Sie sind auf Seite 1von 386

Thin-Walled Structures

AOE 3024

Dr. Eric R. Johnson


Professor Emeritus

Aerospace and Ocean Engineering Department


Virginia Polytechnic Institute and State University
Blacksburg, VA 24061-0203
Copyright ©2008 by Eric Raymond Johnson. All rights reserved.

No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any
means, electronic, mechanical, photocopying, recording, scanning or otherwise, except as permitted under
Sections 107 or 108 of the 1976 United States Copyright Act, without the prior written permission of the author.
Preface

This is the text for AOE 3024, Thin-Walled Structures, a three credit course required of students in the Aerospace
Engineering and Ocean department at Virginia Tech. The prerequisites for AOE 3124 is ESM 2004, Mechanics of
Deformable Bodies. With respect to engineering curriculums in the United States, this course is unusual in the
sense that both aerospace and ocean engineering students are required to take it. Hence, the mechanics of thin-
walled structural elements is motivated by applications to the design of both airplane and ship structures. How-
ever, the students in aerospace and ocean take separate speciality courses in their respective disciplines after
completing AOE 3024. Mathematica® Version 5.0 1 is used to solve some of the example problems in the text.
However, Mathematica is not required, but some engineering software is needed to solve many of the homework
problems in the text.

I acknowledge the interactions with Professors William Hallauer, Raphael Haftka, Rakesh Kapania, and May-
uresh Patil whose contributions to the subject matter of this course have been used in the preparation of the text.
I accept responsibility for any errors in the text, and would appreciate if the reader would inform me of com-
ments and/or corrections via email (erjohns4@vt.edu).

Eric R. Johnson
August, 2008
Warm Springs, Virginia

1. Mathematica is a registered trademark of Wolfram Research, Inc., 100 Trade Center Drive, Cham-
paign, IL 61820

Thin-Walled Structures i
Preface

ii Thin-Walled Structures
Contents

CHAPTER 1 Airplane and ship structures 1


1.1 Structures and Engineering 1
1.2 Principal structural units 2
1.3 Design 4
1.4 Loads 6
1.5 Function of flight vehicle structural members 6
1.6 Ships’ structures 9
1.7 Key words and concepts from Chapter 1 16
1.8 References 16

CHAPTER 2 Axial force, shear force and bending moment


diagrams 17
2.1 Method of sections 17
2.2 Differential equation method 19
EXAMPLE 2.1 Cantilever wing with tip tank 23
EXAMPLE 2.2 The air load acting on a wing given as discrete data. 27
2.3 Semi-graphical method 34
EXAMPLE 2.3 Uniform barge with symmetric load 34
2.4 Buoyancy Force Distribution on Ships 36
2.5 References 39
2.6 Problems 40

CHAPTER 3 Bars subjected to axial loads 43


3.1 Axially loaded bar 43
3.2 The tensile test 45
3.3 Effect of temperature on strain 49
3.4 Bar reference axis 50
3.5 Linear elastic response 52
EXAMPLE 3.1 Axial bar with a specified uniform distributed load and speci-
fied end displacements 53
EXAMPLE 3.2 A bar with fixed ends and subjected to an axial point force. 54
3.6 References 56
3.7 Problems 56

iii
CHAPTER 4 Axial normal stress in pure bending and extension 59
4.1 Pure bending 59
4.2 Geometry of deformation 60
4.3 Bending normal stress — flexure formula 67
EXAMPLE 4.1 Bending normal stress distribution in a cantilever beam with a
thin-walled zee section. 70
EXAMPLE 4.2 Lateral displacements of the zee section beam 72
4.4 Combined extension and pure bending 75
4.5 Moments of areas 76
EXAMPLE 4.3 Thin-walled zee section properties by the composite body tech-
nique 80
EXAMPLE 4.4 Circular cross section of variable thickness 82
EXAMPLE 4.5 Semicircular section with two stringers 83
4.6 Multi-material bars 84
EXAMPLE 4.6 A multi-material beam with a symmetric cross section 88
4.7 Problems 91

CHAPTER 5 Bending of beams under transverse loads 95


5.1 Approximations for slender beams 95
5.2 Beam displacements 96
EXAMPLE 5.1 A statically indeterminate beam with an unsymmetrical cross
section. 101
EXAMPLE 5.2 Cantilevered beam with an equal leg angle section 105
EXAMPLE 5.3 Contact between two cantilever beams. 108
5.3 Problems 112

CHAPTER 6 Shear flows and stresses due to transverse shear


forces 115
6.1 Rectangular cross section 115
6.2 Shear flows in open section beams 118
EXAMPLE 6.1 Shear flow distribution in a tee beam 120
EXAMPLE 6.2 Shear flow in an equal leg angle section 123
6.3 Shear center of an open section 124
EXAMPLE 6.3 Shear center location in an unsymmetrical section 129
6.4 Summary notes on the shear center 132
Procedure to locate the S.C. 132
6.5 Skin-stringer idealization 133
EXAMPLE 6.4 Shear flows in a stringer-stiffened C-section 136
6.6 Problems 140

iv
CHAPTER 7 Bars subjected to torsional loads 143
7.1 Uniform torsion of a circular tube 143
7.2 Uniform torsion of an open section 149
EXAMPLE 7.1 Torsional response of a thin-walled open section and an equiv-
alent closed section 153
7.3 Non-uniform torsion 154
EXAMPLE 7.2 A uniform distributed torque acting on a bar with fixed ends.
155
EXAMPLE 7.3 A point torque acting on a bar with fixed ends. 156
7.4 Unit twist of a single cell section 157
Geometry of the contour 158
Hooke’s law 160
Shear strain-displacement relation 160
Tangential displacement of a typical point on the contour 161
Relation between the shear flow and unit twist 163
7.5 Closed contour with arbitrary shape 164
EXAMPLE 7.4 Torsion of a circular, bi-material section 166
7.6 Shear center of a closed section 167
EXAMPLE 7.5 Shear center location of a single-cell, closed section having one
axis of symmetry. 168
7.7 Multi-cell closed sections 170
EXAMPLE 7.6 Uniform torsion of a two-cell section 173
7.8 Resultant of uniform shear flow 175
EXAMPLE 7.7 Torsion of a five cell closed section; circuit shear flow 177
7.9 Hybrid sections 180
7.10 Determination of the shear stress tangent to the contour – a summary of
chapters 6 & 7 182
7.11 References 183
7.12 Problems 184

CHAPTER 8 Work and energy methods 187


8.1 Hooke’s law and its consequences 187
The principle of superposition for loads at the same point. 188
Principle of superposition 188
Corresponding forces and displacements and the unique meaning of the total work 190
Maxwell’s reciprocal theorem 191
Betti-Rayleigh reciprocal theorem 192
Stiffness influence coefficients 193
8.2 First law of thermodynamics 194
8.3 Castigliano’s theorems 195
Extensions to include a couple and rotation 196
Generalized forces and displacements 197
8.4 Unit action states 199
8.5 Unit displacement states 200

v
8.6 Strain energies in bar theory 201
Solid, mono symmetric cross section 203
A single cell, thin-walled bar having a mono symmetric cross-sectional contour 205
Unit twist formula by Castigliano’s second theorem 210
8.7 Transverse shear deformations – Timoshenko theory 211
8.8 Application to bars in extension 214
EXAMPLE 8.1 Response of a stepped bar by Castigliano s first theorem 214
EXAMPLE 8.2 A suspended bar subjected to self weight 216
8.9 Application to trusses 217
EXAMPLE 8.3 Three bar planar truss 221
EXAMPLE 8.4 Three bar truss with lack of fit 222
8.10 Application to beams 223
EXAMPLE 8.5 Flexibility coefficients for a cantilever beam 223
EXAMPLE 8.6 End rotations of a simply supported beam subject to an end mo-
ment 224
EXAMPLE 8.7 Effect of transverse shear deformations on the tip deflection of
cantilevered spar 226
EXAMPLE 8.8 Shear compliances of a stiffened blade section 229
EXAMPLE 8.9 Tip displacement of a cantilever wing spar a under distributed
load 231
8.11 Application to Frames 233
EXAMPLE 8.10 Strut-braced spar 234
EXAMPLE 8.11 A frame of two slender members 237
8.12 Reference 238
8.13 Problems 239

CHAPTER 9 Criteria for initial yielding 241


9.1 Ductile and brittle behavior 241
9.2 Criteria for yielding of ductile materials 242
9.3 Stress transformation equations for generalized plane stress 245
9.4 Principal stresses and maximum shear stress 247
EXAMPLE 9.1 Maximum shear stress in tensile test 250
EXAMPLE 9.2 Mohr s circle for hydrostatic stress state 250
EXAMPLE 9.3 Principal stresses and maximum shear stress 251
9.5 Octahedral shear stress 253
9.6 Mises criterion for initiation of yielding 256
9.7 Maximum shear-stress criterion 258
EXAMPLE 9.4 Factor of safety against initial yielding 260
EXAMPLE 9.5 Stress responses of a stringer-stiffened, single cell beam. 261
EXAMPLE 9.6 Minimum weight design of the beam in Example 9.5 subject
to a constraint on initial yielding 265
9.8 References 268

vi
CHAPTER 10 Matrix structural analysis 269
10.1 Flexibility and stiffness matrices 269
EXAMPLE 10.1 Two springs in series restrained at one end. 269
10.2 Unrestrained structural stiffness matrix 273
10.3 Assembly of unrestrained structural stiffness matrices 275
10.4 Prescribed nodal displacements and forces 277
10.5 Solution for the unknown nodal variables 279
10.6 Stress matrix 281
10.7 Direct stiffness method 282
10.8 Planar trusses 283
EXAMPLE 10.2 A three-bar truss 285
EXAMPLE 10.3 Restrained three-bar truss of Example 10.2. 287
EXAMPLE 10.4 Five bar truss 291
EXAMPLE 10.5 Using symmetry to reduce problem size 293
10.9 Self strained plane truss 297
EXAMPLE 10.6 A simple statically indeterminate truss subject to heating 302
10.10 Structures containing beam elements 307
EXAMPLE 10.7 Multi-span beam 314
10.11 Beam element with distributed loading; fixed end actions 319
EXAMPLE 10.8 Clamped-clamped, stepped beam restrained by a spring 322
10.12 Frame structures 327
EXAMPLE 10.9 A two-bar frame subjected to a distributed load 333
10.13 Reference 339
10.14 Problems 340

CHAPTER 11 Buckling 347


11.1 One-degree of freedom model 347
Static equilibrium 348
Stability analysis 349
11.2 Perfect Columns 351
EXAMPLE 11.1 Critical load for clamped-free boundary conditions (B) 354
11.3 Imperfect columns 357
Eccentric load 357
Geometric imperfection 359
11.4 Column Design Curve 361
Inelastic buckling 362
11.5 Bending of thin plates 365

vii
viii
CHAPTER 1 Airplane and ship
structures

The objective of Thin-Walled Structures is to provide an understanding of the basic concepts of stress and
deformation 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 definitions 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 defined 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 Hooke’s
discoveries is discussed in chapter two of an informative and readable book on structures by Gordon (1978). Gor-
don’s text is the source for what is written here. By Newton’s 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 fine scale of molecules where there is a large resistance to
stretching and compressing of chemical bonds. Hooke’s 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 deflections 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

Thin-Walled Structures 1
Article 1.2

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 deflection, 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 defined when a stress (internal force
intensity) exceeds the yield strength of the material, or failure can mean excessive displacements which occur
during buckling. The stiffness and strength of a structure depend on its geometrical configuration, 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 fiber-reinforced composites are made from stiff, strong, continuous fibers 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 fixed-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 figures, 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, or shell 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 ordinary differential equations. Aero-hydrodynamic and other loads that act on the structure cause
extensional, bending, 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 contrast 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.

2 Thin-Walled Structures
Principal structural units

(a) Global Swift

(b) Taylorcraft airplane


Fig. 1.1 Principal structural units of light airplanes (from Aircraft Basic Science, 1948)

Thin-Walled Structures 3
Article 1.3

Fig. 1.2 Principal structural units of a Douglas airplane (from Aircraft Basic Science, 1948)

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 configuration,
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).

4 Thin-Walled Structures
Design

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

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 first 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

Thin-Walled Structures 5
Article 1.4

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

1.4 Loads
The first step in preliminary design is to determine the external loads acting on the structure. Maneuvering flight
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 sufficiently
complex such that their determination is done by specialists rather than by designers.

Loads on a vehicle structure may be classified 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 influence 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 configuration 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 flexible vehicles the loads greatly
influence 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 flexible 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 flight vehicle structural members


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

The structure of a flight 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 efficient location for the
structural material is at the outer surface of the vehicle. As a result, the structures of most flight
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 efficient 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 first glance these structures appear to differ considerably, functionally there are simi-
larities. Both have thin-sheet coverings, longitudinal stiffening members, and transverse sup-

6 Thin-Walled Structures
Function of flight vehicle structural members

Longitudinal
stringers

Cover Skin
Transverse
frames
(a)

Cover skin

Transverse
Spar web rib

Spar cap Longitudinal


stringers
(b)

Fig. 7-5 Typical semimonocoque construction. (a) Body


structures: (b) aerodynamic surface structures.

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.

Thin-Walled Structures 7
Article 1.5

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 significant 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
simplification of an actual structure into an analytical model represents a compromise, since
elaborate models which nearly simulate the actual structure are usually difficult 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,

8 Thin-Walled Structures
Ships’ structures

Actual skin and web


carries axial and
shear stresses

(a) Effective longitudinals


(axial stress only)
Idealized webs
(shear stress only)

(b)
Fig. 7-6 Idealization of semimonocoque structure. (a) Actual
structure; (b) idealized structure

Effective longitudinals
(axial stress only)

Wall carries axial


Idealized web
and shear stresses
(shear stress only)
(a) (b)
Fig. 7-7 Idealization of a monocoque shell. (a) Mono-
coque shell; (b) idealization

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 significant
structural or loading discontinuities in the region where the stresses are computed. In many
flight 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 ship’s 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

Thin-Walled Structures 9
Article 1.6

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.
Wave crest

Wave crest

Wave crest

Wave crest
Fig. 4 Ship moving diagonally across waves

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.

A.P. F.P.
1/4 length 3 length
/4

Fig. 5 Wave surface at various


Amidships positions in length

Consider now a transverse section of a ship as shown in Fig. 6. This section is subject first
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 ship’s structure.

10 Thin-Walled Structures
Ships’ structures

Cargo load

Cargo load

Water pressure load

Fig. 6 Distortion of transverse section due to static loading

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 floors or longitudinals. Fig. 7 shows a strip of such
plating between two floors or longitudinals. The tendency is for the plating to bend as a beam

Inner bottom

Floors

Outer bottom

Water pressure
Fig. 7 Distortion of bottom plating due to water pressure

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
defined as local deformation.
Summarising this section, it is clear that the overall problem of the strength of the ship’s
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

Thin-Walled Structures 11
Article 1.6

be very complex. It is for this reason that, in a first study at least of the strength of the ship’s
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 ship’s 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 ship’s structure which will resist this bend-
ing; or in other words there must be material distributed in the fore and after direction to fulfil
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.

Upper deck
plating

2nd deck
plating

Side
shell

Inner bottom
plating
Margin Centre
plate girder
Bottom shell

Fig. 8 Section through ship showing material resisting longitudinal bending

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 fit 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 floors 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 floors, 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

12 Thin-Walled Structures
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.

Tween deck Upper deck


frame Tween deck
beam
pillar

Second deck
beam Hold
pillar

Floor
plate
Tank side Centre
bracket girder
Side girder

Fig. 9 Section through ship showing transverse structure

Another function of these transverse rings is to prevent transverse distortion of the struc-
ture, so that the floors, 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 floor 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 floors in the

Thin-Walled Structures 13
Article 1.6

Tween deck
frame
Upper deck
longitudinals

Deep transverses
2nd deck spaced 6 to 12 ft.
longitudinals apart
Side frame

Inner bottom
longitudinals

Outer bottom
longitudinals
Fig. 10 Section through ship with longitudinally stiffened decks and bottom

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

Flat bar deck


longitudinals

Transverses spaced Side Centre girder


about 10 ft. apart girder
Wing bulkhead
Flat bar side longitudinals Flat bar longitudinals
Side girder
Center girder

Flat bar bottom


longitudinals
Fig. 11 Section through large modern oil tanker

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

14 Thin-Walled Structures
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 fitted 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-

Hatch coaming
Girder
Pillars
Hatch coaming

Girder
Bulkhead

Bulkhead
Pillars

Inner bottom

Outer bottom
Elevation through hold

Pillars
Girder

Deck
beams

Girder
Pillars

Plan at deck
Fig. 12 Wide-spaced pillar and girder arrangement in transversely framed ship

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

Thin-Walled Structures 15
Article 1.7

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 floors 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 briefly the functions which the various parts of the
ship’s 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, floor 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 don’t 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.

16 Thin-Walled Structures
CHAPTER 2 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.

2.1 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 first
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 3c, and is simply supported at
two locations as shown in Fig. 2.1. 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 figure, and the x-axis perpendicular to the plane of the figure. The
origin is at the left end of the bar; 0 ≤ z ≤ 3c . The bar is in equilibrium subject to the external loads F and p y ( z )
which act in the z-y plane. Neglect the weight of the bar relative to the loads F and p y ( z ) , 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 Fz and Fy. The load

Thin-Walled Structures 17
Article 2.1

F Fy py ( z )

z
Fz

c 2c

Fig. 2.1 A slender bar simply supported at two locations and subjected to force F
and distributed load p y ( z ) .

p y ( z ) 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 find the internal axial
force N(z), shear force Vy(z), and bending moment Mx(z). We write the internal actions as N(z), Vy(z), and Mx(z),
since they are mathematically functions of the coordinate z. The basis for their determination is equilibrium.

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. 2.2. The internal actions N, Vy, and Mx are shown as well as the unknown support reactions Ay,
Bz, and By. Let us assume the distributed load p y ( z ) vanishes in this discussion. Internal actions N, Vy, and Mx
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 Vy acts in the positive y-direction, and the posi-
tive moment Mx 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, Vy, and Mx 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), Vy(z), and Mx(z), is to first draw an overall free-body diagram of the
bar removed from its supports and find the unknown support reactions, and then section the bar at various z-loca-
tions to find N, Vy, and Mx. The reader should verify that the support reactions in Fig. 2.2 are A y = ( – 3 ⁄ 2 )F y ,
B z = F z , and B y = F y ⁄ 2 , where Fy and Fz are the known applied loads. Note that the force Ay 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. 2.3. 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 Vy(z) and Mx(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
force: V y ( c + ) – V y ( c - ) = ( 3 ⁄ 2 )F y . The shear force is a piecewise constant for this problem. The bending

18 Thin-Walled Structures
Differential equation method

Fy

Fz Bz

Ay By
c 2c

overall FBD
Fy
Vy

Fz N Bz

z Mx Mx
Ay By
Vy
z 2c

Left and right FBD’s for cuts in the range 0 < z < c

Fig. 2.2 Free body diagrams (FBDs) of the slender bar with p y ( z ) = 0.

moment is piecewise linear; i.e., Mx(z) = z Fy, 0 < z < c, and Mx(z) = (3c - z)(Fy/2), c < z < 3c. The bending
moment is continuous at z = c, but has a discontinuity in slope M' x ( c + ) – M' x ( c - ) = ( 3 ⁄ 2 )F y , where
M' x = dM x ⁄ dz . As illustrated in Fig. 2.2, there are two free-body diagrams for each cut. Either one may be
used to find N, Vy, and Mx. If the left free-body diagram is used to find N, Vy, and Mx, then equilibrium of the
right free-body diagram will give the same values of N, Vy, and Mx. 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.

2.2 Differential equation method


The distributed load intensity p y ( z ) , the shear force Vy(z), and bending moment Mx(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.

Consider a portion of a straight beam subjected to distributed load whose intensity is p y ( z ) as shown in Fig.
2.4. By convention p y ( z ) is positive upwards and negative downwards. A free-body diagram of a portion ∆z-long
of the beam is shown in Fig. 2.5. The shear force and bending moment change with z, and so their values at z +

Thin-Walled Structures 19
Article 2.2

Fy Fy/2
Fz z
c 2c
3/2 Fy
N
Fz

0 z

Vy

Fy/2
0 z

- Fy Vy

N + N
Mx Mx Mx
Vy

0 z

- cFy

Fig. 2.3 Axial force N, shear force Vy, and bending moment Mx
diagrams for the slender bar with p y ( z ) = 0,

∆z are different from their values at z. The distributed load acting on the segment ∆z is replaced by single force of
magnitude p y ∆z acting long a line of action given by z = z*, where z < z* < z + ∆z. Mathematically this is per-
missible by the mean value theorem (for integrals) for continuous functions of a single variable, and from this
theorem we have
( z + ∆z )
1
p y = ------ p y ( ζ ) dζ
∆z ∫
z

20 Thin-Walled Structures
Differential equation method

Note that in the limit as ∆z → 0 , z * → z , and p y → p y .

py ( z )
y

Fig. 2.4 Distributed load intensity shown acting in the positive sense

py ( z )

Vy + ∆Vy

Fig. 2.5 Mx

Vy Mx + ∆Mx
z
z + ∆z

p y ( z * )∆z

Vy + ∆Vy
Fig. 2.6
Mx
Vy Mx + ∆Mx
z
z*
z + ∆z

In the free-body diagram of Fig. 2.6 we sum forces vertically, divide by ∆z, and take the limit as ∆z → 0 to
get

dV y
= –py (2.1)
dz
Often it is more convenient to use an integrated form of eq. (2.1). Integrating it from z1 to z we obtain

V y ( z ) = V y ( z 1 ) – p y ( ζ ) dζ
∫ (2.2)
z1

Thin-Walled Structures 21
Article 2.2

The integrated form is valid only if there are no point loads between z1 and z. If we sum moments at z in the free-
body diagram, divide by ∆z, and take the limit as ∆z → 0 we get

dM x
= Vy (2.3)
dz
The integrated form of eq. (2.3) is
z

M x ( z ) = M x ( z 1 ) + V y ( ζ ) dζ
∫ (2.4)
z1

which is valid if no point couples act between z1 and z. Equations (2.1) and (2.3) are the differential equations of
equilibrium for the beam. Equation (2.1) is valid at z if no point force acts there, and eq. (2.3) is valid at z if no
point couple acts there. Considering separately the equilibrium of a point force F0 and a point couple with
moment magnitude C0 acting at z = z0, as shown in Fig. 2.7, we obtain

F0

y C0

V y ( z 0+ )
z V y ( z 0- ) ε

M x ( z 0+ )
M x ( z 0- )
z0

Fig. 2.7 Point force and couple acting at z0 on an inifintesimal beam element

V y ( z 0+ ) – V y ( z 0- ) = – F 0 M x ( z 0+ ) – M x ( z 0- ) = – C 0 (2.5)

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 = z1 to a typical value of z, z > z1, which has a
distributed load with intensity p y ( z ) acting on it is shown in Fig. 2.8. Using ζ as a dummy variable to measure
the axial position, the resultant force F y ( z ) is

Fy ( z ) =
∫ p ( ζ ) dζ
y (2.6)
z1

and the center of pressure z p ( z ) is given by

z p ( z )F y =
∫ ζp ( ζ ) dζ
y (2.7)
z1

Both the resultant force F y ( z ) and center of pressure z p ( z ) are functions of z. Equations (2.6) and (2.7) are con-

22 Thin-Walled Structures
Differential equation method

py ( z )
Fy ( z )
Vy ( z1 ) Vy ( z ) Vy ( z1 ) Vy ( z )

Mx ( z1 ) Mx ( z ) Mx ( z1 ) Mx ( z )
z1 z1

ζ
zp ( z )
z z

Fig. 2.8 Resultant force at the center of pressure statically equivalent to a distributed load.

ditions of statical equivalence for the distributed load p y ( z ) .

EXAMPLE 2.1 Cantilever wing with tip tank

Consider the cantilever wing with tip tank as shown in Fig. 2.9. Given the weight of the tip tank and its contents

z
p y ( z ) = p 0 ---
L

z
e

L
W

Fig. 2.9 Cantilever wing with tip tank.

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 p 0 at the wing root, determine the shear force and bending moment along the span. The solution to this prob-
lem is given by a Mathematica program listed below.

Thin-Walled Structures 23
Article 2.2

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.


z
py = p0 ;
L
The shear force and bending moment distributions are determined from eqs. (4.2) and
(4.4) with z1 = 0. The shear force and bending moment at the wing tip (z = 0) are
denoted by Vy0 and Mx0 , respectively.

Vy = Vy0 - py z

z2 p0
- + Vy0
2L

Mx = Mx0 + Vy z

z3 p0
Mx0 - + z Vy0
6L
Boundary conditions at the wing tip, obtained from equilibrium of the tip tank, deter-
mined the shear force Vy0 and moment Mx0 .
bc1 = HVy . z fi 0L - W
bc2 = HMx . z fi 0L - e W
-W + Vy0
-e W + Mx0
slv1 = Solve@bc1 == 0, Vy0 D
88Vy0 fi W<<
Vy0 = Vy0 . slv1@@1DD
W
slv2 = Solve@bc2 == 0, Mx0 D
88Mx0 fi e W<<
Mx0 = Mx0 . slv2@@1DD
eW

24 Thin-Walled Structures
Differential equation method

Print@ "Shear force Vy HzL = ", Vy D


Print@"Bending moment Mx HzL = " Mx D
z2 p0
Shear force Vy HzL = W -
2L
z3 p0 z
Bending moment Mx HzL = i
j
je W + W z -
y
z
k 6L {
Plot the shear force and bending moment diagrams for the following parameter
values: L = 144 in, p0 = 70 lb/in, W = 500 lbs, and e = 6 in. (Plots labled p1 and
p2 have been suppressed using the DisplayFunction option.)
p1 =
Plot@HVy . 8 L fi 144, p0 fi 70, W fi 500, e fi 6<L,
8z, 0, 144<,
PlotRange fi 81000, -5000<,
GridLines fi Automatic,
AxesLabel fi 8"z,inches", "Vy , lbs"<,
PlotLabel fi
StyleForm@"Shear force diagram",
"Section"D,
DisplayFunction -> IdentityD
Graphics
p2 =
Plot@HMx . 8 L fi 144, p0 fi 70, W fi 500, e fi 6<L,
8z, 0, 144<,
GridLines fi Automatic,
AxesLabel fi 8"z,inches", " Mx , lb-in"<,
PlotLabel fi
StyleForm@"Bending moment diagram",
"Section"D,
DisplayFunction -> IdentityD
Graphics
Show@GraphicsArray@88p1<, 8p2<<DD
(The plots are are shown on the next page.)
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
Mx . 8 L fi 144, p0 fi 70, W fi 500, e fi 6, z fi 144<
-166920

Thin-Walled Structures 25
Article 2.2

1 2 3 4 5 6

Vy , lbs Shear force diagram


1000

z,inches
20 40 60 80 100 120 140
-1000

-2000

-3000

-4000

-5000

Mx , lb-in Bending moment diagram

10000

z,inches
20 40 60 80 100 120 140
-10000
-20000
-30000
-40000
-50000

Shear force and bending moment diagrams for the cantilever wing with tip tank

26 Thin-Walled Structures
Differential equation method

The shear force and bending moment at the wing tip are determined from equilibrium of the tip tank, which
gives Vy(0) = W and Mx(0) = eW. Note that eq. (2.1) shows that the slope on the shear diagram is equal to the
distributed load intensity. For example at z = 0, p y ( 0 ) = 0 , so that the slope of the shear diagram is zero at z = 0.
Similarly, the slope on the moment diagram dM x ⁄ dz is equal to the shear force as given by eq. (2.3). 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 Mx may be either a local maximum, minimum, or a value corre-
sponding to horizontal inflection point. It is important to compute the largest magnitude of the bending moment,
and this is accomplished by checking the bending moments where Vy = 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 M x = – 166920 lb-in . The bending moment changes sign, and hence vanishes, at
z = 81.4 inches. In a plot of the beam deflection versus z, which is not shown above, the location z = 81.4 inches
is called an inflection point because the curvature is changing from concave down (positive Mx) to concave up
(negative Mx) as z increases through z = 81.4 inches.

EXAMPLE 2.2 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 program.
Since the air load on the wing is given at discrete spanwise locations and not as a mathematical function, it is use-
ful to use Mathematica’s list manipulation capabilities to effect the solution. Before giving the problem state-
ment, 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 { item 1, item 2, item 3, …, item n } . It is usually more efficient 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
{ 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 ] ] }
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.,
Length [ { 5, 8, 11 } ] = 3 . A summary of Mathematica’s 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 progres-
sion)
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]

Thin-Walled Structures 27
Article 2.2

Summary of list manipulation functions in Mathematica (taken from Blachman, 1992)

Function Description
Sort [list] Sorts elements of list into canonical order
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 first n elements from list
Take [list, n] Takes the first n elements from list
First [list] Give the first 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 first 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 [ list1, list2, ... Gives the complement, i.e., those elements in list1 but not in list
] 2, ...
Intersection [ list1, list2, ... ] Gives a sorted list of all the elements common to all list1, list2, ...
Union [list1, list2, ... ] Gives a sorted list of the distinct elements
Join [list1, list2, ... ] 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

28 Thin-Walled Structures
Differential equation method

Problem statement: The aerodynamic loads on an airplane wing cannot be represented by a simple equa-
tion. The load per inch of span, p y ( z ) = p ( z ) , of the airplane wing shown in Fig. 2.10 is 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.

z
Fig. 2.10

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
the centerline of the airplane, as shown in Fig. 2.10. These distances are measured along the wing rather than
horizontally, since the air loads are perpendicular to the wing. The distances between stations, ∆z , are computed
as a list in the program. 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 p ave and the base ∆z . The change in the shear ∆V y between two stations is equal to the
area of the load curve between the stations. The shear V y is then obtained by summation of the ∆V y -values. The
change in the bending moment ∆M x 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 ∆M x -values. Plots
of (the airy,load, shear
, ppforce, and bending
) moment distributions are shown at the end of the Mathematica program

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<

Thin-Walled Structures 29
Article 2.2

1. 2 3 4 5 6 7

Compute average airload intensity in each interval between stations.

In[4]:= pave = 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 DVy - values is computed by a direct multiplication of lists Dz and pave .

In[5]:= DVy = - Dz * pave

Out[5]= 8-2480., -2430., -2360., -2270., -2160.,


-2030., -1870., -1690., -1510., -1290., -930., -87.5<

tip HdV dzL z, and V HtipL= 0, the shear force at the root is the negative of
Since Vy HtipL - Vy HrootL = root y y
the sum the elements in the DVy -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@DVy D

Out[6]= List

In[7]:= Vy0 = - Apply@Plus, DVy D

Out[7]= 21107.5

The shear force at station i is the the partial sum of the of the DVy - values over the intevals from 1 to i; i.e.,
Vy HiL = ij=1 DV y HjL + Vy0 . We use the Table function to generate a list of the shear force values at each
station.

In[8]:= Vy = Table@HSum@DVy @@jDD, 8j, 1, i<D + Vy0 L, 8i, 1, Length@DVy 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
zi - station, including the root.

In[9]:= Vy = PrependTo@Vy , Vy0 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.<

30 Thin-Walled Structures
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]:= Vave = Take@N@HRotateLeft@Vy D + Vy L 2D, Length@Vy 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]:= DMx = Dz * Vave


Out[11]= 8397350., 348250., 300350., 254050., 209750.,
167850., 128850., 93250., 61250., 33250., 11050., 218.75<

tip HdM dzL z, and M HtipL= 0, the bending moment at the root is the
Since Mx HtipL - Mx HrootL = root x x
negative of the sum the elements in the DMx -list.

In[12]:= Mx0 = -Apply@Plus, DMx D

Out[12]= -2.00547 · 106

Compute the bending moment at each station.

In[13]:= Mx = Table@HSum@DMx @@jDD, 8j, 1, i<D + Mx0 L,


8i, 1, Length@DMx D<D;
Mx = PrependTo@Mx , Mx0 D
Out[13]= 9-2.00547 · 106 , -1.60812 · 106 , -1.25987 · 106 ,
-959519., -705469., -495719., -327869., -199019.,
-105769., -44518.7, -11268.8, -218.75, 0.=

In[14]:= TableForm@Transpose@8z, p, Vy , Mx <D,


TableHeadings fi
8None, 8"z,in", "p,lb in",
"Vy ,lb", "Mx ,lb-in"<<D

Thin-Walled Structures 31
Article 2.2

1 2 3 4 5 6 7

Out[14]//TableForm=

z,in p,lb in Vy ,lb Mx ,lb-in


0 125 21107.5 -2.00547 · 106
20 123 18627.5 -1.60812 · 106
40 120 16197.5 -1.25987 · 106
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 fi 8PointSize@0.02D<,
DisplayFunction -> IdentityD
p12 = ListPlot@Transpose@8z, p<D,
PlotJoined -> True,
DisplayFunction -> IdentityD
p1 = Show@p11, p12,
AxesLabel fi 8"z, in.", "lb in"<,
PlotLabel fi "Airload distribution",
DisplayFunction -> IdentityD
p21 = ListPlot@Transpose@8z, Vy <D,
PlotStyle -> 8PointSize@0.02D<,
DisplayFunction -> IdentityD
p22 = ListPlot@Transpose@8z, Vy <D,
PlotJoined -> True,
DisplayFunction -> IdentityD
p2 = Show@p21, p22,
AxesLabel fi 8"z, in.", "lb"<,
PlotLabel fi "Shear force",
DisplayFunction -> IdentityD
p31 = ListPlot@Transpose@8z, Mx <D,
PlotStyle -> 8PointSize@0.02D<,
DisplayFunction -> IdentityD

32 Thin-Walled Structures
Differential equation method

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

lb in Airload distribution
120

100

80

60

40

20

z, in.
50 100 150 200

lb Shear force
20000

15000

10000

5000

z, in.
50 100 150 200

lb-in Bending Moment


z,in
50 100 150 200

-500000

6
-1·10

6
-1.5·10

6
-2·10

Thin-Walled Structures 33
Article 2.3

2.3 Semi-graphical method


The semi-graphical method to draw the shear force and bending moment diagrams is best illustrated by doing an
example.

EXAMPLE 2.3 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. 2.11. This is an example of a structure with no boundary supports, and is typical of aero-

5m 5m 5m 5m
15kN (total) 15kN (total)
10kN

Fig. 2.11 Uniform section barge in still water with symmetric load.

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 p b represent the distributed load intensity due to buoyancy, and p b 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 first sketch the distributive load p y ( z ) , then the shear force Vy(z), and finally the bending moment
Mx(z). Equations (2.1) and (2.3) 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. (2.2) and
(2.4) set at z = z2 give

z2

V y ( z 2 ) – V y ( z 1 ) = – p y ( z ) dz
∫ (2.8)
z1

z2

Mx ( z2 ) – Mx ( z1 ) =
∫ V ( z ) dz
y (2.9)
z1

Equation (2.8) is interpreted in a graphical sense to mean that the difference in the shear force between z2
and z1 is the area under the distributed loading diagram from z1 to z2. This is not geometrical area. The area
between the p y ( z ) curve and the z-axis has units of force, and may be positive, zero, or negative. Similarly eq.
(2.9) 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 p b = ( 2kN ) ⁄ m .
The total distributed load intensity is the difference between p b and the magnitude of the downward acting

34 Thin-Walled Structures
Semi-graphical method

applied loading intensity. The distributed loading intensity diagram is constructed in this manner as shown in Fig.
2.12.

py
10 kN
2

kN/m 15
0 z, m
5 10 20
-1

kN Shear force diagram

z, m
5 10 15 20

-2

-4

kN-m Bending moment diagram

17.5
15
12.5
10
7.5
5
2.5
z, m
5 10 15 20

Fig. 2.12 Shear force and bending moment diagrams for the barge in still water

Thin-Walled Structures 35
Article 2.4

The point force of 10kN acting at z = 10 m is shown schematically in the p y ( z ) -diagram as a downward
pointing arrow. Actually, p y → – ∞ as z → 10m , because a point force is a finite 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.
2.12.The shear force diagram is drawn below the loading intensity diagram in Fig. 2.12. Equilibrium at z = 0
requires Vy(0) = 0, and the slope dVy/dz at z = 0 is equal to 1kN/m. The slope is constant between 0 < z < 5m ,
thus Vy(z) is a straight line in this range of z. The difference in the shear force between z = 5 m and z = 0 is equal
to the negative of the area under the p y ( z ) curve which is 5 kN. Thus Vy(5) = 5 kN since Vy(0) = 0. At z = 5+m
the loading intensity jumps to +2 kN/m. The slope of the shear force jumps from 1 kN/m to -2 kN/m at z = 5m,
but the shear force is itself continuous. The difference Vy(10) - Vy(5) is equal to the negative of the area between
the p y ( z ) -curve and the z-axis between z = 5 m and z = 10 m. Thus Vy(10) - Vy(5) = -10 kN, so Vy(10) = -5 kN.
Note the shear force is zero at z = 7.5 m. At z = 10 m the point force of 10 kN acts. According to the first of eqs.
(2.5) Vy(10+) - Vy(10-) = 10 kN, so that Vy(10+) = 5 kN. The slope of the shear at z = 10 m is +2 kN/m, and
remains constant until z = 15 m. The difference Vy(15) - Vy(10+) = -10 kN, so that Vy(15) = -5 kN. Finally, the
slope changes to +1 kN/m at z = 15+m and remains constant in the range 15 < z < 20. The difference Vy(20) -
Vy(15) = 5 kN, so that Vy(20) = 0. Checking vertical equilibrium at z = 20 m verifies that Vy(20) should be zero.
Moment equilibrium at z = 0 shows Mx(0) = 0. The slope of Mx at z = 0 is equal to the shear force at z= 0.
Hence ( dM x ) ⁄ ( dz ) = 0 at z = 0, as shown in Fig. 2.13. The slope on the moment diagram increases linearly
from zero at z = 0 to 5 kN at z = 5 m. Thus Mx(z) is parabolic from z = 0 to z = 5. The difference Mx(5) - Mx(0) is
equal to the area under shear diagram from z = 0 to z = 5 m. Hence, Mx(5) - Mx(0) = 12.5 kNm, and Mx(5) = 12.5
kNm since Mx(0) = 0. From z = 5 to z = 7.5 the slope of the moment decreases from 5 kN to zero. At z = 7.5, Mx
is a local maximum with a magnitude of 18.75 kNm. The slope of Mx(z) for 7.5 < z < 10 is negative, decreasing
linearly from zero to -5 kN. The difference Mx(10) - Mx(7.5) = -6.25 kNm, so that Mx(10) = 12.50 kNm. The
slope of Mx(z) at z = 10 m jumps from a -5 kN to a +5 kN as shown in Fig. 2.13, but the moment itself is contin-
uous. 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 = 10 m and the bending moment is sym-
metric about z = 10 m. This follows from the symmetrical loading on the barge and equilibrium eqs. (2.1) and
(2.3).

2.4 Buoyancy Force Distribution on Ships


The simple uniform buoyancy distribution acting on the barge in Example 2.3 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 p y ( z ) . In ship design three conditions are recognized to compute p y ( z ) for the same ship.
These conditions are called
• the still water condition,
• sagging condition, and

36 Thin-Walled Structures
Buoyancy Force Distribution on Ships

• the hogging condition.

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. 2.13, and a section between z and z + dz is shown in Fig. 2.14.

Fig. 2.13

A ( z )dz
dF b

Fig. 2.14

Archimedes' principle asserts that the buoyant upthrust is equal to the weight of the fluid displaced. Let A(z)
denote the submerged cross section at z, and let γ denote the specific weight (force per volume) of the fluid. The
differential buoyancy force dFb acting on the ship over a differential length dz is

dF b = γA ( z )dz (2.10)

Consequently, the buoyant upthrust per unit ship length, which we designate p b , is equal to γA(z); i.e.,

dF
p b = --------b- = γA ( z ) (2.11)
dz
A curve of p b for a ship as well as the weight per unit length is shown in Fig. 2.15. 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
case for the barge in Example 2.3, the distribution of the buoyancy per unit length p b 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. 2.16. (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. 2.16. The immersed cross section is the largest at or near the wave crests, and is least near the trough. The

Thin-Walled Structures 37
Article 2.4

weight/length

buoyancy/length

Fig. 2.15

intensity of the buoyancy distribution reflects 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.

concentrated weight

pb

Fig. 2.16 Sagging

The hogging condition is depicted in Fig. 2.17. 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 p b , shown in Fig. 2.17, reflects 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.

38 Thin-Walled Structures
References

concentrated weights

pb

Fig. 2.17 Hogging

2.5 References
Blachman, N.R., 1992, Mathematica: A Practical Approach, Prentice Hall, Englewood Cliffs, New Jersey,
p. 133.

Muckle, W., 1967, Strength of Ship’s 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.

Thin-Walled Structures 39
Article 2.6

2.6 Problems
1. Draw the shear force and bending moment diagram for the simply-supported beam with a moment at the
center. See the figure below.

2 N-m

1m 1m

2L 2
2. The cantilever wing is subjected to a distributed air load p y ( z ) = ------------- 1 – ( z ) , where the total lift (2
πz max
wings) L = 20, 000lbs at cruise, wing length z max = 32.5 ft., and z = z ⁄ z max . Also, the wing supports an
engine weighing 1000 lbs. Plot the loading diagram, shear force diagram V y ( z ) , and bending moment diagram
M x ( z ) as functions of z for 0 ≤ z ≤ 32.5 ft. Partial answer: V y ( 0 ) = 9, 000 lb. and M x ( 0 ) = – 131934 lb-ft.

y
py ( z )

py Vy

z
Mx Vy
+ Mx

6 ft. 1000 lb engine

32.5 ft.

fuselage

3. A proposed solar airplane called Centurion is being designed to achieve semi-perpetual flight (Aviation Week
& Space Technology, May 4, 1998, p.54). Centurion is a flying wing with a span of 206 ft., an 8-ft. chord, and no
taper or sweep. The wing has five 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 avion-
ics, 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 along the
span. For steady level flight, determine the shear force and bending moment diagrams from the centerline of the
wing to its tip, and show them in a sketch. Label significant points. The front view of half of the Centurion is
shown below.

40 Thin-Walled Structures
Problems

z
z1 4.0 ft.
z2 12.0 ft. z1
z3 19.2 ft. z2
z3
z4 32.7 ft. z4
z5 46.3 ft. z5
z6
z6 60.0 ft.
z7
z7 67.1 ft. z8
z8 80.6 ft. z9
z 10
z9 93.4 ft.
Centerline
z10 103.0 ft.

4. The barge shown below has a uniform cross section along its length and is subjected to a uniformly distrib-
uted load of intensity p y ( z ) = – P , in which P has dimensional units of force/length. Also it is subjected to buoy-

z
ancy for the extreme hogging condition py ( z ) = P 1 – cos  π --- . Draw the shear force,
buoyancy  L

V y ( z ) , and bending moment, M x ( z ) , diagrams. Label significant points. Note that M x = 2P ( L ⁄ π ) 2 .


max

z
L L

waterline

5. 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)

Thin-Walled Structures 41
Article 2.6

40 ft 40 ft 40 ft 40 ft 40 ft 40 ft

empty 400 950 950 400 empty


tons tons tons tons 30 ft

6. A barge of uniform rectangular construction has a length of 30m, breadth of 10m, depth of 5m, floats 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/m3)

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 d 0 = 3.7 m , and
that the draft at z = L = 30m is d L = 4.3m

b) Plot the loading intensity diagram, p y ( z ) , where the loading intensity is in N/m, Newton/meter.
(The specific weight in N/m3 is g times the mass density in kg/m3. If for simplicity g is taken as 10
(instead of 9.8) then specific weight of fresh water is 10m/sec 2 × 1000kg/m 3 = 10, 000N/m 3 .)

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.

trim angle
z
5m
2m d0 200t 245t dL
30m 155t

unloaded loaded

42 Thin-Walled Structures
CHAPTER 3 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 defined in Section 3.4. 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 3.1 to
Section 3.5. These three fundamental steps of static structural mechanics are
• equilibrium conditions,
• strain-displacement conditions, or conditions of geometric fit, and
• a material law, or constitutive behavior.

3.1 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 0 ≤ z ≤ L . The bar is subjected to the following loads: a distributed force per unit length of intensity p z ( z ) ,
either an axial force Q 1 or axial displacement q 1 at the left end, and to either an axial force Q 2 or axial dis-
placement q 2 at the right end. The distributed force intensity p z ( z ) , forces Q1 and Q2, and the corresponding
displacements q1 and q2, respectively, are defined positive if they act in the positive in the positive z-direction.
See Fig. 3.1. Under the imposed loads, the bar is in tension and/or compression.

Thin-Walled Structures 43
Article 3.1

y
y
pz ( z )
Q 1, q 1 Q 2, q 2
z, w x

L
Cross section
Fig. 3.1 Axially loaded bar

Equilibrium The internal normal force, or axial force, acting in the z-direction is denoted by function N ( z ) ,
and N is positive if tensile and negative if compressive. See Fig. 3.2. If we consider an interior element of the bar

p z ( z * )dz
Q1 N(ε) N N + dN N(L – ε) Q2

ε→0 z dz
ε ε ε→0
z L
z < z * < z + dz
left end right end

Fig. 3.2 Free body diagrams for equilibrium of the bar

as shown in the center sketch of Fig. 3.2 and a positive normal force is defined 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. A free body diagram of a generic interior element of differential length dz is shown in the figure, where the
distributed force intensity acting on the element is evaluated at z* , where z < z* < z + dz . Axial force equilib-
rium of this element in the z-direction as dz → 0 , where in the limiting process dN → 0 and z * → z , gives the
following differential equation of equilibrium at z:
dN
------- + p z ( z ) = 0 0<z<L (3.1)
dz

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. 3.3. The axial
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 z = 0 are

44 Thin-Walled Structures
The tensile test

either w ( 0 ) = q 1 or N ( 0 ) = – Q 1 but not both (3.2)

and the boundary conditions at z = L are

either w ( L ) = q 2 or N ( L ) = Q 2 but not both (3.3)

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. 3.2.

Strain-displacement Extensional deformation of an element of the bar is shown in Fig. 3.3. Assuming the bar

y
z z + dz z + w(z) z + dz + w + dw
dz dz + dw
z - axis
w + dw
w(z)

Fig. 3.3 Extensional deformation of an element of the bar

does not fracture, the axial displacement function w ( z ) is a continuous function of z. The axial normal strain due
to extension is defined by

( dz + dw ) – dz dw
ε z ≡ Limit ------------------------------------ = -------
dz dz
dz → 0 (3.4)
The axial normal strain εz 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 ≤ z ≤ L, then the bar displaces as a
rigid body and the normal strain is zero, as is evident from eq. (3.4).

To complete the analysis for the response of the bar, we need to relate the internal axial force N to the axial
normal strain εz, 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.

3.2 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 exerted on the specimen and an exten-
someter, or electrical resistance strain gage, to measure the elongation/strain of the gage length. The load frame
may be servo-hydraulically controlled using a feedback system to maintain specified 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-

Thin-Walled Structures 45
Article 3.2

ally the load is increased monotonically in time until fracture of the specimen. In addition, servo-hydraulic load
frames can apply cyclic loads at frequencies of around10Hz for fatigue testing. Electrical signals 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 σ z in the gage section is defined as the axial force N divided by the original cross-
sectional area A of the specimen. The engineering normal strain ε z is defined as the elongation ∆L 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 definition given in eq. (3.4). Typical engineering stress-strain plots for a low carbon steel
and an aluminum alloy are shown in Fig. 3.4. For most engineering materials there is a linear relationship

L N
σ z = ----
A
N N

σz L ( 1 + εz ) σz

σu
σu
σ yu
fracture σy fracture
σp σ yl ( ε f, σ f ) σp ( ε f, σ f )

E
ε 0.2 = 0.002
1
εz εz
0 0
(a) some steels (b) aluminum alloy

Fig. 3.4 Typical stress-strain curves for steel and aluminum alloy from tensile tests

between the axial normal stress σ z and normal strain ε z near the origin of the plot as is shown in the figure. The
slope of this linear portion is a material property called the modulus of elasticity, or Young’s modulus, and it is
denoted by E. The value of the stress where the stress-strain plot deviates from a straight line is called the propor-
tional limit, and is denoted by σ p . The proportional limit is difficult 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 defined 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
difficult to measure precisely since the specimen must be unloaded and reloaded to determine it. Just after the lin-
ear region of the stress-strain plot for some low carbon steels, there is a relative maximum stress followed by a
relative minimum, 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 maximum value of the stress is called the upper yield point σ yu , and the stress at the subsequent rel-
ative minimum value is called the lower yield point σ yl . Values of the upper yield point for metals are sensitive to

46 Thin-Walled Structures
The tensile test

the loading rate and accidental bending stresses. Yielding is associated with plastic deformation of the material.
Plastic deformation is defined as deformation which is independent of time and remains on release of the load.
The principal physical mechanism causing plastic deformation in metals is slippage between planes of atoms in
the crystal grains of the material (Dowling, 1993).

Loading, unloading, and reloading of a metallic specimen beyond its yield point is depicted in Fig. 3.5. The

σz

σy ε 0B = ε 0A + ε AB

ε 0B = total strain
ε 0A = plastic strain
ε AB = elastic strain

0 εz
A B
Fig. 3.5 Loading and unloading a tensile test specimen beyond the yield stress

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 stress art yield, or
the yield strength. 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
σz
steels; i.e., there is no stress just after the linear elastic portion where the σ yield
stress-strain curve has a zero slope. See Fig. 3.4b. 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
E E
define an offset yield stress. A straight line is drawn parallel to the linear elas-
1 1
tic portion of the stress-strain curve starting from a strain ε = ε 0.2 = 0.002 εz
on the strain axis. The stress at the intersection of this straight line with the 0 0.002 ε yield
stress-strain curve is defined to be the yield strength σ yield of the material. Fig. 3.6 0.2% offset
Note that the strain of 0.002, or 0.2% (percent strain is defined as 100ε ), is yield strength
plastic strain, since unloading the specimen from the point ( ε yield, σ yield ) on
the stress strain-curve would follow the straight dashed line in Fig. 3.6 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 defined quantity. The offset yield stress
is generally the most satisfactory means of defining the yielding event for engineering materials.

Thin-Walled Structures 47
Article 3.2

After yielding the load may have to increase to cause further plastic defor-
mation of the specimen. (Also, the elastic deformation increases.) The increase
N N
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
necking deformation
area is called the ultimate tensile strength, or just the tensile strength, of the
material and is denoted by σ u . 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 σ f and the engineering strain at fracture is denoted by ε f . Recall that the
engineering stress is defined with respect to the original cross-sectional area.

In addition to the axial elongation of the bar in the tensile test, most engi-
ε x or ε y σ neering materials exhibit a lateral contraction of the cross section. In axial
ε z = -----z
E compression, the cross section expands. In the linear elastic range of mate-
0 rial behavior, the ratio of the lateral normal strain to the axial normal strain
εz
1 is found to be a constant called Poisson’s ratio. Poisson’s ratio is denoted
ν by ν, and it is a dimensionless quantity. See Fig. 3.7. It usually 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
Fig. 3.7 Lateral normal denoted by ε x and ε y . That is, measurements of the diameter of the test
strains in the tension test
specimen in the tensile test reveal that the normal strain in the x- and y-
directions are
ε x = ε y = – νε z

The negative sign is introduced to make the Poisson’s ratio positive for lateral contraction under uniaxial exten-
sion. It is implicit in the above expression that the Poisson’s ratio is the same along the x- and y-directions. If a
material is isotropic, then the Poisson’s 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 σ z to the three nor-
mal strains is
ε x = – νσ z ⁄ E ε y = – νσ z ⁄ E εz = σz ⁄ E (uniaxial loading only) (3.5)

48 Thin-Walled Structures
Effect of temperature on strain

Typical values of the modulus of elasticity and Poisson’s ratio are listed in the table below for selected metals.

Room temperature modulus of elasticity and Poisson’s ratios for selected metals and metal
alloys (from Callister, 1997)

Modulus of elasticity
Poisson’s ratio,
Material 106 psi GPa dimensionless
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

3.3 Effect of temperature on strain


In the elastic region of material behavior changes in temperature can cause two effects:
• The elastic constants (E and ν) 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 ε t . 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 ε zt .
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 T 0 to T , then the thermal strain is approximated by

ε zt = α ( T – T 0 ) (3.6)

in which α is called the linear coefficient of thermal expansion. The coefficient of thermal expansion is abbrevi-
0 0
ated as CTE. The dimensional units of α are 1 ⁄ F or 1 ⁄ C , since strain is a dimensionless quantity. Values of
the linear coefficient 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,

σ
ε z = -----z + α ( T – T 0 )
E





{

mechanical thermal (3.7)

Thin-Walled Structures 49
Article 3.4

Room temperature linear coefficients of thermal expansion (from Callister, 1997)

Coefficients of thermal expansion

0 0
Material 10 –6 ⁄ F 10 –6 ⁄ C
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

3.4 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 σ z is uniform over
the cross section.With respect to an arbitrary coordinate system x , y in the cross section, this uniform normal
stress distribution is statically equivalent to an axial force N and a moment with components M x and M y as
shown in Fig. 3.8.

y y
y

yc
Mx C

σ z dA
statically equivalent
N
N x
xc
My
x x

Fig. 3.8 Statical equivalence of the normal stress to the bar resultants.

The relationships of static equivalence are

N =
∫ ∫ σ dA
A
z Mx =
∫ ∫ yσ dA
A
z My =
∫ ∫ xσ dA
A
z (3.8)

50 Thin-Walled Structures
Bar reference axis

Substitute the material law for the axial normal stress from eq. (3.7) into these conditions of statical equiva-
lence to get

N =
∫ ∫ [E(ε
A
z – α ( T – T 0 ) ) ] dA

Mx =
∫ ∫ y[E(ε
A
z – α ( T – T 0 ) ) ] dA My =
∫ ∫ x[E(ε
A
z – α ( T – T 0 ) ) ] dA

Since the bar is homogenous, the material properties are independent of the cross-sectional coordinates. Also, the
extensional strain ε z 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

N = EA [ ε z – α ( T – T 0 ) ]
(3.9)
Mx = Qx [ E ( εz – α ( T – T0 ) ) ] My = Qy [ E ( εz – α ( T – T0 ) ) ]

where Q x denotes the first moment of the cross-sectional area about the x -axis and Q y denotes the first area
moment about the y -axis. These first area moments are defined by the integrals

Qx =
∫ ∫ y dA
A
Qy =
∫ ∫ x dA
A
(3.10)

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 while maintaining static equivalence. The coordinates ( x c, y c ) of the point in
the cross section where the moments due to a uniform distribution of axial normal strain vanish is called the cen-
troid. Static equivalence yields the resultants at the centroid as

N = N
Mx = Mx – yc N = 0 (3.11)

My = My – xc N = 0
Solving the last two of eqs. (3.11) for the coordinates of the centroid, and using eqs. (3.9) we get

xc = Qy ⁄ A yc = Qx ⁄ A (3.12)

The location of the centroid depends on distribution of the area of the section about the x - y coordinate system.
The origin of the parallel Cartesian system x, y is at the centroid of the cross section as is shown in Fig. 3.8. The
two coordinate systems are related by

x = xc + x y = yc + y (3.13)

The first area moments have dimensional units of L3. Equations (3.12) 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. If eqs.
(3.13) are substituted into (3.10),which are in turn substituted into eqs. (3.12), we find

Thin-Walled Structures 51
Article 3.5

1 1
x c = x c + ---
A ∫ ∫ x dA
A
y c = y c + ---
A ∫ ∫ y dA
A

These last equations imply that the first moments of the cross-sectional area about the centroidal x and y axis sys-
tem vanish; i.e.,

Qx =
∫ ∫ y dA = 0
A
Qy =
A
∫ ∫ x dA = 0
From the first of eqs. (3.9) and the first of eqs. (3.11), the material law for the axial force is given by
t
N = EA ( ε z – ε zt ) = EAε z – N (3.14)

where the product the modulus of elasticity and the cross-sectional area, or EA, is called the axial stiffness, or
extensional stiffness, of the bar, and N t denotes the so-called thermal force. The thermal axial force is defined by
t
N = EAε zt = EAα ( T – T 0 ) (3.15)

Note that eq. (3.14) 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
( q 1 = q 2 = 0 ), the distributed load intensity vanishes ( p z ( z ) = 0 ), and there is a spatially uniform change in
temperature from the stress free state, then N = – N t for 0 ≤ z ≤ L .

3.5 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 specified loads: equilibrium conditions, strain-displacement relations, and material laws. For the uniaxially
loaded bar problem, we assume that it is made of a linear elastic, isotropic, homogeneous material, and summa-
rize these fundamental equations as follows.
dN
– equilibrium, eq. (3.1): ------- + p z ( z ) = 0 0<z<L
dz
dw
– strain-displacement, eq. (3.4): ε z = -------
dz
t
– material law, eq. (3.14): N = EAε z – N

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. (3.4) for
the strain in eq. (3.14), and in turn substitute the resulting equation into the equilibrium eq. (3.1) for the axial
force to get

d  dw
EA = –pz ( z ) 0<z<L (3.16)
dz dz 
where we used dN t ⁄ dz = 0 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. (3.16) becomes

52 Thin-Walled Structures
Linear elastic response

2
dw
EA 2 = – p z ( z ) 0<z<L (3.17)
dz

Equation (3.16) is the governing linear, 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 deter-
mine the two constants that arise in the general solution of eq. (3.16). These boundary conditions are given in
eqs. (3.2) and (3.3). Using the material law and the strain displacement equations these boundary conditions
become
dw
either w ( 0 ) = q 1 or EA ------- – N t = – Q 1 but not both (3.18)
dz z=0

dw
either w ( L ) = q 2 or EA ------- – N t = Q 2 but not both (3.19)
dz z=L

EXAMPLE 3.1 Axial bar with a specified uniform distributed load and specified end displacements

Consider a linear elastic, uniform bar subjected to a uniformly dis-


T – T0 ≠ 0
tributed load with intensity p z = p 0 , specified end displacements q1 p0
and q2, and a uniform change in temperature T - T0 from the stress free
state as shown in Fig. 3.9. Determine the axial displacement, normal
strain, and normal force of the bar. z, w L

Solution The governing differential equation (3.17) for the displace- w ( 0 ) = q1 w ( L ) = q2


ment reduces to Fig. 3.9 Bar of Example 3.1
2
dw p
= – ------0- 0<z<L
dz2 EA
Integrate this equation twice with respect to z to get
p z2
w ( z ) = – ------0-  ---- + c 1 z + c 2
EA  2 
where c1 and c2 are constants obtained in the indefinite integration. Substitute this expression into the boundary
condition at the left end and right end to get
p 02 p L2
– ------0-  ----- + c 1 0 + c 2 = q 1 – ------0-  ----- + c 1 L + c 2 = q 2
EA  2  EA  2 
Solve these last two equation for constants c1 and c2 to find

q2 – q1 p0 L
c 1 = ---------------
- + ----------- c2 = q1
L 2EA
Substitute these results for the constants of integration in the solution for the axial displacement to get

z z p0 2
w ( z ) = q 1  1 – --- + q 2 --- – ----------
- ( z – Lz ) (3.20)
 L  L 2EA

Thin-Walled Structures 53
Article 3.5

The axial strain in the bar is defined by eq. (3.4), and substituting eq. (3.20) for w(z) we get
( q2 – q1 ) p0
ε z = -------------------- - ( 2z – L )
- – ---------- (3.21)
L 2EA
Finally, substitute the strain given by eq. (3.21) into eq. (3.14) to find the distribution of the axial force as

EA L
N ( z ) = ------- ( q 2 – q 1 ) – p 0  z – --- – N t (3.22)
L  2

where N t = EAα ( T – T 0 ) . Equations (3.20) to (3.22) constitute the complete solution for the linear thermoelas-
tic response of the uniform bar subject to a uniform distributed load intensity p0, uniform temperature change
T – T 0 , and end displacement q1 and q2.

We can use eq. (3.22) and the relations N ( 0 ) = – Q 1 and N ( L ) = Q 2 (refer to eqs. (3.2) and (3.3)), to get

EA p0 L t
Q 1 = – ------- ( q 2 – q 1 ) – --------
-+N (3.23)
L 2

EA p0 L
- – Nt
Q 2 = ------- ( q 2 – q 1 ) – -------- (3.24)
L 2
Addition of these results gives
Q1 + Q2 + p0 L = 0

which is the condition of overall axial force balance for the bar.

EXAMPLE 3.2 A bar with fixed ends and subjected to an axial point force.

F Determine the axial displacement and normal force for the uniform bar
subjected to a point force F as shown in Fig. 3.10. Each end of the bar is
z fixed to a rigid support. There is no change in temperature from the stress
a b free state, and the specified data are, the axial stiffness EA, dimensions a, b,
L and L, and F.
Fig. 3.10 Point force acting on
a clamped bar 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 (3.17) 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 infinite value of the distributed load intensity pz acting over a zero length.
Hence, we will need to solve the governing differential equation separately in each open interval. That is
2
dw
EA = 0 0<z<a and a<z<L (3.25)
dz2
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 fixed, we have
w(0) = 0 w(L) = 0 (3.26)

The displacement of the bar must be continuous at z = a, so

54 Thin-Walled Structures
Linear elastic response

w( a- ) = w( a+ ) (3.27)

where a - implies the limit of z → a for values of z < a , and a + implies the limit of z → a for values of z > a .
The last boundary condition comes from axial force equilibrium at the point of application of the force.
dw dw
– N( a- ) + N( a+ ) + F = 0 or – EA ------- + EA ------- +F = 0 (3.28)
dz z → a-
dz z → a+

The solution to eqs. (3.25) is


w ( z ) = c1 z + c2 0<z<a
w ( z ) = c3 z + c4 a<z<L

where c1, c2, c3, and c4 are constants to be determined from the four boundary conditions. From the two bound-
ary conditions given by eqs. (3.26), we find c 2 = 0 and c 4 = – Lc 3 . Using these results in the continuity condi-
tion given by eq. (3.27) we get
c1 a = c3 ( a – L )

and in the force balance given by eq. (3.28) we get


– EAc 1 + EAc 3 + F = 0

These last two equations can be solved for the constants c1 and c3, and the solution is c 1 = ( bF ) ⁄ ( EAL )
c 3 = – ( aF ) ⁄ ( EAL ) , where we used b = L - a. Hence the displacement function is

bF
 ---------- -z 0≤z≤a
 EAL
w(z) =  (3.29)
aF
 ----------
-(L – z) a≤z≤L
 EAL

The normal force in each interval is obtained from N = EA ( dw ⁄ dz ) , so we have

 --b- F 0≤z≤a
L
N(z) =  (3.30)
 – --a- F a≤z≤L
 L
The displacement and normal force are plotted with respect to the axial coordinate in Fig. 3.11 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-
a b
tinuity in the normal force is N ( a + ) – N ( a - ) =  – --- F –  --- F = – F , which satisfies the first of eqs. (3.28).
 L  L 

Thin-Walled Structures 55
Article 3.6

w
abF
-----------
EAL

z
0
a L
N Fig. 3.11 Displacement and normal force
b distributions for the clamped bar
--- F subjected to a point force.
L

z
0 a L
a
– --- F
L

3.6 References
Callister, W.D., 1997, Materials Science and Engineering, Fourth Edition, John Wiley & Sons, Inc., New York,
pp. 777-779, 788, 789.

Dowling, N.E., 1993, Mechanical Behavior of Materials, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, p.
110.

3.7 Problems
1. Take the bar of Example 3.1 to have a solid circular cross section with a diameter of 0.50 inches, and take its
6
length L = 20 inches. The material is aluminum alloy with a modulus of elasticity E = 10 ×10 psi , yield
3
strength in tension and compression of magnitude σ yield = 47 ×10 psi , and thermal coefficient of expansion
–6
α = 13 ×10 /°F .

a) For the bar subjected to the loads p 0 = 265 lb/in. , q 1 = 0 in. , q 2 = 0.0060 in. , and
T – T 0 = 100 °F , 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 p 0 = ( 265 lb/in. ) λ , q 1 = 0 in. , q 2 = ( 0.0060 in. )λ , and
T – T 0 = ( 100 °F )λ , where λ is a dimensionless load factor. Determine the largest positive value of λ
such that normal stress σ z is between the limits – σ yield ≤ σ z ≤ σ yield and the maximum displacement

56 Thin-Walled Structures
Problems

w max 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 (900) by the wrenches, estimate the tensile force
6
in the bolt if the bolt and sleeve are of steel ( E = 29 ×10 psi ), and the cross-sectional areas are: Bolt area = 1.00
in2, Sleeve area = 0.60 in2.

6 in.

3. A 7 m long AH-1W Supercobra helicopter blade is rotating at 300 rpm and has a mass of 300 kg. Centrifugal
forces due to the rotation of the blade lead to tension in the blade. Plot the distributed axial force intensity and the
internal axial force distribution on the blade. Calculate the stress at the root for a blade cross-sectional area of
0.02 m2. (Assume that the mass is evenly distributed and the center of mass of the cross-section coincides with
the tension axis)

4. A Saturn V rocket carrying an Apollo spacecraft is fired from ground. The launch vehicle is 110 m long, has
a total take-off mass of 27 million kilograms and the thrust generated at take-off is 340 million Newtons. Plot the
internal axial force distribution at take-off as a function of the length of the rocket. (Assume that the mass is
evenly distributed along the length of the launch vehicle)

Thin-Walled Structures 57
Article 3.7

58 Thin-Walled Structures
CHAPTER 4 Axial normal stress in
pure bending and
extension

In the first four sections the flexure 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 flexure formula (eq. (4.27)) 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. (4.79) to (4.81).

4.1 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 α as shown Fig.
4.1. 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
M x = M cos α and M y = M sin α . The sign convention for positive bending moments Mx and My is such that
they tend to cause elongation of the fibers parallel to the z-axis that have positive x and y coordinates (first quad-
rant of the x-y plane in the cross section). Thus Mx is positive if directed along the positive x-axis by the right-
hand screw rule, and My is positive if directed along the negative y-axis by the right-hand screw rule.

Thin-Walled Structures 59
Article 4.2

y
α
plane of loading
y

M sinα A x Mx
M cosα
z α
A
M cosα
M sinα
M
My
Section A-A
Fig. 4.1 A uniform beam subjected to equal and opposite couples at each end.

4.2 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.

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.

Plane of bending The plane in which the beam bends is denoted the y – z plane where the y -axis is inclined by
the angle β from the y-axis as shown in Fig. 4.2. Note that the plane of bending is not assumed to coincide with
the plane of loading; i.e., α ≠ β. Coordinate axes x and y are rotated through the angle β relative to the x and y
coordinate axes as shown in the figure. A point in the cross section can be located by x and y coordinates and by
x and y coordinates. Consider a typical point A with coordinates x and y. The x and y coordinates of point A are
determined by the geometry shown in Fig. 4.3. From Fig. 4.3, the relationship between these coordinates is

x = x cos β – y sin β
(4.1)
y = x sin β + y cos β
The cosines of the angles between the positive coordinate axes, or direction cosines, can be written in tabular
form as shown in Fig. 4.2. For example, the direction cosine of the angle between the x -axis and the y-axis is
cos ( 90° + β ) = – sin ( β ) . Note that the sum of the squares of the direction cosines in a row or column equals
unity. Equations (4.1) show that the rows of the direction cosine table are associated with the x - y coordinates.
The inverse of eqs. (4.1) is

60 Thin-Walled Structures
Geometry of deformation

y
α
y
β direction cosines
x x y
α x cos β – sin β
plane of bending
β y sin β cos β
M
plane of loading
x

Fig. 4.2 Projection of the plane of loading and the plane of bending onto
the cross section of the beam.

y x sin β
90°

y cos β β A
y Fig. 4.3 Point A located in two Cartesian
90° β coordinate systems, where one system is
rotated through angle β relative to the other.
β
90°

β x

x 90°

y sin β
x cos β

x = x cos β + y sin β
y = – x sin β + y cos β
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 ρ in
the plane of bending, which we call the z-curve. See Fig. 4.4. If the beam is considered to be made up of longitu-
dinal fibers, then fibers parallel to the z-axis in the undeformed beam deform into concentric circular arcs. Some
of these parallel fibers are elongated and some are shortened. Consequently, the fibers 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

Thin-Walled Structures 61
Article 4.2

undeformed beam. We will show later that the neutral surface in the undeformed beam coincides with the x -z
plane.

z-curve in the neutral surface.

z
ρ

Fig. 4.4 Beam deformed into a circular arc in the plane of bending.

)
An element in the undeformed beam of length ∆z bounded by two parallel line segments PQ and RS is
)

shown in Fig. 4.5. Line PQ is part of the z-axis and line RS is characterized by a constant y -coordinate value.

–θ
– θ + ∆θ

y R* y
P* S*
R S ρ ∆θ
θ y Q*
z P Q
z ∆z

Fig. 4.5 Deformation of an element in the plane of bending.


)

In the deformed geometry these line segments map to concentric circular arcs P * Q * and R * S * . Let θ denote the
rotation of a line element originally parallel to the y -direction in the undeformed beam, positive clockwise.
Since cross sections remain perpendicular to the neutral surface in the deformed beam, the angle θ is also equal
)

to the rotation of the line element coinciding with z-axis. In Fig. 4.5, the line P * R * is shown to rotate counter-
)

clockwise which corresponds to a – θ , and line Q * S * is shown to rotate clockwise through angle – θ + ∆θ .
)

Then, the included angle between sections P * R * and Q * S * is ∆θ. The radius of the arc P * Q * is ρ as described
in the previous paragraph. It is assumed that the distortion of the cross section in the plane of bending is negligi-

62 Thin-Walled Structures
Geometry of deformation

)
ble, so that the coordinate of circular arc R * S * with respect to arc P * Q * remains equal to y . Hence, the radius
)

)
of arc R * S * is ρ + y . The strain of line element RS is defined as

)
R * S * – RS
ε z = ------------------------ (4.2)

)
RS
From Fig. 4.5

)
R * S * = ( ρ + y )∆θ (4.3)

)
)

)
and since line element P * Q * is in the neutral surface P * Q * = PQ = R S = ρ∆θ . Thus, eq. (4.2) becomes
)

)
R*S* – P*Q* ( ρ + y )∆θ – ρ∆θ y
ε z = -------------------------------- = ----------------------------------------- = --- (4.4)
ρ∆θ ρ
)

P*Q*
)

)
Also, we have P * Q * = ρ∆θ = ∆z , since the line element P * Q * is in the neutral surface. From this relation we
have in the limit as ∆z → 0 and ∆θ → 0 , the following equation for the curvature of the z-curve in the deformed
beam
1 dθ
--- = ------ = θ′ (4.5)
ρ dz
where the prime denotes a derivative with respect to z. Equation (4.5) is consistent with the definition of curva-
ture as the change in the slope angle of the curve with respect to its arc length. Hence,

ε z = yθ′ (4.6)

Equation (4.6) 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 ( y > 0 ) are stretched ( ε z > 0 ) if the beam is bent con-
vex side up ( 1 ⁄ ρ > 0 ). Fibers below the z-axis are shortened ( ε z < 0 ) if the beam is bent convex side up (Fig.
4.4).

Projection of the z-curve onto the x-z and y-z planes Now we need to express the normal strain given by eq.
(4.6) in terms of the x-y coordinate system. This is accomplished by projecting the z-curve in the deformed beam,
which lies in the y – z 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 u ( z ), v ( z ), and w ( z ) 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 ( 0, 0, z ) displaces to the point P* with coordinates ( u ( z ), v ( z ), z + w ( z ) ) . See
Fig. 4.6.

Also the displacements of the particle at P can be represented in the x - y -z coordinate system. Let u ( z )
v ( z ) denote the displacements in the x - and y -directions, respectively, of the particle at point at P. These dis-
placement components are related to components u ( z ) and v ( z ) 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 P 3* in Fig. 4.6 and Fig.

Thin-Walled Structures 63
Article 4.2

v(z)

P 1*
P: ( 0, 0, z ) P 3*
θx x

P*
w(z)
u(z)

Q*
z P 2*
θy

Fig. 4.6 Displacements of the particle from P to P*, and the rotations of the
projections of a line element P*Q*.

4.7. From Fig. 4.7, the displacement components in the x - y coordinates are related to those in the x-y coordi-

y v

v P 3*
y
β

x u
P
plane of bending
u
N.B. u = 0

Fig. 4.7 Lateral displacement components of the particle at point P.

nates by

u = u cos β – v sin β
(4.7)
v = u sin β + v cos β
which is of the same form as the coordinate transformations given by eq. (4.1) with the coordinates replaced by
their respective displacements. Since the z-curve lies in the y – z plane, displacement u = 0 for all z. That is,
point P 3* is in the y -z plane. Also, the angle β 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

64 Thin-Walled Structures
Geometry of deformation

bending to the rotations its projections onto the x-z and y-z planes.

)
)
The displacement and rotation of an infinitesimal line element PQ on the z-axis to P * Q * on the z-curve is
shown in Fig. 4.8. Since the line element does not change length we have by the Pythagorean theorem in the

P * dz + dw
v w θ
– dv
dz
Q*
v + dv w + dw

P Q
z
z z + dz z* z * + dz *

Fig. 4.8 Displacement and rotation of an infinitesimal line element on the z-axis
in the plane of bending.
2
deformed state that ( dz ) 2 = ( dz + dw ) 2 + ( – dv ) . Hence, in the limit as dz → 0 , the derivative of the axial dis-
placement is related to the derivative of the lateral displacement by
2
( 1 + w′ ) 2 + ( – v′ ) = 1 (4.8)

As is shown in Fig. 4.8, the rotation, or slope, of the z-curve is related to displacement v by

dv
θ ≅ sin θ = – ------ = – v′ (4.9)
dz
where we assume the angle θ is small so that the sine of it can be replaced by the angle itself in radians. Differ-
entiating eqs. (4.7) with respect to z, and noting that u = 0 and the angle β is constant for all z, we get

0 = u′ cos β – v′ sin β
(4.10)
– θ = u′ sin β + v′ cos β

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 P 1* , and its projection onto the x-z plane is labeled P 2* , as is shown in Fig. 4.6. Points P 1*
and P 2* and the projections of the differential line element along the z-curve onto the coordinate planes shown in
Fig. 4.6 are also shown in Fig. 4.9. From the differential geometry of Fig. 4.9, we have in the limit as dz → 0
that
– v′ – u′
tan θ x = --------------- tan θ y = --------------- (4.11)
1 + w′ 1 + w′
Since v′ is assumed small with respect to unity, eq. (4.8) yields that w′ is also small with respect to unity. Thus,
we neglect w′ with respect to one in the denominators of eqs. (4.11), and approximate the rotations of the projec-
tions as

Thin-Walled Structures 65
Article 4.2

y x

P 1* dz + dw P 2* dz + dw
θ x – dv θ y – du
v u
v + dv u + du
z z
z* z * + dz * z* z * + dz *

projection in y-z plane projection in x-z plane

Fig. 4.9 Rotations of the projections of the z-curve onto the cartesian planes.

θ x = – v′ θ y = – u′ (4.12)

where the tangent of a small angle is approximated by the angle itself in radians. Thus, eqs. (4.10) become
0 = – θ y cos β + θ x sin β
(4.13)
θ = θ y sin β + θ x cos β
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
θ y = θ sin β θ x = θ cos β (4.14)

Equations (4.14) 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. (4.6) is written in terms of the x-y system by
substituting the second of eqs. (4.1) for y , and the derivative of the second of eqs. (4.13) for θ′ , to get

ε z = ( x sin β + y cos β ) ( θ′ y sin β + θ′ x cos β )

Rearrange this equation to


ε z = x ( θ′ y sin2 β + θ′ x sin β cos β ) + y ( θ′ y cos β sin β + θ′ x cos2 β )

Differentiating the first of eqs. (4.13), recalling that angle β is independent of z, results in the relation
θ′ y cos β = θ′ x sin β

Now use this result in the previous equation to get the strain expressed as
ε z = xθ′ y + yθ′ x (4.15)

where the curvatures of the projected curves are determined from eqs. (4.12) to be
θ′ y = – u″ θ′ x = – v″ (4.16)

66 Thin-Walled Structures
Bending normal stress — flexure formula

4.3 Bending normal stress — flexure formula


The strain ε z is related to the normal stress σ z acting on the cross section by the y
material law, and we consider the beam to be made from a linear elastic, isotropic σy
material. For uniaxial loading only, this material law is given by eq. (3.5) on p. 48 . σx
Under multi-axial loading the material law is obtained by superpostion of the normal
strains ε x, ε y, and ε z for line elements parallel to the x-, y-, and z-directions due to σz x
z
the combined action of normal stresses σ x, σ y, and σ z acting on faces of an infinites-
imal element normal to to the x-, y-, and z-directions. Superposition of normal strains using eq. (3.5) on p. 48
yields

1 ν ν
--- – --- – ---
εx E E E σ
x
= – --ν- 1 ν (4.17)
εy --- – --- σ y
E E E
εz σz
ν ν 1
– --- – --- ---
E E E
where E is the modulus of elasticity and ν is Possion’s ratio. The material law for a linear elastic, isotropic mate-
rial is called Hooke's law. Material properties E and ν 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 4.6. However, in this section we
limit consideration to a beam made of a single homogeneous material, so that E and ν are independent of the spa-
tial coordinates x, y and z. Now we make the third basic assumption of beam theory:

Assumption regarding the material law


3. Lateral normal stresses σ x and σ y are assumed small with respect to the axial normal
stress σ z , and hence are neglected in Hooke’s law.

Neglecting stress components σ x and σ y , Hooke's law becomes

σ z = Eε z (4.18)

Eliminating the strain in eq. (4.18) via eq. (4.15) we get


σ z = E ( xθ′ y + yθ′ x ) (4.19)

The distribution of normal stresses given by eq. (4.19) over the cross section is, in general, statically equiva-
lent to the resultants N, Mx, and My as shown in Fig. 4.10. The axial load N, and bending moments Mx and My
are related to the normal stress by

N =
∫ ∫ σ dA
A
z Mx =
∫ ∫ yσ dA
A
z My =
∫ ∫ xσ dA
A
z (4.20)

where dA = dxdy is an area element in the cross-sectional area denoted by A. Substituting eq. (4.19) into eq.
(4.20) and integrating over the cross section we obtain

Thin-Walled Structures 67
Article 4.3

y
dA ⇔ Mx
x N
z σz z
My
x

Fig. 4.10 Statical equivalence of the normal stress to the beam resultants.

N = EQ x θ′ x + EQ y θ′ y
M x = EI xx θ′ x + EI xy θ′ y (4.21)

M y = EI xy θ′ x + EI yy θ′ y

in which we defined

Qx =
∫ ∫ y dA
A
Qy =
∫ ∫ x dA
A
(4.22)

and

I xx = 2 2
∫ ∫ y dA
A
I yy =
∫ ∫ x dA
A
I xy =
∫ ∫ xy dA
A
(4.23)

In obtaining the results shown in eqs. (4.21), 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. (4.22) and
(4.23) are purely geometric quantities. They reflect how the area is distributed with respect to the x- and y-coordi-
nate axes. The quantities Qx and Qy in eqs. (4.22) are called first area moments since the power of the coordi-
nate in the integrand is unity, while the Ixx and Iyy in eqs. (4.23) are called second area moments since the power
of the coordinates in the integrand is two. The quantity Ixy is called the product area moment.

For the case of pure bending, only moments Mx and My are non-zero, so the axial force N must vanish. Set-
ting N = 0 in the first of eqs. (4.21) implies that
( N = 0 ) ⇒ Qx = Qy = 0 (4.24)

The first 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. (This result
was also obtained in Chapter 3 for a bar subject to axial loads only. Refer to discussion prior to eq. (3.14) on p. 52
.) For centroidal coordinates in the cross section, the moment-curvature relationship of eq. (4.21) can be written
in the matrix form

68 Thin-Walled Structures
Bending normal stress — flexure formula

Mx I xx I xy θ′ x
= E (4.25)
My I xy I yy θ′ y

where the second area moments are computed for centroidal coordinates in the cross section. The curvature-
moment relationship is the inverse of this, or

θ′ x 1 I – I xy M x
2 )
- yy
= --------------------------------- (4.26)
θ′ y E ( I xx I yy – I xy – I xy I xx M y

Equations (4.25) and (4.26) 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 σz in eq. (4.19) can be written in terms of the bending moments
Mx and My by using the curvature-moment relation in eq. (4.26). Substituting the curvatures from eq. (4.26) into
eq. (4.19) we get

– I xy M x + I xx M y I yy M x – I xy M y
σ z =  ------------------------------------
- x +  --------------------------------
- y (4.27)
 I I – I2   I I – I2 
xx yy xy xx yy xy

or we can re-write this as

I yy y – I xy x  I xx x – I xy y 
σ z =  ------------------------
- M +  ------------------------
- M (4.28)
I I – I2  x I I – I2  y
xx yy xy xx yy xy

Equation (4.27), or in its alternate form (4.28), is called the flexure 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 Mx and
My by first differentiating eqs. (4.14) with respect to z and recognizing that the z-curve lies in the plane of bend-
ing (angle β is independent of z for pure bending). Second, solve the resulting equations for β to get
θ′
tan β = -------y (4.29)
θ′ x
Now use eq. (4.26) to eliminate the curvatures in eq. (4.29) to find

– I xy M x + I xx M y
tan β = -------------------------------------
- (4.30)
I yy M x – I xy M y

Thus, knowing the bending moment components and the second area moments of the cross section, the angle β,
and hence the plane of bending, is determined by eq. (4.30). The axis normal to the plane of bending through the
centroid is labeled as the x axis in Fig. 4.2. On the x axis y = 0 , so that from the strain-curvature relation given
by eq. (4.6) the normal strain ε z = 0 on the x axis. Also, from Hooke’s law this means that the bending normal

Thin-Walled Structures 69
Article 4.3

stress σ z = 0 on the x axis. Since ε z = 0 on the x axis, the x axis is called the neutral axis . The particles in

x -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 y = 0 in the first of eqs. (4.1). We get

0 = ( x sin β + y cos β ) NA
or y NA = – x NA tan β (4.31)

where the tangent of β is determined by eq. (4.30).

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 N ≠ 0 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. (4.31).
From Fig. 4.1, recall that the plane of loading is located by
M
tan α = -------y
Mx

Using this result in eq. (4.30), we get


– I xy + I xx tan α
tan β = ----------------------------------- (4.32)
I yy – I xy tan α

Equation (4.32) shows that, in general, the plane of loading and the plane of bending do not concide ( α ≠ β ). The
plane of loading and the plane of bending coincide if I xy = 0 and M x = 0 , or I xy = 0 and M y = 0 , or
I xy = 0 and I xx = I yy . For I xy = 0 and M x = 0 the angles α = β = 90° . For I xy = 0 and M y = 0 the
angles α = β = 0° , and for I xy = 0 and I xx = I yy the tan α = tan β .

EXAMPLE 4.1 Bending normal stress distribution in a cantilever beam with a thin-walled zee section.

The cantilever beam shown in Fig. 4.11 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

8 2
I xx = --- a 3 t I yy = --- a 3 t I xy = – a 3 t (4.33)
3 3
determine the neutral axis in the cross section and the distribution of the bending normal stress.

Solution First, note that the components of the specified bending moment are M x = – M and M y = 0 . Thus,
the plane of the couple whose moment is M is the y-z plane, or the angle α = π for the plane of loading shown
in Fig. 4.1. From eq. (4.30)
– I xy –( –a 3 t ) 3
tan β = --------
- = ------------------ = --- so β = 56.31°
I yy 2 3 2
--- a t
3

70 Thin-Walled Structures
Bending normal stress — flexure formula

y y
a

M
z centroid
a
C x

t, typical a

a
Fig. 4.11 Pure bending of a cantilevered beam with a zee cross section
Cross Section

3
and from eq. (4.31) the equation of the neutral axis in the cross section is y NA = – --- x NA .
2

The flexure formula, eq. (4.28), for this example becomes

 2--- a 3 t y – ( – a 3 t )x
3  M
σ z = --------------------------------------------------------- ( – M ) = – ( 6y + 9x ) ---------
-
3t
8 2
 --- a 3 t  --- a 3 t – ( – a 3 t ) 2 7a
3  3 

We plot the bending normal stress on the contour only. Coordinates x and y are related on the contour. That is,

top flange: –a ≤ x ≤ 0 y = a

web: x = 0 –a ≤ y ≤ a
bottom flange: 0≤x≤a y = –a

The neutral axis and the bending normal stress distribution are shown in Fig. 4.12.

Thin-Walled Structures 71
Article 4.3

6M
– --- ------
-
3M 7 a2t
--- ------
-
7 a2t
y

x Fig. 4.12 Bending normal stress


distribution along the contour of the
M zee section

σz
3M
– --- ------
-
7 a2t

6M 3
--- ------
- y NA = – --- x NA
7 a2t 2

EXAMPLE 4.2 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 4.1.

Solution First note that this a statically determinate problem. So the equilibrium equations are satisfied for
M x = – M and M y = 0 for all z ∈ ( 0, L ) , where M denotes the specified end moment. To find the lateral dis-
placements u ( z ) and v ( z ) , 0 ≤ z ≤ L , we begin with the curvature-moment relations given by eqs. (4.26). These
equations are repeated below.

θ′ x 1 I – I xy M x
- yy
= ---------------------------------
2
θ′ y E ( I xx I yy – I xy ) – I xy I xx M y

From equilibrium we have M x = – M and M y = 0 for 0 ≤ z ≤ L . From the given data for the second area
8 2
moments we know I xx = --- a 3 t , I yy = --- a 3 t , and I xy = – a 3 t . Thus,
3 3

2 = 8
det ( I ) = I xx I yy – I xy
2 7
--- × --- – ( – 1 ) 2 a 6 t 2 = --- a 6 t 2 (4.34)
3 3 9
Substitute these data into curvature-moment relations above to get
6 M 9 M
θ′ x = – --- ----------
- θ′ y = – --- ----------
- (4.35)
7 Ea 3 t 7 Ea 3 t

Integrate these equations with respect to z once to get

72 Thin-Walled Structures
Bending normal stress — flexure formula

6 M  9 M 
θ x = – ---  ----------
- z + c1 θ y = – ---  ----------
- z + c2 (4.36)
7  Ea 3 t 7  Ea 3 t
The constants due to indefinite 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 θ x ( 0 ) = 0
and θ y ( 0 ) = 0 . Substituting these boundary conditions into eqs. (4.36), we find c 1 = 0 and c 2 = 0 . Hence,
the rotations are

6 M  9 M 
θ x = – ---  ----------
- z θ y = – ---  ----------
- z 0≤z≤L (4.37)
7  Ea 3 t 7  Ea 3 t

The rotation-displacement relations are given in eqs. (4.12), and are repeated below.
θ x = – v′ θ y = – u′

Substitute these relations into eqs. (4.37) to get

6 M  9 M 
v′ = ---  ----------
- z u′ = ---  ----------
- z (4.38)
7  Ea 3 t 7  Ea 3 t
Integrate these equations with respect to z to get

3 M  2 9 M  2
v = ---  ----------
- z + c3 u = ------  ----------
- z + c4 (4.39)
7  Ea 3 t 14  Ea 3 t
The constants of indefinite integration are evaluated by the boundary conditions. Since the beam is fixed to the
rigid wall at z = 0, we have v ( 0 ) = 0 and u ( 0 ) = 0 . Substitute these boundary conditions into eqs. (4.39) to
find that c 3 = 0 and c 4 = 0 . Hence the lateral displacements of the beam are

3 M  2 9 M  2
v = ---  ----------
- z u = ------  ----------
- z 0≤z≤L (4.40)
7  Ea 3 t 14  Ea 3 t

A plot of the z-axis in the deformed beam, or z-curve, is shown in Fig. 4.13. In the plot the scaled lateral dis-
placements are defined by

Thin-Walled Structures 73
Article 4.3

Ea 3 t Ea 3 t
v = v  ----------2- u = u  ----------2-
 ML   ML 

z-curve

0.4
0.3

v
0.2
0.1
0 0.6
0
0.4
0.25
è
uu
0.5 0.2
zêzL 0.75
0
1

Fig. 4.13 The z-axis in the deformed, zee section beam .

The view of the lateral displacements at the end of the beam are shown in Fig. 4.14.

y
y
M
β = 56.3°

v(L)

x
u(L)

Fig. 4.14 Displacement of the tip of the cantilevered, zee section beam.

74 Thin-Walled Structures
Combined extension and pure bending

4.4 Combined extension and pure bending


The axial normal strain for extension and pure bending of a bar is the superposition of the extensional strain in
eq. (3.4) on p. 45 and the bending strain in eq. (4.15); i.e.,
ε z = w′ + yθ′ x + xθ′ y






extension bending (4.41)

where w ( z ) is the axial displacement of the particle at z on the reference axis, θ x ( z ) is the rotation of the cross
section at station z about the x-axis, θ y ( z ) is the rotation of the cross section at station z about the negative y-axis,
and the prime denotes the ordinary derivative with respect to z. Neglecting the lateral normal stress components,
Hooke’s law is σ z = Eε z . Hence, the axial normal stress is

σ z = E ( w′ + yθ′ x + xθ′ y ) (4.42)

Now substitute eq. (4.42) for the normal stress in the expressions (4.20) for the cross-sectional resultants, and for
a homogeneous cross section we get

N A Q x Q y w′
M x = E Q x I xx I xy θ′ x
My Q y I xy I yy θ′ y

where the geometric properties of the cross-sectional area are

( A, Q x, Q y, I xx, I yy, I xy ) = 2 2
∫ ∫ ( 1, y, x, y , x , xy ) dA
A

The origin of the x-y system is at the centroid of the cross section so that first area moments Q x = Q y = 0 .
Thus, Hooke’s law for the cross section reduces to

N A 0 0 w′
Mx = E 0 I xx I xy θ′ x centroidal coordinates (4.43)

My 0 I xy I yy θ′ y

The inverse of eq. (4.43) is

1⁄A 0 0
w′ I yy – I xy N
1 0 ------------------------
- ------------------------
-
θ′ x = --- 2
I xx I yy – I xy I xx I yy – I xy M x 2 (4.44)
E
θ′ y – I xy I xx My
0 ------------------------
2
- ------------------------
2
-
I xx I yy – I xy I xx I yy – I xy

Now substitute eq. (4.44) for w′ , θ′ x , and θ′ y , in (4.42) to get

N ( I xx M y – I xy M x )x ( I yy M x – I xy M y )y
σ z = ---- + ----------------------------------------
2 )
- + ----------------------------------------
2 )
- (4.45)
A ( I xx I yy – I xy ( I xx I yy – I xy

Thin-Walled Structures 75
Article 4.5

which can be written in the alternate form

N I yy y – I xy x  I xx x – I xy y 
σ z = ---- +  ------------------------
- M +  ------------------------
2  x
- M (4.46)
A  I xx I yy – I xy I I – I2  y
xx yy xy

For a homogenous, Hookean material in the cross section, the axial normal stress due to extension and bending is
given by either formula (4.45) or (4.46).

4.5 Moments of areas


First and second area moments of the cross-sectional area need to be determined before using the flexure 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 y
coincide with the centroid of the cross-sectional area, which y y = yc + y

is labeled C in Fig. 4.15. The second cartesian system x and x = xc + x


y has its origin at an arbitrary point O, the x axis is parallel
C
yc x
to the x axis, and the y axis is parallel to the y axis. The x
and y system should not be confused the rotated system
introduced earlier in the discussion of the plane of bending O
x
(Fig. 4.2). The x and y system in this section is not rotated xc
relative to the x and y system. The location of the centroid in
the x and y system is denoted by coordinate values ( x c, y c ) . Fig. 4.15 Parallel cartesian axes systems.

Usually the x and y system is selected as something conve-


nient to start with, and the first and second area moments with respect to the x and y system are computed or
looked-up in tables. Then the ( x c, y c ) coordinates of the centroid are computed and the parallel axis theorem is
used to find the second area moments in the x and y system.

In the x and y system, the first area moments are defined as

Qx =
A
∫ ∫ y dA Qy =
∫ ∫ x dA
A
(4.47)

in which the area element dA = dxdy . 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

x = xc + x y = yc + y (4.48)

The relationship between the area elements is dA = dA where dA = dxdy , since the ( x c, y c ) values are fixed.

76 Thin-Walled Structures
Moments of areas

If eqs. (4.48) are substituted into eqs. (4.47), we get

Qx = yc A + Qx Qy = xc A + Qy (4.49)

where

A =
∫ ∫ 1 dA
A
Qx =
∫ ∫ y dA
A
Qy =
∫ ∫ x dA
A
(4.50)

Since the origin of the x and y system is at the centroid, the first moments Qx and Qy are zero by definition. Set-
ting Q x = 0 and Q y = 0 in eqs. (4.49), we can solve to find the location of the centroid as

Q Q
x c = ------y y c = ------x (4.51)
A A
These results were also obtained for bars subjected to axial loads only. Refer to eq. (3.12) on p. 51 .

In the x and y coordinate system the second area moments are

2 2
I xx =
∫ ∫ y dA
A
I yy =
∫ ∫ x dA
A
I xy =
∫ ∫ xy dA
A
(4.52)

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. (4.52) 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. (4.48) for the x and y coordinates into eqs. (4.52) to get
2
I xx = y c A + 2y c Q x + I xx
2
I yy = x c A + 2x c Q y + I yy (4.53)

I xy = x c y c A + x c Q x + y c Q y + I xy
where the second area moments in the x and y system are defined as

I xx = 2 2
∫ ∫ y dA
A
I yy =
∫ ∫ x dA
A
I xy =
∫ ∫ xy dA
A
(4.54)

Since the x and y coordinates are centroidal, Qx = Qy = 0, and eqs. (4.53) reduce to
2 2
I xx = y c A + I xx I yy = x c A + I yy I xy = x c y c A + I xy (4.55)

Equations (4.49) and (4.53) are the generalized parallel axis theorem, but in problem solving we use eqs. (4.51)
to find the centroid and then the parallel axis theorem reduces to the use of eqs. (4.55). Note that eqs. (4.54) show
that the Ixx and Iyy are always positive in value with dimensional units of L4. The product area moment Ixy can be
positive, zero, or negative in value. The product area moment Ixy 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 Ixx and Iyy to define radii of gyration
by the definitions

Thin-Walled Structures 77
Article 4.5

rx = I xx ⁄ A ry = I yy ⁄ A (4.56)

The radii of gyration rx and ry have dimensional units of length. However the radii of gyration do not locate a
physically significant point in the cross section. For example, r x ≠ y c + r x , where r x it the radius of gyration with

respect to the x -axis. (Using the parallel axis theorem, the relation between the radius of gyration about the x -
2 2
axis to the x-axis is r x = y c + r x2 .)

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 simplifications with respect to computing
the first and second area moments without loss of significant 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 flanges as is
shown in Fig. 4.16 The gap and overlap at the corners are ignored. Instead of an actual drawing of the zee sec-

t t
0 < --- « 1
h

Fig. 4.16 Modeling a thin-walled zee section with rectangular elements for the web and flanges

tion, or a sketch of the mathematical representation as shown in Fig. 4.16, 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. 4.11. The term “branch” is used to mean either
a flange or web of the zee section. The slope of the contour is continuous within the branch. At the junctions
between branches, like the web-flange 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.

78 Thin-Walled Structures
Moments of areas

For branches that can be represented by a thin-walled


y ds rectangular area, we can obtain simple formulas for the
s second area moments. Consider a thin rectangular area,
where 0 < t << l, represented by a straight line contour
φ inclined at a angle φ as is shown in Fig. 4.17. The con-
x
tour coordinate is denoted by s, and the area element is
dA = tds . The x and y coordinates of the point s on
t the contour are given by x = s cos φ and y = s sin φ .
l/2 l Hence, the second area moments are computed from
l⁄2
l3t
I xx = ( s sin φ ) 2 t ds = ------ sin2 φ
Fig. 4.17 A thin rectangular area inclined with
∫ 12
–l ⁄ 2
respect to the centroidal coordinates. l⁄2
l3t
I yy = ( s cos φ ) 2 t ds = ------ cos2 φ
∫ 12
(4.57)
–l ⁄ 2
l⁄2
l3t
I xy = ( s sin φ ) ( s cos φ )t ds = ------ sin φ cos φ
∫ 12
–l ⁄ 2

Composite body technique The composite body technique for


y
computing the centroid and second area moments is a method 1
applicable to cross-sectional areas that can be subdivided into N
simple geometric shapes whose properties are known. For our
yi 2
example of the thin-walled zee section, we can subdivide it into
yi
three rectangular areas and use the formulas in eqs. (4.57) for xi
the second area moments about the local centroidal axis system Ci
in each rectangle. With the centroid and second area moments i-th sub-area
known for each sub-area, the parallel axis theorem is used to
transfer their properties to a parallel reference system. Then a x
xi
summation of the first 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- Fig. 4.18 Composite body technique
vided into N sub-areas labeled i = 1, 2, …, N (see Fig. 4.18)
are
N N N
A = ∑ Ai Ay c = ∑ yi Ai Ax c = ∑ xi Ai (4.58)
i = 1, 2, … i = 1, 2, … i = 1, 2, …

N
2
I xx = ∑ ( I xx + y A ) i (4.59)
i = 1, 2, …

N
2
I yy = ∑ ( I yy + x A ) i (4.60)
i = 1, 2, …

Thin-Walled Structures 79
Article 4.5

N
I xy = ∑ ( I xy + xyA ) i (4.61)
i = 1, 2, …

where A i, x i, y i, ( I xx ) i, ( I yy ) i, ( I xy ) i 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. (4.58). The second area moments for the reference system deter-
mined from eqs. (4.59) to (4.61) 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 4.3 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 4.1. The section
is sub-divided into three rectangular branches as shown in Fig. 4.19. One branch corresponds to the web and two

y y, y
y3

x3
t = thickness of all branches

a⁄2
a
a
y2
C
x2 x

a y1 yc = a

x1
x x
O a⁄2 O
a

Fig. 4.19 Zee section approximated by three rectangular areas.

branches correspond to the flanges.

Solution First we find the centroid. Equations (4.58) are represented in the table shown below. Summation of the
appropriate columns gives

A = 4at xc A = 0 y c A = 4a 2 t

so that the centroid has coordinates x c = 0 and y c = a .

80 Thin-Walled Structures
Moments of areas

The second area moments are computed for the reference coordinate system ( x, y ) using the second table
shown below. Note that for the local centroidal coordinate systems in each branch we can identify the angle φ in

i Ai xi yi xi Ai yi Ai i
1 at a/2 0 a2t/2 0
2 2at 0 a 0 2a2t
3 at – a/2 2a – a2t/2 2a2t
Sum 4at 0 4a2t
eqs. (4.57) as φ 1 = 0° , φ 2 = 90° , and φ 3 = 0° . These values of the angle φ in each branch are used to com-
pute the local second area moments in each rectangular branch via eqs. (4.57). From the summation of the col-

2 2
i yi Ai ( I xx ) i xi Ai ( I yy ) i xi yi Ai ( I xy ) i
1 0 0 a3t/4 a3t/12 0 0
2 (a2)2at (2a)3t/12 0 0 0 0
3 (2a)2at 0 a3t/4 a3t/12 – a3t 0
Sum 6a3t 2a3t/3 a3t/2 a3t/6 – a3t 0

umns, the second area moments in the ( x, y ) system via eqs. (4.59) to (4.61) are

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
Now we use the parallel axis theorem to transfer these moments to the centroidal system. Equations (4.55) give
2 20 8
I xx = I xx – y c A = ------ a 3 t – a 2 ( 4at ) = --- a 3 t
3 3
2 2 2
I yy = I yy – x c A = --- a 3 t – ( 0 ) ( 4at ) = --- a 3 t
3 3

I xy = I xy – x c y c A = – a 3 t – ( 0 ) ( a ) ( 4at ) = – a 3 t

These are the second area moments given the problem statement of Example 4.1.

Thin-Walled Structures 81
Article 4.5

EXAMPLE 4.4 Circular cross section of variable thickness

The thin-walled cross-section shown has a circular contour with radius r. The wall thickness is t everywhere
except the top-right quarter where it is 2t.
a) Calculate the centroid of the cross-section.
b) Calculate I xy .

2t

r
t

Solution (a) Let the temporary reference be the center of the circle. The first moment of inertia about the center
of the circle can be used to calculate the position of the centroid as:
π⁄2 2π π⁄2 2π

Qx =
∫ y dA = ∫ r sin θ ⋅ 2tr dθ + ∫ r sin θ ⋅ tr dθ = ∫ r sin θ ⋅ tr dθ + ∫ r sin θ ⋅ tr dθ
0 π⁄2 0 0
π⁄2

= tr 2 2
∫ sin θ dθ + 0 = tr
0

π⁄2 2π
 π 3π 5π
A =
∫ 1 ⋅ dA = ∫ 2tr dθ + ∫ r sin θ dθ = tr  2 ⋅ --2- + -----2- = ------tr
2
0 0

Q 2r
y c = ------x = ------
A 5π

2r
Similarly x c = ------ .

(b) The product moment of inertia is given by


π⁄2 2π

I xy =
∫ xy dA = ∫ r cos θ ⋅ r sin θ ⋅ 2tr dθ + ∫ r cos θ ⋅ r sin θ ⋅ tr dθ
0 π⁄2
π⁄2 2π

=
∫ r cos θ ⋅ r sin θ ⋅ tr dθ + ∫ r cos θ ⋅ r sin θ ⋅ tr dθ
0 0

82 Thin-Walled Structures
Moments of areas

π⁄2 π⁄2
π⁄2
sin 2θ tr 3 cos 2θ tr 3
I xy = tr 3 cos θ sin θ dθ + 0 = tr 3 -------------- dθ = ------ ⋅ – --------------
∫ ∫ 2 2 2
-
0
= ------
2
0 0

tr 3 5π 2r 2r 1 2
I xy = I xy – Ax c y c = ------ –  ------tr ⋅ ------ ⋅ ------ = tr 3  --- – ------
2 2 5π 5π  2 5π

EXAMPLE 4.5 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 A f = ( πat ) ⁄ 2 . The section and the reference coordinate system are shown Fig. 4.20. Determine

y y y

Af Af
ds
s
θ O O
x x, x
C
a

t Af Af
Fig. 4.20 Stringer-stiffened, semicircular section.

the location of the centroid and the second area moments.

Solution Since this cross section is symmetric with respect to the x axis, the centroid is located on this axis, and
the product area moment I xy is zero. In thin-walled construction the stringer’s cross-sectional dimensions are
small with respect to the largest dimension of the cross section. Hence, the stringer is modeled by its area As con-
centrated at the stringer’s 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 θ in the moment compu-
tations for the web, where – π ⁄ 2 ≤ θ ≤ π ⁄ 2 , as is shown in the figure. The differential area of the web is
dA = tds = tadθ , and the coordinates to this differential area are x = – a ⋅ cos θ and y = a ⋅ sin θ . The cross-
sectional area is
π
---
2
π
A = = πat + 2  --- at = 2πat
∫ ta dθ + 2A f 2 
π
– ---
2

With reference to the first of eqs. (4.51) and the second of (4.47), the first area moment about the y axis is

Thin-Walled Structures 83
Article 4.6

π⁄2

xc A = ( – acos θ )ta dθ + 2 ( 0 )A s = – 2a 2 t

–π ⁄ 2

Thus, the centroid is located at


– 2a 2 t a
x c = -------------- = – --- yc = 0
2πat π

The second area moment about the x axis is (see the first of eqs. (4.52))
π⁄2 π⁄2
θ 1 3
I xx = ( asin θ ) 2 ta dθ + a 2 A s + ( – a ) 2 A s = a 3 t  --- – --- sin 2θ + πa 3 t = --- πa 3 t
∫ 2 4 
–π ⁄ 2
2
–π ⁄ 2

and the second area moment about the y axis is


π⁄2 π⁄2
θ 1 π
I yy = ( – acos θ ) 2 ta dθ + 2 ( 0 ) 2 A s = a 3 t  --- + --- sin 2θ = --- a 3 t
∫ 2 4 
–π ⁄ 2
2
–π ⁄ 2

Now we use the parallel axis theorem, eqs. (4.55), to get the second area moments about the centroidal system.
2 3 3
I xx = I xx – y c A = --- πa 3 t – ( 0 ) 2 A = --- πa 3 t
2 2

2 π a 2 π 2
I yy = I yy – x c A = --- a 3 t –  – --- 2πat = a 3 t  --- – ---
2  π  2 π

To summarize, the second area moments about the centroidal system are
I xx = 4.712a 3 t I yy = 0.934a 3 t I xy = 0

4.6 Multi-material bars


Strain-displacement Extensional deformation of an element of the beam is shown in Fig. 4.21. The axial nor-

dz dz + dw
z
z w + dw
w(z)

Fig. 4.21 Extensional deformation of a element of the beam

mal strain due to extension is given by

84 Thin-Walled Structures
Multi-material bars

( dz + dw ) – dz
ε z = ------------------------------------ which in the limit as dz → 0 gives ε z = w′
dz
For combined bending and extension, we superpose the extensional strain and the bending strain, eq. (4.15), to
obtained the total strain. (See also Section 4.4.) Hence,

ε z = w′ + xθ′ y + yθ′ x






extension bending (4.62)

Material law Consider Hooke’s 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,
εz = σz ⁄ E + α ( T – T0 )







mechanical thermal (4.63)

in which T denotes the temperature, T 0 the stress-free temperature, α the coefficient of linear thermal expan-
sion (CTE) of the material, and E denotes the modulus of elasticity of the material. If the temperature is given in
degrees fahrenheit ( °F ), then the CTE has dimensional units 1 ⁄ °F . In the material law for the beam given by
eq. (4.63), we have assumed that the lateral stresses σ x and σ y are negligible with respect to the axial normal
stress σ z and that the thermal strains in the x- and y-directions can be neglected. Solving eq. (4.63) for the axial
normal stress we get
σ z = Eε z – Eα∆T (4.64)

where
∆T = T – T 0 (4.65)

In general the material properties E and α can be a function of the temperature as well. However, we assume
that over moderate temperature changes these material properties do not vary significantly from an average value.

In addition to the thermal effects, we assume that the beam cross section is hetrogeneous; i.e., the material
properties E and α are functions of coordinates x and y. In the case of heterogeneity, we define a reference mod-
ulus of elasticity E 0 , 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. (4.64) as

E
σ z = E 0  ------ ε z – Eα∆T (4.66)
 E 0

Resultants We substitute the strain-displacement relation,eq. (4.62), into eq. (4.66) for the axial normal strain to
get

E
σ z = E 0  ------ ( w′ + xθ′ y + yθ′ x ) – Eα∆T (4.67)
 E 0

Recall that the beam resultants, first given by eqs. (4.20), are

Thin-Walled Structures 85
Article 4.6

N =
∫ ∫ σ dA
A
z Mx =
∫ ∫ yσ dA
A
z My =
∫ ∫ xσ dA
A
z

Now substitute eq. (4.67) for the axial normal stress in these beam resultants to get the material law for the beam.
We write the final expression of the material law for the resultants in the matrix form

N A * Q x* Q y* w′ Nt
M x = E 0 Q x* I xx
* I*
xy
θ′ x – M xt (4.68)

My Q y* I xy
* I*
yy
θ′ y M yt

where the modulus weighted area is


E
A* = - dA
∫ ∫ -----
A
E 0
(4.69)

the modulus weighted first area moments are

E E
Q x* = - y dA Q y* = - x dA
∫ ∫ -----
A
E 0
∫ ∫ -----
A
E 0
(4.70)

the modulus weighted second area moments are

* = E 2 * = E 2
I xx - y dA I yy - x dA
∫ ∫ -----
A
E 0
∫ ∫ -----
A
E 0
(4.71)

the modulus weighted product area moment is

* = E
I xy - xy dA
∫ ∫ -----
A
E 0
(4.72)

the thermal axial force is

Nt =
∫ ∫ Eα∆T dA
A
(4.73)

and the thermal bending moments are

M xt = M yt =
∫ ∫ y ( Eα∆T ) dA
A
∫ ∫ x ( Eα∆T ) dA
A
(4.74)

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 first area moments vanish; i.e.,

Q x* = Q y* = 0 (4.75)

We will show how to use this property to find 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. (4.68),
reduces to

86 Thin-Walled Structures
Multi-material bars

N A * 0 0 w′ Nt
M x = E 0 0 I xx
* I*
xy θ′ x
– M xt (4.76)

My * I*
0 I xy yy
θ′ y M yt

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 σ z given by eq. (4.67) in terms of the resultants. Invert eq. (4.76)) and
solve for the extensional strain w′ , and curvatures θ′ x and θ′ y . to get

1
w′ = -----------*- ( N + N t ) (4.77)
E0 A

* –I *
I yy t
θ′ x 1
= -------------------------------------- xy M x + M x (4.78)
E 0 ( I xx * I* – I* 2) * I*
θ′ y yy xy – I xy xx M y + M yt

Now substitute eq. (4.77) for w′ and eqs. (4.78) for θ′ x and θ′ y into eq. (4.67) to eliminate these terms in the
equation for σ z . The equation for the normal stress in the form

σ z = σ zm + σ zt (4.79)

where σ zm is the mechanical portion of normal stress and σ zt is the thermal portion of the normal stress. The for-
mulas for these components are
* M – I * M )x * M – I * M )y
E N ( I xx y xy x ( I yy x xy y
σ zm = ------ ------- + ----------------------------------------- + ----------------------------------------- (4.80)
E 0 A* * *
( I xx I yy – I xy ) *2 * I * – I *2 )
( I xx yy xy

* M t – I * M t )x * M t – I * M t )y
E N t ( I xx y xy x ( I yy x xy y
σ zt = ------ ------- + ----------------------------------------- - – Eα∆T
- + ----------------------------------------- (4.81)
E 0 A* * *
( I xx I yy – I xy ) *2 * I * – I *2 )
( I xx yy xy

In eqs. (4.80) and (4.81) the modulus E and coefficient of thermal expansion α are functions of x and y in gen-
eral.

If the material is homogeneous over the cross-sectional area, then taking E = E 0 the axial normal stresses
due to mechanical and thermal loading are given by

N ( I xx M y – I xy M x )x ( I yy M x – I xy M y )y
σ zm = ---- + ----------------------------------------
2 )
- + ----------------------------------------
2 )
- (4.82)
A ( I xx I yy – I xy ( I xx I yy – I xy

N t ( I xx M yt – I xy M xt )x ( I yy M xt – I xy M yt )y
σ zt = ----- + -----------------------------------------
2 ) 2 )
- – Eα∆T
- + ----------------------------------------- (4.83)
A ( I xx I yy – I xy ( I xx I yy – I xy
The mechanical part of the axial normal stress in eq. (4.82) is the same as the expression given by eq. (4.45).

Thin-Walled Structures 87
Article 4.6

EXAMPLE 4.6 A multi-material beam with a symmetric cross section

A beam of three materials is subjected to a positive bending moment M x only. The maximum tensile normal
stress is 35 ksi. Find the maximum compressive normal stress.
y

2in 1
Mx E 1 = 35msi
x
E 2 = 20msi
2in 2
E 3 = 5msi
*
yc
y 1msi = 10 6 psi
2in 3
x
1in

Fig. 4.22 Bending of a multi-material beam with a symmtric cross section

Solution In this example we take the reference modulus E 0 = 10Msi , so that E 1 ⁄ E 0 = 3.5 , E 2 ⁄ E 0 = 2 , and
E 3 ⁄ E 0 = 0.5 . The modulus weighted area from eq. (4.69) is

3
E Ei
A* =
∫∫ ------ dA =
E0 ∑ -----
E
-A
0
i = 3.5 × 2in 2 + 2 × 2in 2 + 0.5 × 2in 2 = 12in 2
A i=1

Note that the modulus weighted area is twice the geometric area. We choose the x - y system centered at the bot-
tom of the beam as shown in Fig. 4.22. Since the y -axis is an axis of symmetry, both in geometry and in material
properties, the modulus weighted centroid lies on this axis. We need only to find the location on the y -axis of the
*
modulus weighted centroid; i.e., y c . Begin by calculating the modulus weighted first area moment about the x -
axis. See eqs. (4.70). That is,
3
* E Ei
Qx =
∫∫ ------ y dA =
E0 ∑ -----
E
-y A
0
i i = 3.5 × 5in × 2in 2 + 2 × 3in × 2in 2 + 0.5 × 1in × 2in 2
A i=1

so
*
Q x = 48in 3
We now use the parallel axis theorem for the first area moment as given by eqs. (4.49), but generalized to the
multi-material case to write
* *
Q x* = Q x – y c A * = 0 for the modulus weighted centroid

Hence

88 Thin-Walled Structures
Multi-material bars

*
* Q 48in 3
y c = ------*x = ------------2- = 4in
A 12in
Note that the modulus weighted centroid is located at the intersection of materials 1 and 2, which does not coin-
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
* = 0 . Since
that the modulus weighted product area moment about the x-y axes vanishes; i.e., I xy
* = 0 , the formula for the axial normal stress, eqs. (4.79) and (4.80), reduces to
N = M y = I xy

E Mx
-y
σ z = ------ ------
*
2in ≤ y ≤ – 4in
E 0 I xx

* . This computation proceeds as follows


Thus, we only need to compute I xx
3 3 3
E E E Ei
*
I xx = ------ y 2 dA = ∑ ∫∫ ------i y 2 dA = ∑ ∫∫ ------i y 2 dA ∑ -----
-(I + y 2 A )i
∫∫
A
E0
i = 1 Ai
Eo
i=1
Eo
Ai
=
E
i=1
o
xx

In achieving this result we used the fact that the (geometric) second area moment of the i-th area element about
2 + y i2 A i , which is based on the composite body technique and parallel axis theorem.
the x-axis is
∫ ∫ y dA = I
Ai
xxi

See eq. (4.59). Performing the computations we have

1
* = 3.5 ×  ----- 1
I xx - × 1 × 2 3 + ( 1 ) 2 × 2 + 2 ×  ------ × 1 × 2 3 + ( – 1 ) 2 × 2 +
 12   12 
1
0.5 ×  ------ × 1 × 2 3 + ( – 3 ) 2 × 2 = 24in 4
 12 

Only material 1 is in tension for M x > 0 . So the axial normal stress is a maximum in material 1 at
y = 2in ;i.e., σ z = 35ksi Thus,

Mx
35ksi = 3.5 × ------------
- × 2in
24in 4
3
Hence M x = 120 ×10 lb-in . The maximum compressive normal stress in material 2 is

E Mx
σ z = -----2- ------
*
- ( – 2in ) = 2 × 5000lb/in 3 × ( – 2in ) = – 20ksi
E 0 I xx

The maximum compressive normal stress in material 3 is


E Mx
σ z = -----3- ------
*
- ( – 4in ) = 0.5 × 5000 × ( – 4 ) = – 10ksi
E 0 I xx

Therefore, the magnitude of the maximum compressive normal stress is 20ksi.

Thin-Walled Structures 89
Article 4.6

As a function of y, the bending normal stress is given by


3
 17, 500 lb/in × y 0 < y ≤ 2in

σ z = 10, 000 lb/in 3 × y – 2in < y < 0

 2, 500 lb/in 3 × y – 4in ≤ y < – 2in
and this is plotted in Fig. 4.23. 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 w′ , θ x ′ , and θ y ′ is given by eq. (4.62). For this problem N = N t = 0 , so the w′ = 0 from eq.
(4.77); i.e., there is no axial extensional component to the strain. Also, M y = M yt = M xt = I xy
* = 0 , which

* ) and θ ′ = 0 . From eq. (4.62), the axial strain distribution in y -


gives from eq. (4.78) that θ x ′ = M x ⁄ ( E 0 I xx y
coordinate is
Mx 5000 lb/in 3
ε z = yθ x ′ = y ------------
*
= y - = ( 500 × 10 –6 in –1 )y
------------------------------
6
– 4in ≤ y ≤ 2in
E 0 I xx 10 ×10 lb/in 2

The maximum tensile strain is 1000 × 10 –6 in/in = 1000µε at y = 2 in, and the maximum compressive strain is
– 2000 × 10 –6 in/in = – 2000µε at y = -4 in. (The quantity µε is read as microstrain.) The strain is continuous in
the coordinate y as is shown in Fig. 4.23.

y y, in
y, in

2 2
2in 1 Mx > 0
x σ z, ksi ε z, µε
– 20 – 10 – 5 – 2000
35 0 1000
2in 2
–2 –2
2in 3
–4 –4
1in

Fig. 4.23 Distribution of normal stress and strain through the thickness for the multi-material beam

90 Thin-Walled Structures
Problems

4.7 Problems
1. For the thin-walled Y-section shown at right, the y, y
branch thickness t is much smaller that the branch
length l. 45° 45°

a) Determine the location y c of the centroid.

b) Determine the second area moment I xx of C x


the cross section about the x axis.
yc
c) If the bending moment components Mx > 0 O
x
and My = 0, determine the magnitudes of l l
the largest tensile and compressive normal
stress due to bending. State the coordinates Each of the five branches: length l and thickness t
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 Mx = 3500 y

Nmm, My = 0. Determine the angle β of the neutral


axis in the cross section, and the magnitude of the t
a
maximum bending normal stress.

3. Half of the cross section of a ship is shown on 4a C


x
the next page. Only the material that is effective in
β
the longitudinal bending is illustrated in the figure.
Determine the area A, location of the centroid y c , a neutral axis in the cross
t section
and the second area moment about the x-axis (
I xx )for the full section Use the tabular format for the
computations as given in Example 4.3. All plating
has a thickness t = 14 mm unless other wise noted. The descriptions of the numbered structural elements shown
in the figure 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 L500 × 400 × 25
10 Strength deck plating

Thin-Walled Structures 91
Article 4.7

.
4m

10
11

7
9m

8
9

x
C
5.5 m

yc
2
5 4 3
6
t/2 x
R=1m 1 3.5 m

6.5 m

L500 × 400 × 25

π
R A = --- Rt 500mm
2R ⁄ π 2
C π 1 4 C
t x I xx = --- R 3 t  --- – ----- x
2  2 π 2 139mm

400mm
A = 0.0225m 2

4. The cross-section shown in the adjacent sketch is circular with radius r with a ver-
tical member of length r. The thickness of the cross-section is t everywhere.
a)Calculate the centroid of the cross-section.
r
b)Calculate I xx
r
t

92 Thin-Walled Structures
Problems

5. Calculate the second area moments,or moments of inertia, I xx , I yy , and I xy , for the cross-section shown
below. The position of the centroid is marked by the dash-dot line.

1⁄8
2 1⁄8

All dimensions in inches


2 1⁄4

1 1

6. For the wing described in Problem 2 on page 40, calculate the stresses at the root if the airfoil section is a
NACA 0012 with chord length of 3 ft. The wing is made up of composite skin over foam core. The composite
skin is 0.25 in thick symmetric, quasi-isotropic layup. (Assume that the foam is very flexible relative to the com-
posite and thus does not take any load)

Thin-Walled Structures 93
Article 4.7

94 Thin-Walled Structures
CHAPTER 5 Bending of beams under
transverse loads

5.1 Approximations for slender beams


In the last chapter the special case of pure bending of a beam caused equal and opposite couples applied to the
ends of the beam was discussed. In pure bending, the bending moment interior to the beam is equal to the
moment of the applied end couples and the transverse 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 p y ( z ) , then the shear
force Vy(z) is non-zero by equilibrium eq. (2.1) on p. 21 . Thus, moment equilibrium, eq. (2.3) on p. 22 , requires
the bending moment to change with z along the beam. The presence of the transverse shear force causes plane
cross sections 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 4.2. 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 analy-
sis is too complex for routine calculations, and in structural mechanics a simpler theory is used. The simpler the-
ory to compute stresses in beams under transverse loading is called engineering theory, or technical theory, or
Bernoulli-Euler theory.

Assumption in the engineering theory of beams


In the engineering theory of beams it is assumed that the distribution of normal stress σz
given by the flexure formula, eq. (4.27) on p. 69 , or equivalently eq. (4.28), 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.

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.

Thin-Walled Structures 95
Article 5.2

In the following Section 5.2 we discuss the governing boundary value problem for the lateral displacements
of a beam subject to transverse loads. Statically indeterminate beams in transverse bending are considered in
Example 5.1 and Example 5.3.

5.2 Beam displacements


Neglecting the displacements due to transverse shear, the formulation of the pure bending problem discussed in
Section 4.1 and Section 4.2 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 fibers orig-
inally parallel to the z-axis in the undeformed beam.

Equilibrium Let p x ( z ) and p y ( z ) , denote the lateral load intensities (F/L) in the x-direction and y-direction,
respectively, acting on the beam. These external load intensities are defined positive if they act in their respective
positive coordinate directions as shown in Fig. 5.1. Also shown in the figure, are the positive shear forces, V x
py y, v
Vy

px x, u
Vx
z
Mx

My

Fig. 5.1 External distributed loads and internal actions

and V y , and the positive bending moments, M x and M y , 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 dz → 0 ,
the sum of forces leads to

dV y
--------- + p y = 0 (5.1)
dz
and the sum of the moments leads to

dM
----------x – V y = 0 (5.2)
dz

96 Thin-Walled Structures
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 dz → 0 , the sum of forces in the x-direction leads to

dV x
--------- + p x = 0 (5.3)
dz
and the sum of moments about the negative y-direction leads to

dM
----------y – V x = 0 (5.4)
dz
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
2
d Mx
= –py ( z ) (5.5)
dz2
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
2
d My
= –px ( z ) (5.6)
dz2

Strain-displacement The normal strain of line elements parallel to the z-axis in the undeformed beam is deter-
mined in Section 4.2, and is given by eq. (4.15) on p. 66 . Repeating this result we have

dθ dθ
ε z = x  --------y + y  --------x (5.7)
 dz   dz 

in which the rotation of the projection of the z-curve into the y-z plane (Fig. 3.9) is
dv
θ x = – ------ (5.8)
dz
and the rotation of the projection of the z-curve into the x-z plane is
du
θ y = – ------ (5.9)
dz
These rotations are depicted in Fig. 4.9 on page 66. 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

2 2
 d u  d v
ε z = x  – 2 + y  – 2 (5.10)
 dz   dz 

Material law For a linear elastic, isotropic material the normal stress-strain relation is σ z = Eε z , 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

Thin-Walled Structures 97
Article 5.2

homogenous material, it was shown in eq. (4.25) on p. 69 that statical equivalence of the normal stress distribu-
tion to the bending moments leads to

Mx I xx I xy θ′ x
= E (5.11)
My I xy I yy θ′ y

where Ixx, Iyy, and Ixy 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

θ′ x 1 I – I xy M x
2 )
- yy
= --------------------------------- (5.12)
θ′ y E ( I xx I yy – I xy – I xy I xx M y

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

2 2 2 2
d  d v d  d u
2
 EI xx 2
 + 2
 EI xy 2  = p y ( z ) 0<z<L (5.13)
dz  dz  dz  dz 

2 2 2 2
d  d v d  d u
2
 EI xy 2
 + 2
 EI yy 2  = p x ( z ) 0<z<L (5.14)
dz  dz  dz  dz 
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 0 < z < L , 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 Ixy 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 EI xx, EI yy , and EI xy are independent of coordinate z, then the governing equations
reduce to
4 4
dv du
EI xx + EI xy 4 = p y 0<z<L (5.15)
dz4 dz
4 4
dv du
EI xy + EI yy 4 = p x 0<z<L (5.16)
dz4 dz

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 v or shear force V y , but not both

b) either rotation θ x or bending moment M x , 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

98 Thin-Walled Structures
Beam displacements

y, v, V y

θ x, M x 2 2
 d v  d u
z v = 0 M x = EI xx  –  + EI xy  –  = 0
2
 dz   d z 2
(1) simple support

z v = 0 θ x = – v′ = 0
(2) clamped

2 2
d  d v  d u 
V y = -----  EI xx  –  + EI xy  –   = 0
dz  2
 dz   d z 2 
(3) free z 2 2
 d v  d u
M x = EI xx  –  + EI xy  –  = 0
 d z 2  d z 2

2 2
z d  d v  d u 
(4) free in y V y = -----  EI xx  –  + EI xy  –   = 0 θx = 0
dz  2
 dz   d z 2 
clamped about x

Fig. 5.2 Standard boundary conditions for bending in the y-z plane.

a) either displacement u or shear force V x , but not both

b) either rotation θ y or bending moment M y , but not both

The so-called standard boundary conditions for bending in the x-z plane are shown in Fig. 5.3.

Thin-Walled Structures 99
Article 5.2

x, u, V x

θ y, M y 2 2
 d v  d u
z u = 0 M y = EI xy  –  + EI yy  –  = 0
 d z 2  d z 2
(1) simple support

z u = 0 θ y = – u′ = 0
(2) clamped

2 2
x d  d v  d u 
V x = -----  EI xy  –  + EI yy  –   = 0
dz  2
 dz   d z 2 
(3) free z
2 2
 d v  d u
M y = EI xy  –  + EI yy  –  = 0
2
 dz   d z 2

2 2
(4) free in x z d  d v  d u 
V x = -----  EI xy  –  + EI yy  –   = 0 θy = 0
clamped about y dz   d z 2  d z 2 

Fig. 5.3 Standard boundary conditions for bending in the x-z plane.

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

1 1
----------- – -----------
– v″ = ER yy ER xy M x
(5.17)
– u″ 1 1 My
– ----------- -----------
ER xy ER xx

where the second area moment ratios are defined by

I xx I yy – I xy 2 I xx I yy – I xy 2 I xx I yy – I xy 2
R xx = ------------------------
- R yy = ------------------------
- R xy = ------------------------
- (5.18)
I xx I yy I xy

100 Thin-Walled Structures


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
4
dv py ( z ) px ( z )
= -----------
- – ------------ 0<z<L (5.19)
dz4 ER yy ER xy
4
du py ( z ) px ( z )
4
= – -----------
- + ------------ 0<z<L (5.20)
dz ER xy ER xx

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 p 0 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-
8 2
ond area moments for the cross section are I xx = --- a 3 t , I yy = --- a 3 t , and I xy = – a 3 t , where dimension a is as
3 3
shown in the figure and t denotes the wall thickness. Determine the displacements u ( z ) and v ( z ) , and the bend-
ing moments Mx(z) and My(z).

y,v x,u y
p0
a
a x
z z
a
L L
a

Fig. 5.4 Statically indeterminate beam with a zee cross section.

Solution For this cross section the second area moment ratios defined by eqs. (5.18) are

7 7 7
R xx = ------ a 3 t R yy = --- a 3 t R xy =  – --- a 3 t
24 6  9

The governing equation for bending in the y-z plane is obtained from eq. (5.19) as
4
dv p0
= ----------- 0<z<L (5.21)
dz4 ER yy
and the boundary conditions are
v(0) = 0 (5.22)

Thin-Walled Structures 101


Article 5.2

θ x ( 0 ) = – v′ ( 0 ) = 0 (5.23)

v(L) = 0 (5.24)

M x ( L ) = EI xx ( – v″ ) + EI xy ( – u″ ) = 0 (5.25)
z=L

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
4
du p0
4
= – ----------- 0<z<L (5.26)
dz ER xy
and the boundary conditions are
u = u′ = 0 at z = 0 and z = L (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
p0 z 4 z3 z2
u = – ----------- ------+ c 1 ---- + c 2 ---- + c 3 z + c 4
ER xy 24 6 2

where c1,..., c4 are arbitrary constants. The boundary conditions, eq. (5.27) at z = 0 require that constants c3 = c4
= 0. The boundary conditions at z = L lead to

L3 L2 p0 L 4
( u ( L ) = 0 ) → -----c 1 + -----c 2 = -----------------
-
6 2 24ER xy
L2 po L 3
( u′ ( L ) = 0 ) → -----c 1 + Lc 2 = --------------
-
2 6ER xy

These simultaneous equations are solved for constants c1 and c2 to get

p0 L p0 L 2
c 1 = --------------
- c 2 = – -----------------
-
2ER xy 12ER xy

Hence, the solution for u(z), after substitution for Rxy, is

3p 0 2
u ( z ) = -----------------z ( L – z )2 (5.28)
56Ea 3 t

The general solution of eq. (5.21) for v(z) is


p0 z 4 z3 z2
v = ----------- ------ + c 5 ---- + c 6 ---- + c 7 z + c 8
ER yy 24 6 2

where c5,..., c8 are arbitrary constants. Satisfaction of boundary conditions (5.22) and (5.23) requires c7 = c8 = 0.
Boundary condition (5.25) can be re-written as
R xx
v″ ( L ) + -------u″ (L) = 0
R xy

Substituting for v(z) and u(z), eq. (5.28), this boundary condition leads to

102 Thin-Walled Structures


Beam displacements

p 0 L 2 R xx p 0 L 2
Lc 5 + c 6 = ------------------
2
- – ---------------
12ER xy 2ER yy

Boundary condition (5.24) leads to

L3 L2 p0 L 4
-----c 5 + -----c 6 = – -----------------
-
6 2 24ER yy

Solving these last two equations for c5 and c6, we get

L ( – 5R xy 2 + R R )p L 2 ( 3R xy 2 – R R )p
xx yy 0 xx yy 0
c 5 = ----------------------------------------------------
2 R
- c 6 = -------------------------------------------------
2 R
-
8ER xy yy 24ER xy yy

Substituting for the second area moment ratios gives


p0
v ( z ) = -------------------- z 2 ( 39L 2 – 71Lz + 32z 2 ) (5.29)
896Ea 3 t
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 v ( z ) . The out-of-plane lateral displace-
ment u ( z ) is non zero in this problem since the product area moment I xy ≠ 0 . If I xy = 0 , then u ( z ) = 0
∀z ∈ ( 0, L ) . See eqs. (5.18) and (5.26). That is, the beam would not bend out of the plane of loading if
I xy = 0 . However, as can be seen in Fig. 5.5, the displacement in the plane of loading v ( z ) is much larger than
the out of plane displacement u ( z ) 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
p
M x = ----0- ( – L 2 + 5Lz – 4z 2 )
8
p0 L
M y = -------- - ( L – 3z )
64
The bending moment diagrams for Mx and My are shown in Fig. 5.6. Note that the moment for bending in the y-
z plane, or Mx, is larger than the moment for bending in the x-z plane, or My. Moment My would vanish if
I xy = 0 .

Thin-Walled Structures 103


Article 5.2

33 44
Ea
Ea ttvvڐpp0 L
0L

2.5
2

1.5
1
0.5

zzê
⁄ LL
0.2 0.4 0.6 0.8 1

Ea3 3 tu ⁄ p L 4 4
Ea t uêp0 0 L

0.003
0.0025
0.002
0.0015
0.001
0.0005
zz ê⁄ L
0.2 0.4 0.6 0.8 1

Fig. 5.5 Lateral displacements of the unsymmetrical section beam

104 Thin-Walled Structures


Beam displacements

MMxxê⁄pp00LL22

0.05

zz ê⁄ L
L
0.2 0.4 0.6 0.8 1

-0.05

-0.1

2
Myyê⁄ pp00 L2
M

0.01

z ê⁄ LL
z
0.2 0.4 0.6 0.8 1
-0.01

-0.02

-0.03
Fig. 5.6 Bending moment diagrams for example 4.1

Fig. 5.6 Bending moment diagrams for the unsymmetrical section beam.

Fi 5 6 B di t di f l 41
EXAMPLE 5.2 Cantilevered beam with an equal leg angle section

The uniform cantilevered beam shown below has a length L and a cross section in the form of an equal leg angle.
The vertical and horizontal legs have the same length a and thickness t. The beam is subjected to a uniformly dis-
tributed load with intensity p 0 in the y-direction, and the modulus of elasticity is E.

5a 3 t 5a 3 t a3t
a) Show that I xx = ---------- , I yy = ---------- , and I xy = ------- for centroidal coordinates in the cross section.
24 24 8
b) Write the differential equations for the beam displacements and the corresponding boundary conditions.
c) Calculate the displacement functions u ( z ) and v ( z ) .

Thin-Walled Structures 105


Article 5.2

p0 a

p0 a⁄4
y
a y
---
4 x t
z C
a
L
t

Solution (a) Denote the vertical branch, or leg, by the number one and the horizontal branch by the number 2.
The table below is used to compute the second area moments relative to the centroidal coordinates in the cross
section.

i y i2 A i ( I xx ) i x i2 A i ( I yy ) i xi yi Ai ( I xy ) i
1 2 a3t 2 0 0
 – a--- at -------  – a--- at  – a---  – a--- at
 4 12  4  4  4

2 2 0 2 a3t 0
 a--- at  a--- at -------  a---  a--- at
 4  4 12  4  4

a3t a3t a3t a3t a3t 0


∑ -------
8
-------
12
-------
8
-------
12
-------
8

Thus,

a3t a3t 5a 3 t a3t a3t 5a 3 t a3t a3t


I xx = ------- + ------- = ---------- I yy = ------- + ------- = ---------- I xy = ------- + 0 = -------
8 12 24 8 12 24 8 8

(b) Since p x ( z ) = 0 and p y ( z ) = p 0 , ∀z ∈ ( 0, L ) , we have the differential equations for the lateral dis-
placements of the beam given by
4 4
dv p0 du –p0
4
= ----------- 4
= -----------
dz ER yy dz ER xy

For this cantilevered beam, the displacements and slopes are zero at the root while the shear forces and bending
moments are zero at the tip. Thus, the boundary conditions are

106 Thin-Walled Structures


Beam displacements

z = 0 z = L
v = 0 V y = EI xx ( – v′′′ ) + EI xy ( – u′′′ ) = 0

θ x = – v′ = 0 M x = EI xx ( – v′′ ) + EI xy ( – u′′ ) = 0

u = 0 V x = EI xy ( – v′′′ ) + EI yy ( – u′′′ ) = 0

θ y = – u′ = 0 M y = EI xy ( – v′′ ) + EI yy ( – u′′ ) = 0

(c) The solution of the differential equations is given by


p0 z3 z2
- z 4 + c 1 ---- + c 2 ---- + c 3 z + c 4
v = -----------------
24ER yy 6 2

–p0 4 z3 z2
u = -----------------
- z + c 5 ---- + c 6 ---- + c 7 z + c 8
24ER xy 6 2

Substituting the above solutions into the boundary conditions we can calculate the constants. Before we proceed,
it should be noted that the boundary conditions at z = L imply that:

M x = EI xx ( – v′′ ) + EI xy ( – u′′ ) = 0 
 ⇔ v′′ = u′′ = 0
M y = EI xy ( – v′′ ) + EI yy ( – u′′ ) = 0 

V y = EI xx ( – v′′′ ) + EI xy ( – u′′′ ) = 0 
 ⇔ v′′′ = u′′′ = 0
V x = EI xy ( – v′′′ ) + EI yy ( – u′′′ ) = 0 

Satisfaction of the boundary conditions at z = 0 requires


c4 = c3 = c8 = c7 = 0

Vanishing of the transverse shear forces at z = L requires


p0 –p0
-----------L + c1 = 0 -----------L + c5 = 0
ER yy ER xy

Hence,
–p0 L p0 L
c 1 = -----------
- c 5 = -----------
ER yy ER xy

Vanishing of the bending moments at z = L requires


p0 – p 0 L –p0 2  p0 L 
- L 2 +  -----------
-------------- - L + c2 = 0 - L + ----------- L + c 6 = 0
--------------
2ER yy  ER yy 2ER xy  ER xy

Hence,
p0 L 2 –p0 L 2
c 2 = --------------
- c 6 = --------------
-
2ER yy 2ER xy

Thin-Walled Structures 107


Article 5.2

Thus, the final solution for the lateral displacements is


p0 p0 L 3 p0 L 2 2 p0 L 4  z  4 z 3 z 2
- z 4 – --------------
v = ----------------- - --- – 4  --- + 6  ---
- z + --------------- z = -----------------
24ER yy 6ER yy 4ER yy 24ER yy L    L   L

–p0 4 po L 3 p0 L 2 2 R yy I xy


u = ----------------- - z – --------------- z = –  -------
- z + -------------- v = –  -----
- v
24ER xy 6ER xy 4ER xy  R xy   I yy

EXAMPLE 5.3 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 EIxx, 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.
y, v

z P

L/2 L/2
P

R
Fig. 5.7 Two cantilever beams in contact with
each other.

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 p y ( z ) 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
4
dv L L
= 0 0 < z < --- --- < z < L
dz4 2 2
Consequently, the displacement function in the top beam is

z3 z2 L
v ( z ) = v 11 ( z ) = c 1 ---- + c 2 ---- + c 3 z + c 4 0 ≤ z ≤ --- (5.30)
6 2 2
z3 z2 L
v ( z ) = v 12 ( z ) = c 5 ---- + c 6 ---- + c 7 z + c 8 --- ≤ z ≤ L (5.31)
6 2 2
The boundary conditions are
v ( 0 ) = v′ ( 0 ) = 0 M x ( L ) = – EIv″ ( L ) = 0 V y ( L ) = – EIv′′′ ( L ) = – P (5.32)

108 Thin-Walled Structures


Beam displacements

and the conditions of continuity of displacement, rotation, and bending moment are
L
v 11 = v 12 v′ 11 = v′ 12 v″ 11 = v″ 12 at z = --- (5.33)
2
There is a jump in the shear force at z = L/2 due to contact with the lower beam given by

V + L L L L
--- – V y-  --- + R = 0 → – EIv′′′ 12  --- + EIv′′′ 11  --- + R = 0 (5.34)
 y  2  2   2  2

For the lower beam, the governing equation is


4
dv L
= 0 0 < z < ---
dz4 2
and the general solution of this equation is

z3 z2 L
v ( z ) = v 2 ( z ) = c 9 ---- + c 10 ---- + c 11 z + c 12 0 ≤ z ≤ --- (5.35)
6 2 2
The boundary conditions for the lower beam are

L L L L
v ( 0 ) = v′ ( 0 ) = 0 M x  --- = – EIv″  --- = 0 V y  --- = – EIv′′′  --- = – R (5.36)
 2  2  2  2

The condition that the beams are in contact at z = L/2 is

L L
v 11  --- = v 2  --- (5.37)
 2  2

The unknown reaction force R can be eliminated between eq. (5.34) and the third of eqs. (5.36) to get
v′′′ 11 – v′′′ 12 + v″′ 2 = 0 (5.38)
L
z = ---
2

where we have divided by the common factor EI.

For displacement v11, eq. (5.30), to satisfy the first two clamped boundary conditions of eqs. (5.32) constants
c4 = c3 = 0. For displacement v12, eq. (5.31), to satisfy the last two boundary conditions of eqs.(5.32) constants
c5 = P/EI and c6 = L P/EI. For the displacement of the lower beam v2, eq. (5.35), to satisfy the first two clamped
end conditions of eqs. (5.36) constants c11 = c12 =0. For v2 to satisfy the zero moment condition, third of eqs.
(5.36), constant c10 = - c9 L/2. Thus, to this point the displacements are

z3 z2 P LP
v 11 = c 1 ---- + c 2 ---- v 12 = --------- z 3 – --------- z 2 + c 7 z + c 8 (5.39)
6 2 6EI 2EI

z 3 Lz 2
v 2 = c 9  ---- – -------- (5.40)
6 4 
Substitute both of eqs. (5.39) into continuity of moments, second of eqs. (5.33), to get
L LP
c 1 --- + c 2 + --------- = 0
2 2EI

Thin-Walled Structures 109


Article 5.2

Substitute the first of eqs. (5.39) and eq. (5.40) into the contact condition, eq. (5.37), to get
L3 L2 L3
------ c 1 + -----c 2 + ------ c 9 = 0
48 8 24
Substitute the displacements eqs. (5.39) and (5.40) into the contact force continuity condition, eq. (5.38), to get
P
c 1 + c 9 – ------ = 0
EI

These last three equations are solved for c1, c2, and c9. We get c 1 = – P ⁄ ( 4EI ) , c 2 = – ( 3LP ) ⁄ ( 8EI ) , and
c 9 = ( 5P ) ⁄ ( 4EI ) .

Only c7 and c8 remain unknown. The conditions to determine these constants are the continuity conditions
of displacement and rotation, the first and second of eqs. eq. (5.33). Using the results for c1, c2, and c9 these con-
tinuity conditions become

L 3P 5L 2 P
--- c 7 + c 8 = 5L
------------- – c 7 + ------------- = 0
2 96EI 32EI

Thus, we find that c 7 = ( 5L 2 P ) ⁄ ( 3EI ) and c 8 = – ( 5L 3 P ) ⁄ ( 192EI ) .

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

 3LP P L
– ------------ z 2 – ------------ z 3 0 ≤ z ≤ ---
 16EI 24EI 2
v1 ( z ) =  (5.41)
3
5L P 5L P- --------
 – -------------- 2 LP P L
 192EI- + ------------ z – - z 2 + --------- z 3 --- ≤ z ≤ L
32EI 2EI 6EI 2
and the displacement of the lower beam is
5LP 5P
v 2 ( z ) = – ------------ z 2 + ------------ z 3 (5.42)
16EI 24EI

The assumption that the lower beam contacts the upper beam only at its tip needs to be verified. The condi-
tion of no penetration of the contact between the upper and lower beam is
L
v2 ≤ v1 0 ≤ z ≤ ---
2
Substitute eqs. (4.41) and (4.42) into this inequality to get
5LP 5P 3LP P
– ------------ z 2 + ------------ z 3 ≤ – ------------ z 2 – ------------ z 3
16EI 24EI 16EI 24EI
Multiply this equation by the positive factor EI/P, and after rearrangement we find

z2  L 
---- – --- + z ≤ 0 which is valid for 0 ≤ z ≤ L ⁄ 2
4 2 

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

110 Thin-Walled Structures


Beam displacements

M x ( z ) = – EIv″ ( z ) . Using eqs. (5.41), we find for the upper beam

P--- ( 3L + 2z )
L
0 ≤ z ≤ ---
8 2
M x1 = 
 P(L – z) L
--- ≤ z ≤ L
 2
and using eq. (5.42) we find for the lower beam
5P
M x2 = ------- ( L – 2z ) 0≤z≤L⁄2
8
The bending moment distribution in each beam is shown in Fig. 5.9.

v EI/(P L^3)

z/L
0.2 0.4 0.6 0.8 1

-0.05

-0.1

-0.15

-0.2
Fig. 5.8 The displacement functions of the upper beam (solid line) and the lower beam (dashed line)
for the two contacting cantilever beams.

Thin-Walled Structures 111


Article 5.3

M/(P L)

0.6

0.5

0.4

0.3

0.2

0.1

z/L
0.2 0.4 0.6 0.8 1
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.

5.3 Problems
1. Calculate the vertical displacement for the cantilevered beam with a constant distributed load as shown
below. Write the vertical displacement v as a function of the axial coordinate z. (Given: E = 1 GPa ,
–5
I xx = 1 ×10 m 4 , , and I xy = 0.0 m 4 .

y, v
1000 N/m

z
1m

2. For the wing described in Problem 2 on page 40, and in Problem 6 on page 93, calculate the deflection of the
6
wing as a function of the spanwise coordinate. Determine the wing tip displacement. ( E L = 15 ×10 psi )

112 Thin-Walled Structures


Problems

3. The cross section of the uniform beam shown below is the same as in Example 5.2. It is clamped at the root,
or z = 0 . At the tip, or z = L , it is free in the y-z plane and simply supported in the x-z plane. Take
L = 1000 mm , a = 50 mm , t = 3 mm , E = 69000 N/mm 2 , and p 0 = 1 N/mm . Determine the magnitudes
of the largest tensile and compressive axial normal stresses at the root. Show in a sketch of the cross section at the
root the location where these largest tensile and compresive normal stresses occur.

p0
y, v x, u

z z

L L

4. 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. The cross section is symmetric, I xy = 0 , and it bends in the y-z
plane. Neglect the weight of the beam, and determine
v ( z )EI xx
a) The displacement function v ( z ) . Plot the normalized displacement v ( z ) = --------------------
4 )
- versus dimen-
( p 0 z max
sionless coordinate z = z ⁄ z max for 0 ≤ z ≤ 1 .

Mx
b) Plot the normalized bending moment M x ( z ) = ---------------
2
- for 0 ≤ z ≤ 1 .
p 0 z max

Vy
c) Plot the normalized shear force V y ( z ) = --------------- for 0 ≤ z ≤ 1 .
p 0 z max

y, v
p y ( z ) = p 0 = constant

64EI xx
z K = ---------------
3
-
z max
z max ⁄ 2
z max

Thin-Walled Structures 113


Article 5.3

5. 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 EI xx = EI , the distance between the knife edge and the beam’s con-
nection to the disk is L and the radius of the disk is R. This disk rotates about a fixed smooth pin through its cen-
ter under the action of applied moment Ma as shown. Determine the relation between the applied moment Ma and
rotation angle θ of the disk under the assumption that the angle of rotation θ is small.
Ma, θ
y, v R
θx
EI xx = EI
z

114 Thin-Walled Structures


CHAPTER 6 Shear flows and stresses
due to transverse shear
forces

6.1 Rectangular cross section


Consider bending under transverse loads of a rectangular section beam in the y-z plane as shown in Fig. 6.1. For

y
Vy
τ yz
h/2 Vy z x y
τ zy
Mx σz Mx
Mx z
h/2

Vy
t
Fig. 6.1 Shear and normal stresses in a rectangular section beam under general bending.

a slender beam, it is assumed that the normal stress σ z is determined from the flexure formula, eq. (4.27) on p. 69
, with sufficient accuracy when the beam is subjected to shear. The shear stress component τ zy exists if the shear
force Vy is present. Our purpose is to determine the shear stress in terms of the shear force.

The first subscript on the symbol τ zy 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 τ zy is defined as acting in the positive y-direction on a
positive z-face. By the action/reaction law, a positive value for shear stress τ zy 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 infinitesimal stress ele-
ment about the x-direction requires the shear stress component acting on the y-face in the z-direction, or τ yz , be

Thin-Walled Structures 115


Article 6.1

non-zero and equal to τ zy ; i.e., τ yz = τ zy . Shear stress component τ yz is sometimes called conjugate of τ zy . The
sign convention for shear stress τ yz is defined in the same manner as for shear stress τ zy . Shear stress compo-
nents τ zy and τ yz 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 τ yz 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
h h
at a generic value of y, where – --- ≤ y ≤ --- . The free body diagram of the beam below the y = constant plane is
2 2
shown in Fig. 6.2. The bending normal stress σz acts over a portion of the cross sectional area at z and z + ∆z.

∆F yz
σz ( z ) σ z ( z + ∆z )
y
z F + ∆F
F

h/2

z ∆z

Fig. 6.2 Free body diagram for a selected portion of the beam.

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 σz acting over a portion of the cross-sectional area is

y t⁄2 y t⁄2 Mx M y
F = σ z dx dy = ------- y dx dy = -------x yt dy
∫ ∫
–h ⁄ 2 –t ⁄ 2
∫ ∫
–h ⁄ 2 –t ⁄ 2 I xx I xx ∫ –h ⁄ 2
(6.1)

where we substituted the flexure formula for the bending normal stress σz. The integral in the last of these equa-
tions for F is the first area moment of the lower portion of the cross-sectional area; i.e.,
y
t h 2
Qx ( y ) = yt dy = --- y 2 –  ---
∫ –h ⁄ 2 2  2
(6.2)

Thus, eq. (6.1) becomes


M
F = -------x Q x ( y ) (6.3)
I xx

The action of the top portion of the beam on the bottom portion is represented by the force ∆F yz . Notice that the
surface at y = – h ⁄ 2 is stress free, so that no force or shear stress acts on this surface.; i.e., τ yz = 0 at
y = – h ⁄ 2 . As ∆z → 0 , ∆F yz → 0 , since internal forces are only transmitted over finite areas. This force is
given by

116 Thin-Walled Structures


Rectangular cross section

z + ∆z t⁄2
∆F yz = τ yz dx dz
∫ z
∫ –t ⁄ 2

The integral of the shear stress across the width is defined as the shear flow and is denoted by q; i.e.,
t⁄2
q = τ yz dx
∫ –t ⁄ 2
(6.4)

Thus, the shear force ∆Fyz is


z + ∆z
∆F yz = q dz
∫ z
(6.5)

Divide this last equation by ∆z and take the limit as ∆z → 0 to get

dF yz
----------- = q (6.6)
dz
which shows that the shear flow 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
– F + F + ∆F + ∆F yz = 0

and if we divide this by ∆z and take the limit as ∆z → 0 we get

dF dF yz
------- + ----------- = 0
dz dz
Substitute eq. (6.3) for F and (6.6) for the second term in this equation we get
dM Q x ( y )
----------x -------------
-+q = 0
dz I xx

Moment equilibrium of the beam requires V y = dM x ⁄ dz , so that the shear flow is given by

V
q = – -----y- Q x ( y ) (6.7)
I xx

If we assume that the shear stresses are uniform across the width of the beam, then q = τ yz t , and we get

Vy Qx ( y )
τ yz = – -------------------
- (6.8)
I xx t

Since τ zy = τ yz , eq. (6.8) is also the formula for the shear stress in the cross section at y. It is important to note
that Q x ( y ) is the first area moment of a portion of the cross-sectional area about the centroidal x-axis. For the
rectangular section, I xx = ( h 3 t ) ⁄ 12 and Q x ( y ) is given by eq. (6.2), so that we finally obtain

3 V 2y 2 h h
τ zy = ---  -----y 1 –  ------ – --- ≤ y ≤ ---
2  ht   h 2 2
The shear stress distribution is parabolic in the cross section, with a maximum value of 1.5 times the average

Thin-Walled Structures 117


Article 6.2

shear stress, where τ ave = V y ⁄ A = V y ⁄ ( ht ) , and the shear stress is zero at the bottom and top of the beam since
τ yz = 0 at y = −
+ h ⁄ 2 . This distribution is sketched in Fig. 6.3. The shear stress distribution is statically equiv-

h/2
1.5 τ zy
--------
-
τ ave
h/2

Fig. 6.3 Parabolic shear stress distribution in the rectangular section beam.

alent to the shear force. That is,


h⁄2 h⁄2
t⁄2 t⁄2  3  V y  2y 
2
Vy = τ zy dx dy =  ---  ----- 1 –  ------ dx dy = V y
∫ ∫ –t ⁄ 2
∫ ∫ – t ⁄ 2  2 ht h 
–h ⁄ 2 –h ⁄ 2

6.2 Shear flows in open section beams


Consider the thin-walled open cross section shown in Fig. 6.4. Let s denote the contour coordinate, which is

t
---
Assume 2
Vy y 0 ≤ τ zn « τ zs q = zs dn
τ zs
∫τ
t
Vx τ zn – ---
C 2
x
s n
t s q

Fig. 6.4 Shear forces, shear stresses, and shear flow in a general thin-walled open cross section

defined 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 – t ⁄ 2 ≤ n ≤ t ⁄ 2 , where
t denotes the thickness of the wall. Shear forces Vx and Vy are the resultants of the distribution of the shear stress
components τ zs and τ zn , where τ zs is acts tangent to the contour and τ zn acts in the direction normal to the con-
tour, that is, in the thickness direction. The shear stress component conjugate to τ zn is τ nz , and on the stress-free
lateral surfaces where n = ± t ⁄ 2 component τ nz = 0 . Since τ zn = τ nz , τ zn = 0 at n = ± t ⁄ 2 as well. In thin

118 Thin-Walled Structures


Shear flows in open section beams

wall bar theory we neglect the thickness direction component τ zn with respect to the tangential component τ zs ,
because τ zn = 0 at n = ± t ⁄ 2 and the wall is thin so that it appears that the thickness shear stress component is
not much different from zero. The shear flow is defined as the definite integral across the thickness of the shear
stress component tangent to the contour; i.e.,
t⁄2
q = τ zs dn
∫ –t ⁄ 2
(6.9)

The shear flow 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 flow at s, or q(s), can be related to the shear flow at s = 0, or q(0), by axial equilibrium. A free
body diagram of a beam segment defined 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 σz is

q(s)
F
s
q(s)
q(0) q(0)

F + ∆F
∆z
s = 0

Fig. 6.5 Free body diagram of a segment of a branch in the beam.

s
F =
∫ σ t ds
0
z

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
flexure formula for σz, eq. (4.27) on p. 69 , into the above equation we get

I xx M y – I xy M x I yy M x – I xy M y
F = --------------------------------
2
- Q y ( s ) + --------------------------------
2
- Qx ( s ) (6.10)
I xx I yy – I xy I xx I yy – I xy
where the first area moments with respect to the centroidal axes x and y are
s s
Qx ( s ) =
∫ y ( s )t ds
0
Qy ( s ) =
∫ x ( s )t ds
0
(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., x = x ( s ) and y = y ( s ) on the contour. These coordinate functions x(s) and y(s) are required to compute the
first area moments defined by eqs. (6.11). Note that the thickness of the wall can be, in general, a function of con-
tour coordinates as well, as long as t(s) remains small with respect to the overall cross-sectional dimensions.

Thin-Walled Structures 119


Article 6.2

On the s-faces of the free body diagram in Fig. 6.5, the axial forces are approximated by the shear flow
times the small length ∆z. The shear flows at s = 0 and at s are not the same. Note that the shear flow acting on the
positive z-face is assumed positive in the positive s-direction, so that the positive shear flow on the positive s-face
is in the positive z-direction and a positive shear flow 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
F + ∆F – F + q ( s )∆z – q ( 0 )∆z = 0

Divide this equation by ∆z and then take the limit as ∆z → 0 to get

dF
------- + q ( s ) – q ( 0 ) = 0 (6.12)
dz
The derivative of eq. (6.10) is

dF I xx V x – I xy V y I yy V y – I xy V x
------- = ------------------------------Q
2 y ( s ) + ------------------------------Q
2 x(s)
dz I xx I yy – I xy I xx I yy – I xy

where we used moment equilibrium of the beam about the x-axis, or V y = dM x ⁄ dz , and about the y-axis, or
V x = dM y ⁄ dz . Substituting this last result into eq. (6.12), we get the formula for the shear flow due to trans-
verse shear forces as

I xx V x – I xy V y I yy V y – I xy V x
q ( s ) = q ( 0 ) – ------------------------------Q
2 y ( s ) – ------------------------------Q
2 x(s) (6.13)
I xx I yy – I xy I xx I yy – I xy

EXAMPLE 6.1 Shear flow 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
5
shear forces V x = 0, V y > 0 . The second area moment I xx = ------ h 3 t and the product area moment I xy = 0 .
24
Determine the shear flow distribution and the maximum shear stress magnitude.

y
h/2 h/2 s,q s,q

x
C
h 3
--- h s,q
4
t, typical

Fig. 6.6 Symmetric tee-section beam.

Solution A separate contour coordinate is selected in the web, left flange, and right flange. 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

120 Thin-Walled Structures


Shear flows in open section beams

sense of positive shear flow 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
V
q ( s ) = – -----y- Q x ( s )
I xx

In the web, the cartesian coordinates of the contour are


3
x(s) = 0 y ( s ) = – --- h + s 0≤s≤h
4
The first area moment about the x-axis for the web, refer to the first of eqs. (6.11), is
s s
 3   3 s 2
Qx ( s ) = y ( s )t ds =
∫ 0
∫  – --4- h + s t ds = t  – --4- hs + ---2-
0

Substitute this in to the shear flow formula to get

V Vy  3 18 V 2s s
- – --- h + --s- st = ------ -----y  1 – --- --- ---
q ( s ) = – -----y- Q x ( s ) = – ------------- 0≤s≤h (6.14)
I xx 5 3  4 2  5 h 3 h h
------ h t
24

6V
The web shear flow is quadratic in the branch contour coordinate s, is zero at s = 0, and q ( h ) = --- -----y . We set the
5h
derivative of the shear flow with respect to s equal to zero to find the location where the shear flow may have a
maximum magnitude in the web. Hence

 dq 2 s 2s 1 3
------ = 0 →  – ------ --- +  1 – --- --- --- = 0, or s = --- h
 ds   3h h  3 h h 4

3 27 V
This is the location of the centroid for the entire section, and q  --- h = ------ -----y .
4  20 h

The x and y coordinates of the contour in the left flange are


h h h
x ( s ) = – --- + s y ( s ) = --- 0 ≤ s ≤ ---
2 4 2
The first area moment about the x-axis is
s s
Qx ( s ) = y ( s )t ds =  h 1
∫ 0
∫  --4- t ds = --4- hts
0

and hence the shear flow in the left flange is

Vy h 6V s h
- --- ts = – --- -----y  ---
q ( s ) = – ------------- 0 ≤ s ≤ --- (6.15)
5 4 5 h  h 2
------ h 3 t
24
Thus, the shear flow in the left flange is linear in the contour coordinate, and at the junction with the web
h 3V
q  --- = – --- -----y .
 2 5h

Thin-Walled Structures 121


Article 6.2

The x and y coordinates of the contour in the right flange are


h h h
x ( s ) = --- – s y ( s ) = --- 0 ≤ s ≤ ---
2 4 2
The first area moment about the x-axis is
s s
Qx ( s ) = y ( s )t ds =  h 1
∫ 0
∫  --4- t ds = --4- hts
0

and hence the shear flow in the right flange is

Vy h 6V s h
- --- ts = – --- -----y  ---
q ( s ) = – ------------- 0 ≤ s ≤ --- (6.16)
5 4 5 h  h 2
------ h 3 t
24
Thus, the shear flow in the right flange is linear in the contour coordinate and at the junction with the web
h 3V
q  --- = – --- -----y .
 2 5h

The shear flow 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 flows in the two flanges are given by the same equation, and are neg-
ative for Vy > 0. In Fig. 6.7 positive values of the flange shear flows are shown, which means that the arrow indi-
cating the sense of the shear flow in the flanges is drawn in the opposite direction to what was originally assumed
positive. Note that the shear flow into the junction from the web is equal to the sum of the shear flows going out
of the junction into the flanges. The physical basis of this flow relation at the junction is axial equilibrium at the
junction.

3 Vy
--- -----
5h

6 Vy 3 Vy 3 Vy
--- ----- --- ----- --- -----
Vy q 5h 5h
q 5h

27 V y 6 Vy
q ------ ----- --- -----
3 20 h 5h
--- h
4 Shear flows at the junction
of the web and flanges

Fig. 6.7 Shear flow distribution in the tee-section.

27 V
The maximum shear flow magnitude is equal to ------ -----y , and occurs in the web at the centroid of the cross
20 h
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 flow. Thus,

122 Thin-Walled Structures


Shear flows in open section beams

27 V 27 V
τ max = ------ -----y = ------ 2  -------y- = 2.7τ ave
20 ht 20  2ht
where the average shear stress is defined as the shear force Vy 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!

EXAMPLE 6.2 Shear flow in an equal leg angle section

The figure below (on the left) shows the cross-section of a beam, with the origin of x-y system at the centroid..
5a 3 t a3t
The moments of inertia are I xx = I yy = ---------- , and I xy = ------- . For an applied shear force given by V y :
24 8

a) Calculate the shear flow distribution on segments AB and BC in terms of coordinates s 1 and s 2 (as
shown in the figure on the right).
b) Calculate the position of the point on AB where the shear flow is maximum.
c) Calculate the maximum shear flow and the maximum shear stress.
d) Calculate the point on BC (other than C) where the shear flow is zero.
Vy
a
a⁄4 s2
B C
a y y
---
4 x t x
a
t s1

Solution (a) the shear flow is given by (for V x = 0 )

I xy Q y ( s ) – I xx Q x ( s )
q = q 0 + ----------------------------------------------
2
- Vy
I xx I yy – I xy

For segment AB:


y ( s 1 ) = – 3a ⁄ 4 + s 1 x ( s1 ) = –a ⁄ 4

Q x ( s 1 ) = – 3ats 1 ⁄ 4 + ts 12 ⁄ 2 Q y ( s 1 ) = – ats 1 ⁄ 4

( 1 ⁄ 8 ) ( – ats 1 ⁄ 4 ) – ( 5 ⁄ 24 ) ( – 3ats 1 ⁄ 4 + ts 12 ⁄ 2 ) ( 3 ) ( – 6ats 1 ) – ( 5 ) ( – 18ats 1 + 12ts 12 )


q ( s 1 ) = ------------------------------------------------------------------------------------------------------------------V y = - Vy
----------------------------------------------------------------------------------------
a 3 t [ ( 5 ⁄ 24 ) ( 5 ⁄ 24 ) – ( 1 ⁄ 8 ) 2 ] a3t( ( 5 )( 5 ) – 32 )

18as 1 – 15s 12 3V
q ( s 1 ) = ------------------------------
-V y q B = ---------y
4a 3 4a
For segment BC:

Thin-Walled Structures 123


Article 6.3

y ( s2 ) = a ⁄ 4 x ( s2 ) = – a ⁄ 4 + s2

Q x ( s 2 ) = ats 2 ⁄ 4 Q y ( s 2 ) = – ats 2 ⁄ 4 + ts 22 ⁄ 2

( 1 ⁄ 8 ) ( – ats 2 ⁄ 4 + ts 22 ⁄ 2 ) – ( 5 ⁄ 24 ) ( ats 2 ⁄ 4 ) 3V y ( 3 ) ( – 6ats 2 + 12ts 22 ) – ( 5 ) ( 6ats 2 )


q ( s 2 ) = q B + -----------------------------------------------------------------------------------------------------------
- V y = - Vy
--------- + ----------------------------------------------------------------------------------
a 3 t ( ( 5 ⁄ 24 ) ( 5 ⁄ 24 ) – ( 1 ⁄ 8 ) 2 ) 4a a3t( ( 5 )( 5 ) – ( 3 )2 )

3a 2 – 12as 2 + 9s 22
q ( s 2 ) = -------------------------------------------V y
4a 3

One can check that the above expression gives q C = 0 .

(b) The shear flow is maximum where the slope of the shear flow function is zero. Thus for segment AB:

18as 1 – 15s 12 dq V 3
q ( s 1 ) = ------------------------------
-V y -------- = -------y3- ( 18a – 30s 1 ) = 0 ⇒ s 1 = --- a
4a 3 ds 1 4a 5

(c) The maximum shear flow and corresponding shear stress is:

18a ( 3a ⁄ 5 ) – 15 ( 3a ⁄ 5 ) 2 135V y Vy
q max = ------------------------------------------------------------V y = --------------- = 1.35 -----
4a 3 100a a

q max V
- = 1.35 -----y
τ max = ---------
t at

Note that the average shear stress is the shear force divided by the cross-sectional area, or τ ave = V y ⁄ ( 2at ) . The
maximum shear stress is then 2.7 times the average shear stress, or τ max = 2.7τ ave . A coincidence with Example
6.1.

(d) The shear flow on segment BC is zero when:

3a 2 – 12as 2 + 9s 22 2 2
-------------------------------------------V y = 0 ⇒ a – 4as 2 + 3s 2 = 0
4a 3

s2 = a ⁄ 3 s2 = a

6.3 Shear center of an open section


Consider a thin-walled channel section with a web height denoted by h, with flanges of equal width denoted by b,
and with the web and flanges having the same thickness denoted by t. The section is subjected to a vertical shear
force Vy > 0 and no horizontal shear, or Vx = 0. See Fig. 6.8. The objective is to determine the shear flow distri-
bution in the cross section, and then use static equivalence to determine the resultant of the shear flow 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 Ixy =
0. The second area moment about the x-axis is

124 Thin-Walled Structures


Shear center of an open section

b s,q
y
y
h/2
Vy C C
x x

h/2
t, typical s,q s,q
b
Fig. 6.8 Thin-walled channel section subjected to vertical shear.

1 h 2 1 1
I xx = ------ h 3 t + 2 0 +  --- bt = ------ h 3 t + --- h 2 bt
12  2 12 2
The positive contour coordinate directions and shear flows in each branch of the cross section are defined as
shown in Fig. 6.8. To determine the shear flows for this problem, eq. (6.13) reduces to
V s
q ( s ) = q ( 0 ) – -----y- Q x ( s ) where Q x ( s ) =
I xx ∫ y ( s )t ds
0

In the lower flange, s = 0 at a free edge so that q(0) = 0, and y ( s ) = – h ⁄ 2 for 0 ≤ s ≤ b . Thus, the first area
moment of the lower flange segment is Q x ( s ) = – ( hts ) ⁄ 2 , and the shear flow is

V ht
q ( s ) = -----y- ----- s 0≤s≤b (6.17)
I xx 2

The web shear flow at the junction with the lower flange, q(0), is equal to the shear flow of the lower flange
at s = b. This, again, is a consequence of axial equilibrium at the junction. Hence, for the web we have:
q ( 0 ) = ( V y hbt ) ⁄ ( 2I xx ) from eq. (6.17), y ( s ) = – h ⁄ 2 + s for 0 ≤ s ≤ h , and

s
 h  t 2
Qx ( s ) =
∫  – --2- + s t ds = --2- ( – hs + s )
0

Combining these results, the shear flow in the web is


Vy t
- ( hb + hs – s 2 )
q ( s ) = -------- 0≤s≤h (6.18)
2I xx

The upper flange shear flow at the junction with the web, qweb(0), is equal to the web shear flow at s = h.
Hence, from eq. (6.18) we get q web ( 0 ) = q flange ( b ) = ( V y hbt ) ⁄ ( 2I xx ) . For the upper flange: y ( s ) = h ⁄ 2 for
0 ≤ s ≤ b , so Q x ( s ) = ( hts ) ⁄ 2 . The shear flow in the upper flange is

Thin-Walled Structures 125


Article 6.3

Vy t
q ( s ) = --------
- ( hb – hs ) 0≤s≤b (6.19)
2I xx
Note that the upper flange shear flow at s = b is zero. This is as it should be, since the upper flange 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 flow computa-
tions, and we would have to go back at this point to find it. The shear flow distributions given by eqs. (6.17) to
(6.19) are shown schematically in Fig. 6.9(a).
V y hbt
--------------
2I xx V y hb 2 t
F 3 = ----------------
4I xx

y y

V y hbt  h
C C
- 1 + -----
------------- - x ⇒ x
2I xx  4b F2 = Vy
static equivalence

V y hb 2 t
F 1 = ----------------
V y hbt 4I xx
--------------
2I xx

a) Shear flow distribution b) Branch forces


Fig. 6.9 Shear flow distribution due to the transverse shear force and the statically equivalent branch
forces for the channel section.

Next we determine the resultant of the shear flows in several steps using static equivalence. First, for
straight branches we can integrate the shear flow along the contour to get the branch force. The line of action of
b
the branch force is the contour line. For the lower flange, the branch force is defined by F 1 =
∫ q ( s ) ds . Substi-
0
tute eq. (6.17) for the shear flow to get
b V y ht 
 ----- V y hb 2 t
F1 = - ----- s ds = ----------------
∫ 
0 I xx 2
 4I xx
(6.20)

h
For the web, the branch force is defined by F 2 =
∫ q ( s ) ds . Substitute eq. (6.18) for the shear flow to get
0

 Vy t Vy t V 1
h
 h3 h3 h3t
F2 = 2 2
- ( hb + hs – s ) ds = --------- h b + ----- – ----- = -----y- --- h 2 bt + -------
∫  --------
0 2I xx
 2I 2 3xx I xx 2 12

1 1
but I xx = ------ h 3 t + --- h 2 bt , so that
12 2

126 Thin-Walled Structures


Shear center of an open section

F2 = Vy (6.21)

b
For the upper flange, the branch force is defined by F 3 =
∫ q ( s ) ds . Substitute eq. (6.19) for the shear flow
0
to get
b Vy t
 -------- V y hb 2 t
F3 = - ( hb – hs ) ds = ----------------
∫ 
0 2I xx
 4I xx
(6.22)

Note that the branch forces in each flange 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 flows. Since the flange branch forces are the same magnitude, we let F1 = F3 =
1 1
Fflange, and use I xx = ------ h 3 t + --- h 2 bt to find
12 2

 
Vy  hb 2 t 3b 2 
 = V  --------------------
F flange = ----
- ---------------------------------- y 2 - (6.23)
41 3 1 2  h + 6hb 
 ------ h t + --- h bt
12 2

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

a a a
b b b
F
F
F
F

F
d(F)
d a d a d
a
b b b

center of the web to resolve the force system, we get


R x = F flange – F flange = 0 Ry = Vy C z = – hF flange (6.24)

where Rx and Ry are the x- and y-direction components of the resultant force, and Cz 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 Vy to a parallel position to eliminate
the couple and maintain force equivalence. Force Vy is moved such that the moment of Vy in its new position
about the center of the web is equal to Cz. This single force along a specific line of action is the resultant of the
coplanar force system. That is,

Thin-Walled Structures 127


Article 6.3

eV y = hF flange

where e denotes the perpendicular distance between the lines of action of Vy in its original position and its final
position. Note that the force Vy must be move to the left of its original position. The result of this final reduction
is shown in Fig. 6.10(b). The resultant is a single force of magnitude Vy 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).

Vy

Vy
y y
hF flange
C S.C. C
x ⇒ x
static equivalence hF flange 3b 2
e = ------------------
- = ---------------
Vy h + 6b

a) Force and couple at the center of the web b) Resultant of the planar force system

Fig. 6.10 Reduction of branch forces to a single resultant acting at the shear center

The location of the shear center (S.C.) in the cross section is determined by the shear flows due to transverse
shear forces. The computations given above for the channel section illustrate the general conclusion that the S.C.
location depends on the pattern of the shear flow distribution, and not on the magnitude of the shear force. Shear
dV
forces act in the plane of loading to equilibrate the applied loads. (Recall that ---------y = – p y ( z ) and
dz
dV x
--------- = – p x ( z ) .) The line of action of the lateral loads p y ( z ) and p x ( z ) must pass through the shear center if the
dz
beam is to bend without 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.
S.C.

S.C.

S.C.

Fig. 6.11 Shear center locations for open sections with straight branches and one junction

128 Thin-Walled Structures


Shear center of an open section

The location of the shear center for a closed section beam cannot be determined using only the steps out-
lined above for the open section. In a closed section the shear flow at the origin of the contour in eq. (6.13) is an
additional unknown. To determine the shear flow 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 167 .

EXAMPLE 6.3 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

a s,q

y y free edge
2t
C C
2a x x
a 3a/8
2t t

A
2a
Fig. 6.12 Unsymmetrical thin-walled channel section

the centroid is shown in Fig. 6.12, and the second area moments about the centroidal axes are

16 53
I xx = ------ a 3 t I yy = ------ a 3 t I xy = – a 3 t
3 24

Solution To find the vertical location of the shear center we can consider shear forces V x > 0 and V y = 0 . We
do not need to compute the shear flows in each branch to find 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 con-
venient to use the junction of the lower flange 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 flange contributes to the moment
about point A. Then static equivalence of the moment of the branch force and Vx about point A will give the loca-
tion of the line of action for shear force Vx.

First the shear flow in the upper flange is determined using the positive senses for q and the contour coordi-
nate s as shown in Fig. 6.12. For Vy = 0 and q(0) = 0, eq. (6.13) reduces to

( – I xx Q y + I xy Q x )V x
q ( s ) = -----------------------------------------------
2
- (6.25)
I xx I yy – I xy
The cartesian coordinates of the contour coordinate s in the upper flange are
5
x ( s ) = --- a – s y(s) = a 0≤s≤a (6.26)
8

Thin-Walled Structures 129


Article 6.3

The first area moments of the portion of the contour from s = 0 to s of the upper flange are
s
Qx =
∫ ( a )2t ds = 2ast
0
(6.27)

s
5  5 s 2
Qy =
∫  --8- a – s 2t ds = 2t  --8- as – ---2-
0
(6.28)

2 = ( 97a 6 t 2 ) ⁄ 9 . Substituting these results into eq.


The factor in the denominator of eq. (6.25) is I xx I yy – I xy
(6.25) we get

9V 32 5 s2
q ( s ) = ------ ----6-x – 2a 4 s – ------ a 3  --- as – ---- 0≤s≤a (6.29)
97 a 3  8 2
The branch force in the upper flange is the integral of this shear flow over the upper flange; i.e.,
a 23V x
F =
∫ q ( s ) ds = – -----------
0 97
- (6.30)

Static equivalence means that the moment of the shear force Vx about point A must equal the moment of the
branch force F about point A, as is illustrated in the sketch below. Let ey denote the moment arm for the shear

⇔ Vx
2a
ey
A A

force. Static equivalence for the moment about point A gives

2aF = – e y V x (6.31)

Although the branch forces for lower flange 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 ey to get

46
e y = ------ a (6.32)
97
This result for ey locates the line of action of the shear force Vx for beam bending. The shear center is on this par-
ticular line of action of Vx, and we can locate the exact point on this line of action by considering the separate
problem of V x = 0 and V y > 0 .

Now consider V x = 0 and V y > 0 . Equation eq. (6.13) becomes

( I xy Q y – I yy Q x )V y
q ( s ) = ------------------------------------------
2
- (6.33)
I xx I yy – I xy
Substitute eq. (6.27) for Qx and eq. (6.28) for Qy into this equation to get the shear flow in the upper flange as

130 Thin-Walled Structures


Shear center of an open section

9V 53 5 s2
q ( s ) = ------ ----6-y – ------ a 4 s – 2a 3  --- as – ---- 0≤s≤a (6.34)
97 a 12  8 2
The branch force in the upper flange is
a
45
F =
0 194
-V
∫ q ( s ) ds = – -------- y (6.35)

Moment equivalence about point A with the moment arm for shear force Vy denoted as ex, gives

2aF = e x V y (6.36)

See the sketch below. Thus, the location of the line of action of the shear force Vy is

⇔ ex Vy
2a
A A

45
e x = – ------ a (6.37)
97
The location of the shear center is shown in the sketch below.

S.C. 46
------ a
97

45
------ a
97

Thin-Walled Structures 131


Article 6.4

6.4 Summary notes on the


y
C shear center
Locus of S.C.s py x
1.The lines of actions of the lateral loads, e.g., p x ( z )

px z and p y ( z ) , must pass through the shear center if the


beam is to bend and not twist in torsion.
S.C. 2.The shear center location in the cross section
depends on the distribution of the shear flow along
a
the contour and not on the magnitude of the trans-
verse shear forces V x and V y .
y 2t
3.The shear center is located on an axis of symmetry
C in the cross section, if there is one.
x 2a
S.C. 3a/8 a
2t t
6.4.1Procedure to locate the S.C.
1.Take V x ≠ 0 and V y = 0
2a
a)Determine the distribution of the shear flow along
the contour from the equation of axial equilibrium
per unit length of the beam for a portion of the contour from s = 0 to s = s; i.e.,

[ I xx Q y ( s ) – I xy Q x ( s ) ]
q ( s ) = q ( 0 ) – ---------------------------------------------------
2
- Vx
I xx I yy – I xy
b) Torque equivalence about some convenient point, say point A, to find the line of action of shear force
Vx
q1 ( s1 ) Torque about point A, counter-clockwise postive
Tq = –ey Vx
⇔ Vx a
2a
Tq =
A ey
A
∫ ( 2a )q ( s ) ds
1 1
0

2. Take V x = 0 and V y ≠ 0

a) Determine the distribution of the shear flow along the contour from the equation of axial equilibrium per
unit length of the beam for a portion of the contour from s = 0 to s = s; i.e.,

[ – I xy Q y ( s ) + I yy Q x ( s ) ]
q ( s ) = q ( 0 ) – ------------------------------------------------------
2
-V y
I xx I yy – I xy
b) Torque equivalence about some convenient point, say point A, to find the line of action of shear force

132 Thin-Walled Structures


Skin-stringer idealization

Vy
q1 ( s1 )
Torque about point A, counter-clockwise postive

Tq = ex Vy
⇔ ex Vy a
2a
Tq =
A A
∫ ( 2a )q ( s ) ds
1 1
0

6.5 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, Mx
= 15,000 in-lb, and Vy = 10,000 lbs for numerical evaluation. To study what happens when the web thickness t
varies we set t = ε d.

d Area A(s)

y
Vy q(s)
l/2
centroid, C
Mx s
x x-axis

l/2 Web shear flow q(s)


t

d
Geometry of cross section Beam resultants

Fig. 6.13 Two-flanged beam with thin web

M
The flexure formula for the axial normal stress is σ z = -------x y , in which the second area moment of the cross
I xx
section about the x-axis is

2 tl 3 d4 l d 2 2
I xx = d = [ 83.33ε + 60.67 ]d 4
∫ ∫ y dA = -----
A
12
- + 2 ------ + --- + ---
12  2 2
(6.38)

Thin-Walled Structures 133


Article 6.5

The average normal stress in the flange is

M (l + d) M
σ ave = -------x --------------- = -------x ( 5.5d ) (6.39)
I xx 2 I xx
The portion of the total moment carried by the flanges is

 --l- + d  --l- + d
2  2 
M M
( M x ) fl = 2 y ( σ z dA ) = 2 y  -------x y d dy = 2 -------x d y 2 dy
∫ ∫  I xx  I xx ∫
A fl l l
--- ---
2 2

or
( M x ) fl 2d l 3
l 3
-------------- = --- ------  --- + d –  --- = 60.67 ( 83.33ε + 60.67 ) ––1 (6.40)
Mx 3 I xx 2    2
The shear flow in the web is

V Vy  l d 1 l l
q ( s ) = -----y- 2 + ---  --- + s  --- – s t
I xx ∫ ydA = -----
A(s)
- --- + --- d
I  2 2 xx 22  2 
(6.41)
V 1 t l 2 V ε s 2
q ( s ) = -----y- --- ( l + d )d 2 + ---   --- – s 2 = -----y- 5.5 + ---  25 –  ---  d 3
I xx 2 2   2  I xx 2   d 
The portion of the total shear force carried by the web is

l l l l
--- --- --- ---
2 2 2 2
V  l 2 
( V y ) web = τ zy tds = q ( s )ds = 2 q ( s )ds = -----y-  ( l + d )d 2 + t   --- – s 2 ds
∫ ∫ ∫ I xx  ∫   2 

l l 0 0
– --- – ---
2 2

or

( V ) web 1 1 2 l 3 55 + 83.33ε
---------------- = ------ --- ( l + d )ld 2 + --- t  --- = ------------------------------------ (6.42)
Vy I xx 2 3  2 60.67 + 83.33ε
These results are tabulated below for various values of ε.

Flange Flange Web shear flow q, lb/in, eq. (6.41)


Ixx, in4 σave, psi (Mx)fl/Mx (Vy)web/Vy
ε = t/d eq. (6.38) eq. (6.39) eq. (6.40) s=0 s = ±5 in s = ±10, in eq. (6.42)
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
A plot of the shear flow in the web for each ε in the table above is shown in Fig. 6.14.

134 Thin-Walled Structures


Skin-stringer idealization

s ⁄ 10
sêH 10in.inL
1
ε = 0
ε = 1 ⁄ 50
0.5 ε = 1 ⁄ 10
ε = 1⁄4

qêHq500
⁄ 500lb
lb/in.
êinL
0.7 0.8 0.9 1

-0.5

-1

Fig. 6.14 Web shear flow distributions as a function for decreasing web thickness to flange
width ratio.

The assumptions of the skin-stringer approximation are


• The flanges carry all of the bending moment, but no shear force. Note: a flange is considered a stringer.
• Normal stress is uniform over each flange area. In other words, each flange area is assumed to be concentrated
at its centroid for calculation of total section centroid, Ixx, σz, and the first area moment.
• The web carries all of the shear force, but no bending moment.
Only the transfer term in the parallel axis theorem is needed
to compute the second area moment about the x- axis.
y

l+d 2
I xx = 2  ----------- A f = 968in 4
 2 
Af = d2
(l + d)/2 The axial normal stress is uniformly distributed over the
stringer area, and the shear flow is uniformly distributed along
x the web.
(l + d)/2
M l+d
σ z = -------x  ± ----------- = ± 170.45psi
Af I xx  2 

V l+d
q = -----y-  ----------- A f = 454.55 lb/in
I xx  2 

Thin-Walled Structures 135


Article 6.5

EXAMPLE 6.4 Shear flows 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
V x = 0 and V y = 40 kN . The flange of the stringers are assumed to carry only axial forces and each has the
same cross-sectional area A f = 150 mm 2 . Each branch has a thickness t, and dimensions h = 80 mm and
b = 20 mm . Determine the shear flow distributions in each branch for t = 5 mm and t = 0.5 mm and plot
them. Also determine the maximum shear stress for each thickness.

Af b Af
s 3, q 3

y y
h
Vy ---
2

C x C x
S.C.

h
---
2 t, typical s 2, q 2 branch shear flows

s 1, q 1
Af b Af

Fig. 6.15 C-section stiffened by four stringers

Solution From eq. (6.13), the shear flow formula applicable to this problem is
V
q ( s ) = q ( 0 ) – -----y- Q x ( s ) (6.43)
I xx

To use this formula, we should first compute the second area moment I xx of the cross section. Using the thin wall
approximations and the composite body technique, the second area moment about the centroidal x-axis is

h 2 h3t h 2
I xx = 2  --- bt + ------- + 4  --- A f
 2 12  2

136 Thin-Walled Structures


Skin-stringer idealization

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- ε→0
rem affects the second area moment computation for the entire section. Begin
q 1 ( ε )dz
with determining the shear flow in the first branch shown in Fig. 6.15, or the
lower flange. The shear flow at s 1 = 0 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 dN 1 dz

dN lower right stringer


q 1 ( 0 ) = – --------1- (6.44)
dz
But the axial force in the stringer is N 1 = σ z1 A s , and the derivative of this force is dN 1 ⁄ dz = ( dσ z1 ⁄ dz )A f .

h h
From the flexure formula σ z1 = M x  – --- ⁄ I xx , and the derivative dσ z1 ⁄ dz = V y  – --- ⁄ I xx . Thus,
 2  2

dN 1 V
--------- = -----y- Q xs1 (6.45)
dz I xx

where Q xs1 is the first area moment of the lower right stringer about the x-axis given by Q xs1 = ( – h ⁄ 2)A f .
Hence, the shear flow at s 1 = 0 is q 1 ( 0 ) = – ( V y ⁄ I xx )Q xs1 . Substitute this result for q 1 ( 0 ) into eq. (6.43) to
get
V
q 1 ( s 1 ) = – -----y- [ Q xs1 + Q x1 ( s 1 ) ] (6.46)
I xx
s1

The first area moment of the portion of the branch 1 from s 1 = 0 to s 1 is Q x1 ( s 1 ) =


∫ y ( s )t ds
1 1 1, and the y-
0
coordinate to a generic point on the contour is y 1 ( s 1 ) = – h ⁄ 2 . Hence,

h
Q x1 ( s 1 ) = – ---( ts 1 ) 0 ≤ s1 ≤ b
2
Combining these results, the shear flow in branch 1 becomes

V h
q 1 ( s 1 ) = -----y-  --- [ A f + ts 1 ] 0 ≤ s1 ≤ b (6.47)
I xx  2

The shear flow in branch 2, or the web, is determined from the same gen-
eral formula given in eq. (6.43), and the shear flow at s 2 = 0 is determined q 2 ( 0 )dz
separately from the free body diagram of the lower left stringer shown in the
adjacent sketch. Axial equilibrium of the lower left stringer gives
dN q 1 ( b )dz
q 2 ( 0 ) = q 1 ( b ) – --------2- (6.48)
dz dN 2
The derivative of the axial force in the stringer is obtained following the same lower left stringer
procedure used for lower right stringer discussed above, which resulted in eq.
(6.45). The result for the lower left stringer is

Thin-Walled Structures 137


Article 6.5

dN V
--------2- = -----y- Q xs2 (6.49)
dz I xx

where the first area moment about the x-axis of stringer 2 is Q xs2 = ( – h ⁄ 2 )A f . Calculating the shear flow 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 flow at
the beginning of branch 2 from eq. (6.48) is
V
q 2 ( 0 ) = – -----y- [ Q xs1 + Q x1 ( b ) + Q xs2 ] (6.50)
I xx
The term in brackets on the right-hand-side of eq. (6.50) represents the first area moment about the centroidal x-
axis of the lower right stringer, the lower flange or branch 1, and the lower left stringer. For the web, or branch 2,
the y-coordinate to the generic point s 2 is y 2 ( s 2 ) = – h ⁄ 2 + s 2 , so the first area moment of branch 2 is
s2
t
Q x2 ( s 2 ) = = --- ( – hs 2 + s 22 )
∫ y ( s )t ds
2 2 2
2
0

It follows from this result, eqs. (6.43) and (6.50), that the shear flow in branch 2 is

V h t
q 2 ( s 2 ) = -----y- hA f + --- tb – --- ( – hs 2 + s 22 ) 0 ≤ s2 ≤ h (6.51)
I xx 2 2

For the upper flange, or branch 3, the general shear flow formula is again given by eq. (6.43). It follows from
the procedure used to obtain the beginning shear flow in branch 2 that
V
q 3 ( 0 ) = – -----y- [ Q xs3 + Q x2 ( h ) + Q xs2 + Q x1 ( b ) + Q xs1 ] (6.52)
I xx

The y-coordinate to a generic point on branch 3 is y 3 ( s 3 ) = h ⁄ 2 . Hence, the first area moment of the portion of
branch 3 is
s3
h
Q x3 ( s 3 ) =
∫ y ( s )t ds
3 3 3 = --- ts 3
2
0

It follows from this result and eqs. (6.43) and (6.52), that the shear flow in branch 3 is

V h
q 3 ( s 3 ) = -----y-  --- [ A f + tb – ts 3 ] 0 ≤ s3 ≤ b (6.53)
I xx  2

V h
q 3 ( b )dz Note that the shear flow at the end of branch 3 is q 3 ( b ) = -----y-  --- A f . We can
I xx  2 
dz 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
dN 3 gives
upper right stringer dN
– q 3 ( b ) + --------3- = 0
dz
Similar to the results for the derivatives of the stringer forces given in eqs. (6.45) and (6.49), we have

138 Thin-Walled Structures


Skin-stringer idealization

dN V
--------3- = -----y- Q xs3 . Axial equilibrium of the upper right stringer is
dz I xx

V h V h
– -----y- --- A f + -----y- --- A f = 0
I xx 2 I xx 2
which, of course, is satisfied. If axial equilibrium of the upper right stringer were not satisfied, we would have to
go back over our previous calculations to search for an error.

Numerical evaluation and plots Given the data h = 80 mm , b = 20 mm , A f = 150 mm 2 , and


3
V y = 40 ×10 N , the shear flows from eqs. (6.47), (6.51), and (6.53) evaluate as

3
( 40 ×10 ) ( 6000 + 4ts 1 )
q 1 ( s 1 ) = -------------------------------------------------------
- 0 ≤ s 1 < 20 mm (6.54)
320000
960000 + ------------------ t
3

t
( 40 ×10 )  12000 + 800t + 40ts 2 – --- s 22
3
 2 
q 2 ( s 2 ) = ---------------------------------------------------------------------------------------------- 0 ≤ s 2 ≤ 80 mm (6.55)
320000
960000 + ------------------ t
3
and
3
( 40 ×10 ) ( 6000 + 800t – 40ts 3 )
q 3 ( s 3 ) = ---------------------------------------------------------------------------
- 0 ≤ s 3 ≤ 20 mm (6.56)
320000
960000 + ------------------ t
3

The dimensional units of the shear flows are N/mm for the thickness t and branch contour coordinates speci-
fied in mm. These shear flows 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 defined by
s = s1 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

Note that the shear flow distributions exhibit jumps at the stringer locations, and that the shear flow distribution is
more uniform in each branch for the thinner section. The tendency to a spatially uniform shear flow 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 flow occurs at the center of the web. Thus,

 535.7 N/mm t = 5 mm
q max = q 2 ( 40 ) = 
 505.3 N/mm t = 0.5 mm

The maximum shear flow is reduced in the thinner section with respect to the thicker section. The maximum
shear stress is estimated from τ max = q max ⁄ t . Hence,

 107 MPa t = 5 mm
τ max = 
 1011 MPa t = 0.5 mm

Thin-Walled Structures 139


Article 6.6

q, N êmm

500

400

300

200

100

s, mm
20 40 60 80 100 120

Fig. 6.16 Shear flows in the stiffened C-section for t = 5 mm (solid lines) and t =
0.5 mm (dashed lines)

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.

6.6 Problems
1. The thickness of each branch in the thin-walled
y
cross section shown below is 3mm and
I xx = 10 5 mm 4 . The shear force V y = 5 kN .

40 a)Determine the shear flow distribution and sketch


x it on the cross section. Indicate on the sketch the
C
A 10 positive sense along the branch.
50 30 30 50 b)Estimate the shear stress due to the transverse
shear force at point A.
Note: all dimensions in mm
c)Estimate the maximum shear stress due to trans-
verse shear.

140 Thin-Walled Structures


Problems

2. Consider the thin-walled, Y-section beam of Problem 1. on page


y
91, in which we found I xx = 2.0238l 3 t . Also, I yy = l 3 t . Determine the
magnitude, direction, and sense of the shear flow at the centroid if the
45° 45°
shear force components are V x = 0 and V y > 0 .

C x
0.6414l

l l

Note: the length of each branch is l


and the thickness of each branch is t
3. Determine the location of the shear center e from point A for
the thin-walled, open section shown. a


S.C. α A axis of symmetry
α

t

4. Calculate the shear flow distribution for the semi-circular


cross-section shown in the adjacent sketch. (Given:
I xx = πr 3 t ⁄ 2 , V x = 0 , and V y = V . r

5. Show that the shear flow distribution in the cross-section t


shown below for a vertical shear force is given by
6s 6 ( a 2 + as 2 – s 22 ) 6 ( a – s3 )
q ( s 1 ) = -------12-V y q ( s 2 ) = --------------------------------------
- Vy q ( s 3 ) = ---------------------V y
7a 7a 3 7a 2
Calculate the position of the shear center.

a
s3
D C

a
t
s2
A B
s1

Thin-Walled Structures 141


Article 6.6

6. Determine the coordinates (ex,ey), 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

I xx = 0.660155 in. 4 I yy = 0.130305 in. 4 I xy = – 0.127197 in. 4

y
1.5
( x c, y c ) = ( 1.02786, 0.529989 ) y
y S.C.
ey
C x
O x O x
ex
2

t = 0.030 typ.
All dimensions in inches

142 Thin-Walled Structures


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.

y y
a
b
T z T x

T, θ z
L

Fig. 7.1 Circular tube subjected to uniform torsion

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 0 < z < L . 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 configuration, 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.

Thin-Walled Structures 143


Article 7.1

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).

D A

B C
C B
x
∆z ∆z
180°
A D
(a) (b)

Fig. 7.2 Symmetry about the midplane of an element of the torsion tube.

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 fit together in the deformed geometry to form a continuous tube.
Thus, plane sections remain plane.

dish-in bulge-out

∆z ∆z

Fig. 7.3 Potential axial symmetric end surface deformations.

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

144 Thin-Walled Structures


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 conflict 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.

T T

x
∆z ∆z
180°

T T
(a) (b)
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.

Symmetry of deformation has not ruled out symmetrical


expansion or contraction of the circular cross section, or a y
lengthening or shortening of the tube. It does not seem plau- r∆θ
sible that such dilational deformations would be an important ∆r
part of the deformation. Thus, extensional strains of line seg- ∆z r
ments in the axial direction (z-axis), radial direction (r), and
the circumferential direction ( θ ) are assumed zero. See Fig. θ
7.5. That is extensional strains, ε z = 0 , ε r = 0 , and x
εθ = 0 . z

The only motion of a circular cross section is, then, a Fig. 7.5 Cylindrical coordinates and
differential line elements.
rotation about the z-axis. Thus, the right angle between line
elements originally parallel to the r- and θ -directions does
not change, which means the shear strain γ rθ = 0 . 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 γ rz = 0 . In cylin-
drical coordinates the only non-zero strain is γ θz . To compute this shear strain consider an element of the tube of
length ∆z and radius r, where a ≤ r ≤ b . 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 A’B’ as shown in Fig. 7.6. For very small ∆z the seg-
ment A’B’ can be considered a straight line segment forming the angle γ θz with line A’C’ taken parallel to line
AB. This angle is the change of the right angle between originally perpendicular elements ∆z and r∆θ , and is the

Thin-Walled Structures 145


Article 7.1

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

γ θz
A’
θz

B’ A
r θ
∆θ z C’ x
B
z

∆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.

as
r∆θ z = γ θz ∆z

which in the limit as ∆z → 0 gives


γ θz = r --------z (7.1)
dz
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 dθ z ⁄ dz is a constant along a uniform section of
the tube. The quantity dθ z ⁄ dz is the twist per unit length or simply the unit twist.

We use Hooke’s law to relate the shear stress to the shear strain, since we have assumed the material is linear
elastic and isotropic. Hooke’s law is
τ θz = Gγ θz (7.2)

where G denotes the shear modulus of the material. For an isotropic material recall that
E
G = --------------------
2(1 + ν)
where the modulus of elasticity is denoted by E and Poisson’s ratio is denoted by ν. Eliminating the shear strain
by eq. (7.1), we get

τ θz = Gr --------z (7.3)
dz

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 τ θz does not change

146 Thin-Walled Structures


Uniform torsion of a circular tube

in the θ -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 θ -face. Hence the differential
element is in equilibrium. See Fig. 7.7.

y r
θ
τ θz
τ zθ
τ zθ
z x
θ
z τ zθ = τ θz

T
Fig. 7.7 Shear stresses due to torque

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

T =
∫ ∫ r(τ
A
zθ ) dA (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

T = GJ --------z (7.5)
dz
where the torsion constant J is defined as
b 2π
π
J = r 2 dA = r 3 ( dθ ) dr = --- ( b 4 – a 4 )
∫∫
A
∫∫
a 0 2
(7.6)

The torsion constant J is the same as the polar area moment for a circular annulus, and has dimensional units of
L4. From eq. (7.5) we find the twist per unit length is
dθ T
--------z = ------- (7.7)
dz GJ
For tube of length L, the total angle of twist between the ends is found by integrating eq. (7.7); i.e.,
L
T TL
θz ( L ) – θz ( 0 ) = ------- dz = ------
∫ 0 GJ GJ
- (7.8)

Now substitute eq. (7.7) into eq. (7.3) to find


Tr
τ zθ = ------ a≤r≤b (7.9)
J

Thin-Walled Structures 147


Article 7.1

Thin-walled approximations Let the mean radius be denoted by r and the wall thickness be denoted by t. The
relationships between the inside and outside radii to the mean radius and wall thickness are
t
a = r – --- 1
2 r = --- ( a + b )
or the inverse 2 (7.10)
t
b = r + --- t = b–a
2

A thin-walled tube is quantified by 0 < t ⁄ r « 1 ; e.g. t ⁄ r ≤ 1 ⁄ 20 . The torsion constant, eq. (7.6), becomes
4
π t 4 t 4 πr t 4 t 4
J = ---  r + --- –  r – --- = --------  1 + ----- –  1 – -----
2  2  2 2  2r
 
2r

4
πr t t 2 t 3 t 4
J = --------  1 + 4  ----- + 6  ----- + 4  ----- +  -----  –
2   
2r
 
2r
 
2r
  
2r

 1 – 4  ----t- + 6  ----t- 2 – 4  ----t- 3 +  ----t- 4


         
2r 2r 2r 2r

π 4 t t 3 3 t 2
J = --- r 8  ----- + 8  ----- = 2πr t 1 +  -----
2      
2r 2r 2r

Since the quantity 0 < ( t ⁄ 2r ) « 1 , we can neglect this factor squared with respect to one in the above equation to
get
3
J = 2πr t thin-walled circular tube (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

Tr r t r t
τ zθ = ------  -  1 –  ----- ≤ - ≤ 1 +  -----
J  r  
2r r
 
2r

which shows that the range of r ⁄ r 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.

Tr
τ zθ = ------ thin-walled circular tube (7.12)
J
Since the shear stress is assumed to be uniform across the wall thickness, we define the shear flow due torsion q
as q = τ zθ t . Substituting eq. (7.12) for the shear stress and eq. (7.11) for the torsion constant, the shear flow is
written as

Trt T
q = ------------
3
- = --------- (7.13)
2πr t 2A c

2
where the area enclosed by the mean radius is A c = πr . 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

148 Thin-Walled Structures


Uniform torsion of an open section

tube. Equation (7.13) is called the Bredt-Batho formula. These torsion results for the thin-walled circular tube
are depicted in Fig. 7.8., in which we used the contour coordinate s rather than the angle θ to write the shear
stress as τ zθ = τ zs

y T
q = ---------
2A c τ zs = q ⁄ t

r
s = rθ
θ
x
t
T

t
Ac
Fig. 7.8 Shear flow and shear stress in a thin-walled circular tube subjected to a torque.

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

y
T

t
0 < --- « 1 x
b
z
t T
L

b
Fig. 7.9 Uniform torsion of a bar with a rectangular cross section.

problem of uniform torsion of bars with non-circular cross sections was first 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

Thin-Walled Structures 149


Article 7.2

y T y

x x
z
w

T
w ( x, y ) = – ------- xy
GJ

T
(a) Rotation and warping of the cross sections (b) Warping of the cross section
Fig. 7.10 Deformation of a thin-walled rectangular section bar under uniform torsion

non-circular section relative to the circular section bar. (Refer to the third symmetry argument of Section 7.1.)
The out-of-plane displacement w ( x, y ) is called warping function of the cross section (Fig. 7.10b). For the thin-
walled rectangular section this warping displacement is approximated by

w ( x, y ) = – xy --------z (7.14)
dz
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 axial 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
dθ z T
-------- = --------------------- (7.15)
dz 1
G  --- bt 3
3 

dθ T
Comparing this result to the standard torsion formula --------z = ------- , we find the torsion constant is
dz GJ

1
J = --- bt 3 (7.16)
3
Shear stress components due torsion are τzx and τzy, as shown in the adjacent sketch. The dom-
τ zy inate stress under free warping of the narrow rectangular section is the shear stress tangent to
t τ zx the long side, τzx, and it is approximated by

dθ t t
τ zx = – 2G  --------z y – --- ≤ y ≤ --- (7.17)
 dz  2 2
b
The shear stress component τzy, is negligible except near the ends at x = ± --- , where component τzx must vanish
2

150 Thin-Walled Structures


Uniform torsion of an open section

b
because the lateral surfaces of the bar at x = ± --- are stress free. The shear stress distribution through the thick-
2
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
a maximum magnitude at y = ± t ⁄ 2 . Combining eqs. (7.15) and (7.17), the magnitude of the maximum shear
stress is
3T
τ max = ------2- (7.18)
bt

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
dθ z T T
-------- = ---------------------
- τ max = -----------2- (7.19)
dz G ( k 1 bt 3 ) k 2 bt
where the dimensionless parameters k1 and k2 are functions of the ratio of b/t. These parameters are listed for var-
ious b/t ratios in the table below. 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 flow is calculated from the shear stress distribution given by eq. (7.17) we get
t⁄2 t⁄2 dθ z 
 – 2G  ------- dθ t⁄2
q = τ zx dy = - y dy = – 2G  --------z y dy = 0
∫ –t ⁄ 2
∫ –t ⁄ 2
  dz    dz  ∫ –t ⁄ 2

That is, the shear flow due to a linear shear stress distribution through the thickness of the wall is zero. The shear
stresses are not zero, but the shear flow 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.

Table of parameters for torsion of a rectangular cross section bara

Width to thickness ratio b/t Unit twist parameter k1 Shear stress parameter k2
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

a. Timoshenko and Goodier, 1970, p.312

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 flat membrane supported on its edges over an opening, where the edges of the opening are

Thin-Walled Structures 151


Article 7.2

b2 b2 b2

t2 t2
b
b1 t b1
t1 t1

1 1 1
J = --- ( b 1 t 13 + 4b 2 t 23 ) J = --- bt 3 J = --- ( b 1 t 13 + 2b 2 t 23 )
3 3 3

Fig. 7.11 Some thin-walled open sections and their torsion constants

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 deflect. The deflected shape of this pressurized membrane is anal-
ogous to the torsion problem in that level contours on the surface of the deflected 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-
portional to the magnitude to the shear stress. Also, the volume between the x-y plane and the deflected 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 first discovered by Ludwig Prandtl in 1903 and is known’s as Prandtl’s 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 film 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

dθ z dθ
T = ∑
branches
Ti = ∑ GJ  -------
i
dz 
- = GJ  --------z
 dz 

where the torsion constant for the entire cross section is


1 3
J = ∑
branches
Ji = ∑
branches
--- b i t i
3
(7.20)

Note that the twist per unit length is the same for all branches in the open section, because the cross section is
152 Thin-Walled Structures
Uniform torsion of an open section

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 ith branch of the section is given by

3T 3GJ i dθ z 3G 1 T Tt
- -------- = --------2-  --- b i t i3  ------- = ------i
( τ max ) i = --------2-i = ----------- (7.21)
bi ti 2
b i t i dz bi ti 3    GJ  J

That is, the maximum shear stress in the ith branch of the open section is the total torque divided by the torsion
constant for the entire section times the thickness of the ith 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.

small slit

t t
r r

T T

Fig. 7.12 Closed and open thin-walled circular sections

T
- from eq. (7.13), and the torsion constant J = 2πr 3 t
Solution For the closed section the shear flow q = ---------------
2 ( πr 2 )
from eq. (7.11). Hence the maximum shear stress and unit twist are

T
τ closed = ------------
-  dθ
--------z T T
= ------- = ----------------------
-
2πr 2 t  dz  closed GJ G ( 2πr 3 t )

For the open section the developed length b of the contour is essentially 2πr , 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 b = 2πr , we have

3T
τ open = ------------2-  dθ
--------z T
= ---------------------------
2πrt  dz  open 1
G  --- 2πrt 3
3 

Forming the ratio of the maximum shear stress of the open section to the closed section we find

Thin-Walled Structures 153


Article 7.3

τ open 3T 2πr 2 t r
- = ------------2- ⋅ ------------- = 3 - » 1
--------------
τ closed 2πrt T t

Likewise, the ratio of the unit twists are


( dθ z ⁄ dz ) open 3T G2πr 3 t r 2
- = -----------------3- ⋅ ------------------ = 3  -  » 1
----------------------------------
( dθ z ⁄ dz ) closed G2πrt T  t

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


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 tz (units of F-L/L, or F) acts on the bar. This
intensity is a mathematical function of z, so we denote it as tz(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-
tz ( z )

z = 0 z = L

z t z ( z ¥)dz, z < z * < z + dz

T T + dT
z z + dz

Fig. 7.13 Differential element of a bar subjected to a distributed torque

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 figure to get

T + dT – T + t z ( z * )dz = 0 or dT + t z ( z * )dz = 0

Divide the last equation by dz, take the limit as dz → 0 , and note that z * → z in the limit, to get

dT
------ + t z = 0 0<z<L (7.22)
dz

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.

154 Thin-Walled Structures


Non-uniform torsion

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
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,
at z = 0 and z = L prescribe either T or θ z but not both. (7.23)

For a linear elastic material, the constitutive equation of the bar is

dθ z
T = GJ (7.24)
dz
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

d  dθ z
GJ = –tz ( z ) 0<z<L
dz dz 
If the torsional stiffness is uniform along the length, then this equation reduces to
2

GJ 2 z = – t z ( z ) θz = θz ( z ) 0<z<L (7.25)
dz
This is a second order, linear differential equation for the angle of twist θz and requires two boundary conditions
to determine its solution. These boundary conditions are specified 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 fixed ends.

Consider a uniform bar subjected to a uniform distributed torque of intensity t z ( z ) = t z0 = constant , 0 < z < L .
Each end of the bar is fixed to a rigid support which prevents the twisting rotation of the bar so that θ z ( 0 ) = 0
and θ z ( L ) = 0 . Determine distributions of the angle of twist θz(z) and the internal torque T(z) along the length
of the bar.

Solution The governing differential equation, eq. (7.25), reduces to


2
d θz
GJ = – t z0 0<z<L
dz2
divide this equation by the torsional stiffness and integrate twice with respect to z to get
t z0 z 2
- ---- + c 1 z + c 2
θ z = – ------
GJ 2
where c1 and c2 are constants of integration. These constants are determined from the boundary conditions. At
the left end we have

Thin-Walled Structures 155


Article 7.3

t z0 0 2
- ----- + c 1 0 + c 2 = 0
θ z = – ------
GJ 2
So c2 = 0. Using this fact in the boundary condition at the
qz qzmax t z0 L 2 right end, we get
1 θ zmax = -----------
-
8GJ
t z0 L 2
0.8 - ----- + c 1 L = 0
θ z = – ------
GJ 2
0.6
0.4 t z0 L
So c 1 = ---------
- . Hence, the solution for the angle of twist is
0.2 2GJ
0.2 0.4 0.6 0.8 1 z L t z0
- ( Lz – z 2 )
θ z = --------- 0≤z≤L (7.26)
2GJ
T Tmax t z0 L
1 T max = ---------
- The torque is obtained by substituting this result, eq.
2 (7.26), for the twist angle in the material law, eq. (7.24), to
0.5
get
0.2 0.4 0.6 0.8 1 z L t zo
-0.5 T = ----- ( L – 2z ) 0≤z≤L (7.27)
2
-1 The distributions of the angle of twist, eq. (7.26), and
Fig. 7.14 Distributions of the twist angle torque, eq. (7.27), are plotted in Fig. 7.14.
and torque for the bar with fixed ends
and subjected to a uniform distributed
torque.

EXAMPLE 7.3 A point torque acting on a bar with fixed ends.

Consider a uniform bar of length L subjected to an external point torque,


Q denoted by Q, at z = a. Each end of the bar is fixed to a rigid support so that
θ z ( 0 ) = 0 and θ z ( L ) = 0 . Determine the angle of twist θz and torque T in
z the bar for 0 ≤ z ≤ L. See Fig. 7.15.
a b
L Solution In this case the distributed torque intensity tz in the governing dif-
ferential equation, eq. (7.25), is zero in the open intervals 0 < z < a and a < z
Fig. 7.15 Bar with fixed < L. The governing differential equation is not valid at the point of applica-
ends under a point torque 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
2

GJ 2 z = 0 for 0 < z < a and a < z < L
dz
The general solution to this equation is

 c1 z + c2 0≤z<a
θz = 
 c3 z + c4 a<z≤L

156 Thin-Walled Structures


Unit twist of a single cell section

where c1, c2, c3, and c4 are constants to be determined from the boundary conditions. The fixed end condition at
z = 0 yields c2 = 0, and the fixed end condition at z = L yields c4 = - c3 L. The constants c1 and c3 are determined
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., θ z ( a —) = θ z ( a + ) . Imposing this condition of continuity we have

c1 a = –c3 ( L – a ) = –c3 b

since b = L - a. Second, as shown in the accompanying free body diagram, torsional


equilibrium at point z = a yields Q
T ( a —) T( a+ )
T ( a + ) – T ( a —) + Q = 0
Using the material law, eq. (7.24), this equilibrium equation yields the second condi- a
tion for constants c1 and c3, which is

Q
GJ ( c 3 ) – GJ ( c 1 ) + Q = 0 or c 1 – c 3 = -------
GJ
Now we have two linear, independent equations to deter-
θz
Qb Qa
mine c1 and c3. We find c 1 = ------- --- and c 3 = – ------- --- . Qab
GJ L GJ L -----------
GJL

Thus, the solution for the angle of twist is


z
Qb
 ------ 0
- --- z 0≤z≤a a
 GJ L L
θz =  (7.28)
Qa T
 ------
- --- ( L – z ) a≤z≤L
 GJ L
The torque is determined from this result and the mate- b
Q ---
L
rial law, eq. (7.24). We find z
0 a L
 Q --b- 0≤z<a
 L
T =  (7.29)
 – Q --a- a
 L a<z≤L – Q ---
L
Plots of the twist angle and the torque are shown in Fig.
7.16. Fig. 7.16 Distributions of the angle of
twist and torque in Example 7.3.

7.4 Unit twist of a single cell section


Consider a single cell bar with a contour of arbitrary shape and a wall thickness t ( s ) , where s denotes the con-
tour coordinate, as is shown in Fig. 7.17. The bar is made of an isotropic and homogeneous material that follows
Hooke’s Law. Our purpose is to determine the twist per unit length dθ z ⁄ dz in terms of the shear flow distribu-
tion around the contour. In the derivation presented in this section it must be emphasized that the shear flow can
be cause by the transverse shear forces V x and V y , and the torque T. The relationship of the shear flows to the
shear forces is given by eq. (6.13) on page 120, and for the special case of a circular contour the shear flow was
related to the torque via Bredt’s formula, eq. (7.13). In Section 7.5 we will determine how to compute the shear

Thin-Walled Structures 157


Article 7.4

z x
Vy
q(s)
t(s) s
Vx
T, θ z

Fig. 7.17 Thin-walled, single cell cross section with an arbitrary shaped contour

flow due to torsion for a closed contour of arbitrary shape. All we need to keep in mind in this section is that the
shear flow is the superposition of the shear flows due to transverse shear and torsion. To relate the shear flow to
the unit twist of the cross section, we use the material law and the geometry of the deformation.

7.4.1 Geometry of the contour


The contour C is closed curve in the x-y plane as shown in Fig. 7.18. The origin of the x- and y-coordinate axis is
y
C
t(s)
ˆj
x(s)
O ˆi ˆt
x
ds
R(s) dy
θ
s = S Α
s dx
r ˆt nˆ
s = 0
θ(s)
y(s) A

Fig. 7.18 Geometry of the contour C.

at point O. Take the origin of the contour coordinate (s = 0) at some point along the contour with s increasing
counterclockwise. 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 as is shown in Fig. 7.18. A generic point on the contour is labeled A and has cartesian
coordinates [ x ( s ), y ( s ) ] . The directions tangent and normal to the contour at A are denoted by unit vectors ˆt and
nˆ , respectively, as is shown in Fig. 7.18, and the angle between the positive x-direction and the tangent to the
contour is denoted by θ ( s ) . The differential arc length ds is given by the Pythagorean theorem as

158 Thin-Walled Structures


Unit twist of a single cell section

ds 2 = dx 2 + dy 2
The position vector of point A relative to O is

R ( s ) = x ( s )iˆ + y ( s )jˆ

and the unit tangent vector is defined as ˆt ≡ dR ⁄ ds , so we have

ˆt = dx dy
------ ˆi + ------ ˆj
ds ds
The unit tangent vector can written in terms of the angle θ ( s ) as
ˆt = ( cos θ )iˆ + ( sin θ )jˆ

Hence, we obtain the trigonometric relations

dx dy
cos θ = sin θ = (7.30)
ds ds

The unit normal vector can be obtained from the cross product nˆ = ˆt × kˆ , which gives
dy dx
nˆ = ------ ˆi – ------ ˆj
ds ds

The normal coordinate of point A relative to O is defined as r ( s ) , and it is given by r ( s ) = R • nˆ . Hence,

dy dx
r(s) = x –y (7.31)
ds ds
Given the contour coordinate functions [ x ( s ), y ( s ) ] , where s denotes the arc length of the contour, the trigono-
metric functions of the angle θ ( s ) are determined from eq. (7.30) and the normal coordinate of the contour is
determined from eq. (7.31).

A quantity that will appear in subsequent formulas is the area enclosed by the contour C, which is denoted
by A c . As is shown in Fig. 7.19, the area of the triangle with height r ( s ) and differential base of length ds is
( rds ) ⁄ 2 . For each differential arc length around the contour similar triangles can be formed. Summing the dif-
ferential areas of all these triangles we have in the limit as ds → 0 for each triangle, with the number of triangles
approaching infinity, the enclosed area given by
1
A c = --- rds
2 °∫
C
(7.32)

For example, if the contour is a circle of radius r and ds = rdθ , then eq. (7.32) becomes

1 2
A c = --- r ( r ) dθ = πr
2 ∫
0

Thin-Walled Structures 159


Article 7.4

C C

1
dA c = --- rds
2 Ac
O 1
A c = --- rds
2 °∫
C
r
θ
ds
A

Fig. 7.19 Area enclosed by the contour

7.4.2 Hooke’s law


The shear stress is directly related to the shear strain with the constant of proportionality being the shear modulus
G of the material.
τ zs = Gγ zs (7.33)

Multiply this equation by the wall thickness t to get


q = Gtγ zs (7.34)

The deformation due to the shear flow is depicted in Fig. 7.20.


y

s, v t
C C*
z q
A x
t(s) B A*
z, w
q
B*
π
--- – γ zs
2

Fig. 7.20 Element of the wall deformed in shear.

7.4.3 Shear strain-displacement relation


The shear strain γzs 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. 7.20. The shear strain is related to the partial derivatives of
the displacement components of a material point on the contour. Let the vt(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

160 Thin-Walled Structures


Unit twist of a single cell section

to position A*, B*, and C* in the deformed shell as shown in Fig. 7.20 and Fig. 7.21. The displacements of these
adjacent material points are shown in Fig. 7.21. The reduction in the right angle between lines elements origi-

∂w w(s+ds,z)
ds
C* ∂s C*
B* vt(s+ds,z)
B*  1 + ∂v t ds C
 ∂s 
∂v t A*
dz vt(s,z+dz) ds
∂z A* vt(s,z)
 1 + ∂w dz
B dz
 ∂z  A:(s,z)
w(s,z)
w(s,z+dz)

Fig. 7.21 Tangential and axial displacements of three adjacent points in the wall of the shell.

nally parallel to the s- and z-directions is determined from Fig. 7.21 as

∂v t ∂w
dz ds
∂z ∂s
γ zs = --------------------------- + --------------------------- (7.35)
 1 + ∂w dz  1 + ∂v t ds
 ∂z   ∂s 
Since the displacement derivatives are small in magnitude with respect to unity for infinitesimal deformations,
we get

∂v t ∂w
γ zs = + (7.36)
∂z ∂s
Combining eqs. (7.34) and (7.36) to eliminate the shear strain gives

q ∂w ∂v t
------ = + (7.37)
Gt ∂s ∂z

7.4.4 Tangential displacement of a typical point on the contour


The projection of the cross section in the deformed cylindrical shell onto the x-y plane is assumed not to distort
from its original cross-sectional shape in the undeformed shell. However, the projected cross section can displace
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 circular con-
tour (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

Thin-Walled Structures 161


Article 7.4

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 vt of point A consists of a rigid
body translation vt1, in which point O moves to O* and A moves to A1, followed by a rotation θ z about O*. The
tangential displacement of A1 to A* is denoted by vt2. Thus, v t = v t1 + v t2 .

x θz + θ
O* A*

θ θz
v0 θ
r β
uo A1 v
t2
O
r –vn
v t1
β θ
vt
A

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

The displacement of point O has a horizontal component denoted by u0(z) and a vertical component denoted
by v0(z). Since the angle between the x-direction and the tangent to the contour at s is denoted by the function
θ ( s ) , the projection of the horizontal component in the tangent direction is u 0 cos θ , and the projection of the
vertical component in the tangent direction is v 0 sin θ . Also, the displacement of A to A1 is the same as the dis-
placement of O to O*. Thus, the tangential displacement of A to A1 due to translation of the section is the sum of
these projections, or
v t1 = u 0 cos θ + v 0 sin θ (7.38)

The tangential displacement due to the rotation θ z is determined by the projection of line O*A* onto the tangent
direction. First note that the length of lines OA , O * A 1 , and O * A * are the same because of the rigid section
assumption. The angle between the line OA and the normal direction to the contour at point A is denoted by β.
The tangential component of the displacement of A1 to A* is obtained from the geometry shown in Fig. 7.22 as

v t2 = O * A * sin(β + θ z ) – O * A 1 sin β = O * A* [ sin β cos θ z + cos β sin θ z – sin β ]

162 Thin-Walled Structures


Unit twist of a single cell section

We assume that the rotation θ z is small such that sin θ z ≈ θ z and cos θ z ≈ 1 . Also, note that OA cos β = r ,
which is the coordinate of A relative to O in the direction normal to the contour at A. Thus,

v t2 = ( O * A * cosβ )θ z = ( OA cos β )θ z = rθ z

The total tangential displacement of A to A* is then


v t ( s, z ) = u 0 ( z ) cos [ θ ( s ) ] + v 0 ( z ) sin [ θ ( s ) ] + r ( s )θ z ( z )




















translation rotation (7.39)

7.4.5 Relation between the shear flow and unit twist


We can take the partial derivative of the tangential displacement, eq. (7.39), with respect to z to get

∂v t du 0 dv dθ
= cos θ + 0 sin θ + r z (7.40)
∂z dz dz dz
Substitute this result for the derivative of the tangential displacement in eq. (7.37) to get

q ∂w du 0 dv dθ
------ = + cos θ + 0 sin θ + r z
Gt ∂s dz dz dz
Now integrate this last expression completely around the contour to get

q ∂w du dv dθ
------ ds = ds + 0 cos θ ds + 0 sin θ ds + ( rds ) z
°∫ Gt °∫∂s dz °∫dz °∫
dz °∫ (7.41)

From eqs. (7.30) we have

°∫ cos θ ds = °∫ dx = x ( S ) – x ( 0 ) = 0
and

°∫ sin θ ds = °∫ dy = y ( S ) – y ( 0 ) = 0
so that the second and third terms on the right-hand side of eq. (7.41) vanish. The first integral on the right-hand
side of eq. (7.41) can be done exactly to give

∂w
°∫ ∂ s ds = [ w ( S, z ) ) – w ( 0, z ) ] = 0 ,
since the axial displacement of the material point at s = 0 and s = S is unique. From eq. (7.32) the fourth integral
on the right-hand side of eq. (7.41) is identified as twice the enclosed area of the contour. Hence, the final result
from eq. (7.41) is

dθ z 1 q
= ------------ ------ ds
dz 2A c Gt °∫ (7.42)

Equation (7.42) is the formula we set out to derive. It relates the shear flow to the twist per unit length of the bar.
The shear flow can be caused by both transverse shear forces and torque. In this sense, eq. (7.42) is more general
than for torsion only. Equation (7.42) is also derived by energy methods in Article 8.6.3 on page 210.

Thin-Walled Structures 163


Article 7.5

7.5 Closed contour with 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

y
T

z x
t(s)

T, θ z

Fig. 7.23 Uniform torsion of a thin-walled cylindrical shell with an arbitrary shaped contour.

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 dθ z ⁄ dz is constant, since the
torque, the cross section, and material do not change with respect to z. Our purpose is to relate the shear flow q to
the torque T.

Axial equilibrium requires uniform shear flow around the contour We assume the shear stress τ zs is uni-
formly distributed across the thickness of the wall, based on the analysis of a thin-walled circular tube. The shear
flow is given by q = τ zs t ( s ) , and the shear flow acts tangent to the contour. Since there are no axial normal
stresses (no bending nor extension), the shear flow is uniform along the contour. That is,
q = constant with respect to s (7.43)

A free body element from the wall is shown in Fig. 7.24, and axial equilibrium is used to establish the shear flow
is uniform along the contour.

q(s + ds) dz
s
z
s + ds
s q(s) dz
q ( s + ds )dz – q ( s )dz = 0
( ∑ Fz = 0 ) →

Fig. 7.24 A free body diagram of an element of the shell for axial equilibrium.

Resultant of the uniform shear flow In general, the resultant of the shear flow resolved at an arbitrary point in
the cross section (again, labeled O) is a force and a couple. Static equivalence gives

164 Thin-Walled Structures


Closed contour with arbitrary shape

Fx =
°∫ ( cos θ )qds Fy =
°∫ ( sin θ )qds T =
°∫ r ( qds ) (7.44)

in which r(s) is the coordinate in the normal direction to the contour at A measured from point O. See Fig. 7.25.

Fy
1
dA c = --- rds T
2
O O Fx

r qds
θ
ds ds
dy
A θ
dx

Fig. 7.25 Resultant of the shear flow

Since the shear flow is independent of the contour coordinate s, it may be brought outside the integral in eqs.
(7.44). We get, using eqs. (7.30) and (7.32) in the process, that

F x = q cos θds = q dx = q [ x ( S ) – x ( 0 ) ] = 0
°∫ °∫ (7.45)

FY = q sin θds = q dy = q [ y ( S ) – y ( 0 ) ] = 0
°∫ °∫ (7.46)

T = q rds = q 2dA c = 2A c q
°∫ °∫ (7.47)

Equation (7.47) is called Bredt’s formula, or the Bredt-Batho formula, and it relates the torque to the shear flow
via the area enclosed by the contour. Since Bredt’s formula is the principal result, we repeat it below as eq.
(7.48).
T
q = --------- Bredt’s formula (7.48)
2A c

For the shear flow independent of the contour coordinate, we find that the unit twist from eq. (7.42) reduces
to

dθ z q ds
= --------- ------
dz 2A c Gt °∫ (7.49)

Equations (7.48) and (7.49) can be combined by eliminating the shear flow, and we write the result as

dθ z
T = G0 J (7.50)
dz
in which the torsion constant is given by

Thin-Walled Structures 165


Article 7.5

4A 2
J = --------c- single cell section (7.51)
ds
°∫
-----
t*
and the modulus-weighted thickness is given by
G(s)
t * = -----------t (7.52)
G0
Shear modulus G0 is introduced in the definition 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-
posed of a single homogeneous material, then we take G0 = 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. (7.51), reduces to
value of J for the circular tube given in Section 7.1, eq.
(7.11).
t 2
A c = πr A c = ab
r a ds 2(a + b)
ds 2πr t
°∫ ----t- = --------t - °∫ ----t- = -------------------
t
-

4 ( πr )
2 2 4 ( ab ) 2 2 ( ab ) 2 t
3 J = -------------------- = -------------------
J = ------------------- = 2π ( r ) t b 2(a + b) a+b
2πr --------------------
--------- t
t
Fig. 7.26 Torsion constants for thin-walled, single cell sections with circular and rectangular
contours

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 flow, maximum
shear stress, the torsion constant, and the torsional stiffness.
6 6
G = 2.5 ×10 psi G = 5 ×10 psi

10000 lb-in. 2.0 in.

0.250 in. 0.125 in.

Fig. 7.27 Bi-material circular torsion tube

166 Thin-Walled Structures


Shear center of a closed section

Solution The shear flow is computed from Bredt’s formula, eq. (7.48), as
T 10000 lb-in.
q = --------- = ----------------------------
- = 1591.5 lb/in.
2A c 2π ( 1 in. ) 2

The shear stress in the left side wall is τ zs = ( 1591.5 lb-in. ) ⁄ ( 0.250 in. ) = 6366.2 psi , and in the right side
wall the shear stress is τ zs = ( 1591.5 lb-in. ) ⁄ ( 0.125 in. ) = 12732.4 psi . Hence, the maximum shear stress is
12.7 ksi, and it occurs in the 0.125-inch-thick section.

6
Take the reference shear modulus G 0 = 2.5 ×10 psi. First we compute the denominator of the torsion con-
stant given by eq. (7.51).
ds π ( 1 in. ) π ( 1 in. )
- + -------------------------------------------------- = 8π
°∫ ----t - = ------------------------------------------------------
*
2.5 ×10 psi
6
5 ×10 psi
6
---------------------------- ( 0.250 in. ) ----------------------- ( 0.125 in. )
G0 G0

Note that the modulus-weighted thickness of the right side wall is 2 (0.125 in.) = 0.250 in. From eq. (7.51) the
torsion constant is
4 [ π ( 1 in. ) 2 ] 2 π
J = -------------------------------- = --- in. 4
8π 2

T
the torsional stiffness is defined as ------------------ , which equals G 0 J via eq. (7.50). Thus, the torsional stiffness for this
dθ z ⁄ dz
example is

T π
------------------ = G 0 J = ( 2.5 ×10 psi )  --- in. 4 = 3.93 ×10 lb-in. 2
6 6
dθ z ⁄ dz 2 

7.6 Shear center of a closed section


The unit twist formula given by eq. (7.42) is applicable even if the shear flow is a function of the contour
coordinate s, as is the case when the transverse shear forces are non-zero. Hence, eq. (7.42) can be used to com-
pute the unit twist of the section when the shear forces V x and V y , and the torque T act in unison. In the case of
torsion only, we showed in the last section that the shear flow was constant along the contour coordinate based on
axial equilibrium, which led to the simplification given by eq. (7.49). If V x and V y are non-zero, then the compu-
tation for the unit twist is a more involved. We illustrate the use of eq. (7.42) 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 flow distribution around the section due to transverse bending using eq. (6.13) on page
120.
This step involves selecting a contour origin where s = 0. The shear flow q(0) at the contour origin is
unknown.
• Determine the unknown shear flow q(0) by setting the unit twist, eq. (7.42), to zero.
Setting the unit twist to zero enforces the condition of no torsional deformation for the section.

Thin-Walled Structures 167


Article 7.6

• 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 I xx = 300t cm 4 , where t is
specified in cm. Since the shear center lies on this axis of symmetry, we take the shear force components V x = 0
and V y > 0 . Determine the location ex of the shear center from point O, which is located on the x-axis at the ver-

y
Vy s 1, q 1
5 13
C O 5
S.C. x s 2, q 2 12
O
5 q0
t s 3, q 3
1
4 ---
3
12
All dimensions in cm
ex

Fig. 7.28 A closed section with an isosceles triangle contour

tex of the two sides of the triangle that are of equal length.

Solution The shear flows are assumed positive in the counter-clockwise direction with the contour origin at
V
point O. The shear flow at point O is q0, and eq. (6.13) on page 120 reduces to q ( s ) = q 0 – -----y- Q x ( s ) . For the
I xx
first branch shown in Fig. 7.28 we have that the shear flow due to transverse bending is given by

Vy s1 Vy 2
5 
 -----
q 1 ( s 1 ) = q 0 – ----------
- - s t ds = q 0 – -----------
-(s ) 0 ≤ s 1 ≤ 13 cm
300t ∫ 0
 13 1 1 1560 1
(7.53)

The shear flow due to bending in the second branch is


Vy s2
q 2 ( s 2 ) = q 1 ( 13 ) – ----------
- ( 5 – s 2 )t ds 2 0 ≤ s 2 ≤ 10 cm
300t ∫ 0

From eq. (7.53) q 1 ( 13 ) = q 0 – ( 13V y ) ⁄ 120 . Hence the shear flow in the second branch is

Vy  1
- 13 + 2s 2 – --- s 22
q 2 ( s 2 ) = q 0 – -------- 0 ≤ s 2 ≤ 10 cm (7.54)
120  5 

168 Thin-Walled Structures


Shear center of a closed section

The shear flow due to bending in the third branch is


Vy s3
5 
 – 5 + -----
q 3 ( s 3 ) = q 2 ( 10 ) – ----------
- - s t ds 0 ≤ s 3 ≤ 13 cm
300t ∫ 0
 13 3 3

From eq. (7.54) q 2 ( 10 ) = q 0 – ( 13V y ) ⁄ 120 , which is the same value as the shear flow at the junction of
branches one and two. Hence the shear flow in branch three is
Vy  1
- 13 – 2s 3 + ------ s 32
q 3 ( s 3 ) = q 0 – -------- 0 ≤ s 3 ≤ 13 cm (7.55)
120  13 

Note from this result that q 3 ( 13 ) = q 0 , as it should since the shear flow at point O is unique.

To find the shear flow at point O, we impose the condition of zero unit twist from eq. (7.42). Since the prod-
uct of Gt is the same in each branch, eq. (7.42) reduces to
dθ 1
--------z = --------------------
dz 2( Gt ) A c °∫ qds= 0
which requires the contour integral of the shear flow around the entire section to vanish. Note that the unit twist is
positive counter-clockwise and consequently a counter-clockwise shear flow is positive in eq. (7.42). Substituting
eqs. (7.53) to (7.55) into this integral, we get
13 Vy 2 
 q – -----------
10 Vy 
 q – -------- 1
qds = - ( s ) ds + - 13 + 2s 2 – --- s 22  ds 2 +
°∫ ∫ 0
 0 1560 1  1 ∫ 0
 0 120  5 

13 Vy 
 q – -------- 1
- 13 – 2s 3 + ------ s 32  ds 3 = 0
∫ 0
 0 120  13  

After evaluation of the definite integrals, this equation becomes


23 23
36q 0 – ------ V y = 0 Hence q 0 = --------- V y (7.56)
10 360
Substitute q0 from eq. (7.56) into eqs. (7.53) to (7.55) to find that the shear flows due to bending only are

Vy
- ( 299 – 3s 12 )
q 1 ( s 1 ) = ----------- (7.57)
4680
Vy
- ( – 80 – 30s 2 + 3s 22 )
q 2 ( s 2 ) = ----------- (7.58)
1800
Vy
- ( – 208 + 78s 3 – 3s 32 )
q 3 ( s 3 ) = ----------- (7.59)
4680

For each straight branch we can compute the branch force. These branch forces are

13 13 Vy 13
f1 = q 1 ds 1 = - ( 299 – 3s 12 ) ds 1 = ------ V y
-----------
∫ 0
∫ 0 4680 36
(7.60)

Thin-Walled Structures 169


Article 7.7

10 10 Vy 13
f2 = q 2 ds 2 = - ( – 80 – 30s 2 + 3s 22 ) ds 2 = – ------ V y
-----------
∫ 0
∫ 0 1800 18
(7.61)

13 Vy 13
f3 = - ( – 208 + 78s 3 – 3s 32 ) ds 3 = ------ V y
-----------
∫ 0 4680 36
(7.62)

These branch forces are shown in Fig. 7.29. Now torque equivalence is used to determine the location of the

13 Vy
------ V y
36

13
------ V y
18
O ⇔ O

13
------ V y ex
36

Fig. 7.29 Static equivalence to find the shear center of the triangular contour

shear center on the axis of symmetry. Torque equivalence gives

13 2
12 cm  ------ V y = e x V y Hence e x = 8 --- cm
 18  3
The shear center and the centroid are shown in Fig. 7.30.

S.C. C
O
Fig. 7.30 Shear center and centroid
2
location of the triangular section 7 --- cm
3
2
8 --- cm
3

12cm

7.7 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 t1
and t2, respectively. The thickness of the common branch separating the two cells is denoted by t 1 – 2 . All

170 Thin-Walled Structures


Multi-cell closed sections

y
A
T z T

A
L

t1
y
t2

x A
Section A-A A c1 c2

t1 – 2

Fig. 7.31 Torsion of a shell-beam with a cross section consisting of two closed cells

branches are assumed made of the same homogeneous material with the shear modulus denoted by G. Our pur-
pose is to determined the shear flows and shear stresses due to torsion, and to determine the torsion constant J in

the standard formula T = GJ --------z where dθ z ⁄ dz is the twist per unit length.
dz

We assume positive shear flow directions as shown in Fig. 7.32. The shear flows in the exterior branches of

y
q1
t1 q2 t2

q1 – 2
x A c1 A c2
q1 q2
O t1 – 2

Fig. 7.32 Assumed positive shear flows for the two-cell section

cells 1 and 2 are taken positive counter-clockwise and are denoted by q1 and q2, respectively. The shear flow in
the common branch is denoted by q1-2. The common branch shear flow is related to the shear flows 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.
Axial equilibrium at the junction point O gives
q1 – 2 = q1 – q2 (7.63)

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. (7.63). An easy way to determine the axial equilibrium of the junction is to recall that the shear

Thin-Walled Structures 171


Article 7.7

q1 – 2

z
q1
q2 ∆z

q1 O q2

Fig. 7.33 Free body diagram of the junction at point O

flow into the junction must equal the shear flow out of the junction.

Now we determine the torque about point O due to the shear flows. The shear flow q1-2 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
Bredt’s formula, the torque about point O is
T = 2 A c1 q 1 + 2 A c2 q 2 (7.64)

where A c 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 flow distribution is resolved. The fact that the
torque due to the shear flows is the same about any point in the plane is discussed in Section 7.8.

At this point we have two equations, eqs. (7.63) and (7.64), for the three unknown shear flows assuming that
the torque and geometry of the section are known. An additional equation relating the shear flows 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. (7.42). Since the shear modulus
is the same for all branches in the cross section eq. (7.42) reduces to
dθ z 1 q
-------- = -------------
dz 2 A cG °∫ --t- ds
This unit twist formula was derived on the basis that a counter-clockwise shear flow tends to produce a counter-
clockwise unit twist. Apply this equation to cell 1 to get

 dθ 1 S S
--------z = ----------------  --- q 1 +  --- q (7.65)
 dz  1 2A c1G  t  1  t 1 – 2 1 – 2

where S1 is the arc length of the exterior branch of cell 1, and S1-2 is the length of the common branch between
cells 1 and 2. For cell 2, the unit twist is

dθ z
 ------- 1 S S
- = ----------------  --- q 2 –  --- q (7.66)
 dz  2 2A c2G  t  2  t 1 – 2 1 – 2

where S2 is the arc length of the two exterior branches of cell 2. Note that the contribution of the common branch
shear flow is negative in the unit twist formula for cell 2. Relative to an observer in cell 2, a positive value for the
shear flow q1-2 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. (7.65) and eq. (7.66) to get

172 Thin-Walled Structures


Multi-cell closed sections

1 S S 1 S S
----------------  --- q 1 +  --- q 1 – 2 = ----------------  --- q 2 –  --- q (7.67)
2A c1G  
t 1  
t 1–2 2A c2G  
t 2  t 1 – 2 1 – 2

We now have three equations for the three unknown shear flows; eqs. (7.63), (7.64), and (7.67). These equa-
tions are solved for the shear flows in terms of the torque and geometry of the cross section. Once the shear flows
are known, the unit twist for either cell 1, eq. (7.65), or cell 2, eq. (7.66) 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-
dθ T
dard formula --------z = ------- to identify the value of the torsion constant J. For a multi-cell section, the formula
dz GJ
given for the torsion constant of a single cell section, eq. (7.51), is not applicable.

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 πr. 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 flows q1 and q2 in the exterior branches of cells 1 and 2, and
• the torsion constant J.

πr y
q2
q1 q1 – 2
3t
T
r x A c1 A c2

t t q2
∼ 3r

Fig. 7.34 Uniform torsion of a two-cell section

Solution From eq. (7.64) the torque is

πr 2 2r ⋅ 3r
2  -------- q 1 + 2  --------------- q 2 = T
 2   2 

or

πr 2 q 1 + 6r 2 q 2 = T (7.68)

Using eq. (7.65) for the unit twist for cell 1, and eq. (7.63) for the common web shear flow, we get

Thin-Walled Structures 173


Article 7.7

dθ z
 ------- 1 πr 2r
- = ----------------------  ----- q 1 +  ----- ( q 1 – q 2 )
 dz  1 πr 2  t   3t 
2  -------- G
 2 

or

 dθ 1 2
--------z = -----------------  π + 2--- q 1 – --- q 2 (7.69)
 dz  1 ( πrt )G  3 3
Using eq. (7.56) for the unit twist for cell 2, and eq. (7.63) for the common web shear flow, we get

dθ z
 ------- 1 2πr 2r
-  --------- q 2 –  ----- ( q 1 – q 2 )
- = --------------------
 dz  2 2 ( 3r 2 )G  t   3t 

or

dθ z 1 2
- = ----------------- – --- q 1 +  2π + 2--- q 2
 ------- (7.70)
 dz  2 ( 6rt )G 3  3
Now equate eqs. (7.69) and (7.70) to get

1 2 1 2
---  π + 2--- q 1 – --- q 2 = --- – --- q 1 +  2π + 2--- q 2
π  3 3 6 3  3

which after rearrangement gives

2 1
 1 + ----- π 1 2
- + --- q –  --- + --- + ------ q = 0 or 1.3233177q 1 – 1.3705152q 2 = 0 (7.71)
 3π 9 1  3 9 3π 2

Equations (7.68) and (7.71) are two simultaneous equations for the shear flows q1 and q2 in terms of the torque.
The solution to these simultaneous equations determines the shear flows. Hence,
T T
answer : q 1 = --------------------2- q 2 = --------------------2- (7.72)
8.9350r 9.2536r

To find the torsion constant J, we substitute the solutions for the shear flows into eq. (7.69). The result is

dθ dθ T
--------z =  --------z = 0.11273713 ----------
-
dz  dz  1 Gr 3 t
The same result is obtained using the unit twist for cell 2; i.e., eq. (7.70). Comparing the unit twist result given
dθ T
above to the standard torsional relation ------ = ------- , we find that the torsion constant J is
dz GJ

1 0.11273713
--- = ---------------------------
- giving the answer J = 8.870r 3 t (7.73)
J r3t

174 Thin-Walled Structures


Resultant of uniform shear flow

7.8 Resultant of uniform shear flow


In torsion problems it often necessary to find the resultant of a constant shear flow along the contour of a curved
branch. This situation is depicted in Fig. 7.35. The resultant of the shear flow is resolved at point A in this fig-

y y

yB B
qds
dA c
ds dy
yA q A dx
A r
x x
xA xB

(a) Curved contour (b) Differential element on contour

Fig. 7.35 Constant shear flow along a curved branch.

ure. From the differential element shown in Fig. 7.35(b), the differential force and differential torque resolved at
point A are

dx dy
dF x = ( qds )  ------ = qdx dF y = ( qds )  ------ = qdy dT = r ( qds )
 ds  ds

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
B B
Fx = q dx = q ( x B – x A ) Fy = q dy = q ( y B – y A ) T = 2A c q
∫ A
∫ A
(7.74)

where A c 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

F = F x2 + F y2 = qL where L = ( xB – xA ) 2 + ( yB – yA ) 2

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 F y ⁄ F x = ( y B – y A ) ⁄ ( x B – x A ) . The force and torque resolved at point A
are shown in Fig. 7.36(a). The final 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 flow in a curved
branch is a force of magnitude equal to the shear flow 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.

Thin-Walled Structures 175


Article 7.8

y y

yB B B

qL
2A c q Ac qL

L A
yA
A
2A c ⁄ L
x x
xA xB
(a) Force and torque at A (b) Resultant

Fig. 7.36 Steps to resolving the constant shear to a single force along a specific line of action.

From the above discussion, a constant shear flow 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. (7.74). However, a constant
shear flow 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 Bredt’s formula can add as a negative quantity if the
torque produced by the constant shear flow is clockwise over a segment of the branch. About point O in this fig-
ure, the torque produced by the shear flow from point B to A in the right half of the contour is counter-clockwise,
and the torque produced by the shear flow 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.

A ca – A cb = A c
q q
A A Ac
A ca A cb
q

B B
2A c q
O 2A ca q O O
2A cb q

Fig. 7.37 The torque about an arbitrary point O of a constant shear flow in a closed contour is twice the
enclosed area of the circuit times the shear flow.

Finally consider torsion of a multi-cell section. Each branch composing the cross section has a constant
shear flow, but the shear flow from branch-to-branch is, in general, a different value. Axial equilibrium requires
that the shear flows into a junction must equal the shear flow out of the junction. If the shear flows satisfy this
condition at all junctions, then the shear flow 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 flow is

176 Thin-Walled Structures


Resultant of uniform shear flow

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 q 1 S 1 – 2 ,

( q 1 – q 2 )S 1 – 2 F
q1 – q2 T
q2 S1 – 2
A c1 A c2 q1 S1 – 2

⇔ S1 – 2 ⇔
q1
q2

2 A c1 2 A c2
------------ ------------
S1 – 2 S1 – 2

Fig. 7.38 Branch-by-branch statical equivalence of a two-cell section

where S 1 – 2 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 figure, 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 q 2 S 1 – 2 with a line of action parallel to the common branch. The common branch resultant is an upward
force of magnitude ( q 1 – q 2 )S 1 – 2 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
F = – q 1 S 1 – 2 + ( q 1 – q 2 )S 1 – 2 + q 2 S 1 – 2 = 0

2 A c1 2 A c2
T = ----------
- ( q 1 S 1 – 2 ) + ----------
-(q S ) = 2 A c1 q 1 + 2 A c2 q 2
S1 – 2 S1 – 2 2 1 – 2
These resultants confirm eq. (7.64).

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

Consider the five 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 flows q1, q2, q3, q4, and q5, and
• the torsion constant J

Solution It is convenient to define circuit shear flows in each cell. Circuit shear flows are assumed to be posi-
tive counter-clockwise in each cell and are equal to the actual shear flows in the exterior branches of the cell, if
there are any exterior branches. However, the shear flow in a common branch between cells is the difference in
the circuit shear flows sharing the common branch. This procedure automatically satisfies shear flow continuity
at the junctions;. i.e., axial equilibrium at the junctions. See the sketch of the center junction in Fig. 7.39.

Thin-Walled Structures 177


Article 7.8

a a a

q5 q4 q3 a
q4 – q1 q2 – q4
T

q2 q1 – q2
q1 a
Balance of shear flows
at the center junction

3a/2 3a/2

Fig. 7.39 Uniform torsion of a five cell section

Using Bredt’s formula to compute the torque for each cell, then summing these individual torques, assuming
counter-clockwise as positive, results in

3 3
2  --- a 2 q 1 + 2  --- a 2 q 1 + 2 ( a 2 )q 3 + 2 ( a 2 )q 4 + 2 ( a 2 )q 5 = T (7.75)
2  2 

The twist per unit length for each cell, with positive shear flows taken counter-clockwise within a cell, is

 dθ 1 5 a
--------z = ------------------------  --- a q 1 + a ( q 1 – q 2 ) +  --- ( q 1 – q 4 ) + a ( q 1 – q 5 ) (7.76)
 dz  1 3 2   2
2  --- a 2 Gt
2 

 dθ 1 5 a
--------z = ------------------------  --- a q 2 + a ( q 2 – q 3 ) +  --- ( q 2 – q 4 ) + a ( q 2 – q 1 ) (7.77)
 dz  2 3  2   2
2  --- a 2 Gt
2 

dθ z
 ------- 1
- [ ( 2a )q 3 + a ( q 3 – q 4 ) + a ( q 3 – q 2 ) ]
- = ------------------- (7.78)
 dz  3 2 ( a 2 )Gt

dθ z
 ------- 1 a a
- aq 4 + a ( q 4 – q 5 ) +  --- ( q 4 – q 1 ) +  --- ( q 4 – q 2 )
- = ------------------- (7.79)
 dz  4 2 ( a )Gt2  2   2

dθ z
 ------- 1
- [ ( 2a )q 5 + a ( q 5 – q 1 ) + a ( q 5 – q 4 ) ]
- = ------------------- (7.80)
 dz  5 2 ( a 2 )Gt
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

dθ z
 ------- dθ  dθ dθ  dθ dθ  dθ dθ
- =  --------z --------z =  --------z --------z =  --------z --------z =  --------z (7.81)
 dz  1  dz  2  dz  2  dz  3  dz  3  dz  4  dz  4  dz  5
Substitute eqs. (7.76) to (7.80) into eqs. (7.81) to get four equations for the five shear flows. Then add eq. (7.75)
to get the fifth equation. These five equations in matrix form are

178 Thin-Walled Structures


Resultant of uniform shear flow

1 1
2 –2 --- 0 – ---
( dθ z ⁄ dz ) 1 = ( dθ z ⁄ dz ) 2 : 3 3
q1 0
1 13 7 1
( dθ z ⁄ dz ) 2 = ( dθ ⁄ dz ) 3 : – --- ------ – --- --- 0 q2
3 6 3 3 0
( dθ z ⁄ dz ) 3 = ( dθ z ⁄ dz ) 4 : 1 1 1 q3 = 0 (7.82)
--- – 1--- – 2 ------
( dθ z ⁄ dz ) 4 = ( dθ z ⁄ dz ) 5 : 4 4 2 2 q 0
4
1
eq. (7.75): --- – 1--- 0 5
2 – --- q 5 T
4 4 2
3a 3a 2 2a 2
2 2a 2a 2
2

Solving eq. (7.82) for the shear flows, we get


23 T 22 T 55 T 22 T
q 1 = q 2 = --------- ----2- q 3 = --------- ----2- q 4 = --------- ----2- q 5 = --------- ----2-
281 a 281 a 562 a 281 a
The unit twist of the section can be determined by substituting the above shear flows into any one of the eqs.
(7.76) to (7.80). Using eq. (7.76), the unit twist is
dθ z 75 T
-------- = ------------ -----------
-
dz 1124 a 3 tG

dθ T
Compare this to the standard formula --------z = ------- to find
dz GJ

1 75 1 1124
--- = ------------ ------
- Hence J = ------------ a 3 t = 14.9867a 3 t ν
J 1124 a 3 t 75

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. (7.51) the torsion constant is J = 14.40a 3 t . For this example, subdividing the
single cell section into five 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 five 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.

Thin-Walled Structures 179


Article 7.9

7.9 Hybrid sections


Consider a hybrid section composed a single closed cell
and open branches, or fins, as shown in Fig. 7.40 The total
ti bi torque carried by the section is the sum of the torques car-
t 2 ried by the closed cell and open branches. For n open
branches, we have

Ac i =1 n
T T = T closed + ∑T i (7.83)
i=1
n-1
with

i=n dθ dθ
T closed = ( GJ ) closed  --------z T i = ( GJ ) i  --------z (7.84)
 dz   dz 
Fig. 7.40 Torsion of a hybrid section
The torsional stiffness for the closed cell is

4A 2
( GJ ) closed = ---------c- (7.85)
ds
------
Gt °∫
and for each open branch the torsional stiffness is

1
( GJ ) i = G i  --- b i t i3 (7.86)
3 

Combining eqs. (7.83) and (7.84), we have


n
dθ z
T = ( GJ ) closed + ∑ ( GJ ) i --------
dz
i=1


Comparing this result to the standard torsional formula T = ( GJ ) eff  --------z , the effective torsional stiffness for
 dz 
the entire section is
n
( GJ ) eff = ( GJ ) closed + ∑ ( GJ ) i (7.87)
i=1

where the closed and open parts of the torsional stiffness are given by eqs. (7.85) and (7.86).

The shear stress in the closed cell is τ closed = T closed ⁄ ( 2A c t ) , where the portion of the torque carried by the
closed cell is T closed = ( GJ ) closed ( dθ z ⁄ dz ) . The unit twist is given by ( dθ z ⁄ dz ) = T ⁄ ( GJ ) eff . Combining
these results, the shear stress in the closed cell is
( GJ ) closed T
τ closed = ------------------------ ⋅ ----------- (7.88)
( GJ ) eff 2A ct
For the open branches, we use the formula for the shear stress given by the first of eqs. (7.21). Repeating this
equation, using the second of eqs. (7.84) for the torque carried by the branch, and then using eq. (7.86) for the

180 Thin-Walled Structures


Hybrid sections

torsional stiffness of the branch, we can formulate the shear stress as

3T 3 1 dθ 3 1 T
τ i = --------2-i = --------2- ⋅  G i --- b i t i3 --------z = --------2- ⋅  G i --- b i t i3 ⋅ -----------------
bi ti bi ti  3  dz bi ti  3  ( GJ ) eff

The unit twist was eliminated in this equation using the overall section formula for it. After simplification of the
above result we get

Gi ti
-T
τ i = ---------------- (7.89)
( GJ ) eff
If the shear modulus is the same in all branches, then eqs. (7.88) and (7.89) reduce to

J closed T Tt
τ closed = --------------
- ⋅ ----------- τ i = ------i (7.90)
J 2 Ac t J

Thin-Walled Structures 181


Article 7.10

7.10 Determination of the shear stress tangent to the contour – a


summary of chapters 6 & 7
The shear stress tangent to the contour τ zs is the sum of the shear stresses due to
transverse shear force V x , transverse shear force V y , and the torque T

Section Transverse shear forces V x, V y acting


Torque T
type through the shear center
Open τ zs = q ( s ) ⁄ t ( s ) , uniform thru wall thickness dθ z
Single T = ( GJ ) eff
dF dz
Cell q ( s ) = q ( 0 ) – ------- , shear flow from axial equi-
dz
Multi-
Cell librium of a portion of the contour per unit bar
Hybrid length
dF I xx V x – I xy V y I yy V y – I xy V x
- Q y ( s ) + -------------------------------
------- = ------------------------------- - Qx ( s )
dz I xx I yy – I xy 2 I xx I yy – I xy 2

s s
Qx ( s ) = ∫0 y ( s )t ds Qy ( s ) = ∫0 x ( s )t ds
[ x ( s ), y ( s ) ], I xx, I yy, I xy , x-y thru centroid

Open q = 0 at free edges of branches/flanges/webs q = 0 for all s


τ zs linear across wall thickness
( τ zs ) i = ( Tt i ) ⁄ J in i-th branch
max

GJ eff = ∑ G i J i , J i = ( b i t 13 ) ⁄ 3
branches

Single dθ z τ zs = q ( s ) ⁄ t ( s ) , uniform thru wall thickness


1 q
Cell = ------------ ∫ ------ ds = 0 , to find the shear flow
dz 2A c °Gt q = T ⁄ ( 2A c ) , uniform along contour
closed
at the contour origin, q ( 0 ) 4A 2
( GJ ) eff = ---------c-
ds
°∫ -----
Gt
C
-

Multi- q n = the shear flow in exterior webs of cell n,


 dθ z = 0 , n = 1, 2, …, N to find shear flows
Cell dz  n or circuit shear flow of cell n, C.C.W. positive.
Closed
at the contour origin of each cell, q n ( 0 ) . q n – q m = shear flow in common web between

N = num- cells n and m


ber of Shear flows uniform along each branch
cells τ zs = q ⁄ t , where q = q n or q = q n – q m
N
T = ∑ 2Acn qn
n=1

 dθ z =  dθ z , n = 1, 2, …, N – 1
dz  n dz  n + 1

182 Thin-Walled Structures


References

The shear stress tangent to the contour τ zs is the sum of the shear stresses due to
transverse shear force V x , transverse shear force V y , and the torque T

Section Transverse shear forces V x, V y acting


Torque T
type through the shear center
Hybrid q = 0 at free edges of branches/flanges/webs ( GJ ) closed T
with one τ closed = -------------------------
- ⋅ -----------
dθ z ( GJ ) eff 2A ct
closed 1 q
= ------------ ∫ ------ ds = 0 , to find shear flow
dz 2A c °Gt Gi ti
cell and τ i = -----------------T , in the i-th open branch
several q ( 0 ) at the contour origin in closed cell ( GJ ) eff
open n
branches ( GJ ) eff = ( GJ ) closed + ∑ ( GJ )i
i=1
4A 2 1
( GJ ) closed = ---------c , ( GJ ) i = G i  --- b i t i3
ds 3 
°∫ -----
Gt
-

7.11 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.

Thin-Walled Structures 183


Article 7.12

7.12 Problems
3b ⁄ 2
1.Determine torsion constant J and the magnitude of the maximum shear stress
τ max for the section shown.
t
t 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).
b

50 mm 50 mm
b 2t
10mm
20mm

50 mm 1.5mm 50 mm
1.5mm

10mm 20mm
(a) (b)

Figure for problem 2

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
θ z ( z ) , 0 ≤ z ≤ 2500mm .

y 40 N/mm 40N/mm

x 1.2mm G = 18GPa
G = 26GPa y
2.1mm 1.2mm x 250mm
2.1mm

1000mm
θz 2500mm
cross section
z
Figure for problem 3

184 Thin-Walled Structures


Problems

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 I xx = 0.618731r 3 t . All branches have the same thickness t and same shear modulus
G.
a) Determine the shear flow distributions q 1 ( s ), q 2 ( s ), and q 3 ( s ) in each branch due to transverse bend-
ing in terms of shear force V y , the shear flow q 0 at point O, and radius r. Take positive shear flow
counter-clockwise around the section as shown with the contour origin at point O. Note: in the circular
branch 2 take s 2 = rθ , 0 ≤ θ ≤ π ⁄ 2 , and hence the y-coordinate to point s2 is y ( s 2 ) = r cos ( π ⁄ 4 + θ )

b) Determine the shear flow q 0 at point O in terms of V y and r by enforcing zero unit twist.(answer:
q 0 = 0.495416V y ⁄ r )

c) Determine the location of the shear center e from point O in terms of r by torque equivalence.
y
Vy s 2, q 2

s 1, q 1
r
S.C. 45° O x
axis of symmetry
45°
t s 3, q 3

Figure for problem 4


e

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 flows q1 and q2 in the exterior branches.
b) Determine the torsion constant J.
q1 q2

a cell 1 cell 2 t, typ.


T

a 2a

Figure for problem 5

Thin-Walled Structures 185


Article 7.12

6. Calculate the shear flows q 1 and q 2 for the two-cell beam shown below. The applied torque is 1000 lb.-in
(clock-wise).

25 39
t = 1 ⁄ 4 typical

30 q1 q2 all dimensions in inches

20 36

7. A small slit is cut in the lower exterior branch in cell 2 of problem 5.


q

a small slit t, typ.


T

a 2a
Figure for Problem 7
a) Determine the torsion constant J.
b) Determine the maximum shear stress.

8. For the homogeneous cross section shown below, take dimensions b = 30 mm and t = 3 mm .
a) Determine the torsion constant J.
b) For a torque T = 600 Nm , determine shear stress in the closed circular portion and the maximum
shearing stress in the open portion.

9. Determine the location e y of the shear center in terms of dimension b for the section shown below.
y

b
--- the thickness of all branches is t
2
y c = 0.8487b
Figure for Problems 8 and 9
2 π
C x I yy =  --- + --- b 3 t
b SC  3 8
yc
ey

b b

186 Thin-Walled Structures


CHAPTER 8 Work and energy
methods

8.1 Hooke’s law and its consequences1


Historically, the notion of elasticity was first announced in 1676 by Robert Hooke (1635–1703) in the form of an
anagram, ceiinosssttuv. He explained it in 1678 as Ut tensio sic vis, or “the power of any springy body is in the
same proportion with the extension.” Hooke’s law is the constitutive law for a Hookean, or linear elastic, mate-
rial. As stated in the original form, its meaning is not very clear. One way to give it precise meaning is to make
use of the common notion of “springs,” and consider the load–displacement relationship. The second way is to
state it in as an equation relating strain and stress, as is done in the mathematical theory of elasticity. The first
way of “springs” is followed here.

Consider the static equilibrium state of a solid body under the action of P2
external forces, as is shown in Fig. 8.1. Let the body be supported in a man- P1
ner that rigid body motion is impossible, and let it be described with
respect to a rectangular Cartesian reference frame. Consider the three 3
2 1
hypotheses u

1. The body is a continuum.

Under this hypothesis the atomic structure of the body is ignored and Fig. 8.1 Static equilibrium of
a body under external forces
the body is idealized into a geometrical copy in Euclidean space whose
points are identified with the material particles of the body. The continuity
is defined in the mathematical sense with respect to the idealized continuum. Neighboring points remain neigh-
bors under any loading condition. No cracks or holes may open up in the interior of the body under the action of
external load.

To introduce the second hypothesis, consider the action of a set of forces on the body. Let the forces be fixed
in direction and in point of application, and let the magnitudes of all the forces be increased or decreased
together: always bearing the same ratio to each other. Let the forces be denoted by P 1, P 2, …, P n and their mag-

1.Much of this discussion is adapted from: Y.C. Fung, (1965).

Thin-Walled Structures 187


Article 8.1

nitude by P 1, P 2, …, P n . Then, the ratios P 1 :P 2 :P 3 :…:P n remain fixed. When such a system of forces is
applied on the body, the body deforms. Let the displacement at an arbitrary point in an arbitrary direction,
denoted by u , be measure with respect to a rectangular Cartesian reference frame. The second hypothesis is
2. Hooke’s law

u = a1 P1 + a2 P2 + … + an Pn (8.1)

where a 1, a 2, …, a n are constants independent of the magnitude of P 1, P 2, …, P n . The constants a 1, a 2, …, a n


depend on the point where the displacement is measured and on the directions and points of application of the
forces. Hooke’s law in the form of eq. (8.1) can be subjected to experimental examination.

To complete the formulation of elasticity, we need a third hypothesis:


3. There exists a unique unstressed state of the body, to which the body returns whenever all external forces are
removed.

A number of deductions can be drawn from these hypotheses.

8.1.1 The principle of superposition for loads at the same point.

Let P 1 and P 1 ′ be forces in the same direction acting through the same point. Then the resultant displacement u
is equal to the sum of the displacements produced by P 1 and P 1 ′ acting individually, regardless of the order of
application of P 1 and P 1 ′ . This is evident from hypothesis 2 when applied to one load acting on the body, since
constant a 1 is independent of P 1 and P 1 ′ .

8.1.2 Principle of superposition


By a combination of hypotheses 2 and 3, we can show that eq. (8.1) is valid not only for a system of loads for
which the ratios P 1 :P 2 :…:P n remain fixed as originally assumed, but also for an arbitrary set of loads
P 1, P 2, …, P n . In other words, eq. (8.1) holds regardless of the order in which the loads are applied. The constant
a 1 is independent of the loads P 2, P 3, …, P n . The constant a 2 is independent of the loads P 1, P 3, …, P n ; etc.
This the principle of superposition of the load-and-deflection relationship.
• Proof: If a proof of the statement above can be established for an arbitrary pair of loads, then the general the-
orem can be proved by mathematical induction.

Let P 1 and P 2 , with magnitudes P 1 and P 2 , be a pair of arbitrary loads acting at points 1 and 2, respec-
tively. Let the displacement in a specific direction be measured at point 3. See Fig. 8.1. According to
hypothesis 2, if P 1 is applied alone, then at point 3 a displacement u 3 = c 31 P 1 is produced. If P 2 is applied

alone, a displacement u 3 = c 32 P 2 is produced. If P 1 and P 2 are applied together, with ratio P 1 ⁄ P 2 fixed,
then according to hypothesis 2 the displacement can be written as

u 3 = c' 31 P 1 + c' 32 P 2 (a)

188 Thin-Walled Structures


Hooke’s law and its consequences

The question arises whether c' 31 = c 31 and c' 32 = c 32 . The answer is affirmative, as can be shown as

follows. After P 1 and P 2 are applied, we take away P 1 , This produces a change in displacement
– c'' 31 P 1 , and the total displacement is

u 3 = c' 31 P 1 + c' 32 P 2 – c'' 31 P 1 (b)

Now only P 2 acts on the body. Hence, upon unloading P 2 , we shall have

u 3 = c' 31 P 1 + c' 32 P 2 – c'' 31 P 1 – c 32 P 2 (c)

Now all the loads are removed, and u 3 must vanish according to hypothesis 3. Rearranging terms, we
have

( c' 31 – c'' 31 )P 1 = ( c 32 – c′ 32 )P 2 (d)

Since the only possible difference of c' 31 and c'' 31 must be caused by the action of P 2 , the difference

c' 31 – c'' 31 can only be a function of P 2 and not of P 1 . Similarly, c 32 – c' 32 can only be a function of

P 1 . If we write Eq. (d) as

c' 31 – c'' 32 c 32 – c' 32


----------------------- = --------------------
- (e)
P2 P1

then the left-hand side is a function of P 2 alone, and the right-hand side is a function of P 1 alone.
Since P 1 and P 2 are arbitrary numbers, the only possibility for Eq. (e) to be valid is for both sides
equal to a constant k , which is independent of both P 1 and P 2 . Hence,

c' 32 = c 32 – kP 1 (f)
But a substitution of Eq. (f) into (a) yields

u 3 = c' 31 P 1 + c 32 P 2 – kP 1 P 2 (g)

The last term is nonlinear in P 1 and P 2 , and Eq. (g) will contradict hypothesis 2 unless k vanishes.
Hence, k = 0 and c' 32 = c 32 . An analogous procedure shows c' 31 = c'' 31 = c 31 .
Thus, the principle of superposition is established for one and two forces. An entirely similar proce-
dure will show that if it valid for m forces, it is also valid for m+1 forces. Thus, the general theorem
follow by mathematical induction.

The constants c 31, c 32 , etc., are seen to be of significance in defining the elastic property of the solid body.
They are called flexibility influence coefficients.

Thin-Walled Structures 189


Article 8.1

8.1.3 Corresponding forces and displacements and the unique meaning of the total work

Let us now consider a set of external forces P 1, P 2, …, P n acting on the body and define the set of displacements
at the points of application and in the direction of the loads as the displacements “corresponding” to the forces
at these points. The reactions at the points of support are considered as external forces exerted on the body and
included in the set of forces.

Under the loads P 1, P 3, …, P n , the corresponding displacements may be written as

u 1 = c 11 P 1 + c 12 P 2 + … + c 1n P n
u 2 = c 21 P 1 + c 22 P 2 + … + c 2n P n
(8.2)
…………………………………
u n = c n1 P 1 + c n2 P 2 + … + c nn P n
or,
n
ui = ∑c P ij j i = 1, 2, …, n (8.3)
j=1

If we multiply the first equation by P 1 , the second equation by P 2 , etc., and add, we obtain

P 1 u 1 + P 2 u 2 + … + P n u n = c 11 P 12 + c 12 P 1 P 2 + … + c 1n P 1 P n +
c 21 P 1 P 2 + c 22 P 22 + … + c 2n P 2 P n +
……………………………… +
c n1 P 1 P n + c n2 P 2 P n + … + c nn P n2
(8.4)

The quantity above is independent of the order in which the loads are applied. Hence, it has a definite meaning
for each set of loads P 1, P 3, …, P n .

P 1 ⁄ P 2 = constant Now, in a special case, the meaning of the quantity on the left-hand
side of eq. (8.4) is clear. This is the case in which the ratios
P1
P 1 :P 2 :P 3 :…:P n are kept constant and the loading increases very
P2
slowly from zero to the final value. In this case, the corresponding dis-
placements also increase proportionally and slowly. It should be clear
0 u1 0 u2 1
that the work done by the force P 1 is exactly --- P 1 u 1 , that of P 2 is
Fig. 8.2 Load-displacement 2
plots for proportional loading 1
--- P 2 u 2 , etc. See Fig. 8.2.
2

Hence, we conclude from eq. (8.4) that the total work done, W, by the set of forces is independent of the or-
der in which the forces are applied.
n
1
W = ---
2 ∑P u i i (8.5)
i=1

190 Thin-Walled Structures


Hooke’s law and its consequences

8.1.4 Maxwell’s reciprocal theorem


An important property of the influence coefficients of corresponding displacements follows immediately.

The influence coefficients for corresponding forces and displacements are symmetric.
c ij = c ji (8.6)

In other words, the displacement at point i due to a unit load at another point j is equal to the displacement at j
due to a unit load at i, provided that the displacements and forces “correspond,” i.e., that they are measured in
the same direction at each point.

• Proof: Consider two forces P 1 and P 2 . First, apply P 1 slowly with P 2 = 0 . At the final value of P 1 , the
1
displacement of point 1 is c 11 P 1 and the displacement of point 2 is c 21 P 1 . The work done is --- c 11 P 12 . With
2
P 1 held fixed, apply P 2 slowly until P 2 attains its final value. The additional displacement at point 1 is
1
c 12 P 2 and the additional displacement at point 2 is c 22 P 2 . The additional work done is P 1 c 12 P 2 + --- c 22 P 22 .
2
When the forces are applied in the order P 1, P 2 the total work done, as shown in Fig. 8.3, is

1 1
W = --- c 11 P 12 + c 12 P 1 P 2 + --- c 22 P 22
2 2











pt. 1 pt. 2

P1 P2
Fig. 8.3 Load-
displacement plots for the
loading sequence P1, P2.
u1 u2
0 0
c 11 P 1 c 12 P 2 c 21 P 1 c 22 P 2

Second, apply P 2 slowly with P 1 = 0 . At the final value of P 2 , the displacement of point 1 is c 12 P 2 and the
1
displacement of point 2 is c 22 P 2 . The work done is --- c 22 P 22 . With P 2 held fixed, apply P 1 slowly until P 1
2
attains its final value. The additional displacement at point 1 is c 11 P 1 and the additional displacement at point
1
2 is c 21 P 1 . The additional work done is P 2 c 21 P 1 + --- c 11 P 12 . When the forces are applied in the order P 2, P 1
2
the total work done, as shown in Fig. 8.4, is

Thin-Walled Structures 191


Article 8.1

1 1
W′ = --- c 11 P 12 + --- c 22 P 22 + c 21 P 1 P 2
2 2












pt. 1 pt. 2

P1 P2
Fig. 8.4 Load-
displacement plots for the
loading sequence P2, P1.
u1 u2
0 0
c 12 P 2 c 11 P 1 c 22 P 2 c 21 P 1

But according to the deduction presented in Article 8.1.3, W = W′ for arbitrary P 1, P 2 . Hence, c 12 = c 21 ,
and the theorem is proved.

The reciprocal relation may be put into a different form, sometimes more convenient in applications.

8.1.5 Betti-Rayleigh reciprocal theorem

Let a set of loads P 1, P 3, …, P n produce a set of corresponding displacements u 1, u 2, …, u n . Let a second set of

loads P 1 ′, P 2 ′, …, P n ′ acting in the same directions and having the same points of application as those of the
first, produce the corresponding displacements u 1 ′, u 2 ′, …, u n ′ . Then

P 1 u 1 ′ + P 2 u 2 ′ + … + P n u n ′ = P 1 ′u 1 + P 2 ′u 2 + … + P n ′u n (8.7)

In other words, in a linear elastic solid, the work done by a set of forces acting through the corresponding dis-
placements produced by a second set of forces is equal to the work done by the second set of forces acting
through the corresponding displacements produced by the first set of forces.

A straightforward proof is furnished by writing out the u i and u i ′ in terms of P i and P i ′ ( i = 1, 2, …, n ),


with the appropriate influence coefficients. Then form the two equations
n n n n n n

∑ P ′u i i = ∑ ∑c P P ′ ij j i ∑ P u ′ = ∑ ∑ c P ′P
i i ij j i
i=1 i = 1j = 1 i=1 i = 1j = 1

In the second equation, interchange the order of summation of the right-hand side. Then let indices i → j and
j → i . Use symmetry of the flexibility influence coefficients, eq. (8.6), to complete the proof. These steps are
shown below.
n n n n n n n n

∑ Pi ui ′ = ∑∑ c ij P j ′P i = ∑∑ c ji P i ′P j = ∑∑ c ij P i ′P j = ∑ P ′u i i
i=1 j = 1i = 1 i = 1j = 1 i = 1j = 1 i=1

192 Thin-Walled Structures


Hooke’s law and its consequences

8.1.6 Stiffness influence coefficients


Assume that the displacement-force system given by eq. (8.2) can be inverted so that the forces may be expressed
in terms of the displacements as
P 1 = k 11 u 1 + k 12 u 2 + … + k 1n u n
P 2 = k 21 u 1 + k 22 u 2 + … + k 2n u n
(8.8)
…………………………………
P n = k n1 u 1 + k n2 u 2 + … + k nn u n
or
n
Pi = ∑k u ij j i = 1, 2, …, n (8.9)
j=1

where constants k ij are called stiffness influence coefficients. In matrix notation, we write the displacement-force
form of Hooke’s law as
{u} = [c] {P}







n×1

n×n n×1 (8.10)

and the force-displacement form as


{P} = [k] {u}







n×1 n×n n×1 (8.11)

where [c] is called the flexibility matrix and [k] is called the stiffness matrix. Both matrices are square of order
n × n . In matrix algebra the stiffness matrix is the inverse of the flexibility matrix, or

[ k ] = [ c ] –1 (8.12)

The inverse matrix has the property that

[ c ] –1 [ c ] = [ c ] [ c ] –1 = [ I ] (8.13)

where [I] is the n × n identity matrix; i.e., the identity matrix is a square matrix with all diagonal elements equal
to unity and all off-diagonal elements equal to zero.

Maxwell’s theorem, eq. (8.6), shows that the flexibility matrix is symmetric. In matrix algebra symmetry is
written as
T
[c] = [c] (8.14)

where the superscript T means matrix transpose; i.e., the matrix obtained by interchanging its rows with its col-
umns. Since the flexibility matrix is symmetric, the stiffness matrix is also symmetric. That is,

[ k ]T = [ k ] (8.15)

• Proof: By definition
[ c ] –1 [ c ] = [ I ]
Take the transpose of this equation to get

Thin-Walled Structures 193


Article 8.2

T T
( [ c ] –1 [ c ] ) = [ I ]
Use the facts the transpose of the product of two matrices is equal to the product of the transpose of the sec-
ond matrix time the transpose of the first matrix, and that the transpose of the identity matrix is equal to itself,
to get
[ c ] T ( [ c ] –1 ) T = [ I ]
By symmetry of the flexibility matrix this becomes
[ c ] ( [ c ] –1 ) T = [ I ]
Pre-multiply by the inverse of the flexibility matrix
[ c ] –1 [ c ] ( [ c ] –1 ) T = [ c ] –1 [ I ] = [ c ] –1
But
[ c ] –1 [ c ] = [ I ]
Hence,
( [ c ] –1 ) T = [ c ] –1

Again, by definition [ c ] –1 ≡ [ k ] . Thus

[ k ]T = [ k ]

8.2 First law of thermodynamics


Further insight can be gained from the first law of thermodynamics. The body is a closed system of particles, not
exchanging mass across its boundaries. When this body is thermally insolated and thermal expansions are
neglected, the first law states that the work done on the body by the external forces in a certain time interval is
equal to the increase in the kinetic energy and internal energy in the same interval. If the process is so slow that
the kinetic energy can be ignored, i.e., quasi-static, the work done is seen to be equal to the change in internal
energy. If the internal energy is taken as zero in the unstressed state, the stored internal energy is called the strain
energy. Writing U for the strain energy we have from eq. (8.5),
n
1
W = U = ---
2 ∑P u i i (8.16)
i=1

From what was said about the deduction presented in Article 8.1.3, we conclude that the strain energy in a linear
elastic solid is independent of the order in which the given forces are applied.

U* Under the action of a single load P 1 acting on the body, the strain energy
P1
1
U = --- P 1 u 1 is interpreted as the area between the load-displacement relation
U 2
and the abscissa as shown in Fig. 5. The area between the load-displacement
0 relation and the ordinate, denoted as U * , is defined as the complementary strain
u1
Fig. 8.5 Strain Energies energy. From this figure we can see that these energies are related by

194 Thin-Walled Structures


Castigliano’s theorems

U * = – U + P1 u1
The above expression for the body subjected to multiple loads generalizes to
n
U* = –U+ ∑P u j j (8.17)
j=1

Substitute eq. (8.16) for U into this expression, and we see that the complementary strain energy for a linear elas-
tic solid is given by
n
1
U* = ---
2 ∑P u i i (8.18)
i=1

Hence, for a linear elastic solid under isothermal conditions the value of the strain energy is equal to the value of
the complementary strain energy. It is preferred, as will be clear later, to write the strain energy in term of dis-
placements or strains, and to write the complementary strain energy in term of forces, or stresses. To obtain the
strain energy in terms of displacements substitute eq. (8.9) for the forces in eq. (8.16) to get
n n
1
U = ---
2 ∑ ∑k u u ij i j (8.19)
i = 1j = 1

To obtain the complementary strain energy substitute eq. (8.3) for the displacements in eq. (8.18), to get
n n
1
U * = ---
2 ∑ ∑c P P ij i j (8.20)
i = 1j = 1

8.3 Castigliano’s theorems


For the case of two loads, or n = 2 , the strain energy in eq. (8.19) reduces to

1
U = --- ( k 11 u 12 + k 12 u 1 u 2 + k 21 u 2 u 1 + k 22 u 22 ) (8.21)
2
Since the flexibility influence coefficients are independent of the loads, and the stiffness influence coefficients are
determined as the inverse of the flexibility coefficients via eq. (8.12), the stiffness influence coefficients are inde-
pendent of the loads. If the “loads” are interpreted as independently specified displacements, then the stiffness
influence coefficients are independent of the displacements. Now take the partial derivatives of this strain energy,
eq. (8.21), with respect to displacements u 1 and u 2 to get

∂U 1
-------- = k 11 u 1 + --- ( k 12 + k 21 )u 2 = k 11 u 1 + k 12 u 2 = P 1
∂u 1 2
(8.22)
∂U 1
-------- = --- ( k 12 + k 21 )u 1 + k 22 u 2 = k 21 u 1 + k 22 u 2 = P 2
∂u 2 2
where we used symmetry of the stiffness influence coefficients, eq. (8.15), and eq. (8.9) to recognize that the lin-
ear terms in the displacements are equal to the appropriate forces. From the manipulations in eq. (8.22) we see
that

Thin-Walled Structures 195


Article 8.3

∂U ∂U
-------- = P 1 and -------- = P 2 (8.23)
∂u 1 ∂u 2

The complementary strain energy for the case of two loads, or n = 2 , from eq. (8.20) is

1
U * = --- ( c 11 P 12 + c 12 P 1 P 2 + c 21 P 2 P 1 + c 22 P 22 ) (8.24)
2
If we take the partial derivatives of this complementary strain energy with respect to P 1 and P 2 we get

∂U * 1
---------- = c 11 P 1 + --- ( c 12 + c 21 )P 2 = c 11 P 1 + c 12 P 2 = u 1
∂P 1 2
(8.25)
∂U * 1
---------- = --- ( c 12 + c 21 )P 1 + c 22 P 2 = c 21 P 1 + c 22 P 2 = u 2
∂P 2 2
where symmetry of the flexibility influence coefficients, eq. (8.6), was used, and from eq. (8.2) it is recognized
that the linear terms in the loads are equal to the appropriate displacements. From the manipulations in eq. (8.25)
we see that

∂U * ∂U *
---------- = u 1 and ---------- = u 2 . (8.26)
∂P 1 ∂P 2

Extending these results to the case of n-loads applied to the body we obtain the two Castigliano’s theorems.
Castigliano’s first theorem is
∂U
P i = ------- i = 1, 2, …, n (8.27)
∂u i
and Castigliano’s second theorem is

∂U *
u i = ---------- i = 1, 2, … , n (8.28)
∂P i

in which u i and P i are corresponding displacements and forces.

8.3.1 Extensions to include a couple and rotation


P2 We can extend Castigliano’s second theorem, eq. (8.28), and
P1 Hooke’s law, eq. (8.2), to include the moment of a couple acting
on the body and the rotation of the arm connecting the couple.
P3 3
2 1
a First, consider the extension of Castigliano’s second theorem. As
P4
shown in Fig. 8.6, the forces P 3 and P 4 form a couple with an
4
arm of length a if P 3 = P and P 4 = P . That is, forces P 3 and
Fig. 8.6 Static equilibrium of a body
under external forces including a P 4 are functions of the force P . Take the partial derivative of the
couple complementary strain energy with respect to force P and use the
chain rule to get

∂U * ∂U * ∂P ∂U * ∂P
---------- = ---------- --------3- + ---------- --------4-
∂P ∂P 3 ∂P ∂P 4 ∂P

196 Thin-Walled Structures


Castigliano’s theorems

∂U * ∂U * ∂P ∂P
But, u 3 = ---------- , u 4 = ---------- , and --------3- = --------4- = 1 . So
∂P 3 ∂P 4 ∂P ∂P

∂U *
---------- = u 3 + u 4
∂P
For small displacements, u 3 + u 4 = aθ , where θ is the small 3
rotation of the moment arm in radians, as is shown in Fig. 8.7. u3
∂U * θ a u3 + u4
Thus, ---------- = aθ . Divide this last equation by the length of the tan θ = ----------------
-
u4 a
∂P
4
1 ∂U * Fig. 8.7 Rotation of the arm of the
moment arm to get --- ---------- = θ . Lastly, the moment of a couple
a ∂P couple
is M = aP , so we get

∂U *
---------- = θ (8.29)
∂M
Also, the work of this couple acting on the linear elastic body is given by
( P 3 u 3 + P 4 u 4 ) ⁄ 2 = P ( u 3 + u 4 ) ⁄ 2 = P ( aθ ) ⁄ 2 = ( Mθ ) ⁄ 2 (8.30)

We can consider a concentrated couple as the limiting case of two equal and opposite forces acting in a
plane at the surface of the body that approach each other, but maintain a constant moment; i.e.,
Limit ( aP ) = M
a→0

In this limiting process P → ∞ . Then, the angle of rotation θ is interpreted as the rotation of an infinitesimal line
element in the plane of the couple.

Second, consider the extension of Hooke’s law to include a concentrated couples and their corresponding
rotations.

8.3.2 Generalized forces and displacements


Define Q i as the magnitude of the generalized force acting at point i on the body, and let q i denote the corre-
sponding generalized displacement at point i, where i = 1, 2, …, n . The product of Q i q i has dimensional unit
of work, or F-L. If Q i is a force, which we previously designated P i , then q i is the corresponding displacement,
which we previously denoted as u i . If Q i is the moment of a concentrated couple with dimensional unit F-L,
then q i is the corresponding rotation in radians of the infinitesimal line element in the plane of the couple at the
point of its application. By defining generalized forces and moments, we can extend Hooke’s law in eq. (8.2) to
include moments and rotations as well as forces and displacements. That is, Hooke’s law is written as
q 1 = c 11 Q 1 + c 12 Q 2 + … + c 1n Q n
q 2 = c 21 Q 1 + c 22 Q 2 + … + c 2n Q n
(8.31)
…………………………………
q n = c n1 Q 1 + c n2 Q 2 + … + c nn Q n

Thin-Walled Structures 197


Article 8.3

In eq. (8.31), the flexibility influence coefficients can have different dimensional units. For example, if q 1 is a
displacement of point 1 on the body and Q 2 is a moment of a couple acting point 2, then the dimensional unit of
flexibility influence coefficient c 12 is F –1 . Since the generalized displacement q 2 corresponding to Q 2 is a rota-
tion in radians and the generalized force Q 1 acting at point 1 is a force corresponding to q 1 , then the dimen-
sional unit of flexibility influence coefficient c 21 is also F –1 . Also, the dimensional unit of c 11 is LF –1 , and c 22
is F –1 L –1 .

Assume that eq. (8.31) can be inverted to give the generalized forces in terms of the generalized displace-
ments and write this linear relationship as
Q 1 = k 11 q 1 + k 12 q 2 + … + k 1n q n
Q 2 = k 21 q 1 + k 22 q 2 + … + k 2n q n
(8.32)
…………………………………
Q n = k n1 q 1 + k n2 q 2 + … + k nn q n

where the stiffness influence coefficients, k ij , also can have different dimensional units.

Castigliano’s first theorem in terms of generalized displacements and forces is as follows:

Castigliano’s first theorem


If the strain energy of an elastic structure is expressed in terms of the independent
generalized displacement components q i , i = 1, 2, …, n , in the direction of the pre-
scribed generalized point forces Q 1, Q 2, …Q n , then the first partial derivative of the
strain energy with respect to the displacement q i is equal to the corresponding force
Q i or

∂U
Q i = ------- i = 1, 2, …, n
∂q i

To apply this first theorem in the continuum analysis of a structure,


• the displacement components must be represented as continuous functions of the spatial coordinates,
• satisfy any specified displacement conditions on the boundary,
• and equate to the generalized displacements q i , i = 1, 2, …, n , at their defined points of application.
For a continuum the strain energy is an integral over the volume of the body of the strain components as will be
shown for bar theory in the next section. Since the strains are the derivatives of the displacement functions with
respect to the spatial coordinates, it is require that the displacement functions be continuous, so that the strains
are at least piece wise continuous, and that the volume integral of the strains, or strain energy, can be computed.
Displacement functions that satisfy continuity conditions and specified boundary conditions are said to be kine-
matically admissible. Castigliano’s first theorem is particularly useful for the response analysis of trusses.

198 Thin-Walled Structures


Unit action states

Castigliano’s second theorem in terms of generalized displacements and forces is as follows:

Generalized form of Castigliano’s second theorem


If an elastic structure is mounted such that rigid body displacements are impossible
and certain generalized point forces Q 1, Q 2, …, Q n act on the structure, in addition to
distributed loads and thermal strains, the displacement component q i , i = 1, 2, …, n
of the point of application of force Q i in the direction of Q i is determined by the equa-
tion

∂U *
q i = ----------
∂Q i

To apply this second theorem in the continuum analysis of a structure,


• the internal actions such as stresses, stress resultants (shear flows and forces), and stress couples (moments),
must be represented as functions of the spatial coordinates that satisfy the differential equations of equilib-
rium in the volume of the structure,
• satisfy specified tractions, or force intensities, on the boundary,
• and contain the generalized forces Q i , i = 1, 2, …, n , as parameters.
For a continuum the complementary strain energy is an integral over the volume of the body of the internal
actions (stresses, shear flows, forces, moments) as will be shown for bar theory in the next section. Internal force
and moment functions that satisfy the differential equations of equilibrium and specified force boundary condi-
tions are said to be statically admissible.

8.4 Unit action states


The following physical interpretation of the flexibility influence coefficients can be given: Assume generalized
force Q 1 = 1 in the appropriate dimensional units, and all other generalized forces are zero. Then, the general-
ized displacements at points 1, 2, …, n are q i = c i1 , i = 1, 2, …, n . That is, the flexibility influence coeffi-
cients c i1 are the generalized displacements at points 1, 2, …, n for a unit generalized force at acting at point 1
with all other generalized forces equal to zero. This special generalized force state is called unit action state 1, or
UAS 1. We have the following definition:

The flexibility influence coefficient c ij represents the generalized displacement at point i in the direction Q i
due to a unit generalized force Q j , all other generalized forces equal to zero.

Thin-Walled Structures 199


Article 8.5

The extension of Maxwell’s reciprocal theorem, eq. (8.6), to


2
include a concentrated couple and its corresponding rotation is
1
illustrated by the following example: Consider a cantilever beam
subjected to a force Q 1 and a moment Q 2 , as shown in Fig. 8.8.
Q1 = 1 Under UAS 1, the rotation at point 2 is c 21 ⋅ 1 . Under UAS 2, the
UAS 1 displacement at point 1 is c 12 ⋅ 1 . The reciprocal theorem states
c 21 that c 21 = c 12 . Hence for these special loading states, the rotation
in radians at point 2 due to a unit force acting at point 1 is equal in
UAS 2 magnitude to the displacement in units of length at point 1 due to
c 12 a unit moment acting at point 2. This fact is very useful in practi-
cal applications.
Q2 = 1
∂q
Also, note that the flexibility influence coefficients c ij = ---------i ,
Fig. 8.8 Illustration of the reciprocal ∂Q j
theorem using unit action states
which can be follows from eq. (8.31). Further note by Cas-
∂U *
tigliano’s second theorem that q i = ---------- , so we can also write
∂Q i
the flexibility influence coefficient as the second mixed partial derivative of the complementary strain energy;
∂2U*
i.e., c ij = ------------------ .
∂Q i ∂Q j

8.5 Unit displacement states


The following physical interpretation of the stiffness influence coefficients can be given: Assume generalized dis-
placement q 1 = 1 in the appropriate dimensional units, and all other generalized displacements are zero. Then,
the generalized forces at points 1, 2, …, n are Q i = k i1 ⋅ 1 , i = 1, 2, …, n . That is, the stiffness influence coef-
ficients k i1 are the magnitudes of the generalized forces at points 1, 2, …, n for a unit generalized displacement
at point 1 with all other generalized displacements equal to zero. This special generalized displacement state is
called unit displacement state 1, or UDS 1. We have the following definition:

The stiffness influence coefficient k ij represents the generalized force at point i in the direction q i due to a
unit generalized displacement q j , all other generalized displacements equal to zero.

The unit displacement states for an airplane wing spar modeled as a cantilevered beam with four degrees of
freedom are shown in Fig. 8.9. The four generalized forces for each unit displacement state are equal in magni-
tude to the four stiffnesses associated with the specified unit displacement, all other displacements equal to zero.

200 Thin-Walled Structures


Strain energies in bar theory

1 3

2 4

k 31
k 11 [ q 1, q 2, q 3, q 4 ] = [ 1, 0, 0, 0 ]
UDS 1
k 41

k 21

k 12 k 32

UDS 2 [ q 1, q 2, q 3, q 4 ] = [ 0, 1, 0, 0 ]

k 22 k 42

k 13
k 33
UDS 3 [ q 1, q 2, q 3, q 4 ] = [ 0, 0, 1, 0 ]

k 23
k 43

k 14 k 34

UDS 4 [ q 1, q 2, q 3, q 4 ] = [ 0, 0, 0, 1 ]
k 24 k 44

Fig. 8.9 Generalized forces for unit displacement states of a cantilever beam model of a wing spar.

8.6 Strain energies in bar theory


We have proved Castigliano’s second theorem using the flexibility influence coefficients c ij , but we do not know
these coefficients. Similarly, we have proved Castigliano’s first theorem using the stiffness influence coefficients
k ij , but we do not know these coefficients.What has been established is that the flexibility influence coefficients
exist under the hypotheses stated. In order to determine the flexibility influence coefficients in a specific applica-
tion, an expression for the complementary strain energy in terms of the generalized forces is required. Similarly,
to determine the stiffness influence coefficients in a specific application, an expression for the strain energy in
terms of the generalized displacements is required. If we use the assumptions of bar theory that plane cross sec-
tions remain plane in the deformed bar, then an approximate expressions for the strain and complementary strain
energies of this particular solid body can be determined. The strain energies of the solid body are only as accurate

Thin-Walled Structures 201


Article 8.6

as are the assumptions underlying the bar theory, and so too, is the accuracy in influence coefficients only as
accurate as the bar theory

The two primary stress components in bar theory are the axial normal stress σ z and shear stress component
tangent to the contour τ zs . Hooke’s law for these stress components is

σ z = E ( ε z – ε zt ) τ zs = Gγ zs (8.33)

where the modulus of elasticity is denoted by E, the axial normal strain by ε z , the axial thermal strain by ε zt , the
shear modulus by G , and the shear strains denoted by γ zs . The thermal strain is given by

ε zt = α ( T – T 0 ) (8.34)

where the coefficient of thermal expansion is denote by α , and the temperature change from the initial stress-free
state is denoted by T – T 0 . For an isotropic material G = E ⁄ ( 2 ( 1 + v ) ) , where Poisson’s ratio is denoted by ν .
Hooke’s law is plotted in Fig. 8.10. In analogy to the discussion in Article 8.2, the area between the Hooke’s law
and the strain axis is defined as the strain energy density U 0 , and the area between the Hooke’s law and the stress
axis is defined as the complementary strain energy density U 0* . These energy densities are energies per unit vol-

σz τ zs

E G
1 1
0 0
ε zt εz γ zs

– Eε zt

Fig. 8.10 Hooke’s law for normal stress and shear stress

ume, or the energy stored in a particle, with dimensional units of the of (F-L)/L3. Hence, the total energy densi-
ties are the integrals over the volume of the bar given by
L L

U = U* *
∫ ∫ U dA dz
0 =
∫ ∫ U dA dz
0 (8.35)
0 A 0 A

where the length of the bar is denoted by L and the cross-sectional area by A. The sum of the areas between
Hooke’s law and the strain axes is computed from the plots in Fig. 8.10 as
1 1 1 2 1 1 1 2
U 0 = --- σ z ( ε z – ε zt ) + --- ( – Eε zt )ε zt + --- Gγ zs = --- [ E ( ε z – ε zt ) ] ( ε z – ε zt ) + --- ( – Eε zt )ε zt + --- Gγ zs
2 2 2 2 2 2
Expand and simplify this equation to get the expression for the strain energy density as
1 1 2
U 0 = --- Eε z2 – Eε zt ε z + --- Gγ zs (8.36)
2 2
The areas between Hooke’s law and the stress axes is computed from the plots in Fig. 8.10 as

202 Thin-Walled Structures


Strain energies in bar theory

2 σ z + Eε zt 2 σ2 2
1 1 τ zs 1 τ zs 1 τ zs
U 0* = --- [ σ z – ( – Eε zt ) ]ε z + --- -----
- = --- [ σ z – ( – Eε zt ) ] ------------------- - = ------z- + σ z ε zt + --- E ( ε zt ) 2 + ------
- + ------ -
2 2G 2 E 2G 2E 2 2G

It turns out that we can neglect the term E ( ε zt ) 2 ⁄ 2 in this last equation, since it is only a function of the specified
temperature change from the stress-free state, and it is not a function of the generalize forces acting on the bar.
The change in temperature and the generalized forces are assumed independently specified quantities. Recall that
it is the derivative of the complementary strain energy with respect to the generalized forces that appears in Cas-
tigliano’s second theorem. Hence, the complementary strain energy density is given as

σ2 τ zs2
U 0* = ------z- + σ z ε zt + ------
- (8.37)
2E 2G

The strain energy densities have the properties that

∂U ∂U *
σ z = ---------0 ε z = ---------0- (8.38)
∂ε z ∂σ z
which are analogous to the first and second Castigliano’s theorems, respectively, given in Article 8.3.2

8.6.1 Solid, mono symmetric cross section


Consider a straight, homogeneous, linear elastic beam referenced to a rectangular Cartesian system x-y-z. Let the
z-axis lie along the longitudinal centroidal axis of the beam, and let the x- and y-axes lie in a plane parallel to the
cross sections normal to the longitudinal axis. See Fig. 8.11. Further assume the beam is symmetric about the y-z
y
y, v y h2
τ zy dA dy
t(y)
statical equivalence
Vy x
σ z dA dA
x
⇔ C
N x h1 ≤ y ≤ h2
z, w θx Mx
z dA = tdy

h1

Fig. 8.11 A bar symmetric about the y-z plane and subjected to loads in this plane
of symmetry. Stresses, stress resultants, and cross-sectional nomenclature.

plane, that it is subjected to loads acting in this plane of symmetry, and that the specified change in temperature is
symmetric about this plane of symmetry. Then, the response of the beam can be mathematically represented by
variables defined in terms y and z. The internal stress resultants in the beam are the axial force N , transverse
shear force V y , and bending moment M x . These stress resultants are defined in terms of the axial normal stress
σ z and shear stress τ zy by

N =
∫ ∫ σ dA
A
z Vy =
∫∫τ
A
zy dA Mx =
∫ ∫ yσ dA
A
z (8.39)

where A is the cross-sectional area, and dA = dxdy .

Thin-Walled Structures 203


Article 8.6

Strain energy In bar theory the axial normal strain is explicit in the cross-sectional coordinates. From eq. (4.41)
on page 75 the axial strain in this symmetric bar is given by

ε z = w′ ( z ) + yθ x′ ( z ) (8.40)

where w ( z ) is the axial displacement function of the particles located by z on the centroidal axis, and θ x ( z ) is
the rotation function of the cross section about the x-axis at location z. The prime indicates ordinary derivative
with respect to z. In the classical bar theory the deformations due to transverse shear are neglected for slender
bars, or γ zy ∼ 0 . For cross sections remaining plane and perpendicular to the centroidal axis in the deformed bar,
eq. (4.12) on page 66 gives θ x = – v′ , where v ( z ) is the y-direction displacement of the centroidal axis. Neglect-
ing transverse shear deformation means that the strain energy in transverse shear is neglected relative to the strain
energy in extension and bending. The strain energy density given by eq. (8.36) reduces to
1
U 0 = --- E [ w′ ( z ) + yθ x′ ( z ) ] 2 – Eε zt [ w′ ( z ) + yθ x′ ( z ) ]
2
Substitute this energy density into first of Eqs. (8.35) to get
L
1 1
U = --- E ( w′ ) 2 1 dA + Ew′θ x ′ y dA + --- E ( θ x ′ ) 2 y 2 dA – Ew′ε zt 1 dA – Eθ x ′ε zt y dA
∫ 2 ∫
A

A
2 ∫
A

A

A
0

where we assumed the change in temperature is independent of coordinate y as well as x. From the geometry of
the cross section, 1 dA = A , y dA = 0 since the x-axis passes through the centroid, and y 2 dA = I xx , where

A

A

A
I xx is the second moment of the cross-sectional area about the x-axis. Hence, the strain energy for the tempera-
ture change only a function of axial coordinate z is given by
L
1 1
U = --- EA ( w′ ) 2 + --- EI xx ( θ x ′ ) 2 – EAε zt ( w′ ) dz
∫ 2 2
(8.41)
0

Complementary strain energy The assumptions and developments of beam theory for this simple, homoge-
neous cross section with no thermal loading lead to a simplification of the axial normal stress as given by eq.
(4.45) on page 75, and the shear stress as given by eq. (6.8) on page 117. The following relationships for the
stresses in terms of the stress resultants are

N M Vy Qx ( y )
σ z = ---- +  -------x y τ zy = – -------------------
- h1 ≤ y ≤ h2 h2 > h1 (8.42)
A  I xx  I xx t

where t is the thickness of the cross-section in the x-direction, and Q x is the first moment of a portion of the
cross-sectional area about the x-axis that extends from the bottom of the beam where y = h 1 to a generic value
of y. Coordinate y = h 1 at the bottom of the beam, and y = h 2 at the top of the beam. See Fig. 8.11. Note that
for a general cross section, the thickness can be a function of y, or t = t ( y ) . However, even for a cross section of
uniform thickness, the first moment of a partial area of the cross section is a function of y, or Q x = Q x ( y ) . For
the type of cross-section under consideration, we can write these geometric integrals as single integrals in y
rather than double integrals in x and y. That is,

204 Thin-Walled Structures


Strain energies in bar theory

h2 y

I xx = y 2 dA y 2 t dy Qx ( y ) =
∫∫ a
=
∫ ∫ yt dy (8.43)
h1 h1

Note that Q x ( h 2 ) = 0 , since the x-axis passes through the centroid of the cross section. Refer to eq. (6.2) on
page 116.

Substitute the bar theory relations for the normal stress and shear stress from eq. (8.42) into the complemen-
tary strain energy density given by eq. (8.37). Assume the change in temperature is a function of z only and inte-
grate the complementary strain energy density over the volume of the bar. The total complementary strain energy
is
L
M x2 V y2 2
U*
N2
1 dA + ------------- y 2 dA + ------------- Q NM
------x dA + -------------x-
N t
=
∫ ------------2-
2EA ∫∫
A
2EI xx 2 ∫∫
A
2GI xx 2
-
∫∫
A
 t EAI xx ∫ ∫ y dA + ---A- ε ∫ ∫ 1 dA dz
A
z
A
(8.44)
0

2
But,
∫ ∫ 1 dA = A , ∫ ∫ y dA = 0 since the x-axis passes through the centroid, and ∫ ∫ y dA = I
A A A
xx . Hence, the

complementary strain energy is


L
N2 M x2 V y2
U* = - + Nε zt dz
∫ ----------- + ------------
2EA 2EI xx 2GA s
- + ------------ (8.45)
0

where the cross-sectional area effective in shear, A s , is defined by

2
1 1
----- ≡ -----
- Q
------x dA
A s I xx 2 ∫∫
A
 t
(8.46)

8.6.2 A single cell, thin-walled bar having a mono symmetric cross-sectional contour
The complementary strain energy is given by
L

U* *
=
∫ °∫ U 0 t ds dz (8.47)
0 C

where s is the arc-length coordinate along the closed contour C, the dif-
ferential cross-sectional area is approximated as dA = t ( s )ds , and t ( s )
τ zs
is the wall thickness. The thickness can be a function of the arc-length s
coordinate s on the cross-sectional contour C. In thin-walled bar theory
y
the major stress components are the axial normal stress σ z and the shear σz
x
stress tangent to the contour τ zs as shown in Fig. 8.12. These stress
components are assumed uniform in the direction through the thickness z
of the wall. For a wall made of a Hookean material, the complementary t(s)
contour C
strain energy density U 0* is given by eq. (8.37), and the thermal strain is
Fig. 8.12 Closed contour
given by eq. (8.34). Assume the material properties E, G, and α are uni- symmetric about the x-axis
form along the contour C. This assumption of uniform material proper-

Thin-Walled Structures 205


Article 8.6

ties is not strictly necessary; we could let the material properties vary with s, as long as the properties are also
symmetric about the single axis of symmetry in the cross section.

In the cross section of the bar as shown in Fig. 8.13, let the x -axis coincide with the axis of symmetry, and
let the y-axis be normal to the axis of symmetry, such that the x – y – z is a right-handed Cartesian axis system.
The origin of the Cartesian axis system is at a point labeled O where x = 0 and y = 0 . The coordinates of the
cross-sectional contour are represented parametrically by the equations [ x ( s ), y ( s ) ] , where
2
ds 2 = ( dx ) + ( dy ) 2 . The centroid is located at x = x c and the shear center at x = e x as is depicted in Fig.
8.13. The area of the cross section is denoted by A, the second area moment about the axis of symmetry by I xx ,
and the area enclosed by the cross-sectional contour C is denoted by A c . These geometric properties are given by

2 1
A =
°∫ t ( s )ds
C
Qx =
°∫ y ( s )t ( s ) ds = 0
C
I xx =
°∫ y ( s )t ( s )ds
C
A c = --- r ( s )t ( s ) ds
2 °∫
C
(8.48)

Note that the first area moment of the total cross-sectional area, which is denoted by Q x , vanishes, since the x -
axis coincides with the axis of symmetry. From eq. (7.31) on page 159, the coordinate normal to the contour at s
with respect to origin O is

dy dx
r ( s ) = x ( s ) ------ – y ( s ) ------
ds ds

y y
q ( s )ds
Vy
h s
--- y(s)
T 2 Ac r(s)
Mx
x, x x, x
O xc e x ( s.c. ) O xc ex x(s)
h s = 0
---
2

t(s) t(s)

Fig. 8.13 Bar cross-sectional resultants, contour coordinates, and the shear flow

The axial normal stress on the contour is given by

N ( z ) M x ( z )-y
σ z ( s, z ) = ----------- + ------------- (s) (8.49)
A I xx

where the axial normal force is denote by N ( z ) , the bending moment by M x ( z ) , both of which are functions of
the axial coordinate z. Let q ( s, z ) the shear flow distribution tangent to the contour in the cross section at z, taken
positive counter-clockwise as is the contour coordinate. The shear stress is given by
q ( s, z )
τ zs = --------------- (8.50)
t(s)

206 Thin-Walled Structures


Strain energies in bar theory

We assume the temperature distribution in the cross section is at most a linear function of the y-coordinate,
where for a symmetric section – h ⁄ 2 ≤ y ≤ h ⁄ 2 with the height of the cross section denoted by h. Let T 1 – T 0
denote the temperature change at the bottom of the bar where y = – h ⁄ 2 , and let T 2 – T 0 denote the temperature
change at the top of the bar where y = h ⁄ 2 . Then, the temperature in eq. (8.34) at a given cross section is
assumed to have the form

1 y 1 y 1 y 1
T – T 0 = ( T 1 – T 0 )  --- – --- + ( T 2 – T 0 )  --- + --- – --- ≤ --- ≤ ---
 2 h  2 h 2 h 2
and this equation is re-written as

T 2 – T 1
T – T 0 = T – T 0 +  ----------------
- y T = ( T1 + T2 ) ⁄ 2 (8.51)
 h 

where T denotes the average temperature. The thermal strain becomes

T 2 – T 1
ε zt = α T – T 0 +  ----------------
- y (8.52)
 h 

From Eqs. (8.47) and (8.37), the strain energy per unit length of the beam is

*
σ z2 t
τ zs2

°∫ U
C
0 t ds =
C
2E
-+σ ε
°∫ ------ z z
2G
- tds
+ ------

Now substitute eq. (8.49) for the axial normal stress, eq. (8.52) for the thermal strain, and eq. (8.50) for the shear
stress, to get the strain energy per unit length of the bar as

 1 N ( z ) Mx ( z ) 2
N ( z ) M x ( z )-y α ( T2 – T1 ) 1 q 2
- y ( s ) + -------  --- tds
U 0* t ds =  ------- ----------- + --------------y ( s ) + ----------- + ------------- ( s ) α ( T – T 0 ) + -------------------------
°∫
C
°∫
C
 2E A I xx A I xx h 2G  t  

Expanding the terms in this equation we get

 N2 NM M x2 N N
* t ds =  ------------2- tds + -------------x- ytds + ------------- y 2 tds + ---- α ( T – T 0 ) tds + ------- α ( T 2 – T 1 ) ytds +
°∫ U
C
0
 2EA
C
°∫ EAI xx
C
°∫ 2EI 2
xx
C
°∫ A
C
Ah °∫ C
°∫
Mx M q2 
------- α ( T – T 0 ) ytds + --------x-α ( T 2 – T 1 ) y 2 tds +  --------- ds 
I xx °∫
C
I xx h °∫
C
 2Gt °∫
C

Using the expressions for the contour integrals listed in eqs. (8.48), the latter equation reduces to

N2 M x2 T 2 – T 1 q2
- + Nα ( T – T 0 ) + M x α  ----------------
U 0* t ds = ----------- + ------------ - +  --------- ds
°∫
C
2EA 2EI xx  h   2Gt °∫
C
(8.53)

Now the total complementary strain energy as given in eq. (8.47) becomes
L
N2 M x2 T 2 – T 1 q2
U* = - + Nα ( T – T 0 ) + M x α  ----------------
- +  --------- ds dz
∫ ----------- + ------------
2EA 2EI xx  h   2Gt°∫
C
(8.54)
0

Thin-Walled Structures 207


Article 8.6

The shear flow is determined by the superposition of the flow due to the transverse shear force V y ( z ) and the
the flow due to torque T(z). Using Bredt’s formula, eq. (7.48) on page 165, for the portion due to the torque, this
total shear flow is written as
T(z)
q ( s, z ) = q ( s ) Vy ( z )
+ ---------- (8.55)
2A c

where q ( s ) Vy
denotes the shear flow distribution due to the transverse shear force V y ( z ) acting through the
shear center. Axial equilibrium of a segment of the contour from s = 0 to s = s , combined with moment equi-
librium of the bar about the x-axis, determines the shear flow due to the transverse shear force. The details of this
axial equilibrium procedure are presented in Article 6.2 with the pertinent result given by eq. (6.13) on page 120.
For this symmetric cross section under the action of transverse shear force V y ( z ) , eq. (6.13) reduces to

Vy ( z )
q(s) = q 0 ( z ) – ------------
-Q (s) (8.56)
Vy ( z ) I xx x
where the first area moment of the segment of the contour about the x-axis is
s

Qx ( s ) =
∫ y ( s )t ( s ) ds (8.57)
0

and q 0 ( z ) is the shear flow at the contour origin (s = 0). Note that the first area moment evaluates to zero if the
integration in eq. (8.57) is taken completely around the contour, as is shown by the second of Eqs. (8.48). The
shear flow at the contour origin is determined by the condition that the unit twist vanishes for the transverse shear
force acting through the shear center as discussed Article 7.6. From eq. (7.42) this vanishing of the unit twist is
written as
dθ z 1 q(s)
-------- = --------- ---------- ds = 0
dz 2A c Gt °∫
C
(8.58)

Substitute eq. (8.56) for the shear flow into this unit twist condition to find

Qx ( s )
y
∫ -------------ds
V (z) ° C
Gt
q 0 ( z ) = ------------
- ----------------------- (8.59)
I xx ds
°∫
C
------
Gt

Now the shear flow due to the transverse shear force in eq. (8.55) may be written in the form

Qx ( s )
-------------ds
1 C Gt °∫
q(s) Vy ( z )
= F y ( s )V y ( z ) F y ( s ) = ------ ----------------------- – Qx ( s ) (8.60)
I xx ds
Gt °∫
------
C

The function F y ( s ) represents the shear flow distribution on the contour for unit value of the transverse shear
force acting through the shear center, and it has dimensional units of L –1 . Consequently the total shear flow is in
eq. (8.55) is

208 Thin-Walled Structures


Strain energies in bar theory

T(z)
q ( s, z ) = F y ( s )V y ( z ) + ---------- (8.61)
2A c
in which both the transverse shear force and the torque are functions of the axial coordinate z. Substitute eq.
(8.61) for the total shear flow in eq. (8.54) to get

q2 
 -------- V 2 ( Fy ( s ) ) 2 Vy T Fy ( s ) T2 ds
- ds = -----y- -------------------
- ds + --------
- ------------- ds + ------------------2- -----
°∫
C
 2Gt 2 °∫
Gt
C
2A c Gt °∫
C
2 ( 2A c ) Gt
-
°∫
C

Equation (8.58) implies that the unit twist due to shear flow distribution F y ( s ) vanishes; i.e.,

Fy ( s )
- ds = 0
°∫ ------------
C
Gt

Hence,

q2 
 -------- V 2 ( Fy ( s ) ) 2 T2 ds
- ds = -----y- -------------------
- ds + ------------------2- -----
°∫
C
 2Gt 2 °∫
C
Gt 2 ( 2A c ) Gt
-
°∫
C

and combining this result with the total complementary strain energy given by eq. (8.54) we get

L
N2 M x2 T 2 – T 1 V y2 T2
U* = - + Nα ( T – T 0 ) + M x α  ---------------- - + -------------------- dz
- + --------------------
∫ ----------- + ------------
2EA 2EI xx  h  2 ( GA ) eff 2 ( GJ ) eff
(8.62)
0

Equation (8.62) is total complementary strain energy for the bar, and ( GA ) eff denotes the effective transverse
shear stiffness and ( GJ ) eff denotes the effective torsional stiffness. These effective bar stiffnesses are given by
the formulas

Qx ( s ) 2
-------------ds
1 ( Fy ( s ) ) 2 1
- ds = -----
°∫
C
Gt ds
------------------ = ------------------- - – Q x ( s ) ------
( GA ) eff °∫ Gt 2
I xx °∫ -----------------------
ds Gt
(8.63)
C C
°∫
C
------
Gt

and

4A 2
( GJ ) eff = ---------c- (8.64)
ds
Gt °∫
------
C

For the case of uniform shear modulus G around the contour, let A s denote the cross-sectional area effective in
transverse shear. Then, ( GA ) eff = GA s and from eq. (8.63) we get

Qx ( s ) 2
-------------ds
1 ( Fy ( s ) ) 2 1
-------------------- ds = -----
°∫
C
t ds
----- = - – Q x ( s ) -----
As °∫ t 2
I xx °∫ -----------------------
ds t
(8.65)
C C
°∫
C
-----
t

Thin-Walled Structures 209


Article 8.6

Also, for uniform shear modulus, the torsion constant from eq. (8.64) is determined by

4A 2
J = --------c- (8.66)
ds
°∫
-----
C
t

8.6.3 Unit twist formula by Castigliano’s second theorem


Consider a bar with a closed cross-sectional contour under the general deformations of extension, bending, trans-
verse shear, and torsion, but restrained against rigid body displacements. If an imaginary cut of the bar is made
perpendicular to its reference axis at z, then the twist at location z can be determined from Castigliano’s second
theorem. Assuming the twist at z = 0 vanishes, the twist of the cross section at z in terms of the torque T at z is
given by
*
∂U
θ z ( z ) = ---------
∂T
where the complementary strain energy is given by
z
* *
U =
∫ °∫ U 0 t ds dz
0 C

The complementary strain energy per unit length is given in eq. (8.53), and in this equation only the shear flow is
a function of the torque. Hence,

z  z
∂ q2 
 -------- q  ∂q
θ z ( z ) = ------ - ds dz = - ------ ds dz
∂T  ∫ °∫
0 C
 2Gt 

∫ °∫ -----
Gt  ∂T
0 C

From eq. (8.55) ∂q ⁄ ∂T = 1 ⁄ ( 2A c ) . The twist of the cross section at z is


z
1 q
θz ( z ) = --------- ------ ds dz
∫ 2A c Gt°∫
C
0

The unit twist is the derivative of the twist with respect to z, or dθ z ⁄ dz . Using the fundamental theorem of inte-
gral calculus, the derivative of the right-hand side of the last equation with respect to z is the function in the inte-
grand. Hence,

dθ 1 q
--------z = --------- ------ ds
dz 2A c Gt °∫
C

This formula for the unit twist is the same as given by eq. (7.42) on page 163, and it was derived in Article 7.4
using Hooke’s law for shear, the strain-displacement relationship for shear, and the assumption that the cross-sec-
tional contour remains rigid in its own plane.

210 Thin-Walled Structures


Transverse shear deformations – Timoshenko theory

8.7 Transverse shear deformations – Timoshenko theory


Classical beam theory neglects the effects of transverse shear stresses on the deformation of slender beams. That
is, classical theory assumes that plane cross sections before deformation remain plane and perpendicular to the
neutral axis (z-curve) in the deformed beam. For thin-walled beams the effect of transverse shearing deformation
can be significant even for slender beams. An approximate method to account for transverse shearing deforma-
tions 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 flexure formula for the bending normal stress obtained from pure
bending is sufficiently accurate in the presence of transverse shear, and consequently the shear flow formula
derived from equilibrium conditions using the flexure formula is also sufficiently accurate. The beam theory that
is developed in this section is sometimes called Timoshenko beam theory to distinguish it from classical beam
theory. The objective of this article is to determine Hooke’s law for an asymmetric cross section that relates trans-
verse shear deformations to the transverse shear forces using Castigliano’s second theorem. The major results are
given in eqs. (8.69) and (8.79) below. For a mono-symmetric cross section discussed in Article 8.6.2, the trans-
verse shear compliance was derived and it is given by eq. (8.63).

Consider a linear elastic beam element of length dz that is not subjected to distributed loads; i.e, distributed
load intensities p x = p y = p z = t z = 0 . At the end cross sections of the element it is subjected to bending
moments M x and M y , transverse shear forces V y and V x , and no axial normal forces or torques. The work of
these actions on the material in the element is
1 z + dz
W 0 dz = --- [ M x θ x + M y θ y + V x u + V y v ]
2 z

where the work per unit length is denoted by W 0 . Divide by dz , let dz → 0 , and assume the quantity in brackets
is a continuous function of z. Then the work per unit length is

d  1--- 
W0 =  [ Mx θx + My θy + Vx u + Vy v ] 
dz 2 

Take the derivative of the product terms to get

1 dM dM ˙ dV dV dθ dθ du dv
W 0 = --- ----------x θ x + ----------y θ y + ---------x u + ---------y v + M x --------x + M y --------y + V x ------ + V y ------
2 dz dz dz dz dz dz dz dz

Castigliano’s second theorem requires the following equilibrium differential equations be satisfied
dM x dV y dM dV x
---------- = V y --------- = 0 ----------y = V x --------- = 0
dz dz dz dz
Hence, the work done per unit length for the element in equilibrium reduces to
1
W 0 = --- [ M x θ′ x + M y θ′ y + V x ψ x + V y ψ y ] (8.67)
2
where the primes denote derivatives with respect to z, and the beam transverse shear strains are defined by

du dv
ψx ≡ + θy ψy ≡ + θx (8.68)
dz dz

Thin-Walled Structures 211


Article 8.7

The graphical representation of the beam transverse shear strains are shown in Fig. 8.14.

θy θx

x y
du dv
w ------ w ------
dz dz
π π
u --- – ψ x v --- – ψ y
2 2

z z
(a) x-z plane (b) y-z plane
Fig. 8.14 Transverse shear strains in (a) the x-z plane, and (b) the y-z plane

The rotation gradients are related to the bending moments by Hooke’s law which is given by eq. (4.26) on
page 69. This equation is repeated below

θ′ x 1 I – I xy M x
- yy
= ---------------------------------
2 )
θ′ y E ( I xx I yy – I xy – I xy I xx M y

For the linear elastic beam element, the beam shear strains are related to the shear forces by

ψ x = c xx V x + c xy V y
(8.69)
ψ y = c yx V x + c yy V y

where c xx, c xy, c yx, c yy are the flexibility influence coefficients for the cross section of the beam. These influence
coefficients are independent of the shear forces, and by Maxwell’s reciprocal theorem c xy = c yx . (Note: For the
mono-symmetric cross section discussed in Article 8.6.2, the result given in eq. (8.63) identifies
c yy = 1 ⁄ ( GA ) eff .) Substituting these material laws into the work per unit length (8.67) we get

1  M x ( I yy M x – I xy M y ) M y ( – I xy M x + I xx M y ) 
W 0 = ---  --------------------------------------------- - + V x ( c xx V x + c xy V y ) + V y ( c yx V x + c yy V y ) 
+ -------------------------------------------------
2  E ( I xx I yy – I xy 2 ) E ( I xx I yy – I xy 2 )

Now take partial derivatives of the work per unit length with respect to the transverse shear forces to get
∂W 0 ∂W 0
---------- = c xx V x + c xy V y ---------- = c yx V x + c yy V y
∂V x ∂V y

where we used the fact that c xy = c yx . From the material law (8.69), it can be seen that the partial derivative of
the work per unit length with respect to a shear force component equals the corresponding transverse shear strain
of the cross section; i.e.,
∂W ∂W
ψ x = ---------0- ψ y = ---------0- (8.70)
∂V x ∂V y
The work per unit length done by the bending moments and transverse shear forces is equal to the strain energy

212 Thin-Walled Structures


Transverse shear deformations – Timoshenko theory

per unit length stored in the beam. That is,

W0 = *
∫U
C
0 t ds (8.71)

where U 0* is the complementary strain energy density. Neglecting thermal effects, we have from eq. (8.37) that
the complementary strain energy per unit length is given by
σ 2 τ zs 2
U 0* t ds = ------z- + ------
- t ds

C

C
2E 2G

q(s)
The shear stress component tangent to the contour is related to the shear flow by τ zs = ---------- . Hence,
t(s)

σ2 q2
U 0* t ds = ------z- tds + --------- ds

C

C
2E 2Gt ∫
C

In the above expression only the shear flow is a function of the transverse shear forces, so that Castigliano’s sec-
ond theorem as given by eqs. (8.70) and (8.71) results in

∂U 0* q ∂q ∂U 0* q ∂q
ψx = ---------- t ds = ------ --------- ds ψy = ---------- t ds = --------- ds

C
∂V x ∫
C
Gt ∂V x ∫
C
∂V y ∫ -----
C
Gt
-
∂V y
(8.72)

From axial equilibrium of an infinitesimal element of the wall of the beam, and using the flexure formula for
the bending normal stress, we derived the shear flow formula which is given in eq. (6.13) on page 120. We write
this formula in a slightly different form from what is given in eq. (6.13) as

I xx Q y ( s ) – I xy Q x ( s ) I yy Q x ( s ) – I xy Q y ( s )
q ( s ) = q ( 0 ) – ----------------------------------------------
- V x – ----------------------------------------------
- Vy
det ( I ) det ( I )
For various cross-sectional contours not subject to torsion, the shear flow at the contour origin q ( 0 ) can be deter-
mined in terms of the transverse shear forces. In the absence of torque, we can write the shear flow equation as
q ( s ) = F x ( s )V x + F y ( s )V y (8.73)

For an open section having a free edge, the origin of the contour coordinate s can be taken at the free edge so that
q ( 0 ) = 0 . Then, for an open section

I xx Q y ( s ) – I xy Q x ( s ) I yy Q x ( s ) – I xy Q y ( s )
F x ( s ) = – ----------------------------------------------
2
- F y ( s ) = – ----------------------------------------------
2
- (8.74)
I xx I yy – I xy I xx I yy – I xy

If a stringer is located at the contour origin of a branch, then the shear flow q ( 0 ) is obtained in terms of shear
forces V x and/or V y as will be shown in Example 8.8. For a single cell cross section having a closed contour the
vanishing of the unit twist given by eq. (7.42) on page 163 determines the shear flow at the contour origin in
terms of the transverse shear forces. For a single cell
I xx Q y ( s ) – I xy Q x ( s ) I yy Q x ( s ) – I xy Q y ( s )
F x ( s ) = F x ( 0 ) – ----------------------------------------------
2
- F y ( s ) = F y ( 0 ) – ----------------------------------------------
2
- (8.75)
I xx I yy – I xy I xx I yy – I xy
where

Thin-Walled Structures 213


Article 8.8

1 1 I xx Q y ( s ) – I xy Q x ( s )
-  ----------------------------------------------
F x ( 0 ) = --------------- ----------------------------- - ds
ds
------ ( I xx I yy – I xy ) C
2  °∫ Gt 
(8.76)

Gt °∫
C

and

1 1 I yy Q x ( s ) – I xy Q y ( s )
-  ----------------------------------------------
F y ( 0 ) = --------------- ----------------------------- - ds
ds
------ ( I xx I yy – I xy ) C
2  °∫ Gt 
(8.77)

Gt °∫
C

Substitute eq. (8.73) into eqs (8.72) to get


[ F x ( s )V x + F y ( s )V y ] [ F x ( s )V x + F y ( s )V y ]
ψx = - [ F ( s ) ] ds ψy = - [ F ( s ) ] ds
∫ -------------------------------------------------
C
Gt x ∫ -------------------------------------------------
C
Gt y (8.78)

Expanding these results and comparing to Hooke’s law in eq. (8.69), we find the cross-sectional compliances in
transverse shear as

F x2 ( s ) F y2 ( s ) F x ( s )F y ( s )
c xx = ------------- ds c yy = ------------- ds c xy = c yx = -------------------------- ds

C
Gt ∫
C
Gt ∫
C
Gt
(8.79)

The transverse shear compliances for the beam 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 func-
tions F x ( s ) and F y ( s ) are 1/L. The dimensional units of the shear modulus are F/L2, so that the dimensional
units of the shear compliances are 1/F.

8.8 Application to bars in extension


EXAMPLE 8.1 Response of a stepped bar by Castigliano’s first 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 Castigliano’s first theorem.

( EA ) 1 ( EA ) 2

q 1, Q 1
z, w q 2, Q 2

L1 L2

Fig. 8.15 Stepped bar

214 Thin-Walled Structures


Application to bars in extension

Solution The deformation of the bar is a state of uniform strain in each section. For a spatially uniform strain,
the associated axial displacement function is linear in the axial coordinate. The axial displacement is specified to
vanish at z = 0. Kinematically admissible displacements are assumed in the form
z
w ( z ) = q 1 ----- 0 ≤ z ≤ L1
L1

and
z–L z–L
w ( z ) = q 1 1 –  -------------1- + q 2  -------------1- L1 ≤ z ≤ L1 + L2
 L2   L2 

Note that these displacement functions are continuous, w(0) = 0, w(L1) = q1 for both functions, and w(L1 + L2) =
q2. From the strain-displacement relation, these displacements yield the strains in each section as

q q2 – q1
( ε z ) 1 = ----1- ( ε z ) 2 = ---------------
-
L1 L2

The displacement is continuous at z = L1, but the strain is discontinuous at z = L 1 unless the external forces are
applied in such a manner that ( L 1 + L 2 )q 1 – L 2 q 2 = 0 . The strain energy is the sum of the strain energies in each
section as determined from eq. (8.41) with the thermal strain set equal to zero; i.e.,
L1 ( L1 + L2 )
1 1
U = --- ( EA ) 1 ( ε z ) 12 dz + --- ( EA ) 2 ( ε z ) 22 dz
2 ∫ 2 ∫
0 L1

Substituting for the strains in terms of the displacements we get

1 EA 1 EA
U = ---  ------- q 12 + ---  ------- ( q 2 – q 1 ) 2
2 L  1 2 L  2
Castigliano’s first theorem gives

∂U EA EA
Q1 = =  ------- q 1 +  ------- ( q 2 – q 1 ) ( – 1 )
∂ q1  L 1  L 2

∂U EA
Q2 = = 0 +  ------- ( q 2 – q 1 )
∂ q2  L 2

Writing these results in matrix form, we have

Q1 k1 + k2 –k2 q1 EA EA
= where k 1 =  ------- k 2 =  -------
Q2 –k2 k2 q2  L 1  L 2

The 2 x 2 matrix

( k1 + k2 ) –k2
(8.80)
–k2 k2

is the stiffness matrix of the structure. Note that the stiffness matrix is symmetric, consistent with eq. (8.15). A
symmetric stiffness matrix results for any linear elastic structural system undergoing small displacements.

Thin-Walled Structures 215


Article 8.8

EXAMPLE 8.2 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 Q 2 as shown in Fig.
8.16. There is no change in temperature from the stress free state, and the material is linear elastic. The material
has a specific weight denoted by γ which has dimensional units of F ⁄ L 3 . Determine the displacement at the tip
of the bar using Castigliano’s second theorem.

z
N
EA
g L dz
Fig. 8.16 Suspended bar under self γAdz
weight and a tip force

N + dN

q 2, Q 2

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. 8.16, the equilibrium differential equation is

dN
+ γA = 0 0<z<L
dz
The force prescribed at the tip gives the boundary condition for the axial force as
N ( L ) = Q2

Integrating the differential equation with respect to z from z to L gives


L L
dN
dz = N ( L ) – N ( z ) = – γA dz
∫ dz ∫
z z

and from the boundary condition prescribed on the axial force, we find the axial force is
N ( z ) = Q 2 + γA ( L – z )

The complementary strain energy only axial deformations and no thermal load is obtained from eq. (8.62) as
L
1 N2
U * = --- ------- dz
2 EA∫
0

The tip displacement is determined from Castigliano’s second theorem as

216 Thin-Walled Structures


Application to trusses

L L
∂ 1 ∂  N2  N  ∂N 
q2 = ( U * ) = --- ------- dz = dz
∂ Q2 ∫
2 ∂ Q 2  EA ∫ ------
-
EA  ∂ Q  2
0 0

Now substitute for the axial force this equation to get


L
( Q 2 + γA ( L – z ) )
q2 = ----------------------------------------- ( 1 ) dz
∫ EA
0

Thus, after integration, we find the tip displacement is


L γ L2
q 2 = -------Q 2 + ---------
EA 2E

This result for the displacement can be evaluated for Q 2 = 0 , 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 Castigliano’s theorem, then set the point force to zero.

8.9 Application to trusses


Trusses are skeletal structures 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 fifteen bars and eight joints is shown in Fig. 8.17.

4 8 12 16
2 4 6 8 3 7 11 15

2 6 10 14

1 3 5 7 1 5 9 13
joint numbering degree of freedom numbering

Fig. 8.17 A 15 bar truss

Each joint in a planar 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 dis-
placement is denoted by q 2i – 1 and the vertical displacement is denoted by q 2i . The positive directions for the
displacements and corresponding forces in the fifteen bar truss are defined as shown in Fig. 8.17. The original
coordinates of the joints and the sixteen displacements completely define the configuration of the truss in the
deformed state. Castigliano’s first theorem is particularly useful in the analysis of the static structural response of

Thin-Walled Structures 217


Article 8.9

trusses. The displacements q n and the corresponding forces Q n , n = 1, 2, …, 16 , used in the formulation of
Castigliano’s theorem are the displacements and forces at the joints, or nodes.

To use Castigliano’s first 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 L i and elongation in the deformed state is ∆ i .
Since the strain is uniform along the bar the axial strain of the i-th bar is

ε i = -----i (8.81)
Li
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 ε it . For no bending and recognizing w′ = ε i in eq. (8.41), the
strain energy in the i-th bar is
Li
1 2
Ui = --- ε i – ε it ε i dz
∫ ( EA ) i
2
(8.82)
0

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. (8.81) for the strain, and
denote the thermal force in the i-th bar as N it = ( EA ) i ε it , so that the strain energy becomes

1 EA
U i = ---  ------- ∆ i2 – N it ∆ i (8.83)
2 L  i

The strain energy of the assemblage is simply the sum of the strain energies in each bar; i.e.,

1  EA 2
U = ∑
all bars
--- ------- ∆ i – N it ∆ i
2 L  i
(8.84)

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, Castigliano’s first the-
orem for the truss shown in Fig. 8.17 gives
15
 EA  ∂∆ i 
Qn = ∑ ------- ∆ – N it   n = 1, 2, …, 16 (8.85)
 L i i  ∂ q n
i=1

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 Castigliano’s 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. 8.18. In the x-y plane of the truss, the coordinates of the beginning
node i are ( x i, y i ) , and the coordinates of the end node j are ( x j, y j ) . The square of the length of the undeformed
bar is determined by the pythagorean theorem as

L 2 = ( xj – xi ) 2 + ( yj – yi ) 2 (8.86)

218 Thin-Walled Structures


Application to trusses

j*
y
L*
q 2j
i* α+Ω
j
yj
q 2j – 1
L q 2i
i α
yi
q 2i – 1

0 xi xj x

Fig. 8.18 Displacement, rotation, and elongation of a truss bar

and the direction cosines of the undeformed bar are

xj – xi yj – yi
cos α = -------------
- cos ( 90 – α ) = sin α = -------------
- (8.87)
L L

As is depicted in Fig. 8.18, the bar displaces and rotates to a new position in the plane. The coordinates of
the beginning node in the deformed state are ( x i + q 2i – 1, y i + q 2i ) , and the coordinates of the end node in the
deformed state are ( x j + q 2j – 1, y j + q 2j ) . The counterclockwise rotation of the bar is denoted by the angle Ω .
Let L * denote the length of the bar in the deformed state. The square of the length of the bar in the deformed
configuration is

( L * ) 2 = [ ( x j + q 2j – 1 ) – ( x i + q 2i – 1 ) ] 2 + [ ( y j + q 2j ) – ( y i + q 2i ) ] 2 (8.88)

and the direction cosines of the bar in the deformed configuration are

( x j + q 2j – 1 ) – ( x i + q 2i – 1 )
cos ( α + Ω ) = -------------------------------------------------------------- (8.89)
L*
( y j + q 2j ) – ( y i + q 2i )
sin ( α + Ω ) = -------------------------------------------------
- (8.90)
L*
Define the relative elongation of the bar in the x-direction as
( q 2j – 1 – q 2i – 1 )
ε x ≡ ------------------------------------
- (8.91)
L
and the relative elongation of the bar in the y-direction as
( q 2j – q 2i )
ε y = -----------------------
- (8.92)
L
Rearranging eq. (8.88) and using the above definitions we get

( L * ) 2 = L 2 [ ( cos α + ε x ) 2 + ( sin α + ε y ) 2 ] (8.93)

The left-hand sides of eqs. (8.89) and (8.90) are expanded by the trigonometric identity for the sum of angles,

Thin-Walled Structures 219


Article 8.9

and the definitions given by eqs. (8.91) and (8.92) are used in the right-hand sides of eqs. (8.89) and (8.90) to get
L
cos α cos Ω – sin α sin Ω = -----*- ( cos α + ε x ) (8.94)
L
L
sin α cos Ω + cos α sin Ω = ----*- ( sin α + ε y ) (8.95)
L

If ∆ denotes the elongation of the bar, then L * = L + ∆ . The latter expression can be written as
L * = L ( 1 + ε ) where the strain of the bar is denoted by ε = ∆ ⁄ L . Expand the right-hand side of eq. (8.93), use
the trigonometric identity cos2 α + sin2 α = 1 , and let L * = L ( 1 + ε ) to get

( 1 + ε ) 2 = 1 + 2 ( cos α )ε x + 2 ( sin α )ε y + ε x2 + ε y2 (8.96)

Expand the left-hand side of this result, subtract one from each side, and the divide each side by two to get

1 1
ε  1 + --- ε = ( cos α )ε x + ( sin α )ε y + --- ( ε x2 + ε y2 ) (8.97)
 2  2

To find the sine of the rotation angle, multiply eq. (8.94) by – sin α , multiply eq. (8.95) cos α , and then add the
resulting equations. We obtain
1
sin Ω = ---------------- [ ( cos α )ε y – ( sin α )ε x ] (8.98)
(1 + ε)
Given the relative x-direction displacements and the relative y-direction displacements of the nodes, the strain of
the bar is determined from eq. (8.97) and the rotation of the bar is determined by eq. (8.98). For most trusses, the
strains and rotation of the bars are very small. For infinitesimal deformations, the following quantities are very
small
0≤ ε «1 0 ≤ εx « 1 0 ≤ εy « 1

Hence, eq. (8.97) yields the following approximation for the bar strain
ε ≈ ( cos α )ε x + ( sin α )ε y (8.99)

and eq. (8.98) yields the following approximation for the rotation
Ω ≈ ( cos α )ε y – ( sin α )ε x (8.100)

Substitute eqs. (8.91) and (8.92) for the relative elongations into eq. (8.99) to get

∆ q 2j – 1 – q 2i – 1 q 2j – q 2i
--- ≈ cos α  -------------------------------
- + sin α  ------------------
-
L  L   L 

Hence, the elongation is


( q 2j – q 2i )
α
∆ = ( cos α ) ( q 2j – 1 – q 2i – 1 ) + ( sin α ) ( q 2j – q 2i ) (8.101)
j α
For small strain and small rotations of the bar, eq. (8.101) shows that the elon- ( q 2j – 1 – q 2i – 1 )
gation is the sum of the projections of the relative displacements onto the refer- i
α
ence axis of the undeformed bar, as is shown in Fig. 8.19. Fig. 8.19

220 Thin-Walled Structures


Application to trusses

EXAMPLE 8.3 Three bar planar truss

The planar truss shown in Fig. 8.20 consists of three bars and one movable joint. Take the thermal strains to be
zero. Determine the 2 x 2 stiffness matrix using Castigliano’s first theorem.

q 2, Q 2

q 1, Q 1

 EA
-------  EA
-------  EA
-------
 L 1  L 2  L 3
α1 α2 α3

Fig. 8.20 Three bar truss

Solution The elongation of each bar as determined from eq. (8.101) is


∆ i = cos ( α i ) q 1 + sin ( α i ) q 2 i = 1, 2, 3 (8.102)

Equation (8.85) results from the application of Castigliano’s theorem, and applied to this example gives
3
EA
Q1 = ∑  ------L- [ cos ( α ) q
i
i 1 + sin ( α i ) q 2 ] [ cos ( α i ) ]
i=1

3
EA
Q2 = ∑  ------L- [ cos ( α ) q
i
i 1 + sin ( α i ) q 2 ] [ sin ( α i ) ]
i=1

These results are written in the matrix form

Q1 k 11 k 12 q 1
= (8.103)
Q2 k 21 k 22 q 2

where the elements of the stiffness matrix are


3 3
 EA
------- cos2 ( α i )
EA
k 11 = ∑  L i
k 22 = ∑  ------L- sin ( α )
i
2
i
i=1 i=1
(8.104)
3
 EA
------- cos ( α i ) sin ( α i )
k 12 = k 21 = ∑  L i
i=1

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 Q 1 and Q 2 , matrix eq. (8.103) is solved for the
nodal displacements q 1 and q 2 . From eq. (8.102) the elongation of each bar is then computed, and from these

Thin-Walled Structures 221


Article 8.9

elongations the bar forces are determined from

EA
N i =  ------- ∆ i i = 1, 2, 3 (8.105)
 L i

EXAMPLE 8.4 Three bar truss with lack of fit

Consider the same three bar truss of Example 8.3, but now assume that bar 1 was too short and had to be
stretched an amount ∆ 1 in order to connect to the joint. This is a case of lack of fit, and lack of fit is common in
the fabrication of structures. That is, before the external loads are applied ( Q 1 = Q 2 = 0 ), the truss bars experi-
ence initial forces due to the lack of fit of bar 1. Determine the initial forces in the bars using Castigliano’s first
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 fit plus the elongation experienced by bar 1 after the connection to the joint is made.
That is, the total elongation of bar 1 is ∆ 1 + ∆ 1 , where ∆ 1 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
3
1 EA 1 EA
U = ---  ------- ( ∆ 1 + ∆ 1 ) 2 + ∑ --2-  ------L- ∆ 2
i
2 L  1 i
i=2

The elongation of each bar after the connection is made is


∆ i = cos ( α i ) q 1 + sin ( α i ) q 2 i = 1, 2, 3

Use Castigliano’s first theorem to get


3
EA ∂∆ EA ∂∆ i
Q n =  ------- ( ∆ 1 + ∆ 1 ) 1 + ∑  ------L- ∆ ∂ q
i n = 1, 2
 L 1 ∂ qn i n
i=2

or
3
EA ∂∆ i EA  ∂∆ 
Qn = ∑  ------L- ∆ ∂ q
i
i
n
+  -------  1 ∆ 1
 L  1  ∂ q n
n = 1, 2
i=1

Substitute into this equation the relation between the nodal displacements and the elongation of each bar, eq.
(8.102), after the connection is made to get

Q1 k 11 k 12 q 1 EA cos ( α 1 )
= +  ------- ∆ 1
Q2 k 12 k 22 q2  L  1 sin ( α 1 )

where the elements of the stiffness matrix are the same as given in Example 8.3. Setting Q 1 = 0 and Q 2 = 0 ,
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

q1 1 k 22 – k 12 cos ( α 1 )  EA
= -------------------------------- ------- ∆ 1
– k 12 k 11 sin ( α 1 )  L  1
( k 11 k 22 – k 12 2 )
q2

222 Thin-Walled Structures


Application to beams

From this solution we can calculate the elongation of each bar after assembly from
∆ i = cos ( α i ) q 1 + sin ( α i ) q 2 i = 1, 2, 3

and finally calculate the initial bar forces from

EA EA EA
N 1 =  ------- ( ∆ 1 + ∆ 1 ) N 2 =  ------- ∆ 2 N 3 =  ------- ∆ 3
 L 1  L 2  L 3

8.10 Application to beams


EXAMPLE 8.5 Flexibility coefficients for a cantilever beam

Consider the cantilever beam shown Fig. 8.8 and let the length of the beam be denoted by L, with 0 ≤ z ≤ L , and
take point 1 at z = L ⁄ 2 . The cross section is rectangular with h 1 = – h ⁄ 2 and h 2 = h ⁄ 2 , where h denotes the
height of the cross section, and the second area moment I xx = th 3 ⁄ 12 . Neglect the complementary strain energy
due to transverse shear and determine the flexibility influence coefficients.

Solution. This is a statically determinate problem and equilibrium of the beam gives the internal stress result-
ants as
N = 0 0≤z≤L

 Q1 ( L ⁄ 2 – z ) + Q2 0≤z≤L⁄2
Mx = 
 Q2 L⁄2≤z≤L

Thus, equilibrium establishes the relationship of the external loads Q 1 and Q 2 to the internal actions N and
M x . From Castigliano’s second theorem, the generalized displacements corresponding to the loads are then
given by
L L
∂U * M x ∂M x ∂U * M x ∂M x
q 1 = ----------- = - ---------- dz
--------- q 2 = ----------- = - ---------- dz
---------
∂Q 1 ∫ EI xx ∂Q 1 ∂Q 2 ∫ EI xx ∂Q 2
0 0

The partial derivatives of the bending moment are

∂M x (L ⁄ 2 – z) 0≤z≤L⁄2
---------- = 
∂Q 1  0 L⁄2≤z≤L

∂M x
---------- = 1 0≤z≤L
∂Q 2

Hence, the generalized displacements at points 1 and 2 are defined by the definite integrals

Thin-Walled Structures 223


Article 8.10

L L
--- ---
2 2
1 1
q 1 = ---------- [ Q 1 ( L ⁄ 2 – z ) + Q 2 ] [ L ⁄ 2 – z ] dz + ---------- [ Q 1 ( L ⁄ 2 – z ) + Q 2 ] [ 0 ] dz
EI xx ∫ EI xx ∫
0 0

 --L- 
2 L 
1  
q2 = ---------
-  [ Q 1 ( L ⁄ 2 – z ) + Q 2 ] [ 1 ] dz + [ Q 2 ] [ 1 ] dz 
EI xx  ∫ ∫ 
0 L
--- 
 2 

Perform the integrations to get

L3 L2
q 1 = ----------------Q 1 + -------------Q 2
24EI xx 8EI xx

L2 L
q 2 = -------------Q 1 + ---------- Q 2 .
8EI xx EI xx
We can now identify the flexibility influence coefficients as

L3 L2 L
c 11 = ---------------- c 12 = c 21 = ------------- c 22 = ----------
24EI xx 8EI xx EI xx

EXAMPLE 8.6 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 Q 1 at its left end as is shown in Fig.
8.21. The material is homogeneous and linear elastic, the cross section is symmetric ( I xy = 0 ), and there are no
thermal strains. The bending stiffness is EI. Use Castigliano’s second theorem to determine the rotation at (a) the
left end, and (b) the right end. Neglect energy due to transverse shear deformation.

EI
Q 1, q 1 q2
z

Fig. 8.21 Simply supported beam subject to an end moment

Solution This beam is statically determinate. Moment equilibrium about


V
point z of the free body diagram of the left section of the beam gives
Q1 M
M = – Q 1 + zR (8.106)
z
R where R is the reaction force at the support.
Free body diagram of left section

Part (a) Moment equilibrium of the free body diagram of the entire beam
gives the support reaction to be

224 Thin-Walled Structures


Application to beams

Q L
R = -----1-
L
Q1
Hence the moment in eq. (8.106) becomes
R R
z
M = – Q 1  1 – --- (a) Free body diagram of entire beam
 L
It is important to note that we have determined a moment distribution from equilibrium that contains only the
applied moment Q 1 . That is, the support reaction R is not an independent force, it is a function of Q 1 . The com-
plementary strain energy is obtained from eq. (8.62), where the only contribution to the complementary strain
energy is due to bending. Castigliano’s second theorem in the case is
L L
M2
∂ -------- M  ∂M 
q1 = - dz = dz

∂ Q 1 2EI ∫ -----
-
EI  ∂ Q 
1
0 0

Substitute for the bending moment in this equation the function determined from equilibrium. We find
L L
1 z z Q z 2
q 1 = ------ Q 1  1 – ---  1 – --- dz = -----1-  1 – --- dz
EI ∫  L  L EI  ∫ L
0 0

Performing the definite integration on the right-hand side of the last equation we get
Q1 L
q 1 = ---------
- q 1 > 0 clockwise
3EI

Part (b) To find the rotation at the right end by Castigliano’s second theorem when no moment acts at that point,
we assume an external moment Q 2 acts at the right end, apply the theorem to compute the corresponding rota-
tion q 2 , then set Q 2 = 0 . That is,

L
M  ∂M 
q2 = dz
∫ -----
-
EI  ∂ Q 
2
(8.107)
0 Q2 = 0

Equation (8.106) for the moment is still valid in this case with two
L
moments applied to the beam. However, the support reaction R is different Q1 Q2
than in part (a). Moment equilibrium of the free body diagram for the entire
beam gives the support reaction as R R
( Q1 + Q2 ) (b) Free body diagram of entire beam
R = -----------------------
-
L
Substitute this result for the support reaction into eq. (8.106) to find the moment as
z
M = – Q 1 + ( Q 1 + Q 2 ) ---
L
Again, it is important to note that we have determined a bending moment distribution from equilibrium that only
contains the applied moment Q 1 and the fictitious external moment Q 2 . Substitute this bending moment into eq.

Thin-Walled Structures 225


Article 8.10

(8.107) to get
L
1 z z
q 2 = ------ – Q 1 + ( Q 1 + Q 2 ) ---  --- dz
EI ∫ L  L
0 Q2 = 0

Now we can set Q 2 = 0 and obtain

L
1 z z
q 2 = ------ – Q 1  1 – ---  --- dz
EI ∫  L  L
0

Performing the definite integration on the right-hand side of this result we find
Q 1 L
q 2 = –  ---------
- q 2 > 0 clockwise
 6EI 

Note that a clockwise moment Q 1 > 0 applied at the left end of the beam results in a counter clockwise rotation
at the right end q 2 < 0 .

EXAMPLE 8.7 Effect of transverse shear deformations on the tip deflection of cantilevered spar

Determine the percentage of the tip deflection due to transverse shear deformation of the uniform, cantilevered
spar subjected to a tip load F using Castigliano’s second theorem. See the sketch below. The spar is built-up from
a thin web of thickness t and two stringers of the same cross-sectional area A f . The modulus of elasticity is
70GPa and the shear modulus is G = 28GPa .

y, v y
F

Af
z h x
t
Af
L

Solution For no change in temperature and no axial deformations, the complementary strain energy given by eq.
(8.45) reduces to
L
M x2 V y2
U* = - dz
∫ ------------
- + ------------
2EI xx 2GA s
0

Castigliano’s second theorem for the tip displacement v L in the direction of the force F is v L = ∂U * ⁄ ∂F . Thus,

L
M x  ∂M x V y  ∂V y
vL = --------- - --------- dz
- ---------- + ---------
∫ EI xx  ∂F  GA s  ∂F 
0

Equilibrium determines the shear force and bending moment distributions in the beam as

226 Thin-Walled Structures


Application to beams

Vy = F M x = ( L – z )F 0≤z≤L

The partial derivative are ∂V y ⁄ ∂F = 1 and ∂M x ⁄ ∂F = ( L – x ) , and the tip deflection is


L L
F F
v L = ---------- ( L – z ) 2 dz + ---------- 1 dz
EI xx ∫ GA s ∫
0 0

or

FL 3 FL
v L = ------------- + ----------
3EI xx GA s






bending transverse shear

The percentage of the tip deflection due to transverse shear is then


FL
----------
GA s f
%shear = -----------------------------
3
- × 100 = ----------- × 100
FL FL 1 + f
------------- + ----------
3EI xx GA s
where

E I xx
f ≡ 3  ---- ----------
-
 G A L 2
s

The cross-sectional area and the second area moment are, respectively,
A = 2A f + ht

h2A h3t
I xx = ----------f + -------
2 12
To evaluate the cross-sectional area effective in transverse shear deformation given by eq. Af
(8.46), we need the distribution of the first area moment Q x ( y ) of a portion of the area about
h tds
---
the x-axis. A portion of the cross-sectional area from s = y to s = h ⁄ 2 is shown in the 2
s
accompanying sketch, so that y x
h centroid
---
2
h h h
Q x ( y ) = --- A f + s ( t ds ) – --- ≤ y ≤ ---
2 ∫ 2 2
y

Evaluating the integral we get

h t h2 h h
Q x ( y ) = --- A f + ---  ----- – y 2 – --- ≤ y ≤ ---
2 2 4  2 2
The following integral appears in the definition of the area effective in transverse shear

Thin-Walled Structures 227


Article 8.10

h h
--- ---
2 2
2 2
Q Q 1 h t 2 2
1 A f2 h 3 A f h 4 t h 5 t
------x dA = ------x t dy = --- --- A f + ---  h----- – y 2 dy = --- -----------
∫∫
a
 t ∫  t t ∫ 2 2 4  t 3
- + ------------ + ---------
12 120
h h
– --- – ---
2 2

We rearrange eq. (8.46) to solve for the area effective in shear to get
2
I xx 5ht ( 6A f + ht ) 2
A s = ---------------------------
2
- = -----------------------------------------------------------
-
Q x 6 ( 30A f2 + 10A f ht + h 2 t 2 )
------ dA
∫∫
 t
a

5
Note that A s = --- ( ht ) for A f = 0 , which is a well known result in structural mechanics for a solid rectangular
6
cross section. Now define
2A
a f ≡ --------f
A
such that A = Aa f + ht with 0 ≤ a f ≤ 1 . Solve for the web thickness to get t = ( A ⁄ h ) ( 1 – a f ) . For a f = 0
the cross-sectional area A = ht , the area of the web. For a f = 1 the web area is zero and the cross sectional
area is concentrated in the flanges of the two stringers with A f = A ⁄ 2 . For the beam to carry a finite bending
moment in the case of a f = 1 , we take h > 0 so that I xx > 0 , which implies the web thickness t = 0 . Eliminate
the web thickness in the geometric quantities that appear in the formula for the factor f to get

h 2 3 ( 2 + 6a f + 7a f2 )
f =  --- -------------------------------------------
-
 L 8 ( 1 – a f ) ( 1 + 2a f )

where E ⁄ G = 70 ⁄ 28 = 5 ⁄ 2 was used. The percentage of the total deflection due to transverse shear is given
the table below for various values of a f and the beam aspect ratio L ⁄ h . Increasing values of parameter a f from
zero correspond to decreasing values of the web thickness, with the web thickness vanishing at a f = 1 .

af = 0 a f = 0.25 a f = 0.5 a f = 0.75 a f = 0.90

L 42.9% 56.8% 71.1% 86.2% 94.6%


--- = 1
h
L 2.91% 4.99% 9.19% 20.0% 41.2%
--- = 5
h
L 0.74% 1.3% 2.47% 5.9% 14.9%
--- = 10
h
Note that the effect of transverse shearing deformations on the tip deflection is substantial for thin webs and
small aspect ratio beams. Even for an aspect ratio of ten, transverse shearing deformations of a built-up beam
with very thin webs account for 15% of the total tip deflection.

228 Thin-Walled Structures


Application to beams

EXAMPLE 8.8 Shear compliances of a stiffened blade section

Determine the transverse shear compliances from eqs. (8.79) for the blade
y
section of Example 8.7, which is repeated in Fig. 8.22. The section is made
of an isotropic, homogeneous material with a shear modulus G. Also, deter-
mine the influence of increasing the stringer area 2A f relative to the web Af
area ht on the shear stiffness. h h»t>0
---
2
C x
Solution The second area moments for this thin-walled section are
h
h 2 --- t
h3t
I xx = ------- + 2  --- A f I yy ≈ 0 I xy = 0 (8.108) 2
12  2 Af
Since the second area moment about the y-axis is zero in the thin wall
Fig. 8.22
approximation, this section can only resist bending in the y-z plane. Hence,
we consider beam loading such that V y ≠ 0 and V x = 0 . 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

h
x(s) = 0 y ( s ) = – --- + s 0≤s≤h
2
The first area moments of the portion of the cross section below s are
s
h h ht t
Q x ( s ) = – --- A s + y ( s )t ds = – --- A f – ----- s + --- s 2
2 ∫ 2 2 2
0

and
s

Qy ( s ) =
∫ x ( s )t ds = 0
0

From eqs. (8.74) the function F x ( s ) = 0 , and F y ( s ) = – Q x ( s ) ⁄ I xx . Hence, three of the transverse shear com-
pliances given in eqs. (8.79) vanish:
c xx = c xy = c yx = 0

and the only non-zero transverse shear compliance is


h
1 Qx ( s ) 2
c yy = ------ ds
Gt ∫ – -------------
I xx
0

The factor Gt is factored out of the integral in the equation for compliance c yy , since the section is homogeneous
and the thickness of the web is uniform. Substitute for the function Q x ( s ) in the expression for the compliance
c yy to get

Thin-Walled Structures 229


Article 8.10

h
1 h ht t 2
c yy = ------------
- – --- A f – ----- s + --- s 2 ds
GtI xx 2 ∫ 2 2 2
0

Performing the definite integral in this equation results in

1  h 3 A f2 h 4 tA f h 5 t 2
c yy = ------------
- ------------ + ------------ + ---------
2  4
(8.109)
GtI xx 12 120 

where the second area moment about the x-axis through the centroid I xx given by the first of eqs. (8.108). Equa-
tion (8.109) is the result sought. The only non-zero transverse shear compliance is c yy , so the material law for
shear deformation of the beam in the y-z plane is ψ y = c yy V y . We write the inverse of this material law as
V y = k yy ψ y , where k yy = 1 ⁄ c yy is the transverse shear stiffness.

As in Example 8.7, define the ratio of the cross-sectional area of the stringers to the total cross-sectional area
A by
α f = ( 2A f ) ⁄ A

For α f = 0 the stringers are absent ( A f = 0 ), and the web resists both bending normal stress and shear stress. In
the extreme case of α f = 1 the web thickness t = 0 , and the stringers carry all the bending load with the web
carrying only shear force V y . Introducing the definition of the stringer area and web area in terms of α f and the
total area A into the expressions for I xx and k yy we get after some algebra

h2A
I xx = --------- ( 1 + 2α f ) (8.110)
12
and

5 ( 1 – α f ) ( 1 + 2α f ) 2
k yy = --- GA -------------------------------------------
- (8.111)
6 1 + 3α f + 3.5α f2

When there are no stringers ( α f = 0 ), the shear stiffness of the blade section is 5GA ⁄ 6 . The shear stiffness of
the section decreases, or the section becomes more flexible in shear, as the stringer area increases as a larger por-
tion of the total area ( α increasing from zero, but remaining less than one). A plot of the transverse shear stiff-
ness versus α f is shown in Fig. 8.23.

230 Thin-Walled Structures


Application to beams

kyy êHGAL
2A
0.8 α = --------f
A
ht
0.6 ----- = 1 – α f
A

0.4

0.2

α
af
0.2 0.4 0.6 0.8 1

Fig. 8.23 Transverse shear stiffness of the blade section as the area of the stringers increases
relative to the web area.

EXAMPLE 8.9 Tip displacement of a cantilever wing spar a under distributed load

Determine the vertical tip displacement q 1 of a cantilever wing spar subjected to the distributed lift load of inten-
sity p y ( z ) = p 0 ( 1 – z ⁄ L ) , where p 0 is the intensity at the root and L is the span. See Fig. 8.25. The spar mate-

y, v
z
p y ( z ) = p o  1 – ---
 L q 1, Q 1

z
EI
L

Fig. 8.25 Cantilever wing spar under a distributed lift load

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 EI xx = EI is uniform along the span. Use Castigliano’s second theo-
rem, and neglect the complementary strain energy due to transverse shear deformation.

Solution Although there is no tip force acting on the beam, to use Castigliano’s second theorem we imagine a
fictitious force Q 1 corresponding to displacement q 1 to act on the spar. We determine the tip displacement via
the theorem, and then set Q 1 = 0 . The only contribution to the complementary energy in eq. (8.62) is from
bending. In mathematical terms the theorem results in

Thin-Walled Structures 231


Article 8.10

L
M x  ∂M x
q1 =  dz
∫ ------
-
EI  ∂ Q  1
(8.112)
0 Q1 = 0

where EI = EI xx . The bending moment distribution is obtained by solving the differential equations of equilib-
rium, eq. (2.1) on page 21 and eq. (2.3) on page 22, subject to the boundary conditions on the shear force and
bending moment at z = L . Substitute the distributed load intensity into (2.1) to get

dV z
---------y = – p 0  1 – ---
dz  L
Indefinite integration of this equation gives

z2
V y = – p 0  z – ------ + c 1
 2L

The boundary condition on the shear force is V y ( L ) = Q 1 , which allows for the determination of the constant of
integration; i.e.,
p0 L
c 1 = Q 1 + --------
-
2
Hence, the shear force is
p
V y ( z ) = -----0- ( z – L ) 2 + Q 1 (8.113)
2L
Substitute the shear force from eq. (8.113) into the equilibrium differential equation (2.3) for the moment to get
dM x p
---------- = -----0- ( z – L ) 2 + Q 1
dz 2L
Introduce a new independent variable by the definition
ζ≡z–L (8.114)

so that the differential equation for the bending moment becomes


dM x dζ p
---------- ------ = -----0-ζ 2 + Q 1
dζ dz 2L

From eq. (8.114) dζ ⁄ dz = 1 . Indefinite integration of the above equation gives


p
M x = -----0-ζ 3 + Q 1 ζ + c 2
6L

The boundary condition for the bending moment at z = L is M x = 0 . The value z = L corresponds to ζ = 0 .
This boundary condition determines that the constant of integration c 2 = 0 . Thus, the bending moment distribu-
tion in the equilibrium configuration under the distributed load and the fictitious force Q 1 is

p
M x = -----0- ( z – L ) 3 + Q 1 ( z – L ) (8.115)
6L

232 Thin-Walled Structures


Application to Frames

Substitute eq. (8.115) for the bending moment in eq. (8.112), recognize that the bending stiffness is indepen-
dent of z, to get
L
1 p
q 1 = ------ -----0- ( z – L ) 3 + Q 1 ( z – L ) [ z – L ] dz
EI ∫ 6L
0 Q1 = 0

Now set Q 1 = 0 , and change independent variables according to eq. (8.114), to write the expression for the tip
displacement as
0 0
p0 p 0  ζ 5
- ζ 4 dζ = ------------
q 1 = ------------ - -----
6EIL ∫ 6EIL  5 
–L
–L

Finally we get the result that

p0 L 4
q 1 = -----------
- (8.116)
30EI

8.11 Application to Frames


Frames are also skeletal structures composed of slender bars that can transmit axial, bending, and transverse
shear loads. The bars act as beams with a superimposed axial load. Joints in a frame are usually assumed rigid.
For example, in a plane frame the rotation of all bars connected to the joint are the same. Moments can be trans-
ferred through a rigid joint, but not a hinge joint, nor ball-and-socket joint. A frame structure may also contain
some hinge joints

Recall that bars in a truss structure carry only internal axial forces and the external loads are only applied at
the hinge joints. A truss structure may not be unstable if the members are not properly arranged; i.e., the structure
cannot sustain any load and it is a mechanism, or a moving mechanical system. A triangular arrangement three
bars forms a stable truss, and repeated triangular arrangements form a stable built-up truss. A rectangular
arrangement of four bars connected by smooth hinges is an unstable truss. If a fifth diagonal member is added to
the four bar structure it is a stable truss. A rectangular arrangement of four bars connected by rigid joints, or a
frame, is a stable structure capable of sustaining load. See Fig. 8.26.

unstable four bar truss stable truss four bar frame with rigid joints (stable)

Fig. 8.26 Unstable and stable skeletal structures

Thin-Walled Structures 233


Article 8.11

EXAMPLE 8.10 Strut-braced spar

Consider Example 8.9 again, but with a strut attached between the fuselage (simulated by a wall) and tip of the
spar as shown in Fig. 8.27. 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
y, v
z
p y ( z ) = p o  1 – ---
 L q 1, Q 1

z
α Fig. 8.27 Strut-braced spar
EA, EI

( EA ) b

q 2, Q 2
L

tip of the spar using Castigliano’s second theorem.

Solution The structural assembly is statically indeterminate. For


y example, the overall free body diagram of the assembly has three
z
p y ( z ) = p o  1 – ---
 L unknown forces of support R y, R z, Q 2 and a moment C x as
Ry
shown in the adjacent sketch. For a coplanar force system there
Rz are three independent equations of equilibrium. Hence, we have
z
one more unknown force than independent equations of equilib-
Cx rium. This structural system is said to have one redundant force.
To proceed in using Castigliano’s second theorem, remove the
strut from the fuselage support and replace the action of the sup-
Q2
port on the strut with force Q 2 . Force Q 2 is selected as the redun-
dant in this approach. In addition to the applied fictitious force
Assembly free body diagram
Q 1 , we regard the redundant force Q 2 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 U * in terms of the
independent forces Q 1 and Q 2 . (Note: support reactions R y, R z, and C x are not independent, since the three
equilibrium equations obtained from the free body diagram of the assembly relate them to forces Q 1 and Q 2 .)
Then, we apply Castigliano’s second theorem to determine the corresponding displacements q 1 and q 2 . 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 fictitious force Q 1 is set to zero after performing the differentiations in Castigliano’s
theorem. Hence, the mathematical statements of Castigliano’s method to this structural assembly are

∂U *
q 1 = ---------- (8.117)
∂Q 1
Q1 = 0

234 Thin-Walled Structures


Application to Frames

∂U *
q 2 = ---------- = 0 (8.118)
∂Q 2
Q1 = 0

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
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 N denote the axial force in the spar and N b denote
the axial force in the brace strut, both positive in tension. Then the complementary strain energy is
L
 N2 M x2  N b2  L 
U* = - + --------- dz + -----------------
- ------------
∫  ----------
2EA 2EI 2 ( EA ) b  cos α








0







spar strut (8.119)

Equilibrium Free body diagrams of the joints between the strut and
Q1
fuselage and strut and spar tip are shown in the adjacent figure. Equilib-
rium at the joint between the strut and fuselage yields Vy
Nb = Q2 (8.120) N α
Mx
Equilibrium at the joint between the spar and strut results in the follow- Nb
ing boundary conditions at z = L in the spar. z
L
– N ( L ) – N b cos α = 0 (8.121)
Nb
– V y ( L ) + Q 1 – N b sin α = 0 (8.122)
Q2
My ( L ) = 0 (8.123)
Free body diagrams of the joints
The differential equations of equilibrium for the spar are
dN
------- = 0 0<z<L (8.124)
dz
dV z
---------y + p 0  1 – --- = 0 0<z<L (8.125)
dz  L

dM y
---------- – V y = 0 0<z<L (8.126)
dz
The solution to eq. (8.124) subject to boundary condition (8.121) is
N ( z ) = – Q 2 cos α (8.127)

in which we used eq. (8.120) to write the axial force in the strut in terms of Q 2 . The solution to eqs. (8.125) and
(8.126) subject to boundary conditions (8.122) and (8.123) is the same procedure explained in Example 8.9.
Omitting the details, the solution for the bending moment in the spar is (refer to eq. (8.115))

Thin-Walled Structures 235


Article 8.11

p
M x ( z ) = -----0- ( z – L ) 3 + ( Q 1 – Q 2 sin α ) ( z – L ) (8.128)
6L
Equations (8.120), (8.127), and (8.128) represent the equilibrium solutions in terms of the known distributed load
and the independent forces Q 1 and Q 2 . This step is crucial to using Castigliano’s theorem.

Evaluation of displacement equations Substitute the complementary strain energy, eq. (8.119), into eq.
(8.117) to get the following expression for displacement q 1 .

L
N ∂N M ∂M N b  L   ∂N b
q1 = -------  --------- + -------x  ----------x dz + -------------- ------------ ---------
∫ EA ∂Q 1   EI ∂Q 1   ( EA ) b  cos α  ∂Q 1
0 Q1 = 0

From eq. (8.127) ∂N ⁄ ∂Q 1 = 0 , from eq. (8.128) ∂M x ⁄ ∂Q 1 = z – L , and from eq. (8.120) ∂N b ⁄ ∂Q 1 = 0 .
Thus, the equation for displacement q 1 reduces to

L
1 p
q 1 = ------ -----0- ( z – L ) 3 + ( – Q 2 sin α ) ( z – L ) [ z – L ] dz
EI ∫ 6L
0

in which we used eq. (8.128) again to eliminate M x . Performing the definite integral on the right-hand-side
Q1 = 0
of the previous equation we get

p 0 L 4 ( Q 2 sin α )L 3
q 1 = -----------
- – ----------------------------- (8.129)
30EI 3EI
The first term on the left-hand side of this equation is the tip displacement obtained in Example 8.9 when there
was no strut. The second term on the left-hand-side of eq. (8.129) represents the reduction the tip displacement
due to the presence of the strut.

Next substitute the complementary strain energy, eq. (8.119), into eq. (8.118) to get the following expression
for displacement q 2 .

L
N ∂N M ∂M N b  L   ∂N b
q2 = -------  --------- + -------x  ----------x dz + -------------- ------------ ---------
∫ EA  ∂Q 2 EI  ∂Q 2 ( EA ) b  cos α  ∂Q 2
0 Q1 = 0

From eq. (8.127) ∂N ⁄ ∂Q 2 = – cos α , from eq. (8.128) ∂M x ⁄ ∂Q 2 = – ( z – L ) sin α , and from eq. (8.120)
∂N b ⁄ ∂Q 2 = 1 . Use these same equations, (8.127), (8.128), and (8.120), evaluated for Q 1 = 0 , to replace N ,
M x , and N b , respectively, in the equation for displacement q 2 . The resulting expression for q 2 is

L
 ( – Q 2 cos α ) 1 p0 
q2 = ( – cos α ) + ------------ ( z – L ) 3 + ( – Q 2 sin α ) ( z – L ) [ – ( z – L ) sin α ] dz +
∫  ---------------------------
EA EI 6L 
0

Q2  L 
-------------- ------------ ( 1 )
( EA ) b  cos α

236 Thin-Walled Structures


Application to Frames

Evaluating some of the terms in this expression we get


L L
Q 2 L cos2 α p o sin α Q 2 sin2 α Q2 L
- – ----------------- ( z – L ) 4 dz + --------------------
q 2 = ------------------------ ( z – L ) 2 dz + ---------------------------
EA 6EIL ∫ EI ∫ ( EA ) b cos α
0 0

Perform the definite integrations in this equation, rearrange the terms, and recall that q 2 = 0 to get

L cos2 α L 3 sin2 α L p0 L 4
Q 2 ------------------ + ------------------- + --------------------------- – -----------
- sin α = 0 (8.130)
EA 3EI ( EA ) b cos α 30EI

We can solve eq. (8.130) for the redundant force Q 2 in terms of the distributed load intensity p 0 . Then, in
turn substitute the result for Q 2 into eq. (8.129) to find displacement q 1 .

EXAMPLE 8.11 A frame of two slender members

The frame shown Fig. 8.28 in consists of two slender bars AB and BC that lie in the same plane and are con-
nected perpendicular to each other at point B by a rigid joint. The length of bar AB is denoted by a and the length
of bar BC is denoted by b. End C of bar BC is fixed to a rigid support, and bar AB is subjected to a uniformly dis-
tributed, downward load of intensity p. Use Castigliano’s second theorem to determine the downward tip dis-
placement q at point A due to the distributed load. Neglect deformations due to transverse shear.

C C
y2
z2
b B
B
a y1 x2
x1
A z1
A
p

q, Q

Fig. 8.28 Two bar frame subject to a distributed load

Solution A fictitious force Q is applied at point A corresponding to the requested displacement. From the second
theorem we have

∂U *
q = ----------
∂Q Q=0

where U * is the total complementary strain energy of the frame. Bar AB is in bending and bar BC is combined
bending and torsion due to the downward directed loads. Both the distributed load and the point force act in the
negative y-direction as shown in Fig. 8.28. Neglecting transverse shear deformation and using the coordinate sys-
tems shown in Fig. 8.28, the total complementary strain energy is

Thin-Walled Structures 237


Article 8.12

a b
M x1 2 M x2 2 T 22
U* = - dz +
---------- - dz
∫ 2EI 1 1 ∫ ----------
- + ------------
2EI 2 2GJ 2 2
0 0

where M x1 is the bending moment in bar AB, M x2 is the bending moment in bar BC, and T 2 is the torque in bar
BC. From Castigliano’s second theorem we have
a b

q = M x1  ∂M x1
- ------------ dz +
-------- M x2  ∂M x2 T 2   ∂T 2
- ------------ +  --------
-------- - -------- dz 2
∫  EI 1   ∂Q  1 ∫  EI 2   ∂Q   GJ 2  ∂Q 
0 0 Q=0

The bending moment distribution in bar AB is determined from equilibrium of the


y1 V y1 free body diagram as shown in the accompanying sketch. The result is
pz 1
M x1 z
Q M x1 = pz 1  ---1- + Qz 1 0 < z1 < a
 2
z1 ⁄ 2
z1

V y2 The bending moment and torque in bar BC are determined from equilibrium of the free
T2 body diagram shown in the accompanying sketch. The results are
z2 y2
M x2 M x2 = ( pa + Q )z z 0 ≤ z2 ≤ b
x2
and

pa a
T 2 = aQ + pa  --- 0 ≤ z2 ≤ b
Q a⁄2 a  2

Substitute the bending moments and torque into the displacement equation to get
a b b
1 pz 2 1 1 pa 2
q = --------  -------1- ( z 1 ) dz 1 + -------- ( paz 2 ) ( z 2 ) dz 2 + ---------  -------- ( a ) dz 2
EI 1  2  ∫ EI 2 ∫ GJ 2  2 ∫
0 0 0

Performing the definite integrations, we get

pa 4 pab 3 pa 3 b
q = ----------- + ------------ + -------------
8EI 1 3EI 2 2GJ 2
The contributions to the displacement of point A are shown by the terms on the right-hand side in this result; the
first term is from bending of bar AB, the second term is from bending of bar BC, and the third term is from tor-
sion of bar BC.

8.12 Reference
Y.C. Fung, 1965, Foundations of Solid Mechanics, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, pp. 1-7.

238 Thin-Walled Structures


Problems

8.13 Problems
1. Each bar in the truss shown has a cross-sectional
2 100 in. 1
area of 1.0 in.2, and a modulus of elasticity of 107 psi.
There is no change in temperature. Use Castigliano’s 45°
first theorem to find 60°

a) the horizontal and vertical displacements of


3 10,000 lb.
node 1, 45°
y
b) the stress in psi in each bar, and
4 x
c) the horizontal and vertical support reactions
at node 5. 30°

5
2. The bars in the truss shown have the following
cross-sectional areas: A 1 = 1.0 in. 2 , 1
A 2 = A 4 = 2.0 in. 2 , A 3 = 1 ⁄ 2 in. 2 ,
A 5 = A 6 = 3 ⁄ 2 in. 2 . The modulus of elasticity of 2 3 4
40 in.
each bar is 107 psi. Compute the vertical displacement
of the right-hand joint using Castigliano’s second theo-
rem. Note this truss is statically determinate and all bar 5 6
forces can be determined in terms of external load Q.
30 in. 30 in.
3. Use Castigliano’s’ second theorem to compute the
horizontal displacement of the right-hand joint of the Q = 30, 000 lb.
previous problem.

4. The plane truss shown represents a single bay of a


–6
wing spar truss. For all bars: E = 75 GPa and α = 23.0 ×10 ⁄ °C . The cross-sectional areas of the bars are:
2580 mm2 for the horizontal bars, 387 mm2 for the vertical bars, and 2690 mm2 for the diagonal bars. The upper
horizontal bar is heated to 250°C 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.

45kN
250°C
4 2

810 mm

3 1
1080 mm 45kN
Use Castigliano’s first theorem to find
a) stiffness matrix in kN/mm,

Thin-Walled Structures 239


Article 8.13

b) displacement of all joints in mm,


c) all boundary reactions in kN, and
d) the stresses in MPa in each bar.
z2 5.A coplanar bar is subjected to an end force Q 1 as
L
shown. The bar is uniform with axial stiffness EA and
bending stiffness EI. Use Castigliano’s second theorem
z1 L to find

q3 a) the end rotation q 2 , and


3
b) the vertical displacement q 3 at the joint.
q2 4

Q1

6. Consider the statically indeterminate, uniform beam shown below that is subjected to a uniform, downward
distributed load of intensity p. For small displacements assume that only the complementary strain energy is
bending is significant. If the center support moves downward by the amount 0.01pL 4 ⁄ EI xx and remains attached
to the beam, use Castigliano’s second theorem to find the reactions at the left and right supports.

y, v p

z, w
L L
--- ---
2 2

7. The frame consists of three slender, uniform bars of length L, and two right angle bends. Assume the bends
are rigid joints. Each member has a solid circular cross section of diameter d. A force P acts in the global X-direc-
tion at point A. Find the three displacement components u A, v A, w A of point A in terms of P, L, d, and E using
Castigliano’s second theorem. Assume G = 0.4E . Neglect deformations due to transverse shear deformations.
Y, v

D L
C
Z, w X, u
L
B

A P

240 Thin-Walled Structures


CHAPTER 9 Criteria for initial
yielding

The yield strength of a material is determined from the tensile test as discussed in Article 3.2 page 45. The yield
strength is an important material property in the design of structures made of ductile materials, so ductility is dis-
cussed in this chapter. Also, the tensile test is conducted under axial loading such that the state of stress in the
gage length of the specimen is uniaxial tension. So the questions arises as to how to use the yield strength deter-
mined in simple laboratory 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.

9.1 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. 9.1. A measure of ductility is the engineer-
σ σ

fracture
fracture

0 ε 0 ε
ductile material brittle material
Fig. 9.1 Ductile and brittle material responses

ing strain at fracture, which is usually report in percent: i.e, 100ε f . A second, widely reported measure of ductil-
ity is the percent reduction in area, denoted as %RA, that is defined by

Thin-Walled Structures 241


Article 9.2

A i – A f d f2 – d i2
%RA = 100  ---------------
- = 100  ----------------
-
 Ai   d2 
i

where A i is the initial cross-sectional area, A f is the minimum cross-sectional area at fracture (occurring in the
necked-down region), d i , and d f 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 Institute. 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, fillets, etc.) and/or
material discontinuities. Geometric discontinuities are commonly referred to as notches. The capacity to redis-
tribute stresses at stress raiser 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.

9.2 Criteria for yielding of ductile materials


The yield strength σ y is determined from the tensile test data, but the tensile test is designed to produce a uniax-
ial 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. 9.2, 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 defined and discussed. Mises cri-

242 Thin-Walled Structures


Criteria for yielding of ductile materials

n
σn
τ ns
τ nz
τ sn

τ zn σs
σ z τ zs τ zn
σz σ = τ sz σ s τ sn
s
P τ nz τ ns σ n
τ sz

τ zs

z
Fig. 9.2 Three dimensional state of stress at a point in the material.

terion and the maximum shear-stress criterion are presented in Article 9.6 and Article 9.7, respectively.

Consider an infinitesimal rectangular parallelepiped at a material point, denoted by P, in a thin-walled bar.


The element is referenced to the three mutually perpendicular directions z, s, and n, where the z-direction is par-
allel to the longitudinal axis of the bar, the s-direction is tangent to the contour in the cross section, 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 these coordinate directions. These
nine stress components can be written in the matrix form

σ z τ zs τ zn
σ = τ sz σ s τ sn (9.1)

τ nz τ ns σ n

in which the first 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 find 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. 9.3. The principal stresses are denoted by σ 1, σ 2, and σ 3 . The determi-
nation of the principal stresses is discussed Article 9.4. 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 find three mutually perpendicular
directions at the point P, labeled x 1, x 2, and x 3 , 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
τ zs = τ sz . Likewise, moment equilibrium about the z-axis and the s-axis at P leads to τ sn = τ ns and τ nz = τ zn ,
respectively. In matrix notation moment equilibrium at a point implies symmetry of the stress matrix, which is
written as

Thin-Walled Structures 243


Article 9.2

x2
n
σn
τ ns
τ nz σ2
τ sn

τ zn σs

σz P
s
P x1
τ sz
σ1
τ zs
σ3
z x3

σ z τ zs τ zn σ1 0 0
σ = τ sz σ s τ sn [ σp ] = 0 σ2 0
τ nz τ ns σ n 0 0 σ3

Fig. 9.3 Equivalent descriptions of the state of stress at a material point in the body.

σ z τ zs τ zn σ z τ sz τ nz
[ σ ] = [ σ ] T or τ sz σ s τ sn = τ zs σ s τ ns (9.2)

τ nz τ ns σ n τ zn τ sn σ n

where the superscript T denotes matrix transpose. Thus, there only six independent stress components at a point.

The mean normal stress, denoted by σ m , is defined by

1 1
σ m = --- ( σ x + σ y + σ z ) = --- Tr [ σ ] (9.3)
3 3
where Tr [ σ ] denotes the trace of the stress matrix. The trace of a square matrix is defined as the sum of its diag-
onal elements. We will show in Article 9.3, eq. (9.15), 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
1
σ m = --- ( σ 1 + σ 2 + σ 3 ) (9.4)
3
In the case of fluids at rest the shear stress components vanish, and each normal stress component is equal to the
pressure p. The stress matrix for any three mutually perpendicular axes at a point in the fluid at rest is

–p 0 0
[σ] = 0 –p 0 (9.5)

0 0 –p

244 Thin-Walled Structures


Stress transformation equations for generalized plane stress

The hydrostatic stress, denoted by σ h is defined as the negative of the mean normal stress; σ h = – σ m . Hence,
for the fluid at rest under pressure p we have σ h = p . (See Example 9.2).

9.3 Stress transformation equations for generalized plane stress


For thin-walled structures we neglect the shear stress components in the thickness direction, τ zn and τ sn with
respect to the shear stress component tangent to the contour τ zs . The stress matrix in the z, s, n system reduces to
what is called generalized plane stress. That is,

σ z τ zs 0
σ = τ sz σ s 0 (9.6)

0 0 σn

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 find
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 z′, s′, n′ which is rotated about the n-axis through
the angle θ , defined positive counter-clockwise looking down the positive n-axis . The n′-axis and n-axis coin-
cide in this transformation of coordinates. The face normal to the z′ -axis is the oblique face of an infinitesimal
wedge at point P. Looking down the positive n-axis, the wedge appears as a triangle as is shown in Fig. 9.4. You

σz s′
σ z dA z τ sz dA s

τ zs θ
P s′ s′
τ sz s
θ θ θ
σs τ zs ′ τ zs dA z σ s dA s
θ θ
σz ′
θ z′ z′
z′ z-face components s-face components
z

Fig. 9.4 Wedge element as seen looking down the positive n-axis

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 σ z ′, τ zs ′ , respectively. Also, the positive stress components acting on the negative z-
and negative s-faces are shown on the wedge element in the figure. If the area of the wedge face is denoted by dA,
then from the geometry of Fig. 9.4, the area of the z-face dA z = dA cos θ , and the area on the s-face
dA s = dA sin θ . 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 z′-direction gives

σ z ′dA – ( σ z dA z ) cos θ – ( τ zs dA z ) sin θ – ( σ s dA s ) sin θ – ( τ sz dA s ) cos θ = 0

Thin-Walled Structures 245


Article 9.3

Divide this equation by dA, take the limit as the wedge shrinks down to point P, and recognize that
dA z ⁄ dA = cos θ , dA s ⁄ dA = sin θ , and τ zs = τ sz , to get

σ z ′ = σ z cos2 θ + σ s sin2 θ + τ zs 2 sin θ cos θ (9.7)

Force equilibrium in the s′ – direction on the wedge element gives

τ zs ′dA + ( σ z dA z ) sin θ – ( τ zs dA z ) cos θ – ( σ s dA s ) cos θ + ( τ sz dA s ) sin θ = 0

Again divide by dA and let the wedge element shrink to point P in the limit to get

τ zs ′ = – σ z cos θ sin θ + σ s sin θ cos θ + τ zs ( cos2 θ – sin2 θ ) (9.8)

The wedge element used to determine the stress components on the s′-face is shown in Fig. 9.5. For the element

σs ′ s′
τ sz
θ
σs
τ sz ′ θ
P s Fig. 9.5 Wedge element with an oblique face
τ zs normal to the s’-axis
σz

θ
z′
z

in Fig. 9.5, the areas are related by dA z = dA sin θ and dA s = dA cos θ . Following procedures similar to those
resulting in eq. (9.7), force equilibrium of this wedge element in the s′-direction leads to

σ s ′ = σ z sin2 θ + σ s cos2 θ – τ zs 2 sin θ cos θ (9.9)

It is also possible to obtain eq. (9.9) from eq. (9.7) by letting θ → θ + 90° and σ z ′ → σ s ′ , since this would rotate
the z′-axis to the original direction of the s′-axis . Equilibrium of the wedge element in Fig. 9.5 in the
z′-direction leads to

τ sz ′ = – σ z sin θ cos θ + σ s cos θ sin θ + τ zs ( cos2 θ – sin2 θ ) (9.10)

which when compared to eq. (9.8) shows that τ zs ′ = τ sz ′ , as would be expected for moment equilibrium about
the n-axis at point P. The stress transformation eqs. (9.7) to (9.10) are written in matrix form as

σ z ′ τ zs ′ 0 cos θ sin θ 0 σ z τ zs 0 cos θ – sin θ 0


τ sz ′ σ s ′ 0 = – sin θ cos θ 0 τ sz σ s 0 sin θ cos θ 0 (9.11)

0 0 σn ′ 0 0 1 0 0 σn 0 0 1

or in the more compact form as


T
σ′ = T σ T (9.12)

246 Thin-Walled Structures


Principal stresses and maximum shear stress

in which [ T ] is the matrix of direction cosines as given in the table below. Since the n′-axis and the n-axis

z s n
z′ cos θ sin θ 0

s′ – sin θ cos θ 0

n′ 0 0 1

coincide in the rotation of ( z′, s′, n′ ) relative to ( z, s, n ) , the normal stress σ n ′ = σ n . The trace of stress matrix
in the coordinates ( z′, s′, n′ ) is given by

σ z ′ τ zs ′ 0
Tr τ sz ′ σ s ′ 0 = σz ′ + σs ′ + σn ′ (9.13)

0 0 σn ′

Using eqs. (9.7), (9.9), and σ n ′ = σ n , we find

σz ′ + σs ′ + σn ′ = σz + σs + σn (9.14)

In matrix notation this is written as

σ z ′ τ zs ′ 0 σ z τ zs 0
Tr τ sz ′ σ s ′ 0 = Tr τ sz σ s 0 (9.15)

0 0 σn ′ 0 0 σn

In other words, the trace of the stress matrix is invariant with respect to a coordinate rotation.

9.4 Principal stresses and maximum shear stress


Mohr’s circle The graphical representation of the stress transformation equations is called the Mohr’s circle.
Mohr’s 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 first use the
trigonometric identities
1 1
cos2 θ = --- ( 1 + cos 2θ ) sin2 θ = --- ( 1 – cos 2θ ) 2 sin θ cos θ = sin 2θ
2 2
to write eqs. (9.7) to (9.9) as
1 1
σ z ′ = --- ( σ z + σ s ) + --- ( σ z – σ s ) cos 2θ + τ zs sin 2θ (9.16)
2 2

1
τ zs ′ = – --- ( σ z – σ s ) sin 2θ + τ zs cos 2θ (9.17)
2

Thin-Walled Structures 247


Article 9.4

1 1
σ s ′ = --- ( σ z + σ s ) – --- ( σ z – σ s ) cos 2θ – τ sz sin 2θ (9.18)
2 2
Define the average stress σ ave and radius R by

2
1 1 2
σ ave = --- ( σ z + σ s ) R = --- ( σ z – σ s ) + τ zs (9.19)
2 2
Using these definitions we write eqs. (9.16) and (9.17) as
1
--- ( σ z – σ s )
2 τ zs
σ z ′ = σ ave + R ------------------------- cos 2θ + -----
- sin 2θ
R R

1
--- ( σ z – σ s )
2 τ zs
τ zs ′ = R – ------------------------- sin 2θ + -----
- cos 2θ
R R

Define an angle ψ via the trigonometric relations 1


--- ( σ z – σ s )
2
1
--- ( σ z – σ s )
2 τ zs ψ
cos ψ = ------------------------- sin ψ = -----
- (9.20)
R R
τ zs
R
which is depicted in the triangle at right. Using this definition of ψ we get

σ z ′ = σ ave + R [ [ cos ψ cos 2θ + sin ψ sin 2θ ] ]

τ zs ′ = R [ – cos ψ sin 2θ + sin ψ cos 2θ ]

Using the trigonometric identities


cos ( ψ – 2θ ) = cos ψ cos 2θ + sin ψ sin 2θ

sin ( ψ – 2θ ) = sin ψ cos 2θ – cos ψ sin 2θ


we finally get the form
σ z ′ = σ ave + R cos ( ψ – 2θ ) (9.21)

τ zs ′ = R sin ( ψ – 2θ ) (9.22)

Similar manipulations applied to eq. (9.18) results in


σ s ′ = σ ave – R cos ( ψ – 2θ ) (9.23)

The center of the circle is at ( σ, τ ) = ( σ ave, 0 ) and its radius is R. A positive shear stress on the positive z-face,
or τ zs , is defined positive downward on the ordinate. A positive shear stress on a positive s-face, or τ sz , is
defined positive upward on the ordinate. The circle is shown in Fig. 9.6. 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. 9.6 assuming the stresses are all
positive in value. (Set θ = 0 in eqs. (9.21) and (9.22) 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.

248 Thin-Walled Structures


Principal stresses and maximum shear stress

τ sz
s – face
τ sz

s′-face
τ sz ′ σ ave
σ2 σ1
0 σz σ z’
σ
σs ′ σs

τ zs ψ
z ’ – face
R 2θ
τ zs
z – face

τ zs
Fig. 9.6 Mohr’s circle

Hence, these diametrically opposite points are represent principal stresses σ 1 and σ 2 at point P. Principal stress
σ 3 = σ n since no shear stresses act on the n-faces . Hence, the principal stresses are given by

σ 1 = σ ave + R σ 2 = σ ave – R (9.24)

The rotation about the n-axis to the principal stress directions occurs when 2θ = ψ . From eqs. (9.20) we get

τ zs
tan 2θ p = ------------------------
- (9.25)
1
--- ( σ z – σ s )
2
where the rotation angle to the principal direction is denoted θ p . The definition for plotting positive shear in the
Mohr’s 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 Mohr’s circle is twice the rotation in the physical
plane. If the angle θ p is substituted for θ in the matrix transformation eq. (9.11) we get

σ z ′ τ zs ′ 0 σ1 0 0
τ sz ′ σ s ′ 0 = 0 σ2 0 θ → θp (9.26)

0 0 σn ′ 0 0 σ3

As stated above, the third principal stress σ 3 = σ n ′ = σ n .

Thin-Walled Structures 249


Article 9.4

EXAMPLE 9.1 Maximum shear stress in tensile test

Draw Mohr’s circle and determine the maximum shear stress in the uniaxial tension test; i.e, the stress matrix is
given by

σ z τ zs 0 σz 0 0
[ σ ] = τ sz σ s 0 = 0 00 (9.27)

0 0 σn 0 00

Solution First note that the principal stresses are σ 1 = σ z , σ 2 = 0 , and σ 3 = 0 .The average stress
σ ave = σ z ⁄ 2 and radius R = σ z ⁄ 2 . Mohr’s circle is shown in Fig. 9.7 The z-, s-, and n-directions are principal

s σz ⁄ 2 s′
τ sz s′-face σz ⁄ 2
σz ⁄ 2 σz ⁄ 2
σz σz
0 σz ⁄ 2 σ P
45° σz ⁄ 2
σz z
P
90°
σz ⁄ 2
principal stress element z′
τ zs z′-face
maximum shear stress element

Fig. 9.7 Mohr’s circle and stress elements in the tension test

stress directions. The maximum shear stress is τ max = σ z ⁄ 2 , and the maximum shear stress element is rotated
45° clockwise about the n-axis as shown in Fig. 9.7. Note that the normal stresses on the planes where the max-
imum shear stress occurs are non-zero and equal to σ z ⁄ 2 as well.

EXAMPLE 9.2 Mohr’s circle for hydrostatic stress state

Draw Mohr’s circle for the hydrostatic stress state given by

σ z τ zs 0 –p 0 0
[ σ ] = τ sz σ s 0 = 0 –p 0
0 0 σn 0 0 –p

where p is the pressure.


τ sz
Solution The average stress σ ave = – p and the radius R = 0 . Hence,
Mohr’s circle reduces to a point on the negative normal stress axis as shown –p σ
0
in Fig. 9.8. The shear stress is zero and the normal stress is compressive
with magnitude p on every plane through point P. The principal stresses are
σ1 = σ2 = σ3 = –p .
Fig. 9.8 τ zs

250 Thin-Walled Structures


Principal stresses and maximum shear stress

EXAMPLE 9.3 Principal stresses and maximum shear stress

Determine the principal stresses and the maximum shear stress for the state of stress given by the matrix

σ z τ zs 0 60 – 12 0
[ σ ] = τ sz σ s 0 = – 12 15 0 MPa
0 0 σn 0 0 0

Sketch the principal stress element and the maximum shear stress element.

Solution The average stress is given by σ ave = ( 60 + 15 ) ⁄ 2 = 37.5MPa , and the radius is

2
1
R = --- ( 60 – 15 ) + 12 2 = 25.5MPa
2

Mohr’s circle is shown in Fig. 9.9. The principal stresses are σ 1 = 63MPa , σ 2 = 12MPa , and σ 3 = 0 . The

stresses in MPa s x2
15 12
14.04°
z-face 12
– 12
R σ
0 60
15 37.5 60 12 63
z n, x 3
n
x1
R = 25.5
τ zs principal stress element

Fig. 9.9 Mohr’s circle and the principal stress element

rotation to the principal stress element is


– 12
tan 2θ p = -------------------------- = – 0.53333333 θ p = – 14.04°
1
--- ( 60 – 15 )
2
The principal stress element is sketched in Fig. 9.9.

The maximum shear stress on the oblique plane of a wedge element obtained by a rotation about the n-axis,
or x3-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 Mohr’s circle for a rotation about the x 2 -axis ,
which would have principal stresses σ 1 and σ 3 defining its diameter. This second Mohr’s circle is centered at
( σ 1 + σ 3 ) ⁄ 2 with a radius ( σ 1 – σ 3 ) ⁄ 2 . The locus of points on this second Mohr’s circle represents all possible
combinations of the normal and shear stresses on wedge elements with oblique faces normal to the x 1 ′-axis ,
where the x 1 ′-axis is rotated from the x 1 -axis counter-clockwise as view down the positive x 2 -axis . See Fig.
9.10. The maximum shear stress on this second Mohr’s circle is simply its radius, and it occurs for counter-clock-

Thin-Walled Structures 251


Article 9.4

stresses in MPa x2
12
τ x 1 ′-face 14.04°
principal stress element

z-face 63
n, x 3
x 3 -face x 2 -face 90° x 1 -face x1
0 63 σ
12 31.5
x1 ′

63 σ 1 ′ = 31.5
45°

x 2, x 2 ′ 12 τ 13 ′ = 31.5

τ x1
45° maximum shear stress
on a wedge element
x3 x3 ′

Fig. 9.10 Mohr’s circle associated with the maximum shear stress

wise rotation of 90° from the x 1 face stress location on the circle. This rotation corresponds to a counter-clock-
wise rotation of 45° in the physical plane as shown in Fig. 9.10. Equilibrium of the wedge element with an
oblique face normal to the x 1 ′-axis determines the shear and normal stresses on the wedge face, which from
Mohr’s 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. 9.10. For convenience, take the area of the oblique face to be unity.
Then the area on the negative x 1 -face is 1 × cos 45° . Note that the negative x 3 face on the wedge element is
stress free. Summing forces in the x 1 ′-direction to zero we get

σ 1 ′ × 1 – [ 63 × 1 × cos 45° ] × cos 45° = 0

Equating forces to zero in the x 3 ′-direction we get

τ 13 ′ × 1 – [ 63 × 1 × cos 45° ] × sin 45° = 0

Hence, the stresses on the x 1 ′-face are

σ 1 ′ = 31.5MPa τ 13 ′ = 31.5MPa (9.28)

The maximum shear stress for the stress state is 31.5MPa.

252 Thin-Walled Structures


Octahedral shear stress

9.5 Octahedral shear stress


Since the hydrostatic stress σ h is observed not to effect yielding, we find the plane at the material point for which
the normal stress is σ h , and then assume that the shear stress acting on this plane, denoted by τ h , is the stress
component responsible for initiating yielding. We start the analysis to find the shear stress τ h with the principal
stress element at point P, and consider equilibrium of an infinitesimal tetrahedron at P. The direction of the unit
normal to the oblique face of the tetrahedron, denoted by n , is unknown initially, but we can determine it from
(n)
the condition that the normal stress acting on the oblique face is the hydrostatic stress. Let S denote the stress
vector acting on the oblique face, and let dA denote the area of the oblique face of the tetrahedron. See Fig.
9.11. The stress vectors acting on the three triangular faces normal to the principal stress directions are
x2
x2

σ 3 ( –ˆi 3 )dA 3

σ 1 ( –ˆi 1 )dA 1
dx 2
– dx 3 ˆi 3 + dx 2 ˆi 2
P dx
1
dx 3 P
x1
– dx 3 ˆi 3 + dx 1 ˆi 1 x1

σ 2 ( –ˆi 2 )dA 2
x3 nˆ
(n) x3 2dAnˆ
S dA

Fig. 9.11 Tetrahedron element at point P.

σ 1 dA 1 ( –ˆi 1 ) , σ 2 dA 2 ( –ˆi 2 ) , and σ 3 dA 3 ( –ˆi 3 ) , where dA 1 , dA 2 , and dA 3 are the areas of the triangular faces of
the tetrahedron normal to the unit vectors ˆi , ˆi , and ˆi along the principal directions, respectively. Force equi-
1 2 3
librium of the tetrahedron gives the vector equation
(n)
S dA – σ 1 dA 1 ˆi 1 – σ 2 dA 2 ˆi 2 – σ 3 dA 3 ˆi 3 = 0 (9.29)

Divide this equation by the area of the oblique face dA and shrink the tetrahedron to point P in the limit to get

(n) dA dA dA
S = σ 1 --------1- ˆi 1 + σ 2 --------2- ˆi 2 + σ 3 --------3- ˆi 3 (9.30)
dA dA dA

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

( – dx 3 ˆi 3 + dx 1 ˆi 1 ) and ( – dx 3 ˆi 3 + dx 2 i 2 ) (9.31)

as are shown in Fig. 9.11. 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

Thin-Walled Structures 253


Article 9.5

two position vectors. Using the above two edge vectors we have by this interpretation

2dAnˆ = ( – dx 3 ˆi 3 + dx 1 ˆi 1 ) × ( – dx 3 ˆi 3 + dx 2 i 2 ) (9.32)

Expanding the cross product in this equation gives

2dAnˆ = ( – dx 3 dx 2 ) ( ˆi 3 × ˆi 2 ) – dx 1 dx 3 ( ˆi 1 × ˆi 3 ) + dx 1 dx 2 ( ˆi 1 × ˆi 2 )

2dAnˆ = dx 2 dx 3 ˆi 1 + dx 1 dx 3 ˆi 2 + dx 1 dx 2 ˆi 3

The areas of the triangular faces of the tetrahedron normal to the principal directions are dA 1 = ( dx 2 dx 3 ) ⁄ 2 ,
dA 2 = ( dx 1 dx 3 ) ⁄ 2 , and dA 3 = ( dx 1 dx 2 ) ⁄ 2 . Hence, the above equation reduces to

dA dA dA
nˆ = --------1- ˆi 1 + --------2- ˆi 2 + --------3- ˆi 3 (9.33)
dA dA dA
In general, a unit vector is represented as

nˆ = n 1 ˆi 1 + n 2 ˆi 2 + n 3 ˆi 3 where n 12 + n 22 + n 32 = 1 (9.34)

Component n 1 is the cosine of the angle between the normal and the positive x 1 -axis with similar interpretations
for components n 2 and n 3 . Comparing eqs. (9.33) and (9.34), it is clear that the direction cosines of the unit nor-
mal also relate the areas of the faces of the tetrahedron as

dA dA dA
n 1 = --------1- n 2 = --------2- n 3 = --------3- (9.35)
dA dA dA
Substitute the direction cosines of the unit normal from eq. (9.35) into eq. (9.30). Thus, we find the stress vector
on the oblique face of the tetrahedron is related to the principal stresses by the vector relation
(n)
S = σ 1 n 1 ˆi 1 + σ 2 n 2 ˆi 2 + σ 3 n 3 ˆi 3 (9.36)

The normal stress component on the oblique face of the tetrahedron is given by the scalar product
(n)
σ nn = nˆ • S . Using eqs. (9.34) and (9.36) this scalar product is

σ nn = σ 1 n 12 + σ 2 n 22 + σ 3 n 32 (9.37)

We need to find the direction of the oblique face, i.e., to find direction cosines n 1, n 2, and n 3 , such that the nor-
mal stress on this face is equal to the hydrostatic stress, σ h . From eq. (9.4) the hydrostatic stress is

1 1 1
σ h = σ 1 --- + σ 2 --- + σ 3 --- (9.38)
3 3 3
Setting σ nn = σ h and substituting eqs. (9.38) into (9.37), determines the square of the direction cosines as

1 1 1
n 12 = --- n 22 = --- n 32 = --- (9.39)
3 3 3
The result for the direction cosines in this equation implies that there are actually eight oblique faces at the point

254 Thin-Walled Structures


Octahedral shear stress

P on which the hydrostatic acts. Unit normal vectors to these eight oblique faces are given the table below. These

Components
Unit normal
vectors n1 n2 n3
n1 1⁄ 3 1⁄ 3 1⁄ 3
n2 –1 ⁄ 3 1⁄ 3 1⁄ 3
n3 –1 ⁄ 3 –1 ⁄ 3 1⁄ 3
n4 1⁄ 3 –1 ⁄ 3 1⁄ 3
n5 1⁄ 3 1⁄ 3 –1 ⁄ 3
n6 –1 ⁄ 3 1⁄ 3 –1 ⁄ 3
n7 –1 ⁄ 3 –1 ⁄ 3 –1 ⁄ 3
n8 1⁄ 3 1⁄ 3 –1 ⁄ 3

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 find 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 find its value, consider the octahedral plane in the first octant of the cartesian space defined by the principal
directions as shown in Fig. 9.12 The stress vector can be written as

x3
1
σ h nˆ nˆ = nˆ 1 = ------- ( ˆi 1 + ˆi 2 + ˆi 3 )
3

(n)
S
P x2

τ h ˆt
x1
Fig. 9.12 Normal stress and shear stress components acting on an octahedral plane.

(n)
S = σ h nˆ + τ h ˆt (9.40)

where ˆt is a unit vector tangent to the octahedral plane and τ h is the octahedral shear stress. The scalar product
of the stress vector with itself yields the square of its magnitude. Using eq. (9.36) we have
(n) (n)
S •S = ( σ1 n1 ) 2 + ( σ2 n2 ) 2 + ( σ3 n3 ) 2 (9.41)

While from eq. (9.40) we have

Thin-Walled Structures 255


Article 9.6

(n) (n)
S •S = σ h2 + τ h2 (9.42)

Equations (9.41) and (9.42) are equated to one another, eq. (9.39) is substituted for the square of the direction
cosines, and then the resulting equation is solved for the octahedral shear stress to get
1 1 1
τ h2 = σ 12 --- + σ 22 --- + σ 32 --- – σ h2 (9.43)
3 3 3
Now substitute eq. (9.38) for the hydrostatic stress in this equation to get

1 1 1 1 1 1 2
τ h2 = σ 12 --- + σ 22 --- + σ 32 --- –  σ 1 --- + σ 2 --- + σ 3 --- , or
3 3 3  3 3 3

2
τ h2 = --- ( σ 12 + σ 22 + σ 32 – σ 1 σ 2 – σ 1 σ 3 – σ 2 σ 3 )
9
This last expression can be written as
1
τ h2 = --- [ ( σ 1 – σ 2 ) 2 + ( σ 2 – σ 3 ) 2 + ( σ 3 – σ 1 ) 2 ]
9
Hence the octahedral shear stress magnitude is
1
τ h = --- [ ( σ 1 – σ 2 ) 2 + ( σ 2 – σ 3 ) 2 + ( σ 3 – σ 1 ) 2 ] (9.44)
3

9.6 Mises criterion for initiation of yielding


Mises Criterion
Yielding begins in a three-dimensional stress state when the octahedral shear stress τ h is
equal to its value at yield initiation in the uniaxial tension test

In the uniaxial tension test the principal stresses at the initiation of yielding are σ 1 = σ y , σ 2 = 0 σ 3 = 0 .
(Refer to Example 9.1). Substitute these principal stress values into eq. (9.44) to get

2
τ h = ------- σ y (9.45)
3
For the three-dimensional state of stress, we substitute eq. (9.45) for the octahedral shear stress in eq. (9.44) to
get

1
--- [ ( σ 1 – σ 2 ) 2 + ( σ 2 – σ 3 ) 2 + ( σ 3 – σ 1 ) 2 ] = σ y (9.46)
2
Equation (9.44) 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 σ 1 , σ 2 , and σ 3 as is
shown in Fig. 9.13 In this space the yield surface given by eq. (9.46) is a right-circular cylinder of radius
2 ⁄ 3 σ y whose axis makes equal angles with respect to the σ 1 , σ 2 , and σ 3 coordinate axes. Points within the
cylinder correspond to stress states which have not initiated yielding, while points outside the cylinder corre-

256 Thin-Walled Structures


Mises criterion for initiation of yielding

axis of cylinder
σ 1 s1
----- 0
σy 1
2

2
s3σ 3 --- σ y radius
----- 3
σy 1

0
1
σ
-----2 2
σy s2

Fig. 9.13 Mises yield surface in principal stress space

spond to stress states that are beyond yielding. Notice that for any stress state on the axis of the cylinder, for
which σ 1 = σ 2 = σ 3 , no yield is predicted no matter the magnitude of the stresses! This reflects the assumption
that hydrostatic stress does not initiate yielding.

In design, it is convenient to define the Mises stress, denoted σ M , by the definition

1
σM = --- [ ( σ 1 – σ 2 ) 2 + ( σ 2 – σ 3 ) 2 + ( σ 3 – σ 1 ) 2 ] (9.47)
2
Then, the constraint that yielding not occur is
σM < σy (9.48)

Also, it is convenient to define the factor of safety, denoted as FS, with respect to yielding as
σ
FS = ------y- > 1 for no yielding (9.49)
σM

Thin-Walled Structures 257


Article 9.7

For the most general state of stress given by

σ z τ zs τ zn
σ = τ sz σ s τ sn
τ nz τ ns σ n

It can be shown after much algebra (which was done using Mathematica) that the Mises stress can be written in
the equivalent form

σM = σ z2 + σ s2 + σ n2 – σ z σ s – σ z σ n – σ s σ n + 3τ zs
2 + 3τ 2 + 3τ 2
zn sn (9.50)

This equation represents a simplification in computing the Mises stress, since we can omit the step of computing
the principal stresses. For the case of generalized plane stress where τ zn = τ sn = 0 , eq. (9.50) reduces to

σM = σ z2 + σ s2 + σ n2 – σ z σ s – σ z σ n – σ s σ n + 3τ zs
2 (9.51)

9.7 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.,
τ max = ( σ max – σ min ) ⁄ 2 (9.52)

where
σ max = Max ( σ 1, σ 2, σ 3 ) σ min = Min ( σ 1, σ 2, σ 3 ) (9.53)

From Example 9.1, the maximum shearing stress in the tensile test at the initiation of yield is
τ max = σ y ⁄ 2 (9.54)

For a three-dimensional state of stress, we substitute eq. (9.52) for the maximum shear stress in eq. (9.50) to get
σ max – σ min = σ y (9.55)

Equation (9.53) 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 σ 3 = 0 is shown in Fig. 9.14. 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.
9.14.

In the first quadrant of Fig. 9.14, both σ 1 and σ 2 are positive. If σ 1 > σ 2 , then
Mohr’s circles appear as shown in the figure to the right, and the maximum shear
σ3 σ2 σ1
stress is σ 1 ⁄ 2 . Therefore, maximum shear-stress criterion plots as a vertical line
0
with σ 1 = σ y for 0 < σ 2 < σ y . (Note that stresses normalized by σ y are plotted in

258 Thin-Walled Structures


Maximum shear-stress criterion

σ2 ⁄ σy

1
Mises criterion

0.5
Maximum
shear-stress criterion

σ1
-----
-1 -0.5 0.5 1 σy

-0.5

σ3 = 0

-1

Fig. 9.14 Criteria for yield initiation in the σ 1 – σ 2 principal stress plane.

Fig. 9.14.) If σ 2 > σ 1 > 0 , then the role of σ 1 and σ 2 interchange on the Mohr’s circles, so the maximum shear
stress is now σ 2 ⁄ 2 . Therefore, maximum shear-stress criterion plots as the horizontal line σ 2 = σ y for
0 < σ1 < σy .

In the second quadrant of Fig. 9.14 σ 1 < 0 and σ 2 > 0 , so that Mohr’s circles
appear as shown in the figure to the right. The maximum shear stress is
σ1 σ3 σ2
( σ 2 – σ 1 ) ⁄ 2 . Therefore, the maximum shear-stress criterion, eq. (9.55), gives
0
σ 2 = σ y + σ 1 in the second quadrant. This is a straight line with a one-to-one slope
intersecting the σ 2 -axis at σ y . Plotting the maximum shear-stress criterion in quad-
rants three and four in Fig. 9.14 proceed in a similar manner.

The maximum shear-stress envelope in Fig. 9.14 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. (9.50), but the maximum shear stress criterion
requires that we compute the principal stresses and find their numerical order. Also note that the maximum shear
stress acts on four planes at the material point, refer to Fig. 9.10, 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. 9.14, 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

Thin-Walled Structures 259


Article 9.7

τ max < σ y ⁄ 2 (9.56)

and we can define a factor of safety against yielding as


σy ⁄ 2
- > 1 for no yielding
FS = ----------- (9.57)
τ max

EXAMPLE 9.4 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. The state of stress is that given in Exam-
ple 9.3.

Solution From Example 9.3, the principal stresses are σ 1 = 63MPa , σ 2 = 12MPa , and σ 3 = 0 , and the
maximum shear stress is 31.5MPa . If we calculate the Mises stress from eq. (9.47), we get

1
σM = --- ( ( 63 – 12 ) 2 + ( 12 – 0 ) 2 + ( 0 – 63 ) 2 ) = 57.94MPa
2
If we calculate the Mises stress via eq. (9.51), we get

σM = 60 2 + 15 2 – 60 × 15 – 0 – 0 + 3 × ( – 12 ) 2 = 57.94MPa

which is the same value as obtained from eq. (9.47). Hence the factor of safety against yield, eq. (9.49), is
325
FS = ------------------- = 5.61
57.9396

The factor of safety against yield using the maximum shear stress criterion, eq. (9.57), is
325 ⁄ 2
FS = ---------------- = 5.16
31.5
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%).

260 Thin-Walled Structures


Maximum shear-stress criterion

EXAMPLE 9.5 Stress responses of a stringer-stiffened, single cell beam.

The thin-walled, prismatic beam of length L shown in Fig. 9.15 is clamped at z = 0. It is subjected to a linearly
distributed load p y ( z ) = p 0 ( 1 – z ⁄ L ) 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 t 1 . The vertical branch, or second branch, has length h and thickness t 2 . The beam is stiff-
py
T
py ( z ) Af t1 y

h s1
T --- s2
2 S.C. C
z x
h t2
---
2 t1
s3
L b
Af

Cross-section

Fig. 9.15 A stringer-stiffened, cantilevered beam. The contour in the cross section is an isosceles
triangle.

ened by two longitudinal stringers of cross-sectional area A f . The overall dimensions are specified 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.
Aluminum alloy 2024-T4

Property Value Units


E, modulus of elasticity 73 GPa
ν, Poisson’s ratio 0.33 none
G, Shear modulus 27 GPa
ρ, density 2800 kg/m3
σyield, yield strength in tension 325 MPa

Take p 0 = 3.0 N/mm , T = 750 N-m , t 1 = t 2 = 1.0 mm , and


A f = 45 mm 2 . For the cross section at z = 0, τ zs s

a. plot the axial normal stress σ z in MPa versus the contour coordinate s in σz s = 0

mm, and
b. the shear stress τ zs in MPa versus the contour coordinate s in mm.
The contour coordinate is related to the branch contour coordinates by

Thin-Walled Structures 261


Article 9.7

 s1 0 ≤ s1 ≤ b

s =  s2 + b 0 ≤ s2 ≤ h

 s3 + b + h 0 ≤ s3 ≤ b

Equations for the stress analysis


y
Geometry
Af q1
hs s1
y 1 ( s 1 ) = --- ---1- 0 ≤ s1 ≤ b
2b s2 C
h x
h q2 q3
y 2 ( s 2 ) = --- – s 2 0 ≤ s2 ≤ h contour origin s = 0
2 s3
h hs Af
y 3 ( s 3 ) = – --- + --- ---3- 0 ≤ s3 ≤ b
2 2b ( a) c

b h b
h 2 h 2
I xx = y 12 ( s 1 )t 1 ds 1 + y 22 ( s 2 )t 2 ds 2 + y 32 ( s 3 )t 1 ds 3 +  --- A s +  – --- A s
∫ ∫ ∫  2  2
( b)
0 0 0

Axial normal stress due to bending

Mx ( z )
σ z ( z, s ) = -------------
-y ( s ) ( c)
I xx

Shear stress tangent to the contour

q ( z, s )
τ zs ( z, s ) = --------------- q ( z, s ) = q b ( z, s ) + q t ( z ) ( d)
t(s)
Shear flow due to torque

T 1
q t = --------- A c = enclosed area of the cell = --- hc ( e)
2A c 2
Shear flow due to transverse the shear force

s1
Vy ( z )
- y 1 ( s 1 )t 1 ds 1
q b1 ( z, s 1 ) = q 0 – ------------
I xx ∫ ( f)
0

s2
V y ( z )  h Vy ( z )
q b2 ( z, s 2 ) = q b1 ( b ) – ------------
- --- A f – ------------
- y 2 ( s 2 )t 2 ds 2
I xx 2   I xx ∫




0
upper stringer ( g)

262 Thin-Walled Structures


Maximum shear-stress criterion

s3
V y ( z )  h Vy ( z )
q b3 ( z, s 3 ) = q b2 ( h ) – ------------- – --- A f - y 3 ( s 3 )t 1 ds 3
– ------------
I xx  2 I xx ∫








0
lower stringer ( h)

To determine the shear flow at the contour origin, q 0 , due to the transverse shear force V y acting through the
shear center, we invoke the condition of no twist. That is,

dθ z q ( s ) ds q(s)
= ---------------- = 0 → ---------- ds = 0
dz °∫
Gt ( s ) t(s) °∫ ( i)

or

b h b
q b1 ( s 1 ) q b2 ( s 2 ) q b3 ( s 3 )
----------------- ds 1 + ----------------
- ds 2 + ----------------
- ds 3 = 0
∫ t1 ∫ t2 ∫ t1
( j)
0 0 0

Equilibrium py
Vy
L
 dV
---------y = – p y ( z ) and V y ( L ) = 0 → V y ( z ) =
 dz  ∫ p ( z ) dz
y ( k)
Mx
z

L
 dM
----------x = V y ( z ) and M x ( L ) = 0 → M x ( z ) = – V y ( z ) dz
 dz  ∫ ( l)
z

To locate the centroid (the point labeled C) y


Af
A = area = 2bt 1 + ht 2 + 2A f ( m)
s1
b b s2 C
h x
Qy = x 1 ( s 1 )t 1 ds 1 + x 3 ( s 3 )t 1 ds 3 xc = Qy ⁄ A xc
∫ ∫ ( n)
s3
0 0
Af
x1 ( s1 ) = c ( 1 – s1 ⁄ b ) x3 ( s3 ) = c ( s3 ⁄ b ) ( o) c
To locate the shear center (the point labeled S.C.)

From the solution of equations (f) to (j), we can find the shear flow in branch two due to the shear force V y
only. Now we use torque equivalence to locate the S.C. Sum torques about the apex to get

– c q b2 ( s 2 ) ds 2 = ( c – x sc )V y
∫ ( p)
0

Thin-Walled Structures 263


Article 9.7

Equation (p) yields a relation to find x sc , or the location of the shear center. The location of the shear center is
independent of the magnitude of the shear force.
y y
Vy
s2
S.C.
h
q b2 ( s 2 )
x ⇔ x
x sc

c – x sc
c c

Results The shear force and bending moment attain maximum magnitudes at the root. At the root cross section
6
V y ( 0 ) = 2700 N and M x ( 0 ) = ( – 1.62 ×10 ) N-mm . The torque at the root section is
3
T ( 0 ) = 750 ×10 N-mm . Note that 1 N/mm2 equals 1 MPa
a) The axial normal stress is plotted as a function of the contour coordinate in graph below.

ê mm2
z, MPa
s zσ,N
150

100

50

s,mm
50 100 150 200 250 300 350
-50

-100

-150

b) The shear stress is plotted as a function of the contour coordinate in the following graph.

264 Thin-Walled Structures


Maximum shear-stress criterion

, MPa
t τzszs,N ê mm2
80
70
60
50
40
30
20
10
s,mm
50 100 150 200 250 300 350

EXAMPLE 9.6 Minimum weight design of the beam in Example 9.5 subject to a constraint on initial
yielding

Consider the design for minimum weight of the aluminum alloy beam in Example 9.5, which is shown in Fig.
9.15. The design is constrained by material yielding with the factor of safety specified as 1.5. Use the material
data for 2024-T4 aluminium as listed in the Example 9.5 problem statement, Mises yield criterion, and take the
value of the local acceleration due to gravity as 9.81 m/s2. The specified dimensions of the beam are
L = 1800 mm , h = 100 mm , and b = 130 mm . Take the value of the applied loads as p 0 = 3 N/mm and
3
T = – 250 ×10 N-mm . (Note the value of the torque is changed with respect to its value in Example 9.5.)

The objective is to minimize the weight subject to no yielding given the loads p 0 and T . That is, what are
the thicknesses t 1 , t 2 and the stringers’ cross-sectional area A f for minimum weight? Parameters t 1 , t 2 and A f
are called design variables. This is a problem in constrained optimization, which is stated mathematically as
minimize W ( t 1, t 2, A f )
such that g i ( t 1, t 2, A f ) > 0

where W ( t 1, t 2, A f ) is the objective function, or weight in this problem, and g i ( t 1, t 2, A f ) are constraint func-
tions. For design against yielding the constraint functions are defined as
σ all – ( σ M )
g i = ----------------------------i
σ all

Thin-Walled Structures 265


Article 9.7

point 2, s 1 = b where the allowable stress, σ all , is defined as

point 3, s 2 = 0 point 1, s 1 = 0 σ all = σ yield ⁄ F.S. and the Mises stress is σ M . These par-
y ticular constraint functions are called static margins, and
point 4, s 2 = h ⁄ 2 x positive values indicate the degree of safety against exceed-
C
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. 9.16 Critical points for yield evaluation Fig. 9.16. That is, compute the margins of safety at the
in the cross section at the root points labeled 1, 2, 3, and 4 in Fig. 9.16. 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 influence of the stringers on the design for minimum weight. For
each stringer area given in the table below, determine the values of thicknesses t1 and t2 for minimum weight.
List these values along with the weight, and the four margins of safety in the table.

Influence of the stringer area on the minimum weight designs


Stringer Beam Thicknesses in mm Margins of safety, dimensionless
area Af in weight in
mm2 N t1 t2 point 1 point 2 point 2 point 4

50
60
70
80
90
100

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 t1 = 2.84 mm, t2 = 3.42
mm, and A f = 45 mm 2 .

266 Thin-Walled Structures


Maximum shear-stress criterion

Results

0.8

t 2 , mm
Minimum static margin = 0
0.7

constant weight lines


0.6
feasible designs

0.5 least weight design

18 N
0.4

infeasible designs

0.3 16 N
14 N

t 1 , mm
0.2 0.3 0.4 0.5

Design plane for A s = 100 mm 2

A f, mm 2 Weight, N t 1, mm t 2, mm g1 g2 g3 g4
50 14.40 0.437253 0.774719 3.2 – 10 0.0029 1.1
9 ×10
60 13.57 0.342651 0.653471 2.1 – 12 0.008 0.82
– 2 ×10
70 13.31 0.279341 0.564789 1.4 0.001 –8 0.57
– 6 ×10
80 13.49 0.239829 0.505728 0.97 0.0024 – 11 0.41
– 2 ×10
90 13.97 0.214902 0.466853 0.72 0.0034 – 12 0.3
3 ×10
100 14.62 0.198474 0.440723 0.57 0.0039 – 10 0.23
– 1 ×10

Thin-Walled Structures 267


Article 9.8

9.8 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.

268 Thin-Walled Structures


CHAPTER 10 Matrix structural
analysis

To introduce the basic methods of matrix structural analysis, the analyses of structures modeled with linear elas-
tic springs is presented in Article 10.1 to Article 10.6. The objective is to illustrate the steps in the direct stiffness
method, which is summarized in Article 10.7. Then, the matrix structural analysis of planar trusses is presented
in Article 10.8 and Article 10.9. Matrix structural analysis for structures modeled with beam elements is pre-
sented in Article 10.10 and Article 10.11. Finally, the matrix structural analysis of frame structures is presented
in Article 10.12. The approach followed here is based on chapters 2, 3, 4, and 6 of Martin (1966).

10.1 Flexibility and stiffness matrices


The following example continues with the discussion presented in “Unit action states” on page 199 and in
“Unit displacement states” on page 200.

EXAMPLE 10.1 Two springs in series restrained at one end.

Construct the flexibility influence matrix C by the method of unit action states (UAS), and the stiffness influ-

ence matrix K , by the method of unit displacement states (UDS) for the two-degree-of freedom structural
model shown in Fig. 10.1. The model consists of two linear elastic springs in series with the left end fixed
against translation. The left spring has stiffness k a and the right spring has stiffness k b .

The flexibility and stiffness matrix relations are of the form

q1 c 11 c 12 Q 1 Q1 k 11 k 12 q 1
= =
q2 c 21 c 22 Q 2 Q2 k 21 k 22 q 2

Solution

Thin-Walled Structures 269


Article 10.1

q 1, Q 1 q 2, Q 2
ka kb

Fig. 10.1 Two linear elastic springs in series restrained against rigid body translation.

UAS 1: Q 1 = 1 and Q 2 = 0 . The free body diagrams of the springs, with the spring forces assumed positive
in tension, and of nodes 1 and 2 are shown below.

q 1 = c 11 q 2 = c 21
Q1 = 1 Q2 = 0

UAS 1
Sa Sa Sa Sb Sb Sb Sb
node 1 node 2

Equilibrium at nodes one and two gives


– Sa + Sb + 1 = 0 Sb = 0

Therefore,
Sa = 1 Sb = 0

The material laws for the linear elastic springs are


Sa = ka ∆a Sb = kb ∆b

where ∆ a is the elongation of spring a and ∆ b is the elongation of spring b. Spring elongations are related to the
nodal displacements by geometric compatibility, and for this example we have
∆a = q1 ∆b = q2 – q1 .

Thus, 1 = k a q 1 and 0 = k b ( q 2 – q 1 ) . Solve for the displacements to get

1 1
q 1 = ---- = c 11 q 2 = ---- = c 21
ka ka

270 Thin-Walled Structures


Flexibility and stiffness matrices

UAS 2 Q 1 = 0 and Q 2 = 1 . The free body diagrams are shown below.

q 1 = c 12 q 2 = c 22
Q1 = 0 Q2 = 1

UAS 2
Sa Sa Sa Sb Sb Sb Sb
node 1 node 2

Equilibrium at the nodes gives


– Sa + Sb = 0 – Sb + 1 = 0

Therefore,
Sa = 1 Sb = 1

The material laws for each spring and the elongation-displacement relations give
1 = ka q1 1 = kb ( q2 – q1 )

Solve for the displacements to get


1 1 1
q 1 = ---- = c 12 q 2 = ---- + ---- = c 22
ka ka kb

From the method of unit action states we have determined the flexibility matrix to be

1 1
---- ----
ka ka
[C] = (10.1)
1 1 1
----  ---- + ----
k a  k a k b

The flexibility matrix is symmetric, which it must be for a linear elastic structure by Maxwell’s reciprocal theo-
rem: See Article 8.1.4.

UDS 1 q 1 = 1 and q 2 = 0 . From the material laws for each spring and the elongation-displacement relations
we have
Sa = ka q1 = ka ⋅ 1 = ka Sb = kb ( q2 – q1 ) = kb ( –1 ) = –kb

Free body diagrams of the two nodes are shown below.


q1 = 1 q2 = 0
Q 1 = k 11 Q 2 = k 21

UDS 1
Sa Sa = ka ka –kb Sb = –kb Sb –kb
node 1 node 2

Thin-Walled Structures 271


Article 10.1

Equilibrium at the nodes gives


– Sa + Sb + Q1 = 0 – Sb + Q2 = 0

But S a = k a and S b = – k b for UDS 1. Also, we identify Q 1 = k 11 and Q 2 = k 21 for UDS 1. So

k 11 = k a + k b k 21 = – k b

UDS 2 q 1 = 0 and q 2 = 1 . From the material laws for each spring and the elongation-displacement relations
we have
Sa = ka ( 0 ) = 0 Sb = kb ( q2 – q1 ) = kb ( 1 – 0 ) = kb

Free body diagrams of the two nodes are shown below.


q1 = 0 q2 = 1
Q 1 = k 12 Q 2 = k 22

UDS 2
Sa Sa = 0 0 kb Sb = kb Sb kb
node 1 node 2

Nodal equilibrium gives


– Sa + Sb + Q1 = 0 – Sb + Q2 = 0

But S a = 0 and S b = k b for UDS 1. Also, we identify Q 1 = k 12 and Q 2 = k 22 for UDS 1. So

k 12 = – k b k 22 = k b

Therefore, the stiffness matrix is

( ka + kb ) –kb
[K] = (10.2)
–kb kb

The stiffness matrix is also symmetric, which was proved in Article 8.1.6 based on symmetry of the flexibility
matrix.

Note that the matrix product

1 1 ( k a + k b ) k b
 --------------------  – k----b + ----
k b
---- ---- - – ----
ka ka ( ka + kb ) –kb  ka k a  k a k a
[C][K] = = = 10
1 1 1 –kb kb ka + kb )
 (-------------------- 1 1 k 1 1 01
----  ---- + ---- - – k b  ---- + ----   – ----b + k b  ---- + ---- 
k a  k a k b  ka  k a k b   k a  k a k b 

That is, the product of flexibility matrix and the stiffness matrix is the identity matrix. In other words, the inverse
of the flexibility matrix is the stiffness matrix.

272 Thin-Walled Structures


Unrestrained structural stiffness matrix

10.2 Unrestrained structural stiffness matrix


The flexibility influence coefficients c ij are defined for a structure restrained against rigid body motion. How-
ever, it is not necessary to impose this rigid body constraint when forming the stiffness influence coefficients k ij
of a structure. Specifying the generalized displacements in the method of unit displacement states encompasses
both rigid body displacements and those causing deformation. Consider a single, linear elastic spring element
with 2 DOFs connected between nodes i and j, where integers i ≠ j , as shown in Fig. 10.2. Let k denote the stiff-

i j

Fig. 10.2 A two degree of freedom spring element

ness of the spring, which has dimensional units of F/L.

The unrestrained structural stiffness matrix is, in general, given by

Qi k ii k ij q i
= (10.3)
Qj k ji k jj q j

in which i ≠ j . Free body diagrams of the nodes and spring are shown in Fig. 10.3. Equilibrium at the nodes

Qi k Qj
S S S S

i j

Fig. 10.3 Free body diagram of the two-degree of freedom spring element

yields
Qi + S = 0 Qj – S = 0 (10.4)

The spring force is related to the nodal displacements by the material law
S = k ( qj – qi ) (10.5)

UDS 1: q i = 1 and q j = 0 . Therefore eq. (10.5) gives S = – k ⋅ 1 . Nodal equilibrium, eq. (10.4), and the
matrix relation, eq. (10.3), give
Q i = – ( – k ⋅ 1 ) = k ii Q j = – k ⋅ 1 = k ji (10.6)

UDS 2: q i = 0 and q j = 1 . Therefore eq. (10.5) gives S = k ⋅ 1 . Nodal equilibrium, eq. (10.4), and the
matrix relation, eq. (10.3), give

Thin-Walled Structures 273


Article 10.2

Q i = – k ⋅ 1 = k ij Q j = k ⋅ 1 = k jj (10.7)

So that the unrestrained structural stiffness matrix is given by

k –k = k 1 –1
K = (10.8)
–k k –1 1

Note that the unrestrained structural stiffness matrix (10.8) has the following properties:
T
1. Matrix K is symmetric; i.e., K = K

2
2. The sum of the column elements equals zero. ∑k ij = 0 for j = 1, 2 . The vanishing of this sum results
i=1
from Q i + Q j = 0 for each UDS.

3. The det K = k 2 – ( – k ) 2 = 0 . That is, the unrestrained structural stiffness matrix is singular. This occurs

because the structure is not restrained against rigid body translation in the horizontal direction.
Under the action of no external loads, i.e., Q i = 0 and Q j = 0 , the structure can translate horizontally at a
constant speed. Rigid body motion can be used to establish constraints between elements of the unrestrained
structural stiffness matrix. For example, let v denote the horizontal speed and let t denote time, then

q i = q j = vt

Equation (10.3) gives

0 = k ii k ij vt
0 k ji k jj vt

or
( k ii + k ij )q = 0 ( k ji + k jj )q = 0 q = vt ≠ 0

Therefore, the constraints between elements of the unrestrained stiffness matrix are
k ii + k ij = 0 k ji + k jj = 0

4. Diagonal elements of K are positive. This must be true based on physical grounds. If Q i > 0 and q j = 0 ,

then we expect q i > 0 . Thus, in the relation Q i = k ii q i , the stiffness influence coefficient k ii > 0 .

274 Thin-Walled Structures


Assembly of unrestrained structural stiffness matrices

10.3 Assembly of unrestrained structural stiffness matrices


Consider the construction of the 3X3 stiffness matrix for the unrestrained structure shown in Fig. 10.4 given the
generic stiffness matrix of the spring element from eq. (10.8).

1 2 3
ka kb

Fig. 10.4 Unrestrained structure composed of two springs in series

Using the results from eq. (10.8) for each separate spring element, we can write the following results
i j
ka Qi ka –ka qi
=
Qj –ka ka qj
(10.9)

k l
kb Qk kb –kb qk
=
Ql –kb kb ql
(10.10)
Assembly of the individual spring element stiffness matrices is accomplished by displacement continuity at the
nodes and equilibrium at the nodes. Displacement continuity requires
qi = q1 qj = qk = q2 ql = q3 (10.11)

A free body diagram of the structure is shown below.

Q2
ka kb Q3
Q1
Qi Qi Qj Qj Qk Qk Ql Ql

Fig. 10.5 Free body diagram of the two springs in series

Equilibrium at the three nodes requires


Q1 = Qi Q2 = Qj + Qk Q3 = Ql (10.12)

Substitute for the displacements of the individual spring elements in eqs. (10.9) and (10.10) the structural dis-
placements in eqs. (10.11). Then substitute, in turn, these results into the nodal equilibrium equations (10.12) to
eliminate individual spring forces Q i, Q j, Q k, Q l . We get

Q1 = ka q1 – ka q2
Q2 = –ka q1 + ka q2 + kb q2 – kb q3
Q3 = – kb q2 + kb q3

Thin-Walled Structures 275


Article 10.3

Write these last equations in matrix form

Q1 ka –ka 0 q1
Q2 = –ka ( ka + kb ) –kb q2
Q3 0 –kb kb q3

Hence, the unrestrained structural stiffness matrix is

ka –ka 0
K = –ka ( ka + kb ) –kb (10.13)

0 –kb kb

Note the following properties of the unrestrained structural stiffness matrix in eq. (10.13):
T
1. The matrix is symmetric, or K = K .

2. Using a co-factor expansion by the third column, the determinate of the matrix is computed as follows:

–ka ( ka + kb ) ka –ka ka –ka


det K = ( 0 )det – ( – k b )det + ( k b )det
0 –kb 0 –kb –ka ( ka + kb )

det K = 0 – k a k b2 + ( k b ) [ k a ( k a + k b ) – k a2 ] = – k a k b2 + k a k b2 = 0

Since the determinate is zero, the matrix is singular. The unrestrained structural stiffness matrix is singular
because rigid body translation is not restrained.
3
3. The sum of the column elements is zero. ∑k ij = 0 for j = 1, 2, 3 .
i=1

4. Diagonal elements are positive k ii > 0 .

Of course, we could have used the method of unit displacement states to determine the unrestrained struc-
tural stiffness matrix (10.13) for the two springs in series rather than the assembly procedure given above.

Another way to obtain the unrestrained structural stiffness matrix is to first expand the element unrestrained
stiffness matrices to size 3X3 by adding rows and columns of zeros, and then add the 3X3 element stiffness
matrices. For spring element stiffness matrix given by eq. (10.9), displacement compatibility, eq. (10.11), identi-
fies q i = q 1 and q j = q 2 . That is, columns one and two of the element stiffness matrix are associated with glo-
bal degrees of freedom one and two. We write the element stiffness matrix as
q1 q2 q3

ka –ka 0
Ka = –k k 0
a a
0 0 0
(10.14)
The global degrees of freedom are written above the appropriate columns of the expanded element stiffness
matrix in eq. (10.14) to aid in keeping the order of the element columns consistent with the ordering of the global
276 Thin-Walled Structures
Prescribed nodal displacements and forces

displacements. For spring element stiffness matrix given by eq. (10.10), displacement compatibility, eq. (10.11),
identifies q k = q 2 and q l = q 3 . That is, columns one and two of the element stiffness matrix are associated
with global degrees of freedom two and three. We write the element stiffness matrix as
q1 q2 q3
0 0 0
[ Kb ] = 0 kb –kb
0 –kb kb
(10.15)

Note that the only non-zero elements in the expanded element stiffness matrix (10.15) are in rows and columns
two and three. Since matrices (10.14) and (10.15) are of the same dimensions we can add them to get the unre-
strained stiffness matrix of the structure; i.e.,
q1 q2 q3

ka –ka 0
[ K ] = [ Ka ] + [ Kb ] = –ka ( ka + kb ) –kb
0 –kb kb
(10.16)

The unrestrained structural stiffness matrix in eq. (10.16) is the same as the matrix in eq. (10.13). Thus, the
superposition of individual element stiffness matrices to obtain the unrestrained structural stiffness matrix is
equivalent to imposing displacement compatibility and equilibrium at the nodes.

10.4 Prescribed nodal displacements and forces


At a node we can prescribed the either the displacement or the corresponding force, but not both. Consider the
unrestrained structure of the last section in which we prescribe the values of the displacement q 1 , force Q 2 , and
force Q 3 . That is, nodal values of q 1, Q 2, Q 3 are known, and nodal values of Q 1, q 2, q 3 are unknown. Nodal
forces Q 2, Q 3 are the applied loads, and nodal force Q 1 is a reactive force, or support reaction. Nodal displace-
ments q 2, q 3 are the unknown, or active, displacement degrees of freedom. We partition the unrestrained stiff-
ness matrix given in eq. (10.16) by drawing lines between rows and/or columns to separate active and reactive
nodal variables. In this example, we partition row 1 and column 1 as

Q1 ka –ka 0 q1
Q2 = –ka ( ka + kb ) –kb q2
Q3 0 –kb kb q3

Now rearrange the order of the equations and the order of the displacements. The equations for the applied loads
Q 2, Q 3 are moved to correspond with rows one and two, and the reactive force equation for Q 1 is put in row
three. Simultaneously, the unknown displacements q 2, q 3 are ordered such that they appear in columns one and
two, and the prescribed displacement q 1 appears in column three. The equations in matrix form now read as

Thin-Walled Structures 277


Article 10.4

Q2 ( ka + kb ) –kb –ka q2
Q3 = –kb kb 0 q3
Q1 –ka 0 ka q1

The rearranged stiffness matrix is


q2 q3 q1
( ka + kb ) –kb –ka
K = –kb kb 0
–ka 0 ka
(10.17)
In terms of matrix algebra, the unrestrained stiffness matrix in eq. (10.16) was rearranged to the matrix in eq.
(10.17) by the following four step sequence: First interchange elements in rows 1 and 3. Second, interchange ele-
ments in columns 1 and 3. Third, interchange elements in rows 1 and 2. Forth, interchange elements in columns 1
and 2.

Let the vector of unknown displacement degrees of freedom be denoted by q α , the vector of prescribed dis-
placements by q β , the vector of applied forces by Q α , and the vector of reactive forces be denoted by Q β . In the
example of this and the last section these vectors are

q2 Q2
qα = qβ = q1 Qα = Qβ = Q1 (10.18)
q3 Q3

The unrestrained structural stiffness matrix is rearranged to separate unknown and known nodal variables. In
general, this partitioned matrix is written in the form

K αα K αβ
K = (10.19)
K βα K ββ

For the example in this and the last section, a comparison to the matrix in eq. (10.17) gives the submatrices in eq.
(10.19) as

( ka + kb ) –kb –ka
K αα = K αβ = K βα = – k a 0 K ββ = k a (10.20)
–kb kb 0
The matrix equations for the structure in partitioned form are now written as

Qα = K αα K αβ qα (10.21)
Qβ K βα K ββ qβ

Submatrix [ K αα ] is called the restrained structural stiffness matrix. It is a square, symmetric matrix. For the
example of this section it is seen from eq. (10.20) that the restrained stiffness matrix is 2X2, and its determinate

278 Thin-Walled Structures


Solution for the unknown nodal variables

is positive; i.e., det [ K αα ] = k a k b > 0 . The restrained structural stiffness matrix is non-singular if the prescribed
nodal displacements are sufficient to prevent rigid body motion of the structure. The restrained structural stiffness
matrix for this example was determined by the method of unit displacement states in Example 10.1. See eq.
(10.2). The submatrix [ K ββ ] is, in general, square and symmetric. For the example in this section, eq. (10.20)
shows [ K ββ ] is 1X1. The off-diagonal submatrices are, in general, rectangular, but they satisfy the relationship

T
K βα = K αβ (10.22)

10.5 Solution for the unknown nodal variables


Multiply out the matrix equations (10.21) following the ordinary matrix product formula to get

Q α = [ K αα ]q α + [ K αβ ]q β (10.23)

Q β = [ K βα ]q α + [ K ββ ]q β (10.24)

Since the applied load vector Q α and the prescribed displacement vector q β are known, solve eq. (10.23) for the
unknown nodal displacement vector to get
–1 –1
q α = K αα Q α – K αα K αβ q β (10.25)

Continuing with the example of the last two sections, the restrained structural stiffness matrix is given in first
of eqs. (10.20). Its inverse can be computed from

adj K αα
[ K αα ] –1 = -----------------------
det K αα

where adj K αα is the adjoint matrix. The adjoint matrix1 is the transpose of the matrix of co-factors of matrix

K αα . For the 2X2 restrained structural stiffness matrix (10.20)1 in this example, the adjoint matrix is simple to
compute. It is
T
kb kb kb kb
adj K αα = =
kb ( ka + kb ) kb ( ka + kb )

Hence, the inverse matrix is

1. Many determinates must be evaluated to compute the adjoint matrix. For large matrices, evaluating many determinates is
computationally inefficient. Other, more efficient methods to solve large linear systems of equations are use in numerical
algorithms.

Thin-Walled Structures 279


Article 10.5

1 1
---- ----
1 k kb ka ka
[ K αα ] –1 = -------------------------------------2 b = (10.26)
( k a + k b )k b – k b k b ( k a + k b ) 1 1 1
 ----
---- + ----
ka  k a k b

Of course, the inverse of the restrained structural stiffness matrix is recognized as the flexibility matrix. Equation
(10.26) was also obtained by the method of unit action state in Example 10.1. See eq. (10.1). Continuing with the
computations indicated in eq. (10.25) for this example, we have

1 1 1 1
---- ---- ---- ----
q2 ka ka Q2 ka ka –ka
= – q1
q3 1 1 1  Q3
 ---- 1 1 1
 ---- 0
---- + ---- ---- + ----
ka  k a k b ka  k a k b

1 1
---- ----
q2 ka ka Q2
= – –1 q1 (10.27)
q3 1 1 1  Q3
 ---- –1
---- + ----
ka  k a k b

Equation (10.27) is the solution for the unknown nodal displacements.

To find the reactive nodal force vector, substitute the solution for the active nodal displacement vector from
eq. (10.25) into eq. (10.24) to get

 –1 –1 
Q β = K βα  K αα Q α – K αα K αβ q β  + K ββ q β
 

Multiply the matrix products and collect terms in the prescribed nodal displacement vector to get

–1  –1 
Q β = K βα K αα Q α +  K ββ – K βα K αα K αβ q β (10.28)
 

Let’s evaluate eq. (10.28) for the example problem. From eqs. (10.20) and (10.26)

 
1 1
 1 1 
---- ---- ---- ----
ka Q2 
ka ka ka –ka 
Q1 = –ka 0 +  ka – –ka 0 q 1
1 1 1  Q3  1 1 1 0 
----  ---- + ----  ----  ---- + ---- 
k a  k a k b k a  k a k b
 

Performing some of the matrix products, we get

Q2  
Q1 = –1 –1 +  k a – k a q 1
Q3  

This last matrix expression is equivalent to the scalar equation

280 Thin-Walled Structures


Stress matrix

Q1 = – Q2 – Q3

The above result for the reactive nodal force is as expected from overall equilibrium of the structure shown in
Fig. 10.4.

10.6 Stress matrix


The stress matrix is a matrix which directly yields the internal forces or stresses in an element in terms of the
nodal displacements. Consider a typical spring element between nodes i and j as shown in Fig. 10.6. From the

i k j

Q ie S ij S ij S ij S ij Q je

Fig. 10.6 Typical spring element and equivalent nodal forces

overall solution for the structural response, the nodal displacement vector q α is determined. Hence, the actual
nodal displacements q i and q j for the typical spring element are known. Define a vector of equivalent nodal
forces as the element stiffness matrix times the known nodal displacement vector, or
e
Qi q
≡ k –k i (10.29)
Qj –k k qj

These equivalent nodal forces are not the actual forces at the nodes, so they are fictitious. From this equation, the
equivalent nodal forces at node are

qi qi
Q ie = k – k Q je = – k k (10.30)
qj qj

From the free body diagram shown in Fig. 10.6, the spring force S ij is related to the equivalent nodal forces by

Q ie = – S ij Q je = S ij (10.31)

Substitute eqs. (10.30) into (10.31) to eliminate the equivalent nodal forces to find that both of eqs. (10.31) lead
to the same expression for the spring element force. The result is

qi qi
S ij = – k k = S (10.32)
qj qj

where S is the stress matrix for the spring element given by

S = –k k (10.33)

Thin-Walled Structures 281


Article 10.7

For the example problem, the stress matrix for the spring between nodes 1 and 2 is – k a k a . Then, force in
the spring element is

q1
S 12 = – k a k a
q2

From eq. (10.27), the displacement at node 2 is


Q Q
q 2 =  -----2- + -----3- + q 1
 ka ka 

Hence, the force in the spring element between nodes 1 and 2 is


Q Q
S 12 = – k a q 1 + k a q 2 = – k a q 1 + k a  -----2- + -----3- + q 1 = Q 2 + Q 3
 ka ka 

The stress matrix for the spring between nodes 2 and 3 is – k b k b . Then the force in this spring element is

q2
S 23 = – k b k b
q3

Substitute the solution for the nodal displacements from eq. (10.27) into previous equation to get

1 1
---- ----
ka ka Q2
S 23 = – k b k b – –kb kb –1 q1
1 1 1  Q3 –1
----  ---- + ----
k a  k a k b

Perform the matrix products in this last equation to find the force in the element between nodes 2 and 3 is

Q Q Q 1 1  –
S 23 = – k b  -----2- + -----3- + k b  -----2- + Q 3  ---- + ---- 0 q1 = Q3
 ka ka   ka  k a k b 

10.7 Direct stiffness method


We have completed all the steps of the direct stiffness method for the structure shown in Fig. 10.4 in the discus-
sions beginning in Article 10.2 through Article 10.6. The method is summarized as follows:
1. Formulate the member stiffness matrix, and expand it to the overall structural degree of freedom by adding
rows and columns of zeros.
2. Assembly of the member stiffness matrices to form the unrestrained structural stiffness matrix.

K = ∑
elements
K element

3. Prescribe boundary displacement restraints q β and applied nodal forces Q α .

282 Thin-Walled Structures


Planar trusses

Qα = K αα K αβ qα
Qβ K βα K ββ qβ

4. Solution for the unknown nodal displacements.


–1 –1
q α = K αα Q α – K αα K αβ q β

5. Solution for the unknown support reactions.

Q β = [ K βα ]q α + [ K ββ ]q β

6. Determine the member forces/stresses

S ij = S q

10.8 Planar trusses


The element stiffness matrix for a truss bar in the x-y plane is developed. A bar of length L located between
nodes i and j is shown in Fig. 10.7. The coordinates of beginning node i are ( x i, y i ) , and coordinates of the end
node j are ( x j, y j ) , in the undeformed state. The angle between the positive x-direction and directed line element
i-j is denoted as θ . The nomenclature and conventions associated with planar trusses was discussed in Article
8.9, and the matrix structural analysis development is an extension of the discussion in that section. For example,
the x-direction displacement of node i is denoted q 2i – 1 , the y-direction displacement of node i q 2i , the axial
stiffness EA , elongation ∆ , and the axial force, positive in tension, is denoted by N .
j*
y
L*
q 2j
i* θ+Ω
j
yj
q 2j – 1
L q 2i
i θ
yi
q 2i – 1

0 xi xj x

Fig. 10.7 Truss bar in the undeformed and deformed states.

Neglecting the thermal force, Hooke’s law for the truss bar is

EA
N =  ------- ∆ (10.34)
 L

Thin-Walled Structures 283


Article 10.8

Refer to eq. (3.14) on page 52 for Hooke’s law, and eq. (8.81) on page 218 for the strain-elongation relation. For
infinitesimal deformations, the elongation is related to the nodal displacements by eq. (8.101) on page 220,
which is repeated below as eq. (10.35).
∆ = ( cos θ ) ( q 2j – 1 – q 2i – 1 ) + ( sin θ ) ( q 2j – q 2i ) (10.35)

This elongation-displacement relation is written in matrix notation as

q 2i – 1
q 2i T
∆ = –c –s c s = b q (10.36)
q 2j
q 2j – 1

where c = cos θ , s = sin θ , and

T T
b = –c –s c s q = q 2i – 1 q 2i q 2j – 1 q 2j (10.37)

Substitute eq. (10.35) for the elongation in eq. (10.34) to find the axial force for bar i-j as

EA
N i – j =  ------- – c – s c s q = S q (10.38)
 L i–j
i–j

where the stress matrix is defined as

EA EA
- – c – s c s =  ------- b
T
S ≡ ------
L  L
(10.39)

Free body diagrams of the bar and nodes i and j are shown Fig. 10.8. External forces in the x- and y-direc-

Q 2j

θ Q 2j – 1
Ni – j
j
 EA
-------
 L i–j

Q 2i
Ni – j

θ Q 2i – 1
i
Fig. 10.8 Free body diagram of the bar and its end nodes.

tions at node i are denoted by Q 2i – 1 and Q 2i , respectively, and external forces in the x- and y-directions at node
j are denoted by Q 2j – 1 and Q 2j , respectively. Equilibrium at nodes i and j yield

284 Thin-Walled Structures


Planar trusses

Q 2i – 1 + N i – j cos θ = 0 Q 2i + N i – j sin θ = 0 Q 2j – 1 – N i – j cos θ = 0 Q 2j – N i – j sin θ = 0

In matrix notation, these equilibrium equations are written as

Q 2i – 1 –c
Q 2i
= –s Ni – j or Qi – j = b Ni – j (10.40)
Q 2j – 1 c
Q 2j s

where Q i – j is the nodal force vector and matrix b is defined in eq. (10.37). Now substitute eq. (10.38) for the
axial force in eq. (10.40) to get

Q 2i – 1 –c q 2i – 1 c2 cs –c 2 – cs q 2i – 1
Q 2i EA –s q 2i EA cs s2 – cs – s 2 q 2i
=  ------- –c –s c s =  ------- (10.41)
 L  L
Q 2j – 1 c q 2j –c 2 – cs c2 cs q 2j
Q 2j s q 2j – 1 – cs –s 2 cs s 2 q 2j – 1

This last equation is also written as Q = K q where the unrestrained stiffness matrix for the truss bar element
is

c2 cs –c 2 – cs
 EA- cs s2 – cs –s 2
K =  ------
L  –c 2
(10.42)
– cs c2 cs
– cs –s 2 cs s2

EXAMPLE 10.2 A three-bar truss

Each bar in the three bar truss shown in Fig. 10.9 has the same axial stiffness EA , and the nodes are numbered as
shown. Determine the unrestrained structural stiffness matrix using the direct stiffness method.

y 6
5
3

2L
Fig. 10.9 Three bar
truss example. L
4
45° x 2
3
1 L 2
1
DOF numbering convention

Thin-Walled Structures 285


Article 10.8

Solution The following table lists the directions cosines and their products for each bar.

bar θ c s c2 s2 cs
1-2 0° 1 0 1 0 0

1-3 90° 0 1 0 1 0

2-3 135° 1⁄2 1⁄2 –1 ⁄ 2


–1 ⁄ 2 1⁄ 2

From eq. (10.42) and the above table the element stiffness matrices are
q1 q2 q3 q4

1 0 –1 0
EA
K1 – 2 =  ------- 0 0 0 0
 L  –1 0 1 0
0 0 0 0

q1 q2 q5 q6

0 0 0 0
EA
K1 – 3 =  ------- 0 1 0 –1
 L 0 0 0 0
0 –1 0 1

q3 q4 q5 q6

1⁄2 –1 ⁄ 2 –1 ⁄ 2 1⁄2
 EA  – 1 ⁄ 2 1⁄2 1⁄2 –1 ⁄ 2
K2 – 3 = ----------
 2L – 1 ⁄ 2 1⁄2 1⁄2 –1 ⁄ 2
1⁄2 –1 ⁄ 2 –1 ⁄ 2 1⁄2

1
Let a = ---------- , so that
2 2
q3 q4 q5 q6

a –a –a a
EA
K2 – 3 =  ------- – a a a –a
 L  –a a a –a
a –a –a a

Instead of adding rows and columns of zeros to each element stiffness matrix to make it 6X6, only the global
degree of freedom is marked above each column. We are identifying the locations in the unrestrained structural
stiffness matrix where the elements in the individual bar matrices will be added. Now assemblage.

K = K1 – 2 + K1 – 3 + K2 – 3

286 Thin-Walled Structures


Planar trusses

q1 q2 q3 q4 q5 q6

1 0 –1 0 0 0
0 1 0 0 0 –1
 EA – 1 0 1+a –a –a a
K =  ------
-
L 0 0 –a a a –a
0 0 –a a a –a
0 –1 a –a –a 1+a
(10.43)

Note that K is symmetric; the sum of column elements equals to zero; diagonal elements are positive; and its
determinate vanishes.

EXAMPLE 10.3 Restrained three-bar truss of Example 10.2.

Consider the truss of the last example supported in such a manner that nodal displacements q 1 = q 5 = q 6 = 0
as is shown in Fig. 10.10. Assume the nodal forces Q 2, Q 3, Q 4 are prescribed. The unknown nodal displacements
are q 2, q 3, q 4 and the unknown nodal reactions are Q 1, Q 5, Q 6 .

2L
L
2 4
45°
3
L

Fig. 10.10 Three bar truss restrained


against rigid body motion

a) Determine the restrained structural stiffness matrix K αα , and submatrices K αβ , K βα , K ββ .

b) Determine the unknown nodal displacements q 2, q 3, q 4 .

c) Determine the unknown support reactions Q 1, Q 5, Q 6 .

d) Determine the bar forces N 1 – 2, N 1 – 3, N 2 – 3 .

Solution part (a) Rearrange the unrestrained stiffness matrix in eq. (10.43) so that the order of the rows and
columns correspond to DOF’s 2, 3, 4, 1, 5, 6.

Thin-Walled Structures 287


Article 10.8

q2 q3 q4 q1 q5 q6

1 0 0 0 0 –1
0 1+a –a –1 –a a
 EA- 0 –a a 0 a –a
K =  ------
L 0 –1 0 1 0 0
0 –a a 0 a –a
–1 a –a 0 –a 1+a

Compare this matrix to the general form given in eq. (10.19) to identify

1 0 0 0 0 –1
EA EA
K αα =  ------- 0 1 + a – a K αβ =  ------- – 1 – a a
 L  L
0 –a a 0 a –a

0 –1 0 1 0 0
EA EA
K βα =  ------- 0 – a a K ββ =  ------- 0 a – a
 L  L
–1 a –a 0 –a 1 + a

The restrained structural stiffness matrix K αα is symmetric, and the sum of its column elements is non-zero in
general. Also note that the restrained stiffness structural matrix can be obtained from the unrestrained structural
stiffness matrix (10.43) by merely crossing out rows and columns 1, 5, and 6.
q1 q2 q3 q4 q5 q6

1 0 –1 0 0 0
0 1 0 0 0 –1
 EA- –1 0 1+a –a –a a
K =  ------
L 0 0 –a a a –a
0 0 –a a a –a
0 –1 a –a –a 1+a

Solution part (b) The formula for the unknown nodal displacement is given by eq. (10.25), which is

–1 –1
adj K αα
q α = K αα Q α – K αα K αβ q β [ K αα ] –1 = -----------------------
det K αα

T T
In this example the prescribed nodal displacement vector vanishes: i.e., q β = q 1 q 5 q 6 = 000 . The

adjoint of the restrained structural stiffness matrix and it determinate are1

1. The det ( k A ) , where k is a scalar and A is an n-by-n matrix, is equal to k n det A .

288 Thin-Walled Structures


Planar trusses

2 a0 0 3 3
EA EA EA
adj K αα =  ------- 0a a det K αα =  ------- ( ( 1 + a )a – a 2 ) = a  -------
 L  L  L
0 a 1+a
So the inverse of the restrained structural stiffness matrix is

10 0
 L 01 1
[ K αα ] –1 = -------
 EA
1
0 1 1 + ---
a
–1
Perform a check of this inverse. Is K αα K αα = I ?

10 0
10 0 10 0
1 0 0 1 0 0 1 100
 EA L 0 1 a – a  --- + 1 + 1
------- 0 1 + a – a  ------- 0 1 1 = 0 1 + a –a 0 1 1 = a  = 010
 L  EA
1 1
0 –a a 0 1 1 + --- 0 – a a 0 1 1 + --- 1 001
a a 0 0 – a + a  --- + 1
a 

–1
Yes, the inverse satisfies K αα K αα = I . The solution for the unknown nodal displacement vector is

q2 10 0 Q2
 L  0 1 1 1
q3 = ------- Q3 a = ----------
 EA 2 2
1
q4 0 1 1 + --- Q 4
a

Solution part (c) The solution for the support reactions is given by Q β = [ K βα ]q α + [ K ββ ]q β . Also see eq.
(10.28). Again in this example the prescribed nodal displacement vector vanishes: i.e.,
T T
qβ = q1 q5 q6 = 0 0 0 . Submatrix K βα is given in part (a). Hence,

Q1 0 –1 0 q2
EA
Q5 =  ------- 0 – a a q 3
 L
Q6 –1 a –a q4

Substitute for the solution in part (b) for the nodal displacement vector to get

Q1 10 0 Q2 10 0 Q2
0 –1 0 0 –1 0
EA L 01 1
Q5 =  ------- 0 – a a  ------
- Q3 = 0 –a a
0 1 1
Q3
 L  EA
1 1
Q6 –1 a –a 0 1 1 + --- Q 4 – 1 a – a 0 1 1 + --- Q 4
a a

Thin-Walled Structures 289


Article 10.8

0 –1 –1
Q1 1 Q2 0 –1 –1 Q2 – Q3 – Q4
0 0 – a + a  --- + 1
Q5 = a  Q = 0 0 1 Q3 = Q4
3
Q6 1 Q4 –1 0 –1 Q4 – Q2 – Q4
– 1 -0 a – a  --- + 1
a 

Q6 A free body diagram of all the nodal forces is shown in the adjacent sketch. Van-
ishing of moments about node 3 gives
Q5
3
Q1 = – Q3 – Q4

L 2L Vanishing of the moments about node 1 gives

Q4 Q5 = Q4
Q2 1 45° Vanishing of the forces vertically gives
2
L Q3 Q6 = – Q2 – Q4
Q1
Hence, the matrix solution is consistent with equilibrium of the free body dia-
gram.

Solution part (d) The axial normal force for the bar between nodes i and j is given by eq. (10.38), or

N i – j = S i – j q i – j ., and the stress matrix is given by eq. (10.39), or  EA


S i – j ≡  ------- . The
L i – j –c –s c s i–j

direction cosines for each bar are listed in the table in Example 10.2. For bar 1-2, the axial normal force is

q1 0
EA q2
EA EA q2
N 1 – 2 =  ------- – 1 0 1 0 =  ------- – 1 0 1 0 =  ------- q 3
 L q3  L q3  L

q4 q4

L
The solution in part (b) gives q 3 =  ------- ( Q 3 + Q 4 ) , Hence,
 EA

EA L
N 1 – 2 =  -------  ------- ( Q 3 + Q 4 ) = Q 3 + Q 4
 L   EA

For bar 1-3, the axial normal force is

q1 0
 EA q2 EA q EA
N1 – 3 = ------- 0 – 1 0 1 = ------- 0 – 1 0 1 2 =  ------- ( – q 2 )

 L q5  L   L
0
q6 0

L
The solution in part (b) gives q 2 =  ------- Q 2 , Hence,
 EA

290 Thin-Walled Structures


Planar trusses

EA L
N 1 – 3 =  ------- –  ------- Q 2 = – Q 2
 L   EA

For bar 2-3, the axial normal force is

q3 q3
EA q EA
N 2 – 3 =  ---------- 1 ⁄ 2 – 1 ⁄ 2 – 1 ⁄ 2 1 ⁄ 2 4 =  ---------- 1 ⁄ 2 – 1 ⁄ 2 – 1 ⁄ 2 1 ⁄ 2 q 4
 2L q5  2L 0
q6 0

EA q
N 2 – 3 =  ------- 1 – 1 3
 2L  q4

The solution in part (b) gives

Q2
q3 L 01 1
=  ------- Q3
q4  EA
0 1 1+2 2
Q4

Hence,

Q2 Q2
EA L 01 1 1 01 1
N2 – 3 =  ------- 1 – 1  ------
- Q 3 = --- 1 – 1 Q3
 2L   EA 2
0 1 1+2 2 0 1 1+2 2
Q4 Q4

Q2 Q2
1
N 2 – 3 = --- 0 0 – 2 3 / 2 Q 3 = 0 0 – 2 Q 3 = – 2 Q 4
2
Q4 Q4

EXAMPLE 10.4 Five bar truss

The five bar truss shown below is restrained against rigid body motion, since nodes 1 and 4 are fixed. All bars
have the same extensional stiffness EA . Determine the restrained structural stiffness matrix K αα .

4 3
L
cos α = ---------------------
L2 + h2
h
y h
α sin α = ---------------------
x L + h2
2
1 2
L

Thin-Walled Structures 291


Article 10.8

Solution The dimension of the restrained structural stiffness matrix is 4X4 in displacement degrees of freedom
q 3, q 4, q 5, q 6 . The direction cosines for the truss bars are listed below.

bar θ c s c2 s2 cs
1-2 0° 1 0 1 0 0

1-3 α cos α sin α cos2 α sin2 α cos α sin α


2-3 90° 0 1 0 1 0

2-4 180° – α – cos α sin α cos2 α sin2 α – cos α sin α


3-4 180° -1 0 1 0 0

From eq. (10.42) the following member stiffness matrices are constructed using the table above. (Note: only ele-
ments in global DOF’s 3, 4, 5, and 6 are of interest in the individual bar matrices, so elements in the matrices in
DOF’s 1, 2, 7, 8 are indicated by an .)
q1 q2 q3 q4

 EA
K 1 – 2 =  -------
L 1 0
0 0

q1 q2 q5 q6

 EA  EA
K 1 – 3 =  ------------------------ =  -------
L ⁄ ( cos α ) cos2 α cos α sin α  L cos3 α cos2 α sin α
cos α sin α sin2 α cos2 α sin α cos α sin2 α

q3 q4 q5 q6

0 0 0 0 0 0 0 0
EA – 1 =  EA
K2 – 3 =  --------------- 0 1 0 ------- 0 cot α 0 – cot α
 L tan α 0 0 0 0  L 0 0 0 0
0 –1 0 1 0 – cot α 0 cot α

q3 q4 q7 q8

cos2 α – cos α sin α cos3 α – cos2 α sin α


EA sin2 θ EA 2 2
K2 – 4 =  ------------------------ – cos α sin α =  ------- – cos α sin α cos α sin α
 L ⁄ ( cos α )  L

292 Thin-Walled Structures


Planar trusses

q5 q6 q7 q8

1 0
 EA 0 0
K3 – 4 = -------
 L

Assemblage of the restrained structural stiffness matrix

K αα = K 1 – 2 + K 1 – 3 + K 2 – 3 + K 2 – 4 + K 3 – 4

q3 q4 q5 q6

( 1 + cos3 α ) – cos2 α sin α 0 0


 EA 2 2
– cos α sin α ( cot α + cos α sin α ) 0 – cot α
K αα =  -------
L 0 0 ( cos3 α + 1 ) cos2 α sin α
0 – cot α cos2 α sin α ( cos α sin2 α + cot α )

Note that the matrix is symmetric and the column elements do not add to zero. If we take α = 30° , then the
restrained structural stiffness matrix reduces to
q3 q4 q5 q6

3 3
1 + ---------- symm
8
3 9
–  --- 3  ---
EA  8  8
K αα =  -------
 L
3 3
0 0 1 + ----------
8
3 9
0 – 3 --- 3  ---
8  8
(10.44)

In the above matrix only the elements on and below the diagonal are listed, and the notation symm means the
matrix is symmetric.

EXAMPLE 10.5 Using symmetry to reduce problem size

Consider the five bar truss problem of Example 10.4 with α = 30° that is subjected to prescribed nodal forces
Q 3, Q 4, Q 5, Q 6 . Use symmetry to reduce the problem size to solve for the unknown nodal displacements.

Solution We note that the structure and boundary conditions are symmetric about a horizontal axis through the
center of the truss. The applied nodal forces can be decomposed into a symmetric and anti-symmetric set of
forces about this horizontal axis of symmetry as shown in the sketch below.

Thin-Walled Structures 293


Article 10.8

Q6
Q5

y Q4
x 30° Q3

Q6 Ya Yb
Q5 Xa Xb

Q4 = + Yb
Q3 Xa Xb
Ya
(a) symmetric (b) anti-symmetric

The symmetric nodal forces are related to the applied nodal forces by
Xa = ( Q3 + Q5 ) ⁄ 2 Ya = ( – Q4 + Q6 ) ⁄ 2

and the anti-symmetric nodal forces are related to the applied nodal forces by
Xb = ( – Q3 + Q5 ) ⁄ 2 Yb = ( Q4 + Q6 ) ⁄ 2

The total problem solution is then obtained by superposition. That is, the nodal displacement vector in the origi-
nal problem is the sum of the nodal displacements from the symmetric problem (a) and the anti-symmetric prob-
lem (b), or

q3 q3 q3
q4 q4 q4
= +
q5 q5 q5
q6 q6 a
q6 b

In the symmetric problem, the nodal displacements also reflect the symmetry about the horizontal axis of sym-
metry so that
q 5a = q 3a q 6a = – q 4a

In the anti-symmetric problem, the nodal displacement reflect this special symmetry so that
q 5b = – q 3b q 6b = q 4b

Instead of inverting a 4X4 matrix (10.44) to solve for the nodal displacement vector, we invert two 2X2 submatri-
ces and superimpose the displacements from problems (a) and (b). For problem (a) we add column 3 to column 1

294 Thin-Walled Structures


Planar trusses

since q 5a = q 3a , and subtract column 4 from column 2 since q 6a = – q 4a , to get the matrix equations

3 3 3 3 3 3
1 + ---------- –  --- 1 + ---------- –  ---
8  8 8  8
X 3  3  9--- – ( – 3 ) 3 17
–  --- –  --- 3  ------
–Y EA  8   8  q 3a EA  8  8 q 3a
=  ------- =  -------
X  L q 4a  L q 4a
3 3 3 3 3 3
1 + ---------- –  --- 1 + ---------- –  ---
Y a 8  8 8  8
3 9 3 17
--- – 3 – 3  --- --- – 3  ------
8  8 8  8

Notice the equations of the top two rows are the same as the equations from the last two rows. Thus the matrix
equation governing problem (a) is

3 3 3
1 + ---------- –  ---
Xa EA 8  8 q 3a
=  -------
–Ya  L
3 17 q
–  --- 3  ------ 4a
 8  8

The solution for the unknown nodal displacements is


–1
3 3 3 17  3---
1 + ---------- –  ---   3  ------
q 3a L 8  8 Xa L  1   8  8 Xa
=  ------- =  -------  -----------------------
q 4a  EA –Ya  EA  9 17 3
3 17  3--- 3 3 –Ya
–  --- 3  ------  --4- + ------------
8 
- 1 + ----------
 8  8  8 8

q 3a L ( Q3 + Q5 ) ⁄ 2
=  ------- 0.620612 0.0632314
q 4a  EA 0.0632314 0.278137 – ( – Q + Q ) ⁄ 2
4 6

Now consider problem (b). In matrix (10.44) we subtract column 3 from column 1 since q 5b = – q 3b , and
add column 4 to column 2 since q 6b = q 4b , to get the matrix equations

3 3 3 3 3 3
1 + ---------- –  --- 1 + ---------- –  ---
8  8 8  8
–Xb 3 9 3 3
–  --- 3  --- + ( – 3 ) –  --- -------
Yb EA  8  8 q 3b EA  8 8 q 3b
=  ------- =  -------
Xb  L q 4b  L q 4b
3 3 3 3 3 3
–  1 + ---------- --- –  1 + ---------- ---
Yb  8  8  8  8
3 9 3 3
–  --- – 3 + 3  --- –  --- -------
 8  8  8 8

Notice the equations of the top two rows are the same as the equations from the last two rows. Thus the matrix
equation governing problem (b) is
Thin-Walled Structures 295
Article 10.8

3 3 3
1 + ---------- –  ---
–Xb EA 8  8 q 3b
=  -------
Yb  L q 4b
3 3
–  --- -------
 8 8

The solution for the unknown nodal displacements is


–1
3 3 3
1 + ---------- –  --- 3
-------
3
---
q 3b  L 8  8 –Xb  L  1  8 8 –Xb
= ------- = ------- --------------
q 4b  EA Yb  EA  3 ⁄ 8 Yb
3 3 3
–  --- ------- --- 1 + 3---------3-
 8 8 8 8

q 3b L 1 1.73205 ( Q 3 – Q 5 ) ⁄ 2
=  -------
q 4b  EA 1.73205 7.6188 ( Q + Q ) ⁄ 2
4 6

The total solution is given by the sum of problems (a) and (b), or

q3 q3 q3 1 0 1 0
q4 q4 q4 q 3a + 0 1 q 3b
= + = 0 1
q5 q5 q5 1 0 q 4a – 1 0 q 4b
q6 q6 q6 0 –1 0 1
a b

q3 1 0
q4 L  0.620612 0.0632314 ( Q3 + Q5 ) ⁄ 2
= 0 1  ------
- +
q5 1 0  EA  0.0632314 0.278137 – ( – Q 4 + Q 6 ) ⁄ 2
q6 0 –1

1 0
0 1 L
 ------
- 1 1.73205 ( Q 3 – Q 5 ) ⁄ 2
–1 0  EA 1.73205 7.6188 ( Q + Q ) ⁄ 2
4 6
0 1

q3  
 0.620612 0.0632314 1 1.73205 
q4 L  ( Q + Q ) ⁄ 2 7.6188 ( Q 3 – Q 5 ) ⁄ 2 
=  -------  0.0632314 0.278137 3 5 + 1.73205 
q5  EA
 0.620612 0.0632314 – ( – Q 4 + Q 6 ) ⁄ 2 –1 – 1.73205 ( Q 4 + Q 6 ) ⁄ 2 
 
q6  – 0.0632314 – 0.278137 1.73205 7.6188 

q3 0.810306Q 3 + 0.897641Q 4 – 0.189694 Q 5 + 0.83441Q 6


q4 L 0.897641Q 3 + 3.94847Q 4 – 0.83441 Q 5 + 3.67033Q 6
=  -------
q5  EA – 0.189694 Q – 0.83441Q + 0.810306Q – 0.897641Q
3 4 5 6
q6 0.83441Q 3 + 3.67033Q 4 – 0.897641 Q 5 + 3.94847Q 6

296 Thin-Walled Structures


Self strained plane truss

10.9 Self strained plane truss


Strain of the bars in a truss can occur due to temperature changes and due to the lack of fit during assembly, even
in the absence of applied nodal forces. If the truss is statically determinate, the straining of the bars due to these
thermal and lack of fit effects results in a stress-free state for each bar. However, for a statically indeterminate
truss thermal and lack of fit effects result in non-zero bar stresses or internal forces. As shown is in Fig. 10.11, if

3 3

2* 4 2*

1 2 1 2

(a) statically determinate (b) statically indeterminate

Fig. 10.11 Bar 1-2 is heated above room temperature and the other bars remain at room temperature

bar 1-2 is heated above room temperature in a statically determinate truss, node 2 displaces to 2* such that the
length of bar 1-2 is increases but bar 2-3 does not change length. If bar 1-2 is heated above room temperature in a
statically indeterminate truss, the length of bar 1-2 increases but bars 2-3 and 2-4 also change length. In the stati-
cally determinate case, the heated bars change length, or are strained, but the other bars do not strain. The internal
bar forces vanish in the statically determinate case under thermal loading or lack of fit. In the statically indetermi-
nate case, all bars are strained and the internal forces in the bars are, in general, non-zero. Lack of fit during
assembly is, for example, illustrated by the fact that bar 1-2 may to too short or too long to connect to node 2.
Assembly results in straining of bar 1-2, which for the statically determinate case results in no bar forces but a
displacement of node 2 to 2*. For the statically indeterminate case, the lack of fit of bar 1-2 results in straining of
bars 1-2, 2-3, and 2-4 with the bar forces non-zero.

The axial normal force in a truss bar including the thermal strain is


N = EA  --- – EAα ( T – T 0 ) (10.45)
 L

where α is the coefficient of thermal expansion of the material and T – T 0 is the change in temperature from the
stress free state. The change in temperature is assumed to be spatially uniform in the truss bar. (Refer to Article
3.3; in particular, eq. (3.14) and eq. (3.15) on page 52, for the effects of temperature on strain and on Hooke’s
law.) The force vector of bar i-j is denoted by


N i – j = c  EA  --- – EAα ( T – T 0 ) (10.46)
s   L 
i–j

Thin-Walled Structures 297


Article 10.9

Q j where the direction cosines are denoted by c = cos θ and


s = cos ( 90° – θ ) = sin θ . See Fig. 10.12. The elongation of the
–Ni – j bar is related to the nodal displacements by eq. (10.36), which is
Ni – j j repeated below
 EA
-------
 L i–j
q 2i – 1
θ q 2i
–Ni – j ∆i – j = –c –s c s (10.47)
Qi i–j
q 2j – 1
Ni – j Fig. 10.12 Free body q 2j
diagram of bar i-j
i The bar force (10.46) consists of a mechanical part caused by the
elongation ∆ of the bar plus a thermal part caused by the change in
temperature from the stress free state. Hence, we write eq. (10.46) as
(m) (t)
Ni – j = Ni – j + Ni – j (10.48)

where mechanical and thermal bar force vectors, respectively, are defined by

(m) ∆ (t)
N i – j ≡ c EA  --- N i – j ≡ – c EAα ( T – T 0 ) (10.49)
s  L s
i–j i–j

Substitute eq. (10.47) for the elongation to write the mechanical part of the bar force in terms of the nodal dis-
placements to get

q 2i – 1 q 2i – 1
(m) EA c q 2i EA –c 2
– cs c2
cs q 2i
Ni – j =  ------- =  ------- (10.50)
 L  i – j s –c –s c s  L i–j
q 2j – cs – s cs s 2
2
i–j q 2j – 1
q 2j – 1 q 2j

Now consider a truss of three bars where only node j is movable, bar i-j is heated above the stress free state,
and the other two bars are not heated, as is shown in Fig. 10.13. Free body diagrams of the nodes is also shown in
this figure.
Qj

Qk –Nk – j j
j Nk – j
k
–Nl – j
T – T0 k
–Ni – j
i
Nl – j
l Qi Ni – j Ql

l
i
Fig. 10.13 Bar i-j is heated while bars k-j and l-j are not. FBD’s of nodes.

298 Thin-Walled Structures


Self strained plane truss

Vector force equilibrium at the four nodes gives

Qi = –Ni – j Qj = Ni – j + Nk – j + Nl – j Qk = –Nk – j Ql = –Nl – j


From eq. (10.48) each bar force is the sum of the mechanical part and the thermal part. Thus, we can write nodal
equilibrium of the truss in matrix form as the vector sum
(m) (t)
Qi –Ni – j –Ni – j
Qj = Ni – j + Nk – j + Nl – j + Ni – j + Nk – j + Nl – j (10.51)
Qk –Nk – j –Nk – j
Ql –Nl – j –Nl – j
The mechanical part of the bar forces are of the form shown in eq. (10.50). Following the assembly procedure,
the mechanical part of the nodal force vector becomes

q
x x x x 0 0 0 0 2i – 1
(m) q 2i
x x x x 0 0 0 0
–Ni – j x x (x + y + z) (x + y + z) y y z z q 2j – 1
Ni – j + Nk – j + Nl – j = s x x (x + y + z) (x + y + z) y y z z q 2j =
K q (10.52)
–Nk – j 0 0 y y y y 0 0 q 2k – 1
0 0 y y y y 0 0 q 2k
–Nl – j
0 0 z z 0 0 z z q
2l – 1
0 0 z z 0 0 z z
q 2l

where K is the 8X8 unrestrained structural stiffness matrix of the truss. The assembled form of the unre-
strained structural stiffness matrix shown in eq. (10.52) is in terms of elements from the individual stiffness
matrices designate by the symbols x, y, z. The locations of the elements from the 4X4 stiffness matrix of bar i-j
are denoted by an x, locations of elements from the 4X4 stiffness matrix of bar k-j by y, and the locations of ele-
ments from the 4X4 stiffness matrix of bar l-j are denoted by z.

The thermal force vectors for the bars in this example are

(t) (t) (t)


N i–j = – c EAα ( T – T 0 ) Nk – j = Nl – j = 0
s 0
i–j

Hence, the thermal part of the nodal displacement vector in eq. (10.51) is

Thin-Walled Structures 299


Article 10.9

–c
(t) –s
–Ni – j c
Ni – j + Nk – j + Nl – j = – s EAα ( T – T 0 ) (10.53)
–Nk – j 0
0
–Nl – j
0
0
This 8X1 vector of nodal forces is computed with all nodes fixed, so it called the fixed-end nodal force vector,
0
and is denoted by Q . That is,

–c
–s
c
0
Q ≡ – s EAα ( T – T 0 ) (10.54)
0
0
0
0
Now the nodal force equation (10.51) is written as
0
Q = K q+Q
8X1 8X8 8X1 8X1 (10.55)

The fixed end action vector is subtracted from each side of this equation, since it is a known vector determined
from the specified temperature changes of the bars. That is, eq. (10.55) is written in the form
0
Q + ( –Q ) = K q
equivalent joint force vector (10.56)
0
The vector – Q is called the equivalent joint force vector.

Following the direct stiffness method, eq. (10.56) is partitioned into the prescribed nodal displacements and
the corresponding reactive nodal forces, and the unknown nodal displacements and the corresponding applied
nodal forces. For this three bar truss, the prescribed nodal displacements and the corresponding reactive nodal
forces are

300 Thin-Walled Structures


Self strained plane truss

q 2i – 1 Q 2i – 1
0
q 21 0 Q 21
q 2k – 1 Q 2k – 1
qβ = = 0 Qβ =
q 2k 0 Q 2k
q 2l – 1 0 Q 2l – 1
q 2l 0 Q 2l

The unknown nodal displacements and the corresponding prescribed nodal forces are

q 2j – 1 Q 2j – 1
qα = Qα = = 0
q 2j Q 2j 0

Equation (10.56) is rearranged into the standard partitioned form

 0  K αα K αβ
Qα +  – Qα  = qα (10.57)
 0 
Qβ  Qβ  K βα K ββ qβ

For the three bar truss, the unknown nodal displacements at node j are determined from the solution of the matrix
equation

0 +  – – c EAα ( T – T ) q
= K αα 2j – 1
0
0  –s i–j q 2j

After solving for the unknown nodal displacement, the reactive nodal forces are determined from

Q 2i – 1  
–c
Q 21  
 –s 
Q 2k – 1  0  EA  q
+ – ------- α ( T – T 0 ) = K βα 2j – 1
Q 2k  0  L   q 2j
 
Q 2l – 1  0 
 0 
Q 2l

The bar forces are determined from eqs. (10.45) and (10.47) as is shown by the following equations for this three
bar truss.

 
 0 
 EA  0 
Ni – j = -------  – EAα ( T – T 0 )
 L  i – j  – c – s c s i–j q 2j – 1  i–j
 
 q 2j 

Thin-Walled Structures 301


Article 10.9

 
 0 
 EA  0 
Nk – j = -------  – EAα ( 0 )
 L  k – j  – c – s c s k–j q 2j – 1 
k–j
 
 q 2j 

 
 0 
EA  0 
Nl – j =  -------  – c – s c s  – EAα ( 0 )
 L  l – j l–j q 2j – 1 
l–j
 
 q 2j 

Lack of fit Due to manufacturing tolerances, say, the length of bar 1-2 in Fig. 10.11 has a length L ± ∆ , where
L denotes its nominal length and ∆ denotes the magnitude of the manufacturing tolerance. Tolerances are usu-
ally such that ∆ « L . Then, to connect bar 1-2 to node 2 it is has to be strained by an amount ± ∆ ⁄ L . After the
connection is made, bar 1-2 will experience an additional elongation ∆ due to the displacement of joint 2. Thus,
we can replace the thermal strain in eq. (10.45) with the strain ± ∆ ⁄ L . That is, the bar force is now


N = EA  --- – EA [ ± ∆ ⁄ L ] (10.58)
 L

The analysis for the fixed-end actions and the response of the truss due to the lack of fit proceeds as in the ther-
mal load case with α ( T – T 0 ) replaced by ± ∆ ⁄ L .

EXAMPLE 10.6 A simple statically indeterminate truss subject to heating

In the truss shown below all three bars have the same extensional stiffness EA , the length of bar 1-3 is L , bar 1-
3 is heated above the stress free temperature, and bars 1-2 and 2-3 are not heated.
6
3
5

45°
T – T0 L
2
4

1 45° 2
1 3
DOF numbering

Determine
a) the fixed-end actions
b) the unknown nodal displacements
c) support reactions

302 Thin-Walled Structures


Self strained plane truss

d) bar forces

Solution part (a): the fixed end actions. Fix nodes 1, 2, and 3, and draw a free body diagram to determine the
fixed-end nodal forces, which is shown below. The mechanical part of the bar forces vanish for all nodes fixed.

(0)
Q3
(t)
–N 1–3 (t)
–N 2–3
(t)
N 1–3 (t)
N 2–3

(t)
–N 1–3 (t)
–N 2–3
(t)
N 1–3 (t)
N 2–3 0
Q2
0
Q1
(t) (t) (t) (t)
N 1–2 –N1 – 2 N 1–2 –N1 – 2

Vector force equilibrium at the nodes results in


0
(t) (t)
Q1 –N 1–2 –N 1–3
= (t) (t)
Q2 N 1–2 – N2 – 3
(t) (t)
Q3 N1 – 3 + N2 – 3

The thermal parts of the bar forces, as obtained from eq. (10.46), are

N 1 – 2 = 1 ( – ( EAα ) 1 – 2 ( 0 ) ) N 1 – 3 = 1 ⁄ 2 ( – ( EAα ) 1 – 3 ( T – T 0 ) ) N 2 – 3 = 0 ( – ( EAα ) 2 – 3 ( 0 ) )


0 1⁄ 2 1

Hence, the fixed end action vector is

0–1⁄ 2 –1 ⁄ 2
0
Q1 0–1⁄ 2 –1 ⁄ 2
0 0 0
Q = = ( – EAα ( T – T 0 ) ) = ( – EAα ( T – T 0 ) )
Q2 0 0
Q3 0+1⁄ 2 1⁄ 2
0+1⁄ 2 1⁄ 2

Solution part (b): unknown nodal displacements The matrix eq. (10.56) for this example is

Thin-Walled Structures 303


Article 10.9

Q1   q1
 1⁄ 2  2+1⁄2
Q2   1⁄2 1⁄2 symm q2
 1⁄ 2 
Q3  0  EA – 2 0 2 q3
+  – EAα ( T – T 0 )  =  ------- (10.59)
Q4  0  L q4
0 0 0 2
 
Q5  –1 ⁄ 2  –1 ⁄ 2 –1 ⁄ 2 0 0 1⁄2 q5
 
Q6  –1 ⁄ 2  –1 ⁄ 2 –1 ⁄ 2 0 – 2 1 ⁄ 2 1 ⁄ 2 + 2 q6

The prescribed nodal displacements and the corresponding reactive nodal forces are

q2 0 Q2
q3 Q3
qβ = = 0 Qβ =
q4 0 Q4
q5 0 Q5

The unknown nodal displacements and the corresponding prescribed nodal forces are

q1 Q1
qα = Qα = = 0
q6 Q6 0

Then, the governing matrix equation for the restrained truss is

 
0 +  – EAα ( T – T ) 1 ⁄ 2  =  EA 2+1⁄2 –1 ⁄ 2 q1
0 -------
0     L
–1 ⁄ 2  –1 ⁄ 2 1 ⁄ 2 + 2 q6

Solve for the unknown nodal displacements

 
q1 L 1 2+1⁄2 1⁄2  – EAα ( T – T ) 1 ⁄ 2 
=  ------- --------------------- 0
q6  EA ( 2 + 2 )  
1⁄2 2+1⁄2  –1 ⁄ 2 

q1 Lα ( T – T 0 ) – 1
= --------------------------
-
q6 (2 + 2) 1

Solution part (c): support reactions The support reactions are determined in the direct stiffness method from
0
Q β + ( – Q β ) = K βα q α . The unknown support reactions are associate with degrees of freedom 2, 3, 4, and 5.

The active displacements are associated with degrees of freedom 1 and 6. Thus, elements of submatrix K βα are
extracted from rows 2, 3, 4, and 5, and columns 1 and 6, of the unrestrained structural stiffness matrix given in
eq. (10.59). The matrix equation for the unknown support reactions is

304 Thin-Walled Structures


Self strained plane truss

Q2   1 ⁄ 2 –1 ⁄ 2
 1⁄ 2 
Q3  0  EA q1
+  – EAα ( T – T 0 )  =  ------- – 2 0
Q4  0  L
0 – 2 q6
 
Q5  –1 ⁄ 2  –1 ⁄ 2 1 ⁄ 2

Substitute the solution for the unknown displacements into this equation to get

Q2 1 ⁄ 2 –1 ⁄ 2 1⁄ 2
Q3 EA Lα ( T – T )
=  ------- – 2 0 --------------------------
0 –1
- + EAα ( T – T 0 ) 0
Q4  L ( 2 + 2 ) 1 0
0 – 2
Q5 –1 ⁄ 2 1 ⁄ 2 –1 ⁄ 2

2
----------------
2+ 2
Q2 2
----------------
Q3
= EAα ( T – T 0 ) 2 + 2
Q4 2
– ----------------
Q5 2+2
2
– ----------------
2+2

The nodal forces are shown in the free body diagram of the truss in the sketch below. Note the overall force and
moment equilibrium are satisfied.

2EAα ( T – T 0 )
---------------------------------------
2+ 2

L
L ⁄ ( 2)
2EAα ( T – T 0 )
---------------------------------------
2+ 2 L ⁄ ( 2) 2EAα ( T – T 0 )
---------------------------------------
2+ 2

2EAα ( T – T 0 )
---------------------------------------
2+ 2

Solution to part (d): bar forces For bar 1-2 the angle θ between the positive x-axis and the directed line 1-2 is
zero so c = 1 and s = 0 . From eqs. (10.45) and (10.47), the axial normal force in bar 1-2 is given by

Thin-Walled Structures 305


Article 10.9

 q1  
  q1 
  
 EA q2  EA
N1 – 2 = -------  – EAα ( T – T 0 ) =  --------------  – 1 – 0 1 0 0  – EAα ( 0 )
 L  1 – 2  – c – s c s 1–2
q3  1–2  L ⁄ 2  
 0 
  
 q4  0 

Substitute the solution given in part (b) for the nodal displacement q 1 to get

EA EA Lα ( T – T 0 ) 2EAα ( T – T 0 )
N 1 – 2 = – 2  ------- q 1 = – 2  ------- – --------------------------
- = --------------------------------------
-
 L  L (2 + 2) (2 + 2)
So bar 1-2 is in tension for T – T 0 > 0 . For bar 1-3 the angle θ between the positive x-axis and the directed line
1-3 is 45° so c = 1 ⁄ 2 and s = 1 ⁄ 2 . From eqs. (10.45) and (10.47), the axial normal force in bar 1-3 is
given by

 q1 
 
EA  1 1 1 1 q 
N1 – 3 =  -------  – ------- – ------- ------- ------- 2  – EAα ( T – T 0 )
 L  1–3
2 2 2 2 q5 
 
 q6 

 
 q1 
EA  1 1 1 1  EA 1 1 q
N1 – 3
 L  - – ------- ------- ------- 0  – EAα ( T – T 0 ) =  ------- – ------- ------- 1 – EAα ( T – T 0 )
=  -------  – ------
2 2 2 2 0  L 2 2 q6
 
 q 6 

Substitute the solution given in part (b) for the nodal displacements q 1 and q 6 to get

EA 1 Lα ( T – T 0 )
1 ------
- - --------------------------- – 1 – EAα ( T – T 0 )
N 1 – 3 =  ------- – ------
 L
2 2 (2 + 2) 1

2
N 1 – 3 = – E Aα ( T – T 0 )  ----------------
 2 + 2

So bar 1-3 is in compression for T – T 0 > 0 . For bar 2-3 the angle θ between the positive x-axis and the directed
line 2-3 is 90° so c = 0 and s = 1 . From eqs. (10.45) and (10.47), the axial normal force in bar 2-3 is given by

 q3   
  0
   
EA q EA
N2 – 3 =  --------------  – 0 – 1 0 1 4  – EAα ( 0 ) =  --------------  – 0 – 1 0 1 0  – EAα ( 0 )
 L ⁄ 2  q5   L ⁄ 2  0 
  
 q6   q6 

Substitute the solution given in part (b) for the nodal displacement q 6 to get

306 Thin-Walled Structures


Structures containing beam elements

EA EA Lα ( T – T 0 ) 2EAα ( T – T 0 )
N2 – 3 = 2  ------- q 6 = 2  -------  --------------------------
- = --------------------------------------
-
 L  L   (2 + 2)  2+ 2
So bar 2-3 is in tension for T – T 0 > 0 .

10.10 Structures containing beam elements


Consider a prismatic, homogeneous beam with at least one axis of symmetry in its cross section, and which is
subjected to transverse loads acting in a plane that passes through the locus of shear centers of the cross sections.
This beam is referenced to a Cartesian coordinate system with the x and y axes in the cross section, and the z-axis
the longitudinal axis, or neutral axis, of the beam. The origin of the x-y axes is at the centroid of the cross section,
and either one, or both, of these axes are axes of symmetry in the cross section so that the product area moment
I xy = 0 . The plane of loading passing through the locus of shear centers is taken parallel to the y-z plane. Hence,
the beam bends in the y-z plane. The governing differential equation for the deflection of the beam is obtained
from eq. (5.13) on page 98 as
4
dv
EI xx = py ( z ) 0<z<L (10.60)
dz4
in which v ( z ) is the lateral displacement of the neutral axis defined positive in the positive y-direction, EI xx is
the bending stiffness about the x-axis in the cross section, and p y ( z ) is the specified distributed load intensity
defined positive in the positive y-direction. The rotation of the cross section about the x-axis at z is denoted by
θ x ( z ) , and for classical beam theory in which plane cross sections remain plane and perpendicular to the neutral
axis in the deformed beam, this rotation is related to the deflection by

dv
θx = – (10.61)
dz
Refer to eq. (5.8) on page 97.

Let q 1 denote the y-direction displacement of the neutral axis at z = 0, q 2 the rotation of the cross section
about the x-axis at z = 0, q 3 the y-direction displacement of the neutral axis at z = L, and let q 4 denote the rota-
tion of the cross section about the x-axis at z = L. Then the boundary conditions at the ends of the beam are writ-
ten as
v ( 0 ) = q1 θx ( 0 ) = q2 v ( L ) = q3 θx ( L ) = q4 (10.62)

Consider the distributed load intensity p y ( z ) , and the boundary displacements


y, v py ( z )
and the rotations q 1, q 2, q 3, q 4 as prescribed quantities. See the adjacent sketch. q1 q3
Our purpose is to solve differential equation (10.60) subject to boundary condi- z
tions (10.62) for the lateral displacement function v ( z ) . The solution is sought q2 L q4
by the method of superposition. Let the lateral displacement be represented by
the sum of displacements in the form
v ( z ) = v0 ( z ) + v1 ( z ) (10.63)

Thin-Walled Structures 307


Article 10.10

The boundary value problem for v 0 ( z ) is selected as

EIv 0 ′′′′ = p y ( z ) 0<z<L


(10.64)
v0 ( 0 ) = 0 –v0 ′ ( 0 ) = 0 v0 ( L ) = 0 –v0 ′ ( L ) = 0

and as a consequence the boundary value problem for v 1 ( z ) is

EIv 1 ′′′′ = 0 0<z<L


(10.65)
v1 ( 0 ) = q1 –v0 ′ ( 0 ) = q2 v0 ( L ) = q3 –v0 ′ ( L ) = q4

dv
In eqs. (10.64) and (10.65) ordinary derivatives with respect to z are denoted by primes; e.g., v′ = . Also, let
dz
EI = EI xx . The boundary value problem (10.64) for displacement function v 0 ( z ) consists of an inhomogeneous
differential equation with homogeneous boundary conditions, while the boundary value problem (10.65) for dis-
placement function v 1 ( z ) consists of a homogeneous differential equation with inhomogeneous boundary condi-
tions. Since the displacements and rotations vanish at the end points of the beam in the boundary value problem
for v 0 ( z ) , the solution for it will lead to fixed end actions in the matrix structural analysis method. That is, the
fixed end action problem accounts for distributed load intensity p y ( z ) , and the discussion of it will be postponed
until the next section. In this section, we will consider the boundary value problem (10.65) for v 1 ( z ) .

The general solution for v 1 ( z ) satisfying the differential equation in boundary value problem (10.65) is a
cubic polynomial in the longitudinal coordinate, which can be written as

z3 z2
v 1 ( z ) = c 3 ---- + c 2 ---- + c 1 z + c 0 (10.66)
6 2
where the constants c 3, c 2, c 1, c 0 are to be determined by the four boundary conditions specified in eq. (10.65).
Substitute the general solution (10.66) into these four boundary conditions to get
03 02
c 3 ⋅ ----- + c 2 ⋅ ----- + c 1 ⋅ 0 + c 0 = q 1
6 2
02
– c 3 ⋅ ----- – c 2 ⋅ 0 – c 1 – c 0 ⋅ 0 = q 2
2
L3 L2
c 3 ⋅ ----- + c 2 ⋅ ----- + c 1 ⋅ L + c 0 = q 3
6 2
L2
– c 3 ⋅ ----- – c 2 ⋅ L – c 1 – c 0 ⋅ 0 = q 4
2
or in matrix notation

0 0 0 1 c3 q1
0 0 –1 0 c2 q2
=
L3 ⁄ 6 L2 ⁄ 2 L 1 c1 q3
–L 2 ⁄ 2 –L –1 0 c0 q4

Solve this last equation for the constants c 3, c 2, c 1, c 0 to get

308 Thin-Walled Structures


Structures containing beam elements

c3 12 ⁄ L 3 – 6 ⁄ L 2 – 12 ⁄ L 3 – 6 ⁄ L 2 q1
c2 2 2 q2
= –6 ⁄ L 4 ⁄ L 6 ⁄ L 2⁄L
c1 0 –1 0 0 q3
c0 1 0 0 0 q4

Substitute this solution for the constants c 3, c 2, c 1, c 0 in eq. (10.66) to get

1 12 6 12 6 1 6 4 6 2
v 1 ( z ) = ---  -----3- q 1 – ----2-q 2 – -----3- q 3 – ----2-q 4 z 3 + ---  – ----2-q 1 + --- q 2 + ----2-q 3 + --- q 4 z 2 + ( – q 2 )z + q 1
6 L L L L  2 L L L L 
Rearrange this equation to the form

z3 z2 z3 z2 z3 z2 z3 z2
v 1 ( z ) =  2 ----3- – 3 ----2- + 1 q 1 +  – ----2- + 2 ---- – z q 2 +  – 2 ----3- + 3 ----2- q 3 +  – ----2- + ---- q 4 (10.67)
 L L   L L   L L   L L
Equation (10.67) is further written in the matrix form

q1
q2
v1 ( z ) = η1 ( z ) η2 ( z ) η3 ( z ) η4 ( z ) = η(z) q (10.68)
q3
q4

The shape functions, or interpolation functions, are defined as

z3 z2 z3 z2 z3 z2 z3 z2
η 1 ( z ) ≡ 2 ----3- – 3 ----2- + 1 η 2 ( z ) ≡ – ----2- + 2 ---- – z η 3 ( z ) ≡ – 2 ----3- + 3 ----2- η 4 ( z ) ≡ – ----2- + ---- (10.69)
L L L L L L L L
The rotation (10.61) associated with the lateral displacement function v 1 ( z ) is given by

q1
q2
θ x1 ( z ) = – v 1 ′ ( z ) = – η 1 ′ ( z ) – η 2 ′ ( z ) – η 3 ′ ( z ) – η 4 ′ ( z ) (10.70)
q3
q4

where

6 z z2 z z2 z z2 z
η 1 ′ ( z ) = -----z 2 – 6 ----- η 2 ′ ( z ) = – 3 ----- + 4 --- – 1 η 3 ′ ( z ) = – 6 ----- + 6 ----- η 4 ′ ( z ) = – 3 ----- + 2 --- (10.71)
L3 L2 L2 L L3 L2 L2 L

These interpolation functions have the following properties at the end points, or nodes, of the beam element:
η1 ( 0 ) = 1 η2 ( 0 ) = 0 η3 ( 0 ) = 0 η4 ( 0 ) = 0
η1 ′ ( 0 ) = 0 η2 ′ ( 0 ) = –1 η3 ′ ( 0 ) = 0 η4 ′ ( 0 ) = 0
η1 ( L ) = 0 η2 ( L ) = 0 η3 ( L ) = 1 η4 ( L ) = 0
η1 ′ ( L ) = 0 η2 ′ ( L ) = 0 η3 ′ ( L ) = 0 η4 ′ ( L ) = –1

These interpolation functions (10.69) and (10.71) are plotted in Fig. 10.14 to Fig. 10.17 for special values of the
nodal displacement vector.

Thin-Walled Structures 309


Article 10.10

.
1
1.4

0.8 1.2
η1 1 –η1 ′
0.6 v θx L
----1- --------- 0 . 8
q1 q1
0.4 0.6

0.4
0.2
0.2
z⁄L z⁄L
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1

Fig. 10.14 Dimensionless distributions of the lateral displacement and the rotation for nodal
displacements q2 = 0, q3 = 0, and q4 =0.

z⁄L
1
0.2 0.4 0.6 0.8 1
-0.02 0.8

-0.04 0.6

v1 - 0 . 0 6 θ 0.4
--------
- ----x-
q2 L - 0 . 0 8 η2 q2 0.2 –η2 ′
z⁄L
-0.1
-0.12 0.2 0.4 0.6 0.8 1
-0.2
-0.14
-0.4

Fig. 10.15 Distributions of the dimensionless lateral displacement and rotation for nodal
displacement q1 = 0, q3 = 0, and q4 = 0.

z⁄L
1
0.2 0.4 0.6 0.8 1
-0.2
0.8
-0.4
0.6
v θx L - 0 . 6
----1- ---------
q3 η3 q3 - 0 . 8 –η3 ′
0.4
- 1
0.2 -1.2
z⁄L -1.4
0.2 0.4 0.6 0.8 1

Fig. 10.16 Distributions of the dimensionless lateral displacement and rotation for nodal
displacements q1 = 0, q2 = 0, and q4 =0.

310 Thin-Walled Structures


Structures containing beam elements

1
0.14
v1 0.8 θ
0.12
--------
- ----x-
0.1
q4 L 0 . 6 q4

0.08 0.4

0.06
η4 –η4 ′
0.2
z⁄L
0.04
0.2 0.4 0.6 0.8 1
0.02
z⁄L -0.2

0.2 0.4 0.6 0.8 1 -0.4

Fig. 10.17 Distributions of the dimensionless lateral displacement and rotation for nodal
displacements q1 = 0, q2 = 0, and q3 = 0.

The bending moment M x ( z ) in the beam is determined from Hooke’s law as given by eq. (5.11) on page 98.
For the cross section considered here the product area moment is zero, so Hooke’s law reduces to
M x = EI xx ( θ x ′ ) = EI xx ( – v′′ )

The transverse shear force V y ( z ) is determined from moment equilibrium per unit Vy
length along the beam as given by eq. (5.2) on page 96. That is, V y = M x ′ . Positive y
values of the moment and shear are shown in the adjacent sketch. From Hooke’s law Mx Mx
z
for the bending moment given above, the transverse shear force for this homoge-
neous, uniform beam is related to the lateral displacement by Vy

V y = EI xx ( – v′′′ )

Let us calculate the distribution of the bending moment, and the distribution of the transverse shear force, for
the boundary value problem governed by (10.65). Then, v = v 1 in the expressions above for the moment and
shear, and the solution for v 1 in terms of nodal displacements and nodal rotations is given by eq. (10.68). Let the
bending moment M x and transverse shear force V y for the lateral displacement function v 1 ( z ) be denoted by
M 1 and V 1 , respectively. Again, let EI = EI xx . In matrix form the transverse shear force and the bending
moment are written as follows:

q1
V1 ( z ) – v 1 ′′′ – η 1 ′′′ – η 2 ′′′ – η 3 ′′′ – η 4 ′′′ q 2
= EI = EI
M1 ( z ) – v 1 ′′ – η 1 ′′ – η 2 ′′ – η 3 ′′ – η 4 ′′ q3
q4

For the shape functions given in eq. (10.69), this last equation evaluates to

Thin-Walled Structures 311


Article 10.10

12 6 12 6 q1
– -----3- ----2- -----3- ----2-
V1 ( z ) L L L L q2
= EI (10.72)
M1 ( z ) 6 12z 4 6z 6 12z 2 6z q3
----- – -------- – --- + ----- – ----- + -------- – --- + -----
L2 L3 L L2 L2 L3 L L2 q4

Since eq. (10.72) relates the internal actions consisting of the transverse shear force and the bending moment to
the nodal displacement vector, it defines the 2X4 stress matrix for the beam element as

12 6 12 6
– -----3- ----2- -----3- ----2-
L L L L
S ( z ) ≡ EI (10.73)
6 12z 4 6z 6 12z 2 6z
----2- – -------3- – --- + ----2- – ----2- + -------3- – --- + ----2-
L L L L L L L L
such that

q1
V1 ( z ) q2
= S(z) 0≤z≤L (10.74)
M1 ( z ) q3
q4

q 1, Q 1 q 3, Q 3 Now consider equilibrium at the nodes z = 0 and z = L. The force in the


positive y-direction Q 1 corresponds to displacement q 1 , the clockwise
q 2, Q 2 z q 4, Q 4
positive moment Q 2 corresponds to the rotation q 2 , the force in the pos-
L
itive y-direction Q 3 corresponds to displacement q 3 , and the clockwise
positive moment Q 4 corresponds to the rotation q 4 . See the adjacent sketch. The free body diagrams at the end
nodes is shown in Fig. 10.18.

Q1 y Q3
V1 ( L )

Q2 M1 ( 0 ) z M1 ( L ) Q4
L
V1 ( 0 ) V1 ( 0 ) V1 ( L )

Fig. 10.18 Free body diagrams of the end nodes of the beam element

Vertical force and moment equilibrium at the nodes results in the relationships between the external and internal
actions as
Q1 = –V1 ( 0 ) Q2 = –M1 ( 0 ) Q3 = V1 ( L ) Q4 = M1 ( L )

From eq. (10.74) these nodal equilibrium equations can be written as

312 Thin-Walled Structures


Structures containing beam elements

q1 q1
Q1 –V1 ( 0 ) q2 Q3 V1 ( L ) q2
= = [ –S ( 0 ) ] = = [S(L)]
Q2 –M1 ( 0 ) q3 Q4 M1 ( L ) q3
q4 q4

Combine these results into one matrix equation to get

Q1 q1
Q2 –S ( 0 ) q2
= or Q = K q
Q3 q3
S(L) 4X1 4X4 4X1
Q4 q4

where the beam element stiffness matrix is defined by

12 6 12 6
-----3- – ----2- – -----3- – ----2-
L L L L
6 4 6 2
– ----2- --- ----2- ---
–S ( 0 ) L L L L
K = which evaluates to K = EI (10.75)
12 6 12 6
S(L) – -----3- ----2- -----3- ----2-
L L L L
6 2 6 4
– ----2- --- ----2- ---
L L L L

The beam element stiffness matrix (10.75) has the following properties:
• It is symmetric, because the material is linear elastic and the displacements and rotations of the beam
are assumed small.
• Its diagonal elements are positive.
• Its determinate vanishes, since the beam element is not restrained against rigid body motion
• Moreover, rigid body motion implies the following relationships between the columns of the unre-
strained stiffness matrix: Columns one and three add to a zero 4X1 column vector. The sum of L/2
times column one, plus column two, minus L/2 times column three, plus column four, add to a zero
4X1 column vector. Note the sum of the columns elements is not equal to zero.
To show the properties in the fourth bullet above, consider the possible rigid body motions of the beam element
under the action of no external nodal forces. Vertical motion at a constant speed is one rigid body mode. The sec-
ond rigid body mode is a rotation about the x-axis at a constant angular speed. These rigid body modes are
depicted in Fig. 10.19.

Thin-Walled Structures 313


Article 10.10

v v L
--- θ x
2
L
θx --- θ x
2

(a) uniform vertical translation (b) uniform clockwise rotation

Fig. 10.19 Rigid body motions of the unrestrained beam element

The displacement vectors for uniform vertical displacement and uniform clockwise rotation are

q1 v q1 L⁄2
q2 q2
= 0 and = 1 θ
x
q3 v q3 –L ⁄ 2
q4 0 q4 1

Let the elements in the 4X4 stiffness matrix for the beam be denoted by k ij , i, j = 1, 2, 3, 4 . For vanishing
external nodal force vectors, substitute the two rigid body nodal displacement vectors into the 4X4 beam element
matrices to get

0 k 11 k 12 k 13 k 14 v k 11 k 13 0
0 = k 21 k 22 k 23 k 24 0
∀v ⇒
k 21
+
k 23
= 0
0 k 31 k 32 k 33 k 34 v k 31 k 33 0
0 k 41 k 42 k 43 k 44 0 k 41 k 43 0

0 k 11 k 12 k 13 k 14 L⁄2 k 11 k 12 k 13
0 k 14
0 = k 21 k 22 k 23 k 24 1 θ L k 21
∀θ x ⇒ --- +
k 22 L k 23
– --- +
k 24
= 0
x
0 k 31 k 32 k 33 k 34 –L ⁄ 2 2 k k 32 2 k 33 k 34 0
31
0 k 41 k 42 k 43 k 44 1 k 41 k 42 k 43 k 44 0

Substituting the values given in eq. (10.75) for the matrix elements k ij in the above equations will show that the
beam element stiffness matrix satisfies the properties described in the fourth bulleted property.

EXAMPLE 10.7 Multi-span beam

Consider a multi-span, uniform beam that is subjected to equal and opposite couples in the y-z plane at z = 0 and
z = L. The magnitude of the moment of these couples is denoted by M a . The bending stiffness EI is the same
constant in each span.
a) Determine the unknown nodal displacements using symmetry to reduce problem size
b) Draw the shear force and bending moment diagrams.

314 Thin-Walled Structures


Structures containing beam elements

c) Determine the support reaction.

y
z
Ma Ma

L L L L

Solution for the unknown nodal displacements The nodes are taken at the support locations and are numbered
one to five from left to right. Hence, there are ten degrees of freedom (DOF’s) as is shown in the top sketch in the
figure below. The support conditions mean the vertical displacements vanish; i.e.,
q1 = q3 = q5 = q7 = q9 = 0

The geometry, boundary conditions, and material properties of the structure are symmetric about the vertical cen-
terline. If the top sketch of the beam and its DOF’s are rotated 180° about this vertical centerline, the bottom
sketch is obtained. The displacements and rotations at the nodes in the top and bottom sketch must be the same.

1 3 5 7 9
2 4 6 8 10
Ma Ma
L L L L

180°
9 7 5 3 1
10 8 6 4 2
Ma Ma
L L L L

Hence, symmetry implies the nodal rotations must satisfy


q 10 = – q 2 q8 = –q4 q6 = –q6

Clearly, the last symmetry condition on the rotations means rotation of the center node vanishes; q 6 = 0 . Then,
the analysis for the response of the beam reduces to a two-span beam, clamped at its right end as is shown below.
1 3 5

2 4 6
Ma

L L

The two active degrees of freedom are rotations q 2 and q 4 . The stiffness matrix for element 1-2 is obtained from

Thin-Walled Structures 315


Article 10.10

eq. (10.75) as
q1 q2 q3 q4 q3 q4 q5 q6

4⁄L 2⁄L 4⁄L


K 1 – 2 = EI K 2 – 3 = EI

2⁄L 4⁄L

The restrained structural stiffness matrix for the symmetric structure is the assemblage of the element 1-2 and 3-
4 stiffness matrices. We get
q2 q4

4⁄L 2⁄L
K αα = EI
2⁄L 8⁄L

The unknown rotations are determined from

Q2 q
= M a = EI 4 ⁄ L 2 ⁄ L 2
Q4 0 2 ⁄ L 8 ⁄ L q4

Solve for the unknown nodal rotations.

q2 L 1 Ma L 4
= ------ ------------------- 8 – 2 M a = -----------
-
q4 EI ( 32 – 4 ) – 2 4 0 14EI – 1

By symmetry the nodal rotation for the entire structure are

q2 4
q4
Ma L –1
q6 = -----------
-
14EI 0
q8 1
q 10 –4

Solution for the shear force and bending moment distributions The shear force and bending moment distri-
bution in beam elements 1-2 and 2-3 are determined from eq. (10.72). For element 1-2, we have

12 6 12 6 q1
– -----3- ----2- -----3- ----2-
V1 ( z ) L L L L q2
= EI
M1 ( z ) 6 12z 4 6z 6 12z 2 6z q3
1–2 ----2- – -------3- – --- + ----2- – ----2- + -------3- – --- + ----2-
L L L L L L L L q4

But q 1 = q 3 = 0 , so

316 Thin-Walled Structures


Structures containing beam elements

6 6
----2- ----2-
V1 ( z ) L L q2
= EI
M1 ( z ) 6z  2 6z q 4
 – --4- + ----
1–2 - – --- + -----
 L L 2  L L 2

Substitute the solution for the nodal rotations to get

6 6 18 9M a
----2- ----2- -----2- ----------
V1 ( z ) L LMa L 4 Ma L L 7L
= EI ------------ = ----------- =
M1 ( z ) 6z  2 6z 14EI – 1
 – --4- + ---- 14 14 z 9M
1–2 - – --- + -----
 L L 2  L L 2 – ------ + 18 ----2- – M a + ---------a-z
L L 7L
Note that the coordinate z is local to the element in the formulas for the shear force and bending moment. For
element 2-3, we have

12 6 12 6 q3
– -----3- ----2- -----3- ----2-
V1 ( z ) L L L L q4
= EI
M1 ( z ) 6 12z 4 6z 6 12z 2 6z q5
2–3 ----- – -------- – --- + ----- – ----- + -------- – --- + -----
L2 L3 L L2 L2 L3 L L2 q6

But q 3 = q 5 = q 6 = 0 , so

6
----2-
V1 ( z ) L
= EI q4
M1 ( z ) 4 6z
2–3 – --- + ----2-
L L
Substitute the solution for rotation q 4 to get

6 3M
----2-
M L – ---------a-
V1 ( z ) L  – -----------
a  7L
= EI - =
M1 ( z ) 4 6z  14EI 2M 3M
2–3 – --- + ----2- ---------a- – ---------a-z
L L 7 7L

Again, note that the z in these formulas for the shear force and bending moment in element 2-3 is local to the
element and runs from zero to L. However, the beginning node 2 corresponds to the global longitudinal coordi-
nate L, and end node 3 corresponds to the global longitudinal coordinate 2L. The relationship between the ele-
ment local coordinate and the global structural coordinate has to be taken into account drawing the shear force
and bending moment diagrams. The shear force and bending moment diagrams are shown in the graph below. In
this example, the shear force diagram in anti-symmetric, and the moment diagram is symmetric, about the center
of the multi-span beam.

Thin-Walled Structures 317


Article 10.10

y
z
Ma Ma

L L L L

9 M M
---
7

VL V
-------
Ma 3
---
7
0
L 2L 3L 4L
3
– ---
7

9
– ---
7

M
-------
Ma
2⁄7
0 2L
–1 ⁄ 7 L 3L 4L

–1

Solution for the support reactions The support reactions can now be obtained by nodal equilibrium. For exam-
ples, free body diagrams of nodes 2 and 3 are shown below.

Q3 Q5
V1 – 2 ( L ) V2 – 3 ( L )
Q4 Q6

M1 – 2 ( L ) M2 – 3 ( 0 ) M2 – 3 ( L ) M3 – 4 ( 0 )

V1 – 2 ( L ) V2 – 3 ( 0 ) V2 – 3 ( 0 ) V2 – 3 ( L ) V3 – 4 ( 0 ) V3 – 4 ( 0 )

Equilibrium at node 2 gives

318 Thin-Walled Structures


Beam element with distributed loading; fixed end actions

9M 3M 12M
Q 3 = V 1 – 2 ( L ) – V 2 – 3 ( 0 ) = ---------a- –  – ---------a- = ------------a-
7L  7L  7L

2M 2M
Q 4 = M 1 – 2 ( L ) – M 2 – 3 ( 0 ) = ---------a- –  ---------a- = 0
7  7 
Equilibrium at node 3 gives

3M 3M 6M
Q 5 = V 2 – 3 ( L ) – V 3 – 4 ( 0 ) = – ---------a- –  ---------a- = – ---------a-
7L  7L  7L

M M
Q 6 = M 2 – 3 ( L ) – M 3 – 4 ( 0 ) = – ------a- –  – ------a- = 0
7  7
All the support reactions are shown in an overall free body diagram of the multi-span beam are shown below.

Ma Ma

9M 12M 6M a 12M 9M
----------a -------------a ---------- -------------a ----------a
7L 7L 7L 7L 7L
L L L L

10.11 Beam element with distributed loading; fixed end actions


For a beam element subjected to a distributed load intensity, the fixed end action vector is computed from the
boundary value problem (10.64) to account for the distributed load in the matrix structural analysis. The total lat-
eral displacement v ( z ) of the beam element consists of the sum of the displacement from the fixed end action
v 0 ( z ) and the displacement from the nodal displacements v 1 ( z ) ; see eq. (10.63). From the free body diagram of
the nodes shown in Fig. 10.18, the total external nodal force vector is

Q1 –V ( 0 ) v′′′ ( 0 ) v 0 ′′′ ( 0 ) v 1 ′′′ ( 0 )


Q2 v ′′ ( 0 ) v 1 ′′ ( 0 )
= – M ( 0 ) = EI v′′ ( 0 ) = EI 0 + EI
Q3 V(L) – v′′′ ( L ) – v 0 ′′′ ( L ) – v 1 ′′′ ( L )
Q4 M(L) – v′′ ( L ) – v 0 ′′ ( L ) – v 1 ′′ ( L )

Let

Q 10 v 0 ′′′ ( 0 ) Q 11 v 1 ′′′ ( 0 )
0 Q 20 v 0 ′′ ( 0 ) 1 Q 21 v 1 ′′ ( 0 )
Q = = EI Q = = EI
Q 30 – v 0 ′′′ ( L ) Q 31 – v 1 ′′′ ( L )
Q 40 – v 0 ′′ ( L ) Q 41 – v 1 ′′ ( L )

0 1
where Q is the 4X1 the nodal force vector from the fixed end action boundary value problem (10.64), and Q is

Thin-Walled Structures 319


Article 10.11

the 4X1 nodal force vector from the boundary value problem (10.65). That is, the total nodal force vector is
0 1
Q = Q + Q . The boundary value problem (10.65) for displacement function v 1 ( z ) was solved in the last sec-
tion, and its solution led to
1
Q = K q

where the 4X4 beam stiffness matrix is given by eq. (10.75) and q is the 4X1 nodal displacement vector. The
superposition of these two boundary value problems is depicted in Fig. 10.20. Hence, the total nodal force vector

py ( z ) py ( z )
Q1 Q3 Q 10 Q 30 Q 11 Q 31

= +
Q2 Q4 Q 20 Q 40 Q 21 Q 41
v0 ( 0 ) = 0 v0 ( L ) = 0 v1 ( 0 ) = q1 v1 ( L ) = q3
–v0 ′ ( 0 ) = 0 –v0 ′ ( L ) = 0 –v1 ′ ( 0 ) = q2 –v1 ′ ( L ) = q4

fixed end actions actual nodal displacements

Fig. 10.20 Superposition of solutions for the beam element response

is given by
0
Q = Q + K q

This equation may be written in the form


0
Q + ( –Q ) = K q (10.76)

0
where the vector – Q is called the equivalent joint force vector. It is the negative of the fixed end action vector.

To summarize, the analysis of a structure composed of beam elements, with some elements subjected to dis-
tributed loads, is as follows:
1. Lock every joint of the structure against translation and rotation, and calculate the fixed end actions.
2. Apply the fixed end actions with the opposite sign.
0
3. Analyze the structure with the applied joint forces and the negative of the fixed end actions; Q + ( – Q ) . Note
that the nodal displacements computed in this step are the actual nodal displacements.
4. Obtain the internal actions consisting of the shear force and bending moment by superposition.

320 Thin-Walled Structures


Beam element with distributed loading; fixed end actions

V ( z ) = EI – v′′′ = EI – v 0 ′′′ + EI – v 1 ′′′


M(z) – v′′ – v 0 ′′ – v 1 ′′

12 6 12 6 q1
– -----3- ----- -----3- ----2-
V ( z ) = EI – v 0 ′′′ + EI L L2 L L q2
M(z) – v 0 ′′ 6 12z 4 6z 6 12z 2 6z q3
----2- – -------3- – --- + ----2- – ----2- + -------3- – --- + ----2-
L L L L L L L L q4

V ( z ) = EI – v 0 ′′′ + S(z) q
M(z) – v 0 ′′






shear force and moment from FEA problem (10.77)

In eq. (10.77) the 2X4 stress matrix S ( z ) was obtained as eq. (10.73) in the last section, and q is the 4X1 nodal
displacement vector of the beam element.

Thin-Walled Structures 321


Article 10.11

Examples of fixed end actions are shown in Fig. 10.21.

0
Q 2i 0
Q 2j
–1 p0 –1
0
Q 2i –1 –p0 L ⁄ 2
0
Q 2i p 0 L 2 ⁄ 12
0
Q 2i 0
Q 2j uniform distributed load =
0
Q 2j –p0 L ⁄ 2
i EI, L j –1
0
Q 2j – p 0 L 2 ⁄ 12

0
Q 2i 0
–1 p 0 Q 2j – 1
0
Q 2i – 3p 0 L ⁄ 20
–1
0
Q 2i p 0 L 2 ⁄ 30
0
Q 2i 0
Q 2j linearly varying load =
0
Q 2j – 7p 0 L ⁄ 20
i EI, L j –1
0
Q 2j – p 0 L 2 ⁄ 20

0
Q 2i F 0
Q 2j
–1 –1
0
Q 2i –1 – Fb 2 ( 3a + b ) ⁄ L 3
0
Q 2i
0
Q 2i 0
Q 2j point force = Fab 2 ⁄ L 2
0
Q 2j – Fa 2 ( a + 3b ) ⁄ L 3
i EI, L j –1
a b 0 – Fa 2 b ⁄ L 2
Q 2j
L

Fig. 10.21 Fixed end actions of a beam element subjected to several lateral loads

EXAMPLE 10.8 Clamped-clamped, stepped beam restrained by a spring

The left half of the beam shown has a uniform flexural stiffness EI , and the right half has a uniform flexural stiff-
ness 2EI . Each half has a length denoted by a . A vertical linear elastic spring of stiffness k = 6EI ⁄ a 3 is con-
nected to midspan. The left half is subjected to a uniform distributed load of intensity p 0 , the right half is
subjected to a point force F at its midspan, and a point force P is applied at the step change stiffness. Model the
response of the beam with two beam elements, one in each half, and a spring element. Determine
a) The restrained structural stiffness matrix
b) The fixed end action vector
c) The unknown nodal displacements
d) The shear force and bending moment distribution in the left half.

322 Thin-Walled Structures


Beam element with distributed loading; fixed end actions

P 1 3 5
p0
F 6
2 4
2EI
EI

k
7

unrestrained structure
a a⁄2 a⁄2
four nodes
seven degrees of freedom

As shown in the figure, the unrestrained structure had four nodes and seven degrees of freedom. The size of the
unrestrained structural stiffness matrix is 7X7. The support conditions impose the vanishing of the following
generalized displacements: q 1 = q 2 = q 5 = q 6 = q 7 = 0 . Hence, the size of the restrained structural stiffness

matrix K αα is 2X2 in degrees of freedom q 3 and q 4 .

Solution of part (a) The general beam element stiffness matrix is given by eq. (10.75). Using this matrix, the
beam element stiffness matrices are
q1 q2 q3 q4 q3 q4 q5 q6

24 ⁄ a 3 – 12 ⁄ a 2
2
K 1 – 2 = EI K2 – 3 = EI – 12 ⁄ a 8⁄a
3
12 ⁄ a 6 ⁄ a 2
6 ⁄ a2 4 ⁄ a

The stiffness matrix for the spring is


q7 q3

1 – 1 = EI 6 ⁄ a 3 – 6 ⁄ a 3
K4 – 2 = k
–1 1 –6 ⁄ a 3 6 ⁄ a 3

Assemblage of the stiffness matrices in degrees of freedom 3 and 4 gives the restrained structural stiffness matrix
as
q3 q4
42 ⁄ a 3 – 6 ⁄ a 2
K αα = EI
– 6 ⁄ a 2 12 ⁄ a

Solution part (b) Lock each node against displacement and rotation. Only the distributed load and the point
force F are considered in the fixed end actions. Point force P is considered an applied load since it acts at node

Thin-Walled Structures 323


Article 10.11

2. The fixed end actions (FEAs) acting on the beam are shown in the following sketch.
Q 10 Q 30 Q 50

p0
Q 40 F

beam FEAs Q 20 Q 60

a a⁄2 a⁄2
From Fig. 10.21, the following free body diagrams of FEAs on the individual beam elements can be drawn:

p0
F

p0 a 2 p0 a 2 Fa Fa
----------- ----------
- ------- -------
12 12 8 8
a⁄2 a⁄2
p0 a p0 a F F
-------
- -------- --- ---
2 2 2 2

From the free body diagrams for each beam element and the external FEAs acting at nodes, the free body dia-
gram of nodes one, two, and three for fixed displacements can be drawn as shown in the next sketch:

Q 30 Q 50
Q 10
Q 40
p0 a2 Q 60
p0 a 2 Fa Fa
Q 20 ----------
- ----------
- ------- -------
12 12 8 8
p0 a F
p0 a -------
- ---
-------- 2 2
2 F
---
2
Node 1 Node 2 Node 3

The force and moment equilibrium at each node then leads to the following fixed end action vector.

Q 10 –p0 a ⁄ 2
Q 20 p 0 a 2 ⁄ 12
Q 30 – p0 a ⁄ 2 – F ⁄ 2
Q 40 = – p a 2 ⁄ 12 + Fa ⁄ 8
0
Q 50 –F ⁄ 2
Q 60 – Fa ⁄ 8
Q 70 0

Solution part (c) The governing matrix equation for the unknown nodal displacements of the restrained struc-
ture is

324 Thin-Walled Structures


Beam element with distributed loading; fixed end actions

Q3 – Q 30 q3
+ = K αα
Q4 – Q 40 q4

The applied nodal force vector, the equivalent joint force vector, and the restrained structural stiffness matrix are

Q3 – Q 30 p0 a ⁄ 2 + F ⁄ 2 42 ⁄ a 3 – 6 ⁄ a 2
= P = K αα = EI
Q4 0 – Q 40 p 0 a 2 ⁄ 12 – Fa ⁄ 8 – 6 ⁄ a 2 12 ⁄ a

Solve for the nodal displacements.

q3 a4 2 P + p0 a ⁄ 2 + F ⁄ 2
= --------------- 12 ⁄ a 6 ⁄ a
q4 468EI 6 ⁄ a 2 42 ⁄ a 3 p a 2 ⁄ 12 – Fa ⁄ 8
0

Perform the matrix multiplication to get the nodal displacements as

a 3 P a 4 p 0 7a 3 F
--------- + ----------- + ------------
q3 1
= ------ 39 72 624
q4 EI 2 3
a P a p0 a 2 F
--------- + ----------- – ---------
78 72 208

Solution part (d) The shear force and bending moment in the left half are determined from

12 6 12 6 q1
– -----3- ----- -----3- -----
V ( z ) = V 0 ( z ) + EI a a2 a a2 q2
M(z) M0 ( z ) 6 12z 4 6z 6 12z 2 6z q3
----2- – -------- – --- + ----- – ----- + -------- – --- + -----
a a3 a a2 a2 a3 a a2 q4

The fixed end actions and the distributed load for the first beam element from Fig. 10.21are shown in the figure
below. Equilibrium differential equations in the fixed end action problem for a typical beam element are

p0
V0 ( 0 )
p0 a 2 p0 a 2 p0 a 2
----------- M0 ( 0 ) ----------- ----------
-
12 12 12
p0 a z
-------- p0 a p0 a
2 -------- -------
-
2 2
z = 0 a

V′ 0 = – p 0 M′ 0 = V 0

Integrate the shear force equation once with respect to z to get


V0 = –p0 z + c1

Thin-Walled Structures 325


Article 10.11

where c 1 is a constant. This constant can be determined from the boundary condition at z = 0 as V 0 ( 0 ) = pa ⁄ 2 ,
so c 1 = pa ⁄ 2 . Hence, V 0 ( z ) = – p 0 z + p 0 a ⁄ 2 . Now the equation for the moment is

M′ 0 = – p 0 z + p 0 a ⁄ 2

Integrate this moment equation to get

p 0 z 2 p 0 az
M 0 = – ---------
- + ----------- + c 2
2 2
The constant c 2 is evaluated from the boundary condition in the fixed end action problem
p0 a 2
M 0 ( 0 ) = – ----------
- = c 2 . Thus the shear force and bending moment in the fixed end action problem for the left
12
half element is

–p0 z + p0 a ⁄ 2
V0 ( z )
= p 0 z 2 p 0 az p 0 a 2 0≤z≤a
M0 ( z ) – ---------
- + ----------- – -----------
2 2 12

The shear force and bending moment in the nodal displacement problem are given by

12 6 12 6 q1
– -----3- ----2- -----3- ----2-
V1 ( z ) a a a a q2
= EI
M1 ( z ) 6 12z 4 6z 6 12z 2 6z q3
----- – -------- – --- + ----- – ----- + -------- – --- + -----
a2 a3 a a2 a2 a3 a a2 q4

But q 1 = q 2 = 0 , and the nodal displacements q 3 and q 4 were determined in part (c). Thus,

12 ⁄ a 3
6
----2- 12 ⁄ a 3
6
----2- a 3 P a 4 p 0 7a 3 F
--------- + ----------- + ------------
V1 ( z ) a q 3 = EI a 1 39 72 624
= EI ------
M1 ( z ) 6 12z 2 6z q 4 6 12z 2 6z EI 2 a 3p
– ----2- + -------
- – --- + ----- – ----2- + -------
- – --- + ----- a P a2F
--------- + ----------0- – ---------
a a3 a a2 a a3 a a2 78 72 208

5 a 11
------ P + --- p 0 + --------- F
V1 ( z ) 13 4 104
=
M1 ( z )  – 7a 5z a 2 az 3a 11z
------ + ------ P +  – ----- + ----- p 0 +  – ------ + --------- F
 39 13  9 4  52 104

The total shear force and bending moment are given by the sum

V ( z ) = V0 ( z ) + V1 ( z )
M(z) M0 ( z ) M1 ( z )

which evaluates to

326 Thin-Walled Structures


Frame structures

5 11
------ P +  3a
------ – z p 0 + --------- F
V(z) = 13 4  104
M(z) 7a 5z
 – ----- 7a 2 3az z 2 3a 11z
- + ------ P +  – -------- + --------- – ---- p 0 +  – ------ + --------- F
 39 13  36 4 2   52 104

10.12 Frame structures


Frame members in a skeletal structure carry applied loads both by axial deformation and bending deformation.
Frames are often modeled by assuming the joints are rigid, which means that members meeting at a joint have the
same rotation. That is, instead of frictionless pins or ball and socket joints used to model trusses, the connections
at a joint under the rigid joint assumption implies that bending moments in the members at the joint do not van-
ish. When distributed lateral loads act on the member, frame elements maybe required even if the joints at the end
of the member are modeled as frictionless pins. In a truss the loads are assumed to only act the joints, and the
members are not subjected to lateral distributed loads. Hence, the stiffness matrix for a frame member is the
superposition of the stiffness matrix for a truss element and the stiffness matrix for a beam element. There are
three degrees of freedom at each end node, or joint, in a plane frame element; two displacements and a rotation as
shown in Fig. 10.22. In this figure, degrees of freedom labeled one and four account for axial deformation,

y, 2 5

3 EA, EI 6
1 4
z
L

Fig. 10.22 Frame element with six degrees of freedom

degrees of freedom two and five account for lateral deformation in bending, and degrees of freedom three and six
account for rotations in bending. These degrees of freedom are referred to Cartesian coordinate directions along
the longitudinal axis and the axis perpendicular to the member. Let the coordinate along longitudinal, centroidal
axis be denoted by z , 0 ≤ z ≤ L , and let y be the coordinate perpendicular to the member.

Consider a typical plane frame element between nodes i and j in a structure. Node i is the beginning node
and node j is the end node, so that z -axis is directed from node i to node j. Then, the 6X1 generalized displace-
ment vector for the frame element in local coordinate directions is uniquely numbered by
T
q = q 31 – 2 q 3i – 1 q 3i q 3j – 2 q 3j – 1 q 3j

These displacement components are shown in Fig. 10.23. The 6X6 frame stiffness matrix in local coordinate
directions is the sum of the truss stiffness matrix and the beam stiffness matrix, where

Thin-Walled Structures 327


Article 10.12

q 3i – 1 q 3j – 1

q 3i q 3j
q 3i – 2 q 3j – 2
i z j

Fig. 10.23 Generalized displacements for a frame element between nodes i and j

q 3i – 1 q 3i q 3j – 1 q 3j
q 3i – 2 q 3j – 2
12EI ⁄ L 3 – 6EI ⁄ L 2 – 12EI ⁄ L 3 – 6EI ⁄ L 2
EA ⁄ L – EA ⁄ L 2 6EI ⁄ L 2
K truss = K beam = EI – 6EI ⁄ L 4EI ⁄ L 2EI ⁄ L
– EA ⁄ L EA ⁄ L – 12EI ⁄ L 3 6EI ⁄ L 2 12EI ⁄ L 3 6EI ⁄ L 2
– 6EI ⁄ L 2 2EI ⁄ L 6EI ⁄ L 2 4EI ⁄ L

Now add these stiffness matrices with due regard to the element locations in the 6X6 frame element stiffness
matrix to get
q 3i – 2 q 3i – 1 q 3i q 3j – 2 q 3j – 1 q 3j

EA ⁄ L 0 0 – EA ⁄ L 0 0
0 12EI ⁄ L 3 – 6EI ⁄ L 2 0 – 12EI ⁄ L 3 – 6EI ⁄ L 2
0 – 6EI ⁄ L 2 4EI ⁄ L 0 6EI ⁄ L 2 2EI ⁄ L
K =
– EA ⁄ L 0 0 EA ⁄ L 0 0
0 – 12EI ⁄ L 3 6EI ⁄ L 2 0 12EI ⁄ L 3 6EI ⁄ L 2
0 – 6EI ⁄ L 2 2EI ⁄ L 0 6EI ⁄ L 2 4EI ⁄ L
(10.78)

The frame element stiffness matrix (10.78) is symmetric and its diagonal elements are positive. Let the 6X1 gen-
eralized nodal force vector corresponding to the generalize displacement vector for the element in local coordi-
nates be denoted by
T
Q = Q 31 – 2 Q 3i – 1 Q 3i Q 3j – 2 Q 3j – 1 Q 3j
Then, the matrix relationship between the generalized force and displacement vectors is

Q = K q
6X1 6X6 6X1 (10.79)
where the frame element stiffness matrix in local coordinate directions is given by eq. (10.78).

The orientation of a frame element in a planar structure can be in a direction that is not horizontal. Therefore,
we need to transform the stiffness matrix referenced to local coordinate directions to the global horizontal and
vertical directions of the overall planar structure. To effect this transformation, consider a frame element whose
longitudinal axis is oriented at an angle θ with respect the horizontal direction. The overall, global structural
coordinate directions are taken to be the Cartesian directions x, y, and z, with the plane of the structure parallel to
the x-y plane. The direction cosines between the local Cartesian directions x, y, z and the global directions x, y, z

328 Thin-Walled Structures


Frame structures

are given in the table below. The nodal displacement vector at node i can be decomposed into components in the

x y z
y
x 0 0 –1 y
z
θ
y – sin θ cos θ 0
θ
x
z cos θ sin θ 0 z, – x

z and y directions, and into components in the x and y directions. Denote the components of this displacement
vector in the global directions be denoted by q 3i – 2 and q 3i – 1 , and denote the rotation by q 3i . These displace-
ment components in the local and global directions at node i are shown in Fig. 10.24. Then, for these global

q 3i – 1
q 3i – 1
q 3i – 2

θ
θ q 3i – 2
node i

Fig. 10.24 Displacement components in the global directions projected in the local directions

direction components projected onto the element local directions we get

q 3i – 2 = q 3i – 2 cos θ + q 3i – 1 cos ( 90° – θ ) = q 3i – 2 cos θ + q 3i – 1 sin θ

q 3i – 1 = q 3i – 2 cos ( 90° + θ ) + q 3i – 1 cos θ = q 3i – 2 ( – sin θ ) + q 3i – 1 sin θ

q 3i = q 3i

Note that the rotation in local coordinate directions is equal to the rotation in the global coordinate direction, both
of which are taken positive clockwise. These transformation equations are written in matrix form as

qi = τ qi
3X1 3X3 3X1

where

q 3i – 2 c s 0 q 3i – 2
qi = q τ = –s c 0 qi = q 3i – 1
3i – 1
0 01 q 3i
q 3i

c = cos θ , and s = sin θ . The 3X3 matrix τ is the matrix of direction cosines. This matrix has the following

Thin-Walled Structures 329


Article 10.12

properties

cos θ – sin θ 0 cos θ sin θ 0 cos2 θ + sin2 θ 0 0 100


T
τ τ = sin θ cos θ 0 – sin θ cos θ 0 = 0 sin2 θ + cos2 θ 0 = 0 1 0 = I
0 0 1 0 0 1 0 0 1 001

and det τ = 1 . Hence, the inverse of matrix τ is equal to its transpose. Matrix τ is said to be an orthogo-
nal matrix. The transformation of the displacements at node j is the same matrix equation as at node i except that
the components of the vectors are those corresponding to node j. Hence, the transformation of the generalized
displacement vector from global to local coordinate directions for frame element i-j can be written in matrix form
as

q 3i – 2 q
c s 0 0 0 0 3i – 2
q 3i – 1 –s c 0 0 0 0 q 3i – 1
q 3i = 0 0 1 0 0 0 q 3i (10.80)
q 3j – 2 0 0 0 c s 0 q 3j – 2
0 0 0 –s c 0 q 3j – 1
q 3j – 1
0 0 0 0 0 1 q
3j
q 3j
Alternatively, we can write the previous equation in the matrix forms

qi = τ 0 qi
q = T q
qj 0 τ qj
6X1 6X6 6X1
6X1 6X6 6X1 (10.81)

where the 6X6 transformation matrix is

c s 0 0 0 0
–s c 0 0 0 0
τ 0 = 0 0 1 0 0 0
T =
0 τ 0 0 0 c s 0
0 0 0 –s c 0
0 0 0 0 0 1

The transformation matrix T is also an orthogonal matrix. That is, its determinate is one and its inverse is equal
to its transpose.

The generalize nodal force vector for frame element i-j transforms form global coordinate directions to local
coordinate directions in the same manner as the generalized displacement vector does for the element. Hence,
from the second of eq. (10.81), the transformation of the 6X1 generalized force vector for element i-j is

Q = T Q (10.82)

330 Thin-Walled Structures


Frame structures

T
where the 6X1 generalized force vector in global directions is Q = Q 3i – 2 Q 3i – 1 Q 3i Q 3j – 2 Q 3j – 1 Q 3j .
Since the 6X6 transformation matrix is orthogonal, the inverse transformation from local to global directions is
T
Q = T Q (10.83)

To obtain the 6X6 frame element stiffness matrix in global coordinate directions, substitute the second of eq.
(10.81) for the generalized displacement vector, and substitute eq. (10.82) for the generalized force vector, into
eq. (10.79) to get

T Q = K T q

T T
Premultiply this equation by T , recognizing that T T = I , to get

T
Q = T K T q = K q
T
The 6X6 matrix K = T K T is the frame element stiffness matrix in global coordinate directions, and is
given by

EA 2 12EI 2  EA 12EI 6EI EA 12EI 6EI


-------c + ------------ s ------- – ------------ cs --------- s –  -------c 2 + ------------ s 2 –  EA 12EI
------- – ------------ cs --------- s
L L3  L L3  L2  L L3   L L3  L2
 EA 12EI EA 2 12EI 2 6EI EA 12EI EA 12EI 6EI
------- – ------------ cs -------s + ------------ c – --------- c  – ------- + ------------ cs –  -------s 2 + ------------ c 2 – --------- c
 L L3  L L3 L2  L L3   L L3  L2
6EI 6EI 4EI 6EI 6EI 2EI
--------- s – --------- c --------- – --------- s --------- c ---------
L2 L2 L L2 L2 L
K = (10.84)
EA 12EI EA 12EI 6EI EA 2 12EI 2  EA 12EI 6EI
–  -------c 2 + ------------ s 2  – ------- + ------------ cs – --------- s -------c + ------------ s ------- – ------------ cs – --------- s
 L L3   L L3  L2 L L3  L L3  L2
EA 12EI EA 12EI 6EI  EA 12EI EA 2 12EI 2 6EI
–  ------- – ------------ cs –  -------s 2 + ------------ c 2 --------- c ------- – ------------ cs -------s + ------------ c --------- c
 L L3   L L3  L2  L L3  L L3 L2
6EI 6EI 2EI 6EI 6EI 4EI
--------- s – --------- c --------- – --------- s --------- c ---------
L2 L2 L L2 L2 L

The frame element i-j referenced to global coordinate directions is shown in Fig. 10.25.
3j – 1

3j Q 3i – 2 q 3i – 2
3j – 2 Q 3i – 1 q 3i – 1
j Q 3i q 3i
3i – 1 = K
y y Q 3j – 2 q 3j – 2
L, EA, EI
3i z Q 3j – 1 q 3j – 1
θ Q 3j q 3j
θ x
3i – 2
i

Fig. 10.25 Frame element with an arbitrary orientation referred to global coordinate directions

Thin-Walled Structures 331


Article 10.12

Stress matrix The stress matrix for the frame element i-j relates the internal axial force N , the transverse shear
force V , and the bending moment M to the generalized nodal displacement vector. We can combine the stress
matrix for the truss element, eq. (10.38), and the stress matrix for the beam element, eq. (10.72), if local coordi-
nate direction displacements are employed. With due regard for the nodal numbering convention for the frame
element relative to the numbering convention of the truss and beam elements, the following relationship can be
obtained from the stress matrices of the truss and beam elements:

q 3i – 2

– EA ⁄ L 0 0 EA ⁄ L 0 0 q 3i – 1
N
0 – 12EI ⁄ L 3 6EI ⁄ L 2 0 12EI ⁄ L 3 6EI ⁄ L 2 q 3i
V =
6 12z 4 6z 6 12z 2 6z q 3j – 2
M i–j 0 EI  ----2- – -------3- EI  – --- + ----2- 0 EI  – ----2- + -------3- EI  – --- + ----2-
L L   L L  L L   L L
i–j q 3j – 1































S(z) q 3j
i–j
(10.85)

Note that the axial coordinate z is a local coordinate in the frame element which is zero at the beginning node i

and equal to the length L of the frame element at end node j. The 3X6 stress matrix S ( z ) in eq. (10.85) is
i–j
referenced to the generalized displacement vector in local coordinate directions. The stress matrix in terms of the
generalized displacement vector in global coordinate directions is obtained by substituting the second of eq.
(10.81) for the displacement vector in eq. (10.85) to get

q 3i – 2 q 3i – 2
q 3i – 1 q 3i – 1
N
q 3i q 3i
V = S(z) T = S(z) 0≤z≤L
i–j
i–j q 3j – 2 i–j q 3j – 2
M i–j
q 3j – 1 q 3j – 1
q 3j q 3j

where

c s 0 0 0 0
– EA ⁄ L 0 0 EA ⁄ L 0 0 –s c 0 0 0 0
0 – 12EI ⁄ L 3 6EI ⁄ L 2 0 12EI ⁄ L 3 6EI ⁄ L 2 0 0 1 0 0 0
S(z) =
i–j 0 0 0 c s 0
6 12z 4 6z 6 12z 2 6z
0 EI  ----- – -------- EI  – --- + ----- 0 EI  – ----- + -------- EI  – --- + -----
 L2 L3   L L 2  L2 L3   L L 2 0 0 0 –s c 0
0 0 0 0 0 1

Perform the matrix multiplication to find

– cEA ⁄ L – sEA ⁄ L 0 cEA ⁄ L sEA ⁄ L 0


s12EI ⁄ L3 – c12EI ⁄ L3 6EI ⁄ L2 – s12EI ⁄ L3 c12EI ⁄ L3 6EI ⁄ L 2
S(z) = (10.86)
i–j
6 12z 6 12z 4 6z 6 12z 6 12z 2 6z
– EIs  ----- – -------- EIc  ----- – -------- EI  – --- + ----- – EIs  – ----- + -------- EIc  – ----- + -------- EI  – --- + -----
 L2 L3   L2 L3   L L 2  L2 L3   L2 L3   L L 2
i–j

332 Thin-Walled Structures


Frame structures

The 3X6 stress matrix S ( z ) , 0 ≤ z ≤ L , for the frame element relates the internal actions in local coordinate
i–j
directions to the generalized displacement vector in global coordinate directions.

EXAMPLE 10.9 A two-bar frame subjected to a distributed load

Consider the frame shown below consisting of a vertical bar 1-2 and a horizontal bar 2-3, which are joined
together by a rigid joint at node 2. The ends of the bars opposite to their joint are clamped. The horizontal bar 1-
2 is subjected to a linearly distributed load. The degree of freedom numbering convention is shown in the second
sketch in the figure.

p0 5 8
6
4 7
2 3
9
L 1, EA 1, EI 1 L 2, EA 2, EI 2 DOF convention
2
1
1
3

a) Determine the restrained structural stiffness matrix K αα .

b) Determine the 9X1 fixed end action vector Q 0 .

c) The following numerical data are specified: the bar


1.25 in.
lengths L 1 = L 2 = 20 in. , Young’s modulus for each
6
bar is E = 10 ×10 psi , and each bar has the same I-sec-
tion whose dimensions are shown in the adjacent sketch.
0.040 in. typical 1.25 in.
Take the load intensity p 0 = 1 lb/in. . Determine the
unknown generalized displacement vector.
d) If the allowable normal stress in tension and compression
is 30,000 psi, what is the largest value of the distributed
load intensity p 0 ?

Solution part (a) The boundary supports restrain the following generalized displacement degrees of freedom to
vanish.
T T
qβ = q1 q2 q3 q7 q8 q9 = 000000

The unknown generalized displacement vector is


T
qα = q4 q5 q6

Hence, the restrained structural stiffness matrix is 3X3 in DOFs 4, 5, 6. Although not requested, the structure of

Thin-Walled Structures 333


Article 10.12

the unrestrained structural stiffness matrix is shown below. Note that the restrained structural stiffness matrix is
in the 3X3 overlap of the member stiffness matrices.
q1 q2 q3 q4 q5 q6 q7 q8 q9

K1 – 2

K =

K2 – 3

For bar 1-2 the angle θ = 90° , so c = 0 and s = 1 in the element stiffness matrix (10.84). The contributions
of the elements 4, 5, 6 in the bar 1-2 stiffness matrix are
q1 q2 q3 q4 q5 q6

K1 – 2 =
12 ( EI ⁄ L 3 ) 1 0 – 6 ( EI ⁄ L 2 ) 1
0 ( EA ⁄ L ) 1 0
– 6 ( EI ⁄ L2 ) 1 0 4 ( EI ⁄ L ) 1

For bar 2-3 the angle θ = 0° , so c = 1 and s = 0 in the element stiffness matrix (10.84). The contributions of
the elements 4, 5, 6 in the bar 2-3 stiffness matrix are
q4 q5 q6 q7 q8 q9

( EA ⁄ L ) 2 0 0
0 12 ( EI ⁄ L 3 ) 2 – 6 ( EI ⁄ L 2 ) 2

K2 – 3 = 0 – 6 ( EI ⁄ L 2 ) 2 4 ( EI ⁄ L ) 2

Now we assemble the individual stiffness matrices in degrees of freedom 4, 5, 6 to get the restrained structural
stiffness matrix as

334 Thin-Walled Structures


Frame structures

q4 q5 q6

12 ( EI ⁄ L 3 ) 1 + ( EA ⁄ L ) 2 0 – 6 ( EI ⁄ L 2 ) 1
K αα = 0 ( EA ⁄ L ) 1 + 12 ( EI ⁄ L 3 ) 2 – 6 ( EI ⁄ L 2 ) 2
– 6 ( EI ⁄ L 2 ) 1 – 6 ( EI ⁄ L 2 ) 2 4 ( EI ⁄ L ) 1 + 4 ( EI ⁄ L ) 2

Solution part (b) The fixed end action vector is due to the distributed load on bar 2-3. The support reactions
required to equilibrate the distributed load on bar 2-3, which is acting downward, can be deduced from second
free body diagram in Fig. 10.21. These support reactions are shown in the free body diagram below.
py
p0
7p 0 L 2 3p 0 L 2 V
--------------
- --------------
-
20 20
N N
p o L 22 p 0 L 22
------------ ------------ M M
20 z 30 V
L2
positive element

From the free body diagram of bar 2-3 above, we can draw the free body diagrams of the three nodes. These dia-
grams are shown below.

Q 50 Q 80

Q 60 node 3 Q 90
node 2 Q 40 Q 70

7p 0 L 2 p 0 L 22 p 0 L 22 3p 0 L 2
--------------
- ------------ -----------
- --------------
-
20 20 30 20
Q 20

Q 30
node 1 Q 10

Equilibrium at the nodes gives the 9X1 fixed end action vector as
T
0 7p 0 L 2 p 0 L 22 3p 0 L 2 p 0 L 22
Q = 0 0 0 0 -------------
- – ----------- 0 -------------- -----------
20 20 20 30

Solution part (c) The governing matrix structural equation for the unknown generalized displacements is

Q4 q4– Q 40
Q5 + 0
–Q5 = K αα q 5
Q6 – Q 60 q6

where the applied generalized force vector Q 4 Q 5 Q 6 = 0 0 0 , the fixed end action vector was determined

Thin-Walled Structures 335


Article 10.12

in part (b), and where the restrained structural stiffness matrix was determined in part (a). For identical bars, the
above matrix equation reduces to

0
7p 0 L 12EI ⁄ L 3 + EA ⁄ L 0 – 6EI ⁄ L 2 q 4
– ------------ =
20 0 EA ⁄ L + 12EI ⁄ L 3 – 6EI ⁄ L 2 q 5
p0 L 2 – 6EI ⁄ L 2 – 6EI ⁄ L 2 8EI ⁄ L q 6
-----------
20
From the given numerical data for the cross section, we compute the area and second area moment about the cen-
troidal horizontal axis of the I-section as follows:

A = 3 ( 1.25 ) ( 0.040 ) = 0.150 in. 2


I = 2 [ ( 1.25 ⁄ 2 ) 2 ( 1.25 ) ( 0.040 ) ] + ( 0.040 ( 1.25 ) 3 ) ⁄ 12 = 0.0456 in. 4





















two flanges web

6
For each bar the extensional stiffness EA = 1.5 ×10 lb. , and the flexural stiffness EI = 455729. lb-in. 2 . The
numerical values of the elements in the restrained structural stiffness matrix are
12EI EA 6EI 8EI
------------ + ------- = 683.59 + 75000 = 75683.59 lb/in. --------- = 6835.938 lb. --------- = 182291.67 lb-in.
L3 L L2 L

The matrix equation for the unknown generalized displacements becomes

0 75683.59 0 – 6835.938 q 4
– 7 lb. = 0 75683.59 – 6835.938 q 5
20 lb-in. – 6835.938 – 6835.938 182291.67 q 6

The solution to this linear set of equations gives

–6
q4 9.6617 ×10 in.
q 5 = – 8.2828 ×10 in. –5

q6 –4
1.0691 ×10 rad.

Solution part (d) The maximum normal stress magnitude at each cross section is either at the top or bottom of
the flanges where y = ± c , and c = 0.625 in. . These normal stresses are equal to the sum of the extensional
stress and bending stress N ⁄ A + M ( ± c ) ⁄ I , where N > 0 in tension and M > 0 clockwise on a positive z cross
section.

For bar 1-2, where c = 0 and s = 1 , the stress matrix (10.86) reduces to

0 – EA ⁄ L 0 0 EA ⁄ L 0
12EI ⁄ L3 0 6EI ⁄ L2 – 12EI ⁄ L3 0 6EI ⁄ L 2
S(z) = 0≤z≤L
1–2
6 12z 4 6z 6 12z 2 6z
– EI  ----- – -------- 0 EI  – --- + ----- – EI  – ----- + -------- 0 EI  – --- + -----
 L2 L3   L L 2  L2 L3   L L 2

336 Thin-Walled Structures


Frame structures

Since node 1 is fixed, q 1 = q 2 = q 3 = 0 , and the generalized displacements at node 2 were determined in part
(c) for p 0 = 1 lb/in. . Hence,

0 EA ⁄ L 0
N q4
– 12EI ⁄ L 3 0 6EI ⁄ L 2
V = q5
6 12z 2 6z
M 1–2 – EI  – ----2- + -------3- 0 EI  – --- + ----2- q 6
 L L   L L

Substitute numerical values for the quantities in this equation to get

–6
N 0 75000 0 9.6617 ×10
V = – 683.59 0 6835.938 – 8.2828 ×10
–5

M 1–2 6835.938 – 683.59z 0 – 45572.92 + 6835.938z 1.0691 ×10–4 .

N – 6.2121 lb.
V = 0.7242 lb. 0 ≤ z ≤ 20 in.
M 1–2 ( – 4.8062 + 0.7242z ) lb-in.
The maximum bending moment in bar 1-2 occurs at z = 20 in. , or the joint with bar 2-3, and is equal to
9.6778 lb-in. . Since bar 1-2 is in axial compression, the largest normal stress is compressive and occurs at the
bottom flange where y = – 0.625 in. . This compressive normal stress is

– 6.2121 ( 9.6778 ( – 0.625 ) )


σz = ------------------- + ------------------------------------------- = – 41.414 – 132.6459 = – 174.06psi
1–2 0.150 0.0456
This normal stress is for p 0 = 1 lb/in. . The stresses are proportional to the applied load intensity, so at the
allowable stress bar 1-2 is critical for the distributed load intensity
p0
 -------------- 1
- = ---------------- → p 0 = 172.36 lb/in.
 30000 174.06

For bar 2-3 the internal actions are the superposition of the fixed end action problem and the nodal displace-
ment problem. The equilibrium differential equations for the bar are

dN dV z dM
------- = 0 ------- – p 0  1 – --- = 0 -------- – V = 0 0≤z≤L
 L 
dz dz dz
where the positive internal actions are shown in the free body diagram of bar 2-3 in part (b). From this free body
diagram of part (b) we have the boundary conditions

7p 0 L p0 L 2
N(0) = 0 V ( 0 ) = – -----------
- M ( 0 ) = ----------
-
20 20
The solutions of the differential equations of equilibrium subject to the boundary conditions at z = 0 for the
fixed end action problem in matrix form are

Thin-Walled Structures 337


Article 10.12

0
N0 p0 L  2
3p 0 L
= - 1 – --z- + -----------
– -------- - 0≤z≤L
V0 2  L  20
M0 3
p0 L 2  3p 0 L 2  p0 L 2
2–3 ----------- 1 – --z- – -------------
- 1 – --z- + ----------
-
6  L 20  L 30
Substitute the numerical values in this last equation to get

0
N0 z
2

= – 10  1 – ------ + 3 0 ≤ z ≤ 20
V0  20
M0 3
z z
2–3 66.667  1 – ------ – 60  1 – ------ + 13.333
 20  20

The internal actions in bar 2-3 for the nodal displacement problem are determined via the stress matrix
(10.86) with c = 1 and s = 0 . The stress matrix reduces to

– EA ⁄ L 0 0 EA ⁄ L 0 0
0 – 12EI ⁄ L 3 6EI ⁄ L 2 0 12EI ⁄ L 3 6EI ⁄ L 2
S(z) =
2–3
6 12z 4 6z 6 12z 2 6z
– EI  ----- – -------- 0 EI  – --- + ----- 0 EI  – ----- + -------- EI  – --- + -----
 L2 L3   L L 2  L2 L3   L L 2

The support conditions at z = L for bar 2-3 mean the generalized displacements q 7 = q 8 = q 9 = 0 . Hence,
the internal actions are determined by the first three columns of the stress matrix, and we have

– EA ⁄ L 0 0
N1 q4
0 – 12EI ⁄ L 3 6EI ⁄ L 2
V1 = q5
 6 12z  4 6z
M1 –0 EI ----2- – -------3- EI – --- + ----2- q 6
2–3 L L   L L

Substitute the numerical value for the elements in the stress matrix and the solution for the generalize displace-
ment vector into the above equation to get

–6
N1 – 75000 0 0 9.6617 ×10
V1 = 0 – 683.5938 6835.938 – 8.2828 ×10
–5

M1 2–3
0 ( 6835.938 – 683.5938z ) ( – 91145.83 + 6835.938z ) 1.0691 ×10–4 .

N1 – 0.7246275
V1 = 0.787451 0 ≤ z ≤ 20 in.
M1 2–3
– 10.3106 + 0.787451z

The total internal action vector is given by

338 Thin-Walled Structures


Reference

N N0 N1
V = V0 + V1
M 2–3 M0 M1
2–3 2–3

Substitute the results from the fixed end action problem and the nodal displacement problem to get

– 0.7246275
2
N z
= – 10  1 – ------ + 3.787451
V  20
M 3
2–3 z z
66.667  1 – ------ – 60  1 – ------ + 0.787451z + 3.022730
 20  20
Extremum values of the bending moment occur at the end points of the bar 2-3 and where the shear force van-
ishes. The bending moment at z = 0 is 9.6897 lb-in. and the bending moment at z = 20 is 18.7718 lb-in. The
shear force vanishes at

z
 1 – ----- 3.787451
- = ---------------------- = 0.6154227 or z = 7.6915 in.
 20 10
The bending moment at the location of vanishing shear is
M 2 – 3 ( 7.6915 ) = 66.667 ( 0.6154227 ) 3 – 60 ( 0.6154227 ) + 0.787451 ( 7.6915 ) + 3.022730
M 2 – 3 ( 7.6915 ) = – 12.306653 lb-in.

Therefore, the maximum magnitude of the bending moment is 18.7718 lb-in at z = 20 in. Since the axial force
in bar 2-3 is compressive, the maximum compressive normal stress occurs at y = – 0.625 in.

– 0.7246275 ( 18.7718 ( – 0.625 ) )


σz = ---------------------------- + ---------------------------------------------- = – 262.11978 psi
2–3 0.150 0.0456
This normal stress is for p 0 = 1 lb/in. . The stresses are proportional to the applied load intensity, so at the
allowable stress bar 2-3 is critical for the distributed load intensity
p0
 -------------- 1
- = ---------------- → p 0 = 114.45 lb/in.
 30000 262.12

Since the maximum value of p 0 for bar 2-3 is less than the maximum value computed for bar 1-2, the largest
value of p 0 for the structure not to exceed the allowable normal stress magnitude in tension or compression is
114.12 lb/in.

10.13 Reference
Martin, Harold C., Introduction to Matrix Methods of Structural Analysis, McGraw-Hill Book Company,
New York, 1966.

Thin-Walled Structures 339


Article 10.14

10.14 Problems
1. For the spring assembly shown below, determine the 2X2 flexibility influence matrix [ C ] by the method of
unit action states.

kb

ka

kc

1 2

2. Consider a three-degree-of-freedom model of the string in tension shown below. Let T denote the horizontal
component of the tension force. The three degrees of freedom are the vertical displacements and corresponding
forces at the quarter points. The analysis of this structure is different from the analyses we have been using in that
we have to take equilibrium on a slightly deflected configuration rather than on the undeformed configuration
even though the displacements are small. A typical free body diagram to be used in the analysis is shown in the
figure. Note that it is the horizontal component of the string tension that is equal to T .

Q
1 2 3

θ1 θ2
T T
a a a a T T
-------------- --------------
cos θ 1 cos θ 2
string-in-tension
typical FBD

a) Use unit action states and the physical definition of flexibility influence coefficients to calculate the 3X3
flexibility matrix [ C ] . Write the elements of the matrix in terms of tension T and dimension a . Recall
3 a
the vertical displacements are assumed small compared to length a . Partial answer: c 11 = ---  --- .
4  T
b) Use unit displacement states and the physical definition of stiffness matrix elements to calculate the 3X3
T
stiffness matrix [ K ] . Partial answer: k 11 = 2  --- .
 a

c) Check the plausibility of the matrices determined in parts a and b. Are they symmetric? Are diagonal
elements positive? Does [ K ] [ C ] = [ I ] ?

340 Thin-Walled Structures


Problems

3. Derive by the method of unit displacement states the 3X3 stiffness matrix [ K ] for the structure shown.
Assume small displacements and rotations of the horizontal rigid bar. Partial answer: k 21 = 2k ⁄ 9 .

2k 2k

L⁄3 2L ⁄ 3

rigid bar

1 k 2

4. Consider the plane truss restrained against rigid body motion and subject to the loads shown in the figure.

3 6
5

8
4 7
100in.

2 4
5kips 45° 30° 3
1 2 1
10kips DOF convention
Four-noded truss

Use the degree-of-freedom numbering convention based on the node numbering as shown in the figure. All four
6
bars have the same modulus of elasticity E = 10 ×10 psi , and the same A ⁄ L where the cross sectional area for
bar 1-2 is 0.5 in. 2 Solve by hand the computations for the
a) unrestrained structural stiffness matrix
b) restrained structural stiffness matrix
c) unknown nodal displacements.

Thin-Walled Structures 341


Article 10.14

5. Consider the plane truss consisting of five bars shown below. Each bar has the same extensional stiffness
EA . Use the degree of freedom numbering convention based on the node numbers labeled in the figure.

4 3

15° 15°
h
y
x 45° 45°

1 2

a) Determine the unrestrained structural stiffness matrix.


b) Nodes 1 and 3 are restrained such that q 1 = q 2 = q 6 = 0 , and assumed the loads are applied in the

remaining degrees of freedom. Determine the submatrix K βα .

6. For the seven bar truss shown below all bars have the same value for EA ⁄ L . The horizontal displacement of
node 5 is prescribed as q 9 = 1 . All applied forces are zero. Use symmetry to reduce the order of the restrained

structural stiffness matrix K αα and then determine the unknown nodal displacements q 3, q 4, q 5, q 6 .

4 3
30° 60°

1 2
L

342 Thin-Walled Structures


Problems

7. In the three bar truss shown below the temperature of bar 1-2 is increased 100°C above ambient tempera-
ture, while bars 1-3 and 1-4 remain at ambient temperature. The bars are made of aluminum alloy with a modulus
–6
of elasticity E = 69GPa and coefficient of thermal expansion α = 23.6 ×10 ⁄ °C . The length of each bar
L = 250mm , and cross-sectional area of each bar A = 400mm 2 .

L
y q2
x
1 q1

L 45° 45° L

3 2

Determine
a) the 8X1 fixed-end action vector
b) the 8X8 unrestrained structural stiffness matrix
c) the nodal displacements q 1 and q 2 of movable node 1

d) the support reactions


e) the bar forces, and state if they are in tension or compression.

8. The uniform, multi-span beam shown is clamped at each end and subjected to vertical point loads at nodes 2
and 4. Use the node numbers indicated in the figure, and the degree of freedom numbering convention associated
with the node numbers.

P P

1 2 4 5

L L L L

a) Use symmetry to reduce the problem size and compute the nodal displacement vector in terms of P, L,
and EI.
b) Determine the shear force and bending moment distributions in each span in terms of P and L. Sketch
the shear force and bending moment diagrams.
c) Determine the support reactions.

Thin-Walled Structures 343


Article 10.14

9. Determine the restrained structural stiffness matrix in problem 8 if node 5 is simply supported.

10. The flexural stiffness of the uniform beam shown below is EI , and it has a of length 2L. It is supported by
linear elastic springs at each end, each with a stiffness k = 6EI ⁄ L 3 . It is subjected to the linearly varying dis-
tributed load whose intensity is p 0 at midspan.

p0
1 3 5

2 4 6

EI EI, L EI, L
k k
DOFs

L L

Use two elements to model the beam, and use the degrees-of-freedom (DOFs) shown in the figure.
a) Use symmetry about the vertical centerline and determine the restrained structural stiffness matrix in
DOFs 1, 2, and 3, and in terms of parameters EI and L .
0
b) Determine the 6X1 fixed end action vector Q in terms of p 0 and L .
T
c) Solve for the unknown nodal displacement vector q 1 q 2 q 3 q 4 q 5 q 6 in terms of EI , L , and p 0 .

11. Consider the wing spar braced by a strut with the spar subjected to a linearly distributed load to approximate
the spanwise airload acting on a wing. The spar is clamped at the root, and the cross section of the spar is the
same I-section as given in part (c) of Example 10.9. The strut is pin-connected to the spar and its support at the
fuselage. The structure is modeled with a frame element between nodes 1 and 2, a beam element between nodes
2 and 3, and a truss element between nodes 4 and 2. The following data is specified: b = 20 in. , h = 4 in. ,
6
p 0 = 30 lb/in. , cross sectional area of the strut A b = 0.02 in. 2 , and Young’s modulus E = 10 ×10 psi for both
the spar and strut.

p0

2 5 7

3 6 8
1 4
node 1 node 2 node 3
h 10
DOFs
node 4 9
L a

344 Thin-Walled Structures


Problems

a) Determine the 5X5 restrained structural stiffness matrix containing L as a parameter in the matrix ele-
ments. Partial answer: For L = 15 in. the restrained structural stiffness matrix is

˙
112028. 3207.42 0 0 0
3207.42 46225.7 – 97222.2 – 43750. – 109375.
0 – 97222.2 486111. 109375. 182292.
0 – 43750. 109375. 43750. 109375.
0 – 109375. 182292. 109375. 364583.

b) Determine the 5X1 fixed end action vector in the unknown degrees of freedom in terms of parameter L.
To aid in this determination, the support reactions for a clamped-clamped beam subjected to a linearly
varying load is given in the free body diagram below. Partial answer: For L = 15 in. the fixed end
action vector is
T
Q 0 = 0 – 120. – 300. – 5.625 – 6.25

pa pb

L2 L2 z z
------ ( 3p a + 2p b ) ------ ( 2p a + 3p b )
60 p y = p a  1 – --- + p b ---
60 z L  L L

L L
------ ( 7p a + 3p b ) ------ ( 3p a + 7p b )
20 20

c) Determine the compressive normal stress in the top flange at the root of the spar for the values of length
L listed in the table below. Note: if the strut is absent, then the spar is a cantilevered beam and the mag-

Thin-Walled Structures 345


Article 10.14

nitude of the compressive normal stress in the top flange at the root is 27,429 psi.

L, inches σ , psi
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15 -15,880.
16
17
18
19

346 Thin-Walled Structures


CHAPTER 11 Buckling

Buckling of a structure means


• failure due to excessive displacements (loss of structural stiffness), and/or
• loss of stability of an equilibrium configuration of the structure

Stability of equilibrium means that the response of the structure due to a small disturbance from its equilib-
rium configuration 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 infinite, then
the equilibrium state is unstable. Practical structures in engineering are stable at no load. Now consider increas-
ing 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 strength 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). Many 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.

11.1 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. 11.1 and has one coordinate θ, – π < θ < π , to describe the configuration 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 θ = 0. Neglect the weight of the rod with respect to the applied load P.

Aerospace Structures 347


Article 11.1

P P

θ θ
l l
l
FBD
K Kθ
Ox
initial deflected Oy

Fig. 11.1 One degree of freedom structural model

From the free body diagram of the rod shown in Fig. 11.1, the equation of motion for rotation about the fixed
hinge is
2

Pl sin θ – Kθ = I 0 θ = θ(t) t>0 (11.1)
dt2
where I0 is the moment of inertia of the rod about the fixed point and t is time.

11.1.1 Static equilibrium

Consider equilibrium states under the static, downward load P which are characterized by the angle θ being
independent of time t. Hence, the inertia term in eq. (11.1) vanishes and we have
Pl sin θ – Kθ = 0 θ <π (11.2)

The solutions to eq. (11.2) are


P 1 : θ = 0 for any P (11.3)

and

K θ
P 2 : P =  ---- ----------- (11.4)
 l  sin θ

Recall from the calculus using l’Hôpital’s rule that the limit of the
P
--------- indeterminate form θ ⁄ ( sin θ ) as θ → 0 is one. The two equilib-
K⁄l
rium paths are plotted in the load-deflection diagram shown in Fig.
2
11.2. Equilibrium path P1 coincides with the load axis in the plot
P1 and is called the primary equilibrium path, or the trivial equilibrium
P2 P2 path. Equilibrium path P2 is called the secondary path and we note
1 it is symmetric about θ = 0. The two equilibrium paths intersect at
(θ,P) = (0,K/l). This intersection of the two paths is called a bifur-
P1
cation point. At no load the rod is vertical and this corresponds to
the origin in the load-deflection diagram. As the load P is slowly
–π 0 π θ increased from zero the rod remains vertical (θ = 0), and at P = K/l
Fig. 11.2 Equilibrium paths. adjacent equilibrium states exists on the secondary path. The exist-
ence of adjacent equilibrium states in the vicinity of the primary

348 Aerospace Structures


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-deflection 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 infinitesimal disturbances present (there always are), but will rotate either to the left or right
depending on type of infinitesimal disturbance. We note that the magnitude of the angle θ 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 P cr . Thus,

P cr = K ⁄ l (11.5)

Small θ analysis Consider the small angles of rotation such that sin θ ≈ θ for θ measured in radians. Equilib-
rium eq. (11.2) becomes
Plθ – Kθ = 0 (11.6)

And the solutions of this equation are


P 1 ′: θ = 0 for any P , and (11.7)

P 2 ′: P = K ⁄ l for any small θ (11.8)

These solutions are shown in the load-deflection plane in Fig. 11.3. The equilibrium P
---------
path P 1 ′ coincides with path P 1 , but path P 2′ is not a good approximation to path K⁄l
P2 ′
P 2 unless θ is very small. However, the bifurcation point is the same as obtained in
1
the large θ-analysis. Hence, the critical load from the small θ-analysis is the same as
obtained in eq. (11.5) from the large θ-analysis. P1 ′

0 θ
11.1.2 Stability analysis Fig. 11.3 Small θ analysis

Let the rotation angle


θ ( t ) = θ0 + ϕ ( t ) (11.9)

where θ 0 is independent of time and satisfies the equilibrium eq. (11.2); i.e.,

Pl sin θ 0 – Kθ 0 = 0 (11.10)

Consider the additional rotation angle ϕ ( t ) to be small in magnitude but a function P


of time. Thus, we are considering small oscillations about an equilibrium state
θ(t)
( P, θ 0 ) as shown in Fig. 11.4. Substitute eq. (11.9) for θ ( t ) in the equation of ϕ(t)
motion, eq. (11.1), to get
θ0
I 0 ϕ̇˙ + K ( θ 0 + ϕ ) – Pl sin ( θ 0 + ϕ ) = 0 (11.11)
l
2

where the dots denote derivatives with respect to time; e.g., ϕ̇˙ = 2 . Using the
dt
trigonometric identity for the sine of the sum of two angles, and performing some Fig. 11.4 Rotations in
minor rearrangements, the last equation becomes the stability analysis

I 0 ϕ̇˙ + Kθ 0 + Kϕ – Pl [ sin θ 0 cos ϕ + cos θ 0 sin ϕ ] = 0 (11.12)

Aerospace Structures 349


Article 11.1

Now expand the trigonometric functions of angle ϕ in a Taylor Series about ϕ = 0 to get

1 1
I 0 ϕ̇˙ + Kθ 0 + Kϕ – Pl sin θ 0 1 – --- ϕ 2 + O ( ϕ 4 ) – Pl cos θ 0 ϕ – --- ϕ 3 + O ( ϕ 5 ) = 0 (11.13)
2 6

in which O ( ϕ n ) means terms of order ϕ n and higher. Arrange eq. (11.13) in powers of ϕ to get

Pl Pl
I 0 ϕ̇˙ + ( Kθ 0 – Pl sin θ 0 ) + ( K – Pl cos θ 0 )ϕ +  ----- sin θ 0 ϕ 2 +  ----- cos θ 0 ϕ 3 + O ( ϕ 4 ) = 0
2  6 






= 0 (11.14)

Note that “coefficient” of the term ϕ 0 vanishes because of the equilibrium condition given by eq. (11.10).

For very small additional rotation angles ϕ ( t ) about the equilibrium configuration, eq. (11.14) is approxi-
mated by

I 0 ϕ̇˙ + ( K – Pl cos θ 0 )ϕ = 0 or ϕ̇˙ + ω 2 ϕ = 0 (11.15)

where

ω 2 = ( K – Pl cos θ 0 ) ⁄ I 0 (11.16)

The solution of the second order differential equation, eq. (11.15), for ω 2 > 0 is

ϕ ( t ) = A 1 sin ( ωt ) + A 2 cos ( ωt ) ω2 > 0 (11.17)

in which constants A 1 and A 2 are determined by initial conditions for ϕ ( 0 ) and ϕ̇ ( 0 ) . The solution given by
eq. (11.17) is a harmonic oscillation about the equilibrium configuration and ω is interpreted as the natural fre-
quency in radians per second. Initial conditions ϕ ( 0 ) and ϕ̇ ( 0 ) 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 ϕ . Thus, ω 2 > 0 is a condition for a stable equilibrium configuration with respect
to infinitesimal disturbances.

The solution of the second order differential equation, eq. (11.15), for ω 2 < 0 is

ϕ ( t ) = A 1 e ωt + A 2 e –ωt ω 2 = – ω 2 = – ( K – Pl cos θ 0 ) ⁄ I 0 (11.18)

For arbitrary initial conditions, the term with the positive exponent in the dominates the solution. This corre-
sponds to large values of the ϕ no matter how small the initial disturbance. Hence, ω 2 < 0 is a condition of
unstable equilibrium configuration with respect to infinitesimal disturbances. The dynamic criterion for structural
stability is

Dynamic criterion for stability of an equilibrium state


The equilibrium state is stable if ω2 > 0
The equilibrium state is critical if ω2 = 0
The equilibrium state is unstable if ω2 < 0

On the primary equilibrium path P 1 given by eq. (11.3), we have from eq. (11.16) that

350 Aerospace Structures


Perfect Columns

ω 2 = ( K – Pl ) ⁄ I 0 on P 1 (11.19)

Thus, equilibrium configurations are stable if P < K ⁄ l , critical if P = K ⁄ l , and unstable if P > K ⁄ l . The pri-
mary equilibrium path ceases to be stable at P = P cr , and P cr is the buckling load.

11.2 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. 11.5. 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
configuration of this column is pure compression. Let N(z) denote the internal axial force. From equilibrium of

y,v

P
z,w

Fig. 11.5 A straight column subjected to a centric, compressive axial force.

a differential element shown below we have dN ⁄ dz = 0 , and from Hooke’s law N = EAε z where ε z is axial
normal strain. The axial normal strain is related to the axial displacement by ε z = dw ⁄ dz as is shown in Fig.
11.6. The boundary conditions for the column are w ( 0 ) = 0 and N ( L ) = – P . Hence, the internal axial load is

dw
y dz + w ( z + dx ) – w ( z ) ≈  1 +  dz
 dz 

z N N + dN
z dz w ( z + dz )
w(z)

Fig. 11.6 An element of the column in the pre-buckling equilibrium state

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

Pz
N = –P w ( z ) = – ------- v(z) = 0 0≤z≤L (11.20)
EA

Aerospace Structures 351


Article 11.2

The end shortening under the compressive load is – w ( L ) , and this is plotted on the load-end shortening plot
shown in Fig. 11.7. The equilibrium configuration of pure compression of the perfect column is called the trivial

Fig. 11.7 Load-end shortening EA


plot in pre-buckling -------
L
1

0 –w ( L )

equilibrium state. Note that in the trivial equilibrium state the lateral displacement of the column, v ( z ) 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 configuration 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 configura-
tions?

To answer this question we consider equilibrium of a slightly deflected element of the column at the same
value of the external load P. The free body diagram of this element is shown in Fig. 11.8. The displacement due

 1 + dw
------- dz
 dz  P V y1 + dV y1
θ x1 dv
– -------1- dz
y dz M x1 P M x1 + dM x1
v1 dv V y1
v 1 + -------1- ( dz )
dz
z

Fig. 11.8 Free body diagram of an element of the column in the buckled state.

to buckling is denoted by v1(z), and all quantities due to buckling are labeled with the subscript 1. Vertical force
equilibrium gives
dV y1
----------- = 0 (11.21)
dz
where V y1 is the y-direction shear force due to buckling. Moment equilibrium about the x-axis gives

dM x1  dw dv
------------ – 1 + ------- V y1 +  – -------1- P = 0
dz  dz   dz 

where M x1 is the bending moment due to buckling. In general, the axial strain in the equilibrium configuration
dw ⁄ dz is very small in magnitude compared to unity and is then neglected with respect to unity in this equation.

352 Aerospace Structures


Perfect Columns

Then, the moment equation becomes

dM x1 dv
------------ – V y1 +  – -------1- P = 0 (11.22)
dz  dz 

Neglecting dw ⁄ dz with respect to unity, and assuming the rotation θ x1 due to buckling is small, implies that the
rotation is given by θ x1 ( z ) = – ( dv 1 ⁄ dz ) . A very important term in eq. (11.22) is the contribution of the axial
compressive load P through the buckling displacement v1 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 deflected column element. Hooke’s law for the bending moment is
2
 dv
M x1 = EI xx  – 21 (11.23)
 dz 
where Ixx 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 area moment of the cross section). If we take the derivative of eq.
(11.22), use eq. (11.21), and then substitute eq. (11.23) for the bending moment due to buckling we get
2 2
2
d  d v 1  d v 1
 – EI xx  + – P = 0 (11.24)
dz2  dz2   dz2 

This is the governing fourth order, ordinary differential equation for the buckling displacement v1(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 v1 by
substituting eq. (11.23) into eq. (11.22) to get
2
d  d v 1 dv 1
V y1 = –  EI  – -------- P (11.25)
d z  d z 2  dz

To determine the buckling displacement v1 we need boundary conditions at z = 0 and z =L in addition to the
o.d.e. given by eq. (11.24). There are four standard boundary conditions. These are

A. Pinned-pinned

v1 ( 0 ) = 0 v1 ( L ) = 0 P
M x1 ( 0 ) = 0 M x1 ( L ) = 0 z
L

B. Clamped-free
P
v1 ( 0 ) = 0 M x1 ( L ) = 0
θ x1 ( 0 ) = 0 V y1 ( L ) = 0 z
L

Aerospace Structures 353


Article 11.2

C. Clamped-clamped
v1 ( 0 ) = 0 v1 ( L ) = 0 P
θ x1 ( 0 ) = 0 θ x1 ( L ) = 0 z
L

D. Clamped-pinned
v1 ( 0 ) = 0 v1 ( L ) = 0 P
θ x1 ( 0 ) = 0 M x1 ( L ) = 0 z
L

One solution to the o.d.e., eq. (11.24), subject to boundary conditions A-D is v 1 ( z ) = 0 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. (11.24) becomes

4 2
dv dv
EI 41 + P 21 = 0
dz dz
or
4 2
d v1 dv
4
+ k 2 21 = 0 0<z<L (11.26)
dz dz
where
P
k 2 = ------ (11.27)
EI
The general solution of eq. (11.26) for k 2 > 0 is

v 1 ( z ) = A 1 sin ( kz ) + A 2 cos ( kz ) + A 3 z + A 4 (11.28)

where A1, A2, A3, and A4 are arbitrary constants to be determined by boundary conditions.

EXAMPLE 11.1 Critical load for clamped-free boundary conditions (B)

Consider the clamped-free boundary conditions; i.e. b.c.’s (B) above. Determine the critical load P cr for
which the perfect column ceases to be stable.

Solution The boundary conditions in this case become

v1 ( 0 ) = 0 v1 ′ ( 0 ) = 0 EIv 1 ″ ( L ) = 0 [ v 1 ′′′ + k 2 v 1 ′ ] z = L = 0
where the primes denote derivatives with respect to z. Taking derivatives of eq. (11.28) we have

354 Aerospace Structures


Perfect Columns

v 1 = A 1 sin ( kz ) + A 2 cos ( kz ) + A 3 z + A 4
v 1 ′ = A 1 k cos ( kz ) – A 2 k sin ( kz ) + A 3
v 1 ″ = – A 1 k 2 sin ( kz ) – A 2 k 2 cos ( kz )
v 1 ′′′ = – A 1 k 3 cos ( kz ) + A 2 k 3 sin ( kz )

Substitute these solutions into the four boundary conditions to get

0 1 0 1 A1
k 0 1 0 A2
= 0 (11.29)
– k 2 sin ( kL ) – k 2 cos ( kL ) 0 0 A3
0 0 k2 0 A4
A non-trivial solution for A1 to A4 requires the determinate of coefficients to vanish

0 1 0 1
det k 0 1 0 = 0 (11.30)
– k 2 sin ( kL ) – k 2 cos ( kL ) 0 0
0 0 k2 0
After expanding this determinate we get

– k 5 cos ( kL ) = 0 (11.31)

which gives “n-values”, n = 1, 2, 3,..., of kL; or

( 2n – 1 ) π P
k n = -------------------- --- = -----n- n = 1, 2, 3, … (11.32)
L 2 EI

For k n L = ( 2n – 1 ) ( π ⁄ 2 ) , the fourth row of matrix eq. (11.29) gives A 3 = 0 ; using this result in the sec-
ond row of matrix eq. (11.29) gives A 1 = 0 ; the first row of matrix eq. (11.29) gives A 2 = – A 4 . Note that the
third row of matrix eq. (11.29) is identically satisfied. So eq. (11.32) implies

π 2 EI
P n = ( 2n – 1 ) --- -----2- n = 1, 2, 3, … , (11.33)
2 L
where Pn are the buckling loads. Equation (11.31) is called the characteristic equation, and the roots of this
equation determine the buckling loads. For each value of n we have an associated buckling mode (A1 = A3 = 0,
A2 = – A4)

v 1n ( z ) = A 4 [ 1 – cos ( k n z ) ] (11.34)

where coefficient A4 is arbitrary. The first three buckling modes are shown in Fig. 11.9. Note that the amplitude
of the buckling mode is not known. However, we can plot its shape. The critical load, denoted by P cr , is the low-
est buckling load. That is

π 2 EI
P cr = P 1 = ----- -----2-
4L

Aerospace Structures 355


Article 11.2

Remember that in design we use the minimum EI for the cross section.

π 2 EI
1 P 1 = ----- -----2-
0.8 4L
v1 ⁄ A4 0.6 n=1
0.4
0.2
0.2 0.4 0.6 0.8 1
2
1.5
v1 ⁄ A4 1
n=2 9π 2 EI
P 2 = --------- -----2-
0.5 4 L
0.2 0.4 0.6 0.8 1
2
1.5
v1 ⁄ A4 n=3 25π 2 EI
1 P 3 = ------------ -----2-
0.5 4 L

0.2 0.4 0.6 0.8 1

Fig. 11.9 First three buckling modes for the clamped-free column.

356 Aerospace Structures


Imperfect columns

The critical loads for boundary conditions A through D and for EI = constant are given in Fig. 11.10.

EI
A. Pinned-pinned P cr = π 2 -----2-
L
L

π 2 EI
P cr = ----- -----2-
4L
B. Clamped-free

EI
C. Clamped-clamped P cr = 4π 2 -----2-
L
L

EI EI
D. Clamped-pinned P cr = 2.046π 2 -----2- = ( 4.49 ) 2 -----2-
L L
L

Fig. 11.10 Buckling loads for the standard boundary conditions A to D.

11.3 Imperfect columns

11.3.1 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. 11.11.
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 influence of axial compression on bending, is determined
by the same analysis that led to eqs. (11.26) to (11.28), except that we drop the subscript “1” on the lateral dis-
placement, 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

Aerospace Structures 357


Article 11.3

4 2
dv dv
+ k2 2 = 0 0<z<L (11.35)
dz4 dz
2 2
dv dv
v(0) = 0 – EI ( 0 ) = eP v(L) = 0 – EI ( L ) = eP (11.36)
dz2 dz2
where k2 is as given in (11.27). Note that the boundary conditions, eqs. (11.36), are inhomogeneous. Thus, eqs.
(11.35) and (11.36) do not have the trivial solution v(z) = 0 for all values of the load P. Using the general solution

y,v

EI P P
z,w

e eP eP
e

P L P

Fig. 11.11 Column subjected to an eccentric axial load

form given by (11.28) for the solution of differential equation (11.35), subject to the boundary conditions
(11.36), we find

kL
v ( z ) = e – 1 + cos ( kz ) + tan  ------ sin kz (11.37)
 2

kL 1 – cos ( kL )
By a trigonometric identity, tan  ------ = ----------------------------- . Let δ denote the midspan displacement; i.e., δ = v(L/2).
 2 sin ( kL )
From eq. (11.37) we get

kL kL kL 1
δ = e – 1 + cos  ------ + tan  ------ sin  ------ = e --------------------- – 1
 2  2  2 kL
cos  ------
 2

The factor kL/2 can be written in terms of the eccentric axial load P and the critical load Pcr for the pinned-
pinned uniform column subjected to centric load as

kL 1 P  π 2 EI  π P
------ =  --- ------ L  -------------  = --- ------- .
2  2 EI  L 2 P cr  2 P cr

Thus, the center deflection of the eccentrically loaded column becomes

358 Aerospace Structures


Imperfect columns

1
δ = e ------------------------------ – 1 (11.38)
π P
cos  --- -------
 2 P cr

The load-displacement response is shown in Fig. 11.12. Note that δ → ∞ as P → P cr for e ≠ 0 . That is,

P/Pcr
e=0
1

EI
increasing e P cr = π 2 -----2-
L

0 δ

Fig. 11.12 Load-deflection curves for an eccentrically loaded column

no matter the magnitude of the eccentricity as P → P cr the value of the center deflection δ gets very large.

11.3.2 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 v 0 ( z ) . The column is subjected to a centric, axial compressive load P. The lateral
displacement of the column is denoted by v ( z ) , so that v ( z ) = v 0 ( z ) when P = 0. Also, the bending moment
in the column is zero under no load. Thus, we write the material law for bending as

2 2
d v d v 
M x = – EI  2 – 20 (11.39)
dz dz 
Vertical force equilibrium of the deflected column leads to a differential equation similar to eq. (11.21) except
that we drop the “1” subscript since it is the configuration 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.
(11.22) with the subscript “1” dropped. Combining these differential equations of vertical force equilibrium and
moment equilibrium via elimination of the vertical shear force gives
2 2
d Mx  d v
+  – 2 P = 0 (11.40)
dz2  dz 
The pinned-pinned boundary conditions are
v(0) = 0 Mx ( 0 ) = 0 v(L) = 0 Mx ( L ) = 0 (11.41)

Aerospace Structures 359


Article 11.3

πz
Consider the imperfection shape v 0 ( z ) = a 1 sin  ----- , where a 1 denotes the amplitude at midspan of the
 L
slightly crooked column. Substitute eq. (11.39) into eq. (11.40) to eliminate the moment to get
4 2 2
dv dv d v0 π 2 πz
+ k2 2 = = a 1  --- sin  ----- (11.42)
dz 4 dz dz 2  L  L

where k 2 is given by eq. (11.27). The boundary conditions, eqs. (11.41), for this imperfection shape lead to
2
dv
v = = 0 at z = 0 and z = L (11.43)
dz2
The solution of the differential equation (11.42) subject to boundary conditions (11.43) is
a1 πz
- sin  -----
v ( z ) = ---------------------- 0≤z≤L (11.44)
kL 2  L
1 –  ------
 π

It is convenient to measure the deflection of the imperfect column under load with respect to its original unloaded
state. That is, let δ define the additional displacement at midspan by δ = v ( L ⁄ 2 ) – v 0 ( L ⁄ 2 ) . Hence,

P
 -------
 P cr
δ = a 1 ---------------------- (11.45)
P
1 –  -------
 P cr

The load-displacement response is sketched in Fig. 11.13.


Note that δ → ∞ as P → P cr for a 1 ≠ 0 . That is, for a non- P ⁄ P cr EI
P cr = π 2 -----2-
zero value of the imperfection amplitude, the displacement L
gets very large as the axial force approaches the buckling 1.0
load of the perfect column. Also, the imperfect column
deflects in the direction of imperfection; e.g., if a 1 > 0 , then
δ>0. increasing a 1

Collectively the eccentric load and the geometric shape


imperfection are called imperfections. All real columns are 0 δ
imperfect. Even for a well manufactured column whose geo-
Fig. 11.13 Load-deflection response plots
metric imperfections are small and with the load eccentricity for geometrically imperfect columns
small, the displacements become excessive as the axial com-
pressive force P approaches the critical load P cr of the per-
fect column. Hence, the critical load determined from the
analysis of the perfect column is meaningful in practice.

360 Aerospace Structures


Column Design Curve

11.4 Column Design Curve


2
Consider the pinned-pinned uniform column whose critical load is given by P cr = π 2 ( EI ⁄ L ) . Let A denote the
cross-sectional area of the column. At the onset of buckling the critical stress is defined as

σ cr = P cr ⁄ A = ( π 2 EI ) ⁄ ( AL 2 ) (11.46)

We write the second area moment as I = r 2 A , where r denotes the minimum radius of
h
gyration of the cross section. For the rectangular section shown in the adjacent sketch,
I min = ( bh 3 ) ⁄ 12 and A = bh , so that r = h ⁄ 12 , where 0 < h < b . Thus, the critical b
stress becomes
E
σ cr = π 2 ----------------2 (11.47)
(L ⁄ r)
and L ⁄ r is called the slenderness ratio. The slenderness ratio is the column length divided by a cross-sectional
dimension significant to bending.

For any set of boundary conditions we define the effective length KL by the formula

EI
P cr = π 2 --------------2 (11.48)
( KL )
The effective lengths for the four standard boundary conditions are as follows:

A pinned-pinned EI EI KL = L K = 1
P cr = π 2 -----2- = π 2 --------------2
L ( KL )
B clamped-free π 2 EI EI KL = 2L K = 2
P cr = ----- -----2- = π 2 --------------2
4L ( KL )
C clamped-clamped EI EI KL = L ⁄ 2 K = 1⁄2
P cr = 4π 2 -----2- = π 2 --------------2
L ( KL )
D clamped-pinned EI EI KL = 0.699L K = 0.7
P cr = 20.2 -----2- = π 2 --------------2
L ( KL )
The definition 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., σ cr versus KL ⁄ r .
For elastic column buckling under all boundary conditions

π2E
σ cr = ---------------2 (11.49)
 KL
-------
 r 

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. 11.14

Aerospace Structures 361


Article 11.4

σ cr

σp
Euler curve, depends only on E

Fig. 11.14 Column curve for elastic


buckling

long columns
KL
0 -------
r
KL E
------- = π -----
r σp

11.4.1 Inelastic buckling


The column curve equation, eq. (11.49), is valid up to the proportional limit of the material, denoted by σ p . The
proportional limit is defined 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. (11.49), cannot be used. The proportional limit is difficult
to measure from test data because its definition 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 fit 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 σ 0.85 where the
secant line drawn from the origin with slope 0.85 E intersects the stress-strain curve, and the stress σ 0.7 where a
second secant line drawn from the origin with slope 0.7E intersects the stress-strain curve. These data are
depicted in Fig. 11.15. Note that the compressive normal strain corresponding to the stress σ 0.7 is usually about
σ

σ 0.7 experimental compressive


stress-strain curve
σ 0.85
Fig. 11.15 Data used to fit the compression 0.7E
stress-strain curve of aluminum alloys. 1
0.85E
1
E
1 ε
0 ≈ 0.002
the 0.2% offset yield strain for the material. Hence, stress σ 0, 7 is close to the 0.2% offset yield stress of the alu-
minum alloy. The Ramberg-Osgood equation is

σ 3 σ n–1
ε = --- 1 + ---  --------- (11.50)
E 7  σ 0.7

where the shape parameter n is given by


362 Aerospace Structures
Column Design Curve

17
ln  ------
 7
n = 1 + ---------------------- (11.51)
σ 0.7 
ln  ---------- -
 σ 0.85

We can re-write eq. (11.50) as

Eε σ 3 σ n
--------- = --------- + ---  --------- (11.52)
σ 0.7 σ 0.7 7  σ 0.7

and plot σ ⁄ σ 0.7 versus ( Eε ) ⁄ σ 0.7 for various values of the shape parameter n., and this plot is shown in Fig.

n = 2
1.2 n = 5
n = 10
1 50 n = 20
0.8 5 n = 50
σ
---------
σ 0.7 0.6 n = 2

0.4

0.2

0.5 1 1.5 2

---------
σ 0.7
Fig. 11.16 A normalized plot of the Ramberg-Osgood material law for various values of
the shape parameter n.

11.16. Some approximate values for common aluminum alloys are given in the table below.

AL E in106psi σ 0.7 in 103psi 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

From the Ramberg-Osgood equation, eq. (11.50), 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 tan-

gent modulus; i.e., ------ = E t where Et is the tangent modulus. Differentiating eq. (11.50) we get

Aerospace Structures 363


Article 11.4

dσ 3 nσ n – 1- dε 1 1 3n σ n–1
dε = ------ + --- --------------- dσ ------ = ----- = --- + --- ---  --------- (11.53)
E 7 Eσ 0.7 n–1 dσ Et E 7 E  σ 0.7

or
E
E t = --------------------------------------- (11.54)
3 σ n–1
1 + --- n  ---------
7 σ 0.7

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. (11.49), with the modulus of elasticity replaced by the tangent
modulus. This inelastic buckling analysis is called the tangent modulus theory. That is,
Et
σ cr = π 2 ---------------
2
(11.55)
 KL
-------
 r 

Now substitute eq. (11.54) for the tangent modulus in this equation, noting that σ = σ cr , to get

π2 E
σ cr = ---------------2 ---------------------------------------
KL 3
 ------- 1 + --- n  -------- σ cr  n – 1
-
 r  7  σ 0.7

After division by σ 0.7 , this equation can be written as

σ cr 3  σ cr  n 1
- + --- n --------- = -------------------------------2-
-------- (11.56)
σ 0.7 7  σ 0.7 ( KL ⁄ r )
------------------------
π E ⁄ σ 0.7
A plot of the column curve given by eq. (11.56) is shown in Fig. 11.17.

364 Aerospace Structures


Bending of thin plates

1.2

n = ∞
1 n = 50
n = 20
0.8
σ cr
--------
-
σ 0.7 0.6
n = 10
0.4 n = 5

0.2 n = 2

0.5 1 1.5 2 2.5 3 3.5


( KL ) ⁄ r
------------------------
π E ⁄ σ 0.7

Fig. 11.17 Column curves for a Ramberg-Osgood material law with different shape factors.

11.5 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 flat
and curved, are common structural elements in flight 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 thin, rectangular, flat plate is shown in Fig. 11.18 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, and 0 ≤ x ≤ a ,
0 ≤ y ≤ b and – t ⁄ 2 ≤ z ≤ t ⁄ 2 . The plane with z = 0 is the midsurface, or reference surface, of the plate.

A beam resists the transverse loads, or lateral loads, primarily by the longitudinal normal stress σx, and the
so-called lateral stresses σy, σz, τyx, and τyz are assumed to be negligible. Transverse loads, acting in the z-direc-
tion applied to the plate are primarily resisted by the in-plane stress components σx, σy, and τxy. Transverse shear
stresses τxz and τyz 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 theory, the transverse normal stress σz is very small
with respect to the in-plane normal stresses and, hence, is neglected. Bending of thin plates is discussed in many
texts on plate theory; For example, see Ugural and Fenster (2003). Only some elements of the plate bending the-
ory are discussed here. The assumptions of the linear theory for thin plates are as follows:
1. The deflection of the midsurface is small with respect to the thickness of the plate, and the slope of the
deflected midsurface is much less than unity.

Aerospace Structures 365


Article 11.5

σz z

τ zy
τ zx
τ xz τ yz
x y
t
σx τ xy τ σy
yx

3-D stress state


τ xy τ a
σx yx σy b a»t>0
b»t>0
primary stresses in plate theory

Fig. 11.18 Illustration of the nomenclature and primary stresses for a flat, rectangular plate

2. Straight lines normal to the midsurface in the undeformed plate remain straight and normal to the midsurface
in the deformed plate, and do not change length.
3. The normal stress component σz is negligible with respect to the in-plane normal stresses and is neglected in
Hooke’s law.

Now consider the deformation, or strains, caused by the normal stresses. Hooke’s law for the normal stresses
and strains in a three-dimensional state of stress is
1
ε x = --- ( σ x – νσ y – νσ z )
E
1
ε y = --- ( – νσ x + σ y – νσ z ) (11.57)
E
1
ε z = --- ( – νσ x – νσ y + σ z )
E
where E is the modulus of elasticity and ν is Poisson’s ratio. From assumption 3 the thickness normal stress σz
is assumed negligible and is set to zero in Hooke’s law. From assumption 2 the thickness normal strain ε z = 0 ,
because the line element normal to the midsurface does not change length. Since the normal stress σ z is also
assumed to vanish, the third of Eqs. (11.57) leads to a contradiction. Hence, the third equation of Hooke’s law is
neglected. The material law for the in-plane normal strains and stresses for thin plates is
1
ε x = --- ( σ x – νσ y )
E
(11.58)
1
ε y = --- ( – νσ x + σ y )
E

Consider pure bending of the plate or beam subjected to moment M. We consider two cases. In the first 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 first case the structure is a beam, and in 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 ρ , and the normal strain in
the x-direction is ε x = z ⁄ ρ . Note that we assumed that the x-axis coincided with the neutral axis in the unde-
formed beam. Hence, longitudinal line elements above the neutral axis, z > 0, are stretched, and line elements

366 Aerospace Structures


Bending of thin plates

z ρ
A ---
ν

A
x ρ
M M Section A-A
z

Fig. 11.19 Pure bending of a beam in the x-z plane and the associated anticlastic curvature of its
cross section.

below the neutral axis, z < 0, are compressed. In the case of a beam, the normal stress in the y-direction, σ y , is
also very small and is neglected with respect to the longitudinal normal stress σ x . That is, the beam resists the
applied bending moment by the longitudinal normal stress σ x . Since σ y = 0 , we get from Hooke’s law, eq.
(11.58), that
σ x = Eε x
ν z
ε y = – νε x = – --- z = – --------- (11.59)
ρ ρ
 ---
 ν

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 Poisson’s ratio times the longitudinal normal strain. The form of the last
expression for εy in eq. (11.59) shows that the line elements in the cross section parallel to the y-axis before
deformation also bend into circular arcs. The line element parallel to the y-direction at z = 0 in the undeformed
ρ
beam has a radius of curvature of --- in the deformed beam. This transverse curvature is called anticlastic curva-
ν
ture, and is illustrated in Fig. 11.19.

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 σy 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 σ y at the free edges. In the central portion of the plate, the associated normal strain is zero. The
suppression of anticlastic curvature is characterized by the vanishing of the normal strain εy. Hence from Hooke’s
law, eq. (11.58), for εy = 0 we get

E
σ x = -------------2-ε x σ y = νσ x (11.60)
1–ν
Since denominator in the expression for σx is positive but less than unity, the plate is stiffer than the beam owing

Aerospace Structures 367


Article 11.6

to the presence of the non-zero transverse normal stress σ y to help in resisting the applied moment. Compare
eqs. (11.59) and (11.60) for the normal stress σx. The quantity E ⁄ ( 1 – ν 2 ) is an effective modulus of the plate.

11.6 Compression buckling of thin rectangular plates


Consider the perfectly flat plate subjected to a longitudinal compressive force of magnitude P x applied in a spa-
tially uniform manner along edges x = 0 and x = a, as shown in Fig. 11.20. The intensity of the distribution load-
y

Px Px b

z t
x
Px Px

Fig. 11.20 Uniformly applied compressive forces applied to opposite longitudinal edges of a
rectangular plate.

ing along the edges is denoted as N x , where N x = P x ⁄ b . The dimensional units of N x are F/L. The equilibrium
response of the plate in linear theory is that of pure compression in the x-y plane with no out-of-plane deflection
of the midsurface. That is, in the pre-buckling equilibrium state the plate remains flat.

At a critical value of the compressive force P xcr the plate will buckle, or deflect out of the flat pre-buckling
equilibrium state. To determine this critical force we have to consider a slightly deflected equilibrium configura-
tion of the plate, similar to the analysis of the perfect column presented in Section 11.2. Refer to Brush and Alm-
roth (1975) for the details of this adjacent equilibrium analysis for the critical force.

Instead of a detailed adjacent equilibrium analysis of the plate, we can make a comparison to the critical
force determined for the pinned-pinned column in Fig. 11.10. The configuration of the plate comparable to the
pinned-pinned column has simply supported, or hinged, edges at x = 0 and x = a , and has free edges at
y = 0 and y = b . The compressively loaded plate for these boundary conditions is called a wide column. The
critical force for the pinned-pinned column is
EI
P cr = π 2 -----2- (11.61)
L
For the plate, replace the modulus of elasticity E in the column formula by E ⁄ ( 1 – ν 2 ) , since the plate is stiffer

368 Aerospace Structures


Compression buckling of thin rectangular plates

than the column. Also set L = a for the plate. The formula for the second area moment of a rectangular cross
section is I = ( bt 3 ) ⁄ 12 . Hence, eq. (11.61) transforms to

E 1 bt 3 Et 3  b
P xcr = π 2  -------------2-  ----2-  ------- = π 2  ------------------------- ----- (11.62)
 1 – ν   a   12   12 ( 1 – ν 2 ) a 2

For the wide column configuration of the plate, the critical load is written in the form

π 2 Db
P xcr = ------------
- (11.63)
a2
where the bending stiffness, or flexural rigidity, of the plate is defined as

Et 3
D = ------------------------- . (11.64)
12 ( 1 – ν 2 )

For other boundary conditions of the plate, the critical compressive load is written in the form

π2D
P xcr = k c ---------- (11.65)
b
where k c is a nondimensional buckling coefficient for compressive loading. This compressive buckling coeffi-
cient is a function of the plate aspect ratio a ⁄ b and the support conditions on the edges x = 0 , x = a , y = 0 ,
and y = b . The transition from column to plate as supports are added along the unloaded edges ( y = 0 and
y = b ) are depicted in Figure 1 of NACA TN 3781 (Gerard and Becker, 1957). A scanned image of this figure is
on page 370.

The compressive buckling coefficient is plotted for various support conditions in Figure 14 of NACA TN
3781, and a scanned image of this figure is on page 371. Note that the some of the curves for the buckling coeffi-
cient exhibit cusps, or discontinuous slopes, at selected values of the aspect ratio. The cusps correspond to
changes in the half wave length of the buckle pattern along the x-direction. In particular, for the plate with simple
support on all four edges, case C in Figure 14, note that k c = 4 for integer aspect ratios.

Aerospace Structures 369


Article 11.6

Fig. 1 from NACA TN 3781.

370 Aerospace Structures


Compression buckling of thin rectangular plates

Fig. 14 from NACA TN 3781

Aerospace Structures 371


Article 11.7

11.7 Buckling of flat rectangular plates under shear loads


A plate subjected to uniformly distributed shear loading is illustrated in Fig. 11.21. The shear force per unit

y
N xy

N xy b

Fig. 11.21 Plate subjected to in-plane shear loading

length is denoted by N xy , and it is the integral of the in-plane shear stress through the thickness of the plate; i.e.,

t⁄2

N xy = τ xy dz
∫ (11.66)
–t ⁄ 2

Note that eq. (11.66) is also the definition of the shear flow in thin-walled bar theory defined by eq. (6.9) on page
119. From Mohr’s circle for plane stress, the state of pure shear is equivalent to tensile and compressive normal
stresses at forty-five degrees to the direction of pure shear. It is this compressive normal stress that leads to buck-
ling of the thin plate subjected to shear.

The critical value of the shear force per unit length, N xycr , is given by the formula

π2D
N xycr = k s ---------
- (11.67)
b2
where k s is a nondimensional buckling coefficient for shear loading. This buckling coefficient is a function of the
plate aspect ratio a ⁄ b and the boundary conditions applied to the plate. Values of the shear buckling coefficient
are given in Figure 22 of NACA TN 3781. A scanned image of this figure is on page 373. The buckling mode
termed the symmetric mode in the figure pertains to a buckled form that is symmetric with respect to a diagonal
across the plate at the node line slope. For a narrow range of aspect ratios the plate buckles in an antisymmetric
mode. For infinitely long strip, or a ⁄ b → ∞ , k s = 5.35 for simply supported, or hinged, edges at y = 0, b , and
k s = 8.98 for clamped edges.

372 Aerospace Structures


Buckling of flat rectangular plates under shear loads

Fig. 22 from NACA TN 3781

Aerospace Structures 373


Article 11.8

11.8 Buckling of flat rectangular plates under combined compression


and shear
A plate subjected to uniformly applied longitudinal compression and shear is shown in Fig. 11.22. The critical
y
N xy

N xy Px ⁄ b b

Fig. 11.22 Plate subjected to longitudinal compression and in-plane shear loading

combination of shear and compression forces under different boundary conditions and different aspect ratios of
the plate can be approximated to a sufficient accuracy by
N xy  2  P x 
 -----------
- + --------- = 1 (11.68)
 N xycr  P xcr

where N xycr and P xcr are the critical values of the separately acting shear force per unit length and the compres-
sion force, respectively. Equation (11.68) is plotted in Fig. 11.23

Px
---------
P xcr
1

0.8

0.6

0.4

0.2
N xy
-----------
-
- 1 -0.5 0 0.5 1 N xycr

Fig. 11.23 Buckling interaction relationship for critical combinations of shear and compression.

374 Aerospace Structures


References

11.9 References
Brush, D. O., and Almroth, B. O., 1975, Buckling of Bars, Plates, and Shells, McGraw-Hill, Inc., New York,
pp. 75-105.

Gerard, G., and Becker, H., 1957, Handbook of Structural Stability, Part I, Buckling of Flat Plates, NACA
TN 3781,

Ugural A. C., and Fenster, S. K., 2003, Advanced Strength and Applied Elasticity, 4th Edition, by Pearson
Education, Inc., Publishing as Prentice Hall Professional Technical Reference, Upper Saddle River, New Jer-
sey, 07458, pp. 472-490.

Aerospace Structures 375


Article 11.9

376 Aerospace Structures

Das könnte Ihnen auch gefallen