Sie sind auf Seite 1von 89

Department of

Mechanical and Industrial Engineering

STRENGTH OF MATERIALS IV
Study guide 1 for SOM401M

HM Ngwangwa
RM Nkgoeng
University of South Africa
Florida

2011 University of South Africa


All rights reserved
Printed and published by the
University of South Africa
Muckleneuk, Pretoria
SOM401M/1/20122014
98789155
InDesign, Florida

CLO_Style

SOM401M/1

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV


CONTENTS
Theme

Page

Shear Forces and Bending Moments

1.1

Types of beams, loads and reactions

1.2

Types of loads

1.3

Reactions

1.4

Shear forces and bending moments

1.5

Sign conventions

1.6

Relationships between loads, shear forces and bending moments

1.7

Shear force and bending-moment diagrams

1.8

Concentrated load

1.9

Uniform load

1.10 Several concentrated loads

1.11

Exercise

Shear Stresses in Beams


(Basic Topics)

2.1

Introduction

2.2

Pure bending and non-uniform bending

2.3

Curvature of a beam

2.4

Longitudinal strains in beams

2.5

Normal stress in beams

2.6

Design of beams for bending stress

10

2.7

Shear stresses in beams of rectangular cross-section

10

2.8

Vertical and horizontal shear stresses

11

2.9

Derivation of shear formula

11

2.10 Calculation of the first moment Q

12

2.11 Distribution of shear stresses in a rectangular beam

12

2.12 Effects of shear strains

13

2.13 Shear stresses in the webs of beams with flanges

14

2.14 Maximum and minimum shear stresses

15

2.15 Built-up beams and shear flow

15

2.16 Shear flow

16

2.17 Exercise

17
iii

iv

Stresses in Beams Advanced Topics

19

3.1

Introduction

19

3.2

Composite beams

19

3.3

Strains and stresses

19

3.4

Neutral axis

20

3.5

Moment-curvature relationship

20

3.6

Normal stresses

21

3.7

Approximate theory for bending of sandwich beams

21

3.8

Transformed-section method

21

3.9

Doubly symmetric beams with inclined loads

21

3.10 Normal stresses (bending stresses)

22

3.11 Neutral axis

22

3.12 Relationship between N.A. and the inclination of the loads

22

3.13 Bending of asymmetric beams

23

3.14 Neutral axis

23

3.15 Procedure for analysing an asymmetric beam

24

3.16 Deflection due to asymmetrical bending

24

3.17 The shear-centre concept

25

3.18 Shear stresses in beams of thin-walled open cross-sections

25

3.19 Shear stresses in wide-flange beams

26

3.20 Shear stresses in the upper flange

26

3.21 Shear stress in the web

27

3.22 Shear centres of thin-walled open sections

27

3.23 Channel section

28

3.24 Angle section

29

3.25 Z-section

29

3.26 Elasto-plastic bending

29

3.27 Yield moment

29

3.28 Plastic moment and neutral axis

30

3.29 Plastic modulus and shape factor

30

3.30 Beams of rectangular cross-section

31

3.31 Beams with wide-flange shape

32

3.32 Exercise

32

Deflections of Beams

33

4.1

Introduction

33

4.2

Differential equation of the deflection curve

33

4.3

Beams with small angles of rotation

33

4.4

Non-prismatic beams

34

4.5

Prismatic beams

34

4.6

Deflection by integration of the bending-moment equation

34

4.7

Deflection by integration of the shear-force and load equations

35

4.8

Method of superposition

35

4.9

Moment-area method

36

4.10 Non-prismatic beams

36

4.11 Strain energy

36

4.12 Deflections caused by a single load

37

4.13 Castiglianos theorem

37

4.14 Derivation of Castiglianos theorem

38

4.15 Application of Castiglianos theorem

38

4.16 Use of a fictitious load

39

4.17 Differentiation under the integral sign

40

4.18 Deflections produced by impact

41

4.19 Temperature effects

41

Statically Indeterminate Beams

5.1

Introduction

43

5.2

Types of statically indeterminate beams

43

5.3

Analysis by the differential equations of the deflection curve

45

5.4

Continuous beam on multiple supports

45

5.5

Reactions at supports

46

5.6

Method of superposition

46

5.7

Analysis with RB as redundant

46

5.8

Temperature effects

47

5.9

Longitudinal displacements at the ends of a beam

47
47

SOM401M/1

5.10 Exercise

43

vi

Columns

49

6.1

Introduction

49

6.2

Buckling and stability

50

6.3

Critical load

50

6.4

Columns with pinned ends

51

6.5

Differential equation for column buckling

52

6.6

Critical loads

53

6.7

Critical stress

54

6.8

Columns with other support conditions

54

6.9

Column fixed at the base and free at the top

55

6.10 Effective lengths of columns

56

6.11 Fixed-free column

56

6.12 Column with both ends fixed against rotation

56

6.13 Column fixed at the base and pinned at the top

57

6.14 Critical loads and effective lengths

57

6.15 Columns with eccentric axial loads

57

6.16 Maximum deflection

59

6.17 Maximum bending moment

59

6.18 The secant formula for columns

59

6.19 Elastic and inelastic column behaviour

59

6.20 Inelastic buckling

60

6.21 Exercise

60

Thick and Compound Cylinders

61

7.1

Introduction

61

7.2

Boundary conditions

61

7.3

Shear stress in a beam

62

7.4

Transverse normal stress in a beam

63

7.5

Stress distribution in a pressurised thick-walled cylinder

64

7.6

Cylinder with end caps

65

7.7

Cylinder with pistons

66

7.8

Internal and external pressure

67

7.9

Internal pressure only

68

7.10 Stress distribution for and r

68

7.11

69

Axial stress and strain

7.12

Cylinder with end caps but free to change in length

69

7.13 Pressure retained by piston in each end of cylinder

69

7.14

Methods of containing high pressure

69

7.15 Stress set up by a shrink-fit assembly

70

7.16

Stress distribution in a pressurised compound cylinder

71

7.17

Using a spreadsheet to determine the stresses in a compound cylinder

71

Theories of Elastic Failure

8.1

Introduction

73

8.2

Yield criteria and fracture criteria

73

8.3

Maximum principal stress theory

74

8.4

Maximum shear stress theory

74

8.5

Maximum principal strain theory

75

8.6

Maximum total strain energy per unit volume theory

76

8.7

Maximum shear strain energy per unit volume theory-distortion energy

76

8.8

Mohr fracture criterion or Mohrs modified theory

76

8.9

Mohr fracture criterion or Mohrs modified theory

77
77

SOM401M/1

8.10 Exercise

73

vii

SOM401M/001



1 L M
,
,
, T
dx,

JG


1 L M
M
dx,
EI

3WL
,
5Gbd

3WL
3WL
,

5Gbd
20Gbd

6WL

5Gbd

2


2
4


,
n
2

2


2
4

A x A x

MA L 2MB L L MC L 6

L
L


Land

2,

anad

2 12 2,

an

, 12 12 4 ,

, an an




an an ,
,
,

, , and



,an

0,

and
8

viii

SOM401M/001

3

4
12

6
4

2
3

2

3

4 3
6

SOM401M/1

ix

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 1
Shear Forces and Bending
Moments

SOM401M/001

Chapter 1 Shear Forces and Bending Moments


The important points to be picked up from this chapter are the following:

The plane of bending

Theimportantpointstobepickedupfromthischapterarethefollowing:
How to work with axially loaded beams
Theplaneofbending
This chapter
is a revision of a chapter in SOM381B and has been included to refresh your
Howtoworkwithaxiallyloadedbeams
memory.
This chapter is a revision in SOM401M. I choose to believe that you have already dealt with this
chapterinSOM381B.Itwasincludedsothatyoucanrefreshyourmind.

1.1 Types of beams, loads and reactions

Types of beams, loads and reactions


Summary
Summary

We
the textbook
onsupported
page 306, beam,
fig. 4.2 Fig1.1(b)
(G&G). The
first figurebeam
shows and
a simply
supported
Figrefer
1.1(a)toshows
a simply
a cantilever
Fig1.1(c)
a beam
beam,thesecondfigureacantileverbeamandthethirdoneabeamwithanoverhang.Makesure
with an overhang. Make sure that you understand the difference in the types of beams
that you understand the difference in the types of beams and reactions. Majority of the study of
and reactions. The study of Strength of Materials and Design mainly involves working with
Strength of Materials and design involves working with beams of different sections and loaded in
beams of different sections which are loaded in various ways.
variousways.

P1 P2

HA A

RA

B
c

HA

MA
RB

P3
12
5

(a)

P4

RA

q2

q1

(b)

M1

a
RA

RB

L
(c)

Figure 1.1

SOM401M/1

Types of loads

Concentratedloadtheseinvolveforcesactingvertically,horizontallyorskewtothesection.
Distributedloadtheyaremeasuredbytheirintensity.Theycanvarylinearlyortheycanbe
constant.Theyareexpressedinunitsofforceperunitdistance(N/m)
Couplesandmomentstheyareexpressedinforceunitdistance(Nm)

Reactions
Findingthereactionsisusuallythefirststepintheanalysisofabeam.Oncethereactionsareknown, 1
the shear forces and bending moments can be found. If a beam is supported in a statically
10

1.2 Types of loads

Concentrated load involves forces acting vertically, horizontally or skew to the section.

Distributed load is measured by its intensity. The load can vary linearly or can be
constant, and is expressed in units of force per unit distance (N/m).

Couples and moments are expressed in force-unit distance (N-m).

1.3 Reactions
Finding the reactions is usually the first step in the analysis of a beam. Once the reactions
are known, the shear forces and bending moments can be found. If a beam is supported
in a statically determinate manner, all reactions can be found from the free-body diagram
and equations of equilibrium.

1.4 Shear forces and bending moments


Summary
When a beam is loaded by forces or couples, stresses and strains are created throughout
the interior of the beam. To determine these stresses and strains, we must first find the
internal forces and internal couples that act on cross-sections of the beam.

Read more on page 313 in the textbook.

1.5 Sign conventions


It is customary to assume that shear forces and bending moments are positive when
they act in directions as shown in fig. 1.2(a). The shear force tends to rotate the material
clockwise and the bending moment tends to compress the upper part of the beam and
elongate the lower part.
According to fig. 1.2(b), a positive shear force acts clockwise against the material and a
negative shear force acts anticlockwise, whereas a positive bending moment compresses
the upper part of the beam and a negative bending moment compresses the lower part.

Figure 1.2
When drawing free-body diagrams for beams, always apply positive shear forces and
bending moments as shown in the figures above.

1.6 Relationships between loads, shear forces and bending


moments
Summary
In this section our objective is to obtain important relationships between loads, shear forces
and bending moments in beams. These relationships are quite useful when investigating
the shear forces and bending moments throughout the entire length of a beam, and they
are helpful when constructing shear-force and bending-moment diagrams.

Read more on pages 320325.

SOM401M/1

Looking at fig. 1.3 and tackling only the case of distributed loading, let us consider an
element of a beam cut out between two cross-sections that are distance dx apart. In this
case, the sign convention for the distributed load is positive when it acts downward on the
beam and negative when it acts upward.


Lookingatfig.1.5(a)andtacklingonlythecaseofdistributedloading...Letusconsideranelementof
abeamcutoutbetweentwocrosssectionsthataredistanceapart.Now,thesignconventionfor
the distributed load is positive when it acts downward on the beam and negative when it acts
upward.

q(x)

q(x)

M+dM

dx

V+dV

/F

vert

= 0:

Figure 1.3

V - qdx - ^V + dV h = 0

dV =

q
dx

(1.1)

(1.1)

Integratingwithrespecttowegetthedistributedload.FromEqn.1.1wecanseethattherate
Integrating
with force
respect
x point
we get
equation
1.1 we can
see
of change
of theVshear
at to
any
onthe
thedistributed
axis of theload.
beamFrom
is equal
to the negative
of the
that the rate of change of the shear force at any point on the axis of the beam is equal to the
intensityofthedistributedloadatthatsamepoint.Ifthereisnodistributedloadonasegmentof

negative
of the
intensity
distributed
atisthat
samein
point.
If there
is no
distributed
the
and
the shearload
force
constant
that part
of the
beam.
If the
the beam,
i.e.,

then of
dv

load on a segment of
the beam, ie q = 0 then

dx
distributedloadisuniformalongpartofthebeam,the

uniform

and the shear force is constant in that

,isalsoconstantandtheshearforce
along part of the beam, the dv = 0 , is
dx

part of the beam. If the distributed load is


also constant and the shear force varies linearly in that part of the beam.

varieslinearlyinthatpartofthebeam.

dV = - # qdx
A

12

(1.2)

VA - VB = - # qdx
A

The change in shear force between two points along the axis of the beam is equal to the
negative of the total downward load between those points. The area of the loading diagram
may be positive (if q acts downward) or negative (if q acts upward).

1.7 Shear force and bending-moment diagrams


Summary
When designing a beam, we usually need to know how the shear forces and bending moments vary throughout the length of the beam. Of special importance are the maximum
and minimum values of these quantities. Information of this kind is usually provided by
graphs in which the shear force and bending moment are plotted as ordinates and the
distance x along the axis of the beam is plotted as the abscissa. Such graphs are called
shear-force and bending-moment diagrams. Read pages 325 to 337. This information is
very important and it is essential that practising engineers know this.

bendingmomentareplottedasordinatesandthedistancealongtheaxisofthebe
theabscissa.Suchgraphsarecalledshearforceandbendingmomentdiagrams.Rea
to 337. This information is very important and forms a strong foundation of y
practicingengineer.

Concentrated load
1.8 Concentrated
load

RA

M0

RB

M
RA

RA

Referringtofig.1.6(a),wearegoingtoconsidertheentirebeamasafreebodyanda
thereactionsofthebeamcanbereadilydeterminedfromequilibrium.
Figure 1.4
beam as a free body and as it will
Referring to fig. 1.4, we
are going to consider the entire

be seen, the reactions of the beam can be readily


determined
from equilibrium.

TheaboveisfoundbytakingthesumofmomentsaboutBandArespectively.

Pb

and

Pa

A =
B =
Thesecondstepwouldbetocutthroughthebeamatacrosssectiontotheleftoft
L
L
a distance x from the support at A. Fig. 1.6(b) shows the cut section with the
The above is found by taking the sum of moments about B and A respectively. Where b is
bendingmomentactingandareherebywrittenasfollows:

distance of P from B and a is distance of P from A.

The second step is to cut through the beam atacross-section


left

tothe

the load
Pand

of

at a distance x from the support at A. Figure 1.4 shows the cut section with the shear force
and bending moment acting and written as follows:

V = RA = Pb and M = R A x = Pbx
L

13

^0 1 x 1 a h

The above expressions of V and M are valid only for the part of the beam to the left of the
load P. You have to repeat the process until the whole beam is covered. Once you have
done this, you will find the maximum moment, which is

Mmax = Pab
L

1.3

SOM401M/1

and it occurs under the concentrated load.

Theaboveexpressionsofandarevalidonlyforthepartofthebeamtotheleftoftheload.
Youhavetorepeattheprocessuntilthewholebeamiscovered.Attheendofthedaywefindthe
maximummomentbeing

anditoccursundertheconcentratedload.

1.9 Uniform load

Uniform load
A

RA

q
B

FreeBodyDiagram A

x
L

RB

1.3

M
V

x
RA

Figure 1.5

Letusconsidernowthefigureaboveinwhichthebeamisloadedwithauniformlydistributedload.

Let us now consider


the figure above in which the beam is loaded with a uniformly distribThebeamanditsloadingaresymmetrichencewecancomfortablydeterminethereactionsas:
uted load. The beam and its loading are symmetric, hence we can comfortably determine

the reactions as:



2
Theshearforceandbendingmomentatdistancefromthelefthandendaregivenby:
qL

R A = RB =

left-hand



The shear force and bending moment
atdistance
x fromthe
2 end are given by:

Theaboveequationsarevalidthroughoutthelengthofthebeam.Themaximummomentisactually
qx 2
V = R A - qx and
M = RA x theareaundertheparabolaanditisfoundtobe

The above equations


are valid throughout the length of
the beam. The maximum moment
(1.4)


8
is actually the area under the parabola and is

Theaboveisfoundfromthefollowing:

M
max =

qL2

8

(1.4)

2
2
2
Now if we set the above to zero, we get2a value of and if we can substitute it in to the
dM = d c qLx - qx m = qL - qx = V
momentequationwewillbeabletoarriveat


The above is calculated from the following:

dx

dx

Now, if we set the above to zero, we get a valueof


x=
0.5 and if we can substitute it into
8
the moment equation we will arrive at
Several concentrated loads qL2


Mmax =
Dothissectiononyourown.

1.10 Several
concentrated loads

14

Do this section on your own.


Examples to be done:

All examples must be done

1.11 Exercise
Problems to be done:

4.314.35; 4.312; 4.51011; 4.514; 4.51617; 4.527 and 4.539.

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 2
Shear Stresses in Beams
(Basic Topics)

2.1 Introduction
This chapter is a revision of SOM381B and has been included to refresh your memory.
In this chapter we will be investigating the stresses and strains associated with those shear
forces and bending moments. Knowing the stresses and strains will allow us the opportunity
to analyse and design beams subjected to a variety of loading conditions. The loads acting on a beam cause the beam to bend (or flex), thereby deforming its axis into a curve.
As an example, we will consider a cantilever beam AB subjected to a load P at the free
end (see fig. 2.1).
The initial straight axis is bent into a curve (fig. 2.1(b)), called the deflection curve of the
beam.
The deflection of the beam at any point along its axis is the displacement of that point
from its original position, measured in the y direction. In this book the authors denote
deflection by v, but you are welcome to use y or anything that you are comfortable with,
just remember to be consistent with your notation.
P

(a)

(b)

Figure 2.1

2.2 Pure bending and non-uniform bending

SOM401M/1

When analysing beams, it is often necessary to distinguish between pure and non-uniform
bending.

Pure bending refers to flexure of a beam under a constant bending moment. It occurs
only in regions of a beam where the shear force is zero because V = dM . Non-uniform
dx
bending refers to flexure in the presence of shear forces, which means that the bending
moment changes as we move along the axis of the beam.

Read more about this on pages 353 and 354.

2.3 Curvature of a beam


When loads are applied to a beam, its longitudinal axis is deformed into a curve. The
resulting strains and stresses in the beam are directly related to the curvature of the deflection curve.

Figure 2.2
We will be using a cantilever to illustrate this concept (see fig. 2.2(a)). The deflection curve
of this beam is shown in fig. 2.2(b). If k is curvature and represents the radius of curvature,

k= 1
t

(2.1)

What is curvature? Curvature is a measure of how sharply a beam is bent. If the load on a
beam is small, the beam will be nearly straight, the radius of curvature will be very large,
and the curvature will be very small. If the load is increased, the amount of bending will
increase, the radius of curvature will become smaller and the curvature will become larger.

k = 1 = di 
t
ds

(2.2)

The above equation 2.2 can also be written as in equation 2.3 because the distance ds
along the curve may be set equal to its horizontal projection distance dx (see fig. 2.2(b)).

Read more on page 356.

k = 1 = di 
t
dx

(2.3)

2.4 Longitudinal strains in beams

Read about this on pages 356360 and do example 51.

2.5 Normal stress in beams


In this chapter, we will be using the stress-strain curve for the material to determine the
stresses from the strains. The stresses act over the entire cross-section of the beam and vary
in intensity depending upon the shape of the stress-strain diagram and the dimensions of
the cross-section. Since x direction is longitudinal as can be seen in fig. 2.3, we will use
to denote these stresses.
SOM401M/001

M0

M0

D
B
C

C m
M e

p D
f y
s

dx

y
M
x

C m
e

M
z

dx

p D
f
y

s
q

Afterdeformation

Beforedeformation

Thenormalstressconsistsoftwostressresultants:
Figure 2.3

normal stress consists of two stress resultants:


The
aforceactinginthedirection,and
Bendingcoupleactingaboutaxis.

A force acting in the x direction

However,
axial couple
force isacting
zero when
the
Bending
aboutaz beam
axis is in pure bending...therefore we can write the
followingtwoequationsofstatics:

SOM401M/1

However, the axial force is zero when a beam is in pure bending, and we can therefore
write the following two equations of statics:

Fx = 0 =

# v dA = # ydA
x

k= 1 - M
t
EI
1

The flexure formula can be derivedonce


the location
of the neutral axis has been deter


mined and moment-curvature relationship has been derived (see the equation below):

Theflexureformulacanbederivedoncethelocationoftheneutralaxishasbeendeterminedand
My
momentcurvaturerelationshiphasbeenderived...seetheequationbelow:

(2.4)

vx =-

Readmoreaboutthederivationfrompg.361367anddoexamples5.25.4.

(2.4)

Read more about the derivation on pages 361367 and do examples 5.25.4.

2.6 Design of beams for bending stress


The process of designing a beam requires that many factors be considered including the
type of structure, the material to be used, the loads to be supported, the environmental
conditions to be encountered and the costs involved. From the standpoint of strength,
the task is eventually reduced to selecting a shape and size of beam such that the actual
stresses in the beam do not exceed the allowable stresses for the material. In this section
we will be considering only the bending stresses as determined from the flexural formula.
Later on we will consider the effects of shear stresses.
When designing a beam to resist bending stresses, we usually begin by calculating the
required section modulus. If the beam has a doubly symmetric section and the allowable
stresses are the same for both tension and compression, we can calculate the required
modulus by dividing the maximum bending moment by the allowable stress for the material.

S = Mmax 
vallow

(2.5)

The allowable stress is based upon the properties of the material and the desired factor
of safety.

Read more on pages 374377 and do examples 5558 from G&G.


Do example 57

2.7 Shear stresses in beams of rectangular cross-section


Let us consider a beam of rectangular section such as the one shown in fig 2.4. We will
make the following assumptions:
1
2

The shear stresses acting on the cross-section are parallel to the shear force, ie parallel to the vertical sides of the cross-section.
The shear stresses are uniformly distributed across the width of the beam, although
they may vary over the height.

In order for us to start with the analysis, we will isolate a small element as shown in
fig 2.4. At any point in the beam the complementary shear stresses are equal in magnitude.
The equality of the horizontal and vertical shear stresses acting on an element leads to
an important conclusion regarding the shear stresses at the top and bottom of the beam.

10

Inorderforustostartwiththeanalysis,weisolateasmallelementasshowninFig526a.Atany
point in the beam, the complementary shear stresses are equal in magnitude. The equality of the
horizontal and vertical shear stresses acting on an element leads to an important conclusion
regardingtheshearstressesatthetopandbottomofthebeam.

(a)

(b)

Figure 2.4
ByreferringtoFig2.4,wecanbeabletoshowtheexistenceofhorizontalshearstressesinabeam.If
By referring to fig 2.4, we can show the existence of horizontal shear stresses in a beam.
thefrictionbetweenthebeamsisnegligible,eachbeamwillbendindependently.Eachbeamwillbe
If the friction between the beams is negligible, each beam will bend independently. Each
incompressionaboveitsownneutralaxisandintensionbelowitsneutralaxis.
beam will be in compression above its own neutral axis and in tension below its neutral axis.
Ifnowthetwobeamsbecomegluedtogetheralongthecontactsurfacesothattheybecomeasingle
If the two beams become glued together along the contact surface, they become a single
solidbeam.Whenthissolidbeam(glued)isloaded,horizontalshearstressesmustdevelopalongthe
solid beam. When this solid beam (glued) is loaded, horizontal shear stresses must develop
along the glued surface in order to prevent the sliding as shown in the bottom figure of
gluedsurfaceinordertopreventtheslidingasshowninthebottomfigureofFig.2.4b.
fig. 2.4b.

Vertical and Horizontal Shear Stresses


Readmoreonpg.387
2.8 Vertical

and horizontal shear stresses

Derivation of Shear Formula


Readmoreonpg.388391.

Read more on page 387.


18

2.9 Derivation of shear formula

Read more on pages 388391.


x=

VQ

Ib

(2.6)

SOM401M/1

The above equation is known as the shear formula and it can be used to determine the
shear stress at any point in the cross-section of a rectangular beam.

11


(2.6)

Theaboveequationisknownastheshearformulaanditcanbeusedtodeterminetheshearstress
atanypointinthecrosssectionofarectangularbeam.

m1

M V
dx

m1

M+dM

p1

y1

dx

V+dV
n1

n1

m1

p1
dx

Figure 2.5

M+dM

y1

2.10 Calculation
the first
Calculation
of the First of
Moment
moment Q

If we are to determine shear stress at any point on the cross-section of the beam, we need
Ifwearetodetermineshearstressatanypointonthecrosssectionofthebeam,weneedtoobtain
to obtain Q by calculating the first moment of the cross-sectional area above the level of
bycalculatingthefirstmomentofthecrosssectionalareaabovetheleveloftheshadedareain
the shaded area in fig. 2.5
Figure2.6

Q = Ay

,whereisthecrosssectionalareaoftheshadedelement;andisthedistancefromthecentroid
where A is the cross-sectional area of the shaded element; and y is the distance from the
ofthebeamtothecentroidoftheshadedelement.
centroid of the beam to the centroid of the shaded element.

Figure 2.6

Distribution of Shear Stresses in a Rectangular Beam


Aswehaveexplainedintheprevioussubsectionabouthowtodetermine,wearenowgoingto
2.11 Distribution of shear stresses in a rectangular beam
takeitfurther.Againwewillreferto2.6
As we have explained in the previous subsection about how to determine Q, we are now

fig.
2.6.

we will refer to
going to take it further. Again

2
h 2
2 y1
2 4
2
Q = b ` h - y1jc y1 + 2
m = b c h - y12 m
2
2
2 4
WhenwesubstitutetheexpressionforQintotheshearformula,weget
When we substitute the expression for Q into the shear formula, we get
19

12

2
x = V c h - y12 m 
2I 4

(2.7)

SOM401M/001
SOM401M

(2.7)

(2.7
2 4 2 4
ByobservingEq.2.7,wecanseethatshearstressvaryquadraticallywiththedistance
.Thisisalso
ByobservingEq.2.7,wecanseethatshearstressvaryquadraticallywiththedistance
By looking
at equation 2.7, we can see that the shear stress varies quadratically with
the .Thisis
shownbyFig.2.7
shownbyFig.2.7
distance
y1. This is also shown in fig. 2.7.

Figure 2.7

Effects of
ShearofStrains
Effects
Shear Strains

2.12 Effects of shear strains

Sincetheshearstressvariesparabolicallyovertheheightofarectangularbeam,itfollowsthatthe
Sincetheshearstressvariesparabolicallyovertheheightofarectangularbeam,itfollowstha
Since the shear
parabolically over the height of a rectangular beam, it follows
stress varies

shearstrain
forshearstrainalsovariesparabolically.
x
forshearstrainalsovariesparabolically.
shearstrain

for shear strain also varies parabolically.


that the shear strain c =

WenowlookatFig.2.8toseetheeffectsofshearstrains.Itcanbeseenthatthecrosssectionsof
Look at
fig. 2.8 to see the effects of shear strains. It can be seen that the cross-sections of
WenowlookatFig.2.8toseetheeffectsofshearstrains.Itcanbeseenthatthecrosssectio
the beam
that were originally plane surfaces become warped. Warping effects will not be
thebeamthatwereoriginallyplanesurfacesbecomewarped.Warpingeffectwillnotbedealtwith
thebeamthatwereoriginallyplanesurfacesbecomewarped.Warpingeffectwillnotbedealt
dealt
with
in either SOM381B or SOM401M.
ineitherSOM381BorSOM401M.
ineitherSOM381BorSOM401M.

m1
m
n
1

m1
p1m
n p
q n1
q1

p1P

q1

Figure 2.8

Doexample511
Do example
511.
Doexample511
Letusconsiderabeamwithacircularsection(solid)withradius.
Let us Letusconsiderabeamwithacircularsection(solid)withradius.
consider a beam with a circular section (solid) with radius r.

r 4, A = rr 2and 4r
I = r
y = 3r4
4
4


, ,

3
4
3
4 3
` Q = 4r

34

The maximum shear stress is determinedasfollows:


3
3
Themaximumshearstressisdeterminedasfollows:
VQ 4V
Themaximumshearstressisdeterminedasfollows:

xmax =

Ib

3A

4 4

3 3
Readonpg.397anddoexample513.
Readonpg.397anddoexample513.

SOM401M/1

Read page 397 and do example 513.

20

20

13

2.13 Shear stresses in the webs of beams with flanges


When a beam as shown in fig. 2.9(a) is subjected to shear forces as well as bending moments, both normal and shear stresses are developed on the cross-sections. By looking at
fig. 2.9 (b), we can see that the shear stresses in the web of a wide flange beam only act
in the vertical direction and are larger than the stresses in the flanges.

Figure 2.9
Shear stresses in the web
Now look at fig. 2.10. The following assumptions are made:
1
2

The shear stresses act parallel to the y axis.


The shear stresses are uniformly distributed across the thickness of the web.

As before, we need to calculate the first moment Q. The area of interest is the area between e f and the top edge of the cross-section of the shaded area as shown in fig. 2.10.

A1 = b ` h - h1 j
2
2

Figure 2.10
Here b is the width of the flange, h is the overall height of the beam, and h1 is the distance
between the insides of the flanges.

A2 = t ` h1 - y1j
2
where t is the thickness of the web and y1 is the distance from the neutral axis to the line ef.

h1 - y1
h - h1
h
1
2
2
e
o
e
o
Q = A1
+
+ A2 y1 + 2
2
2
2

14

Q = b ^h2 - h12h + t ^h12 - 4y12h 


8
8

(2.8)

The shear stress in the web of the beam at distance y1 from the neutral axis is found to be:

x = V 6 b^h2 - h12h + t^h12 - 4y12h@


8It

(2.9)

2.14 Maximum and minimum shear stresses


In order to determine these, you need to know where they occur. The maximum shear
stress occurs at the neutral axis, where y1 = 0, while the minimum shear stress occurs where
the web meets the flanges, ie y1 = ! h1 . We just need to substitute these values of y1 into
2
equation 2.9.

xmax = V ^bh2 - bh12 + th12h and xmin = V ^h2 - h12h 


8It
8It

(2.10)

Check the shear distribution in fig. 2.11b.

2.15 Built-up beams and shear flow


Built-up beams are beams that are fabricated from two or more pieces of material joined
together to form a single beam as shown in fig. 2.12.

SOM401M/1

Figure 2.12

The next example is a steel plate girder of the type commonly used in bridges and large
buildings. The design calculations in these types of beams involve two phases. In the first
phase, the beam is designed as though it was made of one piece, taking into account both
bending and shear stresses. In the second phase, the connections between the parts are
designed to ensure that the beam does indeed behave as a single entity.
15

SOM401M/001

Shear flow

Inorderforustoobtaintheformulaforthehorizontalshearforcesactingbetweenpartsofabeam

2.16
Shear flow
fromthehorizontalequilibriumofthesubelementinFig.542bwedeterminetheforce
actingon

itslowersurface:
In
order for us to obtain the formula for the horizontal shear forces acting between parts
of a beam, from the horizontal equilibrium of the sub-element in fig. 2.13 we determine

(2.11)
the force F3 acting on its lower surface:

Shearflowisthehorizontalshearforceperunitdistancealongthelongitudinalaxisofthebeamand
(2.11)

F3 = dM # ydA 
I
,actsalongthedistance.
Shear flow is the horizontal shear force per unit distance along the longitudinal axis of the
beam,
and F3 acts along the distance dx. 1



f = F3 = dM ` 1 j # ydA
dx
dx,weobtainthefollowingformulaforshearflow:
I
...butweknowthat and

dM
But
we know that dx = V and
shear flow:

m1

p1

obtain the following formula for


# ydA = Q, wethus

f=


M+dM
y1

n1

m1
p

p1

y1 2
x

dx

m
F1

(2.12)

(2.12)

dx
n

VQ

I

m1

p1

F2

F3

dx

y1

2
x

Figure 2.13

Areasused
Used
when
Calculating
the First
Areas
when
calculating
the first moment
Q Moment Q

To
illustrate how Q is calculated, look at fig. 2.14(a). The welds must transmit the horizontal
Toillustratehowiscalculated,letusrefertoFig.2.14(a).Theweldsmusttransmitthehorizonta
shear
between
thethe
flanges
andand
thethe
web.
At the
flange,
the horizontal
shearforces
forcesthat
thatact
act
between
flanges
web.
Atupper
the upper
flange,
the horizontal shea
shear
force
is
the
shear
flow
along
the
contact
surface
aa.
The
shear
flow
is
calculated
forceistheshearflowalongthecontactsurface.Theshearflowiscalculatedbytakingasthe
by
taking Q as the first moment of the cross-sectional area above the contact surface aa.
firstmomentofthecrosssectionalareaabovethecontactsurface.

In fig. 2.14(b) the beam is strengthened by riveting a channel section to each flange. The
InFig.2.14(b),thebeamisstrengthenedbyrivetingachannelsectiontoeachflange.Thehorizonta
horizontal shear force acting between each channel and the main beam must be transmitshear
between
the main
beam must
transmitted
by rivets. We
ted
by force
rivets. acting
We calculate
thiseach
forcechannel
from theand
shear-flow
formula,
with Qbe
calculated
using
calculatethisforcefromtheshearflowformula,withcalculatedusingtheshadedareaasshown.
the
shaded area as shown.
In fig. 2.14(c), we have a wooden beam with two flanges and two webs that are connected
by nails. The total horizontal shear force between23
the upper flange and the webs is the
shear
flow
acting
along
both
cc
and
dd,
and
therefore
Q is calculated for the upper flange.

16

InFig.2.14(c),wehavegotawoodenbeamwithtwoflangesandtwowebsthatareconnectedby
nails. The total horizontal shear force between the upper flange and the webs is the shear flow
actingalongbothand,andthereforeiscalculatedfortheupperflange.
y

(a)
BeamswithAxialLoads

(b)

Exercise

(c)

Figure 2.14

Beams
with axial loads
Problemstobedonefromthechapter
Probs.5.54;5.555.56;5.58;5.5155.517;5.519;5.521&5.61.

2.17 Exercise

Problems to be done from the chapter:

Chapter 3 Stresses in Beams- Advanced Topics

Problems 5.54; 5.555.56; 5.58; 5.5155.517; 5.519; 5.521 and 5.61.

Introduction

Wewillbestudyingthebendingofbeamsbyexaminingseveralspecializedtopicsthatwealready
havedealtwithinthepreviouschaptersuchascurvature,normalstressinbeamsandshearstresses
inbeams.Therestrictionsthatbeams

needtocomposeofonematerial
needtohaveaplaneofsymmetryinwhichtraverseloadsmustbeapplied

Attheendofthechapterwewillbeextendingtheperformanceintotheinelasticrangeofbehaviour
forbeamsmadeofelastoplasticmaterial.

Composite Beams

Beams that are fabricated of more than one material are called composite beams. The examples
includebimetallicbeams,plasticcoatedpipesandwoodbeamswithsteelreinforcingplates.

Many other types of composite beams have been developed in recent years, primarily to save
materialandreduceweight.Readmoreonpg.457.

Strains and Stresses


Thelongitudinalstrain inacompositebeamvarylinearlyfromtoptobottomofthebeam,

SOM401M/1

ReadmoreonPgs.458and459

(3.1)

24

17

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 3
Stresses in Beams Advanced
Topics

3.1 Introduction
We will be studying the bending of beams by examining several specialised topics that we
already dealt with in the previous chapter such as curvature, normal stress in beams and
shear stresses in beams. The following restrictions must be taken into account:

Beams need to be composed of one material.

Beams need to have a plane of symmetry in which traverse loads must be applied.

At the end of the chapter we will be extending the performance into the inelastic range of
behaviour for beams made of elasto-plastic material.

3.2 Composite beams


Beams that are fabricated of more than one material are called composite beams. Examples
include bimetallic beams, plastic-coated pipes and wood beams with steel reinforcing plates.
Many other types of composite beams have been developed in recent years, primarily to
save material and reduce weight.

Read more on page 457.

3.3 Strains and stresses


The longitudinal strain fx in a composite beam varies linearly from top to bottom of the
beam,

fx =-

y
=- ky 
t

(3.1)

SOM401M/1

19

Read more on pages 458 and 459.

SOM401M/001

3.4 Neutral
Neutral
Axis axis
The position of the neutral axis, the z-axis, is found from the condition that the resultant
Thepositionoftheneutralaxistheaxisisfoundfromtheconditionthattheresultantaxialforce
axial force acting on the cross-section is zero; therefore,
actingonthecrosssectioniszero,therefore,
(3.2)
# vx dA + #1 vx dA = 0 
1

(3.2)

where the first integral is evaluated over the cross-sectional area of material 1 and the
...whenitisunderstoodthatthefirstintegralisevaluatedoverthecrosssectionalareaofmaterial1
second integral evaluated over the cross-sectional area of material 2; replacing x1 and
andthesecondintegralevaluatedoverthecrosssectionalareaofmaterial2.Replacing
and
x2 with the following:
withthefollowing:


vx =- E1 ky and vx =- E2 ky

Thereforetheresultingequationisasfollows:
Therefore the resulting equation is as follows:
1

E1#y
2# y
dA+E
dA=0 

(3.3) (3.3)

y
t

h
O
z

Figure 3.1

The integrals in this equation represent the first moments of the two parts of the cross-
Figure31
sectional area with respect to the neutral
axis.

The integrals in this equation represent the first moments of the two parts of the crosssectional
Equation 3.3 is a generalised form of the analogous equation for a beam of one material,
areawithrespecttotheneutralaxis.
ie equation 5.8 in the textbook.
Eq.(3.3)isageneralizedformoftheanalogousequationforabeamofonematerial,i.e.,Eq.(5.8)
If the cross-section of a beam is doubly symmetric, as in the case of a wood beam with
fromthetextbook.
steel cover plates on the top and bottom (see fig. 3.1), the neutral axis is located at the
mid-height of the cross-section and equation 3.3 is not needed.
Ifthecrosssectionofabeamisdoublysymmetric,asinthecaseofawoodbeamwithsteelcover
platesonthetopandbottom...seeFig.3.1,theneutralaxisislocatedatthemidheightofthecross
sectionandEq.(3.3)isnotneeded.

3.5 Moment-curvature relationship


Moment-Curvature Relationship

20

The moment-curvature relationship for a composite beam of the two materials may be
determined from the condition that the moment resultant of the bending stresses is equal
Themomentcurvaturerelationshipforacompositebeamofthetwomaterialsmaybedetermined
to the
bending
moment
M acting
at the
cross-section.
from
the
condition
that the
moment
resultant
of the bending stresses is equal to the bending
momentMactingatthecrosssection.ReadmoreonPg.460.

(3.4)

Read more on page 460.


M

k= 1 =
t
E1 I1 + E2 I2

(3.4)

3.6 Normal stresses

Read about this on page 460.

3.7 Approximate theory for bending of sandwich beams


If the material of the faces (material 1) has a much larger modulus of elasticity than does
the material of the core (material 2), it is reasonable to disregard the normal stresses in
the core and assume that the faces resist all of the longitudinal bending stresses. This assumption is equivalent to saying that the modulus of elasticity E2 of the core is zero. Under
these conditions the flexural formula for material 2 (equation 6.6b) gives vx = 0 to give
2

My

vx =I1
1

(3.5)

which is similar to the ordinary flexural formula. The quantity I1 is the moment of inertia of
the two faces evaluated with respect to the neutral axis.

Read more on pages 461 and 462. Do examples 6.1 and 6.2.

3.8 Transformed-section method


Skip this section.

SOM401M/1

3.9 Doubly symmetric beams with inclined loads

In this section, the bending stresses in the beam are determined by resolving the inclined
load into two components, one acting in each plane of symmetry. The bending stresses
can be obtained from the flexure formula for each load component acting separately, and
the final stresses can be obtained by superposing the separate stresses.

21

3.10 Normal stresses (bending stresses)


The normal stresses associated with the individual bending moments My and Mz are obtained from the flexure formula (equation 5.13). These stresses are then superposed to
give the stresses produced by both moments acting simultaneously.
A positive moment My produces tension at this point and a positive moment Mz produ
ces compression; thus, the normal stress at point A is

vz =

Mz y
My z

Iy
Iz

(3.6)

Using the above equation, we can find the normal stress at any point in the cross-section
by substituting the appropriate algebraic values of the moments and the coordinates.

3.11 Neutral axis


SOM401M/001

The equation of the neutral axis can be determined by equating the normal stress vz
(equation 3.6) to zero:
M

y z Mz y
vz =
0 = -
I
I
y
z

Thisequationshowsthattheneutralaxisisastraightlinepassingthroughthecentroid.Theangle
This equation shows that the neutral axis is a straight line passing through the centroid.
The angle between the neutral axis and the z-axis is determined as follows:
betweentheneutralaxisandtheaxisisdeterminedasfollows:

y
M y Iz

tanb = =
Mz I y
z

(3.7)
(3.7)

Knowing the orientation of the neutralaxisis useful


when determining the points in the
Knowing the
orientation
of
the
neutral
axis
is
useful
when
determining
the points in the cross
cross-section where the normal stresses are the largest.
sectionwherethenormalstressesarethelargest.
See figures 6.11 to 6.15.
SeeFigs.611to615

Relationship
N.A and
the inclination
of the loads
3.12 between
Relationship
between
N.A. and
inclination of the
loads
Acantileverbeamisusedasanexample...;seeFig.3.2(a).Thebeamisloadedbyforceactingin
theplaneoftheendcrosssectionandinclinedatanangletothepositiveaxis.
We will use a cantilever beam as an example (see fig. 3.2). The beam is loaded by force P
acting in the plane of the end cross-section and inclined at an angle to the positive y-axis.
TheloadPcanberesolvedintoitsrespectivecomponents inthepositiveydirectionand
The load P can be resolved into its respective components P cos i in the positive y direction
and acting
cross
in the
Therefore
theTherefore
bending moments
moments
and positive
the positive
z direction.
the bending
My andon
Ma
acting
P sin i indirection.
z
on
a
cross-section
located
at
distance
x
from
the
fixed
support
are
sectionlocatedatdistancefromthefixedsupportare

22

Figure32
Figure 3.2

Figure 3.3
in which L is the length of the beam. The ratio of these moments is

My
= tan i 
Mz

(3.8)

which shows that the resultant moment vector M is at the angle from the x-axis in
Fig 3.3. The resultant moment vector is perpendicular to the longitudinal plane containing
the force P. The angle between the neutral axis and the z-axis is obtained from equation
3.9 which shows that the angle is generally not equal to the angle . Thus, except in
special cases, the neutral axis is not perpendicular to the longitudinal plane containing
the load. These exceptions are mentioned on page 474 where you can read more about
them. Do example 65. Skip example 64.

tanb =

My I z
I
= z tani 
Mz I y
Iy

(3.9)

3.13 Bending of asymmetric beams


In this section we will be dealing with beams with asymmetric cross-sections. We will begin
by investigating beams in pure bending, and then consider the effects of lateral loads. The
assumption that the beams are made of linearly elastic materials still holds.
When bending couples act in a plane of symmetry of the member, either because they act
in a different plane, or because the member does not possess any plane of symmetry, the
situation is referred to as asymmetric bending.

3.14 Neutral axis

SOM401M/1

Read more on pages 480481.

23

3.15 Procedure for analysing an asymmetric beam


n

Myy

Mzz

Figure 3.4
Let us consider fig. 3.4 as shown above, and describe a general procedure for analysing
an asymmetric beam subjected to any bending moment M. We will begin by locating the
centroid C of the cross-section and constructing a set of principal axes at that point (the
y- and z-axes) in the figure. Next, the bending moment is resolved into components My
and Mz , positive and negative, as shown in the figure.

Mx = M sin i and My = M cos i 

(3.10)

in which is the angle between the moment vector M and the z-axis. Since each component
acts in a principal plane, it produces pure bending in that same plane.
The superposition of the bending stresses in order to obtain the resultant stress at any point
in the cross-section is given by the following equation:

vx =

Mz y
M cos iy
My z

= M sin iz +
Iy
Iz
Iy
Iz

(3.11)

in which y and z are the coordinates of the point under consideration.


The equation of the neutral axis is obtained by setting vz = 0 and simplifying:

sin i z - cos i y = 0 
Iy
Iz

(3.12)

The angle between the neutral axis and the z-axis can be obtained from the preceding
equation as follows:

tan b =

y
I
= z tan i 
z
Iy

(3.13)

3.16 Deflection due to asymmetrical bending


In asymmetrical bending, as in symmetrical bending, the deflection will occur perpendicular to the neutral plane. At any section along the beam the deflection may be calculated
using the deflection formula. For example, in the case of a cantilever of length L, carrying
a load P at the free end, the maximum deflection of the centroid of the cross-section will
be given by

24

3
d = PNA L
3EINA

(3.14)

where PNA is the component of the load perpendicular to the neutral axis and the INA is the
second moment of the area about the neutral axis. The most convenient way to obtain INA is
to use the coordinates (Iy , Iyz) and Iy Iyz to construct a Mohrs circle for moments. This then
allows the second moment of area at any angle to the y- or z- direction to be determined.

PNA = P cos b
Do examples 6.6 and 6.7 in the textbook.

3.17 The shear-centre concept


In this section we will be considering fig. 3.5. The beam is called an unbalanced I-beam.
The normal stresses acting on the cross-section have a resultant that is the bending moment
Mo, and the shear stresses have a resultant that is the shear force P. Now, if the material
follows Hookes law, the normal stresses will vary linearly with the distance from the neutral
axis (z-axis) and can be calculated from the flexure formula.
Since the shear stresses acting on a cross-section are determined from the normal stresses
solely from equilibrium considerations, it follows that the distribution of shear stresses over
the cross-section is also determined.

Read more on pages 488489.

o
M0
P
Figure 3.5

3.18 Shear stresses in beams of thin-walled open crosssections


The following formula was derived for calculating shear stresses:
SOM401M/1

x=

VQ

Ib

(3.15)

In the above formula, V represents the shear force acting in the cross-section, I is the moment of inertia of the cross-sectional area, b is the width of the beam at the location where
the shear stress is to be determined, and Q is the first moment of the cross-sectional area
outside the location where the stress is being found. Typical beams are displayed in the
25

textbook on page 489. Now, we will be considering the shear stresses in a special class
of beams known as thin-walled open cross-section. The figure that we need to look at is
fig. 6.31 on page 490.
Please page through for this figure because I have omitted it from the manual.

dM = Mx - Mx = Vy 
dx
dx

x=

Vy
Iz t

# ydA 
s

(3.16)
(3.17)

The above equation, equation 3.18, gives the shear stress at any point in the cross-section
at distances from the free edge. The integral on the right-hand side represents the first moment with respect to the z-axis of the area of the cross-section from s = 0 to s = s. Denoting
this moment by Qz, we can write the equation for the shear stresses in the simpler form:

x=

Vy Qz

Iz t

(3.18)

Read more on pages 491 and 492.

3.19 Shear stresses in wide-flange beams


In this section we will consider a beam as shown in fig. 6.32a, b, c and d on page 493 in
the textbook. This beam is loaded by a force acting in the plane of the web, that is, through
the shear centre, which coincides with the centroid of the cross- section. The dimensions of
the cross-section of the beam are shown in fig. 6.32a while the shear distribution diagram
is shown in fig. 3.5.

3.20 Shear stresses in the upper flange


The shear stress in the flange is given by:

P ` st f h 2 j
Vy Qz
xf =
=
= shP 
Iz t f
Iz t f
2Iz

(3.19)

Make sure that you redo this derivation because I am definitely going to ask something
like this in the examination.

26

1
2

max

2
1

Figure 3.6
The variation of the stresses in the upper flange is shown graphically in fig.3.6, and we
see that the stresses vary from zero at point a to a maximum value T1

x1 = bhP 
4Iz

(3.20)

Note that we have calculated the shear stress at the junction of the centrelines of the flange
and web, using only the centreline dimensions of the cross-section in the calculations. This
approximate procedure simplifies the calculations and is satisfactory for thin-walled cross
sections.

3.21 Shear stress in the web


The next step is to determine the shear stresses acting in the web. The stress that corresponds to the first moment Qz = bt f h is found to be:

x2 =

bht f P

2Iz tw

(3.21)

Read more about this section on pages 494496. This is an important section in the book
and we will be doing more examples in this regard. Do the derivation of the necessary
shear stresses.

SOM401M/1

3.22 Shear centres of thin-walled open sections

In this section we are going to consider fig. 3.7. Only beams with singly symmetric or asymmetric cross-sections will be considered, because we already know that the shear centre
of a doubly symmetric cross-section is located at the centroid. The procedure for locating
shear centre consists of two principal steps: first, evaluating the shear stresses acting on
the cross-section when bending occurs about one of the principal axes, and secondly, de27

termining the resultant of those stresses. The shear centre is located on the line of action
of the resultant. By considering bending about both principal axes, we can determine the
position of the shear centre.
y
V

F1
e

h
2

z
F2

F3

h
2

Figure 3.7

Figure 3.8

3.23 Channel section


As shown in fig. 3.8, the shear stresses in a channel vary linearly in the flanges and parabolically in the web. The expression for e is shown below; you can read more on pages
497500.

e=

3b2 t f

htw + 6bt f

(3.22)

I strongly urge you to redo this derivation and I will put the derivation on the website in
the latter part of the year.

28

3.24 Angle section

Read about this on pages 500501. This is an important section for you to know as a
practising engineer. I have worked many times with structural design and a lot of times I
have used channels and angles to build large structures.

Read about this section on pages 496502 and do example 68 in G&G.

3.25 Z-section

Read more about this on page 502.

3.26 Elasto-plastic bending


This is an important chapter in Mechanical Engineering because it forms part of your life
as a mechanical engineer. What do I mean by this? Well, if you are going to be involved in
design, you will be using this and other theories such as fatigue analysis, fracture analysis,
etc. Make sure that you understand it and apply it in your design project.
Elasto-plastic materials follow Hookes law up to the yield stress y and then yield plastically under constant stress.
Structural steels are excellent examples of elasto-plastic materials because they have sharply
defined yield points and undergo large strains during yielding.

3.27 Yield moment

SOM401M/1

We begin by considering a beam of elasto-plastic material subjected to a bending moment


that causes bending in the xy plane. When the bending moment is small, the maximum
stress in the beam is less than the yield stress, and therefore the beam is in the same condition as a beam in ordinary elastic bending with a linear stress distribution. See fig. 3.9.

29

c
r

y
Figure 3.9

The bending moment in the beam when the maximum stress just reaches the yield stress,
called the yield moment , My can be obtained from the flexure formula:

My = vy Ic = vy S
in which c is the distance to the point farthest from the neutral axis and S is the corresponding section modulus.

3.28 Plastic moment and neutral axis


If we now increase the bending moment above the yield moment, the strains in the beam
will continue to increase and the maximum strain will exceed the yield strain. In fig. 3.9(a)
note that the outer regions of the beam have become fully plastic while a central core
remains linearly elastic. To find the plastic moment, you need to locate the neutral axis of
the cross-section under fully plastic conditions. The resulting equation of the plastic moment is shown below:
y1 + y2h
v A^
(3.23)

MP = y

3.29 Plastic modulus and shape factor


Plastic moment can be expressed as follows:

in which
30

MP = v y Z 

(3.24)

Z=

A^y1 + y2h

2

(3.25)

is the plastic modulus or the plastic section modulus for the cross-section.
The ratio of the plastic moment to the yield moment is solely a function of the shape of
the cross-section and is called the shape factor f:

f = MP = Z 
MY
S

(3.26)

3.30 Beams of rectangular cross-section


Let us consider a beam of cross-section as shown in fig. 3.10. The beam is made of elastoplastic material. The section modulus is
2
S = bh
6

h
2

h
2

Figure 3.10
The yield moment is given as follows:

MY =

vy bh2

6

(3.27)

in which b is the width and h is the height of the cross-section. The distances to the centroids
of the areas above and below the neutral axis are

y1 = y2 = h
4
The plastic modulus is derived as follows:

Z=

2
A^y1 + y2h
= bh ` h + h j = bh
2
2 4 4
4

and the plastic moment is

MP =

vy bh2

4

(3.28)

SOM401M/1

31

Read more about this section on pages 509 510 and do examples 69 and 610 in G&G.

3.31 Beams with wide-flange shape


Read and do the derivation on page 510 in G&G.

3.32 Exercise
You are required to do the following problems from the textbook:
6.211; 6.39; 6.31112; 6.414; 6.46; 6.49; 6.41213; 6.518; 6.512;
6.813; 6.9311; 6.1045, 1113, 1819.

32

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 4
Deflections of Beams

4.1 Introduction
As you will notice most procedures of finding beam deflections are based on the differential
equations of the deflection curve.

4.2 Differential equation of the deflection curve


We are going to consider a cantilever beam with a concentrated load acting upward at
the free end as shown in fig. 4.1(a). The beam deforms into a curve as shown in fig. 4.1(b)
with the x-axis directed to the right and the y-axis directed upward.
The z-axis is directed outward from the figure towards you.

B
v

L
P

(a)

(b)
Figure 4.1

The assumption that we can make here is that xy is a plane of symmetry of the beam and
that all loads act in this plane.
The deflection v is the displacement in the y direction. Because the y-axis is positive upward,
the deflections are also positive when upward.

SOM401M/1

Read more on pages 680681 about the derivation.

4.3 Beams with small angles of rotation


If the material of a beam is linearly elastic and follows Hookes law, the curvature is
33

=1 = M
t
EI

(4.1)

The basic differential equation of the deflection curve of a beam is given by:

d2 v = M 
EI
dx2

(4.2)

The above equation can be integrated in each particular case to find the deflection v,
provided the bending moment and flexural rigidity are known as a function of x.
The sign conventions to be used with the preceding equations are as follows:
1
2
3
4
5

The x- and y-axes are positive to the right and upward, respectively.
The deflection v is positive upward.
The slope dv
dx and angle of rotation are positive when counter-clockwise with
respect to the positive x-axis.
The curvature k is positive when the beam is bent concave upward.
The bending moment M is positive when it produces compression in the upper part
of the beam.

4.4 Non-prismatic beams


The flexural rigidity is variable so the equation is written as follows:

2
EIx d v2 = M 
dx

(4.3)

where the subscript x means that the flexural rigidity may vary with x.

d EIx d2 v = dM = V 
c 2m
dx
dx
dx

(4.4)

d2 EIx d2 v = dV =- q 
c 2m
dx
dx
dx2

(4.5)

4.5 Prismatic beams


The differential equations become

2
3
4
EI d v2 = M, EI d v3 = V and EI d v4 =- q 
dx
dx
dx

(4.6)

These equations are referred to as the bending-moment equation, the shear-force equation
and the load equation respectively.

4.6 Deflection by integration of the bending-moment


equation
The general procedure for solving the differential equations is as follows: For each region of
the beam we substitute the expression for M into the d.e and integrate to obtain the slope v.
Next, we integrate each slope equation to obtain the corresponding deflection v. Each
integration produces a new constant thus there are two constants of integration for each
region of the beam.
34

These constants are evaluated from known conditions pertaining to the slopes and deflections.
The conditions fall into three categories:
1
2
3

Boundary conditions
Continuity conditions
Symmetry conditions

Boundary conditions pertain to the deflections and slope at the supports of a beam. At a
fixed end v = v = 0, and at either the pin or roller support v = 0 only. Each boundary
condition supplies one equation that can be used to evaluate the constants of integration.
Continuity conditions occur at points where the regions of integration meet, such as at
point C as shown in fig. 4.2.

B A

Figure 4.2
Symmetry conditions occur when a beam is simply supported and carries a uniform load
throughout its length. This allows us to know in advance that the slope of the deflection
curve at the midpoint must be zero. This condition supplies us with an extra equation.

4.7 Deflection by integration of the shear-force and load


equations
The equations of the deflection curve in terms of the shear force V and the load q may
also be integrated to obtain slopes and deflections. The procedure for solving either the
load equation or the shear force equation is similar to that for solving the bending-moment
equation, except that more integrations are required.
Conditions of the shear forces are equivalent to conditions on the third derivative (EIv = V).
In a similar manner, conditions on the bending moments are equivalent to conditions on
the second derivative (EIv = M). When the shear-force and bending-moment conditions
are added to those for the slopes and deflections, we always have enough independent
conditions to solve for the constants of integration.
Do examples 94 and 95.

SOM401M/1

4.8 Method of superposition

The effect of a given combined loading on a structure may be obtained by determining


separately the effects of the various loads and then combining the results obtained.
This type of superpositioning is only possible if each effect is linearly proportional to the
load which produces it. Also the deformation resulting from any given load should be small
enough that it does not affect the condition of application of other loads.
35

Let us consider an arbitrary beam subjected to distributed loads (q1, q2 and q3) with
accompanying bending moments at a section of the beam caused by each load when
acting on the beam ie, M1, M2 and M3.
But

M = M1 + M2 + M3 + g

Therefore

M = EI d v2
dx

v= 1
EI
v= 1
EI

# ` # ^M + M
1

+ M3 + gh dxj dx

# ` # M dxj dx + # ` # M dxj dx + # ` # M dxj dx + g


1

v = v1 + v2 + v3 + g
Thus the deflection at a section of a beam subjected to complex loading can be obtained
by the summation of the deflections caused at that section by the individual components
of the loading.
When dealing with problems of superpositioning you might find Appendix G in the textbook
handy, that is if you cannot quickly derive the deflection or slope equation. In the examination I will provide the table with the necessary information that you will need from it.

4.9 Moment-area method


Skip this section.

4.10 Non-prismatic beams


Skip this section.

4.11 Strain energy


Let us consider a simple beam that is in pure bending under the action of two couples as
shown in fig. 4.3.

M
A

L
L
Figure 4.3
As the bending couples gradually increase in magnitude from zero to their maximum
values, they perform work W. This work, equal to the strain energy U stored in the beam,
is given as follows:
36

W = U = Mi 
2

(4.7)

But on the other hand, the angle subtended by the arc as shown in fig. 4.3 is given by the
following formula:

i = ML 
EI

(4.8)

So if we now combine the two equations, we get either one of the following equations:

2
2
U = M L or U = EIi 
2EI
2L

(4.9)

in which the first one expresses the strain energy in terms of the applied moments M, and
the second one expresses the strain energy in terms of the angle . For a strip of element,
2
dU = M dx
2EI

# M2EIdx
2

`U=

the bending moment varies as a function of x.


We can also write the equation that is expressed in the angle as follows:
2
EI^di h2
EI d 2 v
EI d2 v 2

dU =

2dx

dx m =
c
c
m dx
2dx dx2
2 dx2

`U=

d v dx
# EI
c
m
2 dx
2

4.12 Deflections caused by a single load


The strain energy of a beam is equal to the work done by the load, so what we need to
do is to relate and to the load P and bending moment Mo as follows:

U = W = Pd and U = Moi
2
2
The first equation applies to a beam that is loaded with only a load P and the second
equation applies to a beam loaded by only a couple Mo. We are then able to write the
following by just manipulating the above equations:

d = 2U
P

and i = 2U
Mo

Do examples 915 and 916.

4.13 Castiglianos theorem


What is Castiglianos method or theorem and how can it help us? Castiglianos method
is a unique way of analysing deflections of structures from the strain energy and it can be
used to find reactions of indeterminate structures.

SOM401M/1

Let us consider a cantilever beam with a concentrated load, P acting at the free end as
shown in fig. 4.4. The strain energy is found to be

37

2 3
U= P L
6EI

P
A

B
A

L
Figure 4.4
The derivative of the above with respect to the load is given as follows:

dU = PL3
dP
3EI

4.14 Derivation of Castiglianos theorem


I will not discuss this here, but will put the notes of the derivation on the web. Make use
of the formulae.
Theorem 1: When forces act on elastic systems subjected to small displacements, the displacement corresponding to any force, collinear with the force, is equal to the partial derivative of
the total strain energy with respect to that force. The partial derivative of the strain energy of
a structure with respect to any load is equal to the displacement corresponding to that load.
Mathematically put it follows that:

di = 2U 
2Pi

(4.10)

where i is the displacement of the point of application of the force Pi in the direction of Pi.
For rotational displacement

ii = 2U 
2Mi

(4.11)

where i is the rotational displacement, in radians, of the moment Mi in the direction of Mi.

4.15 Application of Castiglianos theorem


Let us consider a cantilever beam as shown in fig. 4.4, with a point load P acting at the
free end. Our aim is to determine the following:
1
2

The strain energy


The deflection due to the applied load

Solution-strain energy:

Mxx =- Px
The strain energy is given by the following formula:

U=

#
0

38

M2 dx =
2EI

#
0

P2 x2 dx
2EI

Therefore, the strain energy for the loaded cantilever beam is


2 3
U= P L
6EI

Solution-deflection:
The deflection at the end of the beam due to load P is
3
dP = 2U = PL
2P
3EI

4.16 Use of a fictitious load


If we were to determine the deflection at a particular point on the beam where there is
no load, then a fictitious load or dummy load corresponding to the desired displacement
must be applied to the beam. Displacements are found by evaluating the strain and taking
the partial derivative with respect to the dummy load. By setting the dummy load equal to
zero, we are able to obtain the displacements produced only by the actual loads.
Let us consider a cantilever beam shown in fig. 4.5. We are required to determine the
deflection at point C but there is no load acting at that point. Therefore we need to apply
a dummy load Q acting downward at point C.

P
A
Mo

C
L

/2

/2

c
Figure 4.5

NB: With beams we are often required to determine the slopes as well. In order to do that,
we are always required to determine the couple at the application of the load in order to
keep the beam stable, ie Mo.

Mxx =- Px - Mo
in which x is the distance from the free end. The strain energy is written as follows:
2
L _- Px - Mo i

U=

#
0

2EI

dx

The deflection at the free end where the load and the couple are acting is given by
2
3
dA = 2U = PL + M0 L
2P
3EI
2EI

The rotational displacement is given by

SOM401M/1

i A = 2U
2Mo

Example No 1- Castiglianos theorem


A simple beam AB supports a concentrated load as shown below in fig 4.6. We are required
to determine the deflection due to the applied load at point C.
39

The reactions are as follows:


P
a

x
A
x

B
z

RA

RB

Figure 4.6

RA = Pb and RB = Pa
L
L
We then cut the beam into two sections XX and YY, with the resulting moments about the
sections being as follows:

Mxx = Pbx and Mzz = Paz


L
L
The deflection is as follows:

d=

d = d1 = d2 =

d= 1
EI

#
0

M 2M ds
EI 2P

Mxx 2Mxx ds +
EI 2P

Pbx bx dx + 1
L L
EI

Mzz 2Mzz ds
EI 2P

Paz az ds
L L

2 2
2 2
` d = Pb2 a ^a + bh = Pb a
3LEI
3L EU

4.17 Differentiation under the integral sign


Using Castiglianos theorem for determining the deflection in a beam may lead to lengthy
and cumbersome integrations, especially if we have more than two loads. After determining the strain energy, we need to differentiate the strain energy to obtain the deflections.
We can actually bypass the step of finding the strain energy by differentiating before
integrating.
Let us begin with the equation for the strain energy:

U=

# 2MEI dx
2

and apply Castiglianos theorem

di = 2U
2Pi
We then differentiate the integral by differentiating under the integral sign:

40

di = d
dPi

M 2M dx 
# 2MEI dx = # ` EI
j c 2P m
2

(4.12)

We can now refer to equation 4.12 as the modified Castiglianos theorem. When using
the modified theorem, we actually integrate the product of the bending moment and its
derivative.
Let us refer to fig. 4.5 Recall that we want to determine the deflection and the angle of
rotation at the free end. The bending moment as has been derived is given by

Mxx =- Px - Mo
Now the partial derivatives of the load as well as the moment are as follows:

2M =- x and 2M =- 1
2P
2Mo
From equation 5.12 we obtain the deflection A and the angle of rotation A,

dA = 1
EI

iA = 1
EI

2
3
^- Px - Moh^- xh dx = PL + Mo L

3EI

2EI

2
^- Px - Moh^- 1h dx = PL + M0 L

2EI

2EI

The advantages of differentiating under the integral sign are even more apparent when
there are more than two loads acting on the structure.
Other types of Castiglianos method will be dealt with later, such as curved beams and
frame structures.

4.18 Deflections produced by impact


Skip this section.

4.19 Temperature effects

SOM401M/1

Skip this section.

41

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 5
Statically Indeterminate Beams

5.1 Introduction
This chapter deals with the analysis of beams in which the number of reactions exceeds the
number of independent equations of equilibrium. The analysis is quite different from that
of statically determinate beams. When a beam is statically determinate, we can obtain all
reactions, shear forces, and bending moments from the free-body diagrams and equations of equilibrium. However, when a beam is statically indeterminate, the equilibrium
equations are not sufficient and additional equations are needed. The most fundamental
method for analysing a statically indeterminate beam is to solve the differential equations
of the deflection curve.
We will also be using the method of superposition, which is applicable to a variety of
structures.
In this method, we supplement the equilibrium equations with compatibility equations and
force displacement equations.
We are not going to do section 6.6 (G&G), but you are welcome to do it on your own. If
you experience any difficulty with it, call me.

5.2 Types of statically indeterminate beams


These types of beams are usually identified by the arrangement of their supports, ie
a beam that is fixed at one end and simply supported at the other end, shown in the
figure 5.1 below. This type of beam is referred to as a propped cantilever.

MA
RxA

P
A

y
RA

B
O

P
O

RB

SOM401M/1

Figure 5.1

The reactions of the beam shown in the figure consist of horizontal and vertical forces at
support A, a moment at support A, and a vertical force at support B. We can see that there
are only three independent equations of equilibrium for this beam; as such it is not possible
to calculate all four of the reactions from equilibrium alone. The number of reactions in

43

excess of the number of equilibrium equations is called the degree of static indeterminacy.
Thus, a propped cantilever beam is statically indeterminate to the first degree.
The excess reactions are called static redundants and must be selected in each particular
case.
For example, the reaction R Aof the propped cantilever beam shown in fig. 5.1 (G&G) may
be selected as the redundant reaction. Since this reaction is in excess of those needed to
maintain equilibrium, it can be released from the structure by removing the support at B.
When support B is removed, we are left with a cantilever beam shown by fig. 5.1 (G&G).
The structure that remains when the redundants are released is called the released structure or the primary structure. The released structure must be stable (so that it is capable
of carrying loads), and it must be statically determinate (so that all force quantities can be
determined by equilibrium alone).
We can opt to choose the reactive moment MA as the redundant from fig. 5.1 (G&G).
When the moment restraint at support A is removed, the released structure is a simple
beam with a pin support at one end and a roller support at the other as shown in fig. 5.1.
These types of beams are usually identified by the arrangement of their supports, ie a beam
that is fixed at one end and simply supported at the other end, shown in fig. 5.1 (G&G).
This type of beam is referred to as a propped cantilever.
Another type of statically indeterminate beam known as a fixed-end beam is shown in
fig. 5.2.
It has fixed supports at both ends, resulting in a total of six unknown reactions. Because
there are only three equations of equilibrium, the beam is statically indeterminate to the
third degree.
We can start by selecting the three reactions at end B as redundants, and if we remove
the corresponding restraints, we are then left with a cantilever beam as shown in fig. 5.2
(G&G). If we release the two fixed-end moments and one horizontal reaction, the released
structure is a simple beam as shown in fig. 5.2.

Figure 5.2

Read more about this in section 6.2 on pages 774777.

44

5.3 Analysis by the differential equations of the deflection


curve
Staticallyindeterminatebeamsmaybeanalysedbysolvinganyoneofthethreedifferential
equationsofthedeflectioncurve:
1
2
3

Thesecond-orderequationintermsofthebendingmoment
Thethird-orderequationintermsoftheshearforce
Thefourth-orderequationintermsoftheintensityofthedistributedload

Theprocedureconsistsofwritingthedifferentialequation,integratingtoobtainitsgeneralsolution,andthenapplyingboundaryandotherconditionstoevaluatetheunknown
quantities.
Theunknownsconsistoftheredundantreactionsaswellastheconstantsofintegration.

5.4 Continuous beam on multiple supports


5.4.1

Bending moments at supports

Thissectionisnotincludedinthebookbutitisveryimportant!
Abeamrestingonmorethantwosupportsistermedcontinuous.Suchabeamisrepresentedinfig.5.3.Changesofcurvatureoccurineachspanowingtonegativebending
momentsatthesupports.

P
qkN/m
A

qkN/m
B

Figure5.3
Insolvingthiskindofproblem,weneedtofirstworkouttheleft-handsideaswellas
theright-handsideofthefollowingequation,sometimesseparatelyespeciallyifyou
becomeuncertain:

- MA L1 - 2MB ^ L1 + L2h - Mc L2 = 6 c A1 x 1 + A2 x 2 m
L1
L2

(5.1)

Forau.d.l,theright-handsideoftheequation5.1

SOM401M/1

qL23

(5.1a)

andforapointload,theright-handsideoftheequation

= Pa ^ L22 - a2h
L2

(5.1b)

45

a is the distance from one support to the location of the load. If the point load is at the
centre of the span, our choice of a becomes quite easy but if the point load is otherwise,
we need to determine whether the point load comes first or last in which case the distance
that we need to take will be ^ L - ah .
This will be explained in a set of examples that will be dealt with in the Sample Problem
Workbook that can be downloaded from the myUNISA website.

5.5 Reactions at supports


In order to calculate the support reactions, we need to take moments about point B and
only considering forces and reactions from either side. NB: We consider the effects of only
the forces from one side, i.e. the left- or right-hand side. The reactions are always taken
as having a positive effect. The way we do this is as follows:

/M

= Rc L2 - P^ L - bh 

(5.1c)

if b is considered the distance from the right-hand support of the continuous beam. The
other reaction can be found from equilibrium of forces.

5.6 Method of superposition


This method is of fundamental importance in the analysis of statically indeterminate bars,
trusses, beams, frames and many kinds of structures. We begin the analysis by taking note
of the degree of static indeterminacy and selecting the redundant reactions. Having done
that, we can then begin writing equations of equilibrium that relate the unknown reactions
to the redundants and the loads.
The next thing that we do is to assume that both the original loads and the redundants
act upon the released structure. This helps us to find the deflections due to the loads and
the redundants. The sum of these deflections must match the deflections in the original
beam. The deflections in the original beam (at the points where restraints were removed)
are either zero, or have known values.
Since the released structure is now statically determinate, we can easily determine its deflections by using the techniques described in chapter 5.
Let us reconsider the beam in example 101 that was analysed in section 103 (G&G).

5.7 Analysis with RB as redundant


We select the reaction R B as redundant at the simple support. The equations of equilibrium
become:

RA = qL - RB MA =

qL2
- RB L
2

The next step is to remove the restraint corresponding to the redundant (ie the support at B).
The released structure that remains is a cantilever beam (see fig.10.12b in G&G). The
uniform load q and the redundant force RB are now applied as loads on the released
structure (see fig. 10.12c and d in the textbook).
46

The deflection at end B of the released structure due to the uniform load is denoted by,
^dBh1 and the deflection at the same point due to the redundant is denoted by ^dBh2 . The
deflection at point B ^dBh is obtained by superposing these two deflections as follows:

dB = ^dBh1 - ^dBh2
Using table G-1 in appendix G, we will get the following expressions for the deflections:

^dBh1 =

qL4
8EI

and

3
^dBh2 = RB L

3EI

Since the deflection at B is zero, the reaction at B can be then determined, ie

dB =

3
qL4
- RB L = 0
8EI
3EI

The remaining reactions (R A and MA) can be found from equilibrium equations:

RA =

5qL
qL2
and MA =
8
8

Complete this by analysing the beam with MA as redundant and finding the slopes in the
process.
In the examination you are free to use any method to solve the question.
Do example 103105 from Gere and Goodno.

5.8 Temperature effects


Skip this section.

5.9 Longitudinal displacements at the ends of a beam


Skip this section.

5.10 Exercise
Do the following problems in chapter 6:

SOM401M/1

10.31, 10.32, 10.35, 10.39, 10.310, 10.41, 10.43, 10.47, 10.410 and 10.421

47

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 6
Columns

This chapter deals with the buckling of slender columns which support compressive
loads in structures. The critical axial load which indicates the onset of buckling is defined
and computed for a number of simple models composed of rigid bars and elastic springs.
The differential equation of the deflection curve is derived and solved to obtain expressions for the Euler buckling load (Pcr) and associated buckled shape for the fundamental
mode. The critical stress cr and slenderness ratio L are defined and the effects of large
r
deflections, column imperfections, inelastic behaviour, and optimum shapes of columns
are defined.

6.1 Introduction
A load-carrying structure may fail in a variety of ways, depending upon the type of structure, the condition of support, the kinds of loads, and the material used. As an example,
an axle in a vehicle may fracture suddenly from repeated cycles of loading, or a beam
may deflect excessively, so that the structure is unable to perform its intended functions.
The strength and stiffness are important factors in design.
We will be dealing strictly with buckling and we are going to consider specifically buckling
of columns, which are long, slender structural members loaded axially in compression, ie
fig. 6.1a. If a compression member is relatively slender, it may deflect laterally and fail by
bending, as shown in fig. 6.1b rather than failing by direct compression of the material.
When lateral bending occurs, we say that the column has buckled. Under an increasing
axial load, the lateral deflections will increase too, and eventually the column will collapse.

P
B

SOM401M/1

(a)

(b)
Figure 6.1
49

6.2 Buckling and stability


Let us consider an idealised structure such as the one shown in fig. 6.2a. The structure
consists of two rigid bars AB and BC, each of length ` L j . They are joined at B by a pin
2
connection and held in a vertical position by a rotational spring with stiffness bR , in which
M = bR i where M is the moment acting on the spring, bR is the rotational stiffness of the
spring and is the angle through which the spring rotates.
Initially, the two bars are perfectly aligned and the axial load P has its line of action along
the longitudinal axis. The spring is initially unstressed and the bars are in direct compression.
P

P
C

BR

BR
B

MB

(a)

(b)

(c)

Figure 6.2
Suppose that the structure is disturbed by some external force that causes point B to move
a small distance laterally as shown in fig. 6.2b. The rigid bar rotates through small angles
and a moment develops in the spring. The direction of this moment is such that it tends to
return the structure to its original straight position, and hence it is called a restoring moment.
Consider what happens when the disturbance force is removed. If the axial force P is
relatively small, the action of the restoring moment will predominate over the action of the
axial force and the structure will return to its straight position. Under these conditions, the
structure is said to be stable. However, if the axial force P is large, the lateral displacement
of point B will increase and the bars will rotate through larger and larger angles until the
structure collapses.
Under these conditions, the structure is unstable and fails by lateral buckling.

6.3 Critical load


The transition between stable and unstable conditions occurs when the value of P reaches
the critical load Pcr. We can determine the critical load in our buckling model by considering the structure in a disturbed position as shown in fig. 6.2b.

M = 2b R i
But since the angle is a small quantity, the lateral displacement of point B is iL

50

` MB - P a iL k = 0
2
PL
a2bR - 2 k i = 0

One solution of this equation is = 0, which is a trivial solution and only means that the
structure is in equilibrium when it is perfectly straight, regardless of the magnitude of the
force P.
A second solution is obtained by setting the term

PL
`2bR - 2 j = 0
and solving for the load P, which will be the critical load:

Pcr =

4b R

L

(6.1)

If P < Pcr, the structure is stable.


If P > Pcr, the structure is unstable.

6.4 Columns with pinned ends


Let us begin the investigation of the stability behaviour of columns by analysing a slender
column with pinned ends as shown in fig. 6.3.
For the purpose of the analysis, let us construct a coordinate system with its origin at support A and with the x-axis along the longitudinal axis of the column. The y-axis is directed
to the left in the figure and the z-axis comes out of the plane of the figure towards us. An
assumption is made that the xy plane is a plane of symmetry of the column and that any
bending takes place in that plane.

P
M

v
L
y

x
A

Figure 6.3
When the axial load has a small value, the column remains perfectly straight and undergoes
direct axial compression. The only stresses are the uniform compressive stress obtained
from the equation

v= P
A

(6.2)

SOM401M/1

The column as mentioned before is in stable equilibrium.

51

As the load is gradually increased, we reach a condition of neutral equilibrium in which the
column may have a bent shape. The corresponding value of the load is the critical load Pcr.
At higher values of the load, the column is unstable and may collapse by buckling, that is
excessive bending.
If P < Pcr: the column is in stable equilibrium in the straight position.
If P = Pcr: the column is in neutral equilibrium in either the straight position or slightly bent
position.
If P > Pcr: the column is in unstable equilibrium in the straight position and will buckle
under the slightest disturbance.

6.5 Differential equation for column buckling


Although both the 4th order D.E (the load equation) and the 3rd order D.E. (the shear force
equation) are suitable for analysing columns, we will elect to use the 2nd order D.E (the
bending moment equation) because its general solution is usually the simplest:

EIv" = M 

(6.3)

in which M is the bending moment at any cross-section, V is the lateral deflection in the y
direction, and EI is the flexural rigidity for bending in the xy plane. (see fig. 6.4)
x
p
x

M
x

Figure 6.4
From equilibrium of moments about point A, we obtain

M + Pv = 0

` M =- Pv 

(6.4)

The differential equation of the deflection curve is therefore


EIv" + Pv = 0 

(6.5)

Solution of the D.E.


We introduce a notation that will help us write the D.E.

k=

P 
EI

in which k has units of the reciprocal of length.


52

(6.6)

We can therefore rewrite equation 6.4 in the form


v" + k2 v = 0 

(6.7)

and from 2nd and 3rd year mathematics, we know that the general solution of this equation is

v = C1 sin kx + C2 cos kx 

(6.8)

in which C1 and C2 are constants of integration (and are evaluated from the boundary
conditions or end conditions of the column).
Deflection is zero at x = 0 and at x = L

v = C1 sin kx

because

C2 = 0

The second condition gives

C1 sin kx = 0
Case 2 in which sinkL = 0 will be dealt with, because our aim is to find all the values of
kL in which

C1 sin kL = 0 

(6.9)

The buckling equation is therefore


sin kL = 0 

(6.10)

2 2
P = n r 2EI 
L

(6.11)

We find the buckling load to be



The deflection curve is

v = C1 sin kx = C1 sin nrx 


L

(6.12)

6.6 Critical loads


The lowest critical load for a column with pinned ends as shown in fig. 6.5(a) is obtained

SOM401M/1

when n = 1

2

Pcr = r EI
L2

(6.13)

with the corresponding buckled shape or mode shape


v = C1 sin rx 
L

(6.14)
53

x
x

C1

B
C1

C1
A

Figure 6.5

6.7 Critical stress


After finding the critical load for a column, we can calculate the corresponding critical
stress by dividing the load by the cross-sectional area.

2
vcr = Pcr = r EI2 
A
AL

(6.15)

in which I is the moment of inertia for the principal axis about which buckling occurs.

r=

I 
A

(6.16)

in which r is the radius of gyration of the cross-section in the plane of bending. The equation for the critical stress therefore becomes

2
vcr = Pcr = r E 2 
A
^ L/r h

(6.17)

in which L/r is a non-dimensional ratio called the slenderness ratio.

6.8 Columns with other support conditions


In practice there are many other end conditions such as fixed ends, free ends and elastic
supports.
The critical loads for columns with various kinds of support conditions can be determined
from the differential equation of the deflection curve by following the same procedure that
was used when analysing a pinned-end column.
The procedure is as follows:
1
2
3

54

The column is assumed to be in the buckled state. With this information, an expression for the bending moment in the column is obtained.
A differential equation of the deflection curve is obtained using the bending-moment
equation ^ EIv" = M h .
A general solution containing two constants of integration plus any other unknown
quantities is obtained.

4
5

Boundary conditions pertaining to the deflection and the slope are applied and a set
of simultaneous equations is obtained.
Finally, we solve those equations to obtain the critical load and the deflected shape
of the buckled column.

6.9 Column fixed at the base and free at the top


Let us consider an ideal column that is fixed at the base, free at the top, and subjected
to an axial load P as shown in fig. 6.6(a). The deflected shape of the buckled column is
shown in fig. 6.6(b).
Pcr

v
L
x

(a)

(b)

Figure 6.6
The bending moment at distance x is written as follows:

M = P ^d - v h 

(6.18)

where is the deflection at the free end of the column. The D.E. of the deflection curve
becomes

EIv" = M = P^d - vh 

(6.19)

in which I is the moment of inertia for buckling in the xy plane. Taking note that k2 = P
EI
, we can write equation 6.18 into the form

v" + k2 = k2 d 

(6.20)

This is a linear differential equation of the second order with constant coefficients.

SOM401M/1

Read more about the derivation on pages 835837.

The lowest critical load is found to be


2

Pcr = r EI
4L2

(6.21)

55

6.10 Effective lengths of columns


The critical loads for columns with various support conditions can be related to the critical
load of a pinned-end column through the concept of an effective length.
The effective length Le for any column is the length of the equivalent pinned-end column,
that is, it is the length of a pinned-end column having a deflection curve that exactly matches
all or part of the deflection curve of the original column.

6.11 Fixed-free column


Le = 2L 

(6.22)

Since the effective length is the length of an equivalent pinned-end column, we can write
a general formula for critical loads as follows: (see fig. 6.7)
2

Pcr = r EI
L2e

(6.23)
P

L=2L

Figure 6.7

6.12 Column with both ends fixed against rotation


Let us consider a column with both ends fixed against rotation in fig. 6.8.

Figure 6.8
The column is free to shorten under an axial load. The effective length of a column with
fixed ends is as follows:
56

Le = L 
2

(6.24)

The equation for the critical load therefore becomes:


2
Pcr = 4r 2EI 
L

(6.25)

An observation can be made that the critical load for a column with fixed ends is four times
that for a column with pinned ends.

6.13 Column fixed at the base and pinned at the top


The critical load and buckled mode shape for a column that is fixed at the base and
pinned at the top (see fig. 6.9), can be determined by solving the differential equation of
the deflection curve.
Skip the derivation of the critical load.

Figure 6.9

6.14 Critical loads and effective lengths


Pinned-pinned columns

Le = L

2

` Pcr = r EI
L2

(6.26)

Fixed-free column

Le = 2L

2

` Pcr = r EI
L2

(6.27)

Fixed-fixed column

SOM401M/1

Le = 0.5L

2
Pcr = 4r 2EI 
L

(6.28)

6.15 Columns with eccentric axial loads


So far we have dealt with columns in which the axial loads acted through the centroid of
the cross-section. Under this condition, the columns remain straight until the critical load
is reached, after which bending may occur.
57

Let us now consider a column being compressed by two loads of magnitude P, applied
with a small eccentricity e, measured from the axis as shown in fig. 6.10(a). Each eccentric
axial load is equivalent to a centric load P and a couple of moments Mo = Pe as shown
in fig. 6.10(b).
In order to analyse this pin-ended column, we need to make the same assumptions as in the
previous sections. The bending moment in the column at distance x from the lower end is

M = Mo + P^- vh = Pe - Pv 

(6.29)

Figure 6.10
where v is the deflection of the column. It must be noted that the deflections of the column
are negative if the eccentricity of the load is positive.
The differential equation of the deflection curve is

EIv" = M = Pe - Pv 

(6.30)

v" + k2 v = k2 e 

(6.31)

or much better

as before. The general solution of the above equation becomes


in which k2 = P
EI

v = C1 sin kx + C2 cos kx + e 

The boundary conditions yield the following: C2 = e while C1 is


e^1 - cos kLh

C1 =
=- e tan kL 

sin kL

(6.32)

(6.33)

The deflection of the curve is


v =- e ` tan kL sin kx + cos kx - 1j 


2

(6.34)

If e and P are known, we can use equation 6.33 to calculate the deflection at any point
along the x-axis. The critical load of the column, since it has pinned ends, is

58

2

Pcr = r EI
L2

(6.35)

6.16 Maximum deflection


The maximum deflection will occur at the midpoint of the column shown in fig. 6.11. It
is found to be

d = e `sec kL - 1j 
2

(6.36)

Figure 6.11

Read more about this section on page 847.

6.17 Maximum bending moment


The maximum bending moment in an eccentrically loaded column occurs where the deflection is a maximum

Mmax = P^e + dh 

(6.37)

When the load P is small, the maximum moment is equal to Pe.


Do examples 111 and 113.

6.18 The secant formula for columns

SOM401M/1

Skip this section.

6.19 Elastic and inelastic column behaviour


Skip this section.
59

6.20 Inelastic buckling


Skip this section.

6.21 Exercise
Do the following problems from this chapter:
11.21, 11.31, 11.39, 11.312, 11.315, 11.316, 11.419, 11.53, 11.510

60

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 7
Thick and Compound Cylinders

7.1 Introduction
It is very important to understand how to apply the principles of using equilibrium, compatibility and boundary conditions in thick cylinder theory. In this chapter we are going to start
with two simple beam-bending situations, followed by thick-walled cylinder similar to the
ones used in high-pressure chemical engineering.

7.2 Boundary conditions


The first and most critical step in any stress analysis is the specification of the correct boundary conditions. With the development of modern numerical and computer-aided techniques,
the solution of a properly defined analysis can often be obtained with little or no effort
on the part of the engineer. This is of little value, however, if the wrong problem is solved.
The boundary conditions for any analysis of the stress equilibrium and compatibility equations will be of two kinds:
1
2

Specified displacements. Over some part of the object boundary the displacement
of the material may be known.
Specified surface tractions. Over the remainder of the object boundary the direct
stress acting normal to the boundary, and /or the shear stress acting on the surface,
may be known.

Let us consider an example: A rubber hose is forced over a rigid bar whose outer diameter
is larger than the bore of the hose, as shown in fig. 7.1. The boundary conditions at the
inner diameter of the rubber hose are that every point has been forced radially outward
by an amount equal to the difference in radii. In terms of the radial and circumferential
displacements and respectively, this implies: at r = a,

and

v=0

SOM401M/1

u=d

61

Figure 7.1
Note that the pressure exerted on the hose at r = a by forcing it over the tube must be found
from the analysis. The radial displacement and normal stress cannot both be specified.
At the outer diameter of the hose the resulting displacement is not known; this must be found
from the analysis. What is known, however, is that on that external surface, no normal or
shear stress is acting. This implies that at r = b,

vr = xri
Note that it is not possible to specify as a boundary condition in this problem, since all
the surfaces on which acts are in the interior of the hose.

7.3 Shear stress in a beam

Figure 7.2
Let us consider a simply supported beam carrying a uniformly distributed load as shown
in fig. 7.2, q over the entire length. The origin of the Cartesian coordinates are taken on
the neutral axis, with x positive left to right and y positive downward. This is treated as a
two-dimensional problem with no variation of stress across the width of the beam, and
therefore only two equations of equilibrium are applicable.

2vx + 2xxy = 0 
2x
2y

(7.1)

2vy 2xxy
+
= 0
2y
2x

(7.2)

Making use of the exact solution of pure bending


62

vx =

My

I

(7.3)

then equation 7.2 is not required, neither is any strain-displacement relationship. This is
because, in the derivative of equation 7.3, the geometry of deformation and the stressstrain relationship were included. Substituting for x in equation #.1 gives

2 ` My I j 2xxy
+
=0
2x
2y
Therefore

y
2xxy =- 2M 2y
2x I
But,

dM = V
dx
the shear force on the section, so that

2xxy =-

Vy
2y 
I

(7.4)

Integrating gives

xxy =- V
I

# y dy + C =- Vy
2I

+ C

(7.5)

At the top and bottom free surface of the beam the shear stress must be zero; therefore

xxy = 0

at

y =! d
2

from which
2
C = Vd
8I

and

xxy =-

Vy2 Vd2

+
2I
8I

(7.6)

At the neutral axis y = 0 and the shear stress has its maximum value which is

2
xxy = Vd 
8I

(7.7)

7.4 Transverse normal stress in a beam


Continuing with the above example, we can further analyse the distribution of the direct
stress in the y direction due to the application of a distributed load q. The equilibrium
equation which is applicable is equation 7.2

2vy 2xxy
+
=0
2y
2x
Now, the shear stress, xy, was determined in equation 7.5, and substituting that value in
equation 7.2 gives

SOM401M/1

Vy2 Vd2
2vy
m)
+ 2 c+
2y
2x
2I
8I

But from dV =- q , therefore


dx

2vy qy2 qd2


+
= 0
2y
2I
8I

(7.8)

63

Or

qy2 qd 2
q y2 d2 y
(7.9)
m dy + C =- c m+ C
I 6
2I
8I
8
Using the boundary condition that at the upper surface y =- d , the compressive stress
2
q , where b is the beam width; then
is vy =b
3
3
q q
C =- + c- d + d m
b
I
48 16

vy =- # c

Substituting

1 = d3
b
12I
we find the value of C to be equal to

C =-

qd 3
24I

Therefore

3
q y3 d 2 y
vy =- c + d m
8
24
I 6

(7.10)

7.5 Stress distribution in a pressurised thick-walled cylinder


This problem is of considerable practical importance in pressure vessels and gun barrels. It
is an application of the cylindrical coordinate system, r, , z. A long hollow cylinder which is
subjected to uniformly distributed internal and external pressure is shown in figures. 7.3(a)
and 7.3(b). The two methods of maintaining the pressure inside the cylinder are either by
end caps which are attached to the cylinder as shown in figure #a or by pistons in each
end of the cylinder (figure 7.3(b)). When considering a cross-sectional slide XX as shown
in fig. 7.4, the deformations produced are symmetrical about the longitudinal axis of the
cylinder, and the small element of material in the wall supports the stress system shown.
For axial symmetry r=0 and is constant at any particular radius. Hence, and r are
principal stresses and additionally are quite independent of the method of end closure of
the cylinder. Considering axial stress z and axial strain z, then both of these occur in the
case of end-cap closure. The axial stress, z, is constant and the axial strain fz = dq
dz

Figure 7.3
For closure by pistons, it is evident that z=0 and z occurs only due to the Poissons ration
effect of r and .
64

Figure 7.4
From which the symmetry of the system and for a long cylinder, we come to the conclusion
that plane cross-sections remain plane when subjected to pressure and therefore axial
dq
deformation, w, across the section is independent of r and
=0
dr

7.6 Cylinder with end caps


The equations of equilibrium for an element of material are

dvr + vr - vi = 0 
dr
r

(7.11)

dvz 0 
=
dz

(7.12)

and since z is constant


The strain-displacement equations are


fr = du 
dr

(7.13)

fi = u 
r

(7.14)

fz =

dq

dz

(7.15)

fr = vr - v ^vi + vzh = du 
E
E
dr

(7.16)

fi = vi - v ^vr + vzh = u 
E
E
r

(7.17)

dq

fz = vz - v ^vi + vrh =
E
E
dz

(7.18)

The stress-strain relationships are

Differentiating equation. 7.17 with respect to r yields

SOM401M/1

E du - u = dvi - v dvz - v dvr


r ` dr
rj
dr
dr
dr

Substituting for du

dr

and u

from equations #.16 and #.17 and simplifying,

1 + v ^vr - vih = dvi = v dvz - v dvr 


r
dr
dr
dr

(7.19)
65

Now, since fz is constant,

dfz
= 0 and differentiating 7.18 yields
dr

dvz v dvi dvr 


= c
+
m
dr
dr
dr

dv

z
Substituting into equation 7.19
for from equation 7.19 and
dr
7.11 and simplifying yields

(7.20)

^vr - vih

from equation

^1 - v2hc dvi + dvr m = 0 

dr

dr

(7.21)

dv

z = 0 and therefore v is constant through


From equations 7.21 and 7.20 we see that
z
dr
the wall thickness. Integrating equation 7.21 shows that

^vr - vih = constant, say 2A

(7.22)

Eliminating between equations 7.22 and 7.11 gives


dvi + 2vr - 2A = 0 
dr
r

(7.23)

from which, multiplying by r2,


2Ar - 2rvr - r2 dvr = 0 


dr

And

2Ar - d ^r2 vrh = 0


dr

By integration,

Ar2 - r2 vr = B

Hence

vr = A - B2 
r

(7.24)

vi = A + B2 
r

(7.25)

and from 9.22


where A and B are constants which may be found using the boundary conditions.

7.7 Cylinder with pistons


In this case, z = 0 and there is a condition of plane stress. This solution has been included
to show that we can arrive at the same expressions for r and by deriving a differential
equation for displacement, u. Putting z = 0 in equations 7.16 and 7.17 and solving for r
and in terms of u yields

vr = ` du + v u j E 2 
dr
r ^1 - v h

(7.26)

vi = `v du + u j E 2 
dr
r ^1 - v h

(7.27)

From equation 7.26,



66

dvi = d2 u + v du - v u
E 
c 2
m
dr
dr
r ^1 - v2h
dr

(7.28)

Substituting equations 7.26, 7.27 and 7.28 into equation 7.11 and simplifying, yields

d 2 u + 1 du - u = 0 
r dr
dr 2
r2

(7.29)

This differential equation expresses radial equilibrium in terms of the displacement u, in


the cylinder wall. The general solution of this equation is

u = Cr + C' 
r

(7.30)

in equations 7.26 and 7.27,


Substituting for ur and du
dr

vr = cC^1 + vh - C2' ^1 - vhm E 2 


^1 - v h
r

(7.31)

vi = cC^1 + vh - C2' ^1 - vhm E 2 


^1 - v h
r

(7.32)

where C and C are constants. These equations may be rewritten with differential constants as

vr = A - B2
r


And

vi = A + B2
r

which are the same constants as equations 7.24 and 7.25.


Boundary conditions:
The next stage is the determination of the constants A and B.

7.8 Internal and external pressure


The boundary conditions of this problem are as follows:
At r = ri = a

r = Pi

At r = ro = b

r = Po

The pressure is negative because at the inside the internal pressure tends to increase the
internal diameter while at the outside; the external pressure tends to decrease the external
diameter.

- Pi = A - rB2

and

- Po = A - rB2
o

from which, eliminating A, we get

B=

^ Pi - Poh r 2i r 2o

r -r
2
o

2
i

and

2
2
A = Pi r 2i - Po2 r o
ro - r i

SOM401M/1

Therefore the radial and hoop stresses become

2
2
^ Pi - Poh r 2i r 2o

vr = Pi r 2i - Po2 r o - 2 2
ro - r i
r ^r o - r 2i h

(9.33a)

2
2
^ Pi - Poh r 2i r 2o

vi = Pi r 2i - Po2 r o + 2 2
ro - r i
r ^r o - r 2i h

(9.33b)

These equations were first derived by Lame and Clapeyron in 1833. If we now let the
radius ratio ro = k , then equations 9.33a and 9.33b may be written as

ri

67

vr =

r2
r2
2
1
; Pi `1 - 2o j - Po k c1 - 2i mE
k -1
r
r

(9.34a)

vi =

r2
r2
2
1
; Pi `1 + 2o j - Po k c1 + 2i mE
k -1
r
r

(9.34b)

It is important to note that the stresses depend on the k-ratio rather than on the absolute
dimensions.

7.9 Internal pressure only


An important special case of the above occurs when the external pressure is atmospheric
only and can be neglected in relation to the internal pressure. Then with Po=0

vo =

vr =

r2
1
; Pi c1 - 02 mE
k -1
r
2

r2
1
; Pi c1 + 02 mE
k -1
r
2

(9.35)
(9.36)

At the inner surface, r and each have their maximum magnitude so that at r = ri,
r = Pi
It is appropriate at this point to note that the radial stress shown on the element in fig 7.4
is in the positive sense, ie tension is in fact in the opposite sense, ie compression. The
circumferential or hoop stress at r = ri = a is
2
2
2
vi = ro2 + r i2 Pi = k2 + 1 Pi
ro - r i
k -1

At the outer surface, where r = ro,

vr = 0 and vi =

2Pi
k -1
2

7.10 Stress distribution for and r


To complete the basic analysis of the elastically deformed thick-walled pressure vessel, the
variation of the two principal stresses and r is shown plotted through the wall thickness
in fig. 7.5 for internal pressure and a k-ratio of 3.

Figure 7.5

68

7.11 Axial stress and strain


Now that expressions have been developed for the radial and circumferential stresses within
the cylinder, the next step is to consider what conditions of stress and strain can exist axially
along the cylinder. These will depend on the boundary conditions at the ends of the cylinder.

7.12 Cylinder with end caps but free to change in length


In this case there must be equilibrium between the force exerted on the end cover by the
internal pressure and the force of the axial stress integrated across the wall of the vessel.
Therefore, from fig. 7.3(a)

v ^rr2o - rr2i h - Pi rr2i = 0


Z

so that

2
vz = 2Pi r i 2 = 2 Pi 
ro - ri
k -1

(7.37)

And

^1 - 2vh Pi
fz = vz - v ^vr = vih =
E
E
E^ k2 - 1h

7.13 Pressure retained by piston in each end of cylinder


Since there is no connection between the piston and the cylinder, the axial force due to
pressure is reacted by the pistons, and therefore there can be no axial stress in the wall
of the cylinder.
Thus

vz = 0 

(7.38)

And

fz =- v ^vr + vih =- 22vPi


E
E^ k - 1h

SOM401M/1

7.14 Methods of containing high pressure

We can observe from fig. 7.5 that there is a marked variation in the stress in the wall of
a thick cylinder subjected to internal pressure, and this situation gets worse when designing for even higher pressures. In order to secure a more uniform stress distribution, one
method is to build up the cylinder by shrinking one tube on the outside of another. In this
case, the inner tube is subjected to hoop compression by shrink fit of the tension. When
the compound is subjected to working pressure, the resultant stresses are the algebraic
sum of that due to the shrinking and that due to the internal pressure. The resultant tensile
stress at the inner surface of the inner tube is not so large as if the cylinder were composed
of one thick tube. The final tensile stress at the inner surface of the outer tube is larger than
if the cylinder consisted of one thick tube.
Thus a more even stress distribution is obtained.

69

The second technique used in special industrial pressure vessels, is to wind around the
outside of a tube a high-tensile strength ribbon of a rectangular section with sufficient
tension to bring the tube into a state of hoop compression. Subsequent internal pressure
can be set up in the tube.
The other method for creating hoop compression at the bore of a cylinder is known as
auto-frettage.
This consists of applying internal pressure to a single cylinder until yielding and a prescribed
amount of plastic deformation occurs at the bore. Since strain increases along the radius
of the cylinder, on the release of pressure, the elastic recovery of the material causes the
bore of the cylinder to be subjected to compressive hoop stress. At the same time tensile
hoop stress occurs in the outer material.

7.15 Stress set up by a shrink-fit assembly


A shrink fit between components is a very important and secure method of assembly. It consists, in the case of two cylindrical objects, of the inner diameter of the outer cylinder being
slightly (by a fraction of a millimetre) larger than the outer diameter of the inner cylinder.
Consequently, when they are at the same temperature the outer cannot be passed over
the inner. However, if the outer cylinder is heated and the inner is cooled then the thermal
expansion and contraction can be made sufficient to allow one cylinder to pass over the
other. On returning each to room temperature there is interference at the mating surface
since they cannot regain their original dimensions at the interface. The two components
are locked firmly together and a system of radial and circumferential stresses is set up at
the interface and through the wall of each cylinder. For elastic conditions the principle of
superposition can be used to add the stresses due to shrink-fit interference to those due
to internal pressure.
For two cylindrical components we require equations 9.24 and 9.25 for each component
given by

vr = A - B2
r

and

vi = A + B2
r

^inner component h 

(7.39a)

vr = C - D2
r

and

vi = C + D2
r

^outer component h 

(7.39b)

where the constants A, B, C and D are determined from the boundary conditions. For
shrink-fit stresses only, the boundary conditions are that r = 0 at the inside of the inner
cylinder and outside of the outer cylinder, and at the mating surface, rm the radial stress
in each vessel must be the same; therefore

A - B2 = C - D2 
rm
rm

(7.40)

Finally, at the mating surface the radial interference is the sum of the displacement of the
inner cylinder inwards, u and the outer cylinder outwards, u; thus

d =- u' + u" = rm ^fi" - fi' h 

Where
70

f = 1 ^vi - vvrh
E

(7.41)

7.16 Stress distribution in a pressurised compound cylinder


The containment of high internal pressure in chemical processes can be achieved more
effectively by shrinking two or more cylinders, one over the other, to give a compound
or multivessel. The analysis simply uses the basic thick cylinder equations for r and
together with the shrink-fit and other boundary conditions.

7.17 Using a spreadsheet to determine the stresses in


a compound cylinder

SOM401M/1

One can write a spreadsheet program to do an analysis but it is important that you understand the theory of thick and compound cylinders before attempting to write excel
spreadsheets.

71

Strength
of Materials
Strength
of MaterialsIV
IV

Strength of Materials IV

Chapter 8
Theories of Elastic Failure

8.1 Introduction
When a ductile metal is subjected to simple axial loading, it is found that beyond a certain
point, stress is no longer proportional to strain, which results in there being a permanent
deformation when the stress is removed as shown in fig. 8.1 below. The material is then
said to have yielded. Knowing the stress at which yielding behaviour commenced, it would
then be a simple matter to design a component from the same material to withstand a
particular axial load without yielding occurring. This example is simple as there is only one
principal stress to consider.

Figure 8.1

SOM401M/1

The problem when designing a pressure vessel, rotating disc, or some component containing a complex principal stress system so that the material remains elastic, ie no yielding,
when under full load is rather more complex. One could adopt a trial and error method
of building a component and testing it to find out when the deformations are no longer
recoverable, but this would obviously be very uneconomical. It is therefore essential to
find some criterion based on stresses or strains or perhaps conditions mentioned above.
If a theoretical criterion can be established which predicts complex material behaviour,
it is then only necessary to establish experimentally the yield point in a simple tension
or compression test.

8.2 Yield criteria and fracture criteria


Yield criteria are the criteria that apply to ductile materials while the fracture criteria apply
to brittle material.

73

A number of theoretical criteria for yielding have been proposed each seeking to obtain
adequate correlation between estimated component life and that actually achieved under
service load conditions for both brittle and ductile material applications. The five main
theories are:
1
2
3
4
5

Maximum principal stress theory (Rankine)


Maximum shear stress theory (Guest-Tresca)
Maximum principal strain (Saint-Venant)
Total strain energy per unit volume (Haigh)
Shear strain energy per unit volume (MaxwellHubervon Mises)

The stress at yield point in a simple tension test is y, 1 2 and 3 and are the principal
stresses in the three-dimensional complex stress system.

8.3 Maximum principal stress theory


This theory assumes that when the maximum principal stress in the complex stress system
reaches the elastic limit stress in simple tension, failure occurs. The criterion of failure is thus

v1 = vy 

(8.1)

It should be noted, however, that failure also occurs in compression if the least principal
stress v3 were compressive and its value reaches the value of the yield stress in compression for the material concerned before the value of vyt is reached in tension. An additional
criterion is therefore

v3 = vy

(compressive) 

(8.2)

We can also write the above as follows:


v1 = vut
n

if v1 2 0 

(8.3)

v3 = vuc
n

if v3 1 0 

(8.4)

And

This theory should not be used for shafts since invariably they are made from a ductile
material and will incorrectly indicate a factor of safety larger than that indicated by two
theories of elastic failure.
While the theory can be shown to hold fairly well for brittle materials, there is considerable
experimental evidence that the theory should not be applied to ductile materials. For example,
even in the case of the pure tension test itself, failure for ductile materials takes place not
because of the direct stresses applied but in shear on planes at 45 to the specimen axis.

8.4 Maximum shear stress theory


Theories of yielding are generally expressed in terms of principal stresses, since these completely determine a general state of stress. The element of material shown in fig. 8.2 is subjected
to three principal stresses and it will be taken that v1 2 v2 2 v3 .
74

This theory was proposed by a French engineer Tresca. He made the assumption that yielding is dependent on the maximum shear stress in the material reaching a critical value. This
is taken as the maximum shear stress at yielding in a uniaxial tensile test. The maximum
shear stress in the complex stress system will depend on the relative values and signs of
the three principal stresses, always being half the difference between the maximum and
minimum. It should be remembered that the minimum stress can be zero or compressive,
in which case it is negative in value.
For a general three-dimensional stress system, or in the two-dimensional case with one
of the stresses tensile, one compressive and the third zero, the maximum shear stress is

xmax = v1 - v2 
2

(8.5)

Under uniaxial tension there is only one principal stress, v1 ^v2 = v3 = oh , so that the
maximum shear stress is
xmax = v1 

(8.6)

and at yield this becomes


xy =

vy

2

(8.7)

Therefore the Tresca criterion states that yielding will occur when the value of 3 being
algebraically the smallest value, ie taking account of sign and the fact that one stress may
be zero. This produces a fairly accurate correlation with experimental results particularly for
ductile materials, and is often used for ductile materials in machine design. The criterion
is often referred to as the Tresca theory and is one of the widely used laws of plasticity.

v1 - v3 = vy 

(8.8)

Figure 8.2

8.5 Maximum principal strain theory


This theory assumes that failure occurs when the maximum strain in the complex stress
equals that at the yield point in the tensile test,

1 ^v1 - vv2 - vv3h = vy 


E
E

(8.9)

SOM401M/1

ie

v1 - vv2 - vv3 = vy 

(8.10)

This theory is contradicted by the results obtained from tests on flat plates subjected to
two mutually perpendicular tensions. The Poissons ratio effect of each tension reduces the
strain in the perpendicular direction so that according to this theory failure should occur at
75

a higher load. This is not always the case. The theory holds reasonably well for cast iron
but is not generally used in design procedures these days.

8.6 Maximum total strain energy per unit volume theory


This theory assumes that failure occurs when the total strain energy in the complex stress
system is equal to that at the yield point in the tensile test.
The criterion of failure is thus,

1 ^v12 + v22 + v23 - ^2vv1 v2 + v2 v3 + v3 v1hh = v 2y 


E
2E

(8.11)

v12 + v22 + v23 - ^2vv1 v2 + v2 v3 + v3 v1h = v 2y 

(8.12)

ie

The theory gives fairly good results for ductile materials but is seldom used. The theory by
Guest and Tresca below is generally preferred.

8.7 Maximum shear strain energy per unit volume theorydistortion energy
This theory states that failure occurs when the maximum shear strain energy component in
the complex stress system is equal to that at the yield point in the tensile test,

^v1 - v2h2 + ^v2 - v3h2 + ^v3 - v1h2 = 2v2y 

(8.13)

This theory has received considerable verification in practice and is widely regarded as the
most reliable basis for design, particularly when dealing with ductile materials. It is often
referred to as the von Mises criterion and is probably the best theory of the five. It is also
sometimes referred to as the distortion energy or maximum octahedral shear stress theory.

8.8 Mohr fracture criterion or Mohrs modified theory


Some materials, such as iron, have much greater strength in compression than in tension.
Mohr proposed that, in the first and third quadrants of Failure locus, a maximum principal
stress theory was appropriate based on the ultimate strength of the material in tension
or compression. In the second and fourth quadrants where the two principal stresses are
of opposite sign the maximum shear-stress theory should apply. The results can be seen
in fig. 8.3 below.

76

2-tension
ut

uo

ut
1-compression

1-tension

0
uc

2-compression

Figure 8.3

8.9 Mohr fracture criterion or Mohrs modified theory

Examples will be posted on myUnisa in lecture no 5 or 6. Other notes have already been
posted on the website, so make sure that you check it: www.unisa.ac.za/myunisa/som401M/
additional~resources.htm

8.10 Exercise

SOM401M/1

1 A shaft subjected to pure tension yields at a torque of 1.2kNm. A similar shaft is


subjected to a torque of 720Nm and a bending moment M. Determine the maximum
allowable value of M according to

a) the maximum shear stress theory

b) the shear strain energy theory

2 A circular steel cylinder of wall thickness 10 mm and internal diameter 200 mm is


subjected to a constant internal pressure of 15MPa. Determine how much

c)

axial tensile load, and

d) axial compressive load


77

can be applied to the cylinder before yielding commences according to the maximum
shear stress theory.
The yield stress of the material in simple tension is 240MPa. Assume that the radial
stress in the wall of the cylinder is zero. Sketch the plane yield stress locus for the
maximum shear stress theory, and show the two points representing the cases above.
3 A thin-walled cylindrical tank has a diameter of 2 m. If it is subjected to an internal
pressure of 1MPa, calculate the required wall thickness for the tank if yielding is not
to occur. The von Mises theory should be used to check for yielding. The yield stress
in simple tension for the material is 280MPa and a factor of safety of 1.5 should
be used.
4 A hollow tube with an external diameter of 100 mm and a wall thickness of 5 mm
is subjected to an axial force of 100kN. If the tensile yield stress for the material
is 280MPa, use the Tresca criterion to establish whether or not a torque of 6kNm
could be applied without causing the tube to yield. A safety factor of 2 should be
used to allow for stress concentrations.
5 A cast-iron cylinder of 60 mm internal and 5 mm wall thickness is to be used to
check the Mohr theory of failure. The tensile and compressive strengths of the material have been measured as 400MPa and 1200MPa respectively. Determine the
following:

a) The internal pressure to cause failure

b) The axial compressive load to cause failure when combining with an internal
pressure of 50MPa

6 A shaft must transmit a torque of 1.2kNm and a bending moment of 1.8kNm.


Calculate the required shaft diameter using both failures for ductile materials. Use
a factor of safety of 2 and yield strength of 280MPa.
7 A solid shaft 75 mm in diameter is subjected to an axial tensile force of 80kN and
a torque of 3.8kNm. If the yield strength of the shaft material is 278MPa, calculate
the safety factors that have been used in the design according to Tresca and von
Mises yield criteria.
8. An elevator pulley of diameter 1 m and negligible mass is mounted on a shaft
midway between bearings which are 800 mm apart. Under the worst conditions
the rope tensions are 25kN on the cage side and 10kN on the counterpoise side
of the pulleys. The yield strength of the shaft material is 250MPa. Use a factor of
safety of 5 and calculate the required shaft diameter using the Tresca and von Mises
criteria.
9 A thin cylinder with a diameter of 0.5 m and a wall thickness of 10 mm is subjected
to an internal pressure of 2MPa and a torque of 30kNm. Calculate the factor of
safety using both criteria for elastic failure. The yield strength is 280MPa.
10 A thick cylinder has inner and outer diameters of 250 mm and 300 mm respectively.
Include the longitudinal stress, and using the Tresca and von Mises criteria, calculate
the internal pressure that will initiate yielding. The yield strength is 300MPa.

78

11 A cast iron cantilever has a length of 0.5 m and is subjected to a point load of
50kN and a moment of 20kNm at the free end. The moment deflects the beam in
the same direction as the point load. The component has a T-cross section with an
overall height of 250 mm and both the web and flange are 50 mm thick. Calculate
the factor of safety of the component using both Rankine and modified Mohr criteria.
vut = 290MPa, and vuc = 950MPa .
12 A thin cylinder 1 m diameter and 3 m long is filled with a liquid to a pressure of
2MPa. Assuming a yield stress for the material of 240MPa in simple tension and
a safety factor of 4, determine the necessary wall thickness of the cylinder, taking
the maximum shear strain energy and the maximum shear stress as the criteria of
failure. E = 207GPa and v = 0.286 .

SOM401M/1

13 Cast iron used in the manufacture of an engineering component has tensile


and compressive strengths of 400MPa and 1200MPa respectively. If the maximum
value of the tensile principal stress is to be limited to one-quarter of the tensile
strength, determine the maximum value and nature of the other principal stress
using Mohrs modified yield theory for brittle materials.

79

Das könnte Ihnen auch gefallen