Sie sind auf Seite 1von 117

Natural Sciences Tripos Part IB

MATERIALS SCIENCE
Course D Mechanics of Materials and Structures

Dr. P.D. Bristowe
Lent Term 2013-14
IB
Name .................................... College .........................
Leonhard Euler
1707-1783
Alan Griffith
1893-1963
Armand Considere
1841-1914
Charles Inglis
1875-1952
Galileo Galilei
1564-1642
Otto Mohr
1835-1918
Simeon Poisson
1781-1840
Thomas Young
1773-1829
Waloddi Weibull
1887-1979
Woldemar Voigt
1850-1919
IB Course D Mechanics of Materials and Structures Lent 2014 DHi

Contents

Introduction ..................................................................................................................................... 1

1 Tensors...................................................................................................................................... 3
1.1 Introduction to Tensors ....................................................................................................... 4
1.2 Stress Components .............................................................................................................. 5
1.3 Transformation of Axes....................................................................................................... 7
1.4 Transformation (Rotation) of the Reference Axes of a Vector ........................................... 8
1.5 Einstein Summation Convention......................................................................................... 9

2 Transformation of Stresses and Mohrs Circle .................................................................. 10
2.1 Transformation of Second Rank Tensors.......................................................................... 10
2.2 Principal Stresses and the Secular Equation...................................................................... 11
2.3 Introduction to Mohrs Circle for Plane Stress ................................................................. 13

3 Mohrs Circle in Practice...................................................................................................... 16
3.1 Construction of Mohrs Circle .......................................................................................... 16
3.2 Simple Examples............................................................................................................... 17
3.3 Example of a Thin-Walled Tube under Applied Torque and Axial Load......................... 18
3.4 Verification using the Secular Equation............................................................................ 21

4 Representation of Strain ....................................................................................................... 23
4.1 Introduction ....................................................................................................................... 23
4.2 The Relative Displacement Tensor ................................................................................... 23
4.3 The Strain Tensor .............................................................................................................. 24
4.4 Some Simple Example Cases ............................................................................................ 26
4.5 Hydrostatic and Deviatoric Components of the Strain Tensor.......................................... 27
4.6 Field and Matter Tensors................................................................................................... 27
4.7 Stiffness and Compliance Tensors .................................................................................... 28
4.8 Relationship to Elastic Constants ...................................................................................... 28
4.9 Poisson Ratio..................................................................................................................... 29

5 Engineering Stresses and Strains and Plasticity................................................................. 31
5.1 Engineering Shear Strains and the Shear Modulus ........................................................... 31
5.2 Bulk Modulus .................................................................................................................... 32
5.3 Relationships between Elastic Constants .......................................................................... 32
5.4 Large Uniaxial Tensile Strains and Plastic Deformation .................................................. 33
5.5 Plastic Instability and Considres Construction............................................................... 34

6 - Beam Bending......................................................................................................................... 38
6.1 Introduction to Beam Bending .......................................................................................... 38
6.2 Beam Curvatures ............................................................................................................... 38
6.3 Bending Moments and Second Moments of Area............................................................. 39
6.4 Beam Stiffness................................................................................................................... 41
6.5 Maximising the Beam Stiffness ........................................................................................ 42
IB Course D Mechanics of Materials and Structures Lent 2014 DHii


7 Examples of Beam Bending and Torsion .............................................................................. 44
7.1 Beam Deflections .............................................................................................................. 44
7.2 The Cantilever Beam......................................................................................................... 44
7.3 Symmetrical 3-Point Bending ........................................................................................... 45
7.4 Introduction to Torsion...................................................................................................... 46
7.5 Torsion of Thin-Walled Circular Tubes............................................................................ 47
7.6 Torsion of Solid Cylindrical Bars ..................................................................................... 48
7.7 Torsion of Thick-Walled Circular Tubes .......................................................................... 50

8 Failure in Compression: Elastic Buckling of Columns ........................................................ 52
8.1 Crushing ............................................................................................................................ 52
8.2 Elastic (Euler) Buckling .................................................................................................... 54
8.2 Freely-hinged columns ...................................................................................................... 54

9 Failure in Compression: the Effect of Constraints and Column Shape ............................. 60
9.1 The effect of End-Constraint ............................................................................................. 60
9.2 The effect of Aspect Ratio................................................................................................. 63
9.3 The effect of Cross-Sectional Shape ................................................................................. 64
9.4 Local Buckling .................................................................................................................. 66
9.5 The effect of Internal Pressure .......................................................................................... 67

10 Brittle Fracture in Tension................................................................................................... 68
10.1 The Ideal Strength ............................................................................................................. 68
10.2 The Effect of Cracks Stress Considerations ................................................................... 69
10.3 Energy Considerations ...................................................................................................... 71

11 Stable Crack Growth............................................................................................................. 76
11.1 Fracture Toughness ........................................................................................................... 76
11.2 The Wedging Geometry .................................................................................................... 77
11.3 Energy Considerations ...................................................................................................... 77
11.4 Crack Healing.................................................................................................................... 80
11.5 The Nature of Cracking..................................................................................................... 81

12 Coping with a Scatter in Stength.......................................................................................... 83
12.1 Introduction ....................................................................................................................... 83
12.2 Weibulls Approach .......................................................................................................... 84
12.3 Using Weibulls Method ................................................................................................... 85
12.4 Practical Solutions............................................................................................................. 87
12.5 Toughening........................................................................................................................ 88

Appendix Summary of Materials Response in Compression and Tension........................... 91
Nomenclature................................................................................................................................. 94
Glossary.......................................................................................................................................... 95

IB Course D Mechanics of Materials and Structures Lent 2014 DH1

Course D: Materials and Structures
Synopsis
Any structure, man-made or biological, and whatever its primary function, must be able to
withstand imposed loads. Building on ideas from the Part IA course which covered the mechanical
behaviour of materials, this course aims to give a grounding in the interplay between the loads
acting on a structure, its shape (design), relevant material properties and the resultant distributions
of stress and strain, and associated deflections. This determines whether the structure is likely to
fulfil its function, which normally requires avoiding excessive elastic deflections or plastic
deformation, and also ensuring that fracture does not occur. In order to do this, its essential for the
manipulation of stresses and strains to be well understood, so the course starts with an introduction
to their representation as second rank tensors. Later parts of the course cover the behaviour of
structures, particularly beams under bending, torsion and axial compression, and the ways in which
they can fail due to buckling, crushing, fast fracture or fatigue.
Lecture Details (12 lectures)
Tensors (Lecture 1)
Introduction to Tensors. The Rank of a Tensor. Stresses as Second Rank Tensors. Transformation
of Axes. Transformation (Rotation) of the Reference Axes of a Vector. The Einstein Summation
Convention.
Transformation of Stresses and Mohrs Circle (Lecture 2)
Transformation of Second Rank Tensors. Principal Stresses and the Secular Equation. Introduction
to Mohrs Circle for Plane Stress.
Mohrs Circle in Practice (Lecture 3)
Construction of Mohrs Circle. Examples of the use of Mohrs Circle. Determination of the
Principal Stresses and their Orientation for a Thin-Walled Tube under Torsion and Axial Tension.
Representation of Strain (Lecture 4)
Deformation of a Body. The Relative Displacement Tensor. Removal of the Rigid Body Rotation
Component and Identification of the Strain Tensor. Hydrostatic and Deviatoric Components of the
Strain Tensor. Stress-Strain Relationships and the Engineering Constants. The Stiffness and
Compliance Tensors and their Relationship to the Elastic Constants. The Poisson Ratio.
Engineering Stresses and Strains and Plasticity (Lecture 5)
Engineering Shear Strains and the Shear Modulus. The Bulk Modulus and Inter-relationships
between the Elastic Constants. Large Uniaxial Strains and Plastic Deformation. True Stress and
Strain. Plastic Instability and Considres Construction.
Beam Bending and Torsion (Lecture 6)
Introduction. Beam Curvatures. Bending Moments and Second Moments of Area. Stress
Distributions in Beams. Beam Stiffness. Maximising the Beam Stiffness.
Examples of Beam Bending and Torsion (Lecture 7)
Beam Deflections. The Cantilever Beam. Symmetrical 3-Point Bending. Torques and Resultant
Twist Angles. Torsion of Thin-Walled Tubes, Thick-Walled Tubes and Solid Cylindrical Bars.
IB Course D Mechanics of Materials and Structures Lent 2014 DH2

Failure in Compression (Lecture 8)
Crushing of Dense and Porous Bodies. Elastic (Euler) Buckling. Freely-Hinged Columns. Energy
Balance Considerations.
Failure in Compression: The Effect of Column Shape (Lecture 9)
The Effect of End-Constraint, Column Slenderness and Cross-Sectional Shape. Local Buckling.
The Effect of Internal Pressure on Buckling of a Hollow Cylinder.
Brittle Fracture in Tension (Lecture 10)
The Ideal Strength. The Effect of Cracks Stress Considerations. Stress Concentrations at Atomic
and Macroscopic Scales. Energy Considerations. Contributions to the Total Energy of the System.
Changes in Mechanical and Surface Energies.
Stable Crack Growth (Lecture 11)
Fracture Toughness. The Wedging Geometry. Energy Considerations. Crack Healing. The Nature
of Cracking. The Importance of Equilibrium. Catastrophic Failure.
Coping with a Scatter in Strength (Lecture 12)
Weibulls Approach. Practical Solutions. Toughening. Crack Bridging. R-curves. Fatigue. S-N
Curves.
Booklist
Several textbooks are relevant to the course, although unfortunately there are sometimes
differences in the conventions and symbolism which they use. The standard book on tensors is
Physical Properties of Crystals Their Representation by Tensors and Matrices, by JF Nye,
although its coverage goes considerably beyond what is required here. The more structural aspects
of the course, such as beam bending, torsion, buckling etc are well described in Mechanics of
Materials by JM Gere and BJ Goodno. Mechanical Metallurgy, by GE Dieter is sound on basic
mechanics and also deals with fracture and fatigue. Perhaps the most relevant book for background
interest reading is Structures by JE Gordon. This omits mathematical treatments, but it does cover
many of the ideas involved and it gives many entertaining examples of their application. It, and the
earlier book of Gordons called The New Science of Strong Materials, should be read by all
Materials Scientists.
Web-based Resources
Most of the material associated with the course (handouts, question sheets, practical scripts etc)
can be viewed on the web (www.msm.cam.ac.uk/teaching) and also downloaded. This includes
model answers to question sheets, which are released after the work concerned should have been
completed. In addition to this text-based material, DoITPoMS resources are also available
(www.doitpoms.ac.uk). The following TLPs are directly relevant to this course:
Stress Analysis and Mohrs Circle Bending and Torsion of Beams
Brittle Fracture Tensors
Finally, there is a CamTools site for the course (https://camtools.cam.ac.uk/ - named MatSci Crs-D
: NSTIB 13_14). Lecture summaries, and pointers to additional material, will be provided there.

IB Course D Mechanics of Materials and Structures Lent 2014 DH3

1 Tensors
Revision from Pt 1A Course D (Mechanical Behavior of Materials)
Normal (!) and shear (") stresses are defined as stresses induced by forces acting either normal
or parallel to the sectional area of a body to which they are applied:



The response of the body to the applied stress is characterised by the strain. The normal strain (!) is
the relative change in length in the direction parallel to the original length. The shear strain (") is
the angular deformation arising from a shear displacement in a particular plane:




IB Course D Mechanics of Materials and Structures Lent 2014 DH4

For elastic deformation, the normal stress and strain and the shear stress and strain are linearly
proportional:
!
" = E# and $ = G% where E and G are the Youngs and shear moduli respectively.

There are two important points:
1) In general, an applied set of forces generates both normal and shear stresses (and strains).
These will, however, differ when different planes within the body are considered. It is therefore
important to clearly specify the reference frame (set of axes) when stating stresses and strains.
2) Both stresses and strains are characterised by a magnitude and two directions. For stress,
these are the direction of the force and the direction of the normal to the face on which it acts. For
strain, these are the direction of the change in length and the direction of the reference length.
Therefore they are NOT vectors and cannot be resolved into components. Characterisation using
tensors is required.

1.1 Introduction to Tensors
The word tensor comes from the latin for extension

, and tensors are strongly associated


with mechanics, although they are also widely used in other branches of science and engineering.
The utility of tensors is mainly concerned with treating differences in the response or characteristics
of a material in different directions within it. The usage is particularly valuable when treating
anisotropic materials and of course most (crystalline) materials are in practice anisotropic to at
least some degree. However, even for isotropic materials, tensor analysis is necessary, or at least
very helpful, when treating most types of mechanical loading. In the present course, only isotropic
materials will be considered.

The term is also used in Robert Hookes famous expression Ut tensio, sic vis As the extension, so the force. This
is the origin of the simple form of Hookes law, relating the normal stress on a material to the resultant normal strain.
Hookes experiments actually involved springs, which deform in a relatively complex manner, with the material being
subjected to pure torsion see lecture 7. Fortunately, there is still a linear relationship between load and extension.

Woldemar Voigt (1850-1919)
was the first to use the term
tensor in its modern context to
describe mechanical stress.
IB Course D Mechanics of Materials and Structures Lent 2014 DH5

A tensor is an n-dimensional array of values, where n is the rank of the tensor. The simplest
type is thus a tensor of zeroth rank, which is a scalar - ie a single numerical value. Properties like
temperature and density are scalars. They are not associated with any particular direction in the
material concerned, and the variable does not require any associated index (suffix

).





A first rank tensor is a vector. This is a 1-D array of values. There are normally 3 values in the
array, each corresponding to one of 3 (orthogonal) directions. Each value has a single suffix,
specifying the direction concerned. These suffices are commonly numerical {1, 2 & 3}, although
sometimes other nomenclature, such as {x, y & z} or {r, # & z}, may be used. Force and velocity
are examples of vectors. The components of a vector can thus be written down in a form such as
(1.1)
in which each of the suffices {1, 2 & 3} refers to a specific direction, such as {x, y & z}.






1.2 Stress Components
There are other variables, such as stress, for which each component requires the specification of
two directions, rather than one, so that two suffices are needed. These are called second rank
tensors.





The term suffix is in common use for these indices, although they are employed as subscripts.
IB Course D Mechanics of Materials and Structures Lent 2014 DH6

In the case of stress, these two suffices specify, firstly, the direction in which a force is being
applied and, secondly, the normal of the plane on which the force is acting

. Stress is thus an
example of a second rank tensor and the components can be written down in the form of a 2-D
array of values
!
ij
=
!
11
!
12
!
13
!
21
!
22
!
23
!
31
!
32
!
33
"
#
$
$
$
%
&
'
'
'
(1.2)
When the suffices i and j are the same, the force acts parallel to the plane normal, ie the component
concerned is a normal stress. When they are different, it is a shear stress (and sometimes the
symbol $ is used instead of % for such components).



Some of the stresses that could act on a body are depicted in Fig.1.1. Provided that the body is
in static equilibrium, which is commonly assumed, then the normal forces acting on opposite faces
so as to generate a normal stress (eg %
33
in Fig.1.1) must be equal in magnitude and anti-parallel in
direction. (If this were not the case, then the body would translate.) For shear stresses, a further
condition applies. Not only must the two forces generating the %
23
stress (see Fig.1.1) be equal and
opposite, but the magnitude of the %
32
stress must be equal to that of the %
23
stress. (If this were not
the case, then the body would rotate.) Shear stresses thus act in pairs. This applies to all shear
stresses, so that %
ij
= %
ji
, and the tensor represented in Eqn.(1.2) must be symmetrical.

Fig.1.1 Illustration of the nomenclature of stresses acting on a body

This is the convention used by Nye. Its also adopted in many other textbooks, and in the Part II courses on Tensors
and Composite Materials. However, the reverse convention ie that the first suffix gives the normal of the plane on
which the force is acting and the second gives the direction of the force can be employed and indeed is used by Dieter
and by Gere and Goodno. Provided a convention is used consistently, problems are unlikely to arise.
IB Course D Mechanics of Materials and Structures Lent 2014 DH7








1.3 Transformation of Axes
It follows from this that there are just 6 independent components in a general stress state 3
normal stresses and 3 shear stresses. The magnitude of these will, of course, depend on the
directions of the axes that have been chosen to provide the frame of reference. However, its clear
that the state of stress itself will be unaffected if we chose an alternative frame of reference. Any
tensor can be transformed so as to refer to a new set of axes, provided the orientation of these with
respect to the original set is specified.
Axes transformations are important in the mechanics of materials for various reasons. For
example we may know the stresses acting on 1-2 planes but really need to know the stresses acting
on planes oriented at some angle & to the 1-axis because these are slip planes in the material. Or
perhaps we need to know the planes on which the normal stresses take their maximal value since
these planes may fail by tensile cracking if the stresses exceed the tensile stress of the material.
It turns out that any stress state can be expressed in terms of normal stresses only (ie all shear
stresses are zero), provided that the appropriate set of axes is chosen. Since its often very helpful
to be able to express a stress state in terms of this unique set of normal stresses, its important to be
clear about the procedure for transforming tensors in this way. We will do this first for a vector
(force) and then move on to treat stress in the second lecture.






IB Course D Mechanics of Materials and Structures Lent 2014 DH8

1.4 Transformation (Rotation) of the Reference Axes of a Vector
Consider a vector, ie a force, F (= [0, F
2
, F
3
]), with components which are referred to the axis set
{1, 2, 3}. A specific reorientation of this set of axes is now introduced, namely a clockwise rotation
by an angle # about the 1-axis, to create a new axis set {1, 2, 3} see Fig.1.2.

Fig.1.2 Rotation, in the 2-3 plane, of the axes forming the reference frame for a vector F
In this case, the new 1-axis coincides with the old 1-axis, but the 2- and 3-axes have been
rotated with respect to the 2- and 3-axes. The values of F
2
and F
3
are found by resolving the
components F
2
and F
3
onto the 2 and 3 axes and adding these resolved components together:

F
2 '
= F
2
cos 2'! 2 ( ) + F
3
cos 2'! 3 ( )
F
3'
= F
2
cos 3'! 2 ( ) + F
3
cos 3'! 3 ( )
(1.3)
where the symbolism (x
^
y) represents the angle between the x and y axes. In terms of the angle #,
these two equations can be written as

F
2 '
= F
2
cos !" ( ) + F
3
cos !90 !" ( ) = cF
2
! sF
3
F
3'
= F
2
cos 90 !" ( ) + F
3
cos !" ( ) = sF
2
+ cF
3
(1.4)
in which c and s represent cos # and sin # respectively.
Clearly these cosines of the angles between new and old axes are central to such transformations.
They are commonly termed direction cosines and are represented by a
ij
, which is conventionally the
cosine of the angle between the new i direction (= i) and the old j direction.

Of course, the rationale can be extended to cases in which all three axes have been reoriented,
leading to the following set of equations

F
1'
= a
11
F
1
+ a
12
F
2
+ a
13
F
3
F
2 '
= a
21
F
1
+ a
22
F
2
+ a
23
F
3
F
3'
= a
31
F
1
+ a
32
F
2
+ a
33
F
3
(1.5)
IB Course D Mechanics of Materials and Structures Lent 2014 DH9

It can be seen that the direction cosines form a matrix and this set of equations can be written more
compactly in matrix form:

!
F
1'
F
2'
F
3'
"
#
$
$
$
%
&
'
'
'
= T
[ ]
F
1
F
2
F
3
"
#
$
$
$
%
&
'
'
'
(1.6)
in which the transform matrix is given by
!
T
[ ]
=
a
11
a
12
a
13
a
21
a
22
a
23
a
31
a
32
a
33
"
#
$
$
$
%
&
'
'
'
(1.7)








1.5 Einstein Summation Convention
Sets of equations such as Eq.(1.6) can be written even more concisely by using the Einstein
summation convention. This states that, when a suffix occurs twice in the same term, then this
indicates that summation should be carried out with respect to that term. For example, in the
equation
F
i '
= a
ij
F
j
(1.8)
j is a dummy suffix, which is to be summed (from 1 to 3). The i suffix, on the other hand, is a free
suffix, which can be given any chosen value. For example, Eqn.(1.8) could be used to create the
equation
F
1'
= a
11
F
1
+ a
12
F
2
+ a
13
F
3
(1.9)
and also the two other equations, corresponding to i being equal to 2 or 3.





IB Course D Mechanics of Materials and Structures Lent 2014 DH10

2 Transformation of Stresses and Mohrs Circle
2.1 Transformation of Second Rank Tensors
Extension of the treatment presented in the previous section to the case of second rank tensors,
such as stress, follows quite logically. However, for a stress, we are concerned, not just with
resolving three single components into new directions (eg F
1
, F
2
and F
3
in the case of a force), but
rather we need to take account of the fact that both the force and the area on which it is acting will
change when we refer them to different axes. In other words, the operation we carried out with
respect to a single suffix in the case of force, F
i
, has to be implemented with respect to two suffices
for a stress, !
ij
. Each stress component thus needs to be multiplied by two direction cosines, rather
than one.
Example of a single crystal under tensile load (see Pt IA course):










IB Course D Mechanics of Materials and Structures Lent 2014 DH11

For the case of a stress, the set of equations corresponding to Eqn.(1.8) can be written
! '
ij
= a
ik
a
jl
!
kl
(2.1)
in which the prime is now applied to the symbol, rather than the individual suffices, to denote the
transformed version. Eqn.(2.1) can be expanded into 9 equations, each giving one of the 9
components of the stress tensor when referred to the new set of axes (although only 6 are required,
since the stress tensor is always symmetrical). Both k and l are dummy suffices in this equation
(since they are repeated), while i and j are free suffices. The first equation of the set, corresponding
to i=1 and j=1, can thus be expanded to

! '
11
= a
11
a
11
!
11
+ a
11
a
12
!
12
+ a
11
a
13
!
13
+a
12
a
11
!
21
+ a
12
a
12
!
22
+ a
12
a
13
!
23
+a
13
a
11
!
31
+ a
13
a
12
!
32
+ a
13
a
13
!
33
(2.2)
while that corresponding to i=3 and j=2 is given by

! '
32
= a
31
a
21
!
11
+ a
31
a
22
!
12
+ a
31
a
23
!
13
+a
32
a
21
!
21
+ a
32
a
22
!
22
+ a
32
a
23
!
23
+a
33
a
21
!
31
+ a
33
a
22
!
32
+ a
33
a
23
!
33
(2.3)






2.2 Principal Stresses and the Secular Equation
One of the main motivations for treating stresses within this rather cumbersome mathematical
framework is that it facilitates identification of the Principal Stresses. These are the normal
stresses acting on the Principal Planes, which are the planes on which there are no shear stresses.
(For the mathematically-inclined, these principal stresses are the eigenvalues of the stress tensor.)
Any general state of stress can thus be transformed such that it can be expressed in the form
!
ij
=
!
1
0 0
0 !
2
0
0 0 !
3
"
#
$
$
$
%
&
'
'
'
(2.4)
with the single suffix being commonly used to denote a principal stress. (Its important to
recognise, however, that these are still second rank tensors, and should thus, strictly speaking,
always have two suffices.) Obtaining these principal stresses, for a general 3-D stress state,
requires diagonalising the stress tensor. First a reminder about determinants:
IB Course D Mechanics of Materials and Structures Lent 2014 DH12










Second rank stress tensors are diagonalised by solving the determinant equation |%
ij
- 'I| = 0
where ' is a parameter and I is a unit matrix

. The solutions for the principal stresses are thus found


from

!
11
" # !
12
!
13
!
21
!
22
" # !
23
!
31
!
32
!
33
" #
= 0 (2.5)
This is a cubic equation in ', the roots of which ('
1
, '
2
, and '
3
) are the principal stresses. It is often
termed the secular equation.






Expanding the terms in Eqn.(2.5) leads to
!
3
" I
1
!
2
+ I
2
! " I
3
= 0 (2.6)
in which the coefficients (termed the invariants, since they do not vary as the axes are changed) are
given by

I
1
= !
11
+ !
22
+ !
33
I
2
= !
11
!
22
+ !
22
!
33
+ !
33
!
11
" !
12
2
" !
23
2
" !
31
2
I
3
= !
11
!
22
!
33
+ 2!
12
!
23
!
31
" !
11
!
23
2
" !
22
!
13
2
" !
33
!
12
2
(2.7)
Solving (2.6) is straightforward if it factorises or if relations between the roots are known, otherwise
numerical methods are needed. However the problem is simplified if one of the principal planes is
known since the analysis becomes two dimensional as shown in the next section. A useful TLP on
tensors elaborates on this further: http://www.doitpoms.ac.uk/tlplib/tensors/index.php

Proof of this is beyond the scope of the course, but in fact this is a fairly routine procedure in matrix algebra.
IB Course D Mechanics of Materials and Structures Lent 2014 DH13

2.3 Mohrs Circle for plane stress
Its fairly straightforward, if a little cumbersome, to find the principal stresses, and the
orientation of the (normals of the) principal planes, for a 3-D stress state specified with respect to an
arbitrary set of axes. It just requires a cubic equation to be solved. However, in practice it is
common to know one principal plane (direction), but to be interested in finding the principal
directions within that plane, or in establishing how the normal and shear stresses vary with
direction in that plane. Provided the plane concerned is a principal one, this problem reduces to
solving a quadratic equation, rather than a cubic. It can therefore be tackled via a geometrical
construction, equivalent to using some simple trigonometry. This is the basis of Mohrs circle,
which was proposed in 1892 by Christian Otto Mohr, a German civil engineer who became a
professor of mechanics in Dresden.






Provided the 1-2 plane is a principal plane, the appropriate version of Eqn.(2.5) is

!
11
" # !
12
0
!
21
!
22
" # 0
0 0 !
33
" #
= 0 (2.8)

and one principal stress is clearly %
33
(= %
3
). Since %
12
= %
21
, the secular equation reduces to
!
"
33
# $
( )
"
11
# $
( )
"
22
# $
( )
#"
12
2
[ ]
= 0
!
"#
2
$ %
11
+%
22
( )
# + %
11
%
22
$%
12
2
( )
= 0 (2.9)

which has the solution
Christian Otto Mohr (1835-1918)
developed the graphical method for
analysing plane stress known as Mohrs
Circle.
IB Course D Mechanics of Materials and Structures Lent 2014 DH14

!
1
, !
2
=
!
11
+ !
22
2
"
#
$
%
&
'

!
11
( !
22
2
"
#
$
%
&
'
2
+ !
12
2
(2.10)

The form of the solution suggests a possible graphical interpretation involving a circle in which
the first term is the circles centre and the second term is its radius. The circle is plotted on a graph
with the normal stress along the abscissa and the shear stress along the ordinate.








To determine how the normal and shear stresses vary with direction in the 1-2 plane we start
with the tensor in diagonalised form and then transform it to give %'
11
, %'
22
and %'
12
, for a specified
orientation
!
"'
11
"'
12
0
"'
12
"'
22
0
0 0 "
3

#
$
%
%
%
&
'
(
(
(
= T
[ ]
"
1
0 0
0 "
2
0
0 0 "
3
#
$
%
%
%
&
'
(
(
(
(2.11)
Since we are rotating about a principal axis, the form of Eqn.(2.1) applicable in this case reduces to
the following

! '
11
= a
11
a
11
!
1
+ a
12
a
12
!
2
+ a
13
a
13
!
3
! '
22
= a
21
a
21
!
1
+ a
22
a
22
!
2
+ a
23
a
23
!
3
! '
12
= a
11
a
21
!
1
+ a
12
a
22
!
2
+ a
13
a
23
!
3
(2.12)
The set of direction cosines applicable here (assuming an anti-clockwise rotation) can be written as
!
T
[ ]
=
a
11
a
12
a
13
a
21
a
22
a
23
a
31
a
32
a
33
"
#
$
$
$
%
&
'
'
'
=
cos( sin( 0
)sin( cos( 0
0 0 1
"
#
$
$
$
%
&
'
'
'
(2.13)
so it follows that these stresses are given by
!
" #
11
= cos
2
$ #
1
+ sin
2
$ #
2
" #
22
= sin
2
$ #
1
+ cos
2
$ #
2
" #
12
= %sin$ cos$ #
1
+ sin$ cos$ #
2
(2.14)
IB Course D Mechanics of Materials and Structures Lent 2014 DH15

These equations can also be written in a form involving 2#, rather than #
!
" #
11
=
#
1
+#
2
2
$
%
&
'
(
) +
#
1
*#
2
2
$
%
&
'
(
) cos2+
" #
22
=
#
1
+#
2
2
$
%
&
'
(
) *
#
1
*#
2
2
$
%
&
'
(
) cos2+
" #
12
= *
#
1
*#
2
2
$
%
&
'
(
) sin2+
(2.15)
These equations can be solved using the circle construction described above. The normal and
shear stresses acting on a plane rotated by an angle # from that on which %
1
acts are given by the
coordinates of a point rotated by 2# around the circumference of a circle centred at the mean of the
two principal stresses, and with a radius equal to half their difference. This is illustrated in
Fig.2.1(b). This construction provides a convenient method of calculating the stresses acting on
particular planes, establishing principal stresses and their orientations, finding the planes on which
peak shear stresses

operate etc.



Fig.2.1 Mohrs circle construction for a principal plane (1-2 plane), showing (a) principal
stresses in the plane (!
1
& !
2
, for a case in which !
1
is tensile and !
2
compressive) and
(b) corresponding Mohrs circle, giving, in addition to the principal stresses, the
normal and shear stresses acting on a plane rotated about the 3-axis by an angle # from
that on which !
1
acts.

Care is needed with the signs of shear stresses, since there is more than one possible convention in use. However, this
is not actually a point of any real concern, since the sign of a shear stress is solely an issue of convention, and has no
physical significance (unlike the sign of a normal stress).
IB Course D Mechanics of Materials and Structures Lent 2014 DH16

3 Mohrs Circle in Practice
3.1 Construction of Mohrs Circle
Mohrs Circle is a useful graphical representation of the plane stress transformations given in the
previous section (eqn. 2.15). It allows visualisation and determination of principal stresses,
maximum shear stresses and the variation of normal and shear stresses on arbitrary planes. Given
values of %
11
, %
22
and %
12
, for example, the principal stresses %
1
and %
2
in the 1-2 principal plane
may be found using the following procedure:
(a) Construct a graph with normal stress along the abscissa and shear stress along the ordinate,
using the same scale on both axes. Use the convention that positive shear stress points downwards.
Along the abscissa normal tensile stresses are positive and point to the right. Normal compressive
stresses are negative and point to the left.








(b) Plot the points (%
11
, -%
12
) and (%
22
, %
12
) and connect them with a straight line. Where the
straight line crosses the abscissa locates the centre of the circle. Label the angle 2& between the
straight line (the circle diameter) and the abscissa.
(c) Draw the circle centred at ((%
11
+ %
22
)/2, 0).
(d) The principal stresses are located at the points where the circle intercepts the abscissa. The
maximum shear stresses are located at the top and bottom of the circle.
(e) The angle of rotation of the diameter (2&) is twice the angle of rotation of the axes in real
space.
(f) The procedure works for any symmetrical second rank tensor and therefore also for strain.
(g) The procedure can be used to represent a three-dimensional stress state transformed into its
principal stresses. The Mohrs circles for the three principal planes are superimposed on the same
graph.

IB Course D Mechanics of Materials and Structures Lent 2014 DH17

3.2 Simple Examples
(a) Uniaxial tension %
11
= %
1
; %
22
= %
12
= 0

!
"
ij
=
"
1
0 0
0 0 0
0 0 0
#
$
%
%
%
&
'
(
(
(






(b) Biaxial stress %
11
= %
1
; %
22
= -%
1
; %
12
= 0

!
"
ij
=
"
1
0 0
0 - "
1
0
0 0 0
#
$
%
%
%
&
'
(
(
(







Representation in the 3 principal planes









IB Course D Mechanics of Materials and Structures Lent 2014 DH18

(c) Triaxial tension %
11
= %
1
; %
22
= %
2
; %
33
= %
3
(%
1
> %
2
> %
3
> 0)

!
"
ij
=
"
1
0 0
0 "
2
0
0 0 "
3
#
$
%
%
%
&
'
(
(
(







3.3 Example of a Thin-Walled Tube under Applied Torque and Axial Load
Consider the following problem. A torque of 100 N m is applied to a thin-walled metal tube
(t = 2 mm), of diameter 30 mm. An axial tensile force of 6 kN is then also applied. Find the
principal stresses, and their orientation, before and after the axial force has been applied.

Firstly, it is convenient to use cylindrical polar coordinates. The appropriate stress tensor is given
by
!
"
ij
=
"
zz
"
z#
"
zr

"
#z
"
##
"
#r
"
rz
"
r#
"
rr
$
%
&
&
&
'
(
)
)
)
where %
zz
, %
##
and %
rr
are the axial, hoop and radial stresses respectively.





The stress tensor can be simplified considerably. Since the wall is thin, and the surfaces are free,
there are no through-thickness stresses ie %
rr
= %
rz
= %
r$
= 0. The plane of the surface of the tube
is thus a principal plane and the Mohrs circle construction can be applied to it. The physical
situation and Mohrs circle for the case where no axial force is applied are shown in Fig.3.1(a).

IB Course D Mechanics of Materials and Structures Lent 2014 DH19


Fig.3.1(a) Physical situation and corresponding Mohrs circle when only torque is applied to a
thin-walled tube.
There are no normal stresses acting in either axial (z) or hoop ($) directions. The Mohrs circle
is centred on the origin - this is sometimes termed a state of pure shear.



There are, however, normal stresses acting on all planes other than those with their normals in
axial or hoop directions. Those with the largest and the smallest (most tensile and most
compressive) normal stresses (ie the principal stresses) are oriented at 45 to these (since they are
90 from the axial and hoop directions in Mohr space) and have values of %
#z.







IB Course D Mechanics of Materials and Structures Lent 2014 DH20

The value of %
#z
is found by recognising that a torque results from a tangential force given by the
product of %
#z
and the sectional area of the tube:
!
T = F
tan
r F
tan
="
#z
2$rt
( )


!
"#
$z
=
T
2%r
2
t
= 35.4MPa

When the %
zz
stress (= +31.8 MPa) is added, the Mohrs circle becomes larger and its centre
moves horizontally in the tensile direction - see Fig.3.1(b). The principal stresses also change in
both magnitude and direction. The shear stresses on the $ and z planes, however, remain unaffected
(%
$z
= %
z$
= 35.4 MPa) and there is no normal stress on the $ plane.


Fig.3.1(b) Physical situation and corresponding Mohrs circle when both a torque and an axial
force are applied to a thin-walled tube.




IB Course D Mechanics of Materials and Structures Lent 2014 DH21

The angle # between %
1
stress and z-axis is given by

!
tan 2" =
#
z$
#
zz
/2
=
35.4
15.9
%" = 32.9 (3.1)

The principal stresses are also readily found, since the radius, R, of the Mohrs circle is given by

!
cos 2" =
#
zz
/2
R
$R =
15.9
cos 65.8
( )
= 38.8 MPa (3.2)
The centre of the circle has been displaced from the origin by %
zz
/2, ie by 15.9 MPa, so the principal
stresses are 15.938.8, ie +54.7 MPa and -22.9 MPa.










3.4 Verification using the Secular Equation
It may be noted that this problem, and indeed all such problems tackled using Mohrs circle,
could alternatively have been solved by manipulating the secular equation (Eqn.(2.5)). For the
torque only situation, this can be written
(3.3)
which can be expanded to
!
"
2
#$
%z
2
= 0
&$
1
= +$
%z
= +35.4 MPa and $
2
= #$
%z
= #35.4 MPa
(3.4)





IB Course D Mechanics of Materials and Structures Lent 2014 DH22


For the torque plus axial load case, the corresponding equation is
(3.5)
which expands to
(3.6)
The orientation of the principal stresses, ie the value of the angle #, can be obtained from Eqn.(3.1)
in the same way as was done using Mohrs circle. (This equation is simply derived from those
presented as Eqn.(2.15).)
However, it should also be noted that, while use of Mohrs circle is never essential, it can
nevertheless be very helpful and convenient, particularly in terms of visualising stress states.
A useful TLP is available which provides an interactive tool for plotting a Mohrs circle
according to a users specified set of stresses:
http://www.doitpoms.ac.uk/tlplib/metal-forming-1/index.php

IB Course D Mechanics of Materials and Structures Lent 2014 DH23

4 Representation of Strain
4.1 Introduction
Stresses tend to deform a body. This deformation could be elastic or plastic, but we are
concerned here, at least in the first half of this course, with elastic (reversible) strains only. (For
most materials, the elastic limit is less than 1%, so elastic strains are normally small.) Care is
needed in defining a strain tensor since it could contain an unwanted rigid body rotation as well as
the genuine shape change of the body. The true strain tensor should reflect only the shape
change. It will become evident that the simple definitions of strain given in the first lecture
(particularly shear strain) do not satisfy this requirement.

4.2 The Relative Displacement Tensor
The application of a set of stresses causes all points in the body to be displaced, relative to their
initial locations, so that their coordinates, in a given frame of reference, will change. Provided that
the body is macroscopically homogeneous, these displacements can be related to the stress state,
via a description of the elastic properties of the material, which may, of course, be anisotropic.
(However, the focus here is on the motion of points in the body, and there is no reference at this
stage to the stress state that produced this motion.)
The relative displacement tensor (sometimes, rather confusingly, called the deformation
tensor), indicates how any point in the body becomes displaced (extended and/or sheared). It is a
second rank tensor, with the first suffix representing the direction in which the displacement has
occurred and the second one the reference direction. This is illustrated in Fig.4.1. All of the terms
are (dimensionless) ratios of two distances. The shear terms (e
ij
, with i(j) can also be considered as
angles - since they are small, they are approximately equal to their tangents. This is how strain was
defined in the first lecture although not using tensor terminology.

Fig.4.1 Illustration, in the y-z (2-3) plane, of how the terms of the relative displacement tensor
are defined, showing (a) a normal term, e
33
, and (b) a shear term, e
23
.
IB Course D Mechanics of Materials and Structures Lent 2014 DH24

The sign of a shear component is taken as positive when the positive axis is rotated towards the
positive direction of the other axis. For example, in Fig.4.1(b) the e
23
component would be positive.
It is clear, however, that for a general two-dimensional strain the relative displacement tensor
will not be symmetric, e.g. e
23
( e
32
. This is illustrated on the left side of Fig.4.2, for shear
deformation in the y-z (2-3) plane. Nevertheless it is possible to decompose the relative
displacement tensor into two parts, one symmetrical (the true strain) and the other anti-symmetrical
(the rigid body rotation). This is seen on the right side of Fig. 4.2.


Fig.4.2 Illustration of how, in the 2-3 plane, the relative displacements e
23
and e
32
can be
represented as the sum of a strain, %
23
(=%
32
) and a rigid body rotation, &
23
(=-&
32
).









4.3 The Strain Tensor
The separation illustrated in Fig.4.2 can be expressed more generally. Any second rank tensor
can be expressed as the sum of a symmetrical tensor and an anti-symmetrical tensor. For the
relative displacement tensor, this can be written as
(4.1)
Since %
ij
is a symmetrical tensor
(4.2)
IB Course D Mechanics of Materials and Structures Lent 2014 DH25

whereas for &
ij
, the rotation tensor, which is anti-symmetrical
(4.3)
Its quite easy to see that such an anti-symmetrical tensor represents only a rotation. For
example, in the 2-3 plane, the components of such a tensor can be written
!
"
ij
=
0 e
23
#e
23
0
$
%
&
'
(
)
(4.4)
and it can be seen in Fig.4.3 that this represents solely rigid body rotation. The residual component
of the relative displacement tensor, ie the !
ij
tensor, represents the strain.

Fig.4.3 Illustration of how an anti-symmetrical relative displacement tensor, in the 2-3 plane,
represents only a rotation, with the body not being subject to any strain.







As in the case of a stress tensor, a strain tensor can be diagonalised to give the principal strains.
These are the normal strains in the principal directions. As with stress, the principal directions are a
unique set of (orthogonal) plane normals. For these directions, there are no shear strains. A strain
tensor can be manipulated in a very similar way to a stress tensor. For example, Mohrs circle can
be used to find the principal strains or to facilitate other calculations or visualisations concerning
the strain.
IB Course D Mechanics of Materials and Structures Lent 2014 DH26

In full the strain tensor is given by
!
"
11
"
12
"
13
"
12
"
22
"
23
"
13
"
23
"
33
#
$
%
%
%
&
'
(
(
(
=
e
11
e
12
+ e
21
( )
/2 e
13
+ e
31
( )
/2
e
21
+ e
12
( )
/2 e
22
e
23
+ e
32
( )
/2
e
31
+ e
13
( )
/2 e
32
+ e
23
( )
/2 e
33
#
$
%
%
%
&
'
(
(
(
Referred to principal axes it is given by
!
"
ij
=
"
1
0 0
0 "
2
0
0 0 "
3
#
$
%
%
%
&
'
(
(
(



4.4 Some Simple Example Cases
We can now consider some familiar examples of states of strain. Fig.4.4 depicts states of pure
shear, pure rotation and simple shear. The latter involves both rotation and strain, whereas the
first involves no rotation and the second involves no strain.


Fig.4.4 Pictorial representation, in 2-D, of combinations of the shear components of the
relative displacement tensor, corresponding to three simple limiting cases.

It will be seen in the next lecture that tensorial shear strains actually differ from the shear strains
that are commonly used in engineering practice. However, for the moment we should just note that
the state of strain in a body is fully and rigorously represented by the strain tensor.

IB Course D Mechanics of Materials and Structures Lent 2014 DH27

4.5 Hydrostatic and Deviatoric Components of the Strain Tensor
A point worth noting at this stage is that the strain tensor can itself be divided into two
components one representing the change in volume of the body and the other the change in its
shape. These are commonly termed the hydrostatic and deviatoric components. Consider a cube
subjected to principal strains !
1
, !
2
and !
3
. The volumetric strain, often called the dilatation, can be
written as (see 1A Course D):
(4.5)

In fact, we can write ) = !
ii
for any strain tensor, since the sum of the diagonal terms of a second
rank tensor is an invariant, and hence independent of the reference axes see Eqn.(2.7). The mean
hydrostatic strain tensor has one third of the dilation for all three of the normal terms, and no shear
terms. The residual part is the deviatoric component

!
"
11
"
12
"
13
"
12
"
22
"
23
"
13
"
23
"
33
#
$
%
%
%
&
'
(
(
(
=
)/ 3 0 0
0 )/ 3 0
0 0 )/3
#
$
%
%
%
&
'
(
(
(
+
"
11
* )/ 3 "
12
"
13
"
12
"
22
* )/ 3 "
23
"
13
"
23
"
33
* )/ 3
#
$
%
%
%
&
'
(
(
(
(4.6)






4.6 Field and Matter Tensors
The tensors we have been treating so far have all been field tensors. These are imposed on
bodies, or regions of space, in some way. Relationships between field tensors, for example between
stress and resultant strain, depend on properties on the material concerned. These are characterised
by matter tensors. Matter tensors reflect the symmetry exhibited by the material. In this course,
we will be treating only isotropic materials, so the arrays of values in the matter tensors we will be
handling will tend to exhibit a high degree of symmetry. However, its helpful to appreciate how
tensors can be used to analyse more complex cases, involving anisotropic materials.




IB Course D Mechanics of Materials and Structures Lent 2014 DH28

4.7 Stiffness and Compliance Tensors
The relationship between a stress (tensor) and the strain (tensor) generated by it can be written
(4.7)
in which C
ijkl
is the stiffness tensor. This is a fourth-rank tensor, with 3
4
(=81) components. (The
rank of a tensor is equal to the number of its suffices, although there are some situations in which a
convention may be used such this is not the case - an example is provided by principal stresses and
strains being given a single suffix.) Eqn.(4.7) is a generalised expression of Hookes law. It
represents 9 equations, generated according to the Einstein summation convention. For example,
the first of these is

(4.8)




The stress-strain relationship can also be expressed in the inverse sense
!
ij
= S
ijkl
"
kl
(4.9)
in which S
ijkl
is the compliance tensor. (Note that the symbols conventionally used for stiffness and
compliance are the reverse of the initial letters of these words.)







4.8 Relationship to Elastic Constants
This looks a little cumbersome and daunting, but in practice the treatment can be simplified. The
symmetry of stress and strain tensors when the body is in static equilibrium means that
C
ijkl
= C
ijlk
= C
jikl
= C
jilk
(4.10)
reducing the number of independent components from 81 to 36. Furthermore, the symmetry
exhibited by the material commonly results in further reductions in this number. In fact, for a cubic
material it reduces to 3. For an isotropic material (eg a polycrystal) where the stiffnesses and
IB Course D Mechanics of Materials and Structures Lent 2014 DH29

compliances are independent of the choice of reference axes, it is reduced to just 2, so the elastic
behaviour of isotropic materials is fully specified by the values of 2 constants.
In terms of engineering constants, the relationship between (normal) stress and strain can be
obtained by considering the application of a single stress %
1
, generating a strain !
1
, so that
(4.11)
in which E is the Youngs modulus. One might be tempted to deduce that E is given by C
1111
.
However, this is incorrect, since the application of a stress %
1
generates, not only a (direct) strain !
1
,
but also Poisson strains !
2
and !
3
(see below), which would appear in the full (tensorial) equation
for %
1
.









Expressed in terms of compliance, however, the equation we need (from Eqn.(4.9)) is
(4.12)
so it follows that
!
E =
1
S
1111
(4.13)

4.9 Poisson Ratio
The second elastic constant specified for an isotropic material is commonly the Poisson ratio, *.
This gives the transverse contraction strain which accompanies an axial extension strain. Again
considering a single (normal tensile) stress %
1
, generating principal strains !
1
, !
2
and !
3
, the Poisson
ratio is defined by
(4.14)
The tensorial expression for !
2
, with only %
1
applied, is
(4.15)
so that
(4.16)
IB Course D Mechanics of Materials and Structures Lent 2014 DH30

Since Poisson strains simply superimpose during multi-axial loading, we can write the
following expressions for the (principal) strains arising from a set of (principal) stresses
(4.17)







For an isotropic material, the 36 components of the compliance tensor can be represented by a 6x6
matrix leading to a pseudovector-matrix form of Eqn (4.9):

!
"
1
"
2
"
3
2"
23
2"
31
2"
12
#
$
%
%
%
&
%
%
%
'
(
%
%
%
)
%
%
%
=
1
E
1 - * - * 0 0 0
-* 1 - * 0 0 0
-* - * 1 0 0 0
0 0 0 2(1+*) 0 0
0 0 0 0 2(1+*) 0
0 0 0 0 0 2(1+*)
+
,
-
-
-
-
-
-
-
.
/
0
0
0
0
0
0
0
1
1
1
2
1
3
1
23
1
31
1
12
#
$
%
%
%
&
%
%
%
'
(
%
%
%
)
%
%
%










IB Course D Mechanics of Materials and Structures Lent 2014 DH31

5 Engineering Stresses and Strains and Plasticity
5.1 Engineering Shear Strains and the Shear Modulus
A minor complication arises when considering shear strains and the shear modulus. One might
imagine that the shear modulus, G, would be defined as shear stress divided by shear strain, eg
%
12
/!
12
. Unfortunately, G ( %
12
/!
12
. The problem is associated with the difference between a
tensorial shear strain, such as !
12
, and the corresponding engineering shear strain, "
12
. The shear
modulus is conventionally (see first lecture) defined by
(5.1)
in which the value of "
12
is simply the measured strain, with no account being taken of the fact that
it really represents the sum of two shear strains, one being generated by each of the pair of shear
stresses which is being applied. This is illustrated in Fig.5.1.

Fig.5.1 Illustration of how a pure shear deformation, in the 1-2 plane, when rotated about the
3-axis, can be viewed as a simple shear, and used to define the shear modulus, G, in
terms of the engineering shear strain, '
12
. The presence of the other shear stress acting
in this plane, !
21
, is neglected in this definition.

The array of engineering strains (%
ii
and '
ij
) thus do not form a tensor and cannot be
transformed using the standard procedures for second rank tensors. However, we can still define
the shear modulus in terms of components of the compliance tensor. With only %
12
and %
21
acting,
Eqn(4.9) gives
!
"
12
= S
1212
#
12
+ S
1221
#
21
= 2S
1212
#
12

!
"G =
#
12
$
12
=
#
12
2%
12
=
1
4S
1212
(5.2)
IB Course D Mechanics of Materials and Structures Lent 2014 DH32

In fact, similar procedures can be used to establish all of the components of the compliance and
stiffness tensors, in terms of (measured) engineering elastic constants. Of course, this procedure is
more complex for anisotropic materials, for which there are more than 2 independent elastic
constants, but it can still be carried out.



5.2 Bulk Modulus
It may finally be noted that the bulk modulus, K, is defined by (see 1A Course D):
!
K =
"
H
#
=
"
1
+"
2
+"
3
( )
/ 3
$
1
+$
2
+$
3
( )
=
"
ii
3$
jj
(5.3)
where %
H
is the hydrostatic component of the stress state. The bulk modulus is a measure of the
resistance of the material to volume change. A general stress tensor, referred to the principal axes,
can be written as the sum of a hydrostatic component (pressure) and a deviatoric stress, in a way
similar to that discussed for strain in section 4.5.

!
"
1
0 0
0 "
2
0
0 0 "
3
#
$
%
%
%
&
'
(
(
(
=
"
ii
/3 0 0
0 "
ii
/3 0
0 0 "
ii
/ 3
#
$
%
%
%
&
'
(
(
(
+
"
1
- "
ii
/3 0 0
0 "
2
- "
ii
/3 0
0 0 "
3
)"
ii
/3
#
$
%
%
%
&
'
(
(
(








5.3 Relationships between Elastic Constants
Since only two elastic constants are required to fully define the behaviour of an (isotropic)
material, it follows that there must be inter-relationships between the 4 that have been presented
here. For example, adding up the 3 equations in Eqn.(4.17) gives
(5.4)
IB Course D Mechanics of Materials and Structures Lent 2014 DH33

Several other relationships are commonly quoted, and can be useful. For example, the shear
modulus is given by
(5.5)





5.4 Large Uniaxial Tensile Strains and Plastic Deformation
Application of a uniaxial tensile stress %
1
generates contraction strains given by

!
"
2
="
3
= #$"
1

The volumetric strain is thus

!
" =#
1
+#
2
+#
3
=# 1$2%
( )
(5.6)
For a ductile material (e.g. a metal) * + 1/3 and a small elastic strain of 0.5% would produce a
volume change of about 0.2%. Eqn. (5.6) indicates that the elastic volume change decreases as *
increases. For a plastically deforming metal subject to large strains the volume change is zero and *
is effectively 1/2 (although * is only properly defined for elastic deformation). Since volume is
conserved during plastic deformation then, for the sample shown below, A
0
l
0
= A
i
l
i









Fig. 5.2 Change in sample dimensions under uniaxial tensile load which results in permanent
deformation.
IB Course D Mechanics of Materials and Structures Lent 2014 DH34

Under constant load the cross-sectional area of the sample decreases and hence the instantaneous
stress increases. This instantaneous or true stress is defined as %
T
= F/A
i
. It is clearly larger than
the nominal (or engineering) stress %
N
= F/A
0
. The difference between the two quantities increases
with increasing plastic deformation.
A true strain can also be defined based on the instantaneous sample length. It is determined by
summing the incremental strains:

!
"
T
=
#l
l
0
+
#l
l
1
+
#l
l
2
+ ......... =
#l
l
i
$
%
&
'
(
)
i
*

where l
1
= l
0
+ ,l, l
2
= l
1
+ ,l etc.
In differential form (1A Course D):

!
"
T
=
dl
l
l
0
l
i
#
= ln
l
i
l
0
$
%
&
'
(
)

The true and nominal stresses and strains can be related as follows:

!
"
T
= ln 1+ "
N
( )
where "
N
=
l
i
# l
0
l
0
$
T
=
F
A
i
=
F
A
0
A
0
A
i
= $
N
A
0
A
i
%
&
'
(
)
*
= $
N
l
i
l
0
%
&
'
(
)
*
thus $
T
= $
N
1+ "
N
( )

For small stresses and strains (i.e. before yielding) %
T
+ %
N
and !
T
+ !
N
and there is no distinction
between the quantities.

5.5 Plastic Instability and Considres Construction
The figure below compares a typical nominal stress-strain curve for a polycrystalline metal with
the corresponding true stress-strain curve.






IB Course D Mechanics of Materials and Structures Lent 2014 DH35

It is seen that the nominal curve passes through a maximum (point a) beyond which continued
straining requires an ever-decreasing applied force. The maximum, sometimes called the ultimate
tensile strength, signals the onset of a plastic or geometric instability. It is geometric in the sense
that it does not reflect changes in material properties. The instability manifests itself in the form of a
neck as shown below:







Fig. 5.3 Illustration of neck formation under large uniaxial tensile deformation.
Necking occurs when the increase in yield stress due to work hardening cannot compensate
for the larger increase in stress due to the reduction in cross-sectional area. Once necking
begins the strain is restricted to the necked region. Voids form where the local stress is large and
eventually the metal fractures at the neck (point b).
The true stress-strain curve shows that %
T
> %
N
beyond the elastic limit (point c) as expected. It is
also concave down but does not exhibit a turning point before fracture (point b). Nevertheless the
onset of necking can still be predicted from a true stress-strain curve using Considres criterion
(1885).
Necking occurs at maximum load, i.e. dF = 0. Therefore necking begins when:

!
dF = A
i
d"
T
+"
T
dA
i
= 0
From conservation of volume we have

!
dV = A
i
dl
i
+ l
i
dA
i
= 0
Hence by combining these equations we find

!
d"
T
"
T
= #
dA
i
A
i
=
dl
i
l
i
= d$
T

and so at the onset of necking it follows that
IB Course D Mechanics of Materials and Structures Lent 2014 DH36


!
d"
T
d#
T
="
T

This is Considres criterion for necking. When the slope of the true stress-strain curve
equals the true stress necking begins. This is represented graphically by a tangent line which
intersects the true strain axis 1 unit away from the relevant strain.








Fig. 5.4 Schematic true stress-strain curve showing Considres tangent construction.
In practice the tangent construction is usually made by plotting the true stress against the
nominal strain (or extension ratio, ' = !
N
+ 1) which also does not exhibit a turning point. From the
relationships above it is easily shown that

!
d"
T
d#
N
=
"
T
#
N
+1

so that in this case the tangent line intersects the nominal strain axis at -1.








Fig. 5.5 Schematic true stress versus nominal strain curve showing Considres tangent
construction.
IB Course D Mechanics of Materials and Structures Lent 2014 DH37


Nominal stress-strain curves are useful in determining strength and ductility data for purposes of
engineering design. True stress-strain curves are useful for understanding the work hardening
process in metals and also for predicting the drawability of different materials. Brittle materials, for
example, exhibit concave upwards curves where no tangent can be constructed and polymers
exhibit sigmoidal curves where two tangents can be constructed. This is discussed further in the Pt
II Materials Science course.



IB Course D Mechanics of Materials and Structures Lent 2014 DH38

6 - Beam Bending
6.1 Introduction to Beam Bending
While mechanics treatments, and mechanical testing procedures, are often based on the uniform
application of a normal stress to a body, such a loading configuration is in fact relatively unusual in
engineering practice. Its much more likely that a component or structure will be subjected to
bending moments. In fact, virtually all structures, including bridges, buildings and many natural
structures (trees, bones etc) are commonly subjected to significant applied moments.





Beam stiffness is an important concept for many types of structure, particularly those with
slender shapes (large span-to-height ratio). Inadequate beam stiffness can lead to large deflections,
and may also cause high, localised, stresses and a danger of failure in that region.
Slightly less common, but still of considerable importance, are situations in which bodies are
subjected to twisting, or torsional moments (torques). This lecture is focussed on beam bending
and the next lecture covers torsion.






6.2 Beam Curvatures (Pt 1A revision)
The concept of the curvature of a beam, -, is central to beam bending. Fig.6.1 shows that the
axial strain in the beam, %
xx
, is given by the ratio y/R, where y is the distance from the neutral axis
(neutral plane in 3-D) and R is the radius of curvature. Equivalently, 1/R (the curvature, -) is
equal to the through-thickness gradient of axial strain.


IB Course D Mechanics of Materials and Structures Lent 2014 DH39




Fig.6.1 Relation between the radius of curvature, R, the beam curvature, (, and the strains
within a beam subjected to a bending moment

6.3 Bending Moments, Stresses and Second Moments of Area
Bending moments are produced by transverse loads applied to beams. The simplest case is the
cantilever beam, widely encountered in balconies, aircraft wings, diving boards etc - see Fig.6.2.
One end of the beam is fixed while the other is subjected to a concentrated load and is free to bend.
Loads distributed along the beams length will have the same effect. In Fig. 6.2 the load is
downwards and the beam is hogging (i.e tension in the top surface, compression in the lower
surface). The mass of the beam is neglected in determining the loading action. In addition to the
bending moment there is a shearing force acting vertically along the length of the beam and in this
case is constant and equal to F.
IB Course D Mechanics of Materials and Structures Lent 2014 DH40


Fig.6.2 Balancing the external and internal moments during the bending of a cantilever beam

The bending moment acting on a section of the beam, due to an applied transverse force, is given
by the product of the applied force and its distance from that section. For the cantilever, this gives
(6.1)
Moments thus have units of N m. Note that it is necessary to adopt a sign convention for M. Here
M is taken as positive when the upper surface is in tension, i.e. the beam shape is concave
downwards and hogging. However, this convention is somewhat arbitrary and could be completely
reversed as seen in many mechanics textbooks. It is thus seen that in the case of the end-loaded
cantilever (Fig. 6.2) the bending moment has a maximum value of +FL at the fixed end and
decreases linearly to zero at the free end.
The moment, as expressed in Eqn.(6.1), may be regarded as being externally imposed, and tending
to rotate the beam. It is balanced by the internal moment arising from the axial stresses generated
in the beam. This moment is given by a summation of all of the internal moments acting on
individual elements within the section. These are equal to the product of the force acting on the
element (stress times area of element) and its distance from the neutral axis, y. The internal
moment can therefore be written
(6.2)
where dA is an element of the section at a uniform distance, y, from the neutral axis. This can be
presented more compactly by defining I (the second moment of area, or moment of inertia) as

!
I = y
2
dA
A
"
(6.3)

IB Course D Mechanics of Materials and Structures Lent 2014 DH41

The units of I are m
4
. The value of I is dependent solely on the beam sectional shape. The
moment can now be written as
(6.4)
Since the product EI is a constant for a given beam shape, the beam curvature - varies in the same
way as M along the length of the beam, which in the case of the cantilever is a linearly decreasing
function of x as seen in Fig. 6.2.

6.4 Beam Stiffness
The product EI is termed the beam stiffness, or sometimes the flexural rigidity. It is a
measure of how strongly the beam resists deflection. The beam curvature, -, is thus given by the
ratio of applied moment to beam stiffness, in an analogous way to axial strain being equal to the
ratio of applied stress to Youngs modulus, under uniaxial loading.










Fig.6.3 Illustration of how the integration is carried out in order to find the expression for I in
the cases of (a) rectangular section and (b) circular section beams
IB Course D Mechanics of Materials and Structures Lent 2014 DH42

The expression for I, for a given sectional shape, is established by carrying out the integration of
Eqn.(6.3). For example, for a rectangular beam of width w and thickness h (see Fig.6.3(a)), it is
given by
(6.5)
For a circular section (see Fig.6.3(b)), the integration is more complex, although the final
expression is again a simple one
(6.6)
Note that these expressions for I are obtained with respect to the neutral (centroidal) axis. Other
expressions would be obtained using a different axis. In general I increases as the reference axis is
moved parallel to itself further from the centroid (centre of mass for a beam of uniform density).
This is indicated by the parallel axis theorem which states
!
I
parallel
= I
neutral
+ d
2
A
where d is the distance between the neutral and parallel axes and A is the area of the section.
i.e.
!
I
parallel
= y + d
( )
"
2
dA = y
2
dA+2d ydA
" "
+ d
2
dA
"


6.5 Maximising the Beam Stiffness
Expressions such as Eqns.(6.5) and (6.6) are clearly useful in beam bending calculations.
However, even without such equations, some simple guidelines can be identified for maximising
the value of I, for a given sectional area (ie for a given beam mass). Its clear that I is raised by
ensuring that much of the section is at large distances from the neutral axis. The classical I-
beam section, commonly used in many constructions (see Fig.6.4) is an example of the application
of this principle.

Fig.6.4 Photo showing I-beams being used in a civil engineering construction
IB Course D Mechanics of Materials and Structures Lent 2014 DH43









Another commonly-employed approach is to use hollow sections, which also tends to ensure that
the sectional area close to the neutral axis is reduced. Calculation of I for hollow beams is very
straightforward, since it is obtained by simply subtracting the I of the missing section from that of
the overall section. For example, that for a cylindrical tube is given by
(6.7)
where D is the outer diameter of the tube and d is the inner diameter.

A similar result is obtained for a hollow box section. The procedure can be particularly useful in
obtaining I for asymmetrical shapes like a T-section. For these shapes the location of the neutral
axis (NA) also has to be determined. This can be done by calculating the coordinates (z, y) of the
centroid of the shape. These are given by the first moment of inertia divided by the area i.e.
!
z' = zdA
"
dA y' = ydA
"
dA
" "








IB Course D Mechanics of Materials and Structures Lent 2014 DH44

7 - Examples of Beam Bending and Torsion
7.1 Beam Deflections
Eqn.(6.4) allows the curvature distribution along the length of a beam (ie its shape), and the
stress distribution within it, to be calculated for any given set of applied forces (ie any variation of
the applied moment along the length). The deflection at any point along the length can therefore be
calculated. In this section we illustrate this by deriving expressions for the beam deflection along
an end-loaded cantilever and along a beam subjected to symmetrical 3-point bending.
The beam support and loading configuration is different in each case. The cantilever is fixed at
one end and therefore experiences a reaction moment at that end. In 3-point bending the ends are
simply-supported which means they rest on knife-edges or frictionless pins perpendicular to the
plane of bending and do not experience reaction moments. The cantilever hogs (positive M)
whereas in 3-point bending the beam sags (negative M). In both cases the deflection is downwards
which is taken to be the positive y-direction and the x-y origin is at the left end of the beam. The
orientation of the axes and location of the origin is somewhat arbitrary and the analysis could
equally be performed with other choices.

7.2 Cantilever beam
The beam curvature, -, is approximately equal to the curvature of the line traced by the neutral axis,
d
2
y/dx
2
, as shown in Fig. 7.1a.

Fig.7.1a Approximation involved in equating beam curvature to the curvature of the neutral axis
IB Course D Mechanics of Materials and Structures Lent 2014 DH45

Hence
!
d
2
y
dx
2
=" =
M
EI

Therefore using Eqn (6.1) for the bending moment of an end-loaded cantilever:
!
EI
d
2
y
dx
2
= M = F(L " x)
The deflection is found by integrating this expression, using the boundary conditions to establish
the integration constants
!
EI
dy
dx
= FLx "
Fx
2
2
+ C
1

!
at x = 0,
dy
dx
= 0, "C
1
= 0
!
EIy =
FLx
2
2
"
Fx
3
6
+ C
2

!
at x = 0, y = 0, "C
2
= 0
The defections along the length of the beam, and specifically at the loaded end, are thus given by
!
y =
Fx
2
6EI
3L " x
( )

!
" =
FL
3
3EI


7.3 Symmetrical 3-point bending


Fig.7.1b Symmetrical 3-point bending


IB Course D Mechanics of Materials and Structures Lent 2014 DH46

The bending moment on the left hand side of the beam (0 . x . L/2) is given by
(7.1)
From Eqn.(6.4), and writing the curvature, -, as d
2
y/dx
2
,
(7.2)
Integrating with respect to x, and applying boundary conditions to find the integration constants
(7.3)
(7.4)
so the equation for the deflection is
(0 . x . L/2) (7.5)
and the deflection of the centre of the beam (x = L/2) is given by
(7.6)

A useful TLP is available which provides an interactive tool for simulating beam bending according
to a users specified loading geometry and beam stiffness:
http://www.doitpoms.ac.uk/tlplib/beam_bending/index.php

7.4 Introduction to Torsion
Torsion is the twisting of a beam under the action of a torque (twisting moment). It is
systematically applied to screws, nuts, axles, drive shafts etc, and is also generated more randomly
under service conditions in car bodies, boat hulls, aircraft fuselages, bridges, springs and many
other structures and components. A torque, T, has the same units (N m) as a bending moment, M.
Both are the product of a force and a distance. In the case of a torque, the force is tangential and the
distance is the radial distance between this tangent and the axis of rotation. While the treatment of
torsion follows logically from that beam bending, there are some differences that need to be noted.
IB Course D Mechanics of Materials and Structures Lent 2014 DH47








7.5 Torsion of Thin-Walled Circular Tubes


Fig.7.2 Application of a torque, T, to a thin-walled tube, showing the stress state experienced
by a small volume element of the tube wall

In section 3.3 we considered the shear stresses acting on a thin-walled tube under torsional
loading. When a torque is applied, the wall is subjected to a state of pure shear, as illustrated in
Fig.7.2. The relation between the torque and the resultant shear stress in the plane of the wall, %
#z

(= %
z#
), was found to be
(7.7)
Because the wall is thin (t << r) it is assumed that this stress does not vary significantly across
the thickness of the wall. This is not true, however, for thick-walled tubes or solid cylinders as
shown below.

IB Course D Mechanics of Materials and Structures Lent 2014 DH48

7.6 Torsion of Solid Cylindrical Bars
In most cases, the cylindrical bodies to which torques are applied are not thin-walled tubes, but
are solid. The analysis is more complex in such cases, since the shear stresses induced will vary
with radial location (distance from the axis of the torque). Furthermore, in order to analyse the
situation properly, we need to consider the response of the cylinder to the torque see Fig.7.3.



Fig.7.3 Torsion of a Cylindrical Bar

In contrast to the case of beam bending, the shape of the cylinder

does not change when a torque


is applied. However, there is a rotation about the torque axis, which would be apparent if there
were reference lines marked along the axis of the cylinder, as in Fig.7.3. It can be seen that the
(engineering) shear strain in an element of the bar is given by

(7.8)

This is, of course, only true for a cylinder. It would not, for example, apply to a square section bar, or other prismatic
bodies with various sectional shapes. Furthermore, torques are commonly applied to bodies which are not prismatic,
but have sections which change along their length such as bolts. However, while the analysis is clearly more complex
in such cases, the principles described here are still applicable to them.
IB Course D Mechanics of Materials and Structures Lent 2014 DH49








This equation applies both at the surface of the bar, as shown, and also for any other radial location,
using the appropriate value of r. Clearly, the shear strain varies linearly with r, from zero at the
centre of the bar to a maximum value at the free surface (r =R)
!
"
#z
max
=
Rd$
dL

The quantity d&/dL is called the rate of twist. In the special case of pure torsion, the rate of twist
equals the total angle of twist divided by the length of the bar:
!
"
#z
max
=
R$
L


The shear stress, %
#z
, at any radial location, is related to the shear strain by
!
"z
= G#
"z
(7.9)
where G is the shear modulus (see Eqn.(5.1). Substituting from Eqn.(7.8), it follows that
!
"z
= Gr
d#
dL
(7.10)
The torque, T, can therefore be written as
T = dT
A
!
= "
#z
r dA
A
!
= G r
2

d$
dL
dA
A
!
(7.11)
As with beam bending, the geometrical integral is represented as a (polar) second moment of area
I
P
= r
2
dA
A
!
(7.12)
Note the relationship between the polar second moment of area and the (rectangular) second
moments of area defined earlier for beam bending:
!
r
2
"
dA = z
2
+ y
2
( ) "
dA = z
2
dA+ y
2
dA
" "

i.e. I
P
= I
z
+ I
y


IB Course D Mechanics of Materials and Structures Lent 2014 DH50


The torque is thus given by
T = G I
P
d!
dL
(7.13)
Comparing this equation with the corresponding one for beam bending (Eqn.(6.4))
M = E I ! (7.14)
it can be seen that the torsional analogue for the curvature of a bent beam is the rate of twist along
the length of the bar, d#/dL.
For a solid cylinder of diameter D, I
P
can be written as

!
I
P
= r
2
dA
A
"
= r
2
2#rdr
0
D / 2
"
=
#r
4
2
$
%
&
'
(
)
0
D / 2
=
#D
4
32
(7.15)
By combining Eqns (7.10) and (7.13) it is seen that the maximum shear stress, which is on the
surface of the cylinder, is given by
!
"
#z
max
=
TR
I
P
=
16T
$D
3
(7.16)

7.7 Torsion of Thick-Walled Circular Tubes
Most of the torsion equations derived for solid cylinders apply to thick-walled circular tubes
with suitable modification for the range of allowable radii (e.g. inner r
1
to outer r
2
). For example the
shear stress is given by
!
"
#z
=
Tr
I
P
where
!
I
P
=
"
32
D
2
4
# D
1
4
( )
=
"
2
r
2
4
# r
1
4
( )
(7.17)

If r is defined as the average radius of the tube equal to (r
1
+ r
2
)/2 then the expression for I
P

becomes
!
I
P
=
"rt
2
4r
2
+ t
2
( )
(7.18)
where t = r
2
r
1
, the tube thickness.
If the thickness of the tube is small (t << r) then Eqn (7.18) reduces to
!
I
P
" 2#r
3
t which is effectively the same as that used in Eqn (7.7).

Thick-walled circular tubes can be more efficient at resisting torsional loads than solid bars
if weight is taken into consideration. For example, for a fixed mass per unit length, a tube will have
a larger outer radius than the solid bar and a larger polar second moment of area resulting in a
IB Course D Mechanics of Materials and Structures Lent 2014 DH51

higher torsional stiffness and a lower maximum shear stress for a given applied torque. As a
consequence large drive shafts, propeller shafts and generator shafts often have tubular cross-
sections.







It is emphasised that the torsion equations for circular cylinders and tubes do not apply to bars
with other shapes. For example bars with rectangular cross-sections do not remain plane under
torsion and their maximum stresses are not located at the furthest point from their centroid.
IB Course D Mechanics of Materials and Structures Lent 2014

DH52
8 Failure in Compression: Elastic Buckling of Columns
8.1 Crushing
8.1.1 Porous bodies
If a body is porous then failure can occur by material being squashed into the pores. This gives rise
to a crease of compacted material that spreads out perpendicular to the compressive stress (i.e. in
the example below, horizontally).








8.1.2 Non-porous bodies
However, if the body is solid it is not at all obvious how it will break when the atoms are pushed
together. There do not seem to be any forces pulling the atoms apart. However, consider the stress
state of uniaxial compression represented with Mohrs circle:







IB Course D Mechanics of Materials and Structures Lent 2014

DH53
This tells us that there are shear stresses within the material, and that these are highest on planes
which lie at 45 to the principal axes. In ductile materials, failure could occur by plastic flow on
these planes. In brittle materials, failure occurs by the growth of cracks, causing the material to
break into pieces on the 45 planes.









Often this occurs by the formation of small cracks that lie parallel to the compression axis. Failure
occurs by the linking of these cracks along the 45 plane.
This failure mechanism is known as crushing and is commonly seen in concrete in compression.







From http://www.tfhrc.gov/safety/pubs/05063/chapt3.htm
IB Course D Mechanics of Materials and Structures Lent 2014

DH54
8.2 Elastic (Euler) Buckling
8.2.1 Freely-hinged columns
We will consider now another way in which a structure might fail in compression under a load less
than that required to cause crushing failure of the material.














A number of important assumptions must be made for a straightforward analysis of elastic
buckling:
1. The column has a uniform prismatic cross section
2. The material is homogeneous, defect-free, isotropic and linear-elastic
3. The forces due to the weight of the column itself are negligible
4. The force is axial with no eccentricity
Also, to begin with, we will assume that the ends are free to rotate
The top pin remains vertically above the bottom pin
IB Course D Mechanics of Materials and Structures Lent 2014

DH55
Let us consider the situation if the column were to deflect.
The resulting scenario is that of a bent beam

If the force F moves downwards by an increment !x, there is a change
in the mechanical energy of the system.



However that will cause an increase in elastic energy in the beam.



If |!U
E
|

> |!U
F
| then the buckling leads to an overall increase in the energy of the system (!U
E
+ !U
F

would be positive) and so it is more favourable for the column to straighten itself out.
If |!U
F
|

> |!U
E
| then the buckling leads to an overall decrease in the energy of the system
(!U
E
+ !U
F
would be negative) and so it is favourable for the column to buckle.

We can now consider the total energy stored in the beam as a function of the sideways
displacement, y.



IB Course D Mechanics of Materials and Structures Lent 2014

DH56
We will consider the case of neutral equilibrium this will allow us to calculate the critical force
(when the energies driving and resisting buckling are equal). Consider some point along the
column, which we will label Q.













Summing moments about point Q, we have the moment due to the applied external force,
(8.1)
which must be balanced by the bending moment arising from the stresses in the beam, which is
equal to the curvature, " multiplied by EI (see section 6.3).
(8.2)
and therefore just by balancing these moments,
(8.3)

IB Course D Mechanics of Materials and Structures Lent 2014

DH57
Eqn (8.3) is similar to eqn (7.2) for symmetrical three-point bending. The important difference,
however, is that for buckling the bending moment is a function of the deflection y. This is not the
case for three-point bending. For buckling the equilibrium equations take account of the geometry
of the deformed structure.
Rearranging (8.3) gives
(8.4)
where . This is an ordinary second order linear homogeneous differential equation with
constant coefficients and has the general solution
(8.5)
where A and B are determined from the boundary conditions.
The boundary conditions are that the end-points of the column do not leave the x-axis during
loading, i.e.



Using eqn. 8.6a,
(8.7)
hence B = 0.

Using eqn. 8.6b,
(8.8)
One solution to this is A = 0, but as B = 0, that means that y = 0 and so the column has not deflected.
If , then the column will have deflected. This is only possible if
(8.9)
, where n = 0, 1, 2, 3.....
IB Course D Mechanics of Materials and Structures Lent 2014

DH58
i.e. k
2
L
2
= n
2
#
2
and k
2
= F/E.I, giving the force required for elastic buckling, F
EB
;
(8.10)
Hence we can sketch possible solutions to the configurations which the column could, in theory
achieve:

n = 0 n = 1 n = 2

Hence for low forces the column will not buckle (for any slight sideways perturbation, the column
will restore itself to reduce the energy). This is a stable equilibrium. However at a critical force,
when F=F
EB
, the n = 1 solution is achieved and the column is in neutral equilibrium. If the force
exceeds F
EB
the column is unstable and any slight sideways perturbation will cause the column to
buckle catastrophically.
The force at which a column in compression will buckle is therefore:
(8.11)
IB Course D Mechanics of Materials and Structures Lent 2014

DH59

Note the critical force depends on the flexural rigidity EI of the column and not on the strength of
the material itself as represented, for example, by the yield stress.
IB Course D Mechanics of Materials and Structures Lent 2014

DH60
9 Failure in Compression: the effect of Constraints and Column
Shape
9.1 The Effect of End-Constraint
So far, we have considered both ends to be hinged. However in many situations this may not be the
case. First we will consider the situation where one end is freely hinged (i.e. free to rotate and to
move laterally) but the other end is fixed (a).











It can be seen that the force which would cause this column to buckle is equal to the force which
would cause a completely freely-jointed column of twice the length (b) to buckle, i.e.
(9.1)
that is to say that the column with one end fixed and one end freely jointed buckles under one
quarter of the load which would cause a freely jointed column of the same length to buckle.
(9.2)


IB Course D Mechanics of Materials and Structures Lent 2014

DH61
Another scenario likely to occur is that in which both ends are clamped (a):










It can be seen that the force which would cause this column to buckle is equal to the force which
would cause a completely freely-jointed column of half the length (b) to buckle, i.e.
(9.3)
that is to say that the column with two fixed ends buckles under four times the load which would
cause a freely jointed column of the same length to buckle.
(9.4)








IB Course D Mechanics of Materials and Structures Lent 2014

DH62
Hence we can write a general expression for elastic buckling as:

(9.5)




c = 1/4



1



4





IB Course D Mechanics of Materials and Structures Lent 2014

DH63
9.2 The effect of aspect ratio (or slenderness)
We wish to consider the effects of shape rather than size, so we must think in terms of the stress,
!
EB
, rather than the force for elastic buckling. The compressive stress under which our column will
buckle is given by
(9.6)
If the column is a solid cylindrical rod, then we can substitute for I and A;

(9.7)
In other words the buckling stress depends on the ratio (L/R)
2
. So as we know from everyday
experience the longer and more slender a rod becomes, the lower is the stress required to make it
buckle. The important shape parameter is L/R, which we call the slenderness ratio. The only
materials property which determines the buckling stress is E, the Youngs Modulus. Hence we can
make a simple plot of the buckling stress for columns as a function of the slenderness.












IB Course D Mechanics of Materials and Structures Lent 2014

DH64
9.3 The effect of cross-sectional shape
As the column is bending when it buckles then we would expect the shape of the cross-section to be
important. We know that if we are designing a beam, an I section can be a good choice. There is
one important difference. In bending the direction of the applied force (and hence the resulting
deformation) is usually known, whereas in buckling it can occur in any direction perpendicular to
the longitudinal axis of the column. It will therefore buckle in the direction in which EI is smallest,
which normally means the direction in which I is smallest.













So there is not much advantage to an I-beam if its job is to resist buckling. Because the buckling
failure will always occur in whichever direction has the lowest EI, we obtain the most efficient use
of material if that value is the same in all directions. Therefore cylindrical columns and those with a
section that is a regular polygon (e.g. equilateral triangle, square etc) will, for a given height and
volume, be more resistant to buckling than rectangular-section columns.

There is a way in which we can increase the second moment of area of a column whilst retaining
uniformity in all directions. That is to have a hollow cross section (moving material away from the
neutral axis). Hence for the same amount of material we can increase the resistance to buckling by
IB Course D Mechanics of Materials and Structures Lent 2014

DH65
using a hollow cylinder (which can therefore have a larger external diameter), increasing the value
of I.







Hence a drinking straw has more resistance to elastic buckling than a solid cylinder of the same
material with the same length and volume.

Does elastic buckling cause failure?
We have described buckling as an elastic phenomenon, which in principle should therefore be
recoverable. However it often leads to other types of failure such as brittle fracture or local buckling
(which we will deal with shortly).
Let us consider a column being tested under displacement control. That is we apply a force to
some end-plates (as much as is needed) and control the distance between them. At first the
displacement can be achieved by uniaxial elastic compression, but quickly the buckling force will
be reached. However the column will not instantaneously collapse at this point. It will buckle in a
stable manner and if the test is stopped the material will be able to recover (as long as the buckling
is not taken so far as to cause some other failure mechanism).




IB Course D Mechanics of Materials and Structures Lent 2014

DH66
Another way we can test a column is under load control. One way we could achieve this is to
apply a constantly increasing force to the top of a column (e.g. by stacking weights on top). Recall
that we derived the force for elastic buckling by saying that it will happen when the energy driving
buckling exceeds the energy resisting it. When this occurs, collapse of the column will be sudden
and instantaneous. However it is possible that the material will remain elastic and be able to recover
once the load is removed.



9.4 Local Buckling
Our analysis of hollow cylinders suggests that we can prevent buckling, even in a very long tube
simply by having a large external diameter. However, assuming we have a fixed amount of
material, the larger we make the external diameter, the thinner the wall will become and we know in
practice that there is a limit to how far we can reduce the wall-thickness. The failure mechanism for
thin-walled tubes is called local buckling. A paper straw kinking is an example. And just like the
straw (and unlike elastic buckling) this is irreversible.
The force required for local buckling to occur is given by:
(9.8)
where k~0.5 but depends on any surface imperfections.
A drinks can is a good example of a thin walled structure which we know fails by local buckling
and it is straightforward to calculate the load which it can support.





IB Course D Mechanics of Materials and Structures Lent 2014

DH67
9.5 The Effect of Internal Pressure
In Part IA Course D you looked at the effect that pressure has on fast fracture of cylinders. Now we
will see how it affects buckling. Consider a pressurised cylinder with some internal pressure which
is greater than the external pressure. We will call this pressure difference $P and in this case it will
have a positive value.





We know that a positive internal pressure has the effect of creating a longitudinal tensile stress in
the wall and so the total stress in the wall will be reduced:











Hence this positive internal pressure helps to prevent buckling by reducing the overall compressive
stress in the wall. A negative internal pressure would lead to buckling under a lower applied load.
IB Course D Mechanics of Materials and Structures Lent 2014

DH68
10 Brittle Fracture in Tension
10.1 The Ideal Strength
We can start by considering when we would expect the atoms within a material to separate. If we
look at a typical force vs. displacement curve between neighbouring atoms we see that bonds
usually break at about 10-20% strain, and hence we should expect a chain of atoms to have a tensile
strength between E/10 and E/5.








Hence for a material with a Youngs modulus of 100 GPa, which is typical for a metal, we might
predict that its tensile strength would be around 10 GPa. In reality however, most practical materials
fail at much lower stresses than this. Having said that, experiments carried out by Griffith
measuring glass fibres showed that the tensile strength was a function of fibre thickness with very
thin samples having strengths approaching the theoretical strength.








IB Course D Mechanics of Materials and Structures Lent 2014

DH69
This is not surprising we should expect that in the limit of a sample which is a linear chain of
atoms it will have the theoretical strength. The question is not why thin samples are so strong, but
why larger samples are so weak. The answer is due to the presence of flaws. There are two ways of
thinking about the effect of the presence of a crack in a material. Inglis considered how the
concentration of stress near the tip of a crack would enable the crack to extend whilst Griffith
considered whether it would energetically favourable for a crack to grow.

10.2 The Effect of Cracks Stress Considerations
10.2.1 Stress concentrations at an atomic scale













From this picture of stresses in a crystal it is clear that where a crack is present, an applied stress
will have to be taken up by other bonds. This stress is not redistributed equally but is concentrated
near to the crack tip. It is apparent that the extent of the stress concentration is dependent not only
on the crack length but also on whether (on a macroscopic scale) the crack is sharp or blunt.
IB Course D Mechanics of Materials and Structures Lent 2014

DH70
10.2.2 Stress concentrations at macroscopic scale
Inglis was able to calculate the stress concentration factor in a continuum material near to a
through-thickness elliptical crack. If the far-field stress is !
0
, the stress at the elliptical opening was
calculated to be:
!
0
(10.1)
where c is the half-length of the crack and R is its radius of curvature. Hence for a circular opening,
the stress locally is raised by a factor of 3.










If we apply this argument to a sharp rather than smooth hole, and assume that the diameter of the
crack tip is of the order of an atomic spacing, this gives us R = 0.1 nm. A surface crack one micron
in depth would then produce a stress concentration of 200 which leads to a more reasonable value
for the strength of glass. However, windows can often contain cracks which are on a cm scale
(producing a concentration factor of 20,000) and large structures such as ships and bridges can have
cracks and openings on even larger length scales. Hence if the stress concentration were the whole
story, materials would end up being weaker than they actually are. The fact is that the stress
concentration is only a mechanism to break a material apart. To work out whether a material will
fracture, we also need to consider whether it is energetically favourable for it to do so.
IB Course D Mechanics of Materials and Structures Lent 2014

DH71
10.3 Energy Considerations
10.3.1 Contributions to the total energy of the system
When a material is subject to a stress, there is some mechanical energy available which could
potentially cause fracture. If the extension of a crack decreases the mechanical energy of the
system, that is clearly a driving force for crack growth and subsequent fracture. However, growth of
a crack produces new surfaces, which have an associated energy. This provides some energetic
resistance to crack growth and fracture. By considering these energies, we can work out whether it
will be favourable for the crack to extend.

10.3.2 The change in mechanical energy of the system, U
M

The mechanical energy of the system can have two parts there is an elastic strain energy stored in
a body under stress, but if the force which is applied is able to do work, we must consider that also.
There are two ways to calculate the value of the change in mechanical energy (driving force for
crack extension). It is possible to put a material in tension with a fixed extension (a simpler
analysis) or apply a fixed load (a more common scenario). These are equivalent in terms of the total
change in mechanical energy (though not in terms of the change in elastic strain energy stored)
when a crack is introduced.

i) Strained material under constant extension









IB Course D Mechanics of Materials and Structures Lent 2014

DH72








ii) Strained material under constant load

















IB Course D Mechanics of Materials and Structures Lent 2014

DH73
Hence we simply need to work out the change in elastic strain energy in a material of Youngs
modulus, E, subjected to a far-field stress ! if it contains an interior crack of length 2c. This is
derived from Inglis analysis and is:
(10.2)

This term is negative and always results in a decrease in the total energy of the system.

We can see qualitatively why the term scales with the square of the crack length by thinking about
the volume of the region in which the stress is relieved.









10.3.3 The energy required to make new, cracked surface, U
S

This is more straightforward the work required to create the 2 new surfaces for an internal crack
of length 2c is:
(10.3)

where " is the amount of energy needed to create 2 new surfaces, each of unit area. This term is
positive (creates new surfaces) and always results in an increase in the total energy of the system.

IB Course D Mechanics of Materials and Structures Lent 2014

DH74
10.3.4 The total energy change as a function of crack length
Combining the mechanical and surface contributions, the change in energy (per unit thickness of
sample) due to the presence of a through thickness crack of length 2c is:
(10.4)
We can see that extending very small cracks actually causes the overall energy of the system to
increase because the creation of new surfaces outweighs the benefit of relieving some elastic strain.
There is then an unstable equilibrium a value of crack length for which the energy used in
creating new surfaces is exactly balanced by the effect of strain energy relief. Any crack which is
larger than the equilibrium length is favoured to grow and will grow catastrophically as the energy
taken to create the new surfaces is more than made up for by the release of strain energy.











It is then straightforward to work out the equilibrium crack length by differentiating:
for c = c
e
. (10.5)

!
c
e
=
E"
#$
2
(10.6)


IB Course D Mechanics of Materials and Structures Lent 2014

DH75
We can plot the gradient (energy per unit crack length) as a function of c:















Clearly we could plot both G and " as positive, in which case
the equilibrium crack length is found where the lines cross.
Recall that " is the energy (per unit area of crack) which is resisting fracture. So we can define G
as the energy (per unit area of crack) driving fracture such that there is a critical value G
c
which
corresponds to G
c
=".
G is often called the strain energy release rate. It is the amount by which the mechanical energy
driving the system varies per unit area of crack (not per unit time).

c
c
e

2"
IB Course D Mechanics of Materials and Structures Lent 2014

DH76
11 Stable Crack Growth
11.1 Fracture toughness
We see from the previous section that there are two materials parameters which determine the stress
at which fracture occurs for a given crack size (or equally which determine the critical crack size
which will propagate for a given tensile stress). So we can write the relationship between the stress
and the crack size as:
(11.1)



!
K
c
=" #c
e
MPa m
( )
(11.2)
So we have now defined a materials property, K
c
as a value which depends on the stiffness and
fracture energy and which tells us about the critical relationship between crack length and stress. In
fact, more strictly we have only really looked at one particular loading geometry (known as mode I
loading) and so we call this value K
Ic
. This mode, is illustrated below, along with other possible
crack opening modes.






K is often described as the stress intensity factor, such that K
c
tells us when the stress reaches a
critical value. Whilst this is evidently true, it is fundamentally based on calculating the energy of
the system subjected to a given stress. Typically G
IIc
>> G
Ic
(and hence K
IIc
>> K
Ic
) since there is
normally a lot of frictional sliding between the crack flanks during mode II loading (i.e. the crack
tip is shielded from much of the applied load). The compressive failure stress of a brittle material is
therefore much higher than the tensile failure stress.
IB Course D Mechanics of Materials and Structures Lent 2014

DH77

11.2 The Wedging Geometry
In 1930, Obreimoff conducted an experiment with mica, where plates of the material were cleaved
apart by inserting a wedge. This clearly creates new surface by propagating cracks, but the
geometry is significantly different from cracking in tension.









Consider a wedge of thickness h inserted beneath a thin layer of material (of thickness d) attached
to a block of unit width causing the thin layer to peel off.


11.3 Energy Considerations
11.3.1 The change in mechanical energy of the system, U
M

One contribution to consider is the work done by the force (equal to F multiplied by the distance
moved in the direction of its operation). F operates at the point where the wedge meets the beam
and does not therefore move in the direction of its operation (it moves sideways) so this
contribution is zero.

IB Course D Mechanics of Materials and Structures Lent 2014

DH78
The only contribution to the mechanical energy is therefore elastic strain energy in the peeling layer
from C to the point where the wedge touches the layer (O):
(11.3)














IB Course D Mechanics of Materials and Structures Lent 2014

DH79
11.3.2 The change in surface energy of the system, U
S

Just as with fracture in tension, breaking the adhesive bond between the mica plates causes an
increase in energy of the surfaces
(11.4)

11.3.3 The change in total energy of the system as a function of crack length
(11.5)
The variation of the individual terms and that of the sum are shown in the figure below.

This equilibrium point (the turning point) can be found by differentiating the overall energy, U(c),
with respect to c and finding the value of c where this differential is equal to zero. That is where
(11.6)
(11.7)
(11.8)
which gives the equilibrium crack length, c
e
, as
(11.9)
IB Course D Mechanics of Materials and Structures Lent 2014

DH80
If E, d and h are known, c
e
can be measured to give a value for ".










!
" =
3Ed
3
h
2
8c
e
4
#
$
%
&
'
(
=
3)2)10
11
)10
*9
)8
2
)10
*6
8)10
*4




11.4 Crack Healing
If our system is truly thermodynamically reversible the crack tip should move backwards (that is the
crack should heal) if we pull the wedge out. To investigate this, Obreimoff cleaved crystals of mica.
On pushing the wedge in he found that the crack grew at a constant distance ahead of the crack tip,
as predicted, and estimated a value of " for mica close to that measured elsewhere. If he tested the
sample in vacuum he found that the crack would heal as he withdrew the wedge.
However when the sample was tested in air this did not occur suggesting that chemical groups
attach themselves to the surface of the mica, lowering, if not eliminating, the ability of the crack to
heal if the wedge is withdrawn. Obreimoff's experiment clearly demonstrated the reversible,
thermodynamic nature of cracking and that the principle outlined by Griffith was correct. All
modern theories of fracture come from this basic idea.
IB Course D Mechanics of Materials and Structures Lent 2014

DH81
11.5 The Nature of Cracking
11.5.1 The importance of equilibrium
The total energy of a cracked body under load, U(c), is
(11.10)
where U
M
is the change in mechanical energy of the system (a combination of the work done by the
applied force and the change in elastic energy on cracking) and U
S
is the work required to create
two new surfaces. Equilibrium will occur when
(11.11)
or
(11.12)
From our expression for cracking in tension
(11.13)
and
(11.14)

Hence
(11.15)


So equilibrium will occur when
(11.16)
We could perform a similar operation for the expression we derived for wedging.
IB Course D Mechanics of Materials and Structures Lent 2014

DH82
11.5.2 Will failure be catastrophic?
It is incorrect to say that failure will necessarily occur when

There will be some cracking but remember that complete failure also requires that
(11.17)
or
(11.18)

In other words failure will be catastrophic when the rate of increase of the driving force with crack
growth is greater than the rate of change in " with crack growth.
So far we have taken " as a constant, so has been zero and any crack propagation would be
catastrophic when >0 (which is always the case when cracks propagate in tension).


But cracking would be stable (even in tension) if
(11.19)
i.e.
(11.20)
So if we can make " increase with crack length, we should be able to get stable crack propagation,
even in tension as long as

IB Course D Mechanics of Materials and Structures Lent 2014

DH83
12 Coping with a Scatter in Strength
12.1 Introduction
Virtually all materials (even metals) fail by cracking. Normally this occurs due to tension or
bending loads, where the equilibrium is unstable and the failure catastrophic. Such cracks normally
develop from pre-existing flaws. However the Griffith expression shows that the stress required for
failure in tension is dependent on the size of the largest flaw, c, according to .

Consider materials subjected to stresses of ~100 MPa
Metals:


Ceramics:



How do we cope with this?
A way around this problem is to determine how likely failure will be under a given applied stress by
testing samples from a given batch of material. To design the component, one must decide what is
an acceptable probability of failure and then set the stress to the associated value.



Having set the tolerable failure level, one then estimates the stress which will produce this
probability of failure and use this as the design strength. You have to accept that a given number are
likely fail. This is very different from designing with a metal, where one has confidence that the
strength will have a given value. The scatter in strength is described using Weibull statistics.

IB Course D Mechanics of Materials and Structures Lent 2014

DH84
12.2 Weibulls Approach
The Weibull approach considers a chain of N links. So the survival of the chain under load requires
that ALL the links survive.







(12.1)
How does the probability of survival/failure of a single link depend on applied stress?

(12.2)
(12.3)
(12.4)
(12.5)
hence
!
lnln
1
S
"
#
$
%
&
'
= mln
(
(
0
"
#
$
%
&
'
+ constant (12.6)
IB Course D Mechanics of Materials and Structures Lent 2014

DH85
Clearly when % = %
0
, S
L
= 1/e. Hence %
0
is the value of % that 37% of the links can be subjected to
without failure.
12.3 Using Weibulls Method
To determine the survival probability associated with each stress we start by testing some samples,
in this case some silicon carbide bars. The strengths are (in MPa)
949, 985, 921, 771, 779, 653, 945, 732, 858
This gives a mean strength of 843 MPa. %
0
= 921 MPa.
First organise the data into descending order with the highest strength being 1, the next highest 2
and the n
th
highest n. A probability of survival, S
n
, is then assigned to each value so that
(12.7)
where n
T
is the total number of samples. If a load is applied corresponding to the highest observed
strength then there would be a relatively low probability of the material surviving, in this case a 1 in 10
chance. If the load corresponds to the lowest strength observed then there is a much higher chance 9
out of 10 of the component surviving. Using eqn (12.6) a plot of against ln(! /!
0
) should
give a straight line with a slope m. Using this relationship it is now possible to work out the value of the
stress corresponding to the survival (or failure) probabilities that are appropriate for the design we have
in mind.
The strength data arranged to estimate the Weibull modulus
n ! (MPa) S=n/n
T
+1 ln (!/!
0
) ln ln (1/S)
1 985 0.1 0.067 0.83403
2 949 0.2 0.030 0.47588
3 945 0.3 0.026 0.18563
4 921 0.4 0 0.08742
5 858 0.5 -0.071 0.36651
6 779 0.6 -0.167 0.67173
7 771 0.7 -0.178 1.0309
8 732 0.8 -0.230 1.4999
9 653 0.9 -0.344 2.2504

IB Course D Mechanics of Materials and Structures Lent 2014

DH86











And we can extrapolate this data back to an acceptable value of S.










For S = 0.99999 (F = 10
-5
): lnln(1/S) = -11.5


This usable strength (so that only 10 per million fail) is much less than the average.
IB Course D Mechanics of Materials and Structures Lent 2014

DH87
12.4 Practical Solutions
If a component is being manufactured to support a given load, there are a number of possible ways
in which to ensure that fewer components fail in service.
1. Redesign
This is rather obvious, but by increasing the cross-sectional area a given load can be supported
under a lower stress. Drawbacks include:




2. Change material
It may be possible to use a material with a higher average failure strength or a larger Weibull
modulus, but this may involve increased cost and there may be other materials properties which
cannot be met. Alternatively it may be possible to toughen the material (increasing its fracture
energy) by producing a composite.

3. Improve processing
If the component is a ceramic then the processing route will play a large part in determining the
flaw size. By changing the processing method (i.e. hot isostatic pressing at larger pressures, higher
temperatures or for longer times) it may be possible to remove the larger flaws.

4. Throw some components away
It may seem odd but the most cost-effective method is simply to throw away (or better, recycle) a
number of the components. If all the components are proof tested up to the stress which they will be
subjected to in service, a small number of them will fail, but these will be the ones which had the
largest flaws and the remainder will have been shown not to contain flaws which propagate under
the service load.
IB Course D Mechanics of Materials and Structures Lent 2014

DH88
12.5 Toughening
12.5.1 Crack bridging
It is commonly observed that cracks can grow stably in a structure over a period of time. This
happens even though the body is being loaded in an unstable manner (i.e. in tension), so that we
would expect any cracking to occur catastrophically. Oddly this phenomenon tends to occur in
tougher materials. This suggests that toughening does not simply increase the magnitude of the
fracture energy but changes the way in which a crack grows.
Consider a material toughened by crack bridging in which intact ligaments across the crack faces
are left behind as the crack grows. Toughening occurs because separating the crack faces then
requires that extra work is done in order to either stretch the ligament or pull it out of the matrix in
which it is embedded. The fibres are normally discontinuous.



IB Course D Mechanics of Materials and Structures Lent 2014

DH89
12.5.2 R-curves
Bridging leads to a change in " with c. This type of behaviour is known as R-curve behaviour. We
can see that as we load the material containing a small flaw, it will begin to grow (under an
increasing applied crack driving force) until the process zone is fully developed. Once the zone has
developed fully then the whole crack will move forward with the process zone size remaining a
constant size. As process zones exist in all toughened materials (that is where " > 2&), we might
expect that they would all show R-curves and this is the case as shown below.

























IB Course D Mechanics of Materials and Structures Lent 2014

DH90




















By controlling the microstructure and properties of the material to give R-curve behaviour then we
can (over certain limits of flaw size) ensure that cracks grow in a stable manner even though the
loading state would normally have caused catastrophic failure.
"
"
c
c
IB Course D Mechanics of Materials and Structures Lent 2014

DH91
Appendix
Summary of Materials Response in Compression and Tension
A.1 In Compression
A.1.1 Structures
If compression elements fail by crushing then the behaviour really just depends on the cross-
sectional area (and the compressive strength), but if it occurs by buckling then the failure stress
depends on (i) the slenderness ratio and (ii) the elastic modulus of the beam








Hence if we decide to change the:
Cross sectional shape:
doesnt affect crushing
subdividing is bad for buckling
hollow tubes help prevent elastic buckling (but worry about local buckling)

Length:
Making longer makes buckling more likely
L/R
compressive failure by
crushing or plastic
yielding
Elastic
buckling
IB Course D Mechanics of Materials and Structures Lent 2014

DH92
A.1.2 Materials
If the structure is short and stocky, we need to design against crushing/yielding. Compressive
strengths of ceramics are very high and so bricks, stone and concrete are a good (cheap!) option as
long as weight is not a factor.




If the structure has a large aspect ratio we need to be more careful.
Difficult to improve materials resistance to buckling (just depends on E).
Use stiff material, such as ceramic (e.g. stone pillars in buildings)
-make it wide enough to have a large safety margin
If applied load is intermittent, may be OK if material buckles and recovers. (e.g. grass)






A.2 In Tension
A.2.1 Structures
If we decide to change the:
Cross sectional shape:
Subdivide no adverse effect on tensile failure can use wire instead of rod
In brittle materials, subdividing is good to prevent possibility of large flaws

IB Course D Mechanics of Materials and Structures Lent 2014

DH93
Length:
Longer structures are more likely to contain significant cracks
(Weibull analysis- more links)

Tensile strengths are usually much larger than those which would cause buckling, hence it is
usually most efficient to use primarily tensile elements.

A.2.2 Materials
Metals
large ', less susceptible to fast fracture than ceramics.
Composites
Addition of fibres improves toughness.

A.3 Conclusions
Tension elements tend to be long and sub-divided. Compression elements tend to be shorter and
made from stiffer materials. It is what we see in structures around us. Think of the difference
between a compression structure such as a stone bridge and a structure with more tensile elements
such as the Viaduc de Millau.




IB Course D Mechanics of Materials and Structures Lent 2014

DH94
Nomenclature
Symbols
A cross-sectional area (m
2
)
c crack length (m)
constant in elastic buckling expression which
depends on end-constraints (-)
d (inner) diameter or linear dimension (m)
D (outer) diameter or linear dimension (m)
eij relative displacement tensor (-)
E Youngs modulus (Pa ( N m
2
)
F force (N)
probability of failure (-)
G strain energy release rate (J m
2
)
shear modulus (Pa ( N m
2
)
h height of beam (m)
I second moment of area (m
4
)
k constant (various)
K stress intensity factor (Pa m
1/2
)
bulk modulus (Pa ( N m
2
)
L distance or length (m)
m Weibull modulus (-)
M bending moment (N m)
N number of chain links (-)
number of cycles to failure (-)
P pressure (Pa)
r radial distance or radius (m)
R radius of curvature (m)
radius (m)
s aspect ratio (-)
S probability of survival (-)
stress amplitude (Pa)
t thickness (m)
time (s)
T temperature (K or C)
torque (N m)
U energy (J)
v velocity (m s
-1
)
V volume (m
3
)
w width (m)
W energy (J)
x distance (along a principal axis) (m)
y distance (along a principal axis) (m)
z distance (along a principal axis) (m)
& (engineering) shear strain (-)
surface energy (J m
2
)
" energy resisting crack growth (J m
-2
)
! deflection (m)
) (normal) strain (-)
)ij strain tensor (-)
)N nominal strain (-)
)T true strain (-)
* angle (rad. or )
" curvature (m
1
)
+ Poisson ratio (-)
, density (kg m
3
)
% normal stress (Pa ( N m
2
)
%ij stress tensor (Pa ( N m
2
)
%N nominal stress (Pa ( N m
2
)
%T true stress (Pa ( N m
2
)
- beam stiffness or flexural rigidity (N m
2
)
. shear stress (Pa ( N m
2
)
/ angle (rad. or )

Subscripts
0 base level
average
1 first, in direction 1 etc
I, II loading mode
c critical
e equilibrium
EB Elastic buckling
E stored elastically
F due to movement of a force
failure
i in the i direction
j in the j direction
L link
LB Local buckling
M mechanical
P polar
r in radial direction
S surface
T total
Y yield
x in x direction
y in y direction
z in z direction
* in hoop direction
IB Course D Mechanics of Materials and Structures Lent 2014 DH95

Glossary
Words in italics provide cross-references to other entries in the glossary.
Anti-symmetrical (2
nd
Rank Tensor). Tensor for which the values of the off-axis components (Sij,
with i0j) are reflected across the diagonal, and their signs changed, so that Sij = -Sji.
Beam Stiffness, EI. Product of Youngs modulus and second moment of area, characterising the
resistance to deflection when subjected to a bending moment. Sometimes termed the flexural
rigidity.
Bending Moment, M. Turning moment generated in a beam by a set of applied forces. The
bending moment is balanced at each point along the beam by the moment of the internal
stresses.
Brittle Fracture. Fracture not involving gross plastic flow (although some local plastic
deformation may occur at the tip of the crack).
Bulk Modulus, K. Constant of proportionality between hydrostatic stress and hydrostatic strain,
for an isotropic body.
Compliance Tensor, Sijkl. Tensor of 4
th
rank reflecting the elastic properties of a material. The
product with the stress tensor is the strain tensor.
Crushing. A compressive failure mode for a brittle material, arising from propagation of cracks
due to shear stresses.
Curvature, ". Reciprocal of the radius of curvature adopted by a beam subject to a bending
moment. Also equal to the through-thickness gradient of strain in the beam.
Deviatoric Component. Component of a second rank tensor representing the shape change (in the
case of strain) or the driving force for shape change (in the case of stress).
Diagonalising (a 2
nd
Rank Tensor). Transforming a tensor so that it is referred to the principal
planes, and hence has non-zero terms only along the diagonal.
Dilation, $. Volume change associated with a strain, given by the sum of the diagonal
components of the strain tensor.
Driving Force for Crack Extension. Any contribution which would decrease the total energy of a
system if a crack were to propagate, e.g. mechanical energy (a negative value if the energy
decreases).
Ductile Fracture. Fracture involving gross plastic flow.
Einstein Summation Convention. Convention relating to abbreviated formulation of sets of
equations, stating that, when a suffix occurs twice in the same term, then this indicates that
summation (from 1 to 3) should be carried out with respect to that term.
Elastic Buckling. See Euler Buckling.
IB Course D Mechanics of Materials and Structures Lent 2014 DH96

Engineering Shear Strain, &ij. Distortional deformation (angle) of a body arising from a shear
stress (actually a pair of shear stresses).
Equilibrium Crack Length. The crack length at which (under a given loading scenario) the
energy of the system is a maximum (for an unstable cracking scenario) or a minimum (for a
stable cracking scenario).
Euler Buckling. An elastic failure mode, whereby a thin column in compression becomes unstable
against lateral perturbations.
Fatigue. A common failure mode in materials subjected to stresses which vary with time. A crack
may propagate gradually over millions of cycles before the reduced cross sectional area of the
component leads to sudden fracture.
Flexural Rigidity. See beam stiffness.
Fracture Energy, Gc. Energy per unit area required to extend a crack. Also termed the critical
strain energy release rate. See surface energy.
Fracture Toughness Kc. Critical value of the stress intensity factor, characterising the resistance of
a material to crack propagation.
Griffith Criterion. Energy-based condition for fracture.
Hookes Law. Stress is proportional to strain during elastic deformation.
Hydrostatic Component. Component of a second rank tensor representing the volume change (in
the case of strain) or the driving force for volume change (in the case of stress).
Invariants (of the Secular Equation). Coefficients of the secular equation, which are independent
of the choice of axes. These combinations of the components of the tensor therefore remain
unchanged during transformation of axes.
Isotropic Material. A material in which the properties of interest are the same in all directions
within it.
Local Buckling. A failure mode for hollow cylinders in compression. If the wall thickness is
small, the walls may crumple under a lower stress than is required for either elastic buckling
or compressive failure. An empty soft-drinks can is a good example.
Mechanical Energy (of a Loaded Body). The total mechanical energy of the system arising from
the stored elastic energy within the material plus the energy contribution to the system if the
load moves and does work .
Mohrs Circle. Geometrical construction for solving the secular equation, applicable only in a
principal plane.
Neutral Axis (of a beam). Axis (strictly, a plane) parallel to the length of a beam, along which
there is no change in axial length on bending.
Nominal (or engineering) strain. The strain based on the original length of the sample.
IB Course D Mechanics of Materials and Structures Lent 2014 DH97

Nominal (or engineering) stress. The stress based on the original cross-sectional area of the
sample.
Normal Strain, )ii. Deformation in which change in length is parallel to original length.
Normal Stress, %ii. Stress induced by a force acting normal to the sectional area to which it is
applied.
Poisson Ratio, #. Ratio of the transverse contraction to the axial extension induced in an isotropic
body by an applied normal stress.
Polar Second Moment of Area, IP. Parameter dependent on sectional shape, characterising the
resistance a body offers to twisting under an applied torsional moment, or torque.
Principal Directions. Normals to the principal planes.
Principal Planes. A unique set of orthogonal planes, on which there are no shear stresses acting
(or on which there are no shear strains).
Principal Strains, )i. A unique set of normal strains, acting on the principal planes, which
represents the strain state of a body.
Principal Stresses, %i. A unique set of normal stresses, acting on the principal planes, which
represents the stress state of a body.
R-curve. A plot of the fracture energy vs. crack length in systems where the fracture energy
increases (towards some upper limit) as the crack grows.
Relative Displacement Tensor, eij. Second rank tensor giving the displacement of points within a
body.
Resistance to Crack Extension. Any contribution which would increase the total energy of a
system if a crack were to propagate, e.g. surface energy (a positive value if the energy
increases).
Rotation Tensor, $ij. Anti-symmetrical second rank tensor representing the rigid body rotation
component of the relative displacement tensor.
Second Moment of Area, I. Parameter dependent on sectional shape, characterising the resistance
a beam offers to deflection under an applied bending moment. Also called the moment of
inertia.
Secular Equation (of a 2
nd
Rank Tensor). A cubic equation, the roots of which are the principal
values of the tensor.
Shear Modulus, G. Constant of proportionality between shear stress and (engineering) shear
strain, when an isotropic body is subjected to a single shear stress (actually a pair of shear
stresses, although this is not taken into account in the definition).
Shear Stress, %ij. Stress induced by a force acting parallel to the sectional area to which it is
applied.
IB Course D Mechanics of Materials and Structures Lent 2014 DH98

Slenderness ratio. The ratio L/R for a column of length L and radius R.
Stiffness Tensor, Cijkl. Tensor of 4
th
rank reflecting the elastic properties of a material. The
product with the strain tensor is the stress tensor.
Strain Energy Release Rate, G. Elastic strain energy released per unit of created crack area.
Strain Tensor, %ij. Symmetrical second rank tensor representing the strain component of the
relative displacement tensor.
Stress Intensity Factor, K. Parameter characterising the crack driving force, in terms of applied
stress level and crack length.
Surface Energy, &. The energy, &, associated with one square metre of surface. Note that the
fracture energy, ", of an ideally brittle material is the energy required to form two new
surfaces, each of unit area. Hence " =2&.
Symmetrical (2
nd
Rank Tensor). Tensor for which the values of the off-axis components (Sij, with
i0j) are reflected across the diagonal, so that Sij = Sji.
Tensor (of nth Rank). Array of values (with n dimensions) representing either an imposed field
(eg velocity, thermal gradient, stress etc) or properties of a specified type of matter (eg
stiffness, conductivity etc).
Torque, T. Twisting moment generated in a body by a set of applied forces, which tend to rotate
the body about an axis lying within it.
Torsional Rigidity. Product of shear modulus and polar second moment of area, characterising the
resistance to twisting when subjected to a torque.
Toughening. Some mechanism which increases the resistance to crack extension. Examples are
crack bridging and plastic deformation.
Transform Matrix. Array of terms which, when concatenated with a tensor, effects a
transformation of axes operation on the tensor.
Transformation of Axes (of a Tensor). Re-evaluation of the components of a tensor, so that they
relate to a different set of reference axes.
True Stress. The stress based on the instantaneous cross-sectional area of the sample.
True Strain. The strain based on the instantaneous length of the sample.
Twisting Moment. See torque.
Wedging. A loading geometry where a wedge exerts a force in a direction normal to the direction
in which it moves.
Weibull Modulus. A parameter which quantifies the variability in failure strengths of components.
IB Course D Mechanics of Materials and Structures Lent 2014 DH99

Weibull Statistics. A method of predicting the failure strength distribution of brittle components.
As there will always be some range of failure strengths, Weibull statistics allow us to predict
the probability of failure at a given stress.
Youngs Modulus, E. Constant of proportionality between normal stress and normal strain, when
an isotropic body is subjected to a single principal stress.

DQ1 IB Materials Science - Course D: Mechanics of Materials and Structures DQ1
-1- PDB: LT14

Question Sheet DQ1

1. At a particular location on the earths surface, magnetic north is 10 E of true north. Write
down a 3!3 transform matrix for converting a velocity measured using a compass to
establish its direction to one which is specified relative to the true geographical directions.
2. A metallic beer can, of radius r and wall thickness t (<<r) contains liquid under a pressure
P. Using the treatment presented in IA (course D) to evaluate the hoop and axial stresses in
such a situation, show that the stress tensor representing the stress state in an element of the
wall can be expressed
!
ij
=
!
1
0 0
0
!
1
2
0
0 0 0
"
#
$
$
$
$
%
&
'
'
'
'

where "
1
is given by Pr/t. Sketch Mohrs circle for each of the 3 principal planes. Hence
derive an expression for the peak shear stress experienced by the material of the can and, on
a sketch of a volume element, indicate the planes on which it acts. If t = 0.3 mm,
r = 30 mm and the shear yield stress of the material is 15 MPa, what pressure would be
needed to cause yielding?
On a hot day, and after being agitated, the pressure in the beer can reaches 2 bar (about two
thirds of that needed to cause yielding). It is then subjected to a torque of 15 N m, applied
about the axis of the can. Calculate the new value of the peak shear stress in the can wall
and decide whether the material will yield under these new conditions.
3. (a) Separate the following tensors into symmetrical and antisymmetrical components:



If these tensors were describing general strains, what would be the significance of the
symmetrical and antisymmetrical components?
(b) The stress tensor is symmetrical. Separate the following stress tensors into hydrostatic
and deviatoric components:



(c) Diagonalise the second stress tensor given in part (b). Separate the diagonalised tensor
into hydrostatic and deviatoric components. Show that the deviatoric stress can be
expressed as the sum of three pure shear stresses.
DQ1 IB Materials Science - Course D: Mechanics of Materials and Structures DQ1
-2- PDB: LT14

4. (a) Describe what is meant by a principal stress. If "
3
is a principal stress write down a
secular equation which can be used to determine the two other principal stresses, "
1
and
"
2
. Solve this equation to show that



where "
ii
and "
ij
are normal and shear stresses respectively.
(b) Describe a possible graphical interpretation of the above solution and how it can be
used to characterize a state of plane stress.
(c) An elastic material, in a two-dimensional stress state, has normal stresses of 85 MPa
(tensile) and 35 MPa (compressive) acting on two mutually perpendicular planes. A
shear stress also acts on these planes. Of the two principal stresses present the largest one
is limited to 110 MPa (tensile). Use Mohrs circle to determine (i) the shear stress (ii) the
second principal stress (iii) the normal stress on the plane of maximum shear (iv) the
maximum shear stress and (v) the orientation of the principal planes. Confirm the results
analytically.
[modified Tripos 2012]

DQ2 IB Materials Science - Course D: Mechanics of Materials and Structures DQ2
PDB: LT14 1

Question Sheet DQ2

1. (a) Show how the following stress states can be represented using Mohrs circles and write
down the corresponding stress tensors: (i) uniaxial compression (ii) pure shear and (iii)
hydrostatic pressure.

(b) Use the relative displacement tensor and suitable sketches to distinguish between states of
pure shear and simple shear and show how this difference affects the definition of the shear
modulus.

(c) Express a general stress tensor, referred to principal axes, in terms of its hydrostatic and
deviatoric components. Show that the deviatoric stress can be written as the sum of three pure
shear stresses.

(d) Determine the pure shear components of the following stress tensor and show that the
hydrostatic component is zero!
1 2 0
2 1 0
0 0 !2
"
#
$
$
$
%
&
'
'
'

[modified Tripos 2013]

2. (a) Define the terms principal stress and principal strain. Describe how these quantities are
computed from a general stress or strain state. Determine the principal stresses of the following
stress state:
!
"
ij
=
" " "
" " "
" " "
#
$
%
%
%
&
'
(
(
(


(b) Consider an isotropic material subjected to a set of principal tensile stresses !
1
, !
2
and !
3
.
Show that the resulting principal strains are given by

!
"
1
=
1
E
#
1
$% #
2
+#
3
( )
( )
"
2
=
1
E
#
2
$% #
1
+#
3
( )
( )
"
3
=
1
E
#
3
$% #
1
+#
2
( )
( )

where E is Youngs modulus and " is Poissons ratio.


DQ2 IB Materials Science - Course D: Mechanics of Materials and Structures DQ2
PDB: LT14 2

(c) An isotropic material is subjected to a compressive load in the 1-direction but restrained
from extending in the 2 and 3-directions. Show that the Youngs modulus of the material
effectively changes from E to
!
E 1"
2#
2
1"# ( )
$
%
&
'
(
)


(d) A uniform bar of isotropic material 10 cm long and 2 cm square is loaded in compression
along its length. If the bar is laterally restrained in the plane normal to the load, determine the
change in length of the bar resulting from a load of 40kN. (For the material E = 100 GPa, " =
0.3).
[modified Tripos 2011]

3. Determine an expression for the second moment of area of the semicircular section shown
below with respect to the neutral axis (NA) parallel to its base.


4. (a) When a uniform beam carries a transverse load it experiences a bending moment and
deflects. Derive a relationship between the bending moment, the beam curvature and the beam
stiffness assuming the defection is small.

(b) A uniform cantilever of length L and stiffness EI (E = Youngs modulus, I = moment of
inertia), is loaded at its free end as shown in the schematic diagram below. Show that the
deflection along the beam is given by

!
y =
F
EI
x
3
6
"
L
2
x
2
+
L
3
3
#
$
%
&
'
(


DQ2 IB Materials Science - Course D: Mechanics of Materials and Structures DQ2
PDB: LT14 3

(c) Where along its length does the beam exhibit its greatest slope (
!
dy dx) and its greatest
curvature (
!
d
2
y dx
2
) and state their values?

(d) If the cantilever is now loaded at its mid-point as shown below, derive an expression for the
deflection at the free end. Note that the relationship derived in (a) now only applies to the left
hand side of the beam.



[Tripos 2011]




DQ3 IB Materials Science - Course D: Mechanics of Materials and Structures DQ3

PDB: LT14
1
Question Sheet DQ3

1. (a) Determine an expression for the second moment of area of the equilateral triangle
shown below with respect to its base (the x-axis).



(b) Consider three freely hinged, unconstrained, columns of the same material and
length but with the following cross-sections (i) a circle (ii) a square and (iii) an
equilateral triangle. If the volume of each column is the same show that the critical
compressive load for elastic buckling of the square and triangular columns is about
5% and 21% greater respectively than that of the circular column.

(c) Three identical columns, which have cross-sections that are equilateral triangles,
are placed together without bonding as shown below and loaded axially in
compression. Assuming freely hinged, unconstrained, conditions show that the
critical load for elastic buckling increases by a factor of five if the columns are
bonded along their lengths.













[Tripos 2012]


DQ3 IB Materials Science - Course D: Mechanics of Materials and Structures DQ3

PDB: LT14
2
2. (a) Define the terms bending moment and beam stiffness. Show that the deflection
y(x) along the axis of a uniform beam subject to a bending moment M is given by

!
y(x) =
M
EI
dx
"
#
$
%
&
'
((
dx + Ax + B
where E is Youngs modulus, I is the second moment of area and A and B are
constants. State the conditions under which this expression is valid and describe how
the constants are determined.

(b) A simply supported uniform beam of length L carries a central concentrated load
F as shown below.









Using the centre of the beam as the origin show that the deflection along its length is
given by
!
y(x) =
F
EI
x
3
12
"
Lx
2
8
+
L
3
48
#
$
%
%
&
'
(
(

Sketch the variation of bending moment and curvature along the length of the beam.

(c) If the loading configuration in (b) is changed to that shown below, write down
expressions for the bending moments along the regions of the beam defined by 0 ! x
! L/6 and L/6 ! x ! L/2.








Hence sketch the new variation of bending moment along the beam. How will this
variation change as each load is successively divided in half and configured
symmetrically along the beam?
[Tripos 2012]
DQ3 IB Materials Science - Course D: Mechanics of Materials and Structures DQ3

PDB: LT14
3

3. A solid cylindrical rod of radius 3 mm is tested in uniaxial compression with its ends
free to rotate.
(a) Describe the failure mechanism of the rod for (i) very short lengths (several
millimetres) and (ii) very long lengths (several metres).
(b) Stating your assumptions, calculate how the failure load depends on sample
length, and plot this on graph paper.
(c) The solid rod is replaced with a hollow rod which has an external diameter of 10
mm, but the same mass per unit length as the solid rod. Plot the failure load vs
length for this hollow rod on the same graph.
[Tripos 2011]
4. Consider a simply supported beam of length L which is subject to symmetrical 4-
point bending as shown below.





Show that the maximum deflection is given by


!
y
max
=
Fa
24EI
3L
2
" 4a
2
( )



DQ4 IB Materials Science - Course D: Mechanics of Materials and Structures DQ4

PDB: LT14
1
Question Sheet DQ4

1. (a) Explain, with an example, how sample size affects the average failure strength of
a brittle material in tension. Using Weibull statistics show that the probability of
failure increases exponentially with sample volume.

(b) If a rod of brittle material is subject to a transverse (bending) load rather than a
tensile load explain whether the average failure strength increases, decreases or
remains the same.

(c) Given below are the tensile strengths (in MPa) of a batch of graphite/epoxy
composite samples. Determine the Weibull modulus and comment on its value.
Estimate the usable strength of the composite if the tolerable failure rate in practice
must be 1 in 10
4
.

509, 452, 571, 555, 536, 522, 498, 468, 477
(d) Estimate how the mean strength of the samples would change if their volume
doubled or halved.
[Tripos 2013]
2. (a) Explain why two samples fabricated from the same brittle material using an
identical process can fail at different tensile loads.
(b) Fifty samples were tested to failure in uniaxial tension by loading them in
increments of 100 g. The accumulated mass at which each sample failed is listed
below. Calculate the mean failure load.
Accumulated mass / g No. of samples which failed
700 0
800 4
900 5
1000 7
1100 10
1200 11
1300 8
1400 4
1500 1
(c) Make a Weibull plot of the data and calculate the Weibull modulus.
DQ4 IB Materials Science - Course D: Mechanics of Materials and Structures DQ4

PDB: LT14
2
(d) Estimate the load for which the probability of failure is a) 0.1 and b) 0.001.
Comment of the relative confidence in these estimates.
(e) A new batch of samples is fabricated using the same processing method. They
have the same cross-sectional area, but are three times longer. Explain whether
the average tensile strength is likely to increase, decrease or remain the same.
Estimate the median tensile failure load of these longer samples.
[Tripos 2009]

3. (a) Explain why, under certain conditions, a crack grows so that the body fails
catastrophically, whereas under other conditions crack growth may start and then
stop.
(b) A wedge 500 m thick is used to cleave a surface coating, with a Youngs
modulus of 200 GPa, off a thick, square substrate with a side length of 40 mm.
Assuming that the fracture energy of the interface between the substrate and the
coating is 1 Jm
-2
, show how increasing the coating thickness from 200 m to 1
mm might influence the way the crack is seen to grow.
[Tripos 2004]
4. (a) Explain how crack bridging gives rise to toughening and comment on the
microstructural features that influence the magnitude of the toughening.
(b) A material contains elongated grains 10 m in length that act as bridging
ligaments and has a fracture energy (!) of 250 Jm
-2
and a Youngs modulus (E) of
380 GPa. Stating your assumptions, estimate the length of the zone where the
ligaments bridge the crack. [In a homogeneous material, the crack opening (2u) is
related to the distance from the crack tip (r ) by the expression

!
u =
4K
E
r
2"
#
$
%
&
'
(
1/ 2
where
K is the fracture toughness].
[modified Tripos 2003]


IB Course D Mechanics of Materials and Structures Lent 2014

Lecture 1
___________________________________________________________________________________
___________________________________________________________________________________
___________________________________________________________________________________
___________________________________________________________________________________
___________________________________________________________________________________
___________________________________________________________________________________
___________________________________________________________________________________
___________________________________________________________________________________

Lecture 2
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________

Lecture 3
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
IB Course D Mechanics of Materials and Structures Lent 2014

Lecture 4
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________

Lecture 5
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________

Lecture 6
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________

IB Course D Mechanics of Materials and Structures Lent 2014

Lecture 7
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________

Lecture 8
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________

Lecture 9
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________

IB Course D Mechanics of Materials and Structures Lent 2014

Lecture 10
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________
____________________________________________________________________________________________

Lecture 11
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________

Lecture 12
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________
___________________________________________________________________________________________

Das könnte Ihnen auch gefallen