Anda di halaman 1dari 125

ME1 Stress Analysis

Lecture notes 2014-15

Daniele Dini
2014 Imperial College London
What is Stress Analysis?
Stress Analysis is the short version of Stress-Strain Analysis, the engineering dis-
cipline related to the theory and the methods developed by physicists, engi-
neers and mathematicians to determine the stresses and strains in materials
and structures subjected to forces or loads. This is one of the fundamental
subjects for civil, mechanical and aeronautical engineers and enables quanti-
tative analysis of the stresses and deformations caused by external loads ap-
plied to a body. In this context, stresses are local measures of forces and are
defined as forces per unit area, while strains are local measures of deforma-
tions and usually defined as linear deformations per unit length.
Given an engineering component or a structure, the properties of the materi-
als they are made of, the links between the parts of the structure, and the typi-
cal forces applied to them, the laws and methods taught in the Stress Analysis
course will enable you to obtain a quantitative description of the stress over all
those parts and joints, and the deformation caused by those stresses.
There is a strong link between stresses and strains as they cannot exist inde-
pendently from each other: stress causes strain and vice versa. The main differ-
ence between Stress Analysis and Mechanics (see ME1 Mechanics) is that in
the Mechanics syllabus deformations are neglected and all parts and struc-
tures are considered as rigid bodies. This is an idealisation which enables to
study the kinematics and dynamics of objects without having to consider de-
formations.

i
Contents
Chapter 1 Static equilibrium . . . . . . . . . . . . . . . . . 1
1.1 Dynamics and statics . . . . . . . . . . . . . . . . . . 1
1.1.1 Dynamics of a particle . . . . . . . . . . . . . . . 1
1.2 Static equilibrium . . . . . . . . . . . . . . . . . . . 2
1.2.1 Example: See-saw in equilibrium . . . . . . . . . . . . 2
1.3 Free body diagram . . . . . . . . . . . . . . . . . . 3
1.3.1 Analysis of free body diagrams . . . . . . . . . . . . . 3
1.3.2 Example: FBD analysis of bicycle . . . . . . . . . . . . 4
1.3.3 Example: FBD analysis of block on wedge . . . . . . . . . 5
1.4 Notes. . . . . . . . . . . . . . . . . . . . . . . 7

Chapter 2 Loaded frames . . . . . . . . . . . . . . . . . 10


2.1 Plane frame supports. . . . . . . . . . . . . . . . . .10
2.2 Example: plane frames . . . . . . . . . . . . . . . . .11
2.3 Pin-jointed frames. . . . . . . . . . . . . . . . . . .13
2.3.1 Analysis of a two-bar frame . . . . . . . . . . . . . .14
2.4 Statical determinacy . . . . . . . . . . . . . . . . . .15
2.4.1 Evaluating statical determinacy . . . . . . . . . . . .16
2.4.2 Example: Statical determinacy of 6-bar frame . . . . . . . .17
2.5 Example: Five-bar frame . . . . . . . . . . . . . . . . .17
2.6 Example: Generalised, statically determinate frame . . . . . . . .19
2.7 Analysing frames by the method of sections . . . . . . . . . .21
2.8 Maximum allowable load on a frame . . . . . . . . . . . . .22
2.8.1 Example: maximum allowable load on a 2-bar frame . . . . . .23
2.8.2 Example: maximum allowable load on a complex frame . . . . .24
2.9 Notes. . . . . . . . . . . . . . . . . . . . . . .26

Chapter 3 Stress . . . . . . . . . . . . . . . . . . . . 29
3.1 Stress and strength . . . . . . . . . . . . . . . . . .29
3.2 Stresses in statically determinate systems . . . . . . . . . . .30
3.2.1 Thin, pressurised cylinders: hoop stress . . . . . . . . . .30
3.2.2 Thin, pressurised cylinders: axial stress . . . . . . . . . .32
3.2.3 Thin, pressurised spherical shell: stress . . . . . . . . . .33
3.2.4 Example: pressure vessel . . . . . . . . . . . . . .34
3.2.5 Example: pressure vessel bolts . . . . . . . . . . . . .34
3.2.6 Thin rotating ring . . . . . . . . . . . . . . . . .35
3.3 Notes. . . . . . . . . . . . . . . . . . . . . . .36

Chapter 4 Elastic strain . . . . . . . . . . . . . . . . . . 39


4.1 Direct strain. . . . . . . . . . . . . . . . . . . . .39
4.2 Elastic stress-strain relationships . . . . . . . . . . . . . .40

Contents ii
4.3 Poisson’s ratio . . . . . . . . . . . . . . . . . . . .40
4.3.1 Typical Poisson’s ratio values . . . . . . . . . . . . .41
4.4 Stress and strain in three dimensions . . . . . . . . . . . . .41
4.4.1 Hydrostatic stress and volumetric strain . . . . . . . . . .43
4.4.2 Relating elastic modulus, Poisson’s ratio and bulk modulus . . . .43
4.4.3 Example: volume change due to stress and temperature . . . .44
4.5 Strains in cylindrical and spherical shells. . . . . . . . . . . .45
4.6 Apparent modulus of elasticity. . . . . . . . . . . . . . .45
4.7 Volumetric strain within shells . . . . . . . . . . . . . . .46
4.7.1 Volumetric strain within a cylindrical shell . . . . . . . . .46
4.7.2 Volumetric strain within a spherical shell . . . . . . . . .47
4.7.3 Stress-strain relationships in polar coordinates . . . . . . . .47
4.8 Elastic strain energy . . . . . . . . . . . . . . . . . .48
4.9 Solving statically indeterminate stress-strain problems . . . . . . .49
4.9.1 Example: statically determinate problem . . . . . . . . .49
4.9.2 Example: statically indeterminate problem . . . . . . . . .50
4.9.3 Example: hydraulic cylinder with tie rods . . . . . . . . .51
4.9.4 Example: pressure vessel with hemispherical ends . . . . . .54
4.10 Notes . . . . . . . . . . . . . . . . . . . . . .54

Chapter 5 Bending of beams . . . . . . . . . . . . . . . . 57


5.1 Beam supports. . . . . . . . . . . . . . . . . . . .57
5.2 Modelling of loaded beams. . . . . . . . . . . . . . . .58
5.3 Shear force and bending moment. . . . . . . . . . . . . .59
5.4 Example: Shear force and bending moment . . . . . . . . . .60
5.4.1 Example: SF and BM diagrams (1) . . . . . . . . . . . .62
5.4.2 Example: SF and BM diagrams (2) . . . . . . . . . . . .64
5.4.3 Example: SF and BM diagrams (3) . . . . . . . . . . . .65
5.5 Moment applied to a beam at a point . . . . . . . . . . . .66
5.6 Relating shear force and bending moment to applied loads . . . . .68
5.6.1 Example: Problem 5.1 . . . . . . . . . . . . . . .68
5.7 Notes. . . . . . . . . . . . . . . . . . . . . . .69

Chapter 6 Bending of beams: stress and strain . . . . . . . . . 72


6.1 Bending stress and strain . . . . . . . . . . . . . . . .72
6.1.1 Stress in beams of symmetrical cross-section . . . . . . . .74
6.1.1.1 Example: bending stress in a rectangular beam . . . . .75
6.1.2 Stress in beams of arbitrary cross-section . . . . . . . . .76
6.1.2.1 Example: stress in beams of arbitrary cross-section . . . .77
6.2 Flexural rigidity of a beam . . . . . . . . . . . . . . . .77
6.2.1 Second moment of area . . . . . . . . . . . . . . .78
6.2.1.1 Composite areas . . . . . . . . . . . . . . .79
6.2.1.2 The parallel axis theorem . . . . . . . . . . . .80
6.2.1.3 Example: Second moment of a non-symmetric area . . . .82

Contents iii
6.3 Example: Problem 5.3 . . . . . . . . . . . . . . . . .83
6.4 Example: stress in an I-beam . . . . . . . . . . . . . . .85
6.5 Bars under combined axial and bending loads . . . . . . . . .86
6.5.1 Example: combined axial and bending loads . . . . . . . .87
6.6 Notes. . . . . . . . . . . . . . . . . . . . . . .88

Chapter 7 Bending of beams: deflections . . . . . . . . . . . 91


7.1 Bending deflections . . . . . . . . . . . . . . . . . .91
7.1.1 Example: Bending deflections (1) . . . . . . . . . . . .92
7.1.2 Example: Bending deflections (2) . . . . . . . . . . . .93
7.1.3 Example: Bending deflections (3) . . . . . . . . . . . .94
7.1.4 Example: Bending deflections (4) . . . . . . . . . . . .95
7.2 Macaulay’s method . . . . . . . . . . . . . . . . . .97
7.2.1 Example: Macaulay’s method (1) . . . . . . . . . . . .98
7.2.2 Macaulay’s method for distributed load . . . . . . . . . .99
7.2.3 Distributed loads over a restricted length . . . . . . . . .99
7.3 Beam bending solutions by superposition . . . . . . . . . . 101
7.3.1 Example: end deflection of a cantilever . . . . . . . . . 101
7.3.2 Example: end loaded cantilever with uniformly distributed load . 101
7.3.3 Example: end-supported cantilever . . . . . . . . . . 102
7.4 Statically indeterminate beams . . . . . . . . . . . . . 103
7.5 Notes. . . . . . . . . . . . . . . . . . . . . . 103

Chapter 8 Torsion of bars . . . . . . . . . . . . . . . . . 106


8.1 What is torsion? . . . . . . . . . . . . . . . . . . 106
8.2 Analysis of torsional stress and strain . . . . . . . . . . . . 108
8.2.1 Torsion of a thin-walled cylinder . . . . . . . . . . . 108
8.2.2 Solid and hollow circular shafts . . . . . . . . . . . 109
8.2.3 Example: Torsion bar design (1) . . . . . . . . . . . 111
8.2.4 Example: Torsion bar design (2) . . . . . . . . . . . 112
8.3 Shafts of varying diameter . . . . . . . . . . . . . . . 113
8.3.1 Example: torsion of a stepped shaft . . . . . . . . . . 114
8.4 Bars of non-circular cross-section . . . . . . . . . . . . . 115
8.4.1 Example: torsion of a non-circular bar . . . . . . . . . 115
8.5 Analysis: torsion of thin-walled, non-circular tubes . . . . . . . 117
8.6 Notes. . . . . . . . . . . . . . . . . . . . . . 118

Contents iv
Chapter 1 Static equilibrium
1
Stress analysis begins by dealing with structures and components which are in
static equilibrium — neither moving nor accelerating rapidly enough for inertial

Static equilibrium
forces to significantly load them.
On successfully completing this section, you will be able to:
• Define statics and dynamics
• Recall Newton’s laws of motion
• Define forces and moments
• Define static equilibrium
• Sketch free body diagrams for very simple systems

1.1 Dynamics and statics


Classical Mechanics is historically divided in three branches:
1. Statics — the science which studies equilibrium of bodies and its relation
to forces;
2. Kinetics — the science which studies motion of bodies and its relation to
forces;
3. Kinematics — the science which deals with the implications of observed
motions of bodies without regard for circumstances causing them.
These three topics have been linked to dynamics in different ways. Earlier ap-
proaches to this subjects combined statics and kinetics under the name, Dy-
namics, which became the branch dealing with determination of the motion
of bodies resulting from the action of specified forces. A more recent approach
— more common in engineering books on mechanics and still in widespread
use among mechanicians — separated Statics, and combined kinetics and kin-
ematics under Dynamics.
In the ME1 Solid Mechanics syllabus we follow the latter approach; however,
while ME1 Mechanics will deal with both Statics and Dynamics but will not
treat deformation of bodies under the action of external forces, ME1 Stress
Analysis will only deal with Statics and will concentrate on the study of the
effect that forces have in terms of deformations, stresses and strains.
Links will also be made between stresses and material characteristics to assess
the potential failure of structures.

1.1.1 Dynamics of a particle


Newton’s laws of motion were originally formulated for a particle, but they apply
in a general sense to all bodies under certain conditions.

Chapter 1 Static equilibrium 1


First law
A particle at rest tends to stay at rest and a particle in motion tends to
stay in motion with the same speed and in the same direction unless ac- 1
ted upon by an unbalanced (net) force.

Static equilibrium
Second law
The acceleration of a particle as produced by a net force is directly pro-
portional to the magnitude of the net force in the same direction as the
net force, and inversely proportional to the mass of the particle:
F = mẍ.

Third law
For every action, there is an equal and opposite reaction.

1.2 Static equilibrium


In a static problem, there is no translational acceleration and no rotational mo-
tion.
For a body of mass m to which several forces Fi and moments Mi are applied:

∑i Fi = ∑ mẍ = 0

and
∑jMj = 0

i.e. both the net force and the net moment in every direction is zero. Such a
system is in static equilibrium.

1.2.1 Example: See-saw in equilibrium


The frictionless see-saw illustrated below is balanced with no linear or angular
acceleration — i.e. in static equilibrium — in the position shown. Find F and N,
neglecting the see-saw’s own mass.
1200 N

1m 2m

1. Draw a free body diagram.


2. Apply the condition that the net vertical force is zero:
∑i Fi = 0 = − F + N − 1200

1.2 Static equilibrium 2


3. Apply the condition that the net moment is zero in the clockwise direc-
tion:
∑i Mi = 0 = − 1 × F + 2 × 1200
1

4. Solve for F and N: F = 2400 N, N = 3600 N.

Static equilibrium
1.3 Free body diagram
To analyse any complex engineering system, it is first broken down into the inter-
acting components it consists of, while noting the points at which each is connec-
ted to others. The whole system can then be predicted from a knowledge of how
each component behaves in isolation.
To analyse mechanisms and structures, they are disassembled into separate
free bodies. All of the points at which each free body contacted those sur-
rounding it are then replaced by equivalent forces and moments, and the free
body is analysed using static equilibrium equations.

1.3.1 Analysis of free body diagrams


If a mechanism or structure is in equilibrium, it can first be represented as a free
body diagram by removing it from outside influences and replacing them by the
forces they exert.

1. Construct a free-body diagram (FBD) of the system, removing any exter-


nal components which interact with it — including constraints such as
supports or enclosing boundaries.
2. Mark onto the diagram all known external forces and moments which act
on the free body.
3. Mark onto the diagram, as variables, any unknown reaction forces and
moments (e.g. from the ground, a wall) which act on it.

Remember
Newton’s Third Law: When one body exerts a force on a second body, the second
body simultaneously exerts a force equal in magnitude and opposite in direction on
the first body.

4. Apply the static equilibrium condition of forces and moments equations


to form:
• For a 2D system, form no more than three equations
• For a 3D system, form no more than six equations.
5. Solve, as far as possible:
• If there are as many equilibrium equations as unknowns, solve com-
pletely

1.3 Free body diagram 3


• If there are more unknowns than equations, separate one section of
the free body as a new free body and return to Step 2.
1
1.3.2 Example: FBD analysis of bicycle

Static equilibrium
Analyse the forces acting on the frame of a bicycle of mass MB loaded by a
rider of mass MR.

Before attempting to analyse stresses using a static equilibrium free body dia-
gram, we first need to be sure that the bike is in static equilibrium — i.e. it is
stationary or being ridden at constant speed, not accelerating or hitting a
bump.
1. Construct a free-body diagram (FBD) of the bicycle:

2. Mark the known forces acting on the bike frame: W 1 = M1g, W B = MBg

3. Add to the diagram variable reaction forces R1, R2 from the ground

4. Apply the static equilibrium condition of forces and moments equations


to form:

1.3.2 Free body diagram 4


a. A horizontal force equilibrium equation ∑i FH = 0

b. A vertical force equilibrium equation ∑i FV = 0 1


c. A moment equilibrium equation about some chosen point.

Static equilibrium
5. In this case there are more unknowns than equations and we continue by
separating one section of the system as a new free body and returning to
Step 2.
Using Newton’s 3rd Law to separate subcomponents

1.3.3 Example: FBD analysis of block on wedge


A block of mass m1 is stationary on a wedge of angle θ and mass m2 (or slides
on it at constant speed). The wedge slides at constant or zero speed along a
horizontal surface. Determine the forces acting along, and normal to each sur-
face.

It is not always necessary to begin with a free body diagram of the whole sys-
tem.
Because both masses move at constant or zero velocity, there are no accelera-
tions and the system is in static equilibrium.
1. Begin by drawing each component — block and wedge — as a separate
free body:

1.3.3 Free body diagram 5


1

Static equilibrium
2. Mark the known forces m1g, m2g acting on each component.

3. Add to the diagram the variable, equal and opposite pairs of action reac-
tion forces (normal R1 and R2, tangential F1 and F2) between the block
and wedge surfaces.

Note
It doesn’t matter which direction is chosen for any action force arrow: if the action
and reaction arrows were both reversed, the resulting value of each force would have
the opposite sign.

4. Apply the static equilibrium condition of forces and moments equations


to form:
a. A horizontal force equilibrium equation for the block:
−F2cosθ + R2sinθ = 0

b. A vertical force equilibrium equation for the block:


−F2sinθ − R2cosθ − m2g = 0

Hence R2 = − m2gcosθ and F2 = − m2gsinθ.

c. A horizontal force equilibrium equation for the wedge:


F2cosθ − R2sinθ + F1 = 0

d. A vertical force equilibrium equation for the wedge:


F2sinθ + R2cosθ − m1g = 0

Hence R1 = m1 + m2 g and F1 = 0.

5. In this case we have fewer unknowns than possible equations: no mo-


ment balance was required.
As a check on the result, a free-body diagram for the entire system could be
drawn, as is usually done first:

1.3 Free body diagram 6


1
θ

Static equilibrium
Hence for vertical force equilibrium: R1 = m1 + m2 g

and for horizontal force equilibrium: F1 = 0

as before.

1.4 Notes

1.4 Notes 7
Static equilibrium
1

8
1.4 Notes
Static equilibrium
1

9
1.4 Notes
Chapter 2 Loaded frames
Pin jointed frames consist of straight bars loaded in tension or compression, as-
sembled with joints which are (or might become) free to rotate. Examples include 2
bridges, cranes and roof support structures for warehouses etc..

Loaded frames
B C D

A G F E

On successfully completing this module, you will be able to:


• Assess the static equilibrium of plane frames
• Recognise different types of frames
• List possible supports and reaction forces
• Build free body diagrams of simple pin-jointed frames
• Discuss static determinacy of pin-jointed frames
• Determine forces in individual bars of pin-jointed frames
• Calculate stresses in individual bars and assess the likelihood of failure
under the prescribed load
• Assess potential failure of the bars by buckling

2.1 Plane frame supports


The forces which act on bars throughout a plane frame may depend on the exact
nature of each of its supports.
A built-in or encastré support can support a vertical force without movement,
and a moment without rotation.

Chapter 2 Loaded frames 10


V V 2 V
H H
M

Loaded frames
A pin joint can support a horizontal or vertical force without movement but
cannot resist rotation.

V V V
H H
M or sliding joint can support only a vertical load without move-
A roller joint
ment.

V V V
H H
M
2.2 Example: plane frames
Find the reaction force at B in the plane frame shown:
B

10 kN 20 kNm 6m

3m
A C

10 m
Free body diagram:
B P in-joint

20
6m 10
A C
HA HC
VA VC

2.2 Example: plane frames 11


Static equilibrium.
Resolve horizontally: HA + 10 − HC = 0

Resolve vertically: V A + V C = 0
2
Take moments clockwise about A: 10 × 3 + 20 − 10V C = 0

Loaded frames
and V C = 5 kN and V A = − 5 kN

+5VA = 0 MB=0 by definition!


10

HA

VA

For member AB, take moments clockwise about B: 10 × 3 − 6HA + 5V A = 0

Hence HA = 55/6 kN and HC = 5/6 kN

and V C = 5 kN and V A = − 5 kN

MB=0 by definition!

20 HC

VC 5

For member BC, take moments clcokwise about B: 20 + 6HC − 5V C = 0

Hence HC = 5/6 kN and HA = 55/6 kN

and V C = 5 kN and V A = − 5 kN.

For forces at B, use free body diagram:


HB B

VB 20 kNm 6m

C
5m HC
VC

Resolve horizontally: HB = HC

Resolve vertically: V B + V C = 0

Hence HC = HB = 5/6 kN

2.2 Example: plane frames 12


and V B = −V C = − 5 kN.

To verify previous calculations:


For forces at B, use free body diagram:
2
VB
B HB

Loaded frames
10 kN 6m
A
HA
5m
VA

Resolve horizontally: 10 + HA = HB

Resolve vertically: V A = V B = − 5 kN.

Take moments clockwise about A: 10 × 3 + 5V B − 6HB = 0

Hence HB = 30 − 25 = 5/6 kN.

2.3 Pin-jointed frames


A pin-jointed frame consists of straight bars, connected at joints which cannot be
relied on to prevent rotation. The joints are therefore assumed to be free to rotate.
As well as assuming that joints are free to rotate, we make two other assump-
tions:
• External forces are applied only at the joints
• The weight of the bars themselves is assumed to be negligible.

The reaction moment at a pin is therefore zero: hence M = Qr = 0 and Q = 0.


Another way to consider the bar is to recognise that the bar is a two-forces
body and, therefore it can only be in equilibrium if the forces are colinear:

2.3 Pin-jointed frames 13


F2

Loaded frames
F1

RELATED LINKS
Canadian Institute of Steel Construction educational website on steel framed structures

2.3.1 Analysis of a two-bar frame


In the frame shown:
0.9m 1.6m

VA VC
HA HC C
A

1.5m 1.2m
2m

B
10kN
What are the tensions in bars AB and BC and the horizontal and vertical reac-
tions at A and C?
1. Draw an FBD of the joint.

TAB
TBC

10kN

2. Resolve horizontally.
−T ABsin α + T BCsin β = 0

Using the information that sin α = 35 , cos α = 45 , sin β = 45 , cos β = 35 :

2.3.1 Pin-jointed frames 14


4 3
T
5 BC
= 5 T AB

hence T BC = 34 T AB.

3. Resolve horizontally. 2
−T ABcos α + T BCcos β − 10 = 0

Loaded frames
4 3 3
T +4
5 AB
× 5 T AB = 10

9
4 + 4 T AB = 50.

4. Hence solve for tensions in bars: T AB = 8 kN and T BC = 6 kN.

5. Resolve at A to find reactions there.


Horizontally: HA+T ABsin α = 0 hence HA = − 4.8 kN.
Vertically: V A+T ABcos α = 0 hence V A = 8 × 45 = 6.4 kN.

6. Resolve at C to find reactions there.


4
Horizontally: HC+T BCsin β = 0 hence HC = 6 × 5
= 4.8 kN.

Vertically: V C−T BCcos β = 0 hence V C = 6 × 35 = 3.6 kN.

Important
Notice that all tensions and reactions have been found by static equilibrium alone: this sys-
tem is statically determinate.

Not all systems are statically determinate.

RELATED LINKS
Animations and additional material on pin-jointed frames

2.4 Statical determinacy


A structure which is statically determinate is one in which all forces can be calcula-
ted from a knowledge of external, applied forces using equilibrium alone. Not all
structures are statically determinate.
Forces in the frame shown below, in which A, D and C are in the same horizon-
tal plane, cannot be found by equilibrium alone. Even if the vertical force at B
is zero, forces can be generated in the three bars if one is slightly shorter or
longer than the diagram suggests.

2.4 Statical determinacy 15


VA VD VC

HA A HD D HC C

Loaded frames
B

Attempting to solve by equilibrium:


• Each joint provides two equations, so there are 2 × 4 = 8 equilibrium
equations
• There are 6 unknown reaction forces and 3 unknown bar tensions, hence
6 + 3 = 9 unknowns

• Hence the problem cannot be solved by equilibrium: it is statically inde-


terminate.
To determine forces for a statically indeterminate structure, we need to consid-
er the compatibility of displacements at B: i.e. the need for the three B ends to
remain in the same position as the three bars stretch under load.

2.4.1 Evaluating statical determinacy


A simple function of the number of bars, joints and reactions in a frame can help
to assess whether it is statically determinate or not.
Consider the problem of determining whether the frame shown below is stati-
cally determinate or not.

A D

C
B

1. Count the number of bars, b.


2. Count the number of reactions at supports, n.
3. Count the number of bar-end joints (i.e. pins), j.
4. Since there are b + n unknown forces and reactions and each joint gives
two equilibrium equations (2 j in total), calculate b + n − 2 j to determine
statical determinacy.
There are three possible distinct results:
• b + n − 2 j = 0 is a necessary but not sufficient condition for statical de-
terminacy (the placing of bars is important).

2.4.1 Statical determinacy 16


• b + n − 2 j > 0 indicates that the structure is statically indeterminate.

• b + n − 2 j < 0 indicates that the structure is a mechanism.


In the system shown b = 3, n = 4 and j = 4. Hence
b + n − 2 j = 3 + 4 − 2 × 4 = − 1 and the structure is a mechanism. 2

2.4.2 Example: Statical determinacy of 6-bar frame

Loaded frames
This example demonstrates that the statical determinacy test for frames must be
used with care.
Determine whether the structure shown is statically determinate:

1. Count the number of bars, b = 6.


2. Count the number of reactions at supports, n = 8.
3. Count the number of bar-end joints, j = 7.
4. b + n − 2 j = 0, which is a necessary condition for statical determinacy.
This is an example of where the rule fails: the structure is not statically deter-
minate. It consists of two regions:
1. The upper three bars form a statically indeterminate structure
2. The lower three bars form a mechanism, the point at which it meets the
upper three bars being effectively a fixed support.
The b + n − 2 j = 0 values for these two regions cancel.

2.5 Example: Five-bar frame


Determine the force in each member of the truss. Indicate whether the mem-
bers are in tension or compression.

2.4.2 Statical determinacy 17


B 10kN

Loaded frames
2m

A 1m D 1m C

60kN

1. Determine whether the structure is statically determinate.


a. b=5
b. n=3
c. j=4
d. b + n − 2 j = 8 implies that the structure is statically determinate,
as neither a statically indeterminate sub-structure nor a mechanism
sub-structure can be identified.
2. Resolve horizontally for the entire structure.
HC = 10 kN

3. Resolve vertically for the entire structure.


V A + V C = 60 kN

4. Take moments around C


10 × 2 + 2V A − 60 × 1 = 0 hence V A = 20 kN and
V C = 60 − V A = 40 kN.

5. Resolve forces at A

TAB
A 45°

20 TAD

40
Vertically: T ABsin 45° + 20 = 0 hence T AB = − √2
= − 28.3 kN

Horizontally: T AD+T ABcos 45° = 0 hence T AD = 20 kN

2.5 Example: Five-bar frame 18


6. Resolve forces at C

TCB

TCD C 10 2

40

Loaded frames
Vertically: T CB = − 40 kN

Horizontally: hence T CD = − 20 kN

7. Resolve forces at D

TDB

20 α=63.44°
D 10
60

Vertically: T DBsin α − 60 = 0 kN
60
hence T DB = sin α
= 67.1 kN.

2.6 Example: Generalised, statically determinate


frame
Find all reaction forces and the tension in every bar for the frame shown. Every
bar is of length L.
B C D

A G F E

HA
W
RA RE

• Note that all angles are 60°


• Initially, assume that all bars are in tension, exerting outward forces on
joints
1. Assume static equilibrium and resolve forces and moments for the entire
frame:

2.6 Example: Generalised, statically determinate frame 19


Moments clockwise about E: RA × 3L − W × 2L = 0, hence RA = 23 W

Horizontal forces: HA = 0

Vertical forces: RA + RE − W = 0, hence RE = 13 W . 2


2. Begin to move through the structure joint by joint, resolving vertically

Loaded frames
and horizontally at each:
In this example, start at A

TAB

60°
TAG

RA
Resolve vertically: T ABsin 60° + RA = 0

Resolve horizontally: T ABcos 60° + T AG = 0


2 2 2 4
Hence T AB = − R
3 A
= −
3
× 3W = −
3 3
W

4 2
and T AG = 3 3
Wcos 60° =
3 3
W

3. If possible, choose the sequence of joints so that each one has only two
remaining unknowns.
In this example A, B, G, C, F, D, E.
a. Sketch an FBD at B:

TBC
60o
60o
TAB TBG

Resolve vertically: T ABcos 30° + T BGcos 30° = 0

Resolve horizontally: T BGcos 60° + T BC − T ABcos 60° = 0


4
Hence T BG = − T AB = 3 3
W

and T BC = 12 T AB − 12 T BG = − 4
3 3
W

b. Sketch an FBD at G:

2.6 Example: Generalised, statically determinate frame 20


TBG TCG

60o! 60o
TAG TGF
2
W

Loaded frames
Resolve vertically: T BGsin 60° + T CGsin 60° = W

Resolve horizontally: T CGcos 60° + T GF − T BGcos 60° − T AG = 0


2 2
Hence T CG = 3
W − T BG =
3 3
W

and T GF = 12 T BG − 12 T CG + T AG = 1
3
W

And then onwards to joins C, F, D and E.


W
Defining k = 3 3
the bar tensions and reactions are:

B -4k C -2k D

-4k
4k 2k -2k 2k -2k

A G! F E
2k 3k k
W
2W/3 W/3

2.7 Analysing frames by the method of sections


The method of sections is particularly useful when the forces in only a few speci-
fied bars are required. An appropriately chosen portion of the frame is cut out and
analysed using a Free Body Diagram.
Find the forces in bars CD and CF for the frame shown. Every bar is of length L.
B C D

A G F E

HA
W
RA RE

2.7 Analysing frames by the method of sections 21


1. Assume static equilibrium and resolve forces and moments for the entire
frame.

Moments clockwise about E: RA × 3L − W × 2L = 0, hence RA = 23 W

Horizontal forces: HA = 0
2

Vertical forces: RA + RE − W = 0, hence RE = 13 W

Loaded frames
2. Cut out from the frame a free body on which the bar forces which you
need to evaluate act as external forces:
In this example, cut bars CD, CF and (to release the free body) GF.
B C
TCD

TCF (√3/2)L

A G TGF F

RA=2W/3 W

3. Apply static equilibrium to the extracted free body.


3
Moments clockwise about F: RA × 2L + T CD × 2
L − WL = 0, hence
2 2
T CD = W − 2RA = − W
3 3 3

Vertical forces: RA − W − T CFsin 60° = 0, hence


RA − W 2
T CF = sin 60° = − W
3 3

2.8 Maximum allowable load on a frame


Possible failure modes in a pin-jointed frame are the shear failure of a pin, tensile
or compressive yield of a bar or buckling of a bar. As the externally applied load is
increased, the mode which occurs first in any member will determine the maxi-
mum allowable load.
For each failure mode, the maximum load on the frame depends on the fol-
lowing factors:

Shear failure of a pin


Proportional to its cross-section and the shear strength of the material;
inversely proportional to the total load on it

2.8 Maximum allowable load on a frame 22


Tensile yield of a bar
Proportional to its cross-section and the tensile yield strength of the ma-
terial; inversely proportional to the tensile load on it
2
Compressive yield of a bar
Proportional to its cross-section and the compressive yield strength of

Loaded frames
the material; inversely proportional to the compressive load on it

Buckling of a bar
Proportional to the second moment of area of its cross-section and the
Young’s modulus of its material; inversely proportional to the square of
its length.
Note
The second moment of area is a function of the beam cross-sectional shape, characterising
its resistance to bending.

2.8.1 Example: maximum allowable load on a 2-bar frame


Both bars in the frame shown are made of the same material and have the
same cross-sectional dimensions. The load W may be applied either upwards,
as shown, or downwards.
0.9m 1.6m

VA VC
HA HC C
A

1.5m 1.2m
2m

B
W

Determine which member will fail first by:


1. Tensile yield
2. Compressive yield
3. Buckling.
1. Use static equilibrium analysis to determine the tension (here compres-
sion, i.e. negative tension) in each bar.
T AB = − 0.8W

T BC = − 0.6W

2.8.1 Maximum allowable load on a frame 23


2. Calculate −TL2 for each beam.

−T ABL2 2
AB = − 0.8W × 1.5 = 1.8W

−T BCL2 2
BC = − 0.6W × 2.0 = 2.4W 2
Using these results, as the magnitude of W increases:

Loaded frames
1. While W acts downwards, bar AB is under the greatest tensile load and
will be the first to fail by tensile yield.
2. While W acts upwards, bar AB is under the greatest compressive load and
will be the first to fail by compressive yield.

3. While W acts upwards, bar BC has the greatest value of −TL2 and will be
the first to fail by buckling.

2.8.2 Example: maximum allowable load on a complex frame


In this example the method of sections is used, and symmetry is exploited, to deter-
mine the bar tensions.
In the symmetrical, pin-jointed structure shown below, all bars have the same
cross-section. Angle ABH is 60°, angle BCH is 45° and all vertical bars are of
length L.
W
B C D

A E
HA H G F
L
VA VE

Determine which bar carries the greatest tensile load and which is most likely
to buckle when
1. W acts downwards
2. W acts upwards.
1. Assume static equilibrium of the entire structure to determine reaction
forces.
Take moments clockwise about A: W L + 3L − 2V E L + 3L = 0,
hence V E = 12 W .

Vertical equilibrium: V A + V E − W = 0, hence V A = 12 W .

Horizontal equilibrium: HA = 0.

2.8.2 Maximum allowable load on a frame 24


2. Make successive cuts across the structure to separate larger and larger
sections around the left-hand end, solving for reactions at each newly
added pin.
a. Cut AB, AH
2
B

Loaded frames
TAB 60o
A
TAH H
W/2

Take moments ↻ B: 12 W × 3L − T AH × L = 0 hence T AH = 1


2
3W .

Resolve vertically: 12 W + T ABcos 60° = 0 hence T AB = − W.

b. Cut BC, BH, AH

B TBC

TBH
A
TAH H
W/2

Resolve vertically: T BH = 12 W
1
Resolve horizontally: T AH + T BC = 0 hence T BC = − 2
3W .

c. Cut BC, HC, HG

B TBC

THC
A
H THG
W/2

Take moments ↻ C: 12 W 3L + L − T HG × L = 0 hence


1
T HG = 2 1 + 3 W

Resolve vertically: 12 W + T HCcos 45° = 0 hence T HC = − 1


2
W.

d. Cut AB, AH and invoke symmetry

2.8 Maximum allowable load on a frame 25


TCG

THG G THG
Resolve vertically: T CG = 0. 2
1. W acts downwards:

Loaded frames
Bar Tension/W Length/L Tension × (length)2 / WL2
AB –1 2 –4.0
AH √3/2 = 0.87 √3 2.61
BH 1/2 = 0.5 1 0.5
BC –√3/2 = –0.87 1 –0.87
HC –1/√2 = 0.71 √2 –1.41
HG (1 = √3)/2 = 1.37 1 1.37

• Maximum tension T = 1.37W in HG (and FG)


2
• Maximum value of − TL 2 = –4.0 in AB (and DE)
WL

2. W acts upwards: all bar tension signs are reversed

Bar Tension/W Length/L Tension × (length)2 / WL2


AB 1 2 4.0
AH –√3/2 = 0.87 √3 –2.61
BH –1/2 = 0.5 1 –0.5
BC √3/2 = –0.87 1 0.87
HC 1/√2 = 0.71 √2 1.41
HG –(1 = √3)/2 = 1.37 1 –1.37

• Maximum tension W in AB (and DE)


2
• Maximum value of − TL 2 = –2.61 in AH (and EF)
WL

2.9 Notes

2.9 Notes 26
Loaded frames
2

27
2.9 Notes
Loaded frames
2

28
2.9 Notes
Chapter 3 Stress
Stress is the mechanical force per unit area of the material through which the force
is transmitted.
On successfully completing this module, you will be able to:
3
• Define stress and strain
• Discuss material strength and links to failure

Stress
• Define statically determinate and statically indeterminate systems
• Analyse statically determinate systems, e.g. pressure vessels and blocks
• Build free body diagrams and sketch stresses acting over surfaces in sim-
ple systems

3.1 Stress and strength


Under increasing stress any solid will first extend (strain) elastically and then, in
most cases, will begin to yield (deform permanently). Ultimately the material will
tear or fracture. The strength of a material is the maximum stress it can withstand
before it fails, whether by yield, tearing or fracture.
The applied load at which a component fails depends on the distribution of
stress which transmits the load through it. The higher the load, the greater is
the stress at every point. The objective of stress analysis is to determine the
load at which the component will fail because its material has at some point
reached a critical stress.
A normal stress or direct stress acts on a surface when a force is applied normal
to it:

Fn
Mean normal stress σn = An
in units of Nm–2 (MN–2 etc.).

Another name for a stress or pressure of one Nm–2 is a pascal, symbol Pa (kPa,
MPa etc.).
Stress is defined as positive when the normal load is tensile:

and tensile when the load is compressive:

Chapter 3 Stress 29
A shear stress acts on a surface area when a force is applied parallel to it:
As
Fs 3

Stress
Fs
The mean shear stress σs = As
has the same units as direct stress.

Strength σmax for some common materials:

Material Strength (MPa)


Iron, steel 300
Aluminium 200
Cu-Be alloy 1000
Nylon 60

RELATED LINKS
Missouri University of Science and Technology MecMovies: see Ch 1, Stress

3.2 Stresses in statically determinate systems


In a statically determinate system under applied load, the load can be found at
every point by considering equilibrium alone. If the component geometry at each
point is known, stresses there can be determined too.

3.2.1 Thin, pressurised cylinders: hoop stress


Many industrially important pressurised structures. e.g. pipelines, can be represen-
ted as thin-walled cylinders. This analysis determines the circumferential (‘hoop’)
stresses acting within the wall.
The cylinder shown below, whose thickness t is much smaller than its mean
diameter 2Rm, is subjected to internal pressure P compared to the pressure
outside it. We need to determine the hoop stress or circumferential stress which
acts on any r, z plane (i.e. a plane of constant θ) within the wall.

3.2 Stresses in statically determinate systems 30


3

Stress
We will assume the wall thickness t to be small enough compared to the mean
radius Rm for the assumption that Rm ≈ Ri, where Ri is the inner radius, is val-
id.
1. The cylinder is in static equilibrium. Remove a semi-circular element of
the wall, Δz in length, and draw the free-body diagram showing the
forces acting on it.
The lower half of the circular section is removed, and replaced by stresses
σθ acting on the upper half which acted across the plane joining them.

2. To resolve forces F vertically upwards, first determine the upward force


due to pressure acting on the internal surface.
Radial force on an element of wall which subtends an angle δα: PRiδθΔz

Vertical force on this element: PRiδθΔzsin θ.

Hence total vertical force due to pressure FP = ∫π0 PRiΔzsin θ ⅆθ

= PRiΔz − cos θ π0 = 2PRiΔz.

Tip
This result can be derived more simply by noting that the total upward force FP due
to pressure is simply the product of the pressure and the area, projected in the verti-
cal direction, on which it acts.

3.2 Stresses in statically determinate systems 31


3. Next, determine the downward force due to circumferential stress acting
on the cut section of wall.
Fσ = 2σθtΔz.

4. Apply vertical equilibrium to show that the total vertical force is zero, and
solve for σθ.
3
Since FP = Fσ,2PRiΔz = 2σθtΔz

Stress
PRi
hence σθ = t
.

3.2.2 Thin, pressurised cylinders: axial stress


The forces due to pressure acting on the closed ends of a pressurised cylinder tend
to push them off. The axial stress in the cylinder wall depends on whether or not
the ends are attached to the wall, which then has to carry these forces.
We will first analyse the case of a cylinder with both ends closed (i.e. both ends
attached to the cylinder wall). The entire cylinder is in equilibrium.
Assume that the wall thickness t is small enough compared to the mean radi-
us Rm to assume that Rm ≈ Ri, where Ri is the inner radius.

1. The pressurised shell is in static equilibrium. Cut across its diameter at


some z = const., remove the right-hand part and draw a free-body dia-
gram of the rest showing forces equivalent to it.

2. To resolve forces F in the horizontall direction, determine the leftward


force due to pressure.

FP = πR2
iP

3. Determine the rightward force due to normal stress σz acting on the cut-
through wall.
Fσ = 2πRmtσz.

4. Apply horizontal equilibrium to show that the total axial force is zero, and
solve for σθ.

Since FP = Fσ, πR2i P = 2πRmtσz


2
P Ri PRi
hence σz = 2t Rm
≈ 2t .

3.2.2 Stresses in statically determinate systems 32


In an open-ended cylinder, the axial pressure force FP is not transferred to
the cylinder wall: one or two internal pistons transfer the force to some
external structure. The axial stress in the cylinder wall is therefore zero.

Stress
Note
A typical example of an open-ended cylinder is a pneumatic or hydraulic actuator,
from which the pressure force is applied to an external load.

3.2.3 Thin, pressurised spherical shell: stress


Thin-walled shells are used for pressure vessels wherever possible. Analysing
stresses in the wall will help to show us why.
We will assume the wall thickness t to be small enough compared to the mean
diameter Rm for the assumption that Rm ≈ Ri, where Ri is the inner radius, to
be valid.
1. The pressurised shell is in static equilibrium. Cut across a horizontal diam-
eter, remove the lower half and draw a free-body diagram of the upper
half showing forces equivalent to the part removed.

t r y r
A
x
Pressure, P z

2. To resolve forces F in the vertical direction, determine the upward force


due to pressure.

FP = πR2
iP

3. Determine the downward force due to normal stress σθ acting on the cut-
through wall circumference of width t.
Fσ = 2πRmtσθ.

4. Apply horizontal equilibrium to show that the total vertical force is zero,
and hence solve for σθ.

Since FP = Fσ,πR2i P = 2πRmtσθ


2
P Ri PRi
hence σθ = 2t Rm
≈ 2t .

3.2.3 Stresses in statically determinate systems 33


Since every diameter in a spherical shell is equivalent, the circumferential
stress anywhere in a spherical shell is PRi /2t. The circumferential stress in a cy-
lindrical shell of the same radius and thickness is twice this value.

RELATED LINKS
Missouri University of Science and Technology MecMovies: see Ch 14, Pressure vessels
3
3.2.4 Example: pressure vessel

Stress
The 300 mm diameter, 6 mm thick pressure vessel shown is to be made up of
two hemispherical shells bolted together at flanges. The maximum allowable
tensile stress in the material to be used is 150 MPa, what is the maximum al-
lowable working pressure?

1. Rearrange the equation for stress in a spherical shell as a function of pres-


sure to give pressure as a function of stress.
2t
σθ = PRi /2t , hence Pmax = R σθmax
i

2 × 0.006
Hence Pmax = 0.15
× 150 × 106 Pa = 12 MPa.

3.2.5 Example: pressure vessel bolts


A 300 mm diameter, 6 mm thick pressure vessel shown is to be made up of
two hemispherical shells bolted together at flanges. Twenty bolts of 6 mm di-
ameter are used to hold the flanges together. What is

3.2.4 Stresses in statically determinate systems 34


3

Stress
3.2.6 Thin rotating ring
If a cylindrical ring rotates about its axis at high speed every element experiences
an outward centrifugal force, just as it would do from internal pressure. Like inter-
nal pressure, this develops a circumferential stress, which can be calculated by
analogy.
This analysis derives an equation for the circumferential stress in a ring of
mean diameter Rm, thickness t ≪ Rm and material density ρ, at a rotational
speed ω.

1. Draw the free body diagram for a circumferential element of included an-
gle δα, and determine the radial centrifugal force on it.

3.2.6 Stresses in statically determinate systems 35


Fr = mRmω2 = ρRmδαtΔz Rmω2

2. In order to use the results of analysis for stress in an internally pressurised


thin cylinder, convert this to an ‘equivalent internal pressure’, Pω.

For the internal-pressure case, Fr = Pω RiδαΔz

so that 3

ρRmδαtΔz Rmω2 = Pω RiδαΔz and hence

Stress
ρR2 2
mω t
Pω = Ri
.

3. Use the result for an internally-pressurised cylinder to derive the circum-


ferential stress due to rotation.
PωRi
σθ = t
= ρR2 2
mω .

3.3 Notes

3.3 Notes 36
3

Stress

3.3 Notes 37
3

Stress

3.3 Notes 38
Chapter 4 Elastic strain
When a solid is put under stress, its shape changes slightly as internal bonds
stretch. The resulting small displacement of particles within a body — without
movement of the body as a whole — are defined as strain; if it disappears when
the stress is removed, the strain is elastic.
On successfully completing this section, you will be able to:
4
• Recall Hooke’s law for uniaxial tension and its generalised form for three-
dimensional systems

Elastic strain
• Determine stress and strains in thin-walled pressure vessels and solid
blocks
• Define volumetric strain
• Determine relationships between elastic constants
• Define and calculate strain energy for individual bodies loaded in tension
and shear
• Identify and analyse statically indeterminate systems

4.1 Direct strain


Direct strain is the extension of a line along a specified direction within the materi-
al.
A bar of initial length L and cross-sectional area A is subject to an axial force F
which causes an increase in length of ΔL.

F
The direct stress at any cross-section is σ = A
.

The direct strain along the axis (in this case the longitudinal, or axial strain) is
ΔL
e= L .

Sign convention for strain: an increase in length L is positive, giving positive


strain. Positive strain is therefore associated with positive (tensile) stress.

RELATED LINKS
Missouri University of Science and Technology MecMovies: see Ch 1, Stress

Chapter 4 Elastic strain 39


4.2 Elastic stress-strain relationships
The relationship between stress and strain is a property of the material to which
either stress or strain is applied.
To measure this relationship, apply an increasing tensile force along the axis of
a bar and measure the increase in length of the bar along the same axis. Plot
as stress versus strain to obtain a uniaxial stress-strain plot:
4

Elastic strain
• If e returns to zero as soon as the load is reduced again to zero (e.g. as for
a rubber), the behaviour of the material is described as elastic.
• If the relationship between σ and e is linear up to some point, the behav-
iour of the material there is described as linearly elastic. Many materials
are linearly elastic at small strains.
For these linearly elastic materials Hooke’s Law holds:
σ
e = E,

where the constant E is the elastic modulus or Young’s modulus.

RELATED LINKS
Animation: linear elastic stress and strain
Missouri University of Science and Technology MecMovies: see Ch 3, Mechanical Properties

4.3 Poisson’s ratio


If a bar is loaded uniaxially in tension, its axial extension is accompanied by con-
traction of its cross-section. The strains in the two directions of its cross-sectional
plane are negative, and proportional to the axial strain.
To demonstrate Poisson’s ratio, load a bar of uniform cross-section in tension
σ x. If the material is linearly elastic, it will extend axially by a strain ex = σ x /E.

Important
For most materials, every cross-sectional plane along the axis of a uniaxially extended bar
will contract by an equal strain in every direction. These in-plane strains will be a constant
proportion −ν of the axial strain.

For example:

4.2 Elastic stress-strain relationships 40


• If the cross-section is rectangular, of width Ly and height Lz, these dimen-
sions will change by a strain −νex: they will contract to Ly 1 − νex and
Lz 1 − νex

• If the cross-section is circular, of radius R, the radius will change by a


strain −νex: it will contract to R 1 − νex .

The ratio ν is known as Poisson’s ratio, and is a constant of the material.


If Poisson’s ratio is 0.5, when the bar is extended its volume will not
4

change (axial extension is balanced by lateral contraction)

Elastic strain
• If Poisson’s ratio is positive and less than 0.5, when the bar is extended in
one direction its volume will increase
• If Poisson’s ratio is negative, when the bar is extended in one direction its
volume will increase. Such auxetic materials are very rare.

4.3.1 Typical Poisson’s ratio values

Material Poisson’s ratio


Rubber 0.5
Lead 0.48 to <0.5
Aluminium 0.44
Polystyrene 0.35
Brass 0.33
Ice 0.33
Polystyrene foam 0.30
Steel 0.29
Beryllium 0.08

4.4 Stress and strain in three dimensions


A stress in one direction creates strains in both of the other dimensions. If the ma-
terial is linearly elastic and stresses are applied in these directions too, the strains
simply add.
To analyse the interaction of stresses and strains in all three dimensions, con-
sider a rectangular block with uniform stresses σ x, σ y, σz applied to opposite
faces along all three directions.

4.3.1 Poisson’s ratio 41


σz

Elastic strain
1. Tabulate the strains due to each stress in turn acting alone.

Direction x y z
Strain ex ey ez

Strain due to σ x σ x /E −νσ x /E −νσ x /E

Strain due to σ y −νσ y /E σ y /E −νσ y /E

Strain due to σz −νσz /E −νσz /E σz /E

2. Add the strain contributions from each stress to give the total strain in
each direction.
σx σy σz 1
ex = E −ν E −ν E = E σ x − ν σ y + σz
σy σz σx 1
ey = E −ν E −ν E = E σ y − ν σz + σ x
σz σx σy 1
ez = E −ν E −ν E = E σz − ν σ x + σ y

3. If the block experiences a uniform temperature increase ΔT , there are


equal strains αΔT in all three directions where α is the coefficient of linear
thermal expansion. Add these to these thermal strains to those due to
stress.
1
ex = E σ x − ν σ y + σz + αΔT
1
ey = E σ y − ν σz + σ x + αΔT
1
ez = E σ z − ν σ x + σ y + αΔT

The resulting equations are referred to as the Generalised Hooke’s Law.

4.4 Stress and strain in three dimensions 42


4.4.1 Hydrostatic stress and volumetric strain
A material is under hydrostatic stress when an equal normal stress acts on it in ev-
ery direction.
In this case we can define the hydrostatic stress σH as

σ H = σ = σ y = σ z.
x

The strain will be equal in every direction too, and will cause a change in vol- 4
ume. The volumetric strain, ev is defined as the change in volume V divided by
the original volume:

Elastic strain
ΔV
ev = V .

The bulk modulus, K is the ratio of volumetric strain to volumetric stress:


σH
K= e
V

and is a material property.

4.4.2 Relating elastic modulus, Poisson’s ratio and bulk


modulus
The generalised Hooke’s law equations can be used to derive the relationship be-
tween these three material constants.
Stresses σ x, σ y and σz are applied to an elemental block of size Δx, Δy and Δz
in the x, y and z directions.
1. Calculate the volumetric strain of the block in terms of strains ex, ey and
ez :

Δx increases by Δx × ex to a new length of Δx′ = Δx 1 + ex

Δy increases by Δy × ey to a new length of Δy′ = Δy 1 + ey

Δz increases by Δz × ez to a new length of Δz′ = Δz 1 + ez

Hence

ΔV= Δx′Δy′Δz′ − ΔxΔyΔz


= Δx 1 + ex Δy 1 + ey Δz 1 + ez

= ΔxΔyΔz 1 + ex + ey + ez + exey + eyez + ezey − 1

Since strains are small, products of strains can be neglected so that

ΔV = ΔxΔyΔz ex + ey + ez

ΔV
and hence eV = V
= e x + e y + ez .

4.4.1 Stress and strain in three dimensions 43


2. Use generalised Hooke’s law expressions to express strains in terms of ap-
plied stresses.
1 1 1
eV= E σ x − ν σ y + σz + E σ y − ν σz + σ x + E σz − ν σ x + σ y
1 ν
= E σ x + σ y + σz − E 2σ x + 2σ y + 2σz
1 − 2ν
= σ x + σ y + σz E

3. Set all three stresses equal to the hydrostatic stress σH to derive the bulk 4
σH
modulus K = eV

Elastic strain
1 − 2ν
Hence eV = 3σH E
and
E
K = 3 1 − 2ν

or E = 3K 1 − 2ν .

4.4.3 Example: volume change due to stress and temperature


Find the increase in volume of a 200 mm long and 100 mm wide rectangular
steel plate, of 50 mm thickness, whose temperature increases by 20°C while a
compressive stress of 120 MPa is applied to both faces and a tensile stress of
60 MPa is applied to the smallest edges. Take the tensile modulus E to be
200 GPa, Poisson’s ratio ν to be 0.3 and α = 10 × 10−6 K–1.
120 MPa
200 mm
100 mm
60 MPa
60 MPa
50 mm
120 MPa

What stress must be applied to the previously unstressed edges if there is to


be no change in the 100 mm dimension due to the combined effect of all
stresses and the increase in temperature?
1. Define coordinate axes.
In-plane x (length) and y (width); thickness z
2. Apply generalised Hooke’s law equations to calculate linear strains
3. Add linear strains to find the volumetric strain, and multiply by the initial
volume to find the change in volume
4. Repeat steps 2 accounting for an additional, unknown stress σ y

5. Set ey to zero to determine σ y.

4.4.3 Stress and strain in three dimensions 44


4.5 Strains in cylindrical and spherical shells
Consider a thin-walled cylinder of radius R and thickness t ≪ R:

4
If the cylinder expands under internal pressure, the axial strain is simply the

Elastic strain
increase in length divided by the original length:
ΔL
ez = L .

Similarly the circumferential strain is the increase in circumference divided by


the original circumference:
2π R + u − 2πR u
eθ = 2πR
=R

where u is the increase in radius caused by extension of the circumference.

Remember
u/R is the circumferential strain, not the radial strain! There is no tensile stress to extend
any material in the radial direction.

For a spherical shell, the circumferential strain for any circumference is also
u/R.
The real radial strain in any shell is the increase in radial length, divided by the
original radial length, of an element of material in the cylindrical shell. The
average radial strain is the increase in wall thickness divided by the original
wall thickness.

4.6 Apparent modulus of elasticity


The extension of a uniaxially loaded bar under axial force will indicate the true
modulus of elasticity only if the bar is free to contract in the lateral, unloaded di-
rections. If it is not, a higher elastic modulus will be measured.
If a stress is applied in the x direction and the lateral surfaces are free, there
will be Poisson contraction in the y and z directions. If they are prevented

4.5 Strains in cylindrical and spherical shells 45


from contracting, there must be tensile stresses σ y and σz of the right magni-
tude to cancel out the contraction, and these stresses will cause a contraction
which reduces the strain in the x direction.

σz

Elastic strain
From the generalised Hooke’s law equations with no temperature change:
1
ex = E σ x − ν σ y + σz + αΔT

The apparent modulus of elasticity is


σx E
Eapparent = e =
x 1 − ν σ y /σ x + σz /σ x

and σ y and σz can be calculated by assuming that ey and ez are zero.

4.7 Volumetric strain within shells


When the wall of a cylindrical or spherical shell stretches, the shell increase in size
and the internal volume increases. Analysing this effect allows us to solve prob-
lems concerning, for example, the additional volume of liquid needed to pressurise
a filled shell.

4.7.1 Volumetric strain within a cylindrical shell


Derivation of the volumetric strain for the internal volume of a cylindrical shell
when its length and circumference increase.
A cylindrical shell of radius R and length L.

1. The volume of the unstrained cylinder is V = πR2L


2. Now let there be small changes in R and L, and determine the resulting
change in volume.
∂V ∂V
Using partial differentiation (δV = ∂R
δR + ∂L δL):

δV = π ⋅ 2RLδR + πR2δL.

4.7 Volumetric strain within shells 46


3. Divide by the original volume to express this change in volume as a volu-
metric strain.
δV π ⋅ 2RLδR + πR2δL δR δL
V
= = 2 R + L = 2eθ + ez
πR2L

Hence ev = 2eθ + ez.

4.7.2 Volumetric strain within a spherical shell


4
Derivation of the volumetric strain for the internal volume of a spherical shell
when its circumference increases.

Elastic strain
A spherical shell has radius R.

1. The volume of the unstrained sphere is V = 43 πR3L

2. Let there be a small changes in R, and determine the resulting change in


volume.
ⅆV
Since δV = ⅆR
δR:

4
δV = 3 π ⋅ 3R2δR.

3. Divide by the original volume to express this change in volume as a volu-


metric strain.
δV δR
V
= 3 R = 3eθ

Hence ev = 3eθ.

4.7.3 Stress-strain relationships in polar coordinates


For any point in a cylindrical solid, the generalised Hooke’s law equations be-
come:
1
eθ = E σθ − ν σr + σz + αΔT
1
er = E σr − ν σz + σθ + αΔT
1
ez = E σ z − ν σ θ + σ r + αΔT

For any point in a spherical solid, the generalised Hooke’s law equations be-
come:
1
eθ = E σ θ − ν σ r + σ ϕ + αΔT
1
er = E σ r − ν σ ϕ + σ θ + αΔT
1
eϕ = E σϕ − ν σθ + σr + αΔT

4.7.2 Volumetric strain within shells 47


4.8 Elastic strain energy
If a material extends elastically when stress is applied to it, potential energy has
been stored in it. When the stress is reduced to zero, the energy is immediately re-
covered.
An axial force is applied to a bar of linearly elastic material, L long and A in
cross-sectional area, the point at which the force is applied will move due to
direct strain. 4

Elastic strain
Load
Extension
The external work done during a small increment of extension δx is Fδx and
the total area under the curve for an extension of ΔL is 21 FΔL.
1
2
W ⅆL
The stored strain energy per unit volume is AL
.

Since strain e = δL/L, stress σ = F/ A and tensile modulus E = σ/e, the elastic
strain energy per unit volume stored by applying a direct stress is 21 σ2 /E.

If a shear force is applied to a bar of linearly elastic material, L long and A in


cross-sectional area, the force will be displaced due to shear strain.

Following the same argument, the elastic strain energy per unit volume stored
by applying a direct stress is 12 τγ = 12 τ2 /G.

4.8 Elastic strain energy 48


4.9 Solving statically indeterminate stress-strain
problems
A statically indeterminate problem cannot be solved by stress or load equilibrium
alone. However, if we know the stress/strain relationship for the material (e.g. line-
arly elastic) it can be solved by considering strain compatibility.
A general method of solution is shown in the flow chart below:
4
System

Elastic strain
More unknowns
than equations
FBD, reactions,
tensions, etc Equations in
Statically terms of strains
indeterminate
Apply static Geometry of
equilibrium deformation

Statically
Stress-strain
determinate relationships

Equations in
terms of stresses
Solve all equations

1. Define the system, identifying all possible external loads and reactions
acting on it.
2. Draw a free body diagram and apply static equilibrium conditions to form
up to three equilibrium equations.
3. Apply the method of sections to successive subcomponents (e.g. bars) of
the system until there are as many equations as unknowns.
4. If there are enough equations, the system is statically determinate: solve
them all.
5. If there are more unknowns than equations, analyse the geometry of de-
formation to derive compatibility equations in terms of strain, deflection
or extension.
6. Use stress-strain (or load-deflection) relatonships to convert these com-
patibility equations into terms of stress (or load), until there are as many
equations as unknowns.
7. Solve all equations.

4.9.1 Example: statically determinate problem

4.9 Solving statically indeterminate stress-strain problems 49


Bar 1, of length L1 and cross-sectional area A1, is made of a material with mod-
ulus E1. It is connected in series with bar 2, of length L2 and cross-sectional
area A2, made of a material with modulus E2. One end of the combined bar is
fixed and a force F is applied at the other. What are the axial displacements
where the bars meet, and at the load point?
1. Take the left-hand end as being fixed.

Elastic strain
2. Draw a free body diagram and apply static equilibrium conditions to form
up to three equilibrium equations.

3. Apply the method of sections to successive subcomponents (e.g. bars) of


the system until there are as many equations as unknowns.
For horizontal equilibrium of each of bars 1 and 2, F1 = F2 = F.

4. If there are enough equations, the system is statically determinate: solve


them all.
This problem is now already solved. Displacements can be calculated, but
they did not have to be calculated in order to solve for forces and stresses.
For each bar, the extension d is:
σ1 F1 d1 F1L1
e1
= E1, so that A = E1 × L and d1 = A E
1 1 1 1

σ2 F2 d2 F2L2
e2
= E2, so that A = E2 × L and d2 = A E .
2 2 2 2

From geometry, the displacement of the load point is


FL1 FL2
d = d1 + d2 = A E + A E .
1 1 2 2

F L1 L2
If A1 = A2 = A: d = d1 + d2 = A E1
+E .
2

4.9.2 Example: statically indeterminate problem


A cylinder, made from a material with modulus E1, has a length L and cross-
sectional area A1. It is is mounted between rigid end-caps. A round bar of

4.9.2 Solving statically indeterminate stress-strain problems 50


equal length, made from a material with modulus E2 and of cross-sectional
area A2, is mounted coaxially between the same end-caps. One end-cap is
fixed and a force F is applied at the other: what is the resulting displacement
at the load point?
1. Take the left-hand end as being fixed.
2. Draw a free body diagram and apply static equilibrium conditions to form
up to three equilibrium equations.
4

Elastic strain
3. Apply the method of sections to successive subcomponents (e.g. bars) of
the system until there are as many equations as unknowns.
Here, for either end-cap F = F1 + F2

4. If there are enough equations, the system is statically determinate: solve


them all.
Here, no: one equation, two unknowns.
5. If there are more unknowns than equations, analyse the geometry of de-
formation to derive compatibility equations in terms of strain, deflection
or extension.
From the geometry of this problem, for the cylinder and bar to remain at-
tached to the rigid end-caps their extensions δ1 and δ2 must be equal so
that δ1 = δ2 = δ.

6. Use stress-strain (or load-deflection) relationships to convert these com-


patibility equations into terms of stress (or load), until there are as many
equations as unknowns.

Hence F1 = E1 A1 × δL and F2 = E2 A2 × δL
δ δ δ
and F = F1 + F2 = E1 A1 × L
+ E2 A2 × L = L E1 A1 + E2 A2 .

7. Solve all equations.


F
Thus the load-point displacement is given by δ = L, from which
E1 A1 + E2 A2

E1 A 1
F1 = F and
E1 A1 + E2 A2

E2 A 2
F2 = F.
E1 A1 + E2 A2

4.9.3 Example: hydraulic cylinder with tie rods

4.9.3 Solving statically indeterminate stress-strain problems 51


A hydraulic cylinder of 80 mm inside diameter and 4 mm wall thickness is wel-
ded to rigid end plates as shown below. The end plates are tied together by
four rods of 8 mm diameter symmetrically arranged around the cylinder.
4 tie rods

Hydraulic cylinder 4

Elastic strain
Calculate the stresses in the rods and the cylinder at the cylinder design pres-
sure of 20 MPa.
The cylinder and tie rods are made from steel with a Poisson’s ratio of 0.3.
1. Define the system, identifying all possible external loads and reactions
acting on it.
There are no significant external loads on the system. Unless the piston
(not shown) is accelerating rapidly, the system is in static equilibrium.

Inside radius of cylinder R = 12 dcyl = 1


2
× 80 = 40 mm

Thickness of cylinder t = 4 mm
Diameter of each rod drod = 8 mm.

2. Draw a free body diagram and apply static equilibrium conditions to form
up to three equilibrium equations.
Free body diagram of either end cap:
end plate
tie rod

Hydraulic cylinder

With stresses in MPa:


πd2
rod
Axial load in each rod Frod = 4
σrod = 50.3σrod N

Axial load in cylinder wall


1
Fcyl = 2π R + 2 t tσz, cyl = 2π 40 + 2 ×4σz, cyl = 1055.6σz, cyl N

4.9 Solving statically indeterminate stress-strain problems 52


Axial force due to pressure = PπR2 = 100.5 × 103 N.

Hence resolving horizontally: 4Frod + Fcyl = PπR2

Substituting: 4 × 50.3σrod + 1056σz, cyl = 100.5 × 103 N

hence σrod + 5.25σz, cyl = 500 MPa.

3. Apply the method of sections to successive subcomponents (e.g. bars) of


the system until there are as many equations as unknowns. 4
In this example there are effectively no subcomponents.

Elastic strain
4. If there are enough equations, the system is statically determinate: solve
them all.
In this example there are two unknowns (σrod, σz, cyl) and one equation:
not enough.
5. If there are more unknowns than equations, analyse the geometry of de-
formation to derive compatibility equations in terms of strain, deflection
or extension.
For the cylinder of thickness t under internal pressure 20 MPa, geometry
of deformation gives
erod = ez, cyl.

6. Use stress-strain (or load-deflection) relatonships to convert these com-


patibility equations into terms of stress (or load), until there are as many
equations as unknowns.
Stress-strain relationships give
σrod
erod = E

1
ez, cyl = E σz, cyl − ν σθ, cyl + er, cyl

PR 20 × 40
eθ, cyl = t = 4
= 200 MPa

σr, cyl ≈ − P ≈ − 20 MPa, which is small compared to σθ, cyl and can be
neglected.
Hence
1
ez, cyl = E σz, cyl − 200ν MPa.

7. Solve all equations.


From deformation we have erod = ez, cyl from which
σrod
E
= σz, cyl − 0.3 × 200 MPa

σrod − σz, cyl = − 60 MPa …………… (Eq. 2)

4.9 Solving statically indeterminate stress-strain problems 53


From equilibrium we have
σrod + 5.25σz, cyl = 500 …………… (Eq. 1)

Solving Eqs. (1) and (2) simultaneously:


6.25σz, cyl = 560 MPa and finally

σz, cyl = 89.6 MPa

σrod = 29.6 MPa. 4

4.9.4 Example: pressure vessel with hemispherical ends

Elastic strain
The pressure vessel on an air-blast circuit breaker has a thin-walled cylindrical
section of mean radius R and thickness t. The vessel has hemispherical ends
and the radial expansion of the cylinder and hemispheres at their junction is
to be the same when it is subjected to pressure P:

1. Calculate the required wall thickness of the hemispheres, assuming


ν = 1/3.
2. If the length of the cylindrical section is 4R, what is the change in volume
of the vessel in terms of R, t and tensile modulus E?
32πPR4
[0.4t, 3tE
]

1. Assume static equilibrium…

RELATED LINKS
What's an air-blast circuit breaker?
Air-blast circuit breaker in action
Typical compressed air tank designs

4.10 Notes

4.9.4 Solving statically indeterminate stress-strain problems 54


Elastic strain
4

55
4.10 Notes
Elastic strain
4

56
4.10 Notes
Chapter 5 Bending of beams
A beam is a structural element which is slender relative to its length, but supports
loads perpendicular to its longitudinal axis by resisting bending. In structural engi-
neering, at least, beams are usually mounted horizontally and are usually loaded
by bearing the weight of other elements.
A beam is generally supported at points other than those at which it is loaded.
Its function is:
1. To limit deflection when the load is applied or moved (e.g. as a vehicle 5
crosses a bridge)
2. To prevent any risk of the load becoming unsupported because excessive

Bending of beams
force breaks the beam (e.g. an aircraft fuselage becoming unsupported
when the wing root breaks).
The objective of our analysis is to predict the deflections and strength of a
beam given its length, cross-section and support conditions. The first step is to
summarise the applied loading as a distribution of bending moment and of
shear stress acting at every point along the beam length.
On successfully completing this module, you will be able to:
• Define a beam
• Recognise different types of beams
• Recognise and list possible supports and reaction forces
• Build free body diagrams of simple beams
• List the possible loads and recognise their engineering symbols
• Determine shear forces and bending moments in simple beams
• Sketch diagrams for shear forces and bending moments
• Derive mathematical expressions to link applied Loads, shear forces and
bending moments.

5.1 Beam supports


Because a beam is not usually designed to be loaded along its axis, only two types
of support are important: built-in, and roller.
A built-in or encastré support can support a vertical force without deflection
and a moments without rotation.

Chapter 5 Bending of beams 57


A cantilevered beam can bend freely except at its support (the wall), where its
angle to the horizontal and its vertical position are fixed. 5
A simple support is equivalent to a pin joint or roller support. It can support a
vertical force without deflection but cannot resist rotation.

Bending of beams
A simply supported beam can bend freely at and between its supports, but it
can deflect only between supports.
Note
We did not consider whether or not each support could support loads along the axis, be-
cause these are beams (intended to resist bending) rather than bars (intended to resist ax-
ial force). A structural element can be both a beam and a bar — but we consider these two
functions separately.

Because it is not important to a beam whether a support can resist axial load
or not, a simple support is drawn simply as an abstract, triangular ‘point sup-
port’.

5.2 Modelling of loaded beams


To model a loaded beam using a free-body diagram, each support is replaced by
the normal (vertical) reaction and moment reaction it applies, loads applied over
a short distance are represented as point forces and pressure-like distributed loads
as functions of position.
The first stage for analysing a loaded beam is to reduce it to a standardised
diagram.
Note the sign convention always used for beam loading: load acting down-
wards is positive.
Example 1: A horizontal beam simply supported at its ends, and loaded by a
single mass placed on it:

5.2 Modelling of loaded beams 58


Example 2: A horizontal built-in at one end (a cantilever), loaded by two mass-
es:

Bending of beams
Example 3: A horizontal beam built-in at both ends ends, and loaded by a con-
tinuous distribution of weight:

w / unit length

5.3 Shear force and bending moment


A loaded beam can be considered as a sequence of connected, infinitesimally nar-
row slices. In transmitting vertical forces horizontally along the beam, each slice is
loaded in shear.
For example, consider the cantilever below:

The beam can be considered as a horizontal row of connected slices:

Each slice is subjected to a downward shear force of F on its left-hand face and,
in reaction, to an upward shear force of F on its right-hand face.
Now section the beam at a point B, distance x from the left-hand end:

5.3 Shear force and bending moment 59


C

A B

and draw regions to the left and right of B as free body diagrams:

A B 5

Bending of beams
For the free body AB:
Resolve vertically: F x = P
Resolve horizontally: H x = 0
Moments about B: M x = Px.
We see that each slice of a beam in static equilibrium must be loaded through
its surfaces by its neighbours in two ways:
1. A shear force created by the equal and opposite vertical loads from the
beam sections to its left and right
2. A bending moment created by the equal and opposite moments acting
on the beam sections to its left and right
Remember
A consistent sign convention is used to represent distributed load w, shear force F and
bending moment M on a beam element. All are positive as shown below:
w

F F
M M

5.4 Example: Shear force and bending moment


A simply supported beam 15 m long is loaded by 12 force units and 10 force
units at distances 3 and 6 units, respectively, from the left hand end. Calculate
the shear force at each point along the beam.
12 10

A B
3 3 6

5.4 Example: Shear force and bending moment 60


Remember
Sign convention:
w

F F
M M

1. Treat the entire beam as a free body to calculate the reaction forces.
Resolve vertically: RA + R = 12 + 10
B
5
Take moments about A: 12 × 3 + 10 × 6 = RB × 12

Bending of beams
hence RB = 8 and RA = 12 + 10 − 8 = 14.

2. Calculate shear force F within each interval of beam length:


a. Within 0 < x < 3:

=14

hence F + 14 = 0, F = − 14.
b. Within 3 < x < 6:

12

=14

hence F + 14 − 12 = 0, F = − 2.
c. Within 6 < x < 12:

12 10

=14

F + 14 − 12 − 10 = 0, F = 8.
3. Draw shear force diagram:

5.4 Example: Shear force and bending moment 61


+8
A –2 B
–14

4. Calculate bending moment M within each interval of beam length:


a. Within 0 < x < 3:

A 5

=14

Bending of beams
hence M + 14x = 0, M = − 14x.
b. Within 3 < x < 6:

12

=14

hence M + 14x − 12 x − 3 = 0, M = − 36 − 2x.


c. Within 6 < x < 12:

12 10

=14

M + 14x − 12 x − 3 − 10 x − 6 = 0, M = − 96 + 8x.
5. Draw a bending moment diagram:

BM
A
B

–42
–48

5.4.1 Example: SF and BM diagrams (1)


A simply-supported uniform beam of length 6 units is subjected to a point
load of 6 units at a point 1 m from the left-hand end.

5.4.1 Example: Shear force and bending moment 62


6

1 5

Draw shear force and bending moment diagrams.


1. Treat the entire beam as a free body to calculate the reaction forces.
5
6

Bending of beams
A B C

1 5

Resolve vertically: RA + RB = 6

Take moments about C: 6RA − 5 × 6 = 0

Hence RA = 5 and RB = 1.

2. Calculate shear force F within each interval of beam length.


Within 0 < x < 1: F + 5 = 0 hence F = − 5.
Within 1 < x < 6: F + 5 − 6 = 0 hence F = 1.
3. Draw shear force diagram.

SF

–5

4. Calculate bending moment M within each interval of beam length.


Within 0 < x < 1: M + 5x = 0 hence M = − 5x.
Within 1 < x < 6: M + 5 − 6 x − 1 = 0 hence M = x − 6.
5. Draw a bending moment diagram.

5.4 Example: Shear force and bending moment 63


BM

–5

5.4.2 Example: SF and BM diagrams (2)


A simply-supported uniform beam of length 10 units is subjected to point 5
loads of 12 units at 3 units at distances of 1 and 4 units from the left-hand end.

Bending of beams
12 3

1 3 2

Draw shear force and bending moment diagrams.


1. Treat the entire beam as a free body to calculate the reaction forces.

12 3

A B C D

1 3 2

Resolve vertically: RA + RB = 15

Take moments about C: 6RA − 5 × 12 − 2 × 3 = 0

Hence RA = 11 and RB = 4.

2. Calculate shear force F within each interval of beam length.


Within 0 < x < 1: F + 11 = 0 hence F = − 11.
Within 1 < x < 4: F + 11 − 12 = 0 hence F = 1.
Within 4 < x < 6: F + 11 − 12 − 3 = 0 hence F = 4
3. Draw shear force diagram.

5.4.2 Example: Shear force and bending moment 64


SF
14

–25

4. Calculate bending moment M within each interval of beam length.


Within 0 < x < 1: M + 11x = 0 hence M = − 11x.
5
Within 1 < x < 6: M + 11x − 12 x − 1 = 0 hence M = x − 12.
Within 4 < x < 6: M + 11x − 12 x − 1 − 3 x − 4 = 0 hence

Bending of beams
M = 4x − 24
5. Draw a bending moment diagram.

BM

–4

–11

5.4.3 Example: SF and BM diagrams (3)


A simply-supported uniform beam of length 10 units is subjected to point
loads of 12 units at 3 units at distances of 1 and 4 units from the left-hand end,
and a load of 24 units uniformly distributed between them.
12 24 total 3

1 3 2

Draw shear force and bending moment diagrams.


1. Treat the entire beam as a free body to calculate the reaction forces.

12 3
8 / unit length
A B C D

1 3 2

5.4.3 Example: Shear force and bending moment 65


Resolve vertically: RA + RD = 12 + 3 + 24
7
Take moments about C: 6RA − 5 × 12 − 2
× 24 − 2 × 3 = 0

Hence RA = 25 and RD = 14.

2. Calculate shear force F within each interval of beam length.


Within 0 < x < 1: F + 25 = 0 hence F = − 25.
Within 1 < x < 4: F + 25 − 12 − 8 x − 1 = 0 hence F = 8x − 21.
Within 4 < x < 6: F + 25 − 12 − 24 − 3 = 0 hence F = 14 5
3. Draw shear force diagram.

Bending of beams
SF
14

–25

4. Calculate bending moment M within each interval of beam length.


Within 0 < x < 1: M + 25x = 0 hence M = − 25x.
1
Within 1 < x < 6: M + 25x − 12 x − 1 − 8 x − 1 2
x − 1 = 0 hence
M = 4x2 − 21x − 8.
5
Within 4 < x < 6: M + 25x − 12 x − 1 − 3 × 8 x − 2
−3 x−4 =0
hence M = 14x − 84
5. Draw a bending moment diagram.

BM
0

–25

–40

5.5 Moment applied to a beam at a point


Shear force and bending moment diagrams reduce the beam to a straight line, to
which only normal loads can be applied. A moment can only be represented as
two equal and opposite, normal, point loads applied close together.

5.5 Moment applied to a beam at a point 66


The simply-supported beam shown below is drawn as having a real height. A
moment could be applied to it as shown if — for example — it was bolted to
another, vertical beam at this point.
A B

When loaded beams are reduced to free-body diagrams in order to model 5


them using shear force and bending moment diagrams, their depth is neglec-
ted. The effect of a bending moment can be seen by representing the moment

Bending of beams
as a force couple F0 separated by an arbitrarily small small distance e, where
F0 = eM0:

Resolving vertically: RA + RB = 0

Taking moments clockwise about A: M0 − R L = 0


B

M0 M0
Hence RB = L
and RA = − L
.

The value of e affects the shear force diagram only within a region of its own
size:
SF

A B

The bending moment diagram is:


BM

B
A

5.5 Moment applied to a beam at a point 67


5.6 Relating shear force and bending moment to
applied loads
Analysis of a uniform beam under general distributed and/or point loading leads
to analytical expressions for the distributions of shear force and bending moment
along it.
Consider the FBD of a small section AB of a uniform beam, where w is the dis-
tributed load per unit length:
5

Bending of beams
A B

1. Resolve vertically for equilibrium of section AB:


ⅆF
F + ⅆx δx − F = wδx

ⅆF
hence w = ⅆx
.

2. Take moments B:
ⅆM 1
M + ⅆx δx − M − Fδx − wδx × 2 δx = 0

ⅆM 1
so that ⅆx
− F − 2 wδx = 0

and as δx 0,
ⅆM
F = ⅆx .

We can immediately see that at local point of maximum or minimum bending


moment, the shear force is zero.

5.6.1 Example: Problem 5.1


A beam 6 m long, simply supported at each end, carries a distributed load
over the whole span. The loading distribution is represented by the equation

w x = ax2 + bx + c
where w is the intensity of load at a distance x along the beam and a, b and c
are constants. The intensity of load is zero at each end and has its maximum
value of 4 kNm–1 at mid-span.
Apply the differential equations of equilibrium relating w, F and M to obtain
the maximum values and distributions of shear force and bending moment.
[8 kN, 15 kNm]

5.6 Relating shear force and bending moment to applied loads 68


1. Treat the entire beam as a free body to calculate the reaction forces.

w x = 0 at x = 0 and x = 6: provides 2 equations


w x = 4 at x = 3: provides 1 equation. 5

Hence w x = − 49 x2 + 83 x.

Bending of beams
ⅆF
2. Apply w = ⅆx
to determine the shear force distribution along the beam.

Hence F(x) = ∫w(x)dx = ∫ − 49 x2 + 83 x dx = − 274 x3 + 86 x2 + C.


ⅆM
3. Apply F = ⅆx
to determine the bending moment distribution along the
beam.

M(x) = ∫F(x)dx = ∫ − 274 x3 + 43 x2 + C dx = − 271 x4 + 49 x3 + Cx


+D

4. Find sufficient boundary conditions to complete the solution.


Both ends are simply supported so that M = 0 at x = 0 and x = 6.

5.7 Notes

5.7 Notes 69
Bending of beams
5

70
5.7 Notes
Bending of beams
5

71
5.7 Notes
Chapter 6 Bending of beams: stress and
strain
Once the length, support conditions and applied loads for a beam have been ex-
pressed as shear force and bending moment distributions, stresses and strains can
be calculated at every point along it. To do so, the cross-sectional shape is ex-
pressed as a single, geometrical property: the second moment of area.
On successfully completing this module, you will be able to:
• Define bending stresses in beams
• List the assumptions made to derive the mathematical expressions for 6
stresses and strains in slender beams

Bending of beams: stress and strain


• Define neutral axis and second moment of area
• Recall mathematical expressions to locate the neutral axis and determine
the second moment of area for regular and arbitrary cross-sections
• Apply the parallel axis theorem to find the second moment of area for a
simple composite shape or non-symmetric cross section
• Determine stresses in beams characterised by regular and arbitrary cross-
sections
• Assess the potential failure of beams by relating the maximum stresses to
the material strength
• Find stresses in beams subjected to combined axial load and bending
moment.

6.1 Bending stress and strain


Deflections of a loaded beam are caused mainly by the distribution of bending
moment along it. At every point, a straight beam bends to a radius of curvature
which, by stress analysis, can be determined from the local bending moment.
Consider an elemental length AB along a beam between two planes A and B
normal to the axis. Under the applied bending moment M, the beam bends to
a sector of a circle whose radius of curvature is to be determined. The beam
can be of any cross-section.

Chapter 6 Bending of beams: stress and strain 72


Cross-section: Side view of beam:
A1 B1 A'1 B'1
A B
A' B'
A0 B0 A'0 B'0

A2 B2 A'2 B'2

Bending of beams: stress and strain


Assumptions:
• There are no axial loads or constraints on the beam.
• There are no out-of-plane loads or constraints on the beam.
• When the beam is bent, every plane cross-section remains plane and nor-
mal to the axis.
• The curvature R−1 remains small.
1. Determine the strain distribution as a function of distance y across the
section:
A′B′ − AB
The axial strain in AB is eAB = AB

This strain increases with radius, but the average strain must be zero if the
beam bears no axial load.
Important
There must be a neutral axis between the top and bottom of the beam at which the
strain is zero.

This axis passes through points A0 and B0, so that


AB = A0B0 = A′0B′0 = Rδθ.

Hence for the element shown at a height y from the neutral axis,

A′B′ = R + y δθ
and
R + y δθ − Rδθ y
eAB = Rδθ
= R.

2. Determine the stress distribution as a function of distance y across the


section:

6.1 Bending stress and strain 73


Since, as usual, we assume there to be no axial or out-of-plane loads on
the beam, axial stress is simply given by:
E
σAB = EeAB = R y.

RELATED LINKS
Animation: assumptions for bending analysis

6.1.1 Stress in beams of symmetrical cross-section


The axial stress in a bent beam varies linearly from bottom to top. If its cross-sec-
tion is symmetrical about a horizontal line midway between bottom and top, then
that line is the neutral axis and stresses can be calculated directly. 6
Consider a beam of rectangular cross-section of depth d and thickness b:

Bending of beams: stress and strain


Tension

A B

Compression

Side-view Cross-section Stress

1. Apply the condition that the beam carries no axial force:


The axial force is given by
+d/2
P= ∫ σ y b ⅆy
−d/2

and we know that the stress is given by σ y = αy where α = E/R.


Hence
1 2 +d/2
αb y = 0.
2 −d/2

2. Calculate the bending moment from the stress distribution:


The bending moment at any cross-section is obtained by integrating the
product of
• The force σ y bδy on every element AB and
• The moment arm, y, of that force.
Hence

6.1.1 Bending stress and strain 74


+d/2
1 3 +d/2 αb d3 d3
∫ σbydy = αb
3
y
−d/2
=
3 8
− −
8
.
−d/2

This must be equal to the local bending moment M, and since from Step 1
α = E/R:

E bd3
M= .
R 12
We have shown that the distribution of stress σ and the local radius of bend-
ing curvature R at any point along a rectangular beam where the moment is M
are given by:
12M σ
= y = R.
E 6
3
bd

Bending of beams: stress and strain


6.1.1.1 Example: bending stress in a rectangular beam
What is the maximum tensile stress in a 2 m long horizontal cantilever of negli-
gible weight, subjected to a vertical end load of 9 kN if it has a rectangular
cross-section 150 mm deep and 40 mm thick?

9 kN 40 mm
150 mm

2m

1. Calculate bending moment distribution.


At any cross-section x, M = 9x.
2. Identify point of maximum tensile stress.
The maximum bending moment is therefore at the beam root, x = 2 m,
where
Mmax = 9 × 2 kNm.

σ 12M
Since y
= ,
bd3

the maximum stress will be at ymax = 0.075 m.

3. Hence calculate maximum stress.


Mmaxymax
σmax =
bd3 /12

bd3 0.04 × 0.153


where = = 1.125 × 10−5 m4. Hence:
12 12

6.1.1.1 Stress in beams of symmetrical cross-section 75


Mmaxymax 9 × 103 × 2 × 0.075
σmax = = = 120 kNm.
bd3 /12 1.125 × 10−5

6.1.2 Stress in beams of arbitrary cross-section


For an arbitrary cross-section:
1. The beam width is not constant, but a function b y of vertical distance
from the neutral axis
2. The cross-section is no longer symmetrical about any horizontal axis
3. The neutral axis cannot therefore be identified at the outset.
6

Bending of beams: stress and strain


tension
A B

compression

Side view Cross-section Stress

Given conditions (1) and (2), our task is to find the location of the neutral axis.
1. Calculate the force transmitted through the cross-section for an arbitrary
choice of neutral axis.
The force transmitted through the elemental strip shown is σb y δy,
where σ = αy where σ = E/R.
Hence the total axial force is
+y1 +y1 +y1

∫ σb y ⅆy = ∫ αyb y ⅆy =α ∫ b y y ⅆy.
−y2 −y2 −y2

2. Determine the neutral axis position by setting the transmitted force to


zero.
Hence the neutral axis position is given by solving the equation
+y1

∫ b y y ⅆy = 0,
−y2

which gives a relationship between y1 and y2.

3. Calculate the bending moment for the same choice of neutral axis.
The moment due to the force calculated in Step 1 is σb y δy × y.
Hence the total bending moment is:

6.1.2 Bending stress and strain 76


+y1 +y1

∫ σb y ydy = α ∫ b y y2dy = α I
−y2 −y2

where I is the second moment of area of the beam cross-section about the
chosen neutral axis.

6.1.2.1 Example: stress in beams of arbitrary cross-section


The procedure for analysing a beam of arbitrary cross-section should work equally
well, of course, for symmetrical cross-sections in which the neutral axis position is
already known.
Show that for a rectangular cross section with depth d and constant thickness 6
b the neutral axis lies at the centre of the beam, and determine the second
moment of area.

Bending of beams: stress and strain


1. Apply Eq. () to the case b y = b:
+y1 +y1 +y1
y2 y2 y2
∫ bydy = b ∫ ydy =b
2
−y2
=b
2
1

2
2
=0
−y2 −y2

hence y1 = y2 = 12 d, defining the neutral axis.

2. Apply Eq. () to the case b y = b and y1 = y2 = 12 d:


+y1 d/2
σ y3 d/2
M=
σ
y ∫ 2 σ
by dy = b
y ∫ y2dy = b
y 3 −d/2
−y2 −d/2 x
3 3 3
σ d/2 −d/2 σ bd
= yb 3
− 3
= y 12

σ M bd3
Hence y
= I where I = 12 as derived previously.

6.2 Flexural rigidity of a beam


The flexural rigidity of a beam (also known as bending stiffness) is a measure of
the resistance of a beam to bending — one of its most important properties. It is
defined as the product of the elastic modulus of the beam material and the second
moment of area of its cross section.
When a moment M is applied to a beam whose second moment of area
around the neutral axis is I and whose material has a modulus E, it will bend to
a radius R where
M E
I
= R.

6.1.2.1 Stress in beams of arbitrary cross-section 77


The greater the radius of curvature, per unit bending moment, of an initally
straight beam the smaller displacements over a given length will be. Thus, EI
characterises the bending stiffness or flexural rigidity of the beam.

6.2.1 Second moment of area


The second moment of area of a beam, alongside the elastic modulus of its mate-
rial, determines the resistance of the beam to bending. It depends on the size and
shape of the beam cross section.
The second moment of area is defined as
y2

Ixx = ∫by2 ⅆy, 6


y1

Bending of beams: stress and strain


where the line y = 0 passes through the centroid and y1 and y2 are the maxi-
mum and minimum extent of the area.
The centroid of an area is identical to the centre of mass of a flat plate of the
same shape and of uniform thickness and density. Its position can be deter-
mined from the shape of the area alone.
In the table below, xx and yy are the orthogonal neutral axes which pass
through the centroid of the section.
Table 6.2-1 Second moments of area: simple shapes

Cross section Second moments of area

x bd3
x Ixx = 12

bd3
Iyy = 12

x x πa4
Ixx = Iyy = 4

6.2.1 Flexural rigidity of a beam 78


Cross section Second moments of area

b1 + b2 h3
Ixx = 36
x x b1 + b2 b2 2
1 + b1b2 + b2 h
Iyy = 36

6.2.1.1 Composite areas 6


Where the shape of a ‘composite’ area can be expressed as the sum or difference of

Bending of beams: stress and strain


component areas whose second moments of area are known for the same axes,
that of the composite area can be derived by adding or subtracting those of the
components.
Two common beam cross-sections which can be treated as composite areas
are a round tube, and a hollow rectangular section (a box section).

Round tube
y

x x

πa4
For the outer area: Iext = 4

πb4
For the inner area: Iint = 4

π
For the composite: Itube = 4
a4 − b4

6.2.1.1 Second moment of area 79


Box section
y

x x

y
6
bd3
For the outer area: Iext = 12

Bending of beams: stress and strain


ec3
For the inner area: Iint = 12

1
For the composite: Ibox = 12
bd3 − ec3 .

6.2.1.2 The parallel axis theorem


Where the shape of a ‘composite’ area can be expressed as the sum or difference of
component areas whose second moments of area are known for axes which are
are parallel but not the same, the axis of one can be shifted using the parallel axis
theorem.
The parallel axis theorem is used to derive the new second moment Iaa of an
area after its axis has been shifted from ‘xx’ by a distance a to another, parallel
axis ‘aa’:

Area A
x x

a a

For the original axis xx passing through the centroid of this area the moment
of area was
y2

Ixx = ∫b y y2 ⅆy
y1

which can be written as

Ixx = ∫ y2 ⅆA.
A

6.2.1.2 Second moment of area 80


The new second moment of area is given by

Iaa = ∫ y + a 2 ⅆA = ∫ y2 ⅆA + 2a∫ y ⅆA + a2∫ ⅆA.


A A A A

But because the original axis passed through the centroid,

2a ∫ y ⅆA = 0
A

by definition; hence

Iaa = Ixx + a2 ∫dA = Ixx + a2A.


A
6
6.2.1.2.1 Example: Parallel axis theorem (1)

Bending of beams: stress and strain


The second moment of area for an I-beam (RSJ) cross section can be analysed us-
ing the parallel axis theorem by two methods subtraction from the outer rectan-
gle, or addition of the component rectangles.
I-beams are designed for high flexural rigidity per unit weight, and are very
widely used in construction.
The cross-section shown has a width of b, a total height of d + 2t and a uni-
form thickness of t. Derive an expression for its second moment of area refer-
red to the neutral axis shown.

web
N A
flange

1. Subtraction method: the I-beam cross-section is treated as two rectangu-


lar side blocks removed from the outer cross-section block. The second
moments of areas of the side blocks are therefore subtracted from that of
the outer block.

6.2.1.2.1 The parallel axis theorem 81


1
b d + 2t 3 2
b − t d3
I = −2
Hence NA 12 12
1
= b d + 2t 3 − b − t d3
12
2. Addition method: the I-beam cross-section is constructed from rectangu-
lar flange blocks added to the web block. The second moments of area of
the flange blocks are therefore added to that of the web block after the
parallel axis theorem has been used to move them away from the NA by a
distance ± d + 12 t .

Bending of beams: stress and strain


td3 bt 3 d+t 2
Hence INA = +2 + bt .
12 12 2

Both methods lead to the same result.

6.2.1.3 Example: Second moment of a non-symmetric area


A T-beam provides one common example of a cross-section for which the neutral
axis position must first be found.
Find a general expression for the second moment of area for the cross-section
shown below.

1. Find location of neutral axis.


Let the location be a distance y from the base. Then
A1+ A2 y = A1y1 + A2y2

d t
td + td y = td × + td × d +
2 2

6.2.1.3 Second moment of area 82


d t
2y = 2 + d + 2 x

hence
3d t
y= + .
4 4
2. Determine second moment of area around the axis now located.
Using the parallel axis theorem

I=I x + A1 y − y1 2 +I x + A2 y − y 2 2
a1a′1 a2a′2

hence
6
td3 3d t d 2 dt 3 t 3d t 2
I= + dt + − + + dt d + − + .
12 4 4 2 12 2 4 4

Bending of beams: stress and strain


6.3 Example: Problem 5.3
A beam ABCD is simply supported at A and C and carries concentrated loads
of 40 kN at B and 50 kN at D, together with a load of 60 kN uniformly distrib-
uted, w, from B to D. AB = BC = 4 m, CD = 2m. Sketch the bending moment
diagram and quote maximum values. [+120 kNm, –60 kNm]
ln order to reduce the greatest bending moment, a jack is placed under the
beam at D. What upward jacking force, P, will reduce the greatest bending
moment in the beam to its smallest possible value? [20 kN]

1.

6.3 Example: Problem 5.3 83


Bending of beams: stress and strain
6

84
6.3 Example: Problem 5.3
6.4 Example: stress in an I-beam
A beam is to bear a uniformly distributed load of 120 kNm–1 over the 5 m span
between its simply supported ends. If the maximum allowable stress for the
material is 160 MPa determine whether an I-beam 528.8 mm in height, 6
200 mm wide and having a second moment of area of 404×10–6 m4 will be

Bending of beams: stress and strain


strong enough.
I
120 kN m–1
528.8 mm
5m

If not, determine the length and thickness of flange reinforcement plates nee-
ded to make it so.
Flange plates are fixed to the outer surfaces of the upper and lower flanges of
an I-beam, and of equal width.
σ M
1. Since y
= I and all variables are specified except the bending moment
M, determine the maximum allowable bending moment:
ymax 160 × 106 × 404 × 10−6
Mallow = σmax = = 245 kNm.
I 528.8/2 × 10−3
2. Treat entire beam as FBD to determine reactions:
By symmetry, RA = RB.

Resolve vertically: 2RA = 5 × 120 hence RA = 300 kN.

3. Determine bending moment distribution:


Take moments around a point x from the left-hand end:
1
M + 300x − 120x × 2 x = 0 hence M = − 300x − 60x2 kNm.

4. Compare maximum bending moment to maximum allowable:


Absolute maximum bending moment is at x = 2.5 m, where

Mmax = − 300 × 2.5 + 60 × 2.52 = − 375 kNm.

6.4 Example: stress in an I-beam 85


5. If the beam will fail anywhere, determine the length over which the maxi-
mum allowable moment must be increased.
The region within which the bending moment exceeds 245 kNm is given
by

M = − 300x − 60x2 = − 245


hence

60x2 − 300x + 245 = 0

300 ± 3002 − 4 × 60 × 245


x= so that x = 2.5 ± 1.47 m.
120
6
6. Determine the necessary length and thickness of flange reinforcement
plates.

Bending of beams: stress and strain


The necessary plate length will be 2 × 1.47 = 2.94 m. Let its thickness be
t.
σmaxI
Using the parallel axis theorem: Mallow = where
ymax

0.2t 3 D+t 2
I = I0 + 2 + 0.2t
12 2

1
ymax = 2 D + t , and

D is the original beam depth (528.8 mm).


Try different values of t…
t = 0 mm: Mallow = 245 kNm

t = 10 mm: Mallow = 405 kNm

t = 8 mm: Mallow = 375 kNm, OK.

6.5 Bars under combined axial and bending


loads
For a bar made of linear elastic material the stress, strain and extension due to ax-
ial load strains and the bending stress and strain due to bending moments can
simply be added together.

6.5 Bars under combined axial and bending loads 86


1 + 2 = 3

Bending Axial tension

6
Note

Bending of beams: stress and strain


The neutral axis for bending of the cross section is determined only by the cross-sectional
shape: it does not move if additional axial load moves the line of zero stress.

There are two practical cases in which combined bending and tension can
arise by a single force:
1. A bar, of finite thickness, end-loaded at a point which does not lie on the
centroid:

Axial stresses along the bar are the sum of those due to
• An axial force θ, plus
• Bending by a clockwise moment M = Fe applied at the end.
2. A bar which is end-loaded at an angle θ to its axis:

Axial stresses along the bar are the sum of those due to
• An axial force Fcos θ, plus
• Bending by a normal end load Fsin θ.

RELATED LINKS
Animation: bending moment, stress and strain

6.5.1 Example: combined axial and bending loads

6.5.1 Bars under combined axial and bending loads 87


Determine the smallest value of end load W which will induce compressive
stress anywhere in a round cantilever bar of diameter 60 mm subjected to ax-
ial tensile force of 10 kN?

1m
10 KN

60 mm

1. Determine axial tensile stress due to 10 kN end load:


10 × 103
σa = = 3.537 MPa.
π 0.03 2 6
2. Determine maximum compressive stress due to bending load W :

Bending of beams: stress and strain


=

σ M
=
y I
Bending moment M at distance x from support
M=W 1−x
hence Mmax = W × 1 Nm at support.

π 0.06 4
I= = 6.362 × 10−7 m4.
64
Hence
Mmax W
σmax = ± ymax = ± × 0.03 Pa
I 6.362 × 10−7
and for this to become equal to σa,

W 3.537 × 106 × 6.362 × 10−7


× 0.03 = = 75 N.
6.362 × 10−7 0.03

This result illustrates that for a typical bar geometry, stresses (and strains) due
to bending are much higher than those developed by stretching. If you want
to break a bar you don’t stretch it: you bend it against a support.

6.6 Notes

6.6 Notes 88
Bending of beams: stress and strain
6

89
6.6 Notes
Bending of beams: stress and strain
6

90
6.6 Notes
Chapter 7 Bending of beams: deflections
Beam analysis allows the radius of curvature to be calculated at every point along
a beam of known material, geometry and loading conditions. The final step is to
integrate this curvature distribution to yield deflections at every point; in statically
indeterminate cases, however, these deflections may themselves influence load-
ing.
On successfully completing this module, you will be able to:
• Define bending deflections and recall the moment-curvature relationship
• Apply the moment-curvature relationship to find slopes and deflections
at any point along the beam
• Solve standard beam problems and recall expressions for significant de- 7
flection values values for simple beams

Bending of beams: deflections


• Use Macaulay’s method to solve for deflections in statically determinate
beams
• Apply superposition principles to find deflections in simple beams
• Use Macaulay’s Method and superposition principles to solve for deflec-
tions in statically indeterminate beams.

7.1 Bending deflections


When a beam is loaded, it deflects — partly due to shear strains and partly due to
axial strains set up by the bending moment. Unless the beam is very short and
deep, bending deflections greatly exceed shear deflections, which can therefore be
ignored.
Given a description of…
• A beam geometry (i.e. length and cross-section)
• The modulus E of its material
• The positions and types (simple, built-in) of its supports and
• The position and distribution of applied loads,
axial stresses at any point x along the beam and at any point y across it can be
determined using the bending equation:
σ x, y M x E
= = .
y I R x
The third term in this equation gives the radius of beam curvature due to load-
ing, and can be used to determine the profile of deflection profilev x .
The curvature at a point for any function is given by
3/2
1 ⅆ2v ⅆv 2
= / 1 +
R ⅆx2 ⅆx

Chapter 7 Bending of beams: deflections 91


ⅆv
but deflections and slopes are small, and so ⅆx
≪ 1 and can be ignored to
give

1 ⅆ2v
=
R ⅆx2
and hence the moment-curvature relationship for a beam

ⅆ2v
EI = M.
ⅆx2

RELATED LINKS
Missouri University of Science and Technology MecMovies: see Ch 10, Beam Deflections

7.1.1 Example: Bending deflections (1) 7

Bending of beams: deflections


End-loaded cantilever

A X B
!

1. Analyse free-body diagram of the entire system.


Resolve vertically: V A = P

Moments about A: MA = PL.

2. Consider FBD for beam cut at an arbitrary point X:

A X

F and M are shear force and bending moment, respectively, at X.


Take moments about X (to avoid F appearing)
M = M A − V Ax = PL − Px = P L − x .

3. Apply the moment-curvature relationship:

ⅆ2v
EI = M = PL − Px.
ⅆx2
4. Integrate twice to obtain the displacement profile:

ⅆv x2
EI = PLx − P +A
ⅆx 2

7.1.1 Bending deflections 92


x2 x3
EIv x = PL −P + Ax + B
2 6
where A and B are arbitrary constants of integration to be determined by
boundary conditions.
5. Apply boundary conditions to determine constants.
a. At x = 0, deflection v = 0 and therefore B = 0.
ⅆv
b. At x = 0, slope ⅆx
= 0 and therefore A = 0.

Thus the deflection of the cantilever under this loading is given by


1 x2 x3
v x = PL −P .
EI 2 6

The end deflection is 7


2 3 3
1 L L PL
PL −P =
EI 2 6 3EI

Bending of beams: deflections


Note the strong dependence of end deflection on L.
The end slope is

ⅆv 1 L2 PL2
= PL2 − P = .
ⅆx EI 2 2EI

7.1.2 Example: Bending deflections (2)


Uniformly-loaded cantilever

A X B

1. Analyse free-body diagram of the entire system.


Resolve vertically: V A = wL

Moments about A: MA = wL × 12 L.

2. Consider FBD for beam cut at an arbitrary point X:

A /unit length
X

F and M are shear force and bending moment, respectively, at X.


Take moments about X (to avoid F appearing)

7.1.2 Bending deflections 93


1 1 1 1
M = MA − V Ax + 2 wx2 = 2 wL2 − wLx + 2 wx2 = 2 w L − x 2.

3. Apply the moment-curvature relationship:

ⅆ2v 1
EI = M = w L − x 2.
ⅆx 2 2

4. Integrate twice to obtain the displacement profile:


ⅆv 1
EI = − w L−x 2+A
ⅆx 6
1
EIv x = w L − x 4 + Ax + B
24
where A and B are arbitrary constants of integration to be determined by
boundary conditions.
7
5. Apply boundary conditions to determine constants.

Bending of beams: deflections


1
a. At x = 0, deflection v = 0 and therefore B = − 24
wL4.

ⅆv 1
b. At x = 0, slope ⅆx
= 0 and therefore A = 6 wL3.

Thus the deflection of the cantilever under this loading is given by


1 L−x 4 L3 L4
v x = w + wx −w .
EI 24 6 24

The end deflection is


1 L3 L4 wL4
wL −w =
EI 6 24 8EI

Note the strong dependence of end deflection on L.


The end slope is

ⅆv wL3
= .
ⅆx 6EI

7.1.3 Example: Bending deflections (3)


Simply supported beam with uniformly distributed load.

A B

1. Analyse free-body diagram of the entire system.

Resolve vertically (and accounting for symmetry): V A = V B = 12 wL

2. Consider FBD for beam cut at an arbitrary point X:

7.1.3 Bending deflections 94


A
X

F and M are shear force and bending moment, respectively, at X.


Take moments about X (to avoid F appearing)
1 1 1 1
M = − V Ax + 2 wx2 = − 2 wLx + 2 wx2 = 2 wx x − L .

3. Apply the moment-curvature relationship:

ⅆ2v w 2
EI =M= x − Lx .
ⅆx 2 2

4. Integrate twice to obtain the displacement profile: 7


ⅆv w x3 x2
EI = −L +A

Bending of beams: deflections


ⅆx 2 3 2

w x4 x3
EIv x = −L + Ax + B
2 12 6

where A and B are arbitrary constants of integration to be determined by


boundary conditions.
5. Apply boundary conditions to determine constants.
a. At x = 0, deflection v = 0 and therefore B = 0.
1
b. At x = L, deflection v = 0 and therefore A = 24
wL3.

Thus the deflection of the cantilever under this loading is given by


wx x3 x2 L3
v x = −L + .
6EI 4 2 4

The deflection at the mid-point (x = 12 L) is


wL L3 L2 L3 5wL4
−L + = .
12EI 32 8 4 384EI

7.1.4 Example: Bending deflections (4)


Simply supported beam with point load:

1. Analyse free-body diagram of the entire system.

7.1.4 Bending deflections 95


aP
Moments about A: V B = L

bP
Resolve vertically: V A + V B = P hence V A = L
.

2. Consider FBD for beam cut at an arbitrary point X:

A
X

F and M are shear force and bending moment, respectively, at X.


a. 0 < x < a:
b
7
M1 = − L Px

Bending of beams: deflections


b. a < x < L:
b
M2 = P x − a − L Px.

3. Apply the moment-curvature relationship:


a. 0 < x < a:

ⅆ 2v 1 bP
EI = − x
ⅆx 2 L

b. a < x < L:

ⅆ 2v 2 bP
EI =P x−a − x
ⅆx 2 L

4. Integrate twice to obtain the displacement profile:


a. 0 < x < a:
ⅆv1 bP 2
EI = − x + A1
ⅆx 2L
bP 3
EIv1 x = − x + A1x + B1
6L
b. a < x < L:
ⅆv2 P bP 2
EI = x−a 2− x + A2
ⅆx 2 2L
P bP 3
EIv2 x = x−a 3− x + A 2 x + B2
6 6L
where A1, A2, B1, B2 are arbitrary constants of integration to be deter-
mined by boundary conditions.
5. Apply boundary conditions to determine constants.

7.1 Bending deflections 96


a. At x = 0, deflection v1 = 0 and therefore B1 = 0.

b. At x = L, deflection v2 = 0 and therefore


P bP 3 P bP 2
0= L−a 3− L + A2L + B2 = b3 − L + A 2L + B 2
6 6L 6 6
bP 2
hence b − L2 + A2L + B2 = 0.
6
c. From continuity conditions at x = a (i.e. there is no break or ‘kink’ in
the beam) at x = a, deflection v1 = v2, and
ⅆv1 ⅆv2
d. At x = a, slope ⅆx
= ⅆx .

bP 3 P bP 3
− a + A1a + 0 = a−a 3− a + A2a + B2
6L 6 6L 7
hence A1a = A2a + B2

Bending of beams: deflections


bP 2 P bP 2
and − a + A1 = a−a 2− a + A2
2L 2 2L
hence A1 = A2, B2 = 0
bP 2
b − L2 + A2L = 0
6
bP 2
A1 = A2 = L − b2
6L
Thus the deflection of the cantilever under this loading is given by
bP 2
EIv1 x = L x − b2x − x3
6L
P bP 2
EIv2 x = x−a 3+ L x − b2x − x3 .
6 6L

7.2 Macaulay’s method


Macaulay’s method is a mathematical technique used to integrate the beam mo-
ment-curvature relationship through a point at which the load changes discontin-
uously. It avoids the need to match deflection and slope values at such points.
This method uses angle brackets to enclose functions which need to be
‘switched off’ when their value is negative. For example:
0 x≤a
f x = x−a n=
x−a n x>a
f(x)

a x

7.2 Macaulay’s method 97


We can integrate these ‘step functions’ w.r.t. x in the same way:

∫0 ⅆx x≤a
∫ x−a n
ⅆx = .
∫ x − a n ⅆx x>a
const. x≤a
= x−a n+1
+ const. x>a
n+1

7.2.1 Example: Macaulay’s method (1)


Point load applied to a simply supported beam:

7
A B

Bending of beams: deflections


1. Determine the bending moment equation in Macaulay form for a point x
which lies beyond the right-most load.

A
X

b
M = − LP x + P x − a

2. Apply the moment-curvature relationship:

ⅆ2v bP
EI =P x−a − x.
ⅆx 2 L

3. Integrate twice to obtain the displacement profile:


ⅆv P bP 2
EI = x−a 2− x +A
ⅆx 2 2L
P bP 3
EIv x = x−a 3− x + Ax + B
6 6L
4. Apply boundary conditions to determine constants:
a. At x = 0, deflection v = 0 and therefore B = 0.
b. At x = L, deflection v = 0 and therefore
P bP 2 P 3 bP 2
A= − L−a 3+ L = − b + L .
6L 6 6L 6
Substitution into the deflection equation reduces this to the result found by
direct integration:

7.2.1 Macaulay’s method 98


bP 2
EIv1 x = L x − b2x − x3 for 0 < x < a
6L
P bP 2
EIv2 x = x−a 3+ L x − b2x − x3 for a < x < L.
6 6L

7.2.2 Macaulay’s method for distributed load


Point load applied to a simply supported beam with distributed load:

A B

1. Analyse free body diagram for the whole system: 7


Taking moments clockwise about A:

Bending of beams: deflections


1 a
RB = w L − a 2 + 2L

Resolve vertically: A

RA + RB = w L − a X
1 a
hence R A = w L − a 2
− 2L .

2. Determine the bending moment equation in Macaulay form for a point x


which lies beyond the right-most load.

X
1 a 1
M = − w L − a 2 − 2L x + 2 wP x − a 2.

3. Apply the moment-curvature relationship


4. Integrate twice to obtain the displacement profile
5. Apply boundary conditions to determine constants
…as before.

7.2.3 Distributed loads over a restricted length


Macaulay's method treats loads at different points along a beam as events which
occur as integration proceeds rightwards from the left-hand. A load distributed
over only part of the beam is treated by following its appearance at its left-hand
limit by the appearance of an equal and opposite distribution at the right-hand
limit.

7.2.2 Macaulay’s method 99


In the case illustrated below, the load is w per unit length between x = a and
x = b.

A B

1. Analyse free-body diagram of the entire system.


Hence RA, RB.

2. Represent the removal of the distributed load w per unit length from
x = b to x = L as the addition of a distributed load −w per unit length.
7

Bending of beams: deflections


A B

= +
A B A B

3. Consider FBD for beam cut at an arbitrary point X:

Hence
1 1
M = − RAx + 2 w x − a 2 − 2 w x − b 2

4. Apply the moment-curvature relationship:

ⅆ2v w w
EI = M = − R Ax + x−a 2− x − b 2.
ⅆx 2 2 2

5. Integrate twice to obtain the displacement profile:

ⅆv x2 w w
EI = − RA + x−a 3− x−b 3+A
ⅆx 2 6 6

7.2 Macaulay’s method 100


x3 w w
EIv x = − R A + x−a 4− x − b 4 + Ax + B
6 24 24
where A and B are arbitrary constants of integration to be determined by
boundary conditions.
6. Apply boundary conditions to determine constants.
a. At x = 0, deflection v = 0
b. At x = L, deflection v = 0
etc..

7.3 Beam bending solutions by superposition


The bending profile of a loaded beam is determined by solving a differential equa-
tion whose terms are proportional to applied loads. As a result, solutions can be
7
superposed.

Bending of beams: deflections


7.3.1 Example: end deflection of a cantilever
Determine the end deflection of the cantilever shown.

1. Determine (e.g. from Handbook of Data and Formulae) end deflection and
slope for a cantilever of length L1.

wL4
1
End deflection δ1 = 8EI

wL3
1
End slope θ = 6EI
.

2. Determine δ2 by geometry.

The beam in L1 < x < L2 is not loaded and therefore has no curvature at
any point.

7.3.2 Example: end loaded cantilever with uniformly


distributed load
Determine the end deflection of the cantilever shown.

7.3 Beam bending solutions by superposition 101


1. Determine the end deflection from distributed load alone.
wL4
δw = 8EI .

2. Determine the end deflection from point load alone.


PL3
δP = 3EI .

3. Determine total deflection as sum of those to individual loads.


7
4 3
wL PL
δw + δP = 8EI + 3EI .

Bending of beams: deflections


7.3.3 Example: end-supported cantilever
The free end of the cantilever shown is supported at the same height as its
root. Determine the maximum deflection of the beam.

1. Treat the entire system as a free body to determine reaction forces.


Resolve vertically: wL = V A + V B

Take moments about A: 12 wL2 − V BL − MA = 0.

Attention
2 equations and 3 unknowns! The reactions cannot be found by static equilibrium
alone.

This problem is statically indeterminate and can only be solved by consid-


ering deflections.
2. Replace the support at B by a force sufficient to cancel out deflection due
to w using known solutions.
The solution to the problem shown below is known:.

7.3.3 Beam bending solutions by superposition 102


wL4 PL3
End deflection δw + δP = 8EI
+ 3EI .

Hence for zero end deflection, P = − 38 wL and V B = − P = 38 wL.

Equilibrium gave two equations to determine the three unknowns and com-
patibility has given a third. Hence V A and MA can now be determined.

7.4 Statically indeterminate beams


If a loaded beam has more constraints than are needed to support the applied
loads, reaction forces cannot be determined by static equilibrium alone. We need
to know the flexural rigidity of the beam and to calculate deflections in order to
know all the forces needed to restrain them.
The beam shown below needs only simple supports to support the load. Be- 7
cause it is built-in at both ends there are also reaction moments, and these

Bending of beams: deflections


make the problem statically indeterminate.

Resolve vertically: P = RA + RB.

Take moments about A: aP − LRB − MA + MB = 0.

To impose the additional condition that the beam slope at both ends is zero,
the displacement profile v x must be determined from the bending moment
distribution
M = − RAx + M A + P x − a .

RELATED LINKS
Missouri University of Science and Technology MecMovies: see Ch 11, Indeterminate beams

7.5 Notes

7.4 Statically indeterminate beams 103


Bending of beams: deflections
7

104
7.5 Notes
Bending of beams: deflections
7

105
7.5 Notes
Chapter 8 Torsion of bars
Torsion is the twisting of an object by an applied moment. It is important in engi-
neering both as an unwanted result of the need to transmit a torque along a shaft,
and as the source of useful compliance in a coil spring.
On successfully completing this module, you will be able to:
• Define torsion, shear stress and shear strain
• Determine stresses and deformations in thin-walled circular and non-cir-
cular shafts under torsion
• Determine stresses and deformations in solid circular and hollow shafts
under torsion
• Define torsional stiffness
• Analyse circular shafts of varying cross-section subjected to torsional load 8
• Recall formulae for the study of stress and deformation in non-circular

Torsion of bars
shafts under torsion.

8.1 What is torsion?


Examples of engineering systems for which torsional analysis is needed include
gearbox shafts, torsion bar suspensions and coil springs.
1. A torque is a moment applied about the axis of a shaft, usually to transmit
power as in a gearbox. The moment is applied to a gearbox transmission
shaft as a force acting on the radius of a gear fixed to the shaft. A reaction
force acts on the radius of a second gear fixed to the same shaft.
Clutch case
Outer Inner
transmission transmission
shaft shaft

FROM
ENGINE

TO
DIFFERENTIAL

Clutch 1
Clutch 2

Chapter 8 Torsion of bars 106


1 2 3

Equilibrium of the system implies: T 1 − T 2 + T 3 = 0

The torque on the shaft 1-2 is: T 1 (equilibrium of shaft 1)

The torque on the shaft 2-3 is: T 3 = T 2 − T 1 (equilibrium of shaft 1).

Note
8
In this case torsion of each shaft is unwanted, and must be minimised.

Torsion of bars
2. Torsion bar suspension for vehicles.

3. Torsion of the wire or bar which a torsion spring is made of allows the
spring to change in length when a force is applied to it.

Note
In both cases 2 and 3, torsion is required — and must be predicted by engineering analy-
sis.

RELATED LINKS
Video on torsion bar suspension
Online textbook on torsion bar suspension

8.1 What is torsion? 107


8.2 Analysis of torsional stress and strain
Torque generates shear stresses, and torsional ‘twist’ generates shear strains. For a
material of given elastic modulus (i.e. stress/strain ratio), the relationship between
torque and twist depends on the cross section of the bar.
Analysis to obtain the relationship between torque, twist and shear stress for a
torsion bar is very much like that to obtain the relationship between bending
moment, radius of curvature and direct stress for a beam.
Shear strain γ (measured in radians) and for small γ ,
τ
γ = G,

where G is the shear modulus. It was shown in that E = 2G 1 + ν .

8.2.1 Torsion of a thin-walled cylinder


If the wall of a long cylinder or tube is sufficiently thin and its cross section is uni-
8
form, stress and strain under torsion can be analysed very simply.

Torsion of bars
Consider a cylinder or tube wall whose thickness t is small compared to its
mean radius Rm

P Q

R S

Assume that:
• Assume the angle of twist is small, so that tan ϕ = Rθ
.
L

• The wall is so thin that shear strain γ can be considered independent of


radius.
• There is axial symmetry, so γ does not vary around the circumference.
• Each cross-section rotates by an angle α without changing shape
1. Calculate the shear strain for the element PQRS, which was rectangular
before torque was applied:

Q'
P Q P'

S'
R S R'

Shear strain γ = tan ϕ ≈ ϕ

8.2 Analysis of torsional stress and strain 108


Roθ
But ϕ = L

Roθ
so that shear strain γ ≈ L
.

2. Calculate the moment (element of torque T ) needed to shear this ele-


ment of the wall:
The shear stress on the element is τ = Gγ , so that the force on the ele-
ment is τtRm ⅆα.

Hence ⅆT = τtRm ⅆα Rm.

3. Integrate circumferentially to obtain the total torque:



T= ∫0 τR2 2
mt ⅆα = τ2πRmt

T
so that τ = . 8
2πR2
mt

4. Use the assumption that t ≪ Rm so that Rm ≈ Ro ≈ Ri ≈ R.

Torsion of bars
τ T Gθ Gϕ
Hence R
= = L = R .
3
2πR t

8.2.2 Solid and hollow circular shafts


To analyse and design solid and hollow circular shafts subjected to torsion, we
need to know the angle of twist for a given torque and the distribution of shear
stress and strain through the material.
We can analyse the torsion of thick-walled and solid circular bars using the re-
sult for a cylinder whose wall thickness is infinitesimal, so that shear stress and
strain can be assumed uniform through it.

8.2.2 Analysis of torsional stress and strain 109


1. Recall the shear stress distribution and torque/twist relationship for an in-
finitely thin-walled tube of radius r and thickness t:
θ
τ = Gr L , and 8
θ
T = 2πr 3t × G L .

Torsion of bars
2. Integrate from the axis r = 0 to the outside radius Ro to obtain the total
torque for a solid shaft consisting of coaxial, thin-walled tubes with thick-
ness ⅆr under equal twist:
Ro

T=G
θ
L ∫2πr3dr = G θL J
0

where
Ro
π 4 R0 πD4
J= ∫ 2πr 3dr =
2
[r ]0 =
32
.
0

3. For a hollow shaft, change the lower limit of the integral for J from zero to
the internal radius Ri:
R0
Ro
J= ∫2πr3dr = 2π [r4]R i
=
π
D4 − D4
32 o i .
0

Result: again we have


τ T Gθ
= =
r J L

πD4
where in this case J = . Compare this with the bending equation
32
σ M E
= = .
y I R

8.2 Analysis of torsional stress and strain 110


Note that in 8.2.1 Torsion of a thin-walled cylinder on page 108, we effectively
approximated J by 2πR3mt. The table below compares this approximation to
the exact result obtained by integration from an infinitesimal thickness.

t/R0 Japprox / J

0.1 0.997
0.2 0.987
0.3 0.97

8.2.3 Example: Torsion bar design (1)


An aluminium alloy tube of constant cross-section is to satisfy:
• Length L = 0.5 m
• Maximum thickness t = 50 MPa
8
• Maximum torque T = 1.0 kNm

Torsion of bars
• Torsional stiffness at least 30 kNm rad–1.
If the outer diameter must be D0 = 50 mm, find the largest allowable inner
diameter, Di.
τ T Gθ
We know that = = , and maximum shear stress at outer diameter
r J L
τmax T max
= .
D0 /2 J

1. Calculate the J value required to satisfy torque and maximum shear stress
limits.
T max D0 /2 1000 × 0.025
J= = = 0.5 × 10−6 m4
τmax 50 × 106

2. Hence determine the necessary inner diameter Di.


π
J= D4 − D4
i = 0.5 × 10
−6
, D4o − D4i = 5.09 × 10−6,
32 o
D4 4
i = 0.05 − 5.09 × 10
−6
= 1.16 × 10−6 so that Di = 32.8 mm.

3. Check whether this Di gives us the required torsional stiffness (K T) of at


least 30 kNm rad–1:
τ T Gθ
Re-arrange = = to give the torsional stiffness condition in terms
r J L
TL T L L
of Di J = = = KT .
Gθ θ G G
Hence for Al alloy of shear modulus G = 26.5 GPa,
0.5
J = 30000 × 9
= 5.66 × 10−7 m4.
26.5 × 10

8.2.3 Analysis of torsional stress and strain 111


π
Hence determine the J = D4 − D4
i = 5.66 × 10
−7
,
32 0
D4 4
0 − Di = 5.76 × 10
−6
and with
D4 4
i = 0.05 − 5.76 × 10
−6
= 4.9 × 10−7 hence with Do = 50 mm
Di = 26.5 mm.

Therefore, for maximum T ≥ 1 kNm: Di ≤ 32.5 mm; but for torsional stiffness
to exceed 30 kNm: Di ≤ 26.5
mm and we must choose the thicker tube on the second condition.

8.2.4 Example: Torsion bar design (2)


An aluminium alloy tube of constant cross-section is to satisfy:
1. Length L = 0.5 m
2. Maximum thickness t = 50 MPa 8
3. Maximum torque T = 1.0 kNm

Torsion of bars
4. Torsional stiffness at least 30 kNm rad–1.
If the internal diameter Di = 50 mm, find the smallest allowable outer diame-
ter, Do.
τ T Gθ
We know that = = , and maximum shear stress at outer diameter
r J L
τmax T max
= .
D0 /2 J

1. Apply the torsional stiffness (K T) required (at least 30 kNm rad–1): re-ar-
τ T Gθ
range = = to give the torsional stiffness condition
r J L
TL T L L
J= = = KT .
Gθ θ G G
2. Hence for Al alloy of shear modulus G = 26.5 GPa,
0.5
J = 30000 × 9
= 5.66 × 10−7 m4.
26.5 × 10
π
3. Hence J = D4 − D4
i = 5.66 × 10
−7
, D40 − D4i = 5.76 × 10−6 and
32 0
with D40 = 6.25 × 10−6 − 5.76 × 10−6 = 1.2 × 10−5 m4 hence
D0 = 58.8 mm is required.

4. Check whether this is strong enough to give the required maximum


T = 1.0 kNm. Maximum shear stress is at outer diameter, hence
τmax T max
= must satisfy maximum shear stress condition τmax = 50
D0 /2 J
MPa.

8.2.4 Analysis of torsional stress and strain 112


50 × 106 1000
5. = gives
D0 /2 π D0 − D4
4
i /32
32 × 1000 × D0
D4 4
0 − Di = 6
= 1.0186 × 10−4D0, hence hence
100π × 10
D4
0 = 6.25 × 10
−6
+ 1.0186 × 10−4D0.

6. To solve this for D0, rewrite as


1/4
D0 = 6.25 × 10−6 + 1.0186 × 10−4D0 and solve by iteration.

Hence D0 = 59.2 mm. To exceed a torsional stiffness of 30 kNm requires


D0 ≥ 58.8 mm, but to ensure a maximum torque of T = 1 kNm requires
D0 = 59.2 mm.

The thicker tube must be chosen, hence D0 = 59.2 mm.

8
8.3 Shafts of varying diameter

Torsion of bars
Torsion of a shaft which consists of several sections with different constant diame-
ters can be analysed by treating each section separately, matching torque and ro-
tation at each junction.
Consider the case illustrated, which consists of just two sections.

A B C

We know that for each section,


T Gθ
J
= L

T2 G2 J2
θ2
= L = K T2 = torsional stiffness of section 2.
2

Hence
TL1 T
θ1 = G J = K and
1 1 T1

TL2 T
θ2 = G J = K
1 2 T2

1. For section 1, between A and B, the rotation θ1 at B relative to that at A


T1 G1θ1
(considered fixed) is J1
= L .
1

T1 G1 J1
2. The torsional stiffness of section 1 is defined as θ1
= L = K T1
1

8.3 Shafts of varying diameter 113


3. For section 2, between B and C, the rotation θ2 at C relative to that at B is
T2 G2θ2
J2
= L .
2

T2 G2 J2
4. The torsional stiffness of section 2 is defined as θ2
= L = K T2
2

T T
Hence the total rotation of C relative to A is θ = θ1 + θ2 = K T1
+K .
T2

T
But θ
≡ K T, the torsional stiffness of the entire shaft.

1 1 1
Hence KT
=K +K .
T1 T2

8.3.1 Example: torsion of a stepped shaft


A torsion bar consists of a solid shaft of length L and diameter D. It is required
to double the torsional stiffness by shrinking onto one end of this shaft a collar 8
of the same material and of diameter 2D. How long should the collar be?

Torsion of bars
a

2D D

1 2

T GJ G πD4
1. The original torsional stiffness was = =
θ L L 32

2GπD4
2. Therefore the required stiffness is .
32L
G π
3. The stiffness of section 1 after the collar is fitted will be K T1 = 2D 4
a 32
G π 4
4. Stiffness of section 2 after the collar is fitted will be K T2 = D
L − a 32
1 1 1
Hence, since = + ,
KT KT KT
1 2

32L 32a 32 L − a
4
= 4
+
2GπD Gπ 2D GπD4
L a
= + L−a
2 16
8L = a + 16 L − a
15a = 8L
8
and the required collar length isa = 15
L.

8.3.1 Shafts of varying diameter 114


8.4 Bars of non-circular cross-section
When a bar of non-axisymmetric cross-section is twisted, its cross-sections do not
remain flat: they warp. This makes the analysis of torsion more complicated, but
its results are no more difficult to use.
The additional shear stresses set up by warpage make the section stiffer.

8
Table 8.4-1 Analytical results for torsion of square, triangular and elliptical cross-sections

Torsion of bars
Shape of cross-section τmax θ
Square 4.81
T
7.10
TL
a3 a4G

Eqilateral triangle 20
T
46
TL
a3 a4G

Ellipse 2 T a2 + b2 TL
π a3
πa3b3 G

8.4.1 Example: torsion of a non-circular bar


The 6061-T6 aluminum shaft shown has cross-sectional area in the shape of
equilateral triangle.

8.4 Bars of non-circular cross-section 115


Determine the greatest torque T that can be applied to the end of this shaft if
the maximum allowable shear stress is τallow = 56 MPa and the maximum al-
lowable twist is θallow = 0.02 rad. Take the shear modulus of 6061-T6 to be
GAl = 26 GPa.

How much torque can be applied to a shaft of circular cross-section made 8


from same amount of material?
1. By inspection, the torque at any internal cross-section along the shaft axis

Torsion of bars
is also T .
20T
2. Use τallow = from Table 8.4-1 Analytical results for torsion of square,
a3
triangular and elliptical cross-sections on page 115 to apply the stress
limit.
This condition requires T = 179.2 Nm.
TL
3. Use θallow = 46 from Table 8.4-1 Analytical results for torsion of
a4GAl
square, triangular and elliptical cross-sections on page 115 to apply the
twist limit.
This condition requires T = 24.12 Nm
4. Comparison these results to find the lowest torque limit:
Torque is thereore limited by the angle of twist, to T = 24.12 Nm.
5. For a circular cross-section, first calculate the radius of the equivalent
cross-section:
For Acircle = Atriangle, R = 14.850 mm.

6. Use results from 8.2.2 Solid and hollow circular shafts on page 109 to im-
plement the stress limit:
TR
From τallow = J
, this requires T = 288.06 Nm.

7. Similarly, impose the twist limit:


TL
Hence from θallow = JGAl
:T = 33.10 Nm.

8.4 Bars of non-circular cross-section 116


Hence in this case the torque is limited by angle of twist. Comparing both re-
sults, we can see that a shaft of circular cross-section can support 37% more
torque than a triangular one.

RELATED LINKS
Missouri University of Science and Technology MecMovies: see Ch 6, Torsion

8.5 Analysis: torsion of thin-walled, non-circular


tubes
Thin-walled tubes of noncircular shape are used to construct lightweight frame-
works such as those in aircraft.
Thin-walled tubes of noncircular shape are used to construct lightweight
frameworks such as those used in aircraft. This section will analyse such shafts
with a closed cross-section. As walls are thin, we assume stress is uniformly dis- 8
tributed across the thickness of the tube.

Torsion of bars
T
Average shear stress: τavg = 2tAm
, where

τavg is the average shear stress acting over thickness of tube (changes with
location if t changes)
T is the resultant internal torque at cross-section
t is the thickness of the tube where τavg is to be determined

Am is the mean area enclosed within the boundary of the centreline of the
tube thickness

8.5 Analysis: torsion of thin-walled, non-circular tubes 117


The angle of twist can be determined using energy methods.
TL ⅆs
θ= ∮ t
4A2
mG

For constant thickness,this is the perimeter of Am.


8
8.6 Notes

Torsion of bars

8.6 Notes 118


Torsion of bars
8

119
8.6 Notes
Torsion of bars
8

120
8.6 Notes

Anda mungkin juga menyukai