You are on page 1of 100

Page 1

SESG 1001 Mechanics of Solids


Lecture 1 and 2 Introduction to Solid Mechanics
Block 1: Statically Determinate Structures
0 Introduction
This module will provide the basic techniques required to analyze and design load-
bearing structures. In general, when designing structures we are concerned with ensuring
their “structural integrity”. This comprises three aspects of structural performance:
1) The structure should not deform excessively under load
2) The structure should not fail (break)
3) The structure should have sufficient life

Solid mechanics covers the analysis techniques to determine how structures transmit load
(force) and how they deform under load. There is a close link to Materials (SESG1003),
which covers aspects of elastic deformation, failure and factors that control life (fatigue,
creep).

Throughout SESG 1001 we will be invoking three basic principles (sometimes called the
“three great principles”):
1) Equilibrium of forces
2) Compatibility of displacements
3) Constitutive behavior of materials
We will see how these apply to several classes of structure during the module.

Finally, we will be constructing models for how structures behave. These are
idealizations (simplifications) of the real structure that enable us to analyze it. A key skill
for an engineer is to be able to develop models for physical situations that allow him/her
to design and analyze engineering systems. For the moment we will rely on the
experience and wisdom of those who have gone before us to examine existing models. In
due course we will start developing our own models.

1.0 Statically Determinate Structures (See BCA, Ch1)


These are structures for which the force distribution can be determined solely by applying
equations of static equilibrium. We will examine three examples during SESG1001: 2-D
trusses, bars of varying thickness and pressure vessels.

1.1 Trusses
A truss is a structural configuration consisting of bars connected at joints. Generally these
are 3-D, but we will only consider 2-D cases in SESG1001. This will start us thinking
about models for structures and how we can analyze internal forces in structures.

1.2 Idealized Planar, Pin-Jointed Truss


First define an idealized planar truss – this is the model
Page 2
1. All bars are straight
2. Joints between bars are frictionless pins
3. Bars are massless and perfectly rigid (for loading analysis)
4. All loads and reactions are applied at the joints.
5. Loads in members are co-linear (axial), i.e. aligned with long axis of bar.
Axial forces in bars also arise from (2).

Fbar = Fpin (by equilibrium)

1.3 Analysis of forces in pin-jointed trusses


At this stage we are interested in determining the internal forces carried by the bars of the
truss. In due course we will look at deformations and other factors related to
determining the structural integrity of trusses

Step 1. Determination of External Reactions


Draw Free Body Diagram and apply equations of equilibrium

EXAMPLE 1: A simple truss (3 bar truss). Note idealizations at support points: pins
and/or pins on rollers:

Draw free body diagram (FBD)


Page 3

Apply planar equations of equilibrium:


Equilibrium of forces in x1, x2 directions, equilibrium of moments about x3 axis (do not
worry about axis symbols for now: x1=x, x2=y, x3=z).

+
! F1 = 0 ""# $ HA + HB = 0 (1)

! F2 = 0 " + # VB + 200N = 0 (2)


VB = !200 N

! M3 = 0 A+ ⇒ 10 • 200 ! 5 HB = 0 (3)
HB = 400N

From (1)
H A = !400 N

Is this statically determinate? – so far yes, we have been able to evaluate all reactions
using equilibrium equations.
Three reactions apply to the three degrees of freedom (translation in x1 and x2 and
rotation about x3) give three equilibrium equations that we can write down.
Page 4
Redrawing FBD

Step 2. Analysis of forces in bars (Method of Joints)


Now we want to evaluate the forces in the bars of the truss. This requires us to isolate
part of the truss, replacing the cut bars by equivalent (equipollent) forces:

Draw FBD for joint C, replace bars by equivalent forces (FCB and FCA)
200 N
5
! = tan "1 = 26.6!
FCB 10
5 1 10 2
sin ! = = , cos! = =
θ 125 5 125 5
FCA

Apply equilibrium
+
!F1 ""# = 0 $ %FCB cos& % FCA = 0
!F2 ' + = 0 $ FCB sin& + 200 = 0
FCB = %200 ( 5 = %447N
NOTE: there is no opportunity to apply equilibrium of moments, all forces meet at C.
Page 5

Substitute in !F1
2
! "("200 5) " FCA = 0
5
⇒ FCA = +400N ⇐

FCB is positive i.e. tensile, pulls out of the joint, FCB is negative, i.e. compression, pushes
into the joint. Redraw joint:
200√5 N
200 N

400 N
Now consider joint, B:

200 N

400 N θ

FBC
FBA

Apply equilibrium

! F2 " + # $200 $ FCB % Sin& $ FBA = 0


1
# FBA = $200 $ ($200 5 % )=0
5
Page 6

FBA = 0
So what is the purpose of FBA ?

If bar BA were not present a


mechanism would result: this is
a dynamic problem, not
statically determinate

As a check, look at joint A

In equilibrium – we have been consistent! Consistent tensile


forces
0 0
400 N FCA = 400
N
Finally, draw structure
Page 7
This approach worked quite well for a truss with only a few bars, or if we want the load
in every bar, then we work progressively through the truss.
- but may not work for more complicated trusses
- or if we only want to isolate a few bars
- use method of sections

Option 2: Method of Sections

Isolate a section (part) of the truss of interest


- Draw FBD, determine reactions
- "cut" truss into sections (through bars)
- Replace cut bars by coincident tensile forces
- Apply planar equilibrium equations (positive tensile, negative compressive)
All as for method of joints, but:
- Can now use equilibrium of moments
- Can now have 3 unknown bar forces in a section

EXAMPLE 2
Problem: to find force in bar ED
E D

L
45°
A C

2L 2L
2P
First draw FBD
E D

x2 45°
A C HC

x1 ~ 2L
~
2L

2P
VA
VC

We are interested in bar ED. So take an appropriate cut.


Page 8

E F ED D

Closed Must
surface include
45° F AB reactions
A crossing
B surface
F EB

2L

We want to find FED - try to find one equation with only one unknown

→ Take moments about B (draw in B)


!M B : +2!P ! !FED = 0
FED = !2P
Suppose we had wanted FEB. Still use same section:
Page 9

!
! F2 = 0 " + # FEBSin45 + P = 0
!= 1 #
" 2$

FEP = + 2P
Suppose we had wanted FAB

→ Take moments about E


! M =0 E :
! PL + LFAB = 0
FAB = P
Redraw FBD. Is this correct?
Page 10

Is this correct?
- Think about cutting members ED - shortens AB – lengthens

NOTE:
In both methods of joints, method of sections, we solve for the internal forces by isolating
part of the structure.

Two Tips:
• Reduce computation by intelligent choice of method and section to analyze
1 equation, 1 unknown
• Check, check and double check as you go!

1.4 Analysis of Truss Deformations


For statically determinate trusses we have been able to use static equilibrium to obtain the
bar forces. In order to calculate deformation (i.e. displacements of nodes due to loading),
we need to invoke the other two great principles: constitutive behavior and compatibility.

Constitutive Behavior of Bars in a Truss


Will later show that the overall change in length of bar is (for a range of materials, loads)

FL
!= + "#TL
AE

Due Due to a change


to in temperature
force

Symbol Meaning Dimensions Si units

A= Cross-section area
[ L2 ] m2

L= Length of bar [L] m

F= Applied force [ML/T2] N/


Page 11

E= Young’s Modulus [ML/T2.L2] N/M2

α= Coefficient of thermal
expansion
[ L L.! ] M
MK

ΔT = Temperature difference [! ] K

this is the same as Hooke’s law for a spring: F = K! for ΔT=0

FL " AE $
!= and rearranging gives # !=F
AE L %

"k" for a bar

[NOTE: a real spring has added geometry]

Compatibility For Trusses


For truss like structures:
• bars extend or contract axially
• bars rotate about pin joints
• bars remain attached at joints
i.e., deflections and rotations must be compatible
Consider three bar truss from lecture 1, assume no temperature change:

Force (N) Length (m) FL


( µm) A = 1! 10 "4 m2
AE
AB 0 5 0 E = 70GPa

AC +400 10 571 6
AE = 7 ! 10 N
BC -447 125 -714
Page 12

Note; the bar extensions are very small compared to bar lengths
Consider what this implies about deformations of 3 bar truss.

we can enlarge the region of interest:


assume deflections are small, circular arcs become straight lines at right angles to
extension/contraction of bars: displacement diagram
Page 13

Can extend to other joints by considering relative displacements.


TIP: Do this on graph paper - measure deflections rather than calculate.

1.4 Concluding Remarks


We have taken a quick look at how we can use equilibrium, constitutive behavior to
examine the internal forces and deformations of a particular class of statically
determinate structure; trusses. Most structures are not made up of discrete bars, instead
they consist of continuous elements, e.g. aircraft skins, ship hulls, pressure vessel walls.
We need to have analysis tools that allow us to deal with such cases. At the heart of these
tools is an understanding of how forces and displacements can be understood in
continuous structures. This leads to the ideas of stress and strain.
Page 1
Lectures 4 and 5 Stress
Block 2: Continuum Mechanics: Stress, Strain and Elasticity
We want to look more closely at how structures transmit loads. In general, structures are not
made up of discrete, 1-D bars, but continuous, complex, 3-D shapes. If we want to understand
how they transmit load we need new tools. To this end we will introduce the concept of stress
(and later strain). See BCA, Chapters 2 and 11.

Definition Stress is the measure of the intensity (per area) of force acting at a point.
Use coordinate system x, y, z ! x1, x2 , x 3

Example:

For each square, increment of force and area given by:


!F1 = F1 nm

!A = A nm
The stress is the intensity, force per unit area, so let m and n go to infinity.

lim $F1 ! force $


!1 = n " # units #
2
&: N/m : Pascals (Pa)
m"# $A 2
" length %

where ! 1 is the stress at a point on the x1 - face


- Magnitude ! 1
- Direction i1
Consider a more general case:

Force, F, is a
vector quantity

Force F acting on x1 face ( i1 is normal to plane of face).


Page 2
Resolve force into three components

F = F1 i1 + F2 i 2 + F3 i 3

Then take the limit as the force on the face is carried by a smaller set of areas.

lim !F
!A " 0 =0 (but stress vector has other components)
!A
! = ! 11 i1 + ! 12 i 2 + ! 13 i 3

!F1 !F2 !F3


!A1 !A1 !A1

Where ! 1 is the stress vector and !1n are the components acting on the x1 normal face, in the x1,
x2 and x3 directions.

We obtain similar results if we looked at the x2 and x3 faces.

! 2 = ! 21 i 2 + ! 22 i 2 + ! 23 i 3 on i2 face
! 3 = ! 31 i1 + ! 32 i 2 + ! 33 i 3 on i3 face

We can write this more succinctly as:

! m = ! mn i n

! mn is the stress tensor, ! m is the stress vector.


This method of representing a set of vector equations is an example of indicial, or tensor,
notation:

Tensor (Indicial) Notation

This is a simple way to write formulae with several terms. Example:


xi = Fij Y j

Subscripts or indices

By Convention
Latin subscripts m, n, i, j, p, q, etc.
Take values 1, 2, 3, (3-D)

Greek Subscripts ! , " , #


Take values 1 or 2 (2-D)
Page 3
Three Rules for Tensor Manipulation
1. A subscript occurring twice is a repeated (or dummy) index and is summed over 1, 2, (and)
3.
Implicit summation
3
For xi = Fij Y j = ! Fij Y j
j=1
i.e: xi = Fi1Y1 + Fi2Y2 + Fi3Y 3

2. A subscript occurring once in a term is called a free index, can take on the range 1, 2, (3) but
is not summed. It represents separate equations.
x1 = F11Y1 + F12Y2 + F13Y 3
x2 = F21Y1 + F22Y2 + F23Y 3
x 3 = F31Y1 + F32Y2 + F33Y 3
3. No index can appear in a term more than twice.

Stress Tensor and Stress Types


! mn is the stress tensor
Consider a differential cubic element:

! mn

Stress acts on
Acts in xn direction
face with
normal in the
xm-direction
Page 4
Types of Stress
We can identify two distinct types of stress:

Normal (Extensional) - acts normal to face


!11, ! 22 , ! 33 - acts to extend element

Shear Stress - acts parallel to face plane


!12 , ! 13, ! 21, ! 23, ! 31, ! 32 - acts to shear element
i.e. 9 components of stress, but it turns out that not all of these are independent:

Symmetry of Stress Tensor


Consider moment equilibrium of differential element:

Take moments about C +

#1 & #1 &
2! 23"x 3 "x 1 % "x 2 ( ) 2! 32"x 2 "x 1 % "x 3 ( = 0
$2 ' $2 '

simplifies to: ! 23 = ! 32

Thus, in general ! mn = ! nm
Stress tensor is symmetric. Six independent components of the stress tensor.
! 11 ! 12 " = ! 21 %
! 22 ! 23 $$ = ! 32 ''
! 33 ! 31 # = ! 13 &
What happens if stress varies with position?

Stress Equations of Equilibrium

Consider equilibrium in x1 direction; let stress vary.


Page 5

x3 !11

#x2
"! 31
! 31 + #x3
"x 3

"! 21
!21 ! 21 + #x2
"x2

#x3 x2

#x1
!31
x1
"! 11 f1
! 11 + #x
" x1 1

fn is a body force - acts over volume - gravity, electromagnetic forces,


acceleration: it is a force per unit volume

Force equilibrium ! F1 = 0 +
#
% ! 11 + "! 11 dx1 &( ( dx2 dx3 ) ) ! 11( dx2 dx3 )
$ "x1 '

# "! &
+ % ! 21 + 21 dx2 ( (dx1dx2 ) ) ! 21 (dx1dx 3 )
$ dx 2 '

# "! &
+ % ! 31 + 31 dx 3( (dx1dx2 ) ) (! 31 )( dx1dx 2 )
$ "x3 '

+ f1dx1dx2 dx3 = 0

dx1 dx2 dx3 common to all terms

Canceling out terms and dividing through by the common: dx1, dx 2 , dx3 .
Gives:
!"11 !" 21 !" 31
+ + + f1 = 0 x1 direction
!x1 !x 2 !x3
Similarly,
!" 12 !" 22 !" 32
+ + + f2 = 0 x2 direction
! x1 ! x2 ! x3
Page 6
!"13 !" 23 !" 33
+ + + f3 = 0 x3 direction
!x1 !x2 !x3

These are the three equations of stress equilibrium.


In tensor form:

!" mn
+ fn + 0
!xm

Stress Transformations
Just as we need to resolve forces with respect to structural axes - need to resolve stress. Identify
directions which "see" maximum stress

Two Dimensional Case:


Consider a unit square (of unit depth dx3)

What is the stress acting on a plane at an arbitrary angle ! ?

Consider equilibrium of a triangular, wedge shaped element:

• Align x˜1 perpendicular to cut face, x˜2 parallel to cut face


Page 7

• The element is of uniform depth dx3

• Must still be in equilibrium of forces:


!F =0
Σ(Stress component x length x depth)

note:dx 2 = dx˜2 cos!


dx1 = dx˜2 sin!
Equation of equilibrium in x˜1 direction:

area Resolved component

!˜11dx3dx˜ 2 " !11dx3 [#x˜ 2 cos$ ]cos $ " ! 22 dx3[#x˜2 sin $ ] sin$ " !12 dx3[#x˜2 sin $ ] cos$ " ! 21dx3[#x˜2 cos$ ] sin$ = 0

cancelling out !˜x2 and combining


!˜11 = cos2 " ! 11 + sin2 " ! 22 + 2 cos" sin " ! 12

Note: two angles: one for area; one for resulting force

This can be done for !˜12 and !˜ 22 , e.g.


(
!˜12 = " !11 cos# sin # + ! 22 sin# cos# + !12 cos2 # " sin2 # )
note one can always take equilibrium of an element

Tensor form
Transformation requires direction cosines (see below)
Stresses are second order tensors (2 subscripts) and require two direction cosines for
transformation.
Thus:
!˜ mn = ! m˜ p! n˜ q! pq
• 2nd order tensor requires 2 direction cosines
• transforms stress from xn system to x˜m system
one angle for area, one for component of force

In 2-D !˜ "# = !"˜$ ! ˜ !$%


#%

Direction Cosines
Page 8
Define direction cosine as:

Angle between

! m˜n = cos[ x m x˜ n ]
such that
xm = ! m˜n x n˜
Tensor equations
or x˜n = !n˜ m x m

where we mean
x1 = !1 ˜1x˜1˜ + !12˜ x˜ ˜2 + !13˜ x˜3˜
x2 = ! 2˜n x˜ n˜
x3 = ! 3˜n x˜n˜

We could use the same idea for forces:


F˜n = ! n˜ m Fm
First order tensor is a vector (one subscript) requires one direction cosine for transformation.
Angles measured positive counter-clockwise.
! m˜ n = !nm˜ Since cosine is an even function cos(! ) = cos(" ! )

But ! m˜ n ! !n˜ m - different angles

2-D Example

y1 y2 y3

y˜1 !˜11 = cos! !1˜ 2 = cos(90 " ! ) = sin! !1˜ 3 = cos90 = 0

y˜2 ! ˜21 = cos("(90 + ! )) ! ˜22 = cos! ! ˜23 = cos90 = 0

y˜3 !3˜ 1 = cos90 = 0 ! ˜32 = cos 90 = 0 ! ˜33 = cos! = 1


Page 9

One further way to transform stress is to use "Mohr's Circle" graphical method for rotating 2nd
order tensors in 2-D coordinate system

Mohr's Circle for Stress


Mohr's circle is a geometric representation of the 2-D transformation of stresses.
Construction: Given the state of stress shown below, with the following definition (by Mohr) of
positive and negative shear:
"Positive shear would cause a clockwise rotation of the infinitessimal element about the element
center."
Thus: ! 2 1 (below) is plotted positive
! 1 2 (below) is plotted negative:

Following these rules for the sign of stress do the following:


1. Plot ! 1 1, - ! 1 2 as point A
2. Plot ! 2 2 , ! 2 1 as point B
3. Connect A and B.
4. Draw a circle with a diameter given by the line AB about the point where the line AB
crosses the horizontal axis (denote this as C).
To read off the stresses for a rotated system:
1. Note that the vertical axis is the shear axis; the horizontal axis is the extensional stress axis.
2. Positive rotations are measured anti-clockwise as referenced to the original system and thus to
the line AB.
3. Rotate line AB about point C by the angle 2θ where θ is the angle between the un-rotated and
rotated system.
Page 10
4. The points D and E where the rotated line intersects the circle are used to read off the stresses
in the rotated system. The vertical location of D is ! "˜ 1 2 ; the horizontal location of D is !˜ 1 1. The
vertical location of E is !˜ 2 1, the horizontal location of E is !˜ 2 2 (recall Mohr's definition).

We can also immediately see the following:


1. The principal stresses, ! I and ! II are defined by the points G and F (along the horizontal axis
!˜ 1 2 = 0 ). The rotation angle to the principal axis is θp which is half the angle from the line AB to
the horizontal line FG.
2. The maximum shear stress is defined by the points H and H' which are the endpoints of the
vertical line. The line is orthogonal to the principal stress line and thus the maximum shear stress
acts along a plane 45° (= 90°/2) from the principal stress system.

Mohr's Circle (Principal Stresses)


Shear

G C F Ext

2!p

A
H'
SESG1001 Lectures 7 & 8: Strain

We have examined stress, the continuum generalization of forces, now let's look at the
continuum generalization of deformations:

Definition of Strain
Strain is the deformation of the continuum at a point. Or, the relative deformation of an
infinitessimal element.

Two ways that bodies deform: By elongation and shear:

Elongation (extension, tensile strain)


Tensile strain can be thought of as the change in length relative to the original length

Ldeformed ! Lundeformed
tensile strain =
Lundeformed

but body can also deform in shear – angles change

Shear
This produces an angle change in the body (with no rotation for pure shear)
Consider infinitesimal element.
Undeformed:

Deformed :

Consider change in angle BPA = φ


φ = undeformed - deformed

NOTE: By convention positive strain is a reduction in angle (consistent with positive


shear stresses)
*! $ ! '-
Δ = , " & " # )/ = # radians
+2 %2 (.

NOTE: Strain is a non-dimensional quantity


(Just as for stress, strain varies with position and direction. Use tensor formulation to
describe)

Formally Define ! mn = strain tensor


There is a general definition of strain, but usually we are concerned with small strains:

change in length < 10% (0.1), change in angles < 5% (0.05)

Good for range of use of most engineering materials, most structural applications.
Allows us to neglect higher order terms and leads to:
Strain - Displacement Relations (for small strains)
Conceptually, want to separate out rigid body translations from
deformations

!11 = relative elongation in x1 direction


! 22 = relative elongation in x2 direction.
! 33 = relative elongation to x3 direction.

Consider infinitesimal element, side length dx1 undergoing displacements and deformations
in the x1 direction defined by
u1 is a field variable = u1( x1,x 2 , x3 ) , i.e. displacements vary with position
let u1 be the displacement of the left-hand side of the infinitesimal element.
" !u1 %
And $ u1 + dx1 'is the displacement of right-hand side
# !x1 &
Recall:
! deformed # ! undeformed
!11 " !1 =
! undeformed
! undeformed
% $u (
' 1 1
dx + u + dx1 # u1* # dx1
& $x1 )
!11 =
dx1

Rigid body
$u translation
= 1
$x1

Similarly, ε22 = elongation in x2

! deformed " ! undeformed


!22 =
! undeformed
$ #u2 '
& 2
dx + u2 + dx 2 " u2 ) " dx2
% #x2 (
=
dx2
"u2
!22 =
"x2

"u 3
and ! 33 =
"u 3

Now for shear strain


Need to be careful which angles we choose to define shear strain
" 1
angle change# = 0/ 12 $% etc.
1
!12 =
2 2

=
1
2
( undeformed - deformed )
1$! $! ''
= & " & " #12 ))
2% 2 % 2 ((
e.g:

Angular charge of the x1 edge in


the x2 direction

!12 + ! 21 = total angle change in x1 " x 2 plane


!13 + ! 31 = total angle change in x1 " x3 plane
! 23 + ! 32 = total angle change in x2 " x3 plane

NOTE: The strain tensor is defined such that there seem to be two parts in each angle
change. But the stress tensor is symmetric and we would like the strain tensor also to be
symmetric. Thus:

1
!12 = ! 21 = ( angle change in x1 " x 2 plane)
2
1
!13 = ! 31 = (angle change in x1 " x3 plane)
2
1
! 23 = ! 32 = ( angle change in x 2 " x3 plane)
2

Consider diagonal of parallelogram undergoing pure shear:


Angles change by same amount.

Formal Definitions

!12 = "1 + "2


Make small angle approximation: tan ! = !

"u1
u1 + dx2 # u1
"x2
!1 =
Hence: dx2
"u
u2 + 2 dx1 # u2
"x1
!2 =
dx1
!12

% 1 +% $u2 ( % $u1 (.5(


' 3 -' 2
u + dx1 # u2* ' 1
u + dx 2 # u1 *03*
1 ' " 3 " -& $x1 ) & $x2 )03*
!12 = #2 # + 6
2' 2 3 2 - dx1 dx2 03*
'' - 03**
34 , /7)
&
1 # "u "u &
!12 = % 2 + 1 ( = !21
2 $ "x1 "x2 '
and similarly:
1 # "u "u &
!23 = % 2 + 3 ( = ! 32
2 $ "x 3 "x 2 '
1 # "u "u &
! 31 = % 3 + 1 ( = !13
2 $ "x1 "x 3 '

1 # "ui "u j &


or !ij = %% + ( The strain tensor for small strains
2 $"x j "xi ('

Strain displacement relation


6 independent strain components

(We have dealt with elongation and shear strains but remember, we were concerned to
eliminate translation and rotational displacements)

(- translation accounted for by subtracting u1, u2 etc. from strain equations)


Rotation
!1 = ! 2

"u 2
u2 + dx # u
"x1 1 2 "u2
!1 = = as before
dx1 "x1

Negative because it acts to increase enclosed


angle - consistent with shear strain

$ #u '
" %u1 + 1 dx2 " u1(
& #x2 ) "#u1
or !2 = =
dx2 #x 2
But may also have shear deformation of the cube !1 " ! 2
Define average rotation (of diagonal)
1 $"u2 "u1 '
! = % # ( note (-) sign
2 & "x1 "x 2 )

Transformation of Strain
Identical to transformation of stress
!˜mn = ! m˜ p ! n˜ q! pq
Mohr's Circle in 2-D
Principal strains/directions
-no associated shear strains
Alternative notations for stress and strain

Engineering Notation (see Benham, Crawford and Armstrong)


x1 ! x
x2 ! y
x3 ! z
Tensor Engineering
!11 !x
! 22 !y
! 33 !z
!12 ! xy
!13 ! xz
! 23 ! yz

In addition τ (tau) is often used for shear stress


! xy = " xy
! xz = " xz
! yz = " yz
For stress this is just the substitution of different symbols, however for strain there is a
fundamental difference from tensor notation.
Engineering shear strain = total angle change
Tensor shear strain = 1/2 (angle change)
Tensor Engineering

!11 " !x

! 22 " !y

! 33 " !z Total angle


change
1 1
!12 " ! or #
2 xy 2 xy
1 1
!13 " ! xz or # xz
2 2
1 1
! 23 " ! or #
2 yz 2 yz

Matrix Notation
"! 1 1 ! 1 2 ! 1 3 %
" !11 !12 !13 % $ '
! mn = $ ! 21 ! 22 ! 23 ' ! mn = $! 2 1 ! 2 2 ! 2 3'
$ '
$# ! ' $#! 3 1 ! 3 2 ! 3 3'&
31 ! 32 ! 33 &
- principal values are eigenvalues
- principal directions are eigenvectors

Next we will link stress and strain by constitutive behavior and to do this we will need to
look at material properties.
SESG1001 Lectures 9 & 10: Elastic Stress-Strain Relations

These lectures will complete our quick journey through continuum mechanics, to provide
you with a continuum version of a constitutive law - at least for linear elastic materials
! pq = E ? " mn

Elasticity
Where does it
come from?

Increasingly, materials are designed along with the structure, you need insight into what
contributes to material properties. These lectures will link to topics you will learn about
in materials lectures and form a foundation for material selection in design.

Engineering Elastic Properties of Materials


In order to understand how we link stress and strain we need to understand that there are
two points of view to this matter. There is the experimental point of view that some
properties (behaviors) are easier to measure than others, and there is the mathematical
point of view that some representations of physical phenomena are mathematically easier
to handle than others. In the present case, engineering elastic constants are derived from
an experimental point of view, whereas the stress and strain tensors, are mathematically
useful. We need to resolve these two points of view.

Young’s modulus and Poisson’s ratio


From materials lectures and the materials and solid mechanics labs you will become
familiar with at least two elastic constants.

If we apply a uniaxial tensile stress ! L to a constant cross-section rod of material, we


will obtain a biaxial state of strain, consisting of an axial tensile strain ! L and a

transverse strain !T . The axial strain will be tensile for a tensile applied stress, and the
transverse strain will usually be compressive. We can measure the strains using
resistance strain gauges.

For many materials, over some range of applied stress, the applied stress and the resulting
strains will follow a linear relationship. This observation is the basis for the definition of
the engineering elastic constants. The Young’s modulus, E, is defined as the constant of
proportionality between a uniaxial applied stress and the resulting axial strain, i.e:
! L = E" L

Note. This only applies for a uniaxial applied stress, and the component of strain in the
direction of the applied stress.

We can also define the Poisson’s ratio, ! , as the ratio of the transverse strain to the axial
strain. Since for the vast majority of materials the transverse strain is compressive for a
tensile applied stress, the Poisson’s ratio is defined as the negative of this ratio, to give a
positive quantity. I.e:
#
! =" T
#L

Typically, for metals, ν = ~ 0.3-0.4.


A similar process, of performing experiments in which a well-defined stress state is
applied and the resulting strain state is measured leads us to define other elastic constants,
we will concern ourselves with only one:

The Shear Modulus


Application of a state of pure shear, leads to a shear strain:

Note angles are exaggerated in the figure.

An applied shear stress leads to an applied shear strain. The shear strain, ! , is defined in
engineering notation, and therefore equals the total change in angle: ! = ".
Consistent with the definition of the Young’s modulus, the Shear modulus. G, is defined
as:
! = G"

Again, note, that this relationship only holds if a pure shear is applied to a specimen.
For reasons that will become apparent shortly, The Young’s modulus, Shear Modulus and
Poisson’s ratio are linked. For most materials Poission’s ratio’s are approximately 0.33,
3
and for these materials G ! E . However, values of the Young’s modulus can vary
8
widely.

The Young’s Moduli of Engineering Materials


A useful way of representing the range of Young’s moduli is to plot them for all the
classes of material against some other material property. Since we are often interested in
lightweight structures, we will choose the density of the materials for this purpose.

From: Material Selection in Mechanical Design, M.F Ashby, Pergamon Press, Oxford,
1992
Note, how the different classes of material tend to cluster: Metals, Polymer, Glasses and
ceramics have similar ranges of density and modulus. There are also some materials that
have quite wide ranges of moduli (and densities), while others are relatively narrowly
banded. Finally note how wide an overall range of moduli is represented, from 0.01 GPA
for foams to 1000 GPA for diamond. The range of densities is somewhat less, but still
spans more than two orders of magnitude. See SESG1003 for more details.

Generalized Hooke's Law


We have met the engineering elastic constants, Young's moduli, Shear Moduli and
Poisson's ratio's, and understand that many structural materials behave elastically over
some range of stress and strain.

Now we want to add a mathematical formalism to this physical basis, i.e. our 3rd great
principle, that of constitutive behavior.

A couple of problems we would like to be able to solve: a) what are the strains due to
stresses at an arbitrary angle to the loading direction:
b) We would also like to be able to deal with any state of multiaxial stress and convert to
the resulting strains, or vice versa. To do this we need to revisit tensor stress and strain.

i.e. we want the elastic property that links the stress tensor to the strain tensor:

σmn=Emnpqεpq
Where Emnpq is the 4th order (i.e. 4 subscripts) ELASTICITY (or STIFFNESS) tensor.
e.g:
!11 = E1111"11 + E1112"12 + E1113"13 ( p = 1,sum on q)

. +E1121" 21 + E1122 "22 + E1123" 23 ( p = 2,sum on q )


+E1131" 31 + E1132" 32 + E1133" 33 ( p = 3,sum on q )
4th order tensor has 81 components, m,n,p,q =1, 2 and 3 therefore 34 = 81 terms
But fortunately there are symmetries which reduce this to only (!) 21 independent
components of the elasticity tensor. In matrix form, this can be written as:

"! 11 & ) E1111 E1122 E1133 2E1123 2E1113 2E1112 , " /11 &
$! $ + E
22 1122 E2222 E2233 2E2223 2E2213 2E2212 . $/ 22 $
$ $ + .$ $
$! 33 $ + E1133 E2233 E3333 2E3323 2E3313 2E3312 . $/ 33 $
# '= + # '
$ ! 23 $ + E1123 E2223 E3323 2E2323 2E1323 2E1223 . $/ 23 $
.
$! 13 $ + E1113 E2213 E3313 2E1323 2E1313 2E1213 . $ /13 $
$ $ + .$ $
%! 12 ( * E1112 E2212 E3312 2E1223 2E1213 2E1212 - %/12 (
Since it links strains to stresses, Emnpq is also termed the "stiffness tensor". By inverting
the matrix we could also obtain the tensor linking stresses to strains, i.e. εmn=Smnpqσpq
Where Smnpq is the "Compliance" tensor (Why S is for compliance and E is used for
stiffness is unclear to me!)

Even with the simplifications, 21 independent terms seems rather too many to have to
deal with!. Let's go back to the engineering elastic constants and see if we can see how to
simplify this list further.
We know that there are several different classes of material. Most metals and ceramics
are isotropic, that is they have the same properties in any direction that you measure. By
contrast, fiber-reinforced composites may have different properties in different directions,
i.e. they are anisotropic, generalized elasticity provides the framework to analyze these.

Elasticity of Isotropic Materials


In SESG1001 we will only look at isotropic materials. However, we do want to be able
to consider the case in which an element of a structure is loaded by all possible
components of stress and we want to know the resulting strains. Also let's ignore thermal
expansion strains for the time being. We have six components of stress producing six
components of strain, therefore we need a six by six matrix to link them:

# !x & #_ _ _ _ _ _& # ) x &


% %
! ( %_ _ _ _ _ _( ) y (
% y( % (% (
% !z ( % _ _ _ _ _ _( % ) z (
% " zy ( = % %
_ _ _ _ _ _ ( * zy (
% " zx ( % (% (
%" ( % _ _ _ _ _ _ ( % * zx (
% xy ( % (% (
$ ' $ _ _ _ _ _ ' $ * xy '

We also know that for small strains, and elastic materials the contributions of the separate
components of stress will superimpose. So let's consider the case of only σx applied and
all the other components of strain are zero:

# !x & #_ _ _ _ _ _ & #) x &


%
!y ( %_ _ _ _ _
%
_( 0 (
% ( % (% (
!
% z ( %_ _ _ _ _ _( % 0 (
% " zy ( = % %
_ _ _ _ _ _( 0 (
% " zx ( % (% (
%" ( % _ _ _ _ _ _( % 0 (
% xy ( % (% (
$ ' $ _ _ _ _ _' $ 0 '
Taking each component in turn:
"x
What does εx equal?. This just reduces to a 1-D tensile test, so ! x =
E

$
And εy , εz are given by the Poisson contractions, so: ! y = ! z = "#! x = "# x
E

This allows us to fill in the first line of the matrix, and also by noticing that we could
have equally well applied σy or σz, and obtained similar relationships, we can fill in all of
the top left hand quadrant of the matrix:

# !x & # 1 )* )*
_ _ _& # + x &
% % E E E (% (
! y ( % )* 1 )*
_ _ _( % + y (
% ( E E E
% z ( % )*
! )* (%+
_( z (
1
=% E _ _
% % (
" zy ( %
E E
,
% ( _ _ _ _ _ _ ( % zy (
% " zx ( % ( % , zx (
_ _ _ _ _ _
% ( %% (( % (
$ " xy ' $ _ _ _ _ _
,
_ ' $ xy '

If instead of applying an extensional stress we applied a shear stress, we know that the
shear stress and shear strain are linked by the shear modulus, so:
# 1 )* )*
_ _ _ &#+
# !x & % E E E ( x&
%
!y ( % )* 1 )*
_ _ _ (%+y (
% ( % )E*
E E
(% (
)* 1
% !z ( % _ _ _ ( % +z (
% =% E E E
% (
" zy ( _ _ _
1
0 0 ( % , zy
% ( % G ( (
% " zx ( % _ _ _ 0
1
0 ( % , zx (
% ( G % (
$ " xy ' % 1 (( $ , xy '
% _
_ _ 0 0
$ G'
Finally we note that for isotropic materials the application of an extensional stress does
not result in a shear strain or vice versa, so the top right and bottom left quadrants are
populated by zeros:

# 1 )* )*
0 0 0& +
# !x & % E E E (# x &
% )* )*
0 (%+y (
1
!y ( % E 0 0
% ( % )* E E
(% (
)* 1
% !z ( % 0 0 0 ( % +z (
% ( =% E E E
% (
0 ( % , zy
" 1
% zy ( % 0 0 0
G
0
( (
% " zx ( % 0 0 0 0
1 % ,
0 ( zx (
% ( G % (
$ " xy ' %% 1 (( $ , xy '
$ 0 0 0 0 0
G'

We have three separate elastic constants required, i.e: E, ν and G. However if we go


back to our knowledge of stress and strain transformation we can reduce this further.
Remember that the application of a shear stress can be thought of as a shear stress
resulting in a shear strain:
"
! =
G
or a biaxial stress state, of a combined tension and compression, at 45 degrees to the axis
of pure shear, i.e:

In terms of the Mohr's circles of stress and strain these appear as (note the factor of two
for the representation of shear strain on the Mohr's circle):
For the biaxial tensile and compressive stress the resulting (principal) strain is given by:

"I " ! II = "! I


!I = # $ II
E E

1
but for the case of pure shear: ! I = " (From Mohr's Circle -remember the
2
factor of two between tensor and engineering shear strain)
and:
! = " II = #" I

'1 &
$% = 2( + )* ! + G =
E
E E 2(1 + & )
So we actually only have two independent elastic constants for an isotropic material.
Note that this only applies for isotropic materials

If we want to go in the reverse direction (i.e. have known strains and want to calculate
stresses) we need to invert the matrix of elastic constants. Note, this situation may arise
because we can experimentally measure strains using strain gauges. The inverse matrix
is usually expressed in terms of groupings of the elastic constant, known as Lamé's
constants, µ & ! , where:
!
µ= =G
2(1+ " )

"E
#=
(1 + " )(1 $ 2" )
# ! x & # ) + 2µ ) ) 0 0 0& # * x &
% ( % % (
! ) ) + 2µ ) 0 0 0( * y
% y( % (% (
% !z( % ) ) ) + 2µ 0 0 0 ( % *z (
Thus: % ( = % % (
" 0 0 0 µ 0 0 ( + xy
% xy ( % (% (
% " xz ( % 0 0 0 0 µ 0 ( % + xz (
% ( % (% (
$ " yz ' $ 0 0 0 0 0 µ' $ + yz '

We can also include the effect of thermal expansion:


# 1 + +
# !x & * * 0 0 0 &#,
% ) ) ) ( x & #. &
%
!y ( %* + 1
*
+
0 0 0 ( % , ( %. (
% ) ) ) y
% ( (% ( % (
+ + 1
% !z ( %* ) * 0 0 0 ( % , z ( %. (
=% ) )
% % ( + % ( /T
" zy ( 0 0 0
1
0 0 ( % - zy 0
% ( % G ( ( % (
% " zx ( % 0 0 0 0
1
0 ( % - zx ( % 0 (
% ( G % ( % (
$ " xy ' % 1 (( $ - xy ' $ 0 '
%
$ 0 0 0 0 0
G'
We are now equipped to analyze stresses and strains in simple continuous structures
SESG1001 Lecture 11 1-D Structures
In the rest of the course we are going to utilize the principles we have learnt so far (i.e the
3 great principles and their embodiment in the equations of elasticity) in order to be able
to analyze simple structural members. These structural members are: Rods, Beams,
Shafts and Columns.

The key feature of all these structural members is that one dimension is longer than the
others (i.e. they are one dimensional).

There is a basic logical set of steps that we will follow for each in turn:

1) We will make general modeling assumptions for the particular class of structural
member
In general these will be on:
a) Geometry
b) Loading/Stress State
c) Deformation/Strain State

2) We will make problem-specific modeling assumptions on the boundary conditions that


apply (such as pins, clamps, rollers that we encountered with truss)

a.) On stresses At certain locations in


b.) On displacements the structure

3) We will apply an appropriate solution method:


1.) Exact/analytical (SESG1001)
2.) Numerical (2 nd , 3rd and 4th year classes) - approximate. Such as energy
methods (finite elements - use computers)

Let us see how this works:


Rods
The first 1-D structure that we will analyzes is that of a rod (or bar), such as we
encountered when we analyzed trusses. We are interested in analyzing for the stresses
and deflections in a rod.

First start with a working definition - from which we will derive our modeling
assumptions:

"A rod (or bar) is a structural member which is long and slender and is capable of
carrying load along its axis via elongation"

Modeling assumptions
a.) Geometry

L = length (x1 dimension)


b = width (x2 dimension)
h = thickness (x3 dimension)

assumption: L much greater than b, h (slender member)

(think about the implications of this - what does it imply about the magnitudes of stresses
and strains)

b.) Loading - loaded in x1 direction only

Results in a number of assumptions on the boundary conditions


! 21 = 0
! 22 = 0
! 23 = 0

Similarly on the x3 face - no force is applied

! 31 = 0 = !13

! 32 = 0 = ! 23
! 33 = 0

on x1 face - take section perpendicular to x1


!12 = 0
!13 = 0

and
" ! 11dA = P
A

"" ! 11dx2 dx3 = P

P P
# ! 11 = =
bh A

c.) deformation

Rod cross-section deforms uniformly (is this assumption justified? - yes, there are no
shear stresses)
So much for modeling assumptions, Now let's apply governing equations and solve.

1. Equilibrium

!" mn
+ fn = 0
!x n

only !11 is non-zero

!" 11
+ f1 = 0 f1 = body force =0
!x1

!" 11
= 0 # " 11 = constant = P
!x1 A

Constitutive Laws
stress - strain equations:

"1%
! 1 1 = $ ' (1 1
# E&
" )* %
! 2 2 = $ ' (1 1
# E&
" )* %
! 3 3 = $ '( 1 1
# E&

Compliance Form

Now apply strain – displacement relations:


1 # "u "u &
! mn = % m + n (
2 $ "x n " x m '
"u "u "u
!11 = 1 , !22 = 2 , ! 33 = 3
"x1 "x2 "x 3

Hence:
! 11 "u1
=
E "x1

"u1
# P AE =
"x1

integration gives:

Px1
u1 = + g(x2 , x3 )
AE

Apply B. C. u1 = 0 @ x1 = 0 ! g(x2 , x 3 ) = 0
Px1
! u1 =
AE
!"P
similarly u2 = x2
AE
!"P
u3 = x3
AE

1 # "u1 "u2 &


check: !12 = % + (=0 √
2 $"x2 "x1 '

Further assumptions

(Closer inspection reveals that our solutions are not exact.)

1) Cross section changes shape slightly. A is not a constant.


If we solved the equations of elasticity simultaneously, we would account for this.
Solving them sequentially is acceptable so long as deformations are small. (δA is
second order.)

2) At attachment point boundary conditions are different from those elsewhere on


the rod.

u1 = 01 , u2 = 01, u3 = 0

We deal with this by invoking St. Venant's principle:

"Remote from the boundary conditions internal stresses and deformations will be
insensitive to the exact form of the boundary condition."

And the boundary condition can be replaced by a statically equivalent condition


(equipollent) without loss of accuracy.

How far is remote?

This is the importance of the "long slender" wording of the rod definition.

This should have been all fairly obvious. Next time we will start an equivalent process
for beams - which require a little more thought.
SESG1001 Lectures 12 and 13
Beam Theory: Bending Moment and Shear Force
Distributions
Reading: Benham, Crawford and Armstrong, 1.10, Ch 6 and Ch 7

A beam is a structural member which is long and slender and is capable of carrying
bending loads. i.e loads applied transverse to its long axis.

Beams are very common structural members and are to be found across all areas of
engineering. Wings, masts, gas turbine blades even whole ships, aircraft or spacecraft
can be treated as beams to varying degrees.

1. We will continue, as for 1-D rods to make a set of modeling assumptions:

a.) Geometry, slender member, L>>b, h

At this stage, we will assume arbitrary, symmetric cross-sections, i.e.:


b) Loading:
Similar to rods (traction free surfaces) but applied loads can be in the z direction

c) Deformation
We will talk about this later

2) Boundary conditions
As for rods, trusses

Pinned, simply supported, point load Cantilever, built in support, distributed load

Analysis approach: Draw free body diagram, apply equilibrium to determine reactions.

3.) Governing equations


Equations of elasticity (stress, strain, constitutive behaviour)

4.) Solution Method


Exact (exactly solve governing differential equations)
Approximate (use numerical solution)

Before moving through this process we need to understand how beams transmit load.

Internal Forces and Moments

Apply methods of sections to beam (also change coordinate system – to x, y, z –


consistent with BCA). Method exactly as for trusses. Cut structure at location where we
wish to find internal forces, apply equilibrium, obtain forces. In the case of a beam, the
structure is continuous, rather than consisting of discrete bars, so we will find that the
internal forces (and moments) are, in general, a continuous function of position.
Internal forces

Opposite
directions on two
faces - equilibrium
where M = Bending movement Beam
S = Shear force Bending "beam bar"
F = Axial force bar, rod

Also drawn as:

Examples of calculating shear force and bending moment distribution along a beam.

Example 1. A cantilever beam:


Free Body Diagram (note moment reaction at root)

Equilibrium: H A = 0, VA = P, M A = !PL

Take cut at point X ,distance x from left hand end (root). 0< x < L. Replace
the effect of the (discarded) right hand side of beam by an equivalent set of forces and
moments (M, S, F) which vary as a function of position, x.

Apply equilibrium
+
! Fx = 0 " F(x) = 0
! Fz = 0 # + P $ S(x) = 0 S( x) = P
! MX = 0 + PL + M( x) $ Px = 0 M( x) = $ P(L $ x)
Draw "sketches" - bending moment, shear force, loading diagrams

Shear Force Diagram

Bending moment diagram:

NOTE: At boundaries values go to reactions (moment at root, applied load at tip).

These representations are useful because they provide us with a visual indication of
where the internal forces on the beam are highest, which will play a role in determining
where failure might occur and how we should design the internal structure of the beam
(put more material where the forces are higher).
Example 2: A simply supported beam

Free body diagram:

take cut at 0< x< L/2


Equilibrium →F(x) = 0

P P
! " S(x) = 0 # S(x) =
2 2 0<x<L
P Px
M x = 0 + # " x + M (x) = 0 # M (x) =
2 2
L
take cut at <x<L
2

Apply equilibrium (moments about X):

P P
" ! F+ = 0 : # P # S(x) = 0 $ S(x) = #
2 2

P % L(
"M =0: # x + P' x # * + M (x) = 0
2 & 2)

P
M (x) = ! (x ! L)
2
P
= (L ! x)
2
Draw Loading, Shear Force and Bending Moment Diagrams

Observations

• Shear is constant between point loads


• Bending moment varies linearly between discrete loads.
• Discontinuities occur in S and in slope of M at point of application of concentrated
loads.
• Change in shear equals amount of concentrated loads.
• Values of S & M (and F) go to values of reactions at boundaries

Distributed loads
e.g. gravity, pressure, inertial loading. Can be uniform or varying with position.

"
q(x) = q0 # 1! $% &= q0 @ x = 0, = 0@ x = L
x
q(x) = q 0
L
[ q ! ] = [force/length]

We deal with distributed loads in essentially the same way as for point loads.
Example: Uniform distributed load, q (per unit length), applied to a simply supported
beam.

Free body diagram:


Apply method of sections to obtain bending moments and shear forces:

Apply equilibrium:

x
qL
"! + # Sx # $ q dx = 0
2 0

qL
% # qx = S(x)
2

qLx x
! Mx + " + Mx + # qxdx = 0
2 0
qL qx2 qLx qx2
" + Mx = = 0 $ Mx = "
2 2 2 2
Plot:

Observations
• Shear load varies linearly due to uniform distributed load.
• Moment varies quadratically (parabolically) over region where distributed load
applied
• This suggests a relationship between M & S & P

General Relation Between q, S, M

Consider a beam under some arbitrary variable, distributed loading q(x):

Consider an infinitesimal element, length, dx, allow F, S, M to vary across element:

!M
M+ dx
!x

!F
F+ dx
!S !x
S + dx
!x
Now use equilibrium, replace partial derivatives by regular derivatives (F, S, M varying
only in x).
+ dF dF
! Fx = 0" : # F + F + dx
dx = 0
dx
=0
dS
! Fz = 0 " + S # S # dx + q(x)dx = 0
dx
dS
= q(x)
dx

dM dx dS dx
+ ! M0 / +M
"M / + dx " S " (S + dx) = 0
dx 2 dx 2
note: q(x) has no net moment about O.
dM 1 dS
dx ! Sdx + (dx) 2 = 0
dx 2 dx

but (dx) 2 is a higher order (small) term


dM
" =S
dx
Summarizing:
dF
=0 (unless a bar)
dx
dS
=q
dx
dM d2M
=S (and = q)
dx dx 2

This is a useful check, and also useful to automate process

Now we will move on to analyze how these internal forces and moments correspond to
the internal stress distribution and deformation of the beam.
Engineering Beam Theory
Reading: Benham, Crawford and Armstrong Ch 6 and Ch 7
We have looked at the statics of a beam and we have seen that loads are transmitted by
internal forces: shear forces and bending moments.
Now look at how these forces imply stresses, strains, and deflections.
Recall model assumptions: slenderness General, symmetric, cross section

Geometry
L >> h,b
Loads
Leads to assumptions on stresses
! yy = ! xy = ! yz = 0
Load in x - z plane →
Also L >> h,b implies ! xx ,! xz >> ! zz

Consider moment equilibrium of a cross section of a beam loaded by some distributed


stress σzz, which is reacted by axial stresses on the cross section, σxx
! zz h
# M X = 0 = ! zz L " ! xx h $ % %0
! xx L
i.e. geometry of beam implies σxx >> σzz

Assumptions on Deformations:
The key to simple beam theory is the Bernoulli - Euler hypotheses (1750)

"Plane sections remain plane and perpendicular to the mid-plane after deformation."

It turns out that this is not really an assumption at all but a geometric necessity, at least
for the case of pure bending.
To see what the implications are of this, consider a beam element, which undergoes
transverse (bending)
deformation.

w= deflection of midplane/midline (function of x only)

Obtain deflection in x-direction, displacement u of point k to K, defined by rotation ! .


Note, key assumption, "plane sections remain plane"
axial displacement, u, of an arbitrary point on the cross-section arises from rotation of
cross, sections
u= -z sin θ

Note, negative sign here due to use of consistent definitions of positive directions for w, x
and dw/dx.

If deformations/angles are small sin ! " !


dw dw
!" # u = $z (1)
dx dx
Hence obtain deformation field
dw
u( x, y,z ) = !z
dx
v( x, y,z ) = 0 nothing happening in y direction
w( x, y,z ) = w( x) Deflection out of original plane
i.e., cross sections remain rigid (2)

Hence we can obtain distributions of strain. (compatible with deformation). In absence of


deformations in transverse directions partial derivatives can be replaced by regular
!_ d_
derivatives, i.e. " . Hence.
!_ d_
"u d 2w
! xx = = # (3)
"x dx 2
"v "w
! yy = = 0, ! zz = =0
"y "z
% "u " v ( % "v "w (
$ xy = 2!12 = ' + * = 0, $ yz = ' + * = 0
& "y " x ) & "z "y )
%"u "w ( % dw dw (
$ xz = ' + * = '# + *=0
& "z "x ) & dx dx )
If no shear - consistent with B-E assumption of plane sections remaining plane.
dM
for constant bending moment: =S=0 (will revisit for S ! 0)
dx
Next use stress - strain (for isotropic material)

" xx (4)
! xx =
E
"
! yy = #$ xx
E
"
! zz = #$ xx
E
& xy
% xy = Note inconsistency -
G ! xz = 0, " xz # 0 (shear
& yz forces are non zero)
% yz =
G
&
% xz = xz
G
The inconsistency on the shear stress/strain arises from the plane/sections remain plane
assumption. Does not strictly apply when there is varying bending moment (and hence

non-zero shear force). However, displacements due to ! xz are very small compared to

those due to ! xx , and therefore negligible.


Finally apply equilibrium:

!" mn
+ fm = 0
!x m
!" xx !# xy !# zx !" xx !# zx
+ + = 0$ + =0 (5)
!x !y !z !x !z
!" xy !# yy !" zy
+ + =0 $0= 0
!x !y !z
!" xz !" yz !# zz !"
+ + = 0 $ xz = 0
!x !y !z !z

Summarizing:
dw
u( x, y,z ) = !z (1)
dx
w( x, y,z ) = w( x) (2)

d 2w
! xx = "z (3)
dx 2
" xx
! xx =
E (4)
!" xx !# zx
+ =0
!x !z (5)

5 equations, 5 unknowns: w, u, ! xx , " xx , # xz

Solutions: Stresses & deflections


First consider relationship between stresses and internal force and moment resultants for
beams (axial force, F, shear force, S, and bending moment M). Consider rectangular
cross section (will generalize for arbitrary cross-section shortly)
Resultant forces and moments are related to the stresses via considerations of
equipollence:
h
2
F = " # xxbdz (6)
!h
2
h
2
S = ! # " xz bdz (7)
!h
2
h
2
M = ! # " xx bzdz (8)
!h
2

Now we combine and substitute between the equations derived above.


combine (3) and (4) to obtain:

d 2w
! xx = E" xx = #Ez 2
dx
Now substitute this in (7)
h2
h
d 2w 2 " d 2 w z2 %
F = !E 2 ( zbdz = $!E 2 b' =0
dx !h $# dx 2 &'
2 !h 2

Which is correct, since there is no axial force in the case of pure beam bending.

Similarly in (8)
+h
2
d w 2
2
M= E 2 " z bdz
dx !h
2 .
We can write this more succinctly as:

d 2w
M = EI 2 this is the moment curvature relation - positive moment results in
dx positive curvature (upward). The quantity “EI” defines the
curvature stiffness of beam – sometimes called “flexural rigidity” – note
that it depends on both the shape and the material property, E.
h
2 2
Where I = " z bdz
!h
2
is the second moment of area (similar to moment of inertia in dynamics)

d 2w M
Recall that ! xx = "Ez 2 = "Ez
dx EI
This is the moment - stress relationship
Mz
! xx = " Implies a linear variation of stress,
I
Maximum stress at edges (top and bottom) of beam

More general definition of I :


See Benham, Crawford and Armstrong Appendix A
For general, symmetric, cross-section
h 2 + y(z)
I yy = ! z 2 dA = ! ! z 2 dydz Iyy is second moment of area about y axis.
A "h 2 "y(z)
Note that there is also a moment of inertia about the z axis – Izz. In general we will be
concerned about the possibility of bending about multiple axes, and the bending of non-
symmetric cross-section. However, we will not consider these cases in SESG1001, see
2nd year structures courses.
1
Note, for particular case of a rectangular section I = bh 3
12
4
[L ]

We need to define centroid: centre of area (analogous to centre of mass). Position


calculated by taking moments of area about arbitrary axis. Hence z position given by:
! zdA
zCentroid = A No net moment about centroid
! dA
A

In bending centroid is unstrained – also known as the “neutral axis” or “neutral plane”
Can also use parallel axis theorem to calculate second moments of area

I total = ! I i + ! Ai d2i

Distance of centroid from total


Second moment area section centroid
of area about
centroid of area
Useful to simplify problems, e.g. for I beam cross-section (structurally efficient form,
effectively used as spars in many aircraft).

is equivalent to:

1
bh 3 b3
12

I total = I y1 y1 + A1d12 + I y 2 y 2 + I y 3 y 3 + A3 d 32

Changing area distribution by moving material from the web to the flanges increases

section efficiency measured as I - resistance to bending for a given amount of material


A
used.
Shear Stresses in Engineering Beam Theory
Note for information – not to be lectured
Reading: Benham Crawford and Armstrong 6.14

Returning to the derivations of simple beam theory, the one issue remaining is to
calculate the shear stresses in the beam. We would like to obtain an expression for:
! zx (z ) .
Recall:
Shear stresses linked to axial (bending) stresses via:
!" xx !# zx !# !"
+ = 0 $ zx = % xx
!x !z !z !x (5)
Also, shear force (for rectangular cross-section, width b) linked to shear stress via:

h
2
S = ! # " xzbdz
!h
2 (7)

Multiply both sides of (5) by b and integrate from z to h/2 (we want to know ! zx (z )

" h%
and we know that ! zx $ ± ' = 0 )
# 2&
h h
2 !" zx !% xx 2
"Mz
#b dz = $ # bdz ! xx =
z !z z !x Iyy
h
# "dM & zb
h 2
[ xz ]
b! (z) = " ) %$ 2
z
dx
(
' I
dz
z

dM
recall =S
dx
# zb &
h

( ( ) )
2
% (
2 " ! xz (z ) = )z S % I ( dz
b ! xz h
$ yy '
h
Note that at z = (top surface) ! zx = 0
2
Also define
h
2
Q = ! zbdz first moment of area above the center
z
Hence:
SQ
! xz (z) = "
I yy b
shear stress -force relation
For a rectangular section:

h h2
2 !z2 $ b !h 2 $
2
Q = ' zbdz = # b& = # (z &
z #" 2 &% 2 #" 4 &%
z

Parabolic, maximum at z = 0 . The centerline. Minimum = 0 at top and bottom


"ligaments"
For non-rectangular sections this can be extended further if b is allowed to vary as a
SQ
function of z. I.e. can write: ! xz (z) = "
I yy b( z)

But need to beware cases in which we have thin walled open or closed sections where

shear is not confined to ! xz . Will also have ! xy (on flanges). E.g.

or

We will defer discussion of these cases until 2nd year courses. Next: Look at examples of
applying Engineering Beam Theory to examples of beams.
Lectures 17 and 18 Examples of Application of Beam
Theory
Reading: Benham, Crawford and Armstrong Chs 6 and 7
Example: Calculate deflected shape and distribution of stresses in beam. Refer back to
example problem in lecture 13 & 14.
Example: Uniform distributed load, q (per unit length), applied to simply supported
beam.

Free Body diagram:

Apply method of sections to obtain bending moments and shear forces:

qL
S(x) = ! qx :
2
qLx qx 2
M (x) = ! :
2 2

Let beam have rectangular cross-section, h thick, b wide. Material, Young’s modulus E:

Deflected shape of the beam: obtain via moment curvature relationship:

d 2w d 2w M (x)
EI =M! =
dx 2 dx 2 EI
Integrate to obtain deflected shape w(x)
M (x)
w( x ) = ! ! dx dx
EI
and for EI constant over length of the beam:
1
! w( x ) = " " M (x)
EI
For our case:
!qx
M (x) = (x ! L)
2
dw 1 qx
= ! " (x ! L)dx
dx EI 2
!q )# x 3 Lx 2 &,
= ++% ! (. + C1
EI *%$ 6 4 ('.-

1 q# x 3 Lx 2 &
So w(x) = "( % ! ( + C1 )dx
EI 2%$ 6 4 ('

!q " x 4 Lx 3 %
w(x) = $ ! + C1 x + C2 '
2EI $# 24 12 '&

C1 and C2 are calculated from values of w(x) and w’(x) at boundary conditions.

Two constants of integration, need two boundary conditions to solve.


x = 0, w = 0 ! C2 = 0
x = L, w = 0
!L4 L4 !L3
0= + + C1 L = 0 " C1 =
24 12 24

Hence:

w(x) = !
q
48EI
[
x 4 ! 2Lx 3 + L3 x ]
dw(x)
Maximum deflection occurs where: =0
dx
dw L
! 4 x 3 " 6 L x2 + L3 = 0 satisified at x =
dx 2

5 qL4
w=!
384 EI yy
Find Stresses:
Mz 1 3
! xx = " , I= bh , substitute for M(x)
I yy 12

L qL2
Max tensile (bending) stress at x = , M max = , z = ±h2
2 8

3qL2
! xx max =
4bh 2

Shear stresses (Not lectured on in 2009):


"SQ qL
! xz = occurs where max shear force S = @ x=0, x=L
Ib 2
for rectangular cross-section
h
2
Q(z) = " z !bdz !
z
h
# bz !2 & 2 b # h 2 &
=% ( = % ) z2(
%$ z ! (' 2 %$ 4 ('
z

bh 2 qL 12 1 3 qL
! xzmax = =
8 2 bh 3 b 4 bh

Note that this is much smaller than ! xx if L/h is large (i.e slender beam).
max

Application of Simple Beam Theory: Discontinuous Loading and


Statically Indeterminate Structures
See Benham, Crawford and Armstrong Ch 8

Example

Free body diagram


Shear forces:

Bending moments:

"Mz
Note in passing M(x) and hence ! xx = = constant for L # x # 2L - useful for
I
mechanical testing of materials – “four point bending”
Could solve for deflections by writing separate equations for: M(x) for 0 < x < L, L <
x < 2L, 2L < x < 3L and integrating each equation separately and matching slope and
displacements at interfaces between sections. This is a tedious process.

Or we can use:.
Macaulay’s Method (BCA 7.3)
Relies on the knowledge that the integral of the moment (i.e. the slope, and deflection)
are continuous, smooth functions. Beams do not develop “kinks”. This is a particular
example of a class of mathematical functions known as “singularity functions” which are
discontinuous functions, defined by their integrals being continuously defined.

Apply method of sections to the right of all the applied loads

Write down equilibrium equation for section, leaving quantities such as {x-L} grouped
within curly brackets { }

M x=x = 0 M + P{x ! 2L} + P{x ! L} ! Px = 0

{x-a} is treated as a single variable, only takes a value for x > a otherwise { } = 0
Proceed as before - integrate moment-curvature relationship twice, applying boundary
conditions
d2w
M = Px ! P(x ! 2L) ! P(x ! L) = EI
dx2

dw Px 2 P 2 P 2
EI = ! {x ! 2L} ! {x ! L} + A
dx 2 2 2
Px 3 P 3 p 3
EIw = ! {x ! 2L} ! {x ! L} + Ax + B
6 6 6

apply boundary conditions: w=0@x=0


w = 0 @ x = 3L
Px3
for x = 0 only 6 is evaluated = 0, therefore B = 0

27PL3 PL3 P
x = 3L !
6
"
6
( )
" 8L3 + A x= 0
6
for
! A = "PL2

1 " Px3 P P %
$ ! { x ! 2 L} ! {x ! L} ! PL x ' (
3 3 2
w( x) =
EI $# 6 6 6 '&

Provides complete description of deflected shape of beam in one equation. Can also use
for distributed loads.

Statically Indeterminate Beams (not lectured in 2009)


Reading: Benham, Crawford and Armstrong Ch 8.

(Two principal ways to approach.)


1.) By simultaneous application of three great principles
i.e., Set up: equilibrium Solve simultaneously
Moment - curvature (reactions remain as unknowns in moment-
curvature equations)
Geometric Constraints

2.) Superposition
Decompose problem into several statically determinate structures.
Replace constraints by applied loads
Solve for loads required to achieve geometric constraints.
Example: cantilever beam under uniform distributed load, with pin at right hand end.
RB = ?

Equivalent to:

(1)

+
(2)

From “standard solutions” or calculation obtain expressions for deflections in case (1)
and case (2):
q! 4
!1 = " ,
8EI
R L3
!2 = + B
3EI

!1 + !2 = 0
qL4 RB L3 3qL
" = " RB =
8EI 3EI 8

This is a powerful technique for solving beam problems – statically indeterminate, or


otherwise complex loading.

Note. For superposition to work we must be superimposing the loads and boundary
conditions on the same beam. Only change one variable in each loading case.
L19 Shafts: Torsion of Circular Shafts
Reading: Benham, Crawford and Armstrong, Ch 4

A shaft is a structural member which is long and slender and subject to a torque
(moment) acting about its long axis. We will only consider circular cross-section shafts
in SESG1001. These have direct relevance to circular cross-section shafts such as drive
shafts for engines of all types. However, the basic principles are more general and will
provide you with a basis for understanding how structures with arbitrary cross-sections
carry torsional moments. Torsional stiffness, and the shear stresses that arise from
torsional loading are important for the design of aerodynamic surfaces such as wings, and
also play a key role in the design of ship hulls, automotive chassis and even hip and knee
replacement prosthetics.
Modelling assumptions
(a) Geometry (as for beam). Long slender, L >> r (b,h)

Note: For the time being we will work in tensor notation since this is all about shear
stresses and tensor notation will make the analysis more straightforward. Remember
we can choose the system of notation, coordinates to make life easy for ourselves!
(b) Loading
Torque about x1 axis, T (units of Force x length). We may also want to consider
the possibility of distributed torques (Force x length/unit length) (distributed aerodynamic
moment along a wing, torques due to individual stages of a gas turbine)

No axial loads (forces) applied to boundaries (on curved surfaces with radial normal, or
on x1 face)
!11 = ! 22 = ! 33 = 0

(c) Deformation
-Cross sections rotate as rigid bodies through twist angle φ, varies with x1 (cf. beams
– plane sections remain plane and perpendicular)
- No bending or extensional deformations in x1 direction

Cross-section:

Torque, T Positive x2, x3


right hand quadrant
rule
positive

u2 = "r# (x1 )sin $


u3 = +r# (x1 )cos $

where r = x22 + x 32

x x
sin $ = 3 cos $ = 2
r r

!
# & x3
so u2 = "% x22 + x 32 () (x1 ) = ") (x1 )x 3
$ '
x22 + x 32

" % x2
u3 = $ x22 + x 32 '( (x1 ) = ( (x1 )x2
# &
! x22 + x 32

So

!
u1 = 0 (1)
u2 = "# (x1 )x 3 (2)
u3 = # (x1 )x2 (3)

Governing Equations
du !
!11 = 1 = 0
dx1
du2
! 22 = =0 Consistent with assumption of no axial,
dx2 or radial stresses, strains
du 3
! 33 = =0
dx 3

1 #% "u1 "u 2 & 1 d*


!12 = + ( = ) x3 (4 )
2 $ "x2 "x1 ' 2 dx1

1 #% "u1 "u3 &( 1 d*


!13 = + = x2 (5)
2 $ "x3 "x1 ' 2 dx1

1 #% "u 2 "u3 &( 1


! 23 = + = ( )*(x1 ) + *(x1 )) = 0
2 $ "x3 "x2 ' 2
Next, apply constitutive laws, i.e. stress-strain relations (assume isotropic),
# mn
remember we are working in tensor notation: " mn = , therefore need “2G”
2G
"
! 23 = 23 = 0
2G
"!
!12 = 12 (6)
2G
"
!13 = 13 (7)
2G

Net moment due to shear stresses must equal resultant torque on section:

equipollent torque, T = $$ ( x2" 13 # x 3" 12 )dx2 dx 3 (8)

"# mn
Apply equilibrium: + fn = 0
"x m
"# 11 "# 21 "# 31
i.e.: + + =0
"x1 "x2 "x 3
! "#
12 + "# 22 + "# 32 = 0
"x1 "x2 "x 3
! "# 13 "# 23 "# 33
+ + =0
"x1 "x2 "x 3
!

!
Retaining non-zero terms we obtain
"# 21 "# 31
+ =0 (9)
" x2 "x 3
"# 12
=0 (10)
"x1
"# 13
=0 (11)
"x1

Solution
Go back to stress-displacement relationships (4,5, 6 and 7)
!
% 1 d$ (*
!12 = 2G"12 = 2G' # x 3
& 2 dx1 )
d#
!12 = "Gx3
dx1

Similarly:
$1 d# ') d#
!13 = 2G13"13 = 2G& x2 = Gx2
%2 dx1 ( dx1

Substitute into equation for resultant moment (8)

T = ## (x 2!13 " x 3!12 )dx2 dx3

% d$ d$ (*
= ## ' x22 + x32 dx dx
& dx1 dx1 ) 2 3
d!
T =G
dx1
( 2
)
2
"" x2 + x3 dA

! 2 2
define J = %% " x2 + x 3 #$ dA

Polar 2nd moment of area

!R 4
= for circular cross-section
2
d!
Hence T = GJ
dx1

Torque-twist relation

d! !2 w
(note, we can compare T = GJ for shafts with M = EI for beams)
dx1 !x 2

Hence, relate stress to torque


Tx
!12 = " 3
J
Tx2
!13 =
J
Express total shear as shear stress resultant, τ

2 2 T Tr
" res = # 12 + # 13 = x 32 + x22 =
J J
Tr $ Mz '
compare " = with &# xx = )
J % I (
!
Model works well for circular cross-sections and cylindrical tubes J = J1 ! J2

!
Does not work for open sections:

We can approximate for other sections, e.g. square cross-section, J = 0.141a4 .

For a full treatment of torsion of slender members see 2nd year structures classes.
L20 Buckling of Simple Columns
Benham Crawford and Armstrong: Chapter 10
Elastic instabilities, of which buckling is the most important example, are a key limitation
on structural integrity. The key feature of an elastic instability is the transition from a
stable mode of deformation with increasing applied load to an unstable one, resulting in
collapse (loss of load carrying capability) and possibly failure of the structure. Examples
of elastic collapse are the buckling of bars in a truss under compressive load, the failure
of columns under compressive load, the failure of the webs of “I” beams in shear, the
failure of skin panels or hull plates in shear and many others. The only particular case we
will consider in SESG1001 is the failure of a bar or column loaded in axial compression,
however, as for the other slender members we have considered, the basic ideas will apply
to more complex structures.
A structure is in stable equilibrium if, for all possible (small)displacements/deformations,
a restoring forces arise.
Now apply this idea to a continuous structural member.
Model of a column
(a) Geometry – identical to a beam, long, straight, slender, symmetric cross-section etc.

L<< b , h
(b) Loading – axial compressive axial forces

(c) Deflection
At low loads - same as a rod, axial (x) stresses, axial deformation only

At higher loads - buckling deflection (transverse) governed by bending relations


d 2w M
i.e., =
2 EI
dx
deflected shape:

Overall free body diagram:

We can take a “cut” (as before)


By equilibrium

+
" Fx = 0 ! P + F(x) = 0 # F ( x) = $P
" F2 = 0 % + S(x) = 0
"MA =0 + M (x) $ F(x)w(x) = 0
M (x) + Pw(x) = 0

Substitute for curvature, obtain second order differential equation

d 2w
EI + Pw = 0
2
dx

Tends to destabilize -
Stabilizing compressive load
(restoring) As w increases

Solution to equation given by:

w = e !x

Write as:

d 2w P
+ w=0
dx 2 EI
P
!2 e !x + e !x = 0
EI
"P P
!2 = # !± i i = "1
EI EI

Can express as -complete homogeneous solution:

P P
W = A sin x + B cos x + Cx + D
EI EI
For the simply supported case:

x=0 w=0
d 2w d 2w
M = EI =0! =0
dx 2 dx 2
x=L w=0
d 2w d 2w
M = EI =0! =0
dx 2 dx 2
d 2w P P P P
=! A sin x! B cos x
dx 2 EI EI EI EI

So using B.C.’s

w( x = 0) = 0 ! B + C = 0 B=0
C=0
d 2w
( x = 0) = 0 ! B = 0
dx 2

P
w( x = L ) = 0 ! + A sin L + DL = 0 A, B, C, D = 0
EI
d 2w P
( x = L ) = 0 ! + A sin L =0
dx 2 EI

! P $
So: A sin# L& = 0
" EI %
Satisfied if A=0, but that is not very interesting!
P P
So: sin L =0! L = n"
EI EI
integer
Thus buckling occurs if:

2 2
P = n ! 2 EI
L

deformed shape

n!x
w = A sin Note: A is not
L defined

n=1

only achieved if 1st mode


n =2 is restrained

n=3

lowest value is the critical buckling load:

! 2 EI
Pcrit =
L2
Other boundary conditions provide other values for A, B, C, & D, solution always of the
form:

P P
w = A sin x + B cos x + Cx + D
EI EI
and buckling load always has the form:

c!2 EI
Pcr =
L2
Simply Supported clamped-free clamped-clamped
c=1 c=1/4 c=4

You might also like