with respect to
torsion of a closed section. Recall that the shear stress distribution is uniform across the thickness of the wall in a
closed section.
Now consider torsion of open section bars of more complex shape as are shown in Fig. 7.11. Understanding
the torsional response of these bars with complex, open cross-sectional shapes is facilitated by an analogy to the
response of an initially at membrane supported on its edges over an opening, where the edges of the opening are
in the same shape as the cross section. The membrane is stretched under a uniform tension, and then subjected to
an internal pressure to cause the membrane to deect. The deected shape of this pressurized membrane is anal-
ogous to the torsion problem in that level contours on the surface of the deected membrane coincide with the
lines of action of the of the shear stresses, and that the slope of the membrane normal to the level contour is pro-
Table of parameters for torsion of a rectangular cross section bar
a
a. Timoshenko and Goodier, 1970, p.312
Width to thickness ratio b/t
Unit twist parameter k
1
Shear stress parameter k
2
1.0 0.1406 0.208
1.2 0.166 0.219
1.5 0.196 0.231
2.0 0.229 0.246
2.5 0.249 0.258
3.0 0.263 0.267
4.0 0.281 0.282
5.0 0.291 0.291
10.0 0.312 0.312
0.333 0.333
y t 2 =

max
3T
bt
2
------- =
d
z
dz
--------
T
G k
1
bt
3
( )
---------------------- =
max
T
k
2
bt
2
------------ =
q
zx
y d
t 2
t 2

2G
d
z
dz
--------


y


y d
t 2
t 2

2G
d
z
dz
--------


y y d
t 2
t 2

0 = = = =

Bars Subjected to Torsional Loads


200 Thin-Walled Structures
portional to the magnitude to the shear stress. Also, the volume between the x-y plane and the deected surface of
the membrane is proportional to the total torque carried by the section. The following text excerpted from Oden
and Ripperger (1981, p. 46) summarizes this analogy.
This analogy was rst discovered by Ludwig Prandtl in 1903 and is knowns as Prandtls membrane anal-
ogy. Prandtl took full advantage of the analogy and devised clever experiments with membranes. By measur-
ing the volumes under membranes formed by a soap lm subjected to a known pressure, he was able to
evaluate torsional constants. By obtaining the contour lines of the membranes he determined stress distribu-
tions.
Torsional constants and the maximum shearing stress can be found for complex cross sections by using the
results for the thin-walled rectangular section. The membrane analogy shows that the torsional load carrying
capacity of the complex open section is nearly the same as the narrow rectangular section, because the volumes
under the membranes are nearly the same if we neglect the small error introduced at the corners or junctions. In
this way, the membrane analogy implies that the complex open cross section has about the same torsional load
carrying capacity as a thin-walled rectangular section with a length equal to the total arc length of the contour of
the complex section.
Since each branch of the open section is equivalent to a narrow rectangular section with the same developed
length and thickness, we can sum the torques carried by each branch in the following way
where the torsion constant for the entire cross section is
(7.20)
Note that the twist per unit length is the same for all branches in the open section, because the cross section is
assumed to be rigid in its own plane. The use of eq. (7.20) for several open sections is shown in Fig. 7.11. Start-
ing from eq. (7.18), the maximum shear stress in the i
th
branch of the section is given by
b
1
t
1
t
2
b
2
b
2
b
1
t
1
b
2
t
2
b
t
J
1
3
--- b
1
t
1
3
4b
2
t
2
3
+ ( ) = J
1
3
---bt
3
= J
1
3
--- b
1
t
1
3
2b
2
t
2
3
+ ( ) =
Fig. 7.11 Some thin-walled open sections and their torsion constants
T T
i
branches

GJ
i
d
z
dz
--------

GJ
d
z
dz
--------


= = =
J J
i
branches

1
3
---b
i
t
i
3
branches

= =
Thin-Walled Structures 201
Uniform torsion of an open section
(7.21)
That is, the maximum shear stress in the i
th
branch of the open section is the total torque divided by the torsion
constant for the entire section times the thickness of the i
th
branch. Note that the largest shear stress magnitude in
a built-up open section occurs in the thickest branch.
EXAMPLE 7.1 Torsional response of a thin-walled open section and an equivalent closed section
A thin-wall circular tube with contour radius r and wall thickness t is subjected to a torque T. The wall of a
second identical tube is cut parallel to its longitudinal axis along its entire length to make the cross section of this
second tube an open circular arc. See Fig. 7.12. Assume the saw kerf is very small. Compare the unit twist and
maximum shear stress in the closed section to the open section.
Solution For the closed section the shear ow from eq. (7.13), and the torsion constant
from eq. (7.11). Hence the maximum shear stress and unit twist are
For the open section the developed length b of the contour is essentially , since the saw kerf is assumed
to be very small. By the membrane analogy the torsional response is the same as the thin-walled rectangular sec-
tion of length b and thickness t. The maximum shear stress is given by eq. (7.18) and the torsion constant is given
by eq. (7.16). For , we have

Forming the ratio of the maximum shear stress of the open section to the closed section we nd

max
( )
i
3T
i
b
i
t
i
2
---------
3GJ
i
b
i
t
i
2
------------
d
z
dz
--------
3G
b
i
t
i
2
---------
1
3
---b
i
t
i
3


T
GJ
-------


Tt
i
J
------- = = = =
r
t
r
t
T T
small slit
Fig. 7.12 Closed and open thin-walled circular sections
q
T
2 r
2
( )
---------------- = J 2r
3
t =

closed
T
2r
2
t
-------------- =
d
z
dz
--------


closed
T
GJ
-------
T
G 2r
3
t ( )
----------------------- = =
2r
b 2r =

open
3T
2rt
2
-------------- =
d
z
dz
--------


open
T
G
1
3
---2rt
3


--------------------------- =

open

closed
---------------
3T
2rt
2
--------------
2r
2
t
T
-------------- 3
r
t
-- 1 = =
Bars Subjected to Torsional Loads
202 Thin-Walled Structures
Likewise, the ratio of the unit twists are
Since the ratio of the radius to thickness is greater than ten for a thin-walled section, the above results show that
the shear stress and unit twist of the open circular section are much larger than for the closed section if both sec-
tions are subjected to the same torque.
Hence, if a bar is to resist torsional loading, a closed section is preferable to an equivalent open section bar.
That is, the unit twist is smaller for the closed section bar (it is stiffer), and the maximum shear stress is smaller,
than for the equivalent open section bar subjected to the same torque.
7.3 Non-uniform torsion; governing boundary value problem
In the last two sections the cross section of the bar and the torque were considered to be uniform along the length
of the bar. Assume that an external distributed torque of intensity t
z
(units of F-L/L, or F) acts on the bar. This
intensity is a mathematical function of z, so we denote it as t
z
(z). In this case the internal torque in the bar is not
uniform along its length, but is a function of the coordinate z; i.e., T(z). First we determine the differential equa-
tion of torsional equilibrium for this case, where a free body diagram of a differential element of the bar is shown
in Fig. 7.13. Sum the torques acting on the differential element in the gure to get
Divide the last equation by dz, take the limit as , and note that in the limit, to get
(7.22)
Equation (7.22) is the differential equation of torsional equilibrium of the bar, and its is applicable in any
interval along the length in which the distributed torque intensity is a continuous function. If an external point
torque acts on the bar, then eq. (7.22) is not applicable at the point of application of this external point torque.
The internal torque T exhibits a jump in value at the point of application of the external torque, and torsional
equilibrium applied to a free body diagram of this point determines the relationship between the value of the
internal torque to the left of this point to the value of the internal torque to the right of this point. The analysis of
d
z
dz ( )
open
d
z
dz ( )
closed
-----------------------------------
3T
G2rt
3
------------------
G2r
3
t
T
------------------ 3
r
t
--


2
1 = =
z z dz +
T T dT +
t
z
z

( )dz z z
*
z dz + < < ,
t
z
z ( )
z
z 0 =
z L =
Fig. 7.13 Differential element of a bar subjected to a distributed torque
T dT T t
z
z
*
( )dz + + 0 = or dT t
z
z
*
( )dz + 0 =
dz 0 z
*
z
dT
dz
------- t
z
+ 0 = 0 z L < <
Thin-Walled Structures 203
Non-uniform torsion; governing boundary value problem
a bar subjected to a point torque is discussed in Example 7.3. The boundary conditions at the ends of the bar are
that we either prescribe the torque or the angle of twist, but not both. That is,
(7.23)
For a linear elastic material, the constitutive equation of the bar is
(7.24)
where GJ is the torsional stiffness of the bar. For the thin-walled circular annulus, the torsional constant J is
given by eq. (7.11), and for the thin-walled open sections the torsion constant is given by eq. (7.20). In subse-
quent sections we will obtain the torsion constant J for more complex cross-sectional shapes.
Substitute eq. (7.24) for the torque in the differential equation of equilibrium, eq. (7.22), to get
If the torsional stiffness is uniform along the length, then this equation reduces to
(7.25)
This is a second order, linear differential equation for the angle of twist
z
and requires two boundary conditions
to determine its solution. These boundary conditions are specied in eq. (7.23), in which the torque is related to
the derivative of the twist angle by eq. (7.24).
EXAMPLE 7.2 A uniform distributed torque acting on a bar with xed ends.
Consider a uniform bar subjected to a uniform distributed torque of intensity , .
Each end of the bar is xed to a rigid support which prevents the twisting rotation of the bar so that
and . Determine distributions of the angle of twist
z
(z) and the internal torque T(z) along the length
of the bar.
Solution The governing differential equation, eq. (7.25), reduces to
divide this equation by the torsional stiffness and integrate twice with respect to z to get
where c
1
and c
2
are constants of integration. These constants are determined from the boundary conditions. At
the left end we have
at z 0 = and z L = prescribe either T or
z
but not both.
T GJ
z d
d
z
=
z d
d
GJ
z d
d
z


t
z
z ( ) = 0 z L < <
GJ
z
2
2
d
d
z
t
z
z =
z

z
z ( ) = 0 z L < <
t
z
z ( ) t
z0
constant = = 0 z L < <

z
0 ( ) 0 =

z
L ( ) 0 =
GJ
z
2
2
d
d
z
t
z0
= 0 z L < <

z
t
z0
GJ
-------
z
2
2
---- c
1
z c
2
+ + =

z
t
z0
GJ
-------
0
2
2
----- c
1
0 c
2
+ + 0 = =
Bars Subjected to Torsional Loads
204 Thin-Walled Structures
So c
2
= 0. Using this fact in the boundary condition at the
right end, we get
So . Hence, the solution for the angle of twist is
(7.26)
The torque is obtained by substituting this result, eq.
(7.26), for the twist angle in the material law, eq. (7.24), to
get
(7.27)
The distributions of the angle of twist, eq. (7.26), and
torque, eq. (7.27), are plotted in Fig. 7.14.
EXAMPLE 7.3 A point torque acting on a bar with xed ends.
Consider a uniform bar of length L subjected to an external point torque,
denoted by Q, at z = a. Each end of the bar is xed to a rigid support so that
and . Determine the angle of twist
z
and torque T in
the bar for 0 z L. See Fig. 7.15.
Solution In this case the distributed torque intensity t
z
in the governing dif-
ferential equation, eq. (7.25), is zero in the open intervals 0 < z < a and a < z
< L. The governing differential equation is not valid at the point of applica-
tion of the point torque Q. We will have to derive the torsional equilibrium
condition separately for the point z = a. The differential equation of equilib-
rium reduces to
The general solution to this equation is
where c
1
, c
2
, c
3
, and c
4
are constants to be determined from the boundary conditions. The xed end condition at
z = 0 yields c
2
= 0, and the xed end condition at z = L yields c
4
= - c
3
L. The constants c
1
and c
3
are determined
0.2 0.4 0.6 0.8 1
zL
-1
-0.5
0.5
1
TT
max
0.2 0.4 0.6 0.8 1
zL
0.2
0.4
0.6
0.8
1
q
z
q
zmax
Fig. 7.14 Distributions of the twist angle
and torque for the bar with xed ends
and subjected to a uniform distributed
torque.

zmax
t
z0
L
2
8GJ
------------ =
T
max
t
z0
L
2
---------- =

z
t
z0
GJ
-------
L
2
2
----- c
1
L + 0 = =
c
1
t
z0
L
2GJ
----------- =

z
t
z0
2GJ
----------- Lz z
2
( ) = 0 z L
T
t
zo
2
------ L 2z ( ) = 0 z L
Q
z
a b
L
Fig. 7.15 Bar with xed
ends under a point torque

z
0 ( ) 0 =
z
L ( ) 0 =
GJ
z
2
2
d
d
z
0 = for 0 z a < < and a z L < <

z
c
1
z c
2
+
c
3
z c
4
+

=
0 z a <
a z L <
Thin-Walled Structures 205
Virtual work and strain energy
from the conditions at z = a. First, the angle of twist at z = a must be continuous for otherwise the bar would be
fractured; i.e., . Imposing this condition of continuity we have
since b = L - a. Second, as shown in the accompanying free body diagram, torsional
equilibrium at point z = a yields
Using the material law, eq. (7.24), this equilibrium equation yields the second condi-
tion for constants c
1
and c
3
, which is
Now we have two linear, independent equations to deter-
mine c
1
and c
3
. We nd and .
Thus, the solution for the angle of twist is
(7.28)
The torque is determined from this result and the mate-
rial law, eq. (7.24). We nd
(7.29)
Plots of the twist angle and the torque are shown in Fig.
7.16.
7.4 Virtual work and strain energy
Consider a bar in equilibrium under a prescribed external torque Q
1
at z = 0, a prescribed external torque Q
2
at z
= L, and a prescribed distributed torque of intensity as shown in Fig. 7.17. From this equilibrium state con-
sider an innitesimal, virtual rotation of the bar in twist denoted by the function . The external virtual
work is
(7.30)
The free body diagrams (FBDs) at each end of the bar shown Fig. 7.17 are used to determine the relation of the

z
a

( )
z
a
+
( ) =
c
1
a c
3
L a ( ) c
3
b = =
T a

( ) T a
+
( )
Q
a
T a
+
( ) T a

( ) Q + 0 =
GJ c
3
( ) GJ c
1
( ) Q + 0 = or c
1
c
3

Q
GJ
------- =
z
z
a
a
L
L
0
0

z
T
Qab
GJL
-----------
Q
b
L
---
Q
a
L
---
Fig. 7.16 Distributions of the angle of
twist and torque in Example 7.3.
c
1
Q
GJ
-------
b
L
--- = c
3
Q
GJ
-------
a
L
--- =

z
Q
GJ
-------
b
L
---z
Q
GJ
-------
a
L
--- L z ( )

=
0 z a
a z L
T
Q
b
L
---
Q
a
L
---

=
0 z a <
a z L <
t
z
z ( )

z
z ( )
W
ext
Q
2

z
L ( ) Q
1

z
0 ( ) t
z
z ( )
z
z ( ) z d
0
L

+ + =
Bars Subjected to Torsional Loads
206 Thin-Walled Structures
external torques to the internal torque in the bar. Thus, torsional equilibrium at each end of the bar requires that
the internal torque T(z) satisfy and . Hence, the external virtual work becomes
This equation can be rewritten in an equivalent form as
Now distribute the derivative of the product in the in the rst term to get
The bar is in equilibrium prior to the consideration of the virtual rotation in twist, so the internal torque and exter-
nal distributed torque intensity satisfy equilibrium equation (7.22). Hence, the last equation reduces to
where we interchanged the variational operator and the derivative as discussed in Chapter 2 with eq. (2.34) on
page 32 as the pertinent result. Since the integrand of this last equation contains variables dened internal to the
bar, we dene it as the internal virtual work. That is,
(7.31)
What the manipulations from eq. (7.30) to (7.31) show is that for a bar in equilibrium, the external virtual
work is equal to the internal virtual work for a kinematically admissible variation in the angle of twist function
. This proves the necessary condition that if the body is in equilibrium then the external virtual work equals
the internal virtual work for a kinematically admissible variation in the angle of twist. In problem solving we
assume that equating the external to the internal virtual work for every kinematically admissible twist function is
z
L
Q
2
t
z
z ( )
Q
1
Q
1
Q
2
T 0 ( ) T L ( )
z 0 = z L =
Fig. 7.17 External torques acting on a bar and the FBDs for the end conditions
FBD at each end of the bar
T L ( ) Q
2
= T 0 ( ) Q
1
=
W
ext
T L ( )
z
L ( ) T 0 ( )
z
0 ( ) t
z
z ( )
z
z ( ) z d
0
L

+ =
W
ext
z d
d
T
z
( ) z d
0
L

t
z
z ( )
z
z ( ) z d
0
L

+ =
W
ext
z d
dT
t
z
+

z
T
z d
d

z
+ z d
0
L

=
W
ext
T
z d
d
z


0
L

dz =
W
int
T
z d
d
z


0
L

dz

z
z ( )
Thin-Walled Structures 207
Virtual work and strain energy
sufcient for equilibrium of the body. See Section 2.6.2 on page 32 for the general statement of the principle of
virtual work.
The principle of virtual work states that the bar is in torsional equilibrium if
(7.32)
where the external virtual work is given by eq. (7.30) and the internal virtual work is given by eq. (7.31) consid-
ering torsion only. The mathematical conditions of kinematic admissibility for the virtual angle of twist function
are that it is continuous, its rst derivative is piecewise continuous, and that it vanishes at points where
the angle of twist is prescribed. That is, the virtual twist function, or variation in the twist function, must be a
possible twist rotation of the bar.
7.4.1 Strain energy density
Lets denote the integrand of the internal virtual work expression in eq. (7.31) as the incremental quantity .
That is,
(7.33)
For a linear elastic material, we can eliminate the torque in this equation using the material law in eq. (7.24) to
get
(7.34)
The form of this incremental quantity suggests there is a functional U
0
of the derivative of the angle of twist such
that its variation is given by eq. (7.34). We denote this functional by , where the prime means ordinary
derivative with respect to z. The variation of a functional is presented in the discussion leading to eq. (2.46) on
page 37. From the latter result, we have that the variation of U
0
is given by
(7.35)
Comparing eqs. (7.34) and (7.35) we get
(7.36)
Regarding as a simple variable, we can integrate this last equation, neglect the constant of integration since it
is immaterial here, to get
(7.37)
Equation (7.37) is the strain energy per unit length of the bar, or the strain energy density. The total strain energy
stored in the bar due to elastic deformation in twist is the integral of the strain energy density over the length of
the bar; i.e.,
W
ext
W
int
= for every kinematically admissible
z
z ( )

z
z ( )
U
0
U
0
T
z d
d
z


=
U
0
GJ
z d
d
z

z d
d
z


=
U
0

z
[ ]
U
0

z

U
0

z
=

z

U
0
GJ
z
=

U
0
1
2
---GJ
z
( )
2
=
Bars Subjected to Torsional Loads
208 Thin-Walled Structures
(7.38)
Notice that the partial derivative of the strain energy density, eq. (7.37), with respect to the derivative of the angle
of twist gives the torque; i.e.,
(7.39)
The fact that is an important property of the strain energy density. In fact, an elastic material can
be dened as one for which a strain energy density function exists.
7.5 Complementary virtual work and energy
Consider the situation in which the angle of twist at each end of the bar is prescribed and the external distributed
torque function is also prescribed. The angle of twist
z
(z) of the bar is a continuous function and its derivative is
at least piecewise continuous; that is, the angle of twist function is compatible. The torque Q
1
at z = 0 and torque
Q
2
at end z = L are unknown reactions to the prescribed twist rotations at each end of the bar. Now consider a
variation in these end torques, or virtual torques, Q
1
and Q
2
. The virtual external torques cause a virtual inter-
nal torque T(z) in the bar. We assume the virtual torque system satises equilibrium; i.e.,
(7.40)
(7.41)
The virtual system of torques satisfying conditions of equilibrium, eqs. (7.40) and (7.41), is
said to be statically admissible. In this case, eqs. (7.40) and (7.41) are satised by
,
so that either Q
1
or Q
2
are independent virtual torques, but not both, since Q
1
+ Q
2
= 0 for overall equilib-
rium of the bar. The external complementary virtual work is dened by
(7.42)
Using the boundary conditions, eq. (7.41), for the internal torque this last equation can be written as
which is equivalent to
Distribute the differential of the product of two functions in the integrand in this equation to get
U U
0
z d
0
L

1
2
--- GJ
z
( )
2
z d
0
L

= =
T

z

U
0
GJ
z
= =
T U
0

z
=
z d
d
T ( ) 0 = 0 z L < <
T 0 ( ) Q
1
= T L ( ) Q
2
=
T z ( ) Q
1
Q
2
, , ( )
T z ( ) constant Q
1
Q
2
= = =
W
ext
*
Q
2

z
L ( ) Q
2

z
0 ( ) + =
W
ext
*
T L ( )
z
L ( ) T 0 ( )
z
0 ( ) =
W
ext
*
z d
d
T
z
( ) z d
0
L

=
Thin-Walled Structures 209
Complementary virtual work and energy
From the differential equation of equilibrium for the virtual force system, eq. (7.40), the rst term in the inte-
grand of this last equation vanishes. Hence, we get
Since this last expression involves only quantities dened internal to the bar, we dene the right-hand side as the
internal complementary virtual work; i.e.,
(7.43)
We have shown in the process of going from eq. (7.40) to (7.43) that, for a compatible angle of twist func-
tion, the external complementary virtual work is equal to the internal complementary virtual work for a statically
admissible variation in the torques. That is, we have proved the necessary condition that if the angle of twist
function is compatible, then the external complementary virtual work equals the internal complementary virtual
work for a statically admissible variation in the torques. In problem solving we assume that equating external
complementary virtual to the internal complementary virtual work for every statically admissible system of vir-
tual torques is sufcient for the angle of twist function of the body to be compatible. This is a statement of the
principle of complementary virtual work for the torsion bar. A general statement of the principle of complemen-
tary virtual work (PCVW) is given on page 47. For the bar in torsion, the PCVW means the angle of twist func-
tion is compatible if
(7.44)
where the external complementary virtual work is given by eq. (7.42) and the internal complementary virtual
work is given by eq. (7.43). The principle of complementary virtual work is independent of the material behavior.
7.5.1 Complementary strain energy
Now assume the material of the bar is elastic and linear. Solve the material law given by eq. (7.24) for the deriv-
ative of the angle of twist to get , and then substitute this result into the internal complementary
virtual work of eq. (7.43) to get
(7.45)
Let the integrand of this equation be denoted by the incremental quantity . So
(7.46)
This form of the incremental quantity suggests there is a functional of the torque whose variation is given by eq.
(7.46). This functional is called the complementary strain energy density, and is denoted by . Based on
W
ext
*
z d
d
T ( )
z
T
z d
d
z
+ z d
0
L

=
W
ext
*
T
z d
d
z


z d
0
L

=
W
int
*
T
z d
d
z


z d
0
L

z
z ( )
W
ext
*
W
int
*
= for every statically admissible virtual torque T z ( )

z
T GJ ( ) =
W
int
*
T
GJ
-------


T z d
0
L

=
U
0
*
U
0
*
T
GJ
-------


T =
U
0
*
T z ( ) [ ]
Bars Subjected to Torsional Loads
210 Thin-Walled Structures
the denition of the variation of a functional given by eq. (2.45) and eq. (2.46) on page 37, we would nd in a
similar manner that the variation of the complementary strain energy density is given by
(7.47)
Equating eq. (7.46) and eq. (7.47) for every statically admissible virtual torque gives
(7.48)
Integrating this expression with respect to T (note that we treat the torque as a simple variable) and taking the
constant of integration to be zero when T = 0, we get the complementary strain energy density as
(7.49)
The important property of the complementary strain energy density is
(7.50)
which can be seen by differentiating eq. (7.49) and using (7.24). A comparison of eq. (7.50) with eq. (7.39) for
the strain energy density shows the dual attributes that these energies possess: the derivative of the complemen-
tary strain energy density with respect to the torque gives the twist rate, and the derivative of the strain energy
density with respect to the twist rate gives the torque.
Return to the internal complementary virtual work given by eq. (7.45), and use eq. (7.46) to write it as
Now interchange the variational operator and the integral operator in accordance with eq. (2.34) on page 32, and
write this equation as
where the complementary strain energy for the bar is dened as
(7.51)
7.6 Unit twist of a single cell beam due to shear ow
Consider a single cell beam with a contour of arbitrary shape and a wall thickness , where s denotes the con-
tour coordinate, as is shown in Fig. 7.18. The beam is made of an isotropic and homogeneous material that fol-
lows Hookes Law. Our purpose is to determine the twist per unit length in terms of the shear ow
distribution around the contour. In the derivation presented in this section it must be emphasized that the shear
ow can be cause by the transverse shear forces , and the torque T. The relationship of the shear ows
to the shear forces was discussed in Section 6.2 (eq. (6.13) on page 160 is the pertinent result), and for the special
U
0
*
T

U
0
*
( )T =
T

U
0
*
( )
T
GJ
------- =
U
0
*
1
2
---
T
2
GJ
------- =

U
0
*
( ) =
W
int
*
U
0
*
z d
0
L

=
W
int
*
U
*
=
U
*
U
0
*
z d
0
L

1
2
---
T
2
GJ
-------


z d
0
L

= =
t s ( )
d
z
dz
V
x
and V
y
Thin-Walled Structures 211
Unit twist of a single cell beam due to shear ow
case of a circular contour the shear ow was related to the torque via Bredts formula, eq. (7.13). In Section 7.7
we will determine how to compute the shear ow due to torsion for a closed contour of arbitrary shape. All we
need to keep in mind in this section is that the shear ow is the superposition of the shear ows due to transverse
shear and torsion. To relate the shear ow to the unit twist of the cross section, we use the material law and the
geometry of the deformation.
7.6.1 Hookes law
The shear stress is directly related to the shear strain with the constant of proportionality being the shear modulus
G of the material.
(7.52)
Multiply this equation by the wall thickness t to get
(7.53)
The deformation due to the shear ow is depicted in Fig. 7.19.

z
T,
x
y
z
t(s)
V
x
V
y
q s ( )
Fig. 7.18 Thin-walled, single cell cross section with an arbitrary shaped contour
s

zs
G
zs
=
q Gt
zs
=
x
y
z
t(s)
A
B
C
A
*

2
---
zs

B
*
C
*
q
q
s v
t
,
z w ,
Fig. 7.19 Element of the wall deformed in shear.
Bars Subjected to Torsional Loads
212 Thin-Walled Structures
7.6.2 Shear strain-displacement relation
The shear strain
zs
is the reduction in the right angle between line elements originally parallel to the s- and z-
directions in the undeformed body as is shown in Fig. 7.19. The shear strain is related to the partial derivatives of
the displacement components of a material point on the contour. Let the v
t
(s,z) denote the displacement tangent
to the contour, or the s-direction, of the material point at (s,z), and let w(s,z) be the displacement of this point in
the axial direction, or z-direction. The three adjacent points labeled A, B, and C in the undeformed shell displace
to position A*, B*, and C* in the deformed shell as shown in Fig. 7.19 and Fig. 7.20. The displacements of these
adjacent material points are shown in Fig. 7.20. The reduction in the right angle between lines elements origi-
nally parallel to the s- and z-directions is determined from Fig. 7.20 as
(7.54)
Since the displacement derivatives are small in magnitude with respect to unity for innitesimal deformations,
we get
(7.55)
Combining eqs. (7.53) and (7.55) to eliminate the shear strain gives
(7.56)
7.6.3 Tangential displacement of a typical point on the contour
Take the origin of the contour coordinate (s = 0) at some point along the contour with s increasing counterclock-
wise. As s transverses the contour and returns to the origin its value is S, where S is the length of the perimeter of
the contour. See Fig. 7.21. A generic point on the contour is labeled A and has cartesian coordinates
w(s,z)
w(s,z+dz)
w(s+ds,z)
v
t
(s,z)
v
t
(s+ds,z)
v
t
(s,z+dz)
A:(s,z)
B
dz
ds
C
A*
B*
C*
s
w
ds
1
s
v
t
+


ds
z
v
t
dz
1
z
w
+


dz
A*
B*
C*
Fig. 7.20 Tangential and axial displacements of three adjacent points in the wall of the shell.

zs
z
v
t
dz
1
z
w
+


dz
---------------------------
s
w
ds
1
s
v
t
+


ds
---------------------------- + =

zs
z
v
t
s
w
+ =
q
Gt
------
s
w
z
v
t
+ =
Thin-Walled Structures 213
Unit twist of a single cell beam due to shear ow
. The directions tangent and normal to the contour at A are denoted by t and n, respectively, as is
shown in Fig. 7.21, and the angle between the positive x-direction and the tangent to the contour is denoted by
. For a differential length ds along the contour at A we have the trigonometric relations
(7.57)
The projection of the cross section in the deformed cylindrical shell onto the x-y plane is assumed not to dis-
tort from its original cross-sectional shape in the undeformed shell. However, the projected cross section can dis-
place in the x-y plane and rotate about the z-direction relative to the cross section in the undeformed shell. The
cross section can warp out of its plane in a manner similar to the warping of the rectangular cross section under
torsion discussed in Section 7.2. This warping displacement of the non-circular contour is in contrast to the circu-
lar contour (Section 7.1) in which plane cross sections remained plane in the deformed shell because of axial
symmetry of the circular section. Again, the loss of axial symmetry results in warping of the contour.
The assumption that the projection of the contour into the x-y plane is in the shape of the undeformed con-
tour results in a kinematic relationship between the tangential displacement of the point A on the contour and the
displacements and rotation of the section. First, we reference the displacement and rotation of the cross section to
an arbitrary point labeled O as shown in Fig. 7.22. The tangential displacement v
t
of point A consists of a rigid
body translation v
t1
, in which point O moves to O* and A moves to A
1
, followed by a rotation about O*. The
tangential displacement of A
1
to A* is denoted by v
t2
. Thus, .
The displacement of point O has a horizontal component denoted by u
0
(z) and a vertical component denoted
by v
0
(z). Since the angle between the x-direction and the tangent to the contour at s is denoted by the function
, the projection of the horizontal component in the tangent direction is , and the projection of the
vertical component in the tangent direction is . Also, the displacement of A to A
1
is the same as the dis-
placement of O to O*. Thus, the tangential displacement of A to A
1
due to translation of the section is the sum of
these projections, or
x s ( ) y s ( ) , [ ]
s ( )
cos
s d
dx
= sin
s d
dy
=
A
t
n
x
y
s 0 =
s S =
s
t(s)
s ( )
ds
dy
dx


t
n
x s ( )
y s ( )
Fig. 7.21 Geometry of the contour.

z
v
t
v
t 1
v
t 2
+ =
s ( ) u
0
cos
v
0
sin
Bars Subjected to Torsional Loads
214 Thin-Walled Structures
(7.58)
The tangential displacement due to the rotation is determined by the projection of line O*A* onto the tangent
direction. First note that the length of lines , , and are the same because of the rigid section
assumption. The angle between the line and the normal direction to the contour at point A is denoted by .
The tangential component of the displacement of A
1
to A* is obtained from the geometry shown in Fig. 7.22 as
We assume that the rotation is small such that and . Also, note that ,
which is the coordinate of A relative to O in the direction normal to the contour at A. Thus,
The total tangential displacement of A to A* is then
(7.59)
7.6.4 Relation between the shear ow and unit twist
We can take the partial derivative of the tangential displacement, eq. (7.59), with respect to z to get
O
A
O*
A
1
A*
u
o
v
0

r
r
x
y

v
n

z
+

z
v
t 1
v
t
v
t 2
Fig. 7.22 The tangential displacement of a generic point on the contour is related to the rigid body
displacement and rotation of the section
v
t 1
u
0
cos v
0
sin + =

z
OA O
*
A
1
O
*
A
*
OA
v
t 2
O
*
A
*
sin(
z
) O
*
A
1
sin + O
*
A*
z
cos sin
z
sin cos sin + [ ] = =

z

z
sin
z

z
cos 1 OA cos r =
v
t 2
O
*
A
*
cos ( )
z
OA cos ( )
z
r
z
= = =
v
t
s z , ( ) u
0
z ( ) s ( ) [ ] cos v
0
z ( ) s ( ) [ ] sin + r s ( )
z
z ( ) + =

translation rotation
Thin-Walled Structures 215
Uniform torsion of a thin-walled closed section with a contour of arbitrary shape
(7.60)
Substitute this result for the derivative of the tangential displacement in eq. (7.56) to get
Now integrate this last expression completely around the contour to get
(7.61)
From eqs. (7.57) we have
and
so that the second and third terms on the right-hand side of eq. (7.61) vanish. The rst integral on the right-hand
side of eq. (7.61) can be done exactly to give
,
since the axial displacement of the material point at s = 0 and s = S is unique. The fourth integral on the right-
hand side of eq. (7.61) (integral of the quantity rds around the contour) is twice the enclosed area of the contour.
Hence, the nal result from eq. (7.61) is
(7.62)
Equation (7.62) is the formula we set out to derive. It relates the shear ow to the twist per unit length of the bar.
The shear ow can be caused by both transverse shear forces and torque. In this sense, eq. (7.62) is more general
than for torsion only.
7.7 Uniform torsion of a thin-walled closed section with a contour of
arbitrary shape
Consider again the thin-walled cylindrical shell with an arbitrary shaped contour subjected to only a torque T and
no transverse shear, as is shown in Fig. 7.23. The cross section and torque are constant along the length of the
shell, and the material is linear elastic and isotropic. However, we permit the shear modulus G of the material to
be a function of the contour coordinate s to model multi-material bars. As in the Section 7.1 and Section 7.2, this
is a uniform torsion problem. Under uniform torsion the twist per unit length is constant, since the
torque, the cross section, and material do not change with respect to z. Our purpose is to relate the shear ow q to
the torque T.
z
v
t
z d
du
0
cos
z d
dv
0
sin r
z d
d
z
+ + =
q
Gt
------
s
w
z d
du
0
cos
z d
dv
0
sin r
z d
d
z
+ + + =
q
Gt
------ds

s
w
ds

z d
du
0
cos ds

z d
dv
0
sin ds

rds

( )
z d
d
z
+ + + =
cos ds

dx

x S ( ) x 0 ( ) 0 = = =
sin ds

dy

y S ( ) y 0 ( ) 0 = = =
s
w
ds

w S z ) , ( ) w 0 z , ( ) [ ] 0 = =
z d
d
z 1
2
-----------
q
Gt
------ds

=
d
z
dz
Bars Subjected to Torsional Loads
216 Thin-Walled Structures
Axial equilibrium requires uniform shear ow around the contour We assume the shear stress is uni-
formly distributed across the thickness of the wall, based on the analysis of a thin-walled circular tube. The shear
ow is given by , and the shear ow is tangent to the contour. Since there are no axial normal
stresses (no bending nor extension), the shear ow is uniform along the contour. That is,
(7.63)
A free body element from the wall is shown in Fig. 7.24, and axial equilibrium is used to establish the shear ow
is uniform along the contour.
Resultant of the uniform shear ow In general, the resultant of the shear ow resolved at an arbitrary point in
the cross section (again, labeled O) is a force and a couple. Static equivalence gives
(7.64)
in which r(s) is the coordinate in the normal direction to the contour at A measured from point O. See Fig. 7.25.
Since the shear ow is independent of the contour coordinate s, it may be brought outside the integral in eqs.
(7.64). We get, using eqs. (7.57) in the process, that
(7.65)

z
T,
T
x
y
z
t(s)
Fig. 7.23 Uniform torsion of a thin-walled cylindrical shell with an arbitrary shaped contour.

zs
q
zs
t s ( ) =
q constant with respect to s =
s
s + ds
q(s + ds) dz
q(s) dz
z
s
q s ds + ( )dz q s ( )dz 0 =
Fig. 7.24 A free body diagram of an element of the shell for axial equilibrium.
F
x
cos ( )qds

= F
y
sin ( )qds

= T r qds ( )

=
F
x
q ds cos

q dx

q x S ( ) x 0 ( ) [ ] 0 = = = =
Thin-Walled Structures 217
Uniform torsion of a thin-walled closed section with a contour of arbitrary shape
(7.66)
(7.67)
The last equation is in eqs. (7.67) is called Bredts formula, or the Bredt-Batho formula, and it relates the torque
to the shear ow via the area enclosed by the contour. Since Bredts formula is the principal result, we repeat it
below as eq. (7.68).
(7.68)
For the shear ow independent of the contour coordinate, we nd that the unit twist from eq. (7.62) reduces
to
(7.69)
Equations (7.68) and (7.69) can be combined by eliminating the shear ow, and we write the result as
(7.70)
in which the torsion constant is given by
(7.71)
and the modulus-weighted thickness is given by
(7.72)
Shear modulus G
0
is introduced in the denition of torsional stiffness as a reference shear modulus. It is selected
for convenience in cross sections where the material can vary from branch to branch. If the cross section is com-
qds
O
A

d
1
2
---rds =
r
ds
O
F
x
F
y
T
Fig. 7.25 Resultant of the shear ow

ds
dx
dy
F
Y
q ds sin

q dy

q y S ( ) y 0 ( ) [ ] 0 = = = =
T q rds

q 2d

2q = = =
q
T
2
-------- = Bredt's formula
z d
d
z q
2
---------
ds
Gt
------

=
T G
0
J
z d
d
z
=
J
4
2
ds
t*
-----

---------- = single cell section


t*
G s ( )
G
0
-----------t =
Bars Subjected to Torsional Loads
218 Thin-Walled Structures
posed of a single homogeneous material, then we take G
0
= G, so that t* is the actual wall thickness t. It is impor-
tant to remember that the formula for the torsion constant J above is only valid for single-cell, closed section.
The torsion constants for circular section and rectangular section made of a single homogeneous material
and uniform wall thickness are shown in Fig. 7.26. Note that the general formula for J, eq. (7.71), reduces to
value of J for the circular tube given in Section 7.1, eq. (7.11).
EXAMPLE 7.4 Torsion of a circular, bi-material section
Consider the circular section made of two materials as shown in Fig. 7.27. Determine the shear ow, maximum
shear stress, the torsion constant, and the torsional stiffness.
Solution The shear ow is computed from Bredts formula, eq. (7.68), as
The shear stress in the left side wall is , and in the right side
wall the shear stress is . Hence, the maximum shear stress is
12.7 ksi, and it occurs in the 0.125-inch-thick section.
r
t
a
t
r
2
=
ds
t
-----

2r
t
--------- =
J
4 r
2
( )
2
2r
t
---------
------------------- 2r
3
t = =
ab =
ds
t
-----

2 a b + ( )
t
-------------------- =
J
4 ab ( )
2
2 a b + ( )
t
--------------------
--------------------
2 ab ( )
2
t
a b +
------------------- = =
Fig. 7.26 Torsion constants for thin-walled, single cell sections with circular and rectangular
contours
b
G 5
6
10 psi = G 2.5
6
10 psi =
10000 lb-in.
0.125 in. 0.250 in.
2.0 in.
Fig. 7.27 Bi-material circular torsion tube
q
T
2
--------
10000 lb-in.
2 1 in. ( )
2
----------------------------- 1591.5 lb/in. = = =

zs
1591.5 lb-in. ( ) 0.250 in. ( ) 6366.2 psi = =

zs
1591.5 lb-in. ( ) 0.125 in. ( ) 12732.4 psi = =
Thin-Walled Structures 219
Shear center of a closed section
Take the reference shear modulus psi. First we compute the denominator of the torsion con-
stant given by eq. (7.71).
Note that the modulus-weighted thickness of the right side wall is 2 (0.125 in.) = 0.250 in. From eq. (7.71) the
torsion constant is
the torsional stiffness is dened as , which equals via eq. (7.70). Thus, the torsional stiffness for
this example is

7.8 Shear center of a closed section
The unit twist formula given by eq. (7.62) is applicable even if the shear ow is a function of the contour
coordinate s, as is the case when the transverse shear forces are non-zero. Hence, eq. (7.62) can be used to com-
pute the unit twist of the section when the shear forces and and the torque T act in unison. In the case of
torsion only, we showed in the last section that the shear ow was constant along the contour coordinate based on
axial equilibrium, which led to the simplication given by eq. (7.69). If and are non-zero, then the com-
putation for the unit twist is a more involved. We illustrate the use of eq. (7.62) for the case of transverse shear
forces without torsion by locating the shear center of a closed section. There are three steps to locating the shear
center of a closed section.
Determine the shear ow distribution around the section due to bending using eq. (6.13) on page 160.
This step involves selecting a contour origin where s = 0. The shear ow q(0) at the contour origin is
unknown.
Determine the unknown shear ow q(0) by setting the unit twist, eq. (7.62), to zero.
Setting the unit twist to zero enforces the condition of no torsional deformation for the section.
Use torque equivalence to locate the shear center.
This procedure is best illustrated by example.
EXAMPLE 7.5 Shear center location of a single-cell, closed section having one axis of symmetry.
Consider the thin-walled closed section shown in Fig. 7.28. The contour of this section is an isosceles triangle
with each branch having the same thickness t and having the same shear modulus G. There is a horizontal axis of
symmetry, which is taken as the x-axis. The second area moment about the x-axis is , where t is
specied in cm. Since the shear center lies on this axis of symmetry, we take the shear force components
G
0
2.5
6
10 =
ds
t
*
-----

1 in. ( )
2.5
6
10 psi
G
0
---------------------------- 0.250 in. ( )
-------------------------------------------------------
1 in. ( )
5
6
10 psi
G
0
----------------------- 0.125 in. ( )
-------------------------------------------------- + 8 = =
J
4 1 in. ( )
2
[ ]
2
8
--------------------------------

2
--- in.
4
= =
T
d
z
dz
------------------ G
0
J
T
d
z
dz
------------------ G
0
J 2.5
6
10 psi ( )

2
--- in.
4


3.93
6
10 lb-in.
2
= = =
V
x
V
y
V
x
V
y
I
xx
300t cm
4
=
Bars Subjected to Torsional Loads
220 Thin-Walled Structures
and . Determine the location e
x
of the shear center from point O, which is located on the x-axis
at the vertex of the two sides of the triangle that are of equal length.
Solution The shear ows are assumed positive in the counter-clockwise direction with the contour origin at
point O. The shear ow at point O is q
0
, and eq. (6.13) on page 160 reduces to . For the
rst branch shown in Fig. 7.28 we have that the shear ow due to bending is given by
(7.73)
The shear ow due to bending in the second branch is
From eq. (7.73) . Hence the shear ow in the second branch is
(7.74)
The shear ow due to bending in the third branch is
From eq. (5.44) , which is the same value as the shear ow at the junction of
branches one and two. Hence the shear ow in branch three is
(7.75)
V
x
0 = V
y
0 >
y
x
C
5
5
12
4
1
3
---
t
V
y
O
e
x
S.C.
s
1
q
1
,
s
2
q
2
,
s
3
q
3
,
q
0
O
All dimensions in cm
Fig. 7.28 A closed section with an isosceles triangle contour
5
13
12
q s ( ) q
0
V
y
I
xx
------Q
x
s ( ) =
q
1
s
1
( ) q
0
V
y
300t
-----------
5
13
------s
1


t s
1
d
0
s
1

q
0
V
y
1560
------------ s
1
2
( ) = = 0 s
1
13 cm
q
2
s
2
( ) q
1
13 ( )
V
y
300t
----------- 5 s
2
( )t s
2
d
0
s
2

= 0 s
2
10 cm
q
1
13 ( ) q
0
13V
y
( ) 120 =
q
2
s
2
( ) q
0
V
y
120
--------- 13 2s
2
1
5
---s
2
2
+


= 0 s
2
10 cm
q
3
s
3
( ) q
2
10 ( )
V
y
300t
----------- 5
5
13
------s
3
+


t s
3
d
0
s
3

= 0 s
3
13 cm
q
2
10 ( ) q
0
13V
y
( ) 120 =
q
3
s
3
( ) q
0
V
y
120
--------- 13 2s
3

1
13
------s
3
2
+


= 0 s
3
13 cm
Thin-Walled Structures 221
Shear center of a closed section
Note from this result that , as it should since the shear ow at point O is unique.
To nd the shear ow at point O, we impose the condition of zero unit twist from eq. (7.62). Since the prod-
uct of Gt is the same in each branch, eq. (7.62) reduces to
which requires the contour integral of the shear ow around the entire section to vanish. Note that the unit twist is
positive counter-clockwise and consequently a counter-clockwise shear ow is positive in eq. (7.62). Substituting
eqs. (7.73) to (7.75) into this integral, we get
After evaluation of the denite integrals, this equation becomes
(7.76)
Substitute q
0
from eq. (7.76) into eqs. (7.73) to (7.75) to nd that the shear ows due to bending only are
(7.77)
(7.78)
(7.79)
For each straight branch we can compute the branch force. These branch forces are
(7.80)
(7.81)
(7.82)
These branch forces are shown in Fig. 7.29. Now torque equivalence is used to determine the location of the
shear center on the axis of symmetry. Torque equivalence gives
q
3
13 ( ) q
0
=
d
z
dz
--------
1
2 Gt ( )
------------------- qds

0 = =
qds

q
0
V
y
1560
------------ s
1
2
( )


s
1
d
0
13

q
0
V
y
120
--------- 13 2s
2
1
5
---s
2
2
+



s
2
d
0
10

+ + =
q
0
V
y
120
--------- 13 2s
3

1
13
------s
3
2
+



s
3
d
0
13

0 =
36q
0
23
10
------V
y
0 = Hence q
0
23
360
---------V
y
=
q
1
s
1
( )
V
y
4680
------------ 299 3s
1
2
( ) =
q
2
s
2
( )
V
y
1800
------------ 80 30s
2
3s
2
2
+ ( ) =
q
3
s
3
( )
V
y
4680
------------ 208 78s
3
3s
3
2
+ ( ) =
f
1
q
1
s
1
d
0
13

V
y
4680
------------ 299 3s
1
2
( ) s
1
d
0
13

13
36
------V
y
= = =
f
2
q
2
s
2
d
0
10

V
y
1800
------------ 80 30s
2
3s
2
2
+ ( ) s
2
d
0
10

13
18
------V
y
= = =
f
3
V
y
4680
------------ 208 78s
3
3s
3
2
+ ( ) s
3
d
0
13

13
36
------V
y
= =
12 cm
13
18
------V
y


e
x
V
y
= Hence e
x
8
2
3
--- cm =
Bars Subjected to Torsional Loads
222 Thin-Walled Structures
The shear center and the centroid are shown in Fig. 7.30.
7.9 Uniform torsion of multi-cell closed sections
Consider a cylindrical shell-beam of length L whose cross section consists of two closed cells. The z-axis is par-
allel to the length of the cylindrical shell-beam, and the x- and y-axes are in a plane parallel to the cross sections.
The shell-beam is subjected to equal and oppositely directed torques T at each end. Hence, equilibrium requires
that each cross section is subjected to a torque equal to T, and the torque is taken positive if it act counter-clock-
wise on a positive z-face. As is shown in Fig. 7.31, the cell on the left side of Section A-A is designated as cell 1
and the cell to the right is designated 2. The thickness of the exterior branches of cells 1 and 2 are denoted by t
1
and t
2
, respectively. The thickness of the common branch separating the two cells is denoted by . All
branches are assumed made of the same homogeneous material with the shear modulus denoted by G. Our pur-
pose is to determined the shear ows and shear stresses due to torsion, and to determine the torsion constant J in
the standard formula where is the twist per unit length.
We assume positive shear ow directions as shown in Fig. 7.32. The shear ows in the exterior branches of
cells 1 and 2 are taken positive counter-clockwise and are denoted by q
1
and q
2
, respectively. The shear ow in
the common branch is denoted by q
1-2
. The common branch shear ow is related to the shear ows in the exte-
rior branches of cells 1 and 2 by axial equilibrium at the junction of the common branch with the exterior
branches of cells 1 and 2. This junction is labeled O in Fig. 7.32, and the free body diagram at the junction is
shown in Fig. 7.33.
O
13
36
------V
y
13
36
------V
y
13
18
------V
y O
V
y
e
x

Fig. 7.29 Static equivalence to nd the shear center of the triangular contour
O
C S.C.
7
2
3
---cm
8
2
3
---cm
12cm
Fig. 7.30 Shear center and centroid
location of the triangular section
t
1 2
T GJ
d
z
dz
-------- = d
z
dz
Thin-Walled Structures 223
Uniform torsion of multi-cell closed sections
Axial equilibrium at the junction point O gives
(7.83)
Axial equilibrium at the upper junction of the common web with the exterior branches of cells 1 and 2 gives the
same result as eq. (7.83). An easy way to determine the axial equilibrium of the junction is to recall that the shear
t
2
t
1
t
1 2
x
y
Section A-A

1

2
L
A
A
z
y
T T
Fig. 7.31 Torsion of a shell-beam with a cross section consisting of two closed cells
t
2
t
1
t
1 2
x
y
q
1 2
q
1
q
1
q
2
q
2
O

1

2
Fig. 7.32 Assumed positive shear ows for the two-cell section
q
1 2
q
1
q
1
q
2
q
2
z
z
O
Fig. 7.33 Free body diagram of the junction at point O
q
1 2
q
1
q
2
=
Bars Subjected to Torsional Loads
224 Thin-Walled Structures
ow into the junction must equal the shear ow out of the junction.
Now we determine the torque about point O due to the shear ows. The shear ow q
1-2
does not contribute to
the torque about point O since the contour of the common branch is straight and it passes through point O. Using
Bredts formula, the torque about point O is
(7.84)
where denotes the area enclosed by a cell. Actually, the same value of the torque is obtained for any point in
the plane of the cross section where the resultant of the shear ow distribution is resolved. The fact that the
torque due to the shear ows is the same about any point in the plane is discussed in Section 7.10.
At this point we have two equations, eqs. (7.83) and (7.84), for the three unknown shear ows assuming that
the torque and geometry of the section are known. An additional equation relating the shear ows is based on the
assumed rigidity of the cross section in its own plane. In torsion, this rigid cross section condition implies the
unit twist of each cell is the same. The unit twist for a single cell is given by eq. (7.62). Since the shear modulus
is the same for all branches in the cross section eq. (7.62) reduces to
This unit twist formula was derived on the basis that a counter-clockwise shear ow tends to produce a counter-
clockwise unit twist. Apply this equation to cell 1 to get
(7.85)
where S
1
is the arc length of the exterior branch of cell 1, and S
1-2
is the length of the common branch between
cells 1 and 2. For cell 2, the unit twist is
(7.86)
where S
2
is the arc length of the two exterior branches of cell 2. Note that the contribution of the common branch
shear ow is negative in the unit twist formula for cell 2. Relative to an observer in cell 2, a positive value for the
shear ow q
1-2
tends to produce a clockwise unit twist, and hence is negative by the convention that counter-
clockwise is positive. Since the unit twist of each cell is the same, we equate eqs. eq. (7.85) and eq. (7.86) to get
(7.87)
We now have three equations for the three unknown shear ows; eqs. (7.83), (7.84), and (7.87). These equa-
tions are solved for the shear ows in terms of the torque and geometry of the cross section. Once the shear ows
are known, the unit twist for either cell 1, eq. (7.85), or cell 2, eq. (7.86) can be determined in terms of the torque,
geometry, and shear modulus. The unit twist computed for one cell is the unit twist for the entire cross section,
since we equated the unit twist of each cell to one another. Finally, compare the computed unit twist to the stan-
dard formula to identify the value of the torsion constant J. For a multi-cell section, the formula
given for the torsion constant of a single cell section, eq. (7.71), is not applicable.
T 2
1
q
1
2
2
q
2
+ =

d
z
dz
--------
1
2G
------------
q
t
---ds

=
d
z
dz
--------


1
1
2
1
G
---------------
S
t
---


1
q
1
S
t
---


1 2
q
1 2
+ =
d
z
dz
--------


2
1
2
2
G
---------------
S
t
---


2
q
2
S
t
---


1 2
q
1 2
=
1
2
1
G
---------------
S
t
---


1
q
1
S
t
---


1 2
q
1 2
+
1
2
2
G
---------------
S
t
---


2
q
2
S
t
---


1 2
q
1 2
=
d
z
dz
--------
T
GJ
------- =
Thin-Walled Structures 225
Uniform torsion of multi-cell closed sections
EXAMPLE 7.6 Uniform torsion of a two-cell section
All branches of the two cell section shown in Fig. 7.31 are made of the same material, whose shear modulus is
denoted by G. The exterior branches have thickness t and the common branch has a thicknesses 3t. The contour
of the nose cell is semi-circular with radius r, and the contour of the second cell is an isosceles triangle with equal
sides of length r. Note that the horizontal length of the isosceles triangle is approximately 3r, and we will use
this approximation to simplify the algebra. Assume the torque T is given in addition to the geometry, and deter-
mine
shear ows q
1
and q
2
in the exterior branches of cells 1 and 2, and
the torsion constant J.
Solution From eq. (7.84) the torque is
or
(7.88)
Using eq. (7.85) for the unit twist for cell 1, and eq. (7.83) for the common web shear ow, we get
or
(7.89)
Using eq. (7.76) for the unit twist for cell 2, and eq. (7.83) for the common web shear ow, we get
or
q
1
q
1 2
q
2
q
2

1

2 r
t
t
3t
r
3r
T
x
y
Fig. 7.34 Uniform torsion of a two-cell section
2
r
2
2
--------


q
1
2
2r 3r
2
----------------


q
2
+ T =
r
2
q
1
6r
2
q
2
+ T =
d
z
dz
--------


1
1
2
r
2
2
--------


G
----------------------
r
t
------


q
1
2r
3t
-----


q
1
q
2
( ) + =
d
z
dz
--------


1
1
rt ( )G
-----------------
2
3
--- +


q
1
2
3
---q
2
=
d
z
dz
--------


2
1
2 3r
2
( )G
---------------------
2r
t
---------


q
2
2r
3t
-----


q
1
q
2
( ) =
Bars Subjected to Torsional Loads
226 Thin-Walled Structures
(7.90)
Now equate eqs. (7.89) and (7.90) to get
which after rearrangement gives
(7.91)
Equations (7.88) and (7.91) are two simultaneous equations for the shear ows q
1
and q
2
in terms of the torque.
The solution to these simultaneous equations determines the shear ows. Hence,
(7.92)
To nd the torsion constant J, we substitute the solutions for the shear ows into eq. (7.89). The result is
The same result is obtained using the unit twist for cell 2; i.e., eq. (7.90). Comparing the unit twist result given
above to the standard torsional relation , we nd that the torsion constant J is
(7.93)
7.10 Resultant of a constant shear ow in a curved branch
In torsion problems it often necessary to nd the resultant of a constant shear ow along the contour of a curved
branch. This situation is depicted in Fig. 7.35. The resultant of the shear ow is resolved at point A in this g-
ure. From the differential element shown in Fig. 7.35(b), the differential force and differential torque resolved at
point A are
From the differential geometry shown in Fig. 7.35, the product of r times ds is twice the enclosed area of the tri-
angle with base ds and height r. Integrating the above expressions from point A to point B on the contour we get
(7.94)
where denotes is the area enclosed by the contour and the chord connecting the end points of the contour.
From the components of the force given above, the magnitude is determined as
d
z
dz
--------


2
1
6rt ( )G
-----------------
2
3
---q
1
2
2
3
--- +


q
2
+ =
1

---
2
3
--- +


q
1
2
3
---q
2

1
6
---
2
3
---q
1
2
2
3
--- +


q
2
+ =
1
2
3
------
1
9
--- + +


q
1

3
---
1
9
---
2
3
------ + +


q
2
0 = or 1.3233177q
1
1.3705152q
2
0 =
answer : q
1
T
8.9350r
2
--------------------- = q
2
T
9.2536r
2
--------------------- =
d
z
dz
--------
d
z
dz
--------


1
0.11273713
T
Gr
3
t
----------- = =
d
dz
------
T
GJ
------- =
1
J
---
0.11273713
r
3
t
---------------------------- = giving the answer J 8.870r
3
t =
dF
x
qds ( )
dx
ds
------


qdx = = dF
y
qds ( )
dy
ds
------


qdy = = dT r qds ( ) =
F
x
q x d
A
B

q x
B
x
A
( ) = = F
y
q y d
A
B

q y
B
y
A
( ) = = T 2q =

Thin-Walled Structures 227


Resultant of a constant shear ow in a curved branch
That is, L denotes the length of the chord connecting the end points of the contour. Moreover, the line of action of
the force is parallel to the chord, since . The force and torque resolved at point A
are shown in Fig. 7.36(a). The nal reduction is to move the force qL to a parallel line of action from point A to
eliminate the torque, and this is shown in Fig. 7.36(b). Thus, the resultant of a constant shear ow in a curved
branch is a force of magnitude equal to the shear ow times the length of the chord connecting the end points of
the contour. The line of action of the resultant force qL is parallel to the chord, offset from the chord by a distance
equal to twice the area enclosed by the contour and the chord divided by the length of the chord, measured per-
pendicular to the chord.
A
x
y
qds
ds
r
dx
dy
A
B
x
x
B
x
A
y
A
y
B
y
q
(a) Curved contour (b) Differential element on contour
Fig. 7.35 Constant shear ow along a curved branch.
d
F F
x
2
F
y
2
+ qL = = where L x
B
x
A
( )
2
y
B
y
A
( )
2
+ =
F
y
F
x
y
B
y
A
( ) x
B
x
A
( ) =
A
B
x
y
qL
A
B
x
x
B
x
A
y
A
y
B
y
qL
L
(a) Force and torque at A (b) Resultant
Fig. 7.36 Steps to resolving the constant shear to a single force along a specic line of action.
2 L
2q
Bars Subjected to Torsional Loads
228 Thin-Walled Structures
From the above discussion, a constant shear ow in a closed contour is statically equivalent to a zero force,
since the beginning and end points of the closed contour coincide. Refer to eqs. eq. (7.94). However, a constant
shear ow in a closed contour is statically equivalent to a torque, and this torque is the same for any point in the
plane about which moments are computed. The fact that the torque is a free vector is depicted in Fig. 7.37, in
which it is shown that some of the enclosed area used in Bredts formula can add as a negative quantity if the
torque produced by the constant shear ow is clockwise over a segment of the branch. About point O in this g-
ure, the torque produced by the shear ow from point B to A in the right half of the contour is counter-clockwise,
and the torque produced by the shear ow from A to B in the left half is clockwise. Hence, the total torque is the
sum of these two torques with due respect to the sign. This summation shows that the total torque is proportional
to the area enclosed by the contour.
Finally consider torsion of a multi-cell section. Each branch composing the cross section has a constant
shear ow, but the shear ow from branch-to-branch is, in general, a different value. Axial equilibrium requires
that the shear ows into a junction must equal the shear ow out of the junction. If the shear ows satisfy this
condition at all junctions, then the shear ow distribution is statically equivalent to a torque and no force. Hence,
if the resultant is only a torque, then the point in the cross section about which we sum moments is immaterial
because the torque is the same about any point. A two cell section is shown in Fig. 7.38. Since the shear ow is
constant in each branch, the resultant for each branch can be determined by considerations of the single curved
branch given above. The resultant of the exterior branch in cell 1 is a downward force of magnitude ,
where is the length of the chord connecting the end points of the contour of the exterior branch. This chord
length, as can be seen in the gure, is the same as the length of the common branch. The line of action of this
resultant is parallel to the chord with the distance from the chord equal to twice the enclosed area of cell 1
divided by the chord length. Similarly, the resultant of the exterior branches of cell 2 is an upward force of mag-
nitude with a line of action parallel to the common branch. The common branch resultant is an upward
force of magnitude whose line of action is the contour of the common branch. From Fig. 7.38,
the force and torque about any point on the contour of the common branch are
These resultants conrm eq. (7.84).
O
A
B
q
q
A
B
O
O
q

a
2
a
q

b
2
b
q

2q

a

b
=
Fig. 7.37 The torque about an arbitrary point O of a constant shear ow in a closed contour is twice the
enclosed area of the circuit times the shear ow.
q
1
S
1 2
S
1 2
q
2
S
1 2
q
1
q
2
( )S
1 2
F q
1
S
1 2
q
1
q
2
( )S
1 2
q
2
S
1 2
+ + 0 = =
T
2
1
S
1 2
----------- q
1
S
1 2
( )
2
2
S
1 2
----------- q
2
S
1 2
( ) + 2
1
q
1
2
2
q
2
+ = =
Thin-Walled Structures 229
Resultant of a constant shear ow in a curved branch
EXAMPLE 7.7 Torsion of a ve cell closed section; circuit shear ow
Consider the ve cell section shown in Fig. 7.39. All branches have the same thickness t and shear modulus G.
For an applied torque T, determine
shear ows q
1
, q
2
, q
3
, q
4
, and q
5
, and
the torsion constant J
Solution It is convenient to dene circuit shear ows in each cell. Circuit shear ows are assumed to be positive
counter-clockwise in each cell and are equal to the actual shear ows in the exterior branches of the cell, if there
are any exterior branches. However, the shear ow in a common branch between cells is the difference in the cir-
cuit shear ows sharing the common branch. This procedure automatically satises shear ow continuity at the
junctions;. i.e., axial equilibrium at the junctions. See the sketch of the center junction in Fig. 7.39.
q
1
q
2

2
1
S
1 2
------------
2
2
S
1 2
------------
q
2
S
1 2
q
1
q
2
( )S
1 2
q
1
S
1 2
q
1
q
2

S
1 2
F
T

1

2
Fig. 7.38 Branch-by-branch statical equivalence of a two-cell section
q
1
q
4
q
5
a a
a
a
T
a
q
2
q
3
3a/2 3a/2
q
4
q
1
q
2
q
4

q
1
q
2

Balance of shear flows


at the center junction
Fig. 7.39 Uniform torsion of a ve cell section
Bars Subjected to Torsional Loads
230 Thin-Walled Structures
Using Bredts formula to compute the torque for each cell, then summing these individual torques, assuming
counter-clockwise as positive, results in
(7.95)
The twist per unit length for each cell, with positive shear ows taken counter-clockwise within a cell, is
(7.96)
(7.97)
(7.98)
(7.99)
(7.100)
Since the cross section is assumed to be rigid in its own plane, the unit twist of each cell must be the same. This
kinematic condition can be written as
(7.101)
Substitute eqs. (7.96) to (7.100) into eqs. (7.101) to get four equations for the ve shear ows. Then add eq.
(7.95) to get the fth equation. These ve equations in matrix form are
(7.102)
Solving eq. (7.102) for the shear ows, we get
The unit twist of the section can be determined by substituting the above shear ows into any one of the eqs.
2
3
2
---a
2


q
1
2
3
2
---a
2


q
1
2 a
2
( )q
3
2 a
2
( )q
4
2 a
2
( )q
5
+ + + + T =
d
z
dz
--------


1
1
2
3
2
---a
2


Gt
------------------------
5
2
---a


q
1
a q
1
q
2
( )
a
2
---


q
1
q
4
( ) a q
1
q
5
( ) + + + =
d
z
dz
--------


2
1
2
3
2
---a
2


Gt
------------------------
5
2
---a


q
2
a q
2
q
3
( )
a
2
---


q
2
q
4
( ) a q
2
q
1
( ) + + + =
d
z
dz
--------


3
1
2 a
2
( )Gt
-------------------- 2a ( )q
3
a q
3
q
4
( ) a q
3
q
2
( ) + + [ ] =
d
z
dz
--------


4
1
2 a
2
( )Gt
-------------------- aq
4
a q
4
q
5
( )
a
2
---


q
4
q
1
( )
a
2
---


q
4
q
2
( ) + + + =
d
z
dz
--------


5
1
2 a
2
( )Gt
-------------------- 2a ( )q
5
a q
5
q
1
( ) a q
5
q
4
( ) + + [ ] =
d
z
dz
--------


1
d
z
dz
--------


2
=
d
z
dz
--------


2
d
z
dz
--------


3
=
d
z
dz
--------


3
d
z
dz
--------


4
=
d
z
dz
--------


4
d
z
dz
--------


5
=
d
z
dz ( )
1
d
z
dz ( )
2
= :
d
z
dz ( )
2
d dz ( )
3
= :
d
z
dz ( )
3
d
z
dz ( )
4
= :
d
z
dz ( )
4
d
z
dz ( )
5
= :
eq. (5.65):
2 2
1
3
--- 0
1
3
---
1
3
---
13
6
------
7
3
---
1
3
--- 0
1
4
---
1
4
--- 2
1
2
---
1
2
---
1
4
---
1
4
--- 0 2
5
2
---
3a
2
3a
2
2a
2
2a
2
2a
2
q
1
q
2
q
3
q
4
q
5
0
0
0
0
T
=
q
1
q
2
23
281
---------
T
a
2
----- = = q
3
22
281
---------
T
a
2
----- = q
4
55
562
---------
T
a
2
----- = q
5
22
281
---------
T
a
2
----- =
Thin-Walled Structures 231
Torsion of hybrid sections
(7.96) to (7.100). Using eq. (7.96), the unit twist is
Compare this to the standard formula to nd

If all the common branches were removed to make the section shown in Fig. 7.38 a single cell of rectangular
section 2a by 3a, then from eq. (7.71) the torsion constant is . For this example, subdividing the
single cell section into ve cell section shown in Fig. 7.38 increases the torsional stiffness by only 4.07% with
respect to the single cell section, while the weight of the ve cell section increases by 60% with respect to the
weight of the single cell section. However, a multi-cell section may be required for improved damage tolerance;
i.e., if we modeled damage as a fracture, or cut, of an exterior branch, then the loss of torsional stiffness of the
single cell would be substantial since it becomes an open section. Damage to an exterior branch of a multi-cell
section on the other hand results in less of a reduction in torsional load carrying capability since some closed
cells remain intact to carry the torsional load.
7.11 Torsion of hybrid sections
Consider a hybrid section composed a single closed cell
and open branches, or ns, as shown in Fig. 7.40 The total
torque carried by the section is the sum of the torques car-
ried by the closed cell and open branches. For n open
branches, we have
(7.103)
with
(7.104)
The torsional stiffness for the closed cell is
(7.105)
and for each open branch the torsional stiffness is
(7.106)
Combining eqs. (7.103) and (7.104), we have
d
z
dz
--------
75
1124
------------
T
a
3
tG
------------ =
d
z
dz
--------
T
GJ
------- =
1
J
---
75
1124
------------
1
a
3
t
------- = Hence J
1124
75
------------a
3
t 14.9867a
3
t = =
J 14.40a
3
t =
b
i
t
i
t
i =1
2
i=n
n-1
T

Fig. 7.40 Torsion of a hybrid section
T T
closed
T
i
i 1 =
n

+ =
T
closed
GJ ( )
closed
d
z
dz
--------


= T
i
GJ ( )
i
d
z
dz
--------


=
GJ ( )
closed
4
2
ds
Gt
------

---------- =
GJ ( )
i
G
i
1
3
---b
i
t
i
3


=
Bars Subjected to Torsional Loads
232 Thin-Walled Structures
Comparing this result to the standard torsional formula , the effective torsional stiffness for
the entire section is
(7.107)
where the closed and open parts of the torsional stiffness are given by eqs. (7.105) and (7.106).
The shear stress in the closed cell is , where the portion of the torque carried by the
closed cell is . The unit twist is given by . Combining
these results, the shear stress in the closed cell is
(7.108)
For the open branches, we use the formula for the shear stress given by the rst of eqs. (7.21). Repeating this
equation, using the second of eqs. (7.104) for the torque carried by the branch, and then using eq. (7.106) for the
torsional stiffness of the branch, we can formulate the shear stress as
The unit twist was eliminated in this equation using the overall section formula for it. After simplication of the
above result we get
(7.109)
If the shear modulus is the same in all branches, then eqs. (7.108) and (7.109) reduce to
(7.110)
7.12 References
Oden, J. T., and Ripperger, E. A., 1981, Mechanics of Elastic Structures, Second Edition, Hemisphere Publish-
ing Corporation, New York, pp.51-56.
Timoshenko, S.P., and Goodier, J.N., 1970, Theory of Elasticity, Third Edition, McGraw-Hill Book Company,
New York, pp. 291-313.
T GJ ( )
closed
GJ ( )
i
i 1 =
n

+
d
z
dz
-------- =
T GJ ( )
eff
d
z
dz
--------


=
GJ ( )
eff
GJ ( )
closed
GJ ( )
i
i 1 =
n

+ =

closed
T
closed
2t ( ) =
T
closed
GJ ( )
closed
d
z
dz ( ) = d
z
dz ( ) T GJ ( )
eff
=

closed
GJ ( )
closed
GJ ( )
eff
-------------------------
T
2t
---------- =

i
3T
i
b
i
t
i
2
---------
3
b
i
t
i
2
--------- G
i
1
3
---b
i
t
i
3


d
z
dz
--------
3
b
i
t
i
2
--------- G
i
1
3
---b
i
t
i
3


T
GJ ( )
eff
------------------ = = =

i
G
i
t
i
GJ ( )
eff
------------------T =

closed
J
closed
J
---------------
T
2 t
---------- =
i
Tt
i
J
------- =
Thin-Walled Structures 233
Problems
7.13 Problems
1. Determine torsion constant J and the magnitude of the maximum shear stress
for the section shown.
2. The wall thickness for each single cell section shown below is 1.5 mm, and
the maximum shear stress for the material is 2.5 MPa. Determine the maximum
torque that can be applied to section (a) and section (b).
3. The uniform, thin-walled, box beam shown below is clamped at the wall and free at the tip. It is subjected to
a uniform distributed load of intensity 40 N/mm along the front web. The shear modulus of the top and bottom
skins is 18 GPa, and the shear modulus of the webs is 26GPa. Determine the distribution of the twist; i.e, function
, .
4. The thin-walled, closed section shown below consists of two straight branches inclined at 45 degrees from
the horizontal axis and a circular branch of radius r. The x-axis is an axis of symmetry and the second area
moment about the x-axis is . All branches have the same thickness t and same shear modulus
G.
a) Determine the shear ow distributions in each branch due to bending in terms
2t
t
t
3b 2
b
b

max
50 mm 50 mm
50 mm
50 mm
10mm
10mm
1.5mm
1.5mm
20mm
20mm
(a) (b)
Figure for problem 2

z
z ( ) 0 z 2500mm
40N/mm
x
y
1.2mm
1.2mm 2.1mm
2.1mm
1000mm
250mm
G 18GPa =
G 26GPa =
cross section
40 N/mm
x
y
z
2500mm

z
Figure for problem 3
I
xx
0.618731r
3
t =
q
1
s ( ) q
2
s ( ) and q
3
s ( ) , ,
Bars Subjected to Torsional Loads
234 Thin-Walled Structures
of shear force , the shear ow at point O, and radius r. Take positive shear ow counter-clockwise
around the section as shown with the contour origin at point O. Note: in the circular branch 2 take
, , and hence the y-coordinate to point s
2
is
b) Determine the shear ow

at point O in terms of and r by enforcing zero unit twist.(answer:
)
c) Determine the location of the shear center e from point O in terms of r by torque equivalence.
5. The two cell closed section bar is subjected to uniform torsion with a torque T as shown. Each branch has the
same thickness t and same shear modulus G.
a) Determine the shear ows q
1
and q
2
in the exterior branches.
b) Determine the torsion constant J.
V
y
q
0
s
2
r = 0 2 y s
2
( ) r 4 + ( ) cos =
q
0
V
y
q
0
0.495416V
y
r =
t
e
S.C.
axis of symmetry
O
45
45
r
x
y
s
1
q
1
,
s
2
q
2
,
s
3
q
3
,
V
y
Figure for problem 4
a
a 2a
t, typ.
q
1
q
2
cell 1 cell 2
T
Figure for problem 5
Thin-Walled Structures 235
Problems
6. A small slit is cut in the lower exterior branch in cell 2 of problem 5.
a) Determine the torsion constant J.
b) Determine the maximum shear stress.
7. For the homogeneous cross section shown below, take dimensions and .
a) Determine the torsion constant J.
b) For a torque , determine shear stress in the closed circular portion and the maximum
shearing stress in the open portion.
8. Determine the location of the shear center in terms of dimension b for the section shown below.
a
a 2a
t, typ.
q
T
small slit
Figure for Problem 6
b 30 mm = t 3 mm =
T 600 Nm =
e
y
b
2
---
b
b b
C
SC
y
c
e
y
x
y
the thickness of all branches is t
y
c
0.8487b =
I
yy
2
3
---

8
--- +


b
3
t =
Figure for Problems 7 and 8
Bars Subjected to Torsional Loads
236 Thin-Walled Structures

Thin-Walled Structures

237

CHAPTER 8

Criteria for Initial
Yielding

The yield stress of a material is determined from the tensile test as discussed in The tensile test on page 19. The
yield stress is an important material property in the design of structures made of ductile materials, so ductility is
discussed in this chapter. Also, the tensile test is conducted under axial loading such that the state of stress in the
gage length is simple tension. So the questions arises as to how to use the yield stress determined in simple labo-
ratory specimens in actual structural components subjected to multi-axial states of stress. Criteria are discussed
in this chapter for the prediction of the initiation of material yield under combined stress states.

8.1 Ductile and brittle behavior

A ductile material, like many metals, is one for which the plastic deformation before fracture is much larger than
the elastic deformation. A brittle material, like ceramics and glasses, exhibits little deformation before fracture;
i.e., fracture occurs without very much plastic deformation. See Fig. 8.1. A measure of ductility is the engineer-
ing strain at fracture, which is usually report in percent: i.e, . A second, widely reported measure of ductil-
ity is the percent reduction in area, denoted as %RA, that is dened by


0 0
fracture
fracture
ductile material brittle material
Fig. 8.1 Ductile and brittle material responses
100
f

Criteria for Initial Yielding

238

Thin-Walled Structures

where is the initial cross-sectional area, is the minimum cross-sectional area at fracture (occurring in the
necked-down region), and and are the corresponding diameters for round specimens. Example values of
the percent reduction in area (Dowling,1992, p. 157) are


35%

RA

for 2024-T4 aluminum,


33%

RA

for 7075-T6 aluminum,


61%

RA

for AISI 1020 steel, and


6%

RA

for AISI 4142 steel as quenched.
The abbreviation AISI stands for the

American Iron and Steel Institut

e. Some high strength alloys (like spring
steel for instance) have ductilities as low as 2%, but even this is enough to insure that the material yields before it
fractures and that fracture, when it occurs, is of tough ductile type (Ashby, 1992, p. 14).
An important attribute in design with ductile materials is their capacity to accommodate stress concentra-
tions through plastic deformation and hence to redistribute the stresses more evenly. Stress concentrations occur
at stress raisers which are either geometric discontinuities (e.g., holes, sharp corners, cracks, llets, etc.) and/or
material discontinuities. Geometric discontinuities are commonly referred to as notches. The capacity to redis-
tribute stresses at stress riser makes a ductile material tough, giving the material a defense mechanism against
stress concentrations. Brittle materials, on the other hand, do not possess such a defense against stress concentra-
tion so that even small scratches and cracks as naturally occur in their fabrication can lead to brittle fracture. For
this reason, brittle materials have to be used with extreme caution in tension structures. Ductile engineering
materials are those for which static strength in engineering applications is limited by yielding and not fracture.

8.2 Criteria for initial yielding of ductile materials

The yield stress is determined from the tensile test data, but the tensile test is designed to produce a uniaxial
state of stress. However, we would like to know what governs yielding under combined states of stress that occur
in structural components under service loads. That is, for a three-dimensional state of stress as shown in Fig. 8.2,
what is condition for initiation of yield?
There is no theoretical way to correlate yielding in a three-dimensional stress state with yielding in the
uniaxial tensile test. Two empirical equations have been proposed which are reasonably simple and are descrip-
tive of the data. These criteria are:


the Mises yield criterion (also known as the distortion energy criterion or the octahedral shear stress crite-
rion), and the


Maximum shear-stress criterion (also known as the Tresca criterion).
Each criterion is based on two observations, which are:


The state of stress can be completely described by the magnitude and direction of principal stresses. For an
isotropic material, the orientation of the principal stresses is unimportant, so only the magnitudes enter the
criteria.


Experiments show that a hydrostatic state of stress does not affect yielding (Dowling, 1993, pp. 106-108).
In the remainder of this section, principal stresses and the hydrostatic stress are dened and discussed. Mises cri-
%RA 100
A
i
A
f

A
i
-----------------


100
d
f
2
d
i
2

d
i
2
-----------------


= =
A
i
A
f
d
i
d
f

y

Thin-Walled Structures

239

Criteria for initial yielding of ductile materials

terion and the maximum shear-stress criterion are presented in Section 8.6 and Section 8.7, respectively.
Consider an innitesimal rectangular parallelepiped at a material point, denoted by

P

, in a thin-walled beam.
The element is referenced to the three mutually perpendicular directions

z

,

s

, and

n

, where the

z

-direction is par-
allel to the beam longitudinal axis, the

s

-direction is tangent to the contour, and the

n

-direction is parallel to the
thickness of the wall measured normal to the contour. In the most general state of stress there are three stress
components acting on each of the three faces normal to the coordinate directions. These nine stress components
can be written in the matrix form

(8.1)

in which the rst row contains stress components acting on the

z

-face, the second row contains components act-
ing on the

s

-face, and the third row contains stresses acting on the

n

-face. Column one contains stress compo-
nents acting in the

z

-direction, the second column contains stress components acting in the

s

-direction and the
third column contains stress components acting in the

n

-direction. At point

P

, it is possible to nd three mutually
perpendicular planes through the point on which only normal stresses act, and the shear stresses vanish. This par-
ticular triple of normal stresses are called principal stresses, and principal stresses are an equivalent description
of the state of stress at point

P

. See Fig. 8.3. The principal stresses are denoted by . The determi-
nation of the principal stresses is discussed Section 8.4. The important point is to realized that although the stress
matrix is fully populated in, say, the

z, s, n

coordinate system, it is possible to nd three mutually perpendicular
directions at the point, labeled , in which the stress matrix is a diagonal matrix. An important part
of the proof of this statement, which is not presented here, is that the stress matrix is symmetric for any three
mutually perpendicular directions at the point. That is, moment equilibrium about the

n

-axis at point

P

leads to
. Likewise, moment equilibrium about the

z

-axis and the

s

-axis at

P

leads to and ,
respectively. In matrix notation moment equilibrium at a point implies symmetry of the stress matrix, which is
written as

s
s
n
z

sn

ns

sz

zs

nz

zn
P

z

zs

zn

sz

s

sn

nz

ns

n
=
Fig. 8.2 Three dimensional state of stress at a point in the material.

z

zs

zn

sz

s

sn

nz

ns

n
=

1

2
and
3
, ,
x
1
x
2
and x
3
, ,

zs

sz
=
sn

ns
=
nz

zn
=

Criteria for Initial Yielding
240 Thin-Walled Structures
(8.2)
where the superscript T denotes matrix transpose. Thus, there only six independent stress components at a point.
Hydrostatic stress, denoted by , is dened by
(8.3)
where denotes the trace of the stress matrix. The trace of a square matrix is dened as the sum of its diag-
onal elements. We will show in Section 8.3, eq. (8.15), that the trace of the stress matrix is invariant with respect
to a rotation of the three mutually perpendicular axes through the point. Thus, we also have
(8.4)
In the case of uids at rest, the pressure, denoted by p, is the hydrostatic stress. The stress matrix for any three
mutually perpendicular axes at a point in the uid is
(8.5)
x
1
x
2
x
3

p
[ ]

1
0 0
0
2
0
0 0
3
=
Fig. 8.3 Equivalent descriptions of the state of stress at a material point in the body.

z

zs

zn

sz

s

sn

nz

ns

n
=

s
s
n
z

sn

ns

sz

zs

nz

zn
P
P
[ ] [ ]
T
= or

z

zs

zn

sz

s

sn

nz

ns

n

z

sz

nz

zs

s

ns

zn

sn

n
=

h
1
3
---
x

y

z
+ + ( )
1
3
---Tr [ ] = =
Tr [ ]

h
1
3
---
1

2

3
+ + ( ) =
[ ]
p 0 0
0 p 0
0 0 p
=
Thin-Walled Structures 241
Stress transformation equations for generalized plane stress
That is, the normal stress on all faces through the point is the negative of the pressure, and all directions at the
point are principal stress directions. (See Example 8.2). Hence, by the denition given in eq. (8.3), the hydro-
static stress .
8.3 Stress transformation equations for generalized plane stress
For thin-walled structures we neglect the shear stress components in the thickness direction, with
respect to the shear stress component tangent to the contour . The stress matrix in the z, s, n system reduces to
what is called generalized plane stress. That is,
(8.6)
For this generalized state of plane stress, we want to determine the principal stresses and the maximum shear
stress at a material point, again labeled P, in the wall. To determine principal stresses we use equilibrium to nd
the normal stress and shear stress on face through point P whose normal is in an arbitrary direction in the z-s
plane. Hence, consider a second orthogonal coordinate system which is rotated about the n-axis through
the angle , dened positive counter-clockwise looking down the positive . The and coin-
cide in this transformation of coordinates. The face normal to the -axis is the oblique face of an innitesimal
wedge at point P. Looking down the positive n-axis, the wedge appears as a triangle as is shown in Fig. 8.4. You
can think of the wedge as having a unit depth in the n-direction. The normal and shear stress components on this
oblique face are denoted by , respectively. Also, the positive stress components acting on the negative z-
and negative s-faces are shown on the wedge element in the gure. If the area of the wedge face is denoted by dA,
then from the geometry of Fig. 8.4, the area of the z-face , and the area on the s-face
. The stresses are converted to differential forces acting on the wedge by multiplying them by the
area of the face on which they act. Force equilibrium in the gives

h
p =

zn
and
sn

zs

z

zs
0

sz

s
0
0 0
n
=
z s n , ,
n-axis n-axis n-axis
z

zs

z
s
z
s

zs

sz
P

s
z

z
dA
z

zs
dA
z

s
z

sz
dA
s

s
dA
s
z-face components s-face components
Fig. 8.4 Wedge element as seen looking down the positive n-axis

z

zs
,
dA
z
dA cos =
dA
s
dA sin =
z-direction
Criteria for Initial Yielding
242 Thin-Walled Structures
Divide this equation by dA, take the limit as the wedge shrinks down to point P, and recognize that
, , and , to get
(8.7)
Force equilibrium in the on the wedge element gives
Again divide by dA and let the wedge element shrink to point P in the limit to get
(8.8)
The wedge element used to determine the stress components on the is shown in Fig. 8.5. For the element
in Fig. 8.5, the areas are related by and . Following procedures similar to those
resulting in eq. (8.7), force equilibrium of this wedge element in the leads to
(8.9)
It is also possible to obtain eq. (8.9) from eq. (8.7) by letting and , since this would rotate
the to the original direction of the . Equilibrium of the wedge element in Fig. 8.5 in the
leads to
(8.10)
which when compared to eq. (8.8) shows that , as would be expected for moment equilibrium about
the n-axis at point P. The stress transformation eqs. (8.7) to (8.10) are written in matrix form as
(8.11)
or in the more compact form as
(8.12)

z
dA
z
dA
z
( ) cos
zs
dA
z
( ) sin
s
dA
s
( ) sin
sz
dA
s
( ) cos 0 =
dA
z
dA cos = dA
s
dA sin =
zs

sz
=

z

z
cos
2

s
sin
2

zs
2 cos sin + + =
s direction

zs
dA
z
dA
z
( ) sin
zs
dA
z
( ) cos
s
dA
s
( ) cos
sz
dA
s
( ) sin + + 0 =

zs

z
sin cos
s
cos sin
zs
cos
2
sin
2
( ) + + =
s-face

sz

s
z
s
P
z

sz

zs

Fig. 8.5 Wedge element with an oblique face


normal to the s-axis
dA
z
dA sin = dA
s
dA cos =
s-direction

s

z
sin
2

s
cos
2

zs
2 cos sin + =
90 +
z

s

z-axis s-axis
z-direction

sz

z
cos sin
s
cos sin
zs
cos
2
sin
2
( ) + + =

zs

sz
=

z

zs
0

sz

s
0
0 0
n

cos sin 0
sin cos 0
0 0 1

z

zs
0

sz

s
0
0 0
n
cos sin 0
sin cos 0
0 0 1
=
T T
T
=
Thin-Walled Structures 243
Principal stresses and maximum shear stress
in which is the matrix of direction cosines as given in the table below. Since the and the
coincide in the rotation of relative to , the normal stress . The trace of stress matrix
in the coordinates is given by
(8.13)
Using eqs. (8.7), (8.9), and , we nd
(8.14)
In matrix notation this is written as
(8.15)
In other words, the trace of the stress matrix is invariant with respect to a coordinate rotation.
8.4 Principal stresses and maximum shear stress
Mohrs circle The graphical representation of the stress transformation equations is called the Mohrs circle.
Mohrs circle is drawn on a plane with the normal stress plotted on the abscissa the shear stress plotted on the
ordinate. To show that the stress transformation equations can be graphed as a circle on this plane, we rst use the
trigonometric identities
to write eqs. (8.7) to (8.9) as
(8.16)
(8.17)
z s n
0
0
0 0 1
T [ ] n-axis n-axis
z cos sin
s sin cos
n
z s n , , ( ) z s n , , ( )
n

n
=
z s n , , ( )
Tr

z

zs
0

sz

s
0
0 0
n

z

s

n
+ + =

n

n
=

z

s

n
+ +
z

s

n
+ + =
Tr

z

zs
0

sz

s
0
0 0
n

Tr

z

zs
0

sz

s
0
0 0
n
=
cos
2
1
2
--- 1 2 cos + ( ) = sin
2
1
2
--- 1 2 cos ( ) = 2 cos sin 2 sin =

1
2
---
z

s
+ ( )
1
2
---
z

s
( ) 2 cos
zs
2 sin + + =

zs

1
2
---
z

s
( ) 2 sin
zs
2 cos + =
Criteria for Initial Yielding
244 Thin-Walled Structures
(8.18)
Dene the average stress and radius R by
(8.19)
Using these denitions we write eqs. (8.16) and (8.17) as
Dene an angle via the trigonometric relations
(8.20)
which is depicted in the triangle at right. Using this denition of we get
Using the trigonometric identities
we nally get the form
(8.21)
(8.22)
Similar manipulations applied to eq. (8.18) results in
(8.23)
The center of the circle is at and its radius is R. A positive shear stress on the positive z-face,
or , is dened positive downward on the ordinate. A positive shear stress on a positive s-face, or , is
dened positive upward on the ordinate. The circle is shown in Fig. 8.6. The locus of points on the circle repre-
sent all possible combinations of normal and shear stresses on the oblique faces through point P. In particular, the
locations of the z-face stresses and s-face stresses are shown on the circle of Fig. 8.6 assuming the stresses are all
positive in value. (Set in eqs. (8.21) and (8.22) to get the z-face stresses.) The diametrically opposite
points of the circle on the normal stress axis represent the oblique faces on which the shear stresses vanish.

1
2
---
z

s
+ ( )
1
2
---
z

s
( ) 2 cos
sz
2 sin =

ave

ave
1
2
---
z

s
+ ( ) = R
1
2
---
z

s
( )
2

zs
2
+ =

z

ave
R
1
2
---
z

s
( )
R
------------------------- 2 cos

zs
R
------ 2 sin + + =

zs
R
1
2
---
z

s
( )
R
------------------------- 2 sin

zs
R
------ 2 cos + =
1
2
---
z

s
( )

zs
R

cos
1
2
---
z

s
( )
R
------------------------- = sin

zs
R
------ =

z

ave
R 2 cos cos 2 sin sin + [ ] [ ] + =

zs
R 2 sin cos 2 cos sin + [ ] =
2 ( ) cos 2 cos cos 2 sin sin + =
2 ( ) sin 2 cos sin 2 sin cos =

z

ave
R 2 ( ) cos + =

zs
R 2 ( ) sin =

s

ave
R 2 ( ) cos =
, ( )
ave
0 , ( ) =

zs

sz
0 =
Thin-Walled Structures 245
Principal stresses and maximum shear stress
Hence, these diametrically opposite points are represent principal stresses and at point P. Principal stress
since no shear stresses act on the . Hence, the principal stresses are given by
(8.24)
The rotation about the n-axis to the principal stress directions occurs when . From eqs. (8.20) we get
(8.25)
where the rotation angle to the principal direction is denoted . The denition for plotting positive shear in the
Mohrs circle plane results in the sense of the rotation on the circle to be the same sense of the rotation about the
n-axis in the physical plane. However, the rotation angle on Mohrs circle is twice the rotation in the physical
plane. If the angle is substituted for in the matrix transformation eq. (8.11) we get
(8.26)
As stated above, the third principal stress .

z
'

ave

zs

zs

zs
'

2 R
z face
z
'
face
s face

sz
s-face

sz

sz

1
2
Fig. 8.6 Mohrs circle

1

2

3

n
= n-faces

1

ave
R + =
2

ave
R =
2 =
2
p
tan

zs
1
2
---
z

s
( )
------------------------- =

z

zs
0

sz

s
0
0 0
n

1
0 0
0
2
0
0 0
3
=
p

3

n

n
= =
Criteria for Initial Yielding
246 Thin-Walled Structures
EXAMPLE 8.1 Maximum shear stress in tensile test
Draw Mohrs circle and determine the maximum shear stress in the uniaxial tension test; i.e, the stress matrix is
given by
(8.27)
Solution First note that the principal stresses are , , and .The average stress
and radius . Mohrs circle is shown in Fig. 8.7 The z-, s-, and n-directions are principal
stress directions. The maximum shear stress is , and the maximum shear stress element is rotated
clockwise about the n-axis as shown in Fig. 8.7. Note that the normal stresses on the planes where the max-
imum shear stress occurs are non-zero and equal to as well.
EXAMPLE 8.2 Mohrs circle for hydrostatic stress state
Draw Mohrs circle for the hydrostatic stress state given by
where p is the pressure.
Solution The average stress and the radius . Hence,
Mohrs circle reduces to a point on the negative normal stress axis as shown
in Fig. 8.8. The shear stress is zero and the normal stress is compressive
with magnitude p on every plane through point P. The principal stresses are

[ ]

z

zs
0

sz

s
0
0 0
n

z
0 0
0 0 0
0 0 0
= =

1

z
=
2
0 =
3
0 =

ave

z
2 = R
z
2 =

z
0

zs

sz
z-face
90
s-face

z
2

z
2

z
2

z
2

z
2
z
s
45
z
s
P
P

z
2

z
2

principal stress element


maximum shear stress element
Fig. 8.7 Mohrs circle and stress elements in the tension test

max

z
2 =
45

z
2
[ ]

z

zs
0

sz

s
0
0 0
n
p 0 0
0 p 0
0 0 p
= =
0
p

zs

sz
Fig. 8.8

ave
p = R 0 =

1

2

3
p = = =
Thin-Walled Structures 247
Principal stresses and maximum shear stress
EXAMPLE 8.3 Principal stresses and maximum shear stress
Determine the principal stresses and the maximum shear stress for the state of given by the matrix
Sketch the principal stress element and the maximum shear stress element.
Solution The average stress is given by , and the radius is
Mohrs circle is shown in Fig. 8.9. The principal stresses are , , and . The
rotation to the principal stress element is
The principal stress element is sketched in Fig. 8.9.
The maximum shear stress on the oblique plane of a wedge element obtained by a rotation about the n-axis,
or x
3
-axis, is 25.5MPa. However, this is not the maximum shear stress at the material point in question. To see
this, begin with the principal stress element and consider a second Mohrs circle for a rotation about the ,
which would have principal stresses and dening its diameter. This second Mohrs circle is centered at
with a radius . The locus of points on this second Mohrs circle represents all possible
combinations of the normal and shear stresses on wedge elements with oblique faces normal to the ,
where the is rotated from the counter-clockwise as view down the positive . See Fig.
8.10. The maximum shear stress on this second Mohrs circle is simply its radius, and it occurs for counter-clock-
[ ]

z

zs
0

sz

s
0
0 0
n
60 12 0
12 15 0
0 0 0
MPa = =

ave
60 15 + ( ) 2 37.5MPa = =
R
1
2
--- 60 15 ( )
2
12
2
+ 25.5MPa = =
63
12
14.04
x
1
x
2
n x
3
,
60
12
12
15
z
s
n
principal stress element
12
37.5 60

zs

0
z-face
R
stresses in MPa
15
R 25.5 =
Fig. 8.9 Mohrs circle and the principal stress element

1
63MPa =
2
12MPa =
3
0 =
2
p
tan
12
1
2
--- 60 15 ( )
-------------------------- 0.53333333 = =
p
14.04 =
x
2
-axis

1

3

1

3
+ ( ) 2
1

3
( ) 2
x
1
-axis
x
1
-axis x
1
-axis x
2
-axis
Criteria for Initial Yielding
248 Thin-Walled Structures
wise rotation of from the stress location on the circle. This rotation corresponds to a counter-clock-
wise rotation of in the physical plane as shown in Fig. 8.10. Equilibrium of the wedge element with an
oblique face normal to the determines the shear and normal stresses on the wedge face, which from
Mohrs circle we know to both have a magnitude of 31.5MPa. It is informative to verify this by summing forces
to zero on the wedge element shown in Fig. 8.10. For convenience, take the area of the oblique face to be unity.
Then the area on the negative is . Note that the negative on the wedge element is
stress free. Summing forces in the to zero we get
Equating forces to zero in the we get
Hence, the stresses on the are
(8.28)
The maximum shear stress for the stress state is 31.5MPa.
63

1
31.5 =

13
31.5 =
x
3
x
1
x
2
x
2
, 12
x
3

x
1

45 45
45
63
12
14.04
x
1
x
2
n x
3
,
principal stress element
maximum shear stress
on a wedge element

0
z-face
stresses in MPa
63 12 31.5
x
1
-face
90
x
3
-face x
2
-face

x
1
-face
Fig. 8.10 Mohrs circle associated with the maximum shear stress
90 x
1
face
45
x
1
-axis
x
1
-face 1 45 cos x
3
face
x
1
-direction

1
1 63 1 45 cos [ ] 45 cos 0 =
x
3
-direction

13
1 63 1 45 cos [ ] 45 sin 0 =
x
1
-face

1
31.5MPa =
13
31.5MPa =
Thin-Walled Structures 249
Octahedral shear stress
8.5 Octahedral shear stress
Since the hydrostatic stress is observed not to effect yielding, we nd the plane at the material point for which
the normal stress is , and then assume that the shear stress acting on this plane, denoted by , is the stress
component responsible for initiating yielding. We start the analysis to nd the shear stress with the principal
stress element at point P, and consider equilibrium of an innitesimal tetrahedron at P. The direction of the unit
normal to the oblique face of the tetrahedron, denoted by , is unknown initially, but we can determine it from
the condition that the normal stress acting on the oblique face is the hydrostatic stress. Let denote the stress
vector acting on the oblique face, and let denote the area of the oblique face of the tetrahedron. See Fig.
8.11. The stress vectors acting on the three triangular faces normal to the principal stress directions are
, , and , where , , and are the areas of the triangular faces of
the tetrahedron normal to the unit vectors , , and along the principal directions, respectively. Force equi-
librium of the tetrahedron gives the vector equation
(8.29)
Divide this equation by the area of the oblique face dA and shrink the tetrahedron to point P in the limit to get
(8.30)
The ratio of the areas of the faces normal to the principal directions to the area of the oblique face is obtained
from geometry of the tetrahedron. Consider two of the edge vectors of the oblique face given by
(8.31)
as are shown in Fig. 8.11. The geometric interpretation of the vector cross product of two position vectors is a
vector normal to the plane of the two vectors with magnitude equal to the area of a parallelogram formed by the

h

h

h
n
S
n ( )
dA
x
1
x
2
x
3

1
i

1
( )dA
1

2
i

2
( )dA
2

3
i

3
( )dA
3
S
n ( )
dA
dx
1
dx
3
dx
2
n
x
1
x
2
x
3
2dAn
dx
3
i

3
dx
1
i

1
+
dx
3
i

3
dx
2
i

2
+
P
P
Fig. 8.11 Tetrahedron element at point P.

1
dA
1
i

1
( )
2
dA
2
i

2
( )
3
dA
3
i

3
( ) dA
1
dA
2
dA
3
i

1
i

2
i

3
S
n ( )
dA
1
dA
1
i

1

2
dA
2
i

2

3
dA
3
i

3
0 =
S
n ( )

1
dA
1
dA
---------i

1

2
dA
2
dA
---------i

2

3
dA
3
dA
---------i

3
+ + =
dx
3
i

3
dx
1
i

1
+ ( ) and dx
3
i

3
dx
2
i
2
+ ( )
Criteria for Initial Yielding
250 Thin-Walled Structures
two position vectors. Using the above two edge vectors we have by this interpretation
(8.32)
Expanding the cross product in this equation gives
The areas of the triangular faces of the tetrahedron normal to the principal directions are ,
, and . Hence, the above equation reduces to
(8.33)
In general, a unit vector is represented as
(8.34)
Component is the cosine of the angle between the normal and the positive with similar interpreta-
tions for components . Comparing eqs. (8.33) and (8.34), it is clear that the direction cosines of the unit
normal also relate the areas of the faces of the tetrahedron as
(8.35)
Substitute the direction cosines of the unit normal from eq. (8.35) into eq. (8.30). Thus, we nd the stress vector
on the oblique face of the tetrahedron is related to the principal stresses by the vector relation
(8.36)
The normal stress component on the oblique face of the tetrahedron is given by the scalar product
. Using eqs. (8.34) and (8.36) this scalar product is
(8.37)
We need to nd the direction of the oblique face, i.e., to nd direction cosines , such that the nor-
mal stress on this face is equal to the hydrostatic stress, . From eq. (8.4) the hydrostatic stress is
(8.38)
Setting and substituting eqs. (8.38) into (8.37), determines the square of the direction cosines as
(8.39)
The result for the direction cosines in this equation implies that there are actually eight oblique faces at the point
2dAn dx
3
i

3
dx
1
i

1
+ ( ) dx
3
i

3
dx
2
i
2
+ ( ) =
2dAn dx
3
dx
2
( ) i

3
i

2
( ) dx
1
dx
3
i

1
i

3
( ) dx
1
dx
2
i

1
i

2
( ) + =
2dAn dx
2
dx
3
i

1
dx
1
dx
3
i

2
dx
1
dx
2
i

3
+ + =
dA
1
dx
2
dx
3
( ) 2 =
dA
2
dx
1
dx
3
( ) 2 = dA
3
dx
1
dx
2
( ) 2 =
n
dA
1
dA
---------i

1
dA
2
dA
---------i

2
dA
3
dA
---------i

3
+ + =
n n
1
i

1
n
2
i

2
n
3
i

3
+ + where n
1
2
n
2
2
n
3
2
+ + 1 = =
n
1
x
1
-axis
n
2
and n
3
n
1
dA
1
dA
--------- = n
2
dA
2
dA
--------- = n
3
dA
3
dA
--------- =
S
n ( )

1
n
1
i

1

2
n
2
i

2

3
n
3
i

3
+ + =

nn
n S
n ( )
=

nn

1
n
1
2

2
n
2
2

3
n
3
2
+ + =
n
1
n
2
and n
3
, ,

h

1
1
3
---
2
1
3
---
3
1
3
--- + + =

nn

h
=
n
1
2
1
3
--- = n
2
2
1
3
--- = n
3
2
1
3
--- =
Thin-Walled Structures 251
Octahedral shear stress
P on which the hydrostatic acts. Unit normal vectors to these eight oblique faces are given the table below. These
eight particular oblique planes at point P are called octahedral planes. The normal stress on the octahedral planes
is the hydrostatic stress.
Now we need to nd the shear stress on the octahedral plane. The shear stress acting on an octahedral plane
is called the octahedral shear stress, and it will have the same value on all eight of the octahedral planes at point
P. To nd its value, consider the octahedral plane in the rst octant of the cartesian space dened by the principal
directions as shown in Fig. 8.12 The stress vector can be written as
(8.40)
where is a unit vector tangent to the octahedral plane and is the octahedral shear stress. The scalar product
of the stress vector with itself yields the square of its magnitude. Using eq. (8.36) we have
(8.41)
While from eq. (8.40) we have
Unit normal
vectors
Components
n
1
n
2
n
3
n
1 1 3 1 3 1 3
n
2 1 3 1 3 1 3
n
3 1 3 1 3 1 3
n
4 1 3 1 3 1 3
n
5 1 3 1 3 1 3
n
6 1 3 1 3 1 3
n
7 1 3 1 3 1 3
n
8 1 3 1 3 1 3

h
n

h
t
S
n ( )
x
1
x
2
x
3
P
n n
1
1
3
------- i

1
i

2
i

3
+ + ( ) = =
Fig. 8.12 Normal stress and shear stress components acting on an octahedral plane.
S
n ( )

h
n
h
t + =
t
h
S
n ( )
S
n ( )

1
n
1
( )
2

2
n
2
( )
2

3
n
3
( )
2
+ + =
Criteria for Initial Yielding
252 Thin-Walled Structures
(8.42)
Equations (8.41) and (8.42) are equated to one another, eq. (8.39) is substituted for the square of the direction
cosines, and then the resulting equation is solved for the octahedral shear stress to get
(8.43)
Now substitute eq. (8.38) for the hydrostatic stress in this equation to get
, or
This last expression can be written as
Hence the octahedral shear stress magnitude is
(8.44)
8.6 Mises criterion for initiation of yielding
In the uniaxial tension test the principal stresses at the initiation of yielding are , .
(Refer to Example 8.1). Substitute these principal stress values into eq. (8.44) to get
(8.45)
For the three-dimensional state of stress, we substitute eq. (8.45) for the octahedral shear stress in eq. (8.44) to
get
(8.46)
Equation (8.44) is the mathematical statement of Mises criterion for the initiation of yielding.
Mises criterion can be visualized in principal stress space, which has orthogonal axis , , and as is
shown in Fig. 8.13 In this space the yield surface given by eq. (8.46) is a right-circular cylinder of radius
whose axis makes equal angles with respect to the , , and coordinate axes. Points within the
cylinder correspond to stress states which have not initiated yielding, while points outside the cylinder corre-
Mises Criterion
Yielding begins in a three-dimensional stress state when the octahedral shear stress is
equal to its value at yield initiation in the uniaxial tension test
S
n ( )
S
n ( )

h
2

h
2
+ =

h
2

1
2
1
3
---
2
2
1
3
---
3
2
1
3
---
h
2
+ + =

h
2

1
2
1
3
---
2
2
1
3
---
3
2
1
3
---
1
1
3
---
2
1
3
---
3
1
3
--- + +


2
+ + =

h
2
2
9
---
1
2

2
2

3
2

2

1

3

2

3
+ + ( ) =

h
2
1
9
---
1

2
( )
2

2

3
( )
2

3

1
( )
2
+ + [ ] =

h
1
3
---
1

2
( )
2

2

3
( )
2

3

1
( )
2
+ + [ ] =

1

y
=
2
0 =
3
0 =

h
2
3
-------
y
=
1
2
---
1

2
( )
2

2

3
( )
2

3

1
( )
2
+ + [ ]
y
=

1

2

3
2 3
y

1

2

3
Thin-Walled Structures 253
Mises criterion for initiation of yielding
spond to stress states that are beyond yielding. Notice that for any stress state on the axis of the cylinder, for
which , no yield is predicted no matter the magnitude of the stresses! This reects the assumption
that hydrostatic stress does not initiate yielding.
In design, it is convenient to dene the Mises stress, denoted , by the denition
(8.47)
Then, the constraint that yielding not occur is
(8.48)
Also, it is convenient to dene the factor of safety, denoted as FS, with respect to yielding as
(8.49)
0
1
2
s1
0
1
2
s2
0
1
2
s3
0
1
2
s3

y
-----

y
-----

y
-----
2
3
---
y
radius
axis of cylinder
Fig. 8.13 Mises yield surface in principal stress space

1

2

3
= =

M
1
2
---
1

2
( )
2

2

3
( )
2

3

1
( )
2
+ + [ ] =

M

y
<
FS

y

M
------- 1 > = for no yielding
Criteria for Initial Yielding
254 Thin-Walled Structures
For the most general state of stress given by
It can be shown after much algebra (which was done using Mathematica) that the Mises stress can be written in
the equivalent form
(8.50)
This equation represents a simplication in computing the Mises stress, since we can omit the step of computing
the principal stresses. For the case of generalized plane stress where , eq. (8.50) reduces to
(8.51)
8.7 Maximum shear-stress criterion
This second empirical criterion assumes that yielding occurs whenever the maximum shear stress attains the
value it has when yielding begins in the tension test. The maximum shear stress is one half of the difference of
the maximum and minimum principal stresses; i.e.,
(8.52)
where
(8.53)
From Example 8.1, the maximum shearing stress in the tensile test at the initiation of yield is
(8.54)
For a three-dimensional state of stress, we substitute eq. (8.52) for the maximum shear stress in eq. (8.50) to get
(8.55)
Equation (8.53) is the mathematical statement for the initiation of yielding by the maximum-shear stress crite-
rion.
A plot of the maximum shear-stress criterion and the Mises criterion for is shown in Fig. 8.14. Cal-
culating the maximum shear stress depends on the numerical order of the principal stresses. In the following
paragraphs this procedure is discussed in terms of plotting the maximum shear-stress criterion shown in Fig.
8.14.
In the rst quadrant of Fig. 8.14, both and are positive. If , then
Mohrs circles appear as shown in the gure to the right, and the maximum shear
stress is . Therefore, maximum shear-stress criterion plots as a vertical line
with for . (Note that stresses normalized by are plotted in

z

zs

zn

sz

s

sn

nz

ns

n
=

M

z
2

s
2

n
2

s

z

n

s

n
3
zs
2
3
zn
2
3
sn
2
+ + + + + =

zn

sn
0 = =

M

z
2

s
2

n
2

s

z

n

s

n
3
zs
2
+ + + =

max

max

min
( ) 2 =

max
Max
1

2

3
, , ( ) =
min
Min
1

2

3
, , ( ) =

max

y
2 =

max

min

y
=

3
0 =

3
0

1

2

1

2
>

1
2

1

y
= 0
2

y
< <
y
Thin-Walled Structures 255
Maximum shear-stress criterion
Fig. 8.14.) If , then the role of and interchange on the Mohrs circles, so the maximum shear
stress is now . Therefore, maximum shear-stress criterion plots as the horizontal line for
.
In the second quadrant of Fig. 8.14 and , so that Mohrs circles
appear as shown in the gure to the right. The maximum shear stress is
. Therefore, the maximum shear-stress criterion, eq. (8.55), gives
in the second quadrant. This is a straight line with a one-to-one slope
intersecting the at . Plotting the maximum shear-stress criterion in quad-
rants three and four in Fig. 8.14 proceed in a similar manner.
The maximum shear-stress envelope in Fig. 8.14 is contained within the Mises envelope. Hence, the maxi-
mum shear stress criterion is conservative from a design perspective, with the largest differences between the
predictions being about 15%. However, in analysis the Mises criterion is easier to implement than the maximum
shear stress criterion. Mises criterion is a single equation, see eq. (8.50), but the maximum shear stress criterion
requires that we compute the principal stresses and nd their numerical order. Also note that the maximum shear
stress acts on four planes at the material point, refer to Fig. 8.10, while the octahedral shear stress acts on eight
planes at the material point. Laboratory tests on thin-walled tubes subject to an axial force, torque, and internal
pressure are often used to study yielding under combined stress states. The experimental data for ductile metal
tubes fall between the maximum shear-stress criterion and the Mises criterion on a plot such as Fig. 8.14, with
the data closer to the Mises prediction (Dowling, 1993, pp. 251 and 252).
In design, the limit state for no yielding by maximum shear-stress criterion is simply
-1 -0.5 0.5 1
-1
-0.5
0.5
1

y
-----

2

y

3
0 =
Mises criterion
Maximum
shear-stress criterion
Fig. 8.14 Criteria for yield initiation in the principal stress plane.
1

2

2

1
0 > >
1

2

2
2
2

y
=
0
1

y
< <

1

3

2
0

1
0 <
2
0 >

2

1
( ) 2

2

y

1
+ =

2
-axis
y
Criteria for Initial Yielding
256 Thin-Walled Structures
(8.56)
and we can dene a factor of safety against yielding as
(8.57)
EXAMPLE 8.4 Factor of safety against initial yielding
Compute the factor of safety against the initiation of yield by Mises criterion and the maximum shear-stress
criterion for 2024-T6 aluminum alloy that has a yield stress of 325 MPa.
Solution From Example 8.3, the principal stresses are , , and , and the
maximum shear stress is . If we calculate the Mises stress from eq. (8.47), we get
If we calculate the Mises stress via eq. (8.51), we get
which is the same value as obtained from eq. (8.47). Hence the factor of safety against yield, eq. (8.49), is
The factor of safety against yield using the maximum shear stress criterion, eq. (8.57), is
In linear structural analysis where the stresses are proportional to the load, the factor of safety means that the load
can be increased by 5.61, in the case of Mises criterion, before the material initiates yielding. In the case of the
maximum shear-stress criterion, the load can be increased by 5.16 before the material begins to yield. The lower
factor of safety predicted by the maximum shear stress criterion illustrates it is slightly conservative with respect
Mises prediction (in this case by about 8%).

max

y
2 <
FS

y
2

max
------------ 1 > = for no yielding

1
63MPa =
2
12MPa =
3
0 =
31.5MPa

M
1
2
--- 63 12 ( )
2
12 0 ( )
2
0 63 ( )
2
+ + ( ) 57.94MPa = =

M
60
2
15
2
60 15 0 0 3 12 ( )
2
+ + 57.94MPa = =
FS
325
57.9396
------------------- 5.61 = =
FS
325 2
31.5
---------------- 5.16 = =
Thin-Walled Structures 257
Maximum shear-stress criterion
EXAMPLE 8.5 Stress responses of a stringer-stiffened, single cell beam.
The thin-walled, prismatic beam of length L shown in Fig. 8.15 is clamped at z = 0. It is subjected to a linearly
distributed load acting through the locus of shear centers, and a torque T applied at z = L.
The cross-sectional contour is an isosceles triangle with branches one and three having the same length b and
the same thickness . The vertical branch, or second branch, has length h and thickness . The beam is stiff-
ened by two longitudinal stringers of cross-sectional area . The overall dimensions are specied as L = 1800
mm, h = 100 mm, and b = 130 mm. The material is 2024-T4 aluminum alloy whose properties are listed in the
table below.
Take , , , and
. For the cross section at z = 0,
d. plot the axial normal stress in MPa versus the contour coordinate s in
mm, and
e. the shear stress in MPa versus the contour coordinate s in mm.
The contour coordinate is related to the branch contour coordinates by
Aluminum alloy 2024-T4
Property Value Units
E, modulus of elasticity 73 GPa
, Poissons ratio 0.33 none
G, Shear modulus 27 GPa
, density 2800 kg/m
3

yield
,

yield strength in tension 325 MPa
p
y
z ( ) p
0
1 z L ( ) =
t
1
s
3
s
1
s
2
t
1
t
2
t
1
h
2
---
h
2
---
b
x
y
p
y
C S.C.
T
Cross-section
A
s
A
s
L
p
y
z ( )
z
T
Fig. 8.15 A stringer-stiffened, cantilevered beam. The contour in the cross section is an isosceles
triangle.
t
2
A
s

zs

z
s
s 0 =
p
0
3.0 N/mm = T 750 N-m = t
1
t
2
1.0 mm = =
A
s
45 mm
2
=

zs
Criteria for Initial Yielding
258 Thin-Walled Structures
Equations for the stress analysis
Geometry
( a)
( b)
Axial normal stress due to bending
( c)
Shear stress tangent to the contour
( d)
Shear ow due to torque
( e)
Shear ow due to transverse the shear force
( f)
( g)
s
s
1
s
2
b +
s
3
b h + +

=
0 s
1
b
0 s
2
h
0 s
3
b
x
y
s
1
s
2
s
3
C
h
c
contour origin s
q
1
q
2 q
3
A
s
A
s
y
1
s
1
( )
h
2
---
s
1
b
---- = 0 s
1
b
y
2
s
2
( )
h
2
--- s
2
= 0 s
2
h
y
3
s
3
( )
h
2
---
h
2
---
s
3
b
---- + = 0 s
3
b
I
xx
y
1
2
s
1
( )t
1
s
1
d
0
b

y
2
2
s
2
( )t
2
s
2
d
0
h

y
3
2
s
3
( )t
1
s
3
d
0
b

h
2
---


2
A
s
h
2
---


2
A
s
+ + + + =

z
z s , ( )
M
x
z ( )
I
xx
--------------- y s ( ) =

zs
z s , ( )
q z s , ( )
t s ( )
--------------- = q z s , ( ) q
b
z s , ( ) q
t
z ( ) + =
q
t
T
2
-------- = enclosed area of the cell
1
2
---hc = =
q
b1
z s
1
, ( ) q
0
V
y
z ( )
I
xx
------------- y
1
s
1
( )t
1
s
1
d
0
s
1

=
q
b2
z s
2
, ( ) q
b1
b ( )
V
y
z ( )
I
xx
-------------
h
2
---


A
s

V
y
z ( )
I
xx
------------- y
2
s
2
( )t
2
s
2
d
0
s
2

upper stringer
Thin-Walled Structures 259
Maximum shear-stress criterion
( h)
To determine the shear ow at the contour origin, , due to the transverse shear force acting through the
shear center, we invoke the condition of no twist. That is,
( i)
or
( j)
Equilibrium
( k)
( l)
To locate the centroid (the point labeled C)
( m)
( n)
( o)
To locate the shear center (the point labeled S.C.)
From the solution of equations (f) to (j), we can nd the shear ow in branch two due to the shear force
only. Now we use torque equivalence to locate the S.C. Sum torques about the apex to get
( p)
q
b3
z s
3
, ( ) q
b2
h ( )
V
y
z ( )
I
xx
-------------
h
2
---


A
s

V
y
z ( )
I
xx
------------- y
3
s
3
( )t
1
s
3
d
0
s
3

lower stringer
q
0
V
y
z d
d
z q s ( ) s d
Gt s ( )
----------------

0 = =
q s ( )
t s ( )
----------ds

0 =
q
b1
s
1
( )
t
1
----------------- s
1
d
0
b

q
b2
s
2
( )
t
2
----------------- s
2
d
0
h

q
b3
s
3
( )
t
1
----------------- s
3
d
0
b

+ + 0 =
V
y
M
x
p
y
dV
y
dz
--------- p
y
z ( ) = and V
y
L ( ) 0 =


V
y
z ( ) p
y
z ( ) z d
z
L

=
dM
x
dz
----------- V
y
z ( ) = and M
x
L ( ) 0 =


M
x
z ( ) V
y
z ( ) z d
z
L

=
x
y
s
1
s
2
s
3
C
h
c
A
s
A
s
x
c
A area 2bt
1
ht
2
2A
s
+ + = =
Q
y
x
1
s
1
( )t
1
s
1
d
0
b

x
3
s
3
( )t
1
s
3
d
0
b

+ = x
c
Q
y
A =
x
1
s
1
( ) c 1 s
1
b ( ) = x
3
s
3
( ) c s
3
b ( ) =
V
y
c q
b2
s
2
( ) s
2
d
0
h

c x
sc
( )V
y
=
Criteria for Initial Yielding
260 Thin-Walled Structures
Equation (p) yields a relation to nd , or the location of the shear center. The location of the shear center is
independent of the magnitude of the shear force.
Results The shear force and bending moment attain maximum magnitudes at the root. A the root
and . The torque at the root section is
. Note that 1 N/mm
2
equals 1 MPa
a) The axial normal stress is plotted as a function of the contour coordinate in graph below.
b) The shear stress is plotted as a function of the contour coordinate in the following graph.
x
sc
x
y
S.C.
c
x
sc
V
y
c x
sc

x
y
s
2
h
c
q
b2
s
2
( )
V
y
0 ( ) 2700 N = M
x
0 ( ) 1.62
6
10 ( ) N-mm =
T 0 ( ) 750
3
10 N-mm =
50 100 150 200 250 300 350
s,mm
-150
-100
-50
50
100
150
sz,Nmm
2
Thin-Walled Structures 261
Maximum shear-stress criterion


EXAMPLE 8.6 Minimum weight design of the beam in Example 8.5 subject to a constraint on initial
yielding
Consider the design for minimum weight of the aluminum alloy beam in Example 8.5, which is shown in Fig.
8.15. The design is constrained by material yielding with the factor of safety specied as 1.5. Use the material
data for 2024-T4 aluminium as listed in the Example 8.5 problem statement, Mises yield criterion, and take the
value of the local acceleration due to gravity as 9.81 m/s
2
. The specied dimensions of the beam are
, , and . Take the value of the applied loads as and
. (Note the value of the torque is changed with respect to its value in Example 8.5.)
The objective is to minimize the weight subject to no yielding given the loads and . That is, what are
the thicknesses and and the stringers cross-sectional area for minimum weight? Parameters , and
are called design variables. This is a problem in constrained optimization, which is stated mathematically as
where is the objective function, or weight in this problem, and are constraint func-
tions. For design against yielding the constraint functions are dened as
50 100 150 200 250 300 350
s,mm
10
20
30
40
50
60
70
80
t zs,Nmm
2
L 1800 mm = h 100 mm = b 130 mm = p
0
3 N/mm =
T 250
3
10 N-mm =
p
0
T
t
1
t
2
A
s
t
1
t
2
A
s
minimize W t
1
t
2
A
s
, , ( )
such that g
i
t
1
t
2
A
s
, , ( ) 0 >
W t
1
t
2
A
s
, , ( ) g
i
t
1
t
2
A
s
, , ( )
g
i

all

M
( )
i

all
---------------------------- =
Criteria for Initial Yielding
262 Thin-Walled Structures
where the allowable stress, , is dened as
and the Mises stress is . These par-
ticular constraint functions are called static margins, and
positive values indicate the degree of safety against exceed-
ing the allowable stress. Due to symmetry about the x-axis
you only need to calculate the margin of safety for yielding
at four points in the cross section at the root as indicated in
Fig. 8.16. That is, compute the margins of safety at the
points labeled 1, 2, 3, and 4 in Fig. 8.16. In addition, the
shear force and bending moment attain their largest magni-
tude simultaneously at z = 0, so the four constraints are evaluated at z = 0.
The intent of this exercise is to study the inuence of the stringers on the design for minimum weight. For
each stringer area given in the table below, determine the values of thicknesses t
1
and t
2
for minimum weight.
List these values along with the weight, and the four margins of safety in the table.
To calibrate the computations, the beam weight is 57.87 N and the margins of safety at points 1 to 4 are
50.3716, 1.92285, 1.92539, and 8.56734, respectively, for the design variable values of t
1
= 2.84 mm, t
2
= 3.42
mm, and A
s
= 45 mm
2
.
Inuence of the stringer area on the minimum weight designs
Stringer
area A
s
in
mm
2
Beam
weight in
N
Thicknesses in mm Margins of safety, dimensionless
t
1
t
2
point 1 point 2 point 2 point 4
50
60
70
80
90
100
x
y
C
point 1, s
1
0 =
point 2, s
1
b =
point 3, s
2
0 =
point 4, s
2
h 2 =
Fig. 8.16 Critical points for yield evaluation
in the cross section at the root

all

all

yield
F.S. =
M
Thin-Walled Structures 263
Maximum shear-stress criterion
Results s

0.2 0.3 0.4 0.5


0.3
0.4
0.5
0.6
0.7
0.8
As Wt. t1 t2 M1 M2 M3 M4
50 14.3955 0.437253 0.774719 3.2 9.4 10
- 10
0.0029 1.1
60 13.5688 0.342651 0.653471 2.1 - 1.7 10
- 12
0.0008 0.82
70 13.3053 0.279341 0.564789 1.4 0.001 - 5.810
- 8
0.57
80 13.4942 0.239829 0.505728 0.97 0.0024 - 2.410
- 11
0.41
90 13.9704 0.214902 0.466853 0.72 0.0034 3.1 10
- 12
0.3
100 14.6189 0.198474 0.440723 0.57 0.0039 - 1. 10
- 10
0.23
14 N
16 N
18 N
A
s
100 mm
2
=
t
1
, mm
t
2
, mm
Minimum static margin 0 =
least weight design
constant weight lines
Design plane for
feasible designs
infeasible designs
Criteria for Initial Yielding
264 Thin-Walled Structures
8.8 References
Ashby, M.F., 1992, Materials Selection in Mechanical Design, Butterworth-Heinemann, Ltd., Oxford.
Dowling, N.E., 1993, Mechanical Behavior of Materials, Prentice-Hall, Inc., Englewood Cliffs, New Jersey.

Thin-Walled Structures

251

CHAPTER 9

Buckling

Buckling of a structure means


failure due to excessive displacements (loss of structural stiffness), and/or


loss of stability of an equilibrium conguration of the structure

Stability of equilibrium

means that the response of the structure due to a small disturbance from its equilib-
rium conguration remains small; the smaller the disturbance the smaller the resulting magnitude of the displace-
ment in the response. If a small disturbance causes large displacement, perhaps even theoretically innite, then
the equilibrium state is unstable. Practical structures are stable at no load. Now consider increasing the load
slowly. We are interested in the value of the load, called the

critical load

, at which buckling occurs. That is, we
are interested in when a a sequence of equilibrium stable states as a function of the load, one state for each value
of the load, ceases to be stable.
If buckling occurs before the elastic limit of the material, which is roughly the yield stress of the material,
then it is called

elastic buckling

. If buckling occurs beyond the elastic limit, it is called

inelastic buckling

, or plas-
tic buckling if the material exhibits plasticity during buckling (mainly metals). Most thin-walled structural com-
ponents buckle in compression below the elastic limit. Therefore, buckling determines the limit state in
compression rather than material yielding. In fact, about 50% of an airplane structure is designed based on buck-
ling constraints.

9.1 One-degree of freedom model

To illustrate the physical nature of buckling as a stability problem and failure by excessive displacements, it is
instructive to analyze the response of a simple structural model to a compressive force. This model is shown in
Fig. 9.1 and has one coordinate

,

, to describe the conguration of the model under the deadweight
load

P

. The model consists of a rigid rod of length l, connected by smooth hinge to a rigid base. The rod can
rotate about the hinge but it is restrained by a linear elastic torsional spring of stiffness K (dimensional units of F-
L/ radian). The spring is unstretched at


= 0. Neglect the weight of the rod with respect to the applied load P.
< <

Buckling

252

Thin-Walled Structures

From the free body diagram of the rod shown in Fig. 9.1, the equation of motion for rotation about the xed
hinge is

(9.1)

where

I

0

is the moment of inertia of the rod about the xed point and

t

is time.

9.1.1 Static equilibrium

Consider equilibrium states under the static, downward load

P

which are characterized by the angle


being
independent of time

t

. Hence, the inertia term in eq. (9.1) vanishes and we have

(9.2)

The solutions to eq. (9.2) are

(9.3)

and

(9.4)

Recall from the calculus using lHpitals rule that the limit of the
indeterminate form as is one. The two equilib-
rium paths are plotted in the load-deection diagram shown in Fig.
9.2. Equilibrium path

P

1

coincides with the load axis in the plot and
is called the primary equilibrium path, or the trivial equilibrium
path. Equilibrium path

P

2

is called the secondary path and we note
it is symmetric about


= 0. The two equilibrium paths intersect at
(


,

P

) = (0,

K/l

). This intersection of the two paths is called a bifurca-
tion point. At no load the rod is vertical and this corresponds to the
origin in the load-deection diagram. As the load

P

is slowly
increased from zero the rod remains vertical (


= 0), and at

P = K/l


adjacent equilibrium states exists on the secondary path. The exist-
ence of adjacent equilibrium states in the vicinity of the primary
l
K

K
O
x
O
y
P
P
initial deflected
FBD
Fig. 9.1 One degree of freedom structural model
l
Pl sin K I
0
t
2
2
d
d
= t ( ) = t 0 >
Pl sin K 0 = <
P
1
: 0 for any P =
P
2
: P
K
l
----
\ )
[

sin
----------- =

0

P
K l
----------
1
2
P
1
P
1
P
2
P
2
Fig. 9.2 Equilibrium paths.
sin ( ) 0

Thin-Walled Structures

253

One-degree of freedom model

equilibrium path has been noted by investigators of structural stability as the onset of buckling. Hence, buckling
is characterized by the bifurcation point on the load-deection diagram. For this reason, the term bifurcation
buckling is used to describe this condition. As we will show later, the rod will not remain vertical for loads

P >
K/l

if there are innitesimal disturbances present (there always are), but will rotate either to the left or right
depending on type of innitesimal disturbance. We note that the magnitude of the angle


becomes large as the
load is increased from

K/l

on the secondary path. The load at the bifurcation point is called the critical load and is
denoted as . Thus,

(9.5)

Small


analysis

(9.6)

Consider the small angles of rotation such that for


measured in radians. Equilibrium eq. (9.2)
becomes

(9.7)

And the solutions of this equation are
, and

(9.8)
(9.9)

These solutions are shown in the load-deection plane in Fig. 9.3. The equilibrium
path coincides with path , but path is not a good approximation to path
unless


is very small. However, the bifurcation point is the same as obtained in
the large


-analysis. Hence, the critical load from the small


-analysis is the same as
obtained in eq. (9.5) from the large


-analysis.

9.1.2 Stability analysis

Let the rotation angle


(9.10)

where is independent of time and satises the equilibrium eq. (9.2); i.e.,

(9.11)

Consider the additional rotation angle to be small in magnitude but a function
of time. Thus, we are considering small oscillations about an equilibrium state
as shown in Fig. 9.4. Substitute eq. (9.10) for q in the equation of motion,
eq. (9.1), to get

(9.12)

where the dots denote derivatives with respect to time; e.g., . Using the
trigonometric identity for the sine of the sum of two angles and performing some
minor rearrangements the last equation becomes
P
cr
P
cr
K l =
sin
Pl K 0 =
P
1
: 0 for any P =
P
2
: P K l = for any small
P
K l
----------
0
1
P
1

P
2

Fig. 9.3 Small analysis


P
1
P
1
P
2
P
2
t ( )
0
t ( ) + =

0
Pl
0
sin K
0
0 =
l
P

0
t ( )
t ( )
Fig. 9.4 Rotations in
the stability analysis
t ( )
P
0
, ( )
I
0
K
0
+ ( ) Pl
0
+ ( ) sin + 0 =

t
2
2
d
d
=
Buckling
254 Thin-Walled Structures
(9.13)
Now expand the trigonometric functions of angle in a Taylor Series about to get
(9.14)
in which means terms of order and higher. Arrange eq. (9.14) in powers of to get
(9.15)
Note that "coefcient" of the term vanishes because of the equilibrium condition given by eq. (9.11).
For very small additional rotation angles about the equilibrium conguration, eq. (9.15) is approxi-
mated by
(9.16)
where
(9.17)
The solution of the second order differential equation, eq. (9.16), for is
(9.18)
in which constants and are determined by initial conditions for and . The solution given by
eq. (9.18) is a harmonic oscillation about the equilibrium conguration and is interpreted as the natural fre-
quency in radians per second. Initial conditions and are considered to be very small but arbitrary to
simulate and arbitrary small initial disturbance. The smaller the initial disturbance, the smaller the maximum
amplitude of the oscillation in . Thus, is a condition for a stable equilibrium conguration with respect
to innitesimal disturbances.
The solution of the second order differential equation, eq. (9.16), for is
(9.19)
For arbitrary initial conditions, the term with the positive exponent in the dominates the solution. This corre-
sponds to large values of the no matter how small the initial disturbance. Hence, is a condition of
unstable equilibrium conguration with respect to innitesimal disturbances. The dynamic criterion for structural
stability is
Dynamic criterion for stability of an equilibrium state
The equilibrium state is stable if
The equilibrium state is critical if
The equilibrium state is unstable if
I
0
K
0
K Pl
0
sin cos
0
cos sin + [ ] + + 0 =
0 =
I
0
K
0
K Pl
0
sin 1
1
2
---
2
O
4
( ) + Pl
0
cos
1
6
---
3
O
5
( ) + + + 0 =
O
n
( )
n

I
0
K
0
Pl
0
sin ( ) K Pl
0
cos ( )
Pl
2
-----
0
sin
\ )
[

2
Pl
6
-----
0
cos
\ )
[

3
O
4
( ) + + + + + 0 =
||||||
0 =

0
t ( )
I
0
K Pl
0
cos ( ) + 0 = or
2
+ 0 =

2
K Pl
0
cos ( ) I
0
=

2
0 >
t ( ) A
1
t ( ) sin A
2
t ( ) cos + =
2
0 >
A
1
A
2
0 ( ) 0 ( )

0 ( ) 0 ( )

2
0 >

2
0 <
t ( ) A
1
e
t
A
2
e
t
+ =
2

2
K Pl
0
cos ( ) I
0
= =

2
0 <

2
0 >

2
0 =

2
0 <
Thin-Walled Structures 255
Perfect Columns
On the primary equilibrium path given by eq. (9.3), we have from eq. (9.17) that
(9.20)
Thus, equilibrium congurations are stable if , critical if , and unstable if . The pri-
mary equilibrium path ceases to be stable at , and is the buckling load.
9.2 Perfect Columns
Consider a perfectly straight, uniform column of length L and cross-sectional area A subjected to a centric end
load P as shown in Fig. 9.5. The column is long relative to its largest cross-sectional dimension, and the column
consists of a homogeneous, linear elastic material whose modulus of elasticity is denoted by E. The equilibrium
conguration of this column is pure compression. Let N(z) denote the internal axial force. From equilibrium of
a differential element shown below we have , and from Hookes law where is axial
normal strain. The axial normal strain is related to the axial displacement by as is shown in Fig.
9.6. The boundary conditions for the column are and . Hence, the internal axial load is
uniform and compressive along the length of the column and equal in magnitude to the applied load P. Summa-
rizing the equilibrium solution we have
(9.21)
P
1

2
K Pl ( ) I
0
= on P
1
P K l < P K l = P K l >
P P
cr
= P
cr
z,w
y,v
L
P
Fig. 9.5 A straight column subjected to a centric, compressive axial force.
dN dz 0 = N EA
z
=
z

z
dw dz =
z
dz
z
y
w z dz + ( )
w z ( )
dz w z dx + ( ) w z ( ) + 1
z d
dw
+
\ )
[
dz
N N dN +
Fig. 9.6 An element of the column in the pre-buckling equilibrium state
w 0 ( ) 0 = N L ( ) P =
N P = w z ( )
Pz
EA
------- = v z ( ) 0 = 0 z L
Buckling
256 Thin-Walled Structures
The end shortening under the compressive load is , and this is plotted on the load-end shortening plot
shown in Fig. 9.7. The equilibrium conguration of pure compression of the perfect column is called the trivial
equilibrium state. Note that in the trivial equilibrium state the lateral displacement of the column, is zero
for all values of the compressive load P. Researchers in structural stability recognized from experience that buck-
ling of the column is associated with the appearance of second, non-trivial, equilibrium conguration at the buck-
ling load. This observation is the basis of the adjacent equilibrium method of stability analysis. The question
characterizing the method of adjacent equilibrium is
What is the value of the load for which the perfect system admits non-trivial equilibrium congura-
tions?
To answer this question we consider equilibrium of a slightly deected element of the column at the same
value of the external load P. The free body diagram of this element is shown in Fig. 9.8. The displacement due to
buckling is denoted by v
1
(z), and all quantities due to buckling are labeled with the subscript 1. Vertical force
equilibrium gives
(9.22)
where is the y-direction shear force due to buckling. Moment equilibrium about the x-axis gives
where is the bending moment due to buckling. In general, the axial strain in the equilibrium conguration
is very small in magnitude compared to unity and is then neglected with respect to unity in this equation.
w L ( )
w L ( )
0
EA
L
-------
1
P
Fig. 9.7 Load-end shortening
plot in pre-buckling
v z ( )
1
dw
dz
------- +
\ )
[
dz
dv
1
dz
--------dz

x1
v
1
v
1
dv
1
dz
-------- dz ( ) +
z
y M
x1
M
x1
dM
x1
+
V
y1
V
y1
dV
y1
+ P
P
Fig. 9.8 Free body diagram of an element of the column in the buckled state.
dV
y1
dz
------------ 0 =
V
y1
dM
x1
dz
------------- 1
dw
dz
------- +
\ )
[
V
y1

dv
1
dz
--------
\ )
[
P + 0 =
M
x1
dw dz
Thin-Walled Structures 257
Perfect Columns
Then, the moment equation becomes
(9.23)
Neglecting with respect to unity, and assuming the rotation due to buckling is small, implies that the
rotation is given by . A very important term in eq. (9.23) is the contribution of the axial
compressive load P through the buckling displacement v
1
to moment equilibrium. This term that couples the
axial compressive load in equilibrium state to the buckling displacement arises only because we took equilibrium
on the slightly deected column element. Hookes law for the bending moment is
(9.24)
where I
xx
is the second area moment of the cross section about the x-axis. (We assume the cross section is sym-
metric about either the x-axis or y-axis, or that these axes are principal axes of the cross section if no symmetry is
present. In design we use the minimum second are moment of the cross section). If we take the derivative of eq.
(9.23), use eq. (9.22), and then substitute eq. (9.24) for the bending moment due to buckling we get
(9.25)
This is the governing fourth order, ordinary differential equation for the buckling displacement v
1
(z). For conve-
nience in writing, we will drop the subscripts on the second area moment in the following developments. Also
note that in the buckling theory the vertical shear force is determined in terms of the buckling displacement v
1
by
substituting eq. (9.24) into eq. (9.23) to get
(9.26)
To determine the buckling displacement v
1
we need boundary conditions at z = 0 and z =L in addition to the
o.d.e. given by eq. (9.25). There are four standard boundary conditions. These are

dM
x1
dz
------------- V
y1

dv
1
dz
--------
\ )
[
P + 0 =
dw dz
x1

x1
z ( ) dv
1
dz ( ) =
M
x1
EI
xx
z
2
2
d
d v
1

\ )
| j
[
=
z
2
2
d
d
EI
xx
z
2
2
d
d v
1

\ )
| j
[
z
2
2
d
d v
1

\ )
| j
[
P + 0 =
V
y1
z d
d
EI
z
2
2
d
d v
1
\ )
| j
[

dv
1
dz
--------P =
z
L
P
A. Pinned-pinned
v
1
0 ( ) 0 = v
1
L ( ) 0 =
M
x1
0 ( ) 0 = M
x1
L ( ) 0 =
z
L
P
B. Clamped-free
v
1
0 ( ) 0 = M
x1
L ( ) 0 =

x1
0 ( ) 0 = V
y1
L ( ) 0 =
Buckling
258 Thin-Walled Structures
One solution to the o.d.e., eq. (9.25), subject to boundary conditions A-D is for all values of the
load P. This is the trivial solution. Are there any other solutions? Can we get them? The answer is yes to both
questions if EI = constant. For EI = constant, eq. (9.25) becomes
or
(9.27)
where
(9.28)
The general solution of eq. (9.27) for is
(9.29)
where A
1
, A
2
, A
3
, and A
4
are arbitrary constants to be determined by boundary conditions.
EXAMPLE 9.1 Critical load for clamped-free boundary conditions (B)
Consider the clamped-free boundary conditions; i.e. b.c.s (B) above. Determine the buckling load for
which the perfect column ceases to be stable.
Solution The boundary conditions in this case become
where the primes denote derivatives with respect to z. Taking derivatives of eq. (9.29) we have
z
L
P
C. Clamped-clamped
v
1
0 ( ) 0 =

x1
0 ( ) 0 =
v
1
L ( ) 0 =

x1
L ( ) 0 =
z
L
P
D. Clamped-pinned
v
1
0 ( ) 0 =

x1
0 ( ) 0 =
v
1
L ( ) 0 =
M
x1
L ( ) 0 =
v
1
z ( ) 0 =
EI
z
4
4
d
d v
1
P
z
2
2
d
d v
1
+ 0 =
z
4
4
d
d v
1
k
2
z
2
2
d
d v
1
+ 0 = 0 z L < <
k
2
P
EI
------ =
k
2
0 >
v
1
z ( ) A
1
kz ( ) sin A
2
kz ( ) cos A
3
z A
4
+ + + =
P
cr
v
1
0 ( ) 0 = v
1
0 ( ) 0 = EI v
1
L ( ) 0 = v
1
k
2
v
1
+ [ ]
z L =
0 =
Thin-Walled Structures 259
Perfect Columns
Substitute these solutions into the four boundary conditions to get
(9.30)
A non-trivial solution for A
1
to A
4
requires the determinate of coefcients to vanish
(9.31)
After expanding this determinate we get
(9.32)
which gives n-values, n = 1, 2, 3,..., of kL; or
(9.33)
For , the fourth row of matrix eq. (9.30) gives ; using this result in the sec-
ond row of matrix eq. (9.30) gives ; the rst row of matrix eq. (9.30) gives . Note that the
third row of matrix eq. (9.30) is identically satised. So eq. (9.33) implies
, (9.34)
where P
n
are the buckling loads. Equation (9.32) is called the characteristic equation, and the roots of this equa-
tion determine the buckling loads. For each value of n we have an associated buckling mode (A
1
= A
3
= 0, A
2
=
A
4
)
(9.35)
where coefcient A
4
is arbitrary. The rst three buckling modes are shown in Fig. 9.9. Note that the amplitude
of the buckling mode is not known. However, we can plot its shape. The critical load, denoted by , is the low-
est buckling load. That is
v
1
A
1
kz ( ) sin A
2
kz ( ) cos A
3
z A
4
+ + + =
v
1
A
1
k kz ( ) cos A
2
k kz ( ) sin A
3
+ =
v
1
A
1
k
2
kz ( ) sin A
2
k
2
kz ( ) cos =
v
1
A
1
k
3
kz ( ) cos A
2
k
3
kz ( ) sin + =
0 1 0 1
k 0 1 0
k
2
kL ( ) sin k
2
kL ( ) cos 0 0
0 0 k
2
0
A
1
A
2
A
3
A
4
0 =
det
0 1 0 1
k 0 1 0
k
2
kL ( ) sin k
2
kL ( ) cos 0 0
0 0 k
2
0
0 =
k
5
kL ( ) cos 0 =
k
n
2n 1 ( )
L
--------------------

2
---
P
n
EI
------ = = n 1 2 3 , , , =
k
n
L 2n 1 ( ) 2 ( ) = A
3
0 =
A
1
0 = A
2
A
4
=
P
n
2n 1 ( )

2
---
2
EI
L
2
------ = n 1 2 3 , , , =
v
1n
z ( ) A
4
1 k
n
z ( ) cos [ ] =
P
cr
P
cr
P
1

2
4
-----
EI
L
2
------ = =
Buckling
260 Thin-Walled Structures
Remember that in design we use the minimum EI for the cross section.
0.2 0.4 0.6 0.8 1
z/L
0.5
1
1.5
2
v1/A4
0.2 0.4 0.6 0.8 1
z/L
0.5
1
1.5
2
v1/A4
0.2 0.4 0.6 0.8 1
z/L
0.2
0.4
0.6
0.8
1
v1/A4
P
1

2
4
-----
EI
L
2
------ =
P
2
9
2
4
---------
EI
L
2
------ =
P
3
25
2
4
------------
EI
L
2
------ =
n = 1
n = 2
n = 3
Fig. 9.9 First three buckling modes for the clamped-free column.
Thin-Walled Structures 261
Imperfect columns
The critical loads for boundary conditions A through D and for EI = constant are given in Fig. 9.10 .
9.3 Imperfect columns
9.3.1 Eccentric load
Consider a uniform (EI = constant), pinned-pinned column subjected to an eccentric axial load P. Let e denote
the perpendicular distance between the line of action of load P and the z-axis. This situation is statically equiva-
lent to a centric axial load P and a moment of magnitude eP applied to the ends of the column. See Fig. 9.11.
Hence, the eccentric axial load will simultaneously subject the column to compression and bending in the equi-
librium state. The analysis for the equilibrium response of the column in compression (axial force N and axial
displacement w(z)) is identical to the case of the perfect column, since the z-axis passes through the centroid of
each cross section (decoupling the axial compression from bending in the material law). The equilibrium
response of the column in bending, which includes the inuence of axial compression on bending, is determined
by the same analysis that led to eqs. (9.27) to (9.29), except that we drop the subscript 1 on the lateral displace-
ment, since in the eccentric load case the lateral displacement refers to an equilibrium state and not to a buckling
mode. The differential equation and boundary conditions for equilibrium displacement v(z) are
L
D. Clamped-pinned
P
cr
2.046
2
EI
L
2
------ 4.49 ( )
2
EI
L
2
------ = =
L
C. Clamped-clamped
P
cr
4
2
EI
L
2
------ =
L
B. Clamped-free
P
cr

2
4
-----
EI
L
2
------ =
L
A. Pinned-pinned
P
cr

2
EI
L
2
------ =
Fig. 9.10 Buckling loads for the standard boundary conditions A to D.
Buckling
262 Thin-Walled Structures
(9.36)
(9.37)
where k
2
is as given in (9.28). Note that the boundary conditions, eqs. (9.37), are inhomogeneous. Thus, eqs.
(9.36) and (9.37) do not have the trivial solution v(z) = 0 for all values of the load P. Using the general solution
form given by (9.29) for the solution of differential equation (9.36), subject to the boundary conditions (9.37), we
nd
(9.38)
By a trigonometric identity, . Let denote the midspan displacement; i.e., = v(L/2).
From eq. (9.38) we get
The factor kL/2 can be written in terms of the eccentric axial load P and the critical load P
cr
for the pinned-
pinned uniform column subjected to centric load as
.
Thus, the center deection of the eccentrically loaded column becomes
z
4
4
d
d v
k
2
z
2
2
d
d v
+ 0 0 z L < < =
v 0 ( ) 0 = EI
z
2
2
d
d v
0 ( ) eP = v L ( ) 0 = EI
z
2
2
d
d v
L ( ) eP =
L
e
e
z,w
y,v
EI
P P
eP
P
eP
P
Fig. 9.11 Column subjected to an eccentric axial load
v z ( ) e 1 kz ( ) cos
kL
2
------
\ )
[
tan kz sin + + =
kL
2
------
\ )
[
tan
1 kL ( ) cos
kL ( ) sin
----------------------------- =
e 1
kL
2
------
\ )
[
cos
kL
2
------
\ )
[
tan
kL
2
------
\ )
[
sin + + e
1
kL
2
------
\ )
[
cos
--------------------- 1 = =
kL
2
------
1
2
---
P
EI
------
\ )
[
L

2
EI
L
2
P
cr
-------------
| |

| |

2
---
P
P
cr
------- = =
Thin-Walled Structures 263
Imperfect columns
(9.39)
The load-displacement response is shown in Fig. 9.12. Note that as for . That is, no
matter the magnitude of the eccentricity as the value of the center deection gets very large.
9.3.2 Geometric imperfection
Consider a uniform, pinned-pinned column that is slightly crooked under no load. The initial shape under no load
is described by the function . The column is subjected to a centric, axial compressive load P. The lateral
displacement of the column is denoted by , so that when P = 0. Also, the bending moment
in the column is zero under no load. Thus, we write the material law for bending as
(9.40)
Vertical force equilibrium of the deected column leads to a differential equation similar to eq. (9.22) except that
we drop the 1 subscript since it is the conguration of the column is one of equilibrium and not a buckling
mode. Similarly, moment equilibrium of the imperfect column leads to a differential equation similar to eq.
(9.23) with the subscript 1 dropped. Combining these differential equations of vertical force equilibrium and
moment equilibrium via elimination of the vertical shear force gives
(9.41)
The pinned-pinned boundary conditions are
(9.42)
e
1

2
---
P
P
cr
-------
\ )
[
cos
------------------------------ 1 =
P/P
cr
0

1
increasing e
e = 0
P
cr

2
EI
L
2
------ =
Fig. 9.12 Load-deection curves for an eccentrically loaded column
P P
cr
e 0
P P
cr

v
0
z ( )
v z ( ) v z ( ) v
0
z ( ) =
M
x
EI
z
2
2
d
d v
z
2
2
d
d v
0

\ )
| j
[
=
z
2
2
d
d M
x
z
2
2
d
d v

\ )
| j
[
P + 0 =
v 0 ( ) 0 = M
x
0 ( ) 0 = v L ( ) 0 = M
x
L ( ) 0 =
Buckling
264 Thin-Walled Structures
Consider the imperfection shape , where denotes the amplitude at midspan of the
slightly crooked column. Substitute eq. (9.40) into eq. (9.41) to eliminate the moment to get
(9.43)
where is given by eq. (9.28). The boundary conditions, eqs. (9.42), for this imperfection shape lead to
(9.44)
The solution of the differential equation (9.43) subject to boundary conditions (9.44) is
(9.45)
It is convenient to measure the deection of the imperfect column under load with respect to its original unloaded
state. That is, let dene the additional displacement at midspan by . Hence,
(9.46)
The load-displacement response is sketched in Fig. 9.13.
Note that as for . That is, for a non-
zero value of the imperfection amplitude, the displacement
gets very large as the axial force approaches the buckling
load of the perfect column. Also, the imperfect column
deects in the direction of imperfection; e.g., if , then
.
Collectively the eccentric load and the geometric shape
imperfection are called imperfections. All real columns are
imperfect. Even for a well manufactured column whose geo-
metric imperfections are small and with the load eccentricity
small, the displacements become excessive as the axial com-
pressive force P approaches the critical load of the per-
fect column. Hence, the critical load determined from the
analysis of the perfect column is meaningful in practice.
v
0
z ( ) a
1
z
L
-----
\ )
[
sin = a
1
z
4
4
d
d v
k
2
z
2
2
d
d v
+
z
2
2
d
d v
0
a
1

L
---
\ )
[
2
z
L
-----
\ )
[
sin = =
k
2
v
z
2
2
d
d v
0 = = at z = 0 and z = L
v z ( )
a
1
1
kL

------
\ )
[
2

-----------------------
z
L
-----
\ )
[
sin = 0 z L
v L 2 ( ) v
0
L 2 ( ) =
a
1
P
P
cr
-------
\ )
[
1
P
P
cr
-------
\ )
[

---------------------- =
0
P P
cr

1.0

P
cr

2
EI
L
2
------ =
increasing a
1
Fig. 9.13 Load-deection response plots
for geometrically imperfect columns
P P
cr
a
1
0
a
1
0 >
0 >
P
cr
Thin-Walled Structures 265
Column Design Curve
9.4 Column Design Curve
Consider the pinned-pinned uniform column whose critical load is given by . Let A denote the
cross-sectional area of the column. At the onset of buckling the critical stress is dened as
(9.47)
We write the second area moment as , where r denotes the minimum radius of
gyration of the cross section. For the rectangular section shown in the adjacent sketch,
and , so that , where . Thus, the critical
stress becomes
(9.48)
and is called the slenderness ratio. The slenderness ratio is the column length divided by a cross-sectional
dimension signicant to bending.
For any set of boundary conditions we dene the effective length by the formula
(9.49)
The effective lengths for the four standard boundary conditions are as follows:
The denition of effective length uses case A boundary conditions as a reference. the concept of effective length
accounts for boundary conditions other than simple support, or pinned-pinned end conditions.
The column curve is a plot of the critical stress versus the effective slenderness ratio; i.e., .
For elastic column buckling under all boundary conditions
(9.50)
which is a hyperbola that depends only on the modulus of elasticity E of the material. This equation governing
elastic buckling is called the Euler curve, and columns that buckle in the elastic range are called long columns.
See Fig. 9.14
A pinned-pinned

B clamped-free
C clamped-clamped
D clamped-pinned
P
cr

2
EI L
2
( ) =

cr
P
cr
A
2
EI ( ) AL
2
( ) = =
h
b
I r
2
A =
I
min
bh
3
( ) 12 = A bh = r h 12 = 0 h b < <

cr

2
E
L r ( )
2
----------------- =
L r
KL
P
cr

2
EI
KL ( )
2
-------------- =
P
cr

2
EI
L
2
------
2
EI
KL ( )
2
-------------- = =
KL L = K 1 =
P
cr

2
4
-----
EI
L
2
------
2
EI
KL ( )
2
-------------- = =
KL 2L = K 2 =
P
cr
4
2
EI
L
2
------
2
EI
KL ( )
2
-------------- = =
KL L 2 = K 1 2 =
P
cr
20.2
EI
L
2
------
2
EI
KL ( )
2
-------------- = =
KL 0.699L = K 0.7 =

cr
versus KL r

cr

2
E
KL
r
-------
\ )
[
2
--------------- =
Buckling
266 Thin-Walled Structures
9.4.1 Inelastic buckling
The column curve equation, eq. (9.50), is valid up to the proportional limit of the material, denoted by . The
proportional limit is dened as the stress where the compressive stress-strain curve of the material deviates from
a straight line. If the stress at the onset of buckling is greater than the proportional limit, then the column is said
to be of intermediate length, and the Euler formula, eq. (9.50), cannot be used. The proportional limit is difcult
to measure from test data because its denition is based on the deviation from linearity. In particular, the com-
pressive stress-strain curves for aluminum alloys typically used in aircraft construction do not exhibit a very pro-
nounced linear range. For aluminum alloys a material law developed by Ramberg and Osgood (1943) is often
used to describe the nonlinear compressive stress-strain curve. The Ramberg-Osgood equation is a three parame-
ter t to the compressive stress-strain curves of aluminum alloys. From the experimental compressive stress-
strain curve we measure the slope near the origin, which is the modulus of elasticity E, the stress where the
secant line drawn from the origin with slope 0.85 E intersects the stress-strain curve, and the stress where a
second secant line drawn from the origin with slope 0.7E intersects the stress-strain curve. These data are
depicted in Fig. 9.15. Note that the compressive normal strain corresponding to the stress is usually about
the 0.2% offset yield strain for the material. Hence, stress is close to the 0.2% offset yield stress of the alu-
minum alloy. The Ramberg-Osgood equation is
(9.51)
where the shape parameter n is given by

cr
KL
r
-------
0
Euler curve, depends only on E

p
KL
r
-------
E

p
------ =
long columns
Fig. 9.14 Column curve for elastic
buckling

0.85

0.7

0
E
0.85E
0.7E
1
1
1

0.85

0.7
0.002
experimental compressive
stress-strain curve
Fig. 9.15 Data used to t the compression
stress-strain curve of aluminum alloys.

0.7

0 7 ,


E
--- 1
3
7
---

0.7
---------
\ )
[
n 1
+ =
Thin-Walled Structures 267
Column Design Curve
(9.52)
We can re-write eq. (9.51) as
(9.53)
and plot versus for various values of the shape parameter n., and this plot is shown in Fig.
9.16. Some approximate values for common aluminum alloys are given in the table below.
From the Ramberg-Osgood equation, eq. (9.51), we can determine the local slope of the compressive stress-
strain curve as a function of the stress. This slope of the compressive stress-strain curve is called the tangent
modulus; i.e., where E
t
is the tangent modulus. Differentiating eq. (9.51) we get
AL E in10
6
psi
in 10
3
psi
n
2014-T6 10.6 60 20
2024-T4 10.6 48 10
6061-T6 10.0 40 30
7075-T6 10.4 73 20
n 1
17
7
------
\ )
[
ln

0.7

0.85
-----------
\ )
[
ln
---------------------- + =
E

0.7
---------

0.7
---------
3
7
---

0.7
---------
\ )
[
n
+ =

0.7
E ( )
0.7

0.5 1 1.5 2
0.2
0.4
0.6
0.8
1
1.2

0.7
---------
E

0.7
---------
n 2 =
n 5 =
n 10 =
n 20 =
n 50 =
n 2 =
5
50
Fig. 9.16 A normalized plot of the Ramberg-Osgood material law for various values of
the shape parameter n.

0.7
d
d
------ E
t
=
Buckling
268 Thin-Walled Structures
(9.54)
or
(9.55)
For intermediate length columns it has been demonstrated by extensive testing that the critical stress is rea-
sonably well predicted using the Euler curve, eq. (9.50), with the modulus of elasticity replaced by the tangent
modulus. This inelastic buckling analysis is called the tangent modulus theory. That is,
(9.56)
Now substitute eq. (9.55) for the tangent modulus in this equation, noting that , to get
After division by , this equation can be written as
(9.57)
A plot of the column curve given by eq. (9.57) is shown in Fig. 9.17.
d
d
E
------
3
7
---
n
n 1
E
0.7
n 1
----------------d + =
d
d
------
1
E
t
-----
1
E
---
3
7
---
n
E
---

0.7
---------
\ )
[
n 1
+ = =
E
t
E
1
3
7
---n

0.7
---------
\ )
[
n 1
+
--------------------------------------- =

cr

2
E
t
KL
r
-------
\ )
[
2
--------------- =

cr
=

cr

2
KL
r
-------
\ )
[
2
---------------
E
1
3
7
---n

cr

0.7
---------
\ )
[
n 1
+
--------------------------------------- =

0.7

cr

0.7
---------
3
7
---n

cr

0.7
---------
\ )
[
n
+
1
KL r ( )
E
0.7

------------------------
2
-------------------------------- =
Thin-Walled Structures 269
Bending of thin plates

9.5 Bending of thin plates
Recall that bars and beams are structural elements characterized by having two orthogonal dimensions, say the
thickness and width, that are small compared to the third orthogonal dimension, the length. Thin plates, both at
and curved, are common structural elements in ight vehicle structures, and they are characterized by one dimen-
sion being small, say the thickness, with respect to the other two orthogonal dimensions, say the width and
length. A at plate with rectangular planform is shown in Fig. 9.18 referenced to cartesian axes x, y, and z,
where the x-direction is parallel to the length, the y-direction is parallel to the width, and the z-axis is parallel to
the thickness of the plate. We denote the length of the plate by a, the width by b, and the thickness by t. Trans-
verse loads, or lateral loads, acting in the z-direction applied to the plate are primarily carried by the in-plane
0.5 1 1.5 2 2.5 3 3.5
0.2
0.4
0.6
0.8
1
1.2
KL ( ) r
E
0.7

------------------------

cr

0.7
---------
n 2 =
n 5 =
n 10 =
n 20 =
n 50 =
n =
Fig. 9.17 Column curves for a Ramberg-Osgood material law with different shape factors.

zy

yz

zx

xz

xy
yx

y

x
a t 0 >
b t 0 >
a
b
t
x
y
z
Fig. 9.18 Illustration of the nomenclature and primary stresses for a at, rectangular plate

xy
yx

y

x
3-D stress state
primary stresses in plate theory
Buckling
270 Thin-Walled Structures
stress components
x
,
y
, and
xy
. Transverse shear stresses
xz
and
yz
are necessary for force equilibrium in the
z-direction under transverse loads, but are smaller in magnitude with respect to the in-plane stresses. In plate the-
ory, the transverse normal stress
z
is small with respect to the in-plane normal stresses and, hence, is neglected.
In contrast a bar carries the transverse load primarily carried by the longitudinal normal stress
x
, and the so-
called lateral stresses
y
,
z
,
yx
, and
yz
are assumed to be negligible. Thus, a plate resists transverse loads
through stress components
x
,
y
, and
xy
while a beam resists transverse loads with only the longitudinal stress

x.
Now consider the deformation, or strains, caused by the normal stresses. Hookes law for the normal stresses
and strains in a three-dimensional state of stress is
(9.58)
where E is the modulus of elasticity and is Poissons ratio. For both the plate and beam, the thickness normal
stress
z
is assumed negligible and is set to zero in Hookes law. We neglect the third of eqs. (9.58), so that these
equations reduce to
(9.59)
Consider pure bending of the plate or beam subjected to moment M. We consider two cases. In the rst case
the cross section is compact with dimension b nearly equal to thickness t, and in the second case with dimension
b is much larger that thickness t. In the rst case the structure is a beam and the second case it is a plate. In pure
bending the neutral axis of the beam deforms into an arc of a circle with radius , and the normal strain in the x-

x
1
E
---
x

y

z
( ) =

y
1
E
---
x

y

z
+ ( ) =

z
1
E
---
x

y

z
+ ( ) =

x
1
E
---
x

y
( ) =

y
1
E
---
x

y
+ ( ) =

---
y
z
Section A-A
A
A
x
z

M
M
Fig. 9.19 Pure bending of a beam in the x-z plane and the associated anticlastic curvature of its
cross section.

Thin-Walled Structures 271


Bending of thin plates
direction is . Note that we assumed that the z-axis coincided with the neutral axis in the undeformed
beam. Hence, longitudinal line elements above the neutral axis, z > 0, are stretched, and line elements below the
neutral axis, z < 0, are compressed. In the case of a beam, the normal stress in the y-direction, , is also very
small and is neglected with respect to the longitudinal normal stress . That is, the beam carries the applied
bending moment by the longitudinal normal stress . Since , we get from Hookes law, eq. (9.59), that
(9.60)
Hence, the longitudinal normal stress is the modulus of elasticity times the longitudinal normal strain, and the
normal strain in the y-direction is just Poissons ratio times the longitudinal normal strain. The form of the last
expression for
y
in eq. (9.60) shows that the line elements in the cross section parallel to the y-axis before defor-
mation bend into circular arcs. The transverse line element at z = 0 in the undeformed beam has a radius of curva-
ture of . This transverse curvature is called anticlastic curvature, and is illustrated in Fig. 9.19.
Now consider pure bending of a plate under the same moment M, where now the dimension b is much larger
than thickness t. In this case experiments show that the transverse line elements remain straight over the central
section of the plate, so that the anticlastic curvature is suppressed. In this central section of the plate the trans-
verse normal stress
y
is non-zero. However, the transverse normal stress must vanish at the free edges at y = 0
and y = b, so that anticlastic curvature develops only in narrow zones near the free edges to adjust to vanishing of
the normal stress there. In the central portion of the plate, the associated normal strain is zero. The suppres-
sion of anticlastic curvature is characterized by the vanishing of the transverse normal strain
y
. Hence from
Hookes law, eq. (9.59), for
y
= 0 we get
(9.61)
Since denominator in the expression for
x
is positive but less than unity, the plate is stiffer than the beam owing
to the presence of the non-zero transverse normal stress to help in carrying the applied moment. Compare
eqs. (9.60) and (9.61) for the normal stress
x
. The quantity E/(1 -
2
) is an effective modulus of the plate.

x
z

--- =

x

y
0 =

x
E
x
=

y

x

---z
z

---
\ )
[
--------- = = =

---

x
E
1
2

--------------
x
=
y

x
=

y
Buckling
272 Thin-Walled Structures
9.6 Compression buckling of thin plates
Thin-Walled Structures 273
Compression buckling of thin plates
Buckling
274 Thin-Walled Structures
Thin-Walled Structures 275
Compression buckling of thin plates

Potrebbero piacerti anche