You are on page 1of 101

STRENGTH OF MATERIALS

CHAPTER 1: INTRODUCTION

SECTION I. EXTERNAL EQUILIBRIUM


The concept of equilibrium is the most central one in the subject of Statics. When the net effect
or the resultant of all the forces (and couples) acting on a system is zero, the system is said to be
in equilibrium. Thus, based on the resultant of all the forces R , and the resultant of all the
moments (couples) M , the vector equations of equilibrium are

R  F  0  M M 0
and (1.1)
The two vector equations of equilibrium can be expressed alternatively as scalar equations of
equilibrium for a system of forces in 3 dimensions ( x , y and z ), as

 F  0,  F  0,  F  0
x y z
(1.2)
 M  0,  M  0,  M  0
x y z
(1.3)

Here,  x represents the algebraic summation of components of all the forces in x-direction.
F

This summation is the same as the resultant (net effect) of all the forces in x-direction.
This set of six equilibrium equations can be narrowed down to three scalar equations in case of
a planer force system (forces acting in two dimensions only)
F x  0, F y  0, M z 0
(1.4)

Figures 1.1 and 1.2 illustrate how resultants are obtained for a two-dimensional (planer) force
system.
C y
P1 P1

3 L
2
x
RAx
A B
P2 P2
L/2 L/2 RAy RBy
Fig. 1.1 Obtaining resultants for a truss

F x  R Ax  P1
F y  R Ay  RBy  P2
3 L 3 L
M A  RBy  L  P1 
2
L  P2  ,
2
M B  R Ay  L  P1 
2
L  P2 
2

1
Introduction

P2 P1

A B x
M1
L L L

P 1y
P2 P1
MA
RAx P1X
M1
RAy
Free body diagram
Fig. 1.2 Obtaining resultants for a cantilever system

F  R
x Ax  P1x
F  R
y Ay  R1 y  P2
M  R A 1y  3L  P2  2 L  M 1  M A

These equations provide the necessary and sufficient forces to keep a system in equilibrium. The
omission of a force that is acting on a system or the inclusion of a force that is not acting on the
system produces erroneous results in analyzing the behavior of the system. Hence, it is of utmost
importance to understand exactly what the mechanical system under consideration is and the
forces that are acting on the specific system. A system is a body or a combination of connected
bodies. The bodies can be either rigid or deformable (even fluids can be treated as body). For
Structural Analysis, we will restrict ourselves to the study of rigid and deformable solids only.
For the important task of identifying the forces (and couples) acting on a system, we take the
help of Free Body Diagrams. Thus, drawing a free body diagram becomes the first and foremost
task in solution of problems in Analysis.

The free body diagram of a body (or its part, or a connected system of bodies) is obtained by
isolating it from the all other surrounding bodies. The diagram detaches the system in
consideration from all mechanical contacts with other bodies and sets it free . The other bodies
are not shown in the diagram, but they are replaced by the forces (and couples) that they apply
on the system for which we are drawing a free body diagram.

2
Introduction

SECTION II: SIMPLE STRESSES AND STRAINS


1. STRESSES
Stress: When some external forces are applied to a body, then the body offers resistance to these
forces. The magnitude of the resisting force is numerically equal to the applied forces. This
internal resisting force per unit area is called 'stress'.
However, we name this subject as "Strength of Materials", but at no stage we try to determine
the strength of the material, we always calculate the stress in the material.
In order to understand the concept of stress, consider a body under the action of a number of
forces as shown in Fig. 1-3 (a). If an imaginary cut is made by passing a cross-section 1-1 and
the left portion is drawn separately as shown in Fig. 1-3 (b), now consider the elementary force
F acting on the elementary area A . Then by definition :

(a) Body subjected to many forces (b) Defining stress at a point


Fig.1.3. Understanding stress
F dF
  Limit 
Stress: A 0 A dA (1.5)
2
The dimensional units of stress are N / m (Newton per metre squared) and is always supposed
to act at a point.

Normal stress: In Fig. 1-3 (b), the force F can be resolved into components such that one of
them is along the outward drawn normal n to the area A (since there can be only one normal
at a point) and the other components lie in the plane of the area A . Let Fn be

Fn dFn
 n  Limit 
A0 A dA (1.6)

The normal stress may be tensile or compressive depending upon the forces acting on the
material to be either of the pull or push type respectively. Tensile and compressive stresses
together are called direct stresses.

Shear stress: The force F may be resolved into infinite number of components in the plane
containing area A , because there are infinite number of directions in the plane containing area
A which are perpendicular to the unit normal n . However, if we restrict our studies to three-

3
Introduction

dimensional co-ordinate system, then we are left with only two directions x and y perpendicular
to each other as shown in Fig. 1-3 (b). Then the shear stresses are defined as:
Fx dFx
 x  Limit 
A0 A dA
Fy dFy
 y  Limit 
A0 A dA (1.7)

2. STRAINS
Strain: It is defined as the change in length per unit length. The strain may be tensile or
compressive depending upon whether the length increases (under tensile load) or decreases
(under compressive load). It is a dimensionless quantity.
Normal strain: It is the strain produced under the action of direct or normal stresses.
Shear strain: It is the strain produced under the action of shear stresses. The shear strain is
measured by the change in the angle. Thus in Fig. 1-4, if dl is the change in the length of
face CD under the action of shear force F, then by definition,
Shear strain  = tan 
For small strains,
dl

tan  =  , thus l (1.8)

Fig.1.4. Shear strain

4
Tension and Compression

CHAPTER 2: TENSION AND COMPRESSION

SECTION I. STRESS AND STRAIN IN TENSION (OR COMPRESSION)

1. Direct stress (  )
It has been noted above that external force applied to a body in equilibrium is reacted by internal
forces set up within the material. If, therefore, a bar is subjected to a uniform tension or
compression, i.e. a direct force, which is uniformly or equally applied across the cross-section,
then the internal forces set up are also distributed uniformly and the bar is said to be subjected
to a uniform direct stress, the stress being defined as stress (  ) = load/area = P/A

Stress σ may thus be compressive or tensile depending on the nature of the load and will be
measured in units of Newton’s per square metre (N/m2) or multiples of this.
In some cases the loading situation is such that the stress will vary across any given section and
in such cases the stress at any point is given by the limiting value of P / A as A tends zero.

2. Direct strain (  )
If a bar is subjected to a direct load, and hence a stress, the bar will change in length. If the bar
has an original length L and changes in length by an amount L , the strain produced is defined
as follows:
strain in length δL


strain ( )= original length L

Strain is thus a measure of the deformation of the material and is non-dimensional, i.e. it has
units; it is simply a ratio of two quantities with the same unit (Fig. 2.1.1).

Since, in practice, the extensions of materials under load are very small, it
is often convenient to measure the strains in the form of strain x 10-6, i.e.
microstrain, when the symbol used becomes  .

3. Sign convention for direct stress and strain


Tensile stresses and strains are considered POSITIVE in sense.
Compressive stresses and strains are considered NEGATIVE in sense.
Thus a negative strain produces a decrease in length:

4. Definitions
 Elastic materials — Hooke's law
A material is said to be elastic if it returns to its original, unloaded dimensions when load is
removed. A particular form of elasticity which applies to a large range of engineering materials,
at least over part of their load range, produces deformations which are proportional to the loads
producing them. Since loads are proportional to the stresses they produce and deformations are

5
Tension and Compression

proportional to the strains, this also implies that, whilst materials are elastic, stress is proportional
to strain. Hooke's law, in its simplest form, therefore states that:
stress

Stress ( )  strain ( ) ; strain constant
It will be seen in later sections that this law is obeyed within certain limits by most ferrous alloys
and it can even be assumed to apply to other engineering materials such as concrete, timber and
non-ferrous alloys with reasonable accuracy.
Whilst a material is elastic the deformation produced by any load will be completely recovered
when the load is removed; there is no permanent deformation.

 Modulus of elasticity — Young's modulus


Within the elastic limits of materials, i.e. within the limits in which Hooke's law applies, it has
stress

been shown that strain constant
This constant is given the symbol E and termed the modulus of elasticity or Young's modulus.
Thus
stress 
E  (2.1)
strain 
P L PL
   (2.2)
A L AL
Young's modulus E is generally assumed to be the same in tension or compression and for most
engineering materials has a high numerical value. Typically, E = 200 x 109 N/m2 for steel, so
that it will be observed from (2.1) that strains are normally very small since


E (2.3)
In most common engineering applications strains do not often exceed 0.001 or 0.1%.
The actual value of Young’s modulus for any materials is normally determined by carrying out
a standard tensile on a specimen of the material.

 Poisson's ratio
When a material is subjected to longitudinal deformation then the lateral dimensions also change.
The ratio of the lateral strain to longitudinal strain is a constant quantity called the Poisson's ratio
and is designated by  or 1/m.
Lateral strain

Longitudinal strain (2.4)
 Modulus of rigidity
It is defined as the ratio of shearing stress to shearing strain, i.e.

G
 (2.5)

6
Tension and Compression

 Factor of safety
Because of uncertainties of loading conditions, we introduce a factor of safety, defined as the
ratio of the maximum stress to the allowable or working stress. The maximum stress is generally
taken as the yield stress. This is also called the ‘factor of ignorance’

 Free body Diagram


The free body diagram of an element of a member in equilibrium is the diagram of only that
member or element, as if made free from the rest, with all the internal and external forces acting
on it.

 Coefficient of Linear Thermal Expansion


Linear thermal strain (  T ) due to change in temperature ( T ) is obtained by using this
coefficient (  ).  has units of per degrees Centigrade (or Fahrenheit)
 T  T (2.6)

5. Bar of Varying Cross-section


Consider a bar of varying circular cross-section as shown in Fig. 2.1.2 and subjected to axial
load P throughout. The area of different cross -sections is:

P P
1 d1 2 d2 d3
3

L1 L2 L3

Fig. 2.1.2 Bar of varying cross-section.


   2
A1  d12 , A2  d 22 , A3  d3
4 4 4

Let  1 ,  2 and 3 be the corresponding stresses, then,
P P P
1  , 2  , 3 
A1 A2 A3
The strains become,
  
1  1 , 2  2 , 3  3
E1 E2 E3
The changes in lengths become,
l1   1l1 , l 2   2 l 2 , l3   3l3 ,
Total change in length,
 l l l  n
l
l  l1  l 2  l3  P 1  2  3  l  P  i
 A1 E1 A2 E 2 A3 E3  or in general, we have i 1 Ai E i

7
Tension and Compression

P1 d1 d2 d3 P4

P2 2 P3
1
3

L1 L2 L3

(a)

P1 d1 1P d2 P4 d3 P4
P1 - P2 2 P1 - P2
1
(P4 - P3) 3

(b) Free body diagrams

Fig.2.1.3. Bar of varying cross-section.

If the loads in different sections of the bar are different as shown in Fig. 2.1.3 (a), then free body
diagrams may be drawn for each section as shown in Fig. 2.1.3 (b), and the net forces acting in
each section may be determined. Thus the stresses, strains and total elongation may be deter-
mined.
P P P
 1  1 , 2  2 , 3  3
A1 A2 A3
  
1  1 , 2  2 , 3  3
E1 E2 E3
l1   1l1 , l 2   2 l 2 , l3   3l3
l  l1  l 2  l3
n
Pi li

i 1 Ai Ei

Worked Example 2.1 A mild steel rod 20mm diameter is subjected to an axial pull of 50KN.
Determine the tensile stress induced in the rod and the elongation if the unloaded length is 5m.
E  210GN / m 2 .
Solution:
Given: d=20mm; P=50KN; l=5m
 2 
A d  (20) 2  10 6  314  10 6 m 2
Area of cross-section of the rod: 4 4
P 50  103  5  103
  6
 159.155MN / m 2
Stress A 314  10
Pl 50  103  5  103
   3.789mm
Elongation AE 314  10 6  210  109

8
Tension and Compression

Worked Example 2.2 A short hollow cast iron cylinder of wall thickness 10mm is to carry a
compressive load of 600KN. Determine the outside diameter of the cylinder if the ultimate
2
crushing stress for material is 540KN / m . Use a factor of safety of 6.

Solution:
Let d 0 be the outside diameter of the cylinder in mm. Then area of cross-section of the cylinder
is,

A  d 02  (d 0  20) 2  10 6   (d 0  10)  10 5 m 2
4
Safe load   (d 0  10)  10 m
5 2

 900  (d 0  10)  600103


 d 0  222.2mm

Worked Example 2.3 A round bar as shown in Fig.2.1.4 is subjected to an axial tensile load of
2
100KN. What must be the diameter ‘d’ if the stress there is to be 100MN / m ? Find also the
total elongation. E  210GN / m .
2

Solution:

100KN d 10cm 8cm 100KN

10cm 15cm 15cm

Fig.2.1.4.
P


d2
Stress: 4
100  103
100  10 6

d2
4
4
d  0.03568m  35.68mm
 Diameter,   103

Total elongation,

9
Tension and Compression

P  l1 l2 l 
l     2 
E  A1 A2 A2 
 
 0.10
100  103 0.15 0.15  1
  3     6
200  109
 10  
 (100) 2  (80) 2  10
 4 4 
 0.0745mm

Worked Example 2.4 A steel bar 25mm diameter is loaded as shown in Fig.2.1.5. Determine
the stresses in each part and the total elongation. E  210GN / m .
2

Solution:
A B C D
40KN 20KN 30KN
10KN

50cm 40cm 20cm

Fig.2.1.5
A B B C C D
40KN 40KN 30KN 30KN
20KN 20KN

Fig.2.1.6 Free body diagrams.


Area of cross-section,
 2 
A d  2
 (25)  10 6  490.87  10 6 m 2
4 4
The free body diagrams for each portion have been shown in Fig.2.1.6.
Stress in various parts are:
40  103
 AB  6
 81.488 MN / m 2
490.87  10
20  103
 BC  6
 40.744 MN / m 2
490.87  10
30  103
 CD  6
 61.166 MN / m 2
490.87  10
1
l 
AE
 Pi li
Total elongation
103
 40  0.5  20  0.4  30  0.2  0.3298mm
490.87  106  210  109

10
Tension and Compression

Exercise Problems
2
E1: A steel bar as shown in Fig.E.1 consists of two parts AB and BC having areas of 4cm and
5cm 2 respectively. It is rigidly fixed at end A and end C at a distance of 1mm from the other rigid
horizontal support. A load of 100KN is applied vertically downward at B. Determine the
reactions produced by the rigid horizontal support and the stress in the parts AB and BC of the
bar. E  210GN / m .
2

ANS; l AB  1.5625mm, P  20KN , AB  200Mpa, BC  40Mpa,

E2: A bar of length 5mm is made of two materials as shown in Fig.E.2. The first 3m of its length
2
is made of brass and is 7.5cm in cross-section and the remainder of its length is of steel and is
5cm 2 in cross-section. Determine the total compression of the bar under of 20KN.
Esteel  210GN / mm 2 , Ebrass  84GN / mm 2 .

ANS; l B  0.95238mm, l S  0.3809mm,   l B  l S  1.333mm

E3: A prismatic bar as shown in Fig.E.3 carries an axial load 10KN. Calculate the reaction at
the supports assuming them rigid.
ANS; RC  10 / 3KN , R A  20 / 3KN
20 KN
A

4cm 2 1.25m
Steel 2m
A B C
B 10KN
100KN
2 1.25m Brass 3m
5cm
1m 2m
C

1 mm

Fig.E.1 Fig.E.2 Fig.E.3

11
Tension and Compression

SECTION II: STRESS –STRAIN DIAGRAM

Ductile Materials:
Fig. 2 -1 shows the stress-strain diagram for a ductile material like mild steel. The curve starts
from the origin 0 showing thereby that there is no initial stress or strain in the test specimen. Up
to point ‘ a ’ Hooke's law is obeyed and stress is proportional to strain. Therefore, oa is a straight
line and point a is called the limit of proportionality and the stress at point a is called the

proportional limit stress, p . The portion of the diagram between ab is not a straight line but up
to point b, the material remains elastic, i.e. on removal of the load, no permanent set is formed
and the path is retraced. The point b is called the elastic limit
point and the stress corresponding to that is called the elastic
limit stress,  c .
In actual practice, the points a and b are so close to each other
that it becomes difficult to differentiate between them.
Beyond the point b, the material goes to the plastic stage
until the upper yield point ‘c’ is reached. At this point the
cross-sectional area of the material starts decreasing and the
stress decreases to a lower value to a point d, called the lower
yield point. Corresponding to point c, the stress is known as

upper yield point stress, yu and corresponding to point d, the

stress is known as lower yield point stress, yi . At point d
the specimen elongates by a considerable amount without
any increase in stress and up to point e. The portion de is
called the yielding of the material at constant stress. From
point e onwards, the strain hardening phenomena becomes
predominate and the strength of the material increases
thereby requiring more stress for deformation, until point f is reached. Point f is called the
ultimate point and the stress corresponding to this point is called the ultimate stress,  u . It is the
maximum stress to which the material can be subjected in a simple tensile test. At point f the
necking of the material begins and the cross-
sectional area starts decreasing at a rapid rate.
Due to this local necking, the stress in the
material goes on decreasing inspire of the fact
that actual stress intensity goes on increasing.
Ultimately the specimen breaks at point g,
known as the breaking point, and the corres-
ponding stress is called the nominal breaking
stress based upon the original area of cross
section. Whereas the true stress at fracture is the
ratio of the breaking load to the reduced area of
cross-section at the neck. The initial portions of
the diagram are shown in Fig. 2.2. on
exaggerated scale.
Sometimes it is not possible to locate the yield point quite accurately in order to determine the
yield strength of the material. For Such materials the yield point stress is defined at some

12
Tension and Compression

particular value of the permanent set. It has been observed that if load is removed in the plastic
range then the unloading path line is parallel to the straight portion of the stress-strain diagram
as shown in Fig.2.2(b). The commonly used value of permanent set for determining the value of
yield strength for mild steel is 0-2 percent of the maximum strain as shown in Fig. 2.3.

Brittle Materials:
The stress-strain diagram for a brittle material like cast iron is shown in Fig. 2.4. There is very
little elongation and reduction in area of the specimen for such materials. The yield point is not
marked at all. The straight line portion of the diagram is also very small.

13
Tension and Compression

SECTION III: EXTENSION OF A BAR UNDER IT’S OWN WEIGHT

1. Bar of uniform area.


Consider a bar of uniform cross-section of area A and length L. Consider an elementary strip of
this bar between cross-sections x and (x+dx) as shown in Fig. 2.3.1. The downward force acting
on this strip is due to the weight of the bar that lies below this element and is equal to Ax .
In order that this elementary strip is in a state of equilibrium, a force equal to  Ax must act
upwards.

Ax
  x
The stress in the strip: A
 x
 
Strain in the strip, E E
x
 dx  dx
Extension of the strip E
x L2
0 E
L
 dx 
Total extension of the bar 2E
Fig. 2.3.1. Uniform bar under its own weight.
If W = Total weight of the bar =  AL
WL
:. Total extension= 2 AE
It may be observed that the total extension produced by the self-weight of the bar is equal to that
produced by a load of half its weight applied at the lower end. Therefore, if the weight of the
bar is to be taken into account for calculating extension, half of the total weight of the bar may
be applied at the lower end.

2. Bar of varying cross-section.


Consider a bar of varying cross-section as shown in Fig. 2.3.2. Consider an elementary strip of
the bar at a distance x from lower end and of length dx.
Total force acting on this strip downwards is equal to the weight
of the bar up to x.
Let A be the area of cross-section at x.

 Weight of the bar up to x:


x
  Adx
0

Stress in the strip:


x
  Adx
 0
A
 x Adx dx
  0
L



0

 Extension of the bar AE

14
Tension and Compression

Worked Example 2.5: A metal bar 5 cm  5 cm section is subjected to an axial compressive


load of 500 KN. The contraction on a 20 cm gauge length is found to be 0.5 mm and the increase
in thickness 0.045 cm. Find the value of Young's modulus and Poisson' s ratio.

Solution:
Given p=500KN ,l = 20 cm,  l= 0.05 cm,  t= 0.0045 cm
Area of cross-section,
A = 5  5 = 25 cm2
l 0.05
   0.0025
Longitudinal strain, l 20 (compressive)
P 500  103
   200MPa
Stress A 25  104 (compressive)
 200  106
E   80GPa
Young’s modulus,  0.0025
l 0.0045
   0.0009
Lateral strain l 5 (tensile)
Lateralstrain 0.0009
   0.36
Poisson’s ratio, Longitudinal strain 0.0025

Worked Example 2.6. A bar of 20mm diameter is subjected to an axial tensile load of 120KN,
4
under which 200mm gauge length of this bar elongates by an amount of 3.5  10 m . Determine
the modulus of elasticity of the bar material. If v=0.3, determine its change in diameter.

Solution:
Modulus of elasticity
P l
E 
A l
120  103 200  103
 
 3.5  104
(20  103 ) 2
4
 218.27GN / m 2
d / d

Poisson’s ratio, l / l
d / 20
0.3  4
(3.5  10 ) /(200  103 )
0.3  20  3.5  101
d   0.0105mm
200

15
Tension and Compression

SECTION IV: TEMPERATURE STRESS

When the temperature of a material is changed, its dimensions change. If this change in
dimensions is prevented, then a stress is set up in the material, which is called a temperature
stress.
Let: l = length of a bar at temperature t0
t = increase in temperature
 = coefficient of linear expansion for the material
Change in length of the material = lt
Expanded length = l (1 +  t)
lt
  t
Strain, l
Stress,  = E  t

When the temperature rises, the material is prevented from expanding and, therefore,
compressive stress is induced in the material. On the other hand, when the temperature decreases,
the material is prevented from contracting and thus tensile stress is induced.

Worked Example 2·7. A rod is 2m long at a temperature of 10 C . Find the expansion of the
rod when the temperature is raised to 80 C . If this expansion is prevented find the stress in the
E  100GN / m 2 and   12  10 / ℃ .
6
material of the rod. Take

Solution:
Rise in temperature , t=80-10=70 C
Strain   t  12  10  70  840  10
6 6

Stress   E  100  10  840  10  84MN / m


9 6 2

Worked Example 2·8. Two parallel walls 6m apart are stayed together by a steel rod 2.5cm
diameter at a temperature of 80 C passing through washers and nuts at each end. Calculate the
pull exerted by the rod when it cooled to 22 C :
(a) if the walls do not yield, and
(b) if the total yield at the two ends is 1.5mm
E=200GN/m2
  11 106C

Solution:
(a) Fall in temperature, t=80-22=58 C
Strain,   t  11 106  58  638 106
Stress,   E  200  109  638 106  127.6MN / m 2

16
Tension and Compression


P  A  127.6  106   6.25  104  62.635KN
Pull exerted, 4

Length at 22 C  l (1  t )  6(1  11 10  58)  5.996172m


6
(b)
Decrease in length =6-5.996172=0.003828m
Due to yielding of walls by 1.5mm,
6
The actual decrease in length =0.003838-0.0015=2328  10 m
2328 106
  388 106
Strain 6

P  EA  200  109  388 106   6.25  106  38.092KN
Pull exerted, 4

1. Composite System of Equal Lengths Subjected to Variation of Temperature.


When a composite system is subjected to a change in temperature then different components
change in length by different amounts due to the difference in their coefficients of linear
expansions. A material having higher coefficient of thermal expansion than the other increases
in length more than the other one. If this increase in length is prevented then the material having
higher coefficient of linear expansion will be in a state of compressive stress and the other
material will be subjected to tensile stress. If no external force is applied then the compressive
force P in the material having higher coefficient of linear expansion will be equal to the tensile
force P in the material having lower coefficient of thermal expansion.

Consider a composite system consisting of


two tubes as shown in Fig. 2·4·1 (a), of equal
length l but of different materials. When the
temperature of the whole system is increased
by t°C, then free increase in length of inner
tube is 1tl and the free increase in the length
of outer tube is  2tl , as shown in Fig.2·4·1
(b) and assuming  1 to be greater than  2 .
Difference in increase in length = (1   2 )tl
Now if this difference is eliminated by
compressing the inner tube by a force P and pulling out the outer tube by an equal tensile force
P as shown in Fig. 2·4·1(c), then
Contraction for inner tube due to P,
pl
l1 
A1 E1
Extension of outer tube due to P,

17
Tension and Compression

pl
l2 
A2 E2
The difference in length is eliminated, when
1 1
(1   2 )tl  l1  l 2  pl (  )
A1 E1 A2 E2
(1   2 )t
 p
1 1
(  )
A1 E1 A2 E2

2. Composite system of equal lengths containing more than two components subjected
to variation of temperature.

Consider three bars of different materials, each of length l , having coefficients of linear
expansions 1 , 2 and  3 and moduli of elasticity E1,E2, E3 respectively, rigidly connected
together at their ends as shown in Fig. 2·4·2. Let t°C be the rise in temperature. Then,
Free extension of bar 1 = 1tl
Free extension of bar 2 =  2tl
Free extension of bar 3 =  3tl
Since their ends are connected together rigidly,
therefore, each bar must elongate by the same amount, l (say).
So actual extension of each of the bars is l .
Thus constrained strains in each of the bars are :

l  1tl l   2tl l   3tl


,
l l and l
Stresses in individual bars are:
 l  1tl   l   2tl   l   3tl 
 
E, 1  E  2   E3
 l   l  and  l 
Forces in each bar are:
 l  1tl   l   2tl   l   3tl 
  E1 A1 ,   E2 A2   E3 A3
 l   l  and  l 
As the composite system is in equilibrium, the total force must be equal to zero.
 l  1tl   l   2tl   l   3tl 
  E1 A1    E2 A2    E3 A3  0
 l   l   l 
l
( A1 E1  A2 E 2  A3 E3 )  t ( 1 A1 E1   2 A2 E 2   3 A3 E3 )
or l
lt (1 A1 E1   2 A2 E2   3 A3 E3 )
l 
or ( A1 E1  A2 E2  A3 E3 )

18
Tension and Compression

or in general for n bars of equal length, we have


n
lt  i Ai Ei
l  i 1
n

AE
i 1
i i

By knowing l , we can determine the magnitudes of stresses and forces in each bar by back
substitution.
If a load P is also applied to the composite system in addition to the temperature rise, then for
the equilibrium of the system,
 l  1tl   l   2tl   l   3tl 
  E1 A1    E2 A2    E3 A3  P
 l   l   l 
l
 ( A1 E1  A2 E2  A3 E3 )  P  t (1 A1 E1   2 A2 E2   3 A3 E3 )
l
l P  t (1 A1 E1   2 A2 E2   3 A3 E3 )
l 
( A1 E1  A2 E2  A3 E3 )
or in general for n bars of equal length, we get
 n

l  P  t   i Ai Ei 
l   n
i 1 

 Ai Ei
i 1

Worked Example 2·9. A weight of 200 KN is supported by three short pillars each 6.25 cm2 in
section as shown in Fig. 2·4·3. The central pillar is of steel and the
two outer ones of copper. The pillars are so adjusted that at a tem-
perature of 15°C, each carries one-third of the total load. The
temperature is then raised to 115°C. Estimate the stress in each pillar
at 15°C and at 115°C. (Take: E =210 GPa, E = 84 GPa  s = s c

12  106 per°C,  c  18.5  10 per °C.


6

Solution,
200103
  106.667 N / mm 2
(a) Stress in each pillars at 15°C 3  625

(b) Temperature rise: t  11515  100C


Let   common strain in the bars.
Strain due to rise in temperature  t

19
Tension and Compression

Net strain due to load carried    t


  (  t) EA  P
(   s t ) E s As  2(   c t ) Ec Ac  P
or
(  12 106 100)  210103  625  2(  18.5 106 100)  84 103  625  200103
 (2.1  2  0.84)  625105    625101 100(12  2.1  2 18.5  0.84)  200103
3.78  625105    62510  56.2  200103
 551.25 103
551.25 103
   2.33333103
3.78  62510 5

 s  (   s t ) E s
 (2333.33106  12 106 100)  210103
 23.807 N / mm 2 (compressive)
 c  (   c t ) Ec
 (2333.33106  18.5 106 100)  84 103
 40.572 N / mm 2 (compressive)

Exercises
E4: A copper bar 25cm long is fixed by means of a sinking support at its ends which yields by
an amount 0.01cm. If the temperature of the bar is raised by 120C , calculate the stresses induced
in the bar. Coefficient of linear expansion for copper ,  c  17.5 10 per C , Ec  98GPa
6

E5: A steel bar 2.5cm diameter is rigidly attached to two parallel supports 8m apart. Find the
pull exerted by the bar on the support when the temperature is increased by 100C (a) if the
supports do not yield, (b) if yielding of both supports is 0.25cm.
 s  12 106 per C , Es  210GPa

20
Shear Force and Bending Moment Diagrams

CHAPTER 3: SHEAR FORCE AND BENDING MOMENT DIAGRAMS

21
Shear Force and Bending Moment Diagrams

Wl
M max 
4 …(3’1)

22
Shear Force and Bending Moment Diagrams

23
Shear Force and Bending Moment Diagrams

24
Shear Force and Bending Moment Diagrams

25
Shear Force and Bending Moment Diagrams

26
Shear Force and Bending Moment Diagrams

27
Shear Force and Bending Moment Diagrams

28
Shear Force and Bending Moment Diagrams

29
Shear Force and Bending Moment Diagrams

30
Shear Force and Bending Moment Diagrams

31
Shear Force and Bending Moment Diagrams

32
Shear Force and Bending Moment Diagrams

33
Shear Force and Bending Moment Diagrams

34
Shear Force and Bending Moment Diagrams

35
Shear Force and Bending Moment Diagrams

ASSIGNMENT (I)
E1:Draw the bending moment and shear force diagrams for the members as shown in Fig.3.15
5 KN 20 KN
10 KN/m 15 KN/m
B 10 KN/m
A C
A B
10 KNm
4m 2m
3m 3m
(a)
(b)

10 KN
°
5 KN/m 45

60°
60
°
B D A E
A B C D
C
10 KNm
3m 2m 2m 2m 2m 2m 2m

d)
(c)

10 KN
10 KN/m 10 KN/m
A C B
B D A B
C D
2m 1m 1m 4m 2m 2m

(e)
(f)

5 KN
10 KN/m
B
A B
2 KNm
2m 2m 4m 2m 2m 2m

(g)

2m

60°

2 KN 10 KN/m

4m 5 KN/m
5m

(i)

(h)

36
Simple Bending of Beams

CHAPTER 4: SIMPLE BENDING OF BEAMS

Definitions
 Centre of Gravity of Plane Figures
The plane geometrical figures (such as T  section, I  section, L  section etc.) have only areas
but no mass. The c.g. of such figures is found out in the same way as that of solid bodies. The
centre of area of such figures is known as centroid, and coincides with the c.g., of the figure.
Let x and y be the co-ordinates of the c.g. with respect to some axis of reference, then
x Ax  Ax
A1 x1  A2 x2  ...
 i i i i

A1  A2  ...A A i
…(4.1)
A y  A y  ...  A y  A y
y 1 1 2
2
 i i i i

A  A  ...
1 2 A A i
…(4.2)
where A1 , A2 ,... etc are the areas, into which the whole figure is divided.
x1 , x2 ,... etc are the respective co-ordinates of the areas A1 , A2 ,... on X  X axis with respect to
same axis of reference.
y1 , y2 ,... etc are the respective co-ordinates of the areas A1 , A2 ,... on Y  Y axis with respect to
same axis of reference.
A = total area of the figure.
 Axis of Reference
The c.g. of a body is always calculated with reference to some assumed axis known as axis of
reference. The axis of reference, of plane figures, is generally taken as the lowest line of the
figure for calculating y and the left line of the figure for calculating x .
 Axis of Symmetry
Sometimes, the given section, whose centre of gravity is required to be found out, is symmetrical
about X  X axis or Y  Y axis. In such cases, the procedure for calculating the c.g. of the body
is very much simplified; as we have only to calculate either x or y . This is due to the reason
that the c.g. of the body will lie on the axis of symmetry.

Worked Example 4.1. Find the centre of gravity of a 10 cm 15 cm  3 cm T  section.


Solution: 10 cm
As the section is symmetrical about Y- Y-axis, bisecting the web, A B
3 cm
therefore the c.g. of the section will lie on this axis. Split up the
H G D C
section into two rectangles ABCH and DEFG as shown in Fig. 4.1.
15 cm
Let y be the distance between c.g. and
the bottom FE, the axis of reference.
F E
A1  10  3  30cm 2
 3
y1  15    13.5cm 3 cm
(i ) Area ABCH  2 Fig.4.1

37
Simple Bending of Beams

A2  12  3  36cm 2
12
y1   6cm
(ii ) Area DEFG: 2

Using the relation,


A1 y1  A2 y2 (30 13.5)  (36  6)
y   9.41cm
A1  A2 30  36

Worked Example 4.2. Find the centre of gravity of a channel section 100  50  15mm .
Solution:
As the section is symmetrical about X  X axis, therefore the c.g. of the section will lie on this
axis. Now split up the whole section into three rectangles ABJH, FGJK and CDEK as shown in
Fig.4.2.
Let x be the distance between c.g. and the face BC, the axis of reference.
(i ) Area ABJH
50
A1  50 15  750mm 2
B A
50 15
x1   25mm J
G H
2
(ii ) Area FGJK 15
100
A2  (100  30)  15  1050mm 2 K
F E
15
x2   7.5mm C D
2
(iii) Area CDEK Fig.4.2
A3  50 15  750mm 2
50
x3   25mm
2
A1 x1  A2 x2  A3 x3 (750 25)  (1050 7.5)  (750 25)
x   17.79mm
Using the relation, A1  A2  A3 750  1050  750

MOMENT OF INERTIA
The moment of inertia or the second moment of the area about a given axis is the product of the
element of area and the square of the distance of the centroid from the axis.
I x   y 2 dA
Thus
I y   x 2 dA
and
or I  AK 2
...(4.3)
where A = area of the figure
K = radius of gyration.

38
Simple Bending of Beams

 Moment of inertia of a rectangular Section


Consider a rectangular section ABCD as shown in Fig.4.3. Y
B
Let b  Width of the section, and
A

d  Depth of the section.


d
X X

Now consider a strip PQ of thickness dy parallel to X  X y


P Q
axis and at a distance y from it as shown in Fig.4.3
 Area of the strip dA  b  dy D C
b
M.I. of the strip about X  X axis is dA  y  bdy  y .
2 2

Y
The M.I. of the whole section about X  X axis can be found out Fig.4.3
d d
 
By integrating for the whole length of d , i.e., from 2 to 2 .
d d d
  
2
y 
2
bd 3 3 2
I XX   by dy b  y dy b   
2 2


d

d  3 d 12
2 2 2

3
db
I YY 
Similarly, 12

 Moment of Inertia of a Hollow Rectangular Section


Consider a hollow rectangular section, in which ABCD is the main section and EFGH is the
cut out section as shown in Fig. 4.4. Y
A B
Let b = Breadth AB of the outer rectangle
d = Depth BC of the outer rectangle
E F
b1 ,d1  Corresponding values for the cut out rectangle
d1 d
We know that the moment of inertia of the outer rectangle X X
b1
bd 3
I XX  H G
ABCD about X-X axis, 12 and moment of inertia of the
b1 d13
I YY  D C

cut out rectangle EFGH about X-X axis, 12 b


 M.I. of the hollow rectangular section about X-X axis, Y

I XX  M.I. of rectangle ABCD- M.I. of rectangle EFGH Fig.4.4

bd 3 b1d13
 
12 12
db3 d1b13
I YY  
Similarly 12 12

39
Simple Bending of Beams

Theorem of perpendicular Axis:


It states, “ If I XX and I YY be the moments of inertia of a plane section about two perpendicular
axes meeting at O, the moment of inertia I ZZ about the axis Z-Z, perpendicular to the plane and
passing through the intersection of X-X and Y-Y is given by the relation”
I ZZ  I XX  I YY

Proof
Consider a small lamina (P) of area da having co-ordinates as x and y along OX and OY two
mutually perpendicular axes on a plane section as shown in Fig.4.5.
Now consider a plane OZ perpendicular to OX and OY. Let r be the distance of the lamina P
from Z-Z axis, so that OP=r. Z

From the geometry of the figure, we find that r  x  y


2 2 2

We know that the moment of inertia of the lamina P about X-X axis, o X
I XX  da  y 2 r y
x p
I YY  da  x 2

And I ZZ  da  r 2  da( x 2  y 2 )  da  x 2  da  y 2  I XX  I YY Y Fig.4.5

 Moment of Inertia of a Circular Section


Consider a circle ABCD of radius r with centre O and XX  and YY  be two axes of reference
through O as shown in Fig.4.6.
Now consider an elementary ring of radius x and thickness dx. Therefore area of ring
da  2x  dx
And M.I. of ring about X  X axis or Y  Y axis
Y

 Area  (Distance)
2
B

 2x  dx  x 2
 2x 3 dx
Now M.I. of the whole section, about the central axis, X O r
X
A x C
can be found out by integration the above equation for
the whole radius of the circle i.e., from 0 to r. Therefore, dx
r
r r
 x4  r 4
I ZZ   2x 3 dx  2  x 3 dx  2    D
0 0  4 0 2
d 4 Y
 Fig.4.6
32
We know from the theorem of perpendicular axis, that
I XX  I YY  I ZZ
I ZZ 1 d 4 d 4
 I XX  I YY    
2 2 32 64

40
Simple Bending of Beams

 Moment of Inertia of a Hollow Circular Section


Consider a hollow circular section as shown in Fig.4.7, whose moment of inertia is required to
be found out.
Let D=Diameter of the outer circle, and
d=Diameter of the cut out circle.
 4
 D
We know that the moment of inertia of the outer circle about X  X axis 64

 d
4 Y

and moment of inertia of the cut-out circle about X  X axis 64


 M.I. of the hollow circular section, about X  X axis,
I XX  M.I. of outer circle - M.I. of inner circle D
d
   X X
  D 4   d 4  (D 4  d 4 )
64 64 64

I YY  ( D 4  d 4 )
Similarly, 64
Y

Parallel axis theorem: Fig.4.7


It states, If the moment of inertia of a plane area about an axis through its centre of gravity, be
denoted by I G , the moment of inertia of the area about axis AB, parallel to the first, and at a
distance h from the centre of gravity is given by:
I AB  I G  Ah 2 ...(4.4)
Where; I AB  M.I. of the area about AB,
IG 
M.I. of the area about c.g.,
A  Area of the section,
h  distance between c.g. of the section and the axis AB.

Worked Example 4.3. An I-section is made up of three rectangles as shown in Fig.4.8. Find
the moment of inertia of the section about the horizontal axis passing through the c.g. of the
section.

Solution:
As the section is symmetrical about Y-Y axis, therefore c.g. of the section will lie on this axis.
Split up the whole section into three rectangles, 1,2 and 3 as shown in Fig.4.8.

Let y be the distance between c.g. of the section and the bottom face.
Rectangle 1
A1  6  2  12cm 2
2
y1  2  10   13cm
2

41
Simple Bending of Beams

Rectangle 2
A2  10  2  20cm 2 6cm
10
y2  2   7cm 1 2cm
2
Rectangle 3
A3  10  2  20cm 2 2cm
10cm
2 2
y 3   1cm
2
Using the relation, 3 2cm

A1 y1  A2 y 2  A3 y3 10cm
y
A1  A2  A3
(12  13)  (20  7)  (20  1) Fig.4.8
  6.08cm
12  20  20
We know that M.I. of rectangle 1 about an axis passing through its c.g. and parallel to X-X axis,
6  23
I G1   4cm 4
12
Distance of c.g. of rectangle 1 from X-X axis,
h1  y1  y  13  6.08  6.92cm
 M.I. of rectangle 1 about X-X axis
 I G1  A1h12  4  12  6.922  578.637cm 4
Similarly M.I. of rectangle 2 about an axis passing through its c.g. and parallel to X-X axis,
2  103
IG2   166.667cm 4
12
Distance of c.g. of rectangle 2 from X-X axis,
h2  y2  y  7  6.08  0.92cm
 M.I. of rectangle 2 about X-X axis
 I G 2  A2 h22  166.667  20  0.922  183.598cm 4
Similarly M.I. of rectangle 3 about an axis passing through its c.g. and parallel to X-X axis,
10  2 3
I G3   6.667cm 4
12
Distance of c.g. of rectangle 3 from X-X axis,
h3  y  y3  6.08  1  5.08cm
 M.I. of rectangle 3 about X-X axis
 I G3  A3 h32  6.667  20  5.082  522.795cm 4
Now moment of inertia of the whole section about X-X axis,
I XX  578.637  183.598  522.795  1285.03cm 4

42
Simple Bending of Beams

 Neutral axis. Neutral axis of a beam is the axis at which the bending stress is zero.
 Simple bending. When the load producing bending lies in the centroidal plane such that
bending is not accompanied by torsion then she bending is said to be simple bending.
 Pure bending. When a beam is subjected to such a system of bending loads so that the shear
force in the beam is zero then the beam is said to be subjected to pure bending. In such a
case the bending moment shall be constant in the beam.

THEORY OF SIMPLE BENDING


When a beam is bent due to the application of a constant bending moment, without being
subjected to shear, it is said to be in a state of simple bending. In simple bending the plane of
transverse loads and the centroidal plane coincide. The following assumptions are made in this
theory:
1. The material is assumed to be homogeneous, perfectly elastic and isotropic.
2. A transverse section of a beam, which is plane before bending, will remain a plane after
bending.
3. The radius of curvature of the beam before bending is very large in comparison to the
transverse dimensions of the beam.
4. The resultant push or pull across a transverse section of the beam is zero.
5. The elastic limit is nowhere exceeded.
6. Young's modulus for the material is same in tension and compression.
7. The transverse section of the beam is symmetrical about an axis passing through the
centroid of the section and parallel to the plane of bending.

Consider the portion of a beam subjected to simple bending as shown in Fig.4.9(a). In the
unstrained state, let GH be a portion of a fibre at a distance y from the centroidal axis KL, its
length being determined by the two transverse parallel planes AD and BC. After bending, the
planes AD and BC assume the positions A1D1 and B1C1 respectively as shown in Fig.4.9(b), being
inclined at an angle  and intersecting at the point O, the centre of curvature. Let R be the radius
of the centroidal surface E1 F1 so that the radius of surface G1 H 1 is (R +y).
G1 H 1 ( R  y ) R  y
 
Now E1 F1 R  R

M D C M 2

y2
E F N A
K L y
G H y1
dA
A B 1

(a) (c) (d)

43
Simple Bending of Beams

D1 R C1 L
K
E1 F1

G1 H1
y
A1 B1

(b)

Fig.4.9 Beam subjected to simple bending

Strain in the fibres at GH is,


G1H1  GH G1H1  EF G1H1 R y
   1  1
GH EF EF R
y

or R
If  = intensity of stress in the fibres, then   E
Ey

or R
 E

or y R (a )
E
For a given load on a beam, is constant (say = k) then,   ky , i.e. the stress in the fibres of a
R

beam at any point in the cross-section is proportional to its distance from the centroidal axis.
Thus the bending stress will be maximum at the boundary of the beam which is at the greatest
distance from the centroidal axis.
In order to locate the position of the neutral axis, consider an element of area dA at a distance y
from the centroidal axis (Fig.4.9 (c)]. Total force on the element is then equal to,
dF   .dA
 1

But y y1
y
  1 
 y1
y
dF   1   dA
Hence y1
Total tensile force on the transverse section below the centroidal axis is then (Fig. 4.9 d).
y 
F1  1   dA  1  ydA
y1 y1

44
Simple Bending of Beams

If the elementary area is chosen on the upper side, then the total compressive force on the
transverse section above the centroidal axis will be,
2
F2 
y2
 ydA
For equilibrium of the beam,
F1  F2
1 2
 
y1 y2

Further, since there is no resultant force across any transverse cross-section, therefore,  ydA  0 ,
i.e. the first moment of the area about the centroidal axis is zero which is possible only if the
neutral axis passes through the centroidal axis. Thus in case of simple bending, the neutral axis
passes through the centroid of the section.
Now the moment of the force acting on the elementary area dA about the neutral axis is,
1 2
dM  y dA
y1
The total moment of all the forces acting on various elements composing the cross-section forms
a couple which is equal to the bending moment M. This total moment is called the moment of
resistance.

M  1  y 2 dA
 y1

Now  y dA  I , the moment of inertia of the cross-section about the neutral axis,
2

1
M  I
 y1
M 1 
 
Hence I y1 y …(b)
Combining Eqs. (a) and (b). we get
M  E
 
I y R …(4.5)
This is the well- known bending formula.
I
M 
Also y

or M  z
I
z
where y ...(4.6)
z is called the section modulus

45
Simple Bending of Beams

Worked Example 4.4 The cross-section of a cast-iron beam is shown in Fig. 4.10 (a) This beam
is simply supported at the ends and carries a uniformly distributed load of 20KN/m. If the span
of the beam is 3m, determine the maximum tensile and compressive stresses in the beam.

Solution:
10cm
c
I 2cm
II 2cm
15cm N A
y
III 3cm
20cm t
(a) (b) Stress distribution

Fig.4.10
Taking moments about the bottom edge of the beam as follows,
(10  2  2 10  20  3) y  10  2 14  2 10  8  20  3 1.5
y  5.3cm
10  23 2  103 20  33
I  10  2(14  5.3) 2   2  10(8  5.3) 2   20  3(5.3  1.5) 2 
12 12 12
 2744.334  108 m4
wl 2 20  32
M   22.5KN  m
8 8
M 

Now I y

Maximum compressive bending stress,


M 22 .5  10 3  9.7  10 2
c  (15  5.3)10  2   79 .527 Mpa
I 2744 .334  10 8
Maximum tensile bending stress,
M 22 .5  10 3  5.3  10 2
t   5.3  10 2   43 .453 Mpa
I 2744 .334  10 8

DISTRIBUTION OF HORIZONTAL SHEAR STRESS IN BEAMS

Fig.4.11

46
Simple Bending of Beams

The vertical shearing force on a beam tends to cause sliding on a vertical section, and the
shearing stress resulting from this is accompanied at any point in the section by a shearing stress
on a horizontal section. The two shears cause tensile and compressive force on mutually
perpendicular planes. The intensity of shear stress on the section of a beam is not constant from
top to bottom of the section, nor it is exactly constant across the width of the section, but for all
practical purposes, we may assume it constant.
The variation of the vertical shearing force may be determined as follows:
Consider a beam of uniform section subjected to bending moment M at the section AC and a
bending moment M  M at the section BD, the two sections being x apart as shown in Fig.
4.11 (a). Let  be the stress at E due to M on a small area of width b and thickness y [Fig. 4.11
(b)] and  1 be the stress at F on a corresponding area of the cross-section. Then
M
 y
I
M  M
1  y
I
Longitudinal pull at E
M
   bdy  bydy
I
Longitudinal pull at F  1bdy
 M  M 
 bydy
 I 
Resultant longitudinal pull on a small elementary slice of length x , width b, and thickness dy
[Fig. 4.11 (c)] becomes
 M  M  M
 bydy  bydy
 I  I
M
 bydy
I
The total force on the elementary length x and area as shown shaded above y is
h / 2 M
 bydy
y I
M
 bydy
I
This resultant pull is resisted by the shearing force on the longitudinal section at EF whose area
is x . b. Let  be the intensity of shear stress on this area, then for equilibrium of AEFB.
M h / 2
I y
b.x  bydy

M 1 h / 2
x Ib y
 bydy

dM 1 h / 2
dx Ib y
 bydy

47
Simple Bending of Beams

dM
F
But dx , the shear force at the section

y bydy  Sum of the moments of the areas of the small strips comprising the shaded
h/2

and
area, about the neutral axis.
If A  Shaded area
y  Distance of its centroid from the neutral axis
h/2

then
y
bydy  Ay

FAy
 
Ib (4.7)

Since a shear is accompanied by a complementary shear of equal intensity of planes


perpendicular to its plane, the intensity of horizontal shear on the cross-section at a distance y
from the neutral axis is given by Eq.(4.7).

(a) Shear stress distribution in a beam of rectangular cross-section.


Consider a beam of rectangular cross-section of width b and depth h as shown in Fig. 4.12 (a).
Let F be the shear force at the cross-section considered.

To determine the shear stress distribution, consider an elementary strip of the beam at a distance
of y and depth dy from the neutral axis.
FAy F  h / 2
   bydy
Now Ib Ib y
h/2
F h / 2 F y2

I 
y
ydy 
I 2 y

2F  h 2
2
  y 
I 4 
F h 2

 3 
 y2 
bh  4 
2
12
6F  h 2 
 3   y2  (4.8)
bh  4 
Fig.4.12 This represents the equation of parabola.
Hence the shear stress distribution is parabolic in nature.
At y  h / 2   0
3 F 3
 max    mean
And at y  0 2 bh 2 (4.9)
Which is the maximum shear stress.

48
Simple Bending of Beams

The distribution of shear stress is shown in Fig. 4.12 (b).


Otherwise:
FAy

Ib
h  1h  1h 
A  b  y  , y  y    y     y 
2  22  22 
1  h2 
 Ay  b  y 2 
2  4 
1  h2 
F b  y 2 
2  4  6F  h2 
      y 2 
bh3 bh3  4 
b
12
which is the same result as obtained above in Eq. (4.8).

(b) Shear stress distribution in a beam of circular cross-section.


Consider a circular beam of radius R. To determine the shear stress distribution, consider an
elementary strip of depth dy at a distance y from the neutral axis as shown in Fig. 4-13 (a).

F R
Ib y
 bydy
Now
Here y  R sin 
b  2 R cos
dy  R cos .d

Fig.4.13
F  /2

I  2 R cos 0  2 R cos  R sin   R cos .d

F  /2

2 IR cos 0  2 R 3 cos2   sin   .d

FR 2  / 2

I cos 0
cos2   sin   .d
 /2
FR 2 cos3 

I cos 3 0

 3 
2
FR
  cos  cos3  
3R 4
 2 
cos
4
4 F cos3  4 F cos2 
  (4.10)
3R 2 cos 3R 2
The variation of shear stress is parabolic in nature

49
Simple Bending of Beams

At   90 ,   0
4F
  0 ,  
and at 3R 2 (4.11)
which is the maximum shear stress. The variation of shear stress distribution is shown in Fig. 4-
13 (b).
F
 mean 
Now R 2
 max 4
 
 mean 3 (4.12)

Worked Example 4.5. A beam of I-section 50 cm deep and 20 cm wide, has equal flanges 2 cm
thick and web 1 cm thick. It carries at a cross-section a shear force of 200 KN. Determine the
shear stress distribution in the beam and the ratio of maximum shear to mean shear.

Solution:
Moment of inertia about neutral axis N-A is,
20  503 19  463
I   208,333.3  154,115.3  54,218cm 4
12 12
FAy

Now Ib
Shear stress in the flange at the junction with the web
200  103  (20  2)  24  10 6

54218 20  10 10
 1.77Mpa
Shear stress in the web at the junction with flange
200  103  (20  2)  24  10 6

54218 1  10 10
 35.4Mpa

2cm 35.4
1.77

N A
45.2
50cm

1cm

1.77
2cm 35.4
20cm

(a) (b)

Fig.4.14

50
Simple Bending of Beams

Shear stress at the neutral axis,


200  103  (20  2)  24  106 200  103  (1  23)  11.5  106
 max  
54128 1  1010 54128 1  1010
 45.2Mpa
The distribution of shear stress is shown in Fig.4.14(b).
Area of the section, A  (20  2)  2  46  1  80  46  126cm
2

F 200  103
 mean    15.8Mpa
A 126  10  4
 45.2
 max   2.85
 mean 15.8

51
Torsion

CHAPTER 5: TORSION

Introduction
In workshops and factories, a turning force is always applied to transmit energy by rotation. This
turning force is applied either to the rim of a pulley, keyed to the shaft, or to any other suitable
point at some distance from the axis of the shaft. The product of this turning force, and the
distance between the point of application of the force, and the axis of the shaft is known as torque,
turning moment or twisting moment. The shaft is said to be subjected to torsion. Due to this
torque, every cross-section of the shaft is subjected to some shear stress.

Assumptions for Finding out Shear Stress in a Circular Shaft, Subjected to Torsion
Following assumptions are made, while finding out shear stress in a circular shaft subjected to
torsion:
(i) The material of the shaft is uniform throughout.
(ii) The twist along the shaft is uniform.
(iii) Normal cross-sections of the shaft, which were plane and circular before twist, remain
plane and circular after twist.
(iv) All diameters of the normal cross-section which were straight before twist, remain straight
with their magnitude unchanged, after twist.
A little consideration will show, that the above assumptions are justified, if the torque applied,
is small and the angle of twist is also small.

Torsional Stresses and Strains


Consider a shaft fixed at one end, and subjected
to a torque at the other as shown in Fig. 5-1.
Let T = Torque in kg-cm,
L = Length of the shaft, and
R = Radius of the shaft.
As a result of this torque, every cross- section of
the shaft will be subjected to shear stresses. Let
the line CA on the surface of the shaft be deformed to CA' and OA to OA' as shown in Fig. 5-1.
Let  ACA' =  in degrees,
 AOA' =  in radians,
 = Shear stress induced at the outermost surface, and
G = Modulus of rigidity, also known as torsional rigidity of the shaft material.
We know that shear strain
= Deformation per unit length
AA'
 tan 
= l
= (  being very small, tan    )

52
Torsion

We also know that the length of the arc AA'  R


AA' R
  
l l …(i)
Moreover deformation
Shear stress
= Modulus of rigidity


G …(ii)
Now from equation (i) and (ii) we find that
 R

G l
 G

or R l
If q be the intensity of shear stress, any layer at a distance r from the centre of the shaft, then

q  G
 
r R l …(iii)

STRENGTH OF A SOLID SHAFT


It means the maximum torque or power the shaft can transmit from one pulley to another. Now
consider a solid circular shaft subjected to some torque.
Let R= Radius of the shaft, and
 =Maximum shear stress developed in the shaft material
Now consider an elementary ring of thickness dx at a x from the centre as shown in Fig. 5-2.
We know that the area of this ring, da  2x  dx
x
x  
and shear stress at this section, R
 Turning force  Stress Area   x  da
x
   da
R
x
    2xdx
R
2 2
 x dx
R
We know that turning moment of this element,
dT = Turning force  Distance of the element from the axis of the shaft
2 2 2 3
dT  x dx  x  x dx
R R
Now the total torque, which the shaft can transmit, may be found out by integrating the above
equation between 0 and R,

53
Torsion

2 3 2 3 2 x 4 R  3 


R R
 T  x dx   x dx  [ ]0  R  D 3
0
R R 0 R 4 2 16
where D is the diameter of the shaft and is equal to 2R.

Worked Example 5.1. A solid shaft is subjected to a torque of 1500kg-metre. Find the necessary
diameter of the shaft, if the allowable shear stress is 600kg/cm2. The allowable twist is 1o for
every 20 diameters length of the shaft. Take G=0.8  106kg/cm2.

Solution.
Given. Torque, T=1500kg-m=150000kg-cm
Allowable shear stress,   600kg / cm
2


  1  rad
Angle of twist, 180
Length of shaft, l=20D
G=0.8  104kg/cm2

Let D= diameter of the shaft.


First of all we shall find out the diameter of the shaft for its strength and stiffness.

(i) For strength


Using the relation

T  D3
16 with usual notations.

150000   600  D 3
16
150000 16
or D3   1273.2
  600
 D  10.84cm ...( i )

(ii) For stiffness


We know that polar moment of inertia of a circular section,

J D
4

32
Using the relation
T G

J l with usual notations.

0.8  10 
6
150000
 180  222.2
 20D D
 D4
32

54
Torsion

150000
D3   2188.7

 222.2
32
 D  13cm …(ii)
From equation (i) and (ii), we find that the necessary diameter of the shaft is 13cm(i.e., greater
of the two values).

STRENGTH OF A HOLLOW SHAFT


It means the maximum torque or power a hollow shaft can transmit from one pulley to another.
Now consider a hollow circular shaft subjected to some torque.
Let R = Outer radius of the shaft,
r = Inner radius of the shaft, and
 = Maximum shear stress developed in the outmost layer of the shaft material.
Now consider an elementary ring of thickness dx at a distance x from the centre as shown in
Fig.5-3. We know that the area of this ring, da  2x  dx
x
 x  
And shear stress at this section, R
 Turning force  Stress Area   x  da
x
   da
R
x
    2xdx
R
2 2
 x dx
R
We know that turning moment of this element,
dT = Turning force  Distance of the element from the axis of the shaft:
2 2 2 3
dT  x dx  x  x dx
R R
The total torque, which the shaft can transmit, may be found out by integrating the above
equation between r and R.
2 3 2 3 2 x 4 R 2 R 4  r 4
R R
 T  x dx   x dx  [ ]r  ( )
r
R R r
R 4 R 4
 D4  d 4
 ( )
16 D
Where D is the external diameter of the shaft and is equal to 2R and d is the internal diameter
of the shaft and is equal to 2r.

Worked Example 5.2. The external and internal diameters of a hollow shaft are 40 cm and 20
cm. Find the maximum torque which the shaft can transmit, if the angle of twist is not to exceed
1 in a length of 10 metres. Take G = 0.8  10 kg / cm .
6 2

55
Torsion

Solution:
Given. External dia. of shaft, D=40cm
Internal dia., d=20cm

  1  rad
Angle of twist 180

Length of shaft l  10m  1000cm


G  0.8  106 kg / cm 2

Let T=Torque transmitted by the shaft.


We know that the polar moment of inertia, of a hollow circular section,
 
J  ( D 4  d 4 )  (404  204 )  75000cm 4
32 32
Using the relation,
T G

J l with usual notations.

0.8  106 
T
 180 40

75000 1000 9

40
 T  75000  3290 103 kg  cm
9

56
Slope and Deflection: Elastic Beam Theory

CHAPTER 6: SLOPES AND DEFLECTIONS

SECTION I. ELASTIC- BEAM THEORY


Introduction
In certain situation it becomes necessary to design a machine component or a structure to minimize
deflection. Thus it becomes essential to determine the deflection of the member. Also for statically
indeterminate or fixed beams it becomes necessary to determinate the slope and deflection at salient
points.

Curvature of the Bending Beam


Consider a beam AB subjected to a bending moment. Let the beam deflect from ACB to ADB into a
circular arc as shorn in Fig. 6.1. E

Let l = Length of the beam AB,


M = Bending moment,
R = Radius of curvature, of the bent up beam,
O
I = Moment of inertia of the beam section,
E = Modulus of elasticity of beam material, i R
L/2 L/2
y = Deflection of the beam (i.e., CD) and
C
B
i = Slope of the beam (angle which the tangent at A makes with AB). A i y

From the geometry of a circle, we know that


AC  CB = EC  CD Fig. 6.1
l l
or   (2 R  y )  y
2 2
l2
  2 Ry  y 2  2 Ry (neglecting y 2 )
4
l2
or y …(i)
8R
We have already discussed that for a loaded beam,
M E EI
 or R 
I R M
Now substituting this value of R in equation (i), we get the value of deflection,
l2 Ml 2
y 
EI 8 EI
8
M
Now from the geometry of the figure, we find that the slope of the beam i at A or B is also equal to
angle AOC.
AC l
 sin i  
OA 2 R
Since the angle i is very small, therefore, sin i may be taken equal to i (in radians)
l
 i radians …(ii)
2R
Again substituting the value of R in equation (ii) we get the value of slope,

57
Slope and Deflection: Elastic Beam Theory

l l Ml
 i   …(iii)
2R EI 2 EI
2
M

Relation between Slope, Deflection and Radius of Curvature

Y C Consider a small portion PQ of a beam, bent into a arc, as shown in Fig. 6-2.
R
Q
d Let ds = Length of the beam PQ,
ds dy R = Radius of the arc, into which the beam has been bent,
P C = Centre of the arc,
dx  = Angle, which the tangent at P makes with x-x axis, and
  d = Angle which the tangent at Q makes with x-x axis.
+d
From the geometry of the figure, we find that
0 X PCG  d and ds  R  d
ds dx
Fig. 6-2  R  (Considering ds=dx)
d d
1 d
 or …(i)
R dx
We know that if x and y be the co-ordinates of point P, then
dy
tan =
dx
Since  is a very small angle, therefore, taking tan = ,
dy
 =
dx
Differentiating the above equation with respect to x,
d d 2 y 1 d
  (  )
dx dx2 R dx
We also know
M E 1
 or M  EI 
I R R
d2y 1
 M  EI (substituting value of )
dx2 R

Double Integration Method for Slope and Deflection


d2y
We have already discussed that the bending moment at a point, M  EI
dx2
dy
dx 
Integrating the above equation,  M EI …(i)

Integrating the above equation once again,


EI. y   M …(ii)
It is thus obvious, that after first integration the original, differential equation, we get the value of slope
at any point. On further integrating, we get the value of deflection at any point.

58
Slope and Deflection: Elastic Beam Theory

Simply Supported Beam with a Central Point Load


Consider a simply supported beam AB of length l and carrying a point load W at the centre of beam C as
shown in Fig. 6-3. From the geometry of the figure , we find that the reaction at A,
W
RA  RB 
2
x
C Consider a section X at a distance x from B. We know
A B that the
yc x
w/2 w/2 bending moment at this section,
L/2 L/2
W
M X  RB . x  x (Plus sign due sagging)
2
Fig.6.3
d 2 y Wx
 EI 2  …(i)
dx 2
dy Wx 2
Integrating the above equation, EI   C1 ...(ii)
dx 4
l dy
where C1 is first the constant of integration. We know when x  ,  0 . Substituting these values
2 dx
Wl 2 Wl 2
in equation (ii), 0  C1 or C1  
16 16
Substituting this value of C1 in equation (ii),
dy Wx 2 Wl 2
 EI …(iii)
dx 4 16
This is the required equation for the slope, at any section, by which we can get the slope at any pint
on the beam. A little consideration will show, that the maximum slope occurs at A and B. Thus, for
maximum slope at B, substituting x  0 in equation (iii),
Wl 2
EI .iB  
16
Wl 2
 iB   (Minus sign means that the tangent at B makes an angle with AB in the negative or
16EI
anticlockwise direction)
Wl 2
or iB  radians
16EI
Wl 2
By symmetry i A  radians
16EI
Integrating the equation (iii) once again,
Wx 3 Wl 2
EI . y   x  C2 ...( iv)
12 16
where C2 is the second constant of integration. We know that when x=0, y=0. Substituting these values
m equation (iv), we get C2=0.
Wx 3 Wl 2
 EI . y   x …(v)
12 16

59
Slope and Deflection: Elastic Beam Theory

This is the required equation for the deflection, at any section, by which we can get the deflection at any
point on the beam. A little consideration will show, that maximum deflection occurs at the mid-point C.
Thus, for maximum deflection, substituting x = l/2 in equation (v),
W l 3 Wl 2 l Wl 3 Wl 3 Wl 3
EI . y  ( )  ( )  
12 2 16 2 96 32 48
Wl 3
or yC   (Minus sign means that the deflection is downwards)
48EI
Wl 3

48EI

Worked Example 6-1. A beam 3 metres long, simply supported at its ends, is carrying a point load (W)
at its centre. If the slope at the ends of the beam is not to exceed 1 , find the deflection at the centre of
the beam.
Solution.
Given. Length, l = 3 m
Central point load = W
1 
Slope at A, iA  1   0.01745 radians
180
We know that slope at an end,
Wl 2
iA 
16EI
and deflection at the centre,
Wl 3 Wl 2 l
yC   
48EI 16EI 3
l Wl 2
 iA  ( iA  )
3 16EI
3
 0.01745  0.01745m  1.745cm Ans.
3

Simply Supported Beam with a Uniformly Distributed Load


w/Unit length x
Consider a simply supported beam AB of length l
A B and carrying uniformly distributed load of w per unit
C x
wL/2 wL/2 length as shown in Fig.6-4. From the geometry of
L the figure, we find that the reaction at A,

Fig.6.4
Wl
RA  RB 
2
Consider a section X at a distance x from B. We know that the bending moment at this section,
Wlx Wx 2
MX   (Plus sign due sagging)
2 2

60
Slope and Deflection: Elastic Beam Theory

d 2 y Wlx Wx 2
 EI   …(i)
dx2 2 2

Integrating the above equation,


dy Wlx2 Wx3
EI    C1 ...(ii)
dx 4 6
l dy
where C1 is first the constant of integration. We know when x  ,  0 . Substituting these values
2 dx
Wl l 2 W l 3 Wl 3
in equation (ii), 0
( )  ( )  C1 or C1  
4 2 6 2 24
Substituting this value of C1 in equation (ii),
dy Wlx 2 Wx3 Wl 3
  EI   …(iii)
dx 4 6 24
This is the required equation for the slope at any section, by which we can get the slope at any point
on the beam. we know that maximum slope occurs A and B. Thus for maximum slope, substituting x =
0 in equation (iii),
Wl 3
EI .iB  
24
Wl 3
 iB   (Minus sign means that the tangent at A makes an angle with AB in the negative or
24EI
anticlockwise direction)
Wl 3
or iB  radians
24EI
Wl 3
By symmetry iA  radians
24EI
Integrating the equation (iii) once again,
Wlx 3 Wx 4 Wl 3 x
EI . y     C2 ...( iv)
12 24 24
where C2 is the second constant of integration. We know that when x=0, y=0. Substituting these
values m equation (iv), we get C2=0.
Wlx 3 Wx 4 Wl 3 x
EI . y    …(v)
12 24 24
This is the required equation for the deflection at any section, by which we can get the deflection at
any point on the beam. We know that maximum deflection occurs at the mid-point C. Thus, for
maximum deflection, substituting x = l/2 equation (v),
Wl l 3 W l 4 Wl 3 l Wl 4 Wl 4 Wl 4 5Wl 4
EI . y C  ( )  ( )  ( )   
12 2 24 2 24 2 96 384 48 384
5Wl 4
or yC   (Minus sign means that the deflection is downwards)
384EI
Wl 4

384EI

61
Slope and Deflection: Elastic Beam Theory

Worked Example 6-2. A timber beam of rectangular section has a span of 4.8 metres and is simply
supported at its ends. It is required to carry a total load of 4500 kg uniformly distributed over the whole
span. Find the maximum values for the breadth (b) and depth (d) of the beam, if maximum bending
stress is not to exceed 70 kg/cm2 and the maximum deflection is limited to 9.5 mm. Take E for timber
as 105  103kg/cm2

Solution.
Given. Span, l = 4.8 m = 480 cm
Total uniformly distributed load, W= wl=4500kg
Bending stress,  = 70 kg/cm2
Maximum deflection, yC  9.5mm  0.95cm
Young's modulus, E = 105  103 kg/cm2
Let b = Breadth of the beam, and .
d = Depth of the beam.
We know that the maximum bending moment in a simply supported beam, carrying a uniformly
distributed load, is at the centre, i.e.
wl 2 Wl 4500 4.8
M    2700kg  m  270000kg  cm
8 8 8
Using the relation,
M 
 with usual notations.
I y
270000 70 bd 3 d
3
 (  I  and y  )
bd d 12 2
12 2
270000  12
 bd 2   23143cm 3 …(i)
70  2
Now using the relation,
5Wl 2
yC  with usual notations.
384EI
5  4500 4803 5184000
0.95  3

bd 7bd 3
384  105 103 
12
5184000
bd 3   779500cm 4 …(ii)
7  0.95
Dividing equation (ii) by (i),
779500
d  33.68cm Ans.
23143
Substituting this value of d in equation (i),
b  33.682  23143
23143
 b  20.4cm Ans.
33.682

62
Slope and Deflection: Elastic Beam Theory

Cantilever with a Point Load at the Free End


x w Consider a cantilever AB of length l and carrying a point
load w at the free end as shown in Fig. 6-5. Consider a
x B
A iB section X, at a distance x from the free end B. We know
yB that bending moment at this section,

L B' M x  W .x (Minus sign due to hogging)

Fig.6.5
d2y
 EI W .x …(i)
dx2
Integrating the above equation,
dy Wx 2
EI   C1 ...(ii)
dx 2
dy
where C1 is first the constant of integration. We know when x  l , 0.
dx
Wl 2 Wl 2
Substituting these values in the equation (ii), 0  C1 or C1 
2 2
Now Substituting this value of C1 in equation (ii),
dy Wx 2 Wl 2
 EI   …(iii)
dx 2 2
This is the required equation for the slope at any section, by which we can get the slope at any point on
the cantilever. A little consideration will show, that maximum slope occurs at the free end. Now using
the abbreviation i for the angle of inclination (in radians) and considering i=tan i, for very small angles.
Now for maximum slope, substituting x = 0 in equation (iii),
Wl 2
EI .iB 
2
Wl 2
 iB  radians ...( iv)
2 EI
Integrating the equation (iii) once again,
Wx3 Wl 2 x
EI . y     C2 ...( v)
6 2
where C2 is the constant of integration. We know that when x=l, y=0. Substituting these values in the
above equation,
Wl 3 Wl 3
0   C2
6 2
Wl 3
or C2  
3
Substituting this value of C2 in equation(v),
Wx3 Wl 2 x Wl 3
EI . y     …(vi)
6 2 3

63
Slope and Deflection: Elastic Beam Theory

This is the required, equation for the deflection, at any section, by which we can get the deflection at any
point on the cantilever. A little consideration will show, that maximum slope occurs at the free end. Now
for maximum deflection, substituting; x = 0 in equation (vi),
Wl 3
EI . yB  
3
Wl 3
or yB   (Minus sign means that the deflection is downwards)
3EI
Wl 3

3EI

Cantilever with a Point Load not at the Free End


Consider a cantilever AB of length l, and carrying a point load W at C at a distance l1, from the fixed end
as shown in Fig. 6-6.
A little consideration will show, that the portion AC of the cantilever will bend into AC', while the
w
portion CB will remain straight, and displaced to C'B' as shown in Fig. 6.6
A C B
The portion AC of the cantilever may be taken as similar to a ic y
C yB
cantilever in above Art. (i.e., load at the free end). C'
B'
Wl 2
 iC  1 L1 L2
2 EI L
Since the portion CB of the cantilever is straight, therefore Fig.6.6
2
Wl
iB  iC  1
2 EI
Wl13
And yC 
3EI
From the geometry of the figure, we find that
Wl13 Wl12
y B  y C  iC (l  l1 )   (l  l1 )
3EI 2 EI
l
If l1 
2
W l 3 Wl l 2 l 5Wl 3
yB  ( )  ( )  
3EI 2 2 EI 2 2 48EI

Worked Example 6-3. A cantilever 15 cm wide and 20 cm deep projects 1.5 m out of a wall, and is
carrying a point load of 5000 kg at the free end. Find the slope and deflection of the cantilever at the
free end. Take E  2.1 106 kg / cm 2 .
Solution.
Given. Width b = 15 cm
Depth, d = 20 cm
bd 3 15  203
 I   10000cm 4
12 12
Length l= 1.5m = 150cm
Load, W=5000kg

64
Slope and Deflection: Elastic Beam Theory

Young's modulus, E = 2.1  l06 kg/cm2

Slope at the free end


Let iB= Slope at the free end.
Using the relation
Wl 2
iB  with usual notations.
2 EI
5000 1502
 rad
2  2.1 106  10000
 0.0027 rad Ans.

Deflection at the free end


Let yB= Deflection at the free end.
Using the relation,
Wl 3
yB  with usual notations.
3EI
5000 1503
  0.27cm Ans.
3  2.1 106  10000

Cantilever with a uniformly distributed Load


x Consider a cantilever AB of length l, and carrying a uniformly
w/unit length distributed load of w per unit length as shown in Fig. 6-7. Consider a
A B section X, at a distance x from the free end B.
iB
yB We know that bending moment at the section,
L wx 2
MX   (Minus sign due to hogging)
2
Fig.6.7
d2y wx 2
 EI   …(i)
dx2 2
dy wx 3
Integrating the above equation, EI   C1 ...(ii)
dx 6
dy
where C1 is the constant of integration. We know when x  l ,  0 . Substituting these values in
dx
wl 3 wl 3
the equation (ii), 0
 C1 or C1 
6 6
Now Substituting this value of C1 in equation (ii),
dy wx 3 wl 3
 
EI  …(iii)
dx 6 6
This is the required equation for the slope at any section, by which we can get the slope at any point
on the cantilever. We know that maximum slope occurs at the free end B. Now for maximum slope,
substituting x = 0 in equation (iii),
wl 3
EI .iB 
6

65
Slope and Deflection: Elastic Beam Theory

wl 3
 iB  radians ...( iv)
6 EI
Integrating the equation (iii) once again,
wx 4 wl 3 x
EI . y     C2 ...( v)
24 6
where C2 is the constant of integration. We know that when x=l, y=0. Substituting these values in the
above equation,
wl 4 wl 4 wl 4
0   C2 or C2  
24 6 8
Substituting this value of C2 in equation(v),
wx 4 wl 3 x wl 4
EI . y     …(vi)
24 6 8
This is the required, equation for the deflection, at any section, by which we can get the deflection at any
point on the cantilever. We know that maximum slope occurs at the free end. Therefore for maximum
deflection, substituting x = 0 in equation (vi),
wl 4 wl 4
EI . y B   or yB   (Minus sign means that the deflection is downwards)
8 8 EI
wl 4

8 EI

Worked Example 6.4. A cantilever 120 mm wide and 200 mm deep is 2.5 metres long. What uniformly
distributed load should the beam carry to produce a deflection of 5 mm at the free end? Take E=200
GN/m2.
Solution.
Given. Width, b = 120 mm
Depth, d = 200 mm
bd 3 120  2003
 I   80  106 mm 4
12 12
Length, l = 2.5 m = 2.5  103 mm
Deflection at the free end,
yB  5mm .
E = 200 GN/m2 = 200  103 N/mm2
Let W = Total uniformly distributed load.
Using the relation,
Wl 3
yB  with usual notations.
8 EI
W (2.5  103 )3 W
5 
8  200  10  80  10
3 6
8192

W  5  8192  40960N  40.96KN Ans.

66
Slope and Deflection: Moment Area Method

SECTION II: Moment Area Method


2.1 Introduction
When a structure is subjected to the action of applied loads each member undergoes deformation
due to which the axis of structure is deflected from its original position. The deflections also
occur due to temperature variations and lack-of-fit of members. The deflections of structures are
important for ensuring that the designed structure is not excessively flexible. The large
deformations in the structures can cause damage or cracking of non-structural elements. The
deflection in beams is dependent on the acting bending moments and its flexural stiffness. The
computation of deflections in structures is also required for solving the statically indeterminate
structures. In this chapter, several methods for computing deflection of structures are considered.

2.2 Moment Area Method


The moment-area method is one of the most effective methods for obtaining the bending
displacement in beams and frames. In this method, the area of the bending moment diagrams is
utilized for computing the slope and or deflections at particular points along the axis of the beam
or frame. Two theorems known as the moment area theorems are utilized for calculation of the
deflection. One theorem is used to calculate the change in the slope between two points on the
elastic curve. The other theorem is used to compute the vertical distance (called tangential
deviation) between a point on the elastic curve and a line tangent to the elastic curve at a second
point.

Consider Figure 6.8 showing the elastic curve of a loaded simple beam. On the elastic curve
tangents are drawn on points A and B . Total angle between the two tangents is denoted as  AB .
In order to find out  AB , consider the incremental change in angle d over an infinitesimal
segment dx located at a distance of x from point B . The radius of curvature and bending
moment for any section of the beam is given by the usual bending equation.
M E

I R (2.1)
where R is the radius of curvature; E is the modulus of elasticity; I is the moment of inertia; and
M denotes the bending moment.
The elementary length dx and the change in angle d are related as,
dx  d  R (2.2)

67
Slope and Deflection: Moment Area Method

A B
A
B
dx d dt
Elastic curve x
t BA AB

(a) Beam B'

xB
M B /EI

M A /EI
c.g.
dx
(b) M/EI diagram

Fig.6.8
Substituting R from Eq. (2.2) in Eq. (2.1)
M
d  dx
EI (2.3)
The total angle change  AB can be obtained by integrating Eq. (2.3) between points A and B
which is expressed as
B B
M
 AB   d   dx
A A
EI (2.4a)
 B   A  Area of M / EI diagram between A and B (2.4b)
The difference of slope between any two points on a continuous elastic curve of a beam is
equal to the area under the M / EI curve between these points.

The distance dt along the vertical line through point B is nearly equal to.
dt  x  d (2.5)
Integration of dt between points A and B yield the vertical distance t BA between the point B and
the tangent from point A on the elastic curve. Thus,
B B
Mx
t BA   xd   dx
A A
EI (2.6)
Since the quantity Mdx/ EI represents an infinitesimal area under the M /EI diagram and
distance x from that area to point B, the integral on right hand side of Eq. (2.6) can be interpreted
as moment of the area under the M/EI diagram between points A and B about point B. This is the
second moment area theorem.

68
Slope and Deflection: Moment Area Method

If A and B are two points on the deflected shape of a beam, the vertical distance of point B
from the tangent drawn to the elastic curve at point A is equal to the moment of bending
moment diagram area between the points A and B about the vertical line from point B ,
divided by EI .

Sign convention used here can be remembered keeping the simply supported beam of Figure 6.8
in mind. A sagging moment is the positive bending moment diagram and has positive area.
Slopes are positive if measured in the anti-clockwise direction. Positive deviation t BA indicates
that the point B lies above the tangent from the point A .

Worked Example 6.5 Determine the end slope and deflection of the mid-point C in the beam
shown below using moment area method.
W

A B
C
L/2 L/2

(a)
WL
4EI

(b)

A B
A C

t CA
t BA

(c)

Figure 6.9
Solution:
The M / EI diagram of the beam is shown in Figure 6.9(a). The slope at A ,  A can be obtained
by computing the t BA using the second moment area theorem i.e.
 A  L  t BA
1  1 WL L  WL2
A     L  
L  2 4 EI 2  16EI (clockwise direction)
The slope at B can be obtained by using the first moment area theorem between points A and B
 B   A   AB
1 WL WL2
B  A   L 
2 4 EI 8EI
WL2 WL2 WL2
B   
8EI 16EI 16EI (anti-clockwise direction)

69
Slope and Deflection: Moment Area Method

WL2
A  
(It is to be noted that the 16EI . The negative sign is because of the slope being in the
clockwise direction. As per sign convention a positive slope is in the anti-clockwise direction)

The deflection at the centre of the beam can be obtained with the help of the second moment
area theorem between points A and C i.e.

L
A    C  t CA
2
WL2 L  1 WL L L 
  C      
16EI 2  2 4 EI 2 6 
WL3
C 
48EI (downward direction)

Worked Example 6.6 Using the moment area method, determine the slope at B and C and
deflection at C of the cantilever beam as shown in Figure 6.10(a). The beam is subjected to
uniformly distributed load over entire length and point load at the free end.
w/unit length W

A C
B
L/2 L/2

(a)
2L/3
M
EI
WL (b)
---
EI

L/2
M
EI
2
2 WL
WL ---
--- 8EI
2EI 3L/4

(c)

tangrnt at A
A=0
B t CA = C
(d)
C

Figure 6.10

70
Slope and Deflection: Moment Area Method

Solution:
The moment curves produced by the concentrated load, W and the uniformly distributed load,w
are plotted separately and divided by EI (refer Figures 6.10(b) and (c)). This results in the simple
geometric shapes in which the area and locations of their centroids are known.
Since the end A is fixed, therefore,  A  0 . Applying the first moment-area theorem between
points A and C
 C   A   AC
1 WL 1 wL2 
 C   A    L 
   L 
2 EI 3 2 EI  (negative sign is due to hogging moment)
 WL 2
wL 
3
 C    
 2 EI 6 EI  (clockwise direction)

The slope at B can be obtained by applying the first moment area theorem between points B and
C i.e.
 C   B   BC
 B   C   BC
 WL2 wL3   1 L WL 1 L wL2 
 B            
 2 EI 6 EI   2 2 2 EI 3 2 82EI 
 3WL2 7 wL3 
 B    
 8 EI 48EI 
(clockwise direction)

The deflection at C is equal to the tangential deviation of point C from the tangent to the elastic
curve at A (see Figure 6.10(d)).
 C  t CA
= moment of areas under M / EI curves between A and C in Figures 6.10(b) and (c)
about C
1 WL 2 L 1 wL2 3L WL3 wL4
  L    L   
2 EI 3 3 2 EI 4 3EI 8EI (downward direction)

71
Slope and Deflection: Moment Area Method

Worked Example 6.7: Determine the end-slopes and deflection at the center of a non-prismatic
simply supported beam. The beam is subjected to a concentrated load at the center.

Solution:
The M/EI diagram of the beam is shown in Figure 6.11(b).
P

A B
I C 2I
L/2 L/2

(a)
PL
4EI
PL
8EI

(b)

A B
A C

t CA
t BA

(c)

Figure 6.11

Applying second moment-area theorem between points A and B,


1 L PL  L L  1 PL L  2 L 
t BA           
2 2 4 EI  2 6  2 8 EI 2  3 2 
PL3 PL3 5 PL3
 AL   
24EI 96EI 96EI
5 PL2
A  
96EI (clockwise direction)

Applying first moment area theorem between A and C .


1 L PL
C   A   
2 2 4 EI
PL2 5PL2 PL2
C   
16EI 96EI 96EI (anti-clockwise direction)

Applying second moment area theorem between A and C .

72
Slope and Deflection: Moment Area Method

PL2 L PL3
t CA   
16EI 6 96EI
5 PL2 L PL3 PL3
C    
96EI 2 96EI 64EI (downward direction)

Geometric Properties of Area

1
h h1 A b(h1  h2 )
C 1 C h2 2
A bh
2 b(2h2  h1 )
x x x
1 b 3(h1  h2 )
b x b
3 Trapezoid
Triangle

h
h C
2
C A bh
3 1
x A  bh
3 3
x b b
x 8 1
Parabolic spandrel x b
b 4

Semi Parabola

73
Slope and Deflection: Conjugate Beam Method

SECTION III: Conjugate Beam Method


The conjugate beam method is an extremely versatile method for computation of deflections in
beams. The relationships between the loading, shear, and bending moments are given by

d 2 M dV
   w( x)
dx 2 dx (3.1)
where M is the bending moment; V is the shear; and w ( x ) is the intensity of distributed load.
Similarly, we have the following
d 2 d M
 
dx 2 dx EI (3.2)

A comparison of two set of equations indicates that if M / EI is the loading on an imaginary


beam, the resulting shear and moment in the beam are the slope and displacement of the real
beam, respectively. The imaginary beam is called as the “conjugate beam” and has the same
length as the original beam.

There are two major steps in the conjugate beam method. The first step is to set up an additional
beam, called "conjugate beam,” and the second step is to determine the “ shearing forces ” and
“ bending moments ” in the conjugate beam.
The loading diagram showing the elastic loads acting on the conjugate beam is simply the
bending-moment diagram of the actual beam divided by the flexural rigidity EI of the actual
beam. This elastic load is downward if the bending moment is sagging.

For each existing support condition of the actual beam, there is a corresponding support
condition for the conjugate beam. Table 3.1 shows the corresponding conjugate beam of
different types of actual beams. The actual beam as well as the conjugate beam are always in
static equilibrium condition.

The slope of (the centerline of) the actual beam at any cross-section is equal to the “shearing
force” at the corresponding cross-section of the conjugate beam. This slope is positive or anti-
clockwise if the “shearing force” is positive — to rotate the beam element anti-clockwise — in
beam convention. The deflection of (the centerline of) the actual beam at any point is equal to
the “bending moment” of the conjugate beam at the corresponding point. This deflection is
downward if the “bending moment” is positive — to cause top fiber in compression — in beam
convention. The positive shearing force and bending moment are shown below in Figure 3.1.

+ ve + ve

Figure 3.1 Positive shearing force and bending moment

Table 3.1 Real and Conjugate beams for different structures

74
Slope and Deflection: Conjugate Beam Method

75
Slope and Deflection: Conjugate Beam Method

Worked Example 3.1 Determine the slope and deflection of point A of a cantilever beam AB of
length L and uniform flexural rigidity EI. A concentrated force P is applied at the free end of
beam.
P

A B

(a) A cantilever beam (actual beam)

PL
EI
A B

(b) Conjugate beam (additional beam) corresponding to the actual beam

PL
MA EI
A B
L
VA
(c) Free-body diagram for the conjugate beam
P
A
B

(d) Deflections of the cantilever beam (actual beam)


Figure 3.2
Solution:
The conjugate beam of the actual beam is shown in Figure 3.2(b). A linearly varying distributed
upward elastic load with intensity equal to zero at A and equal to PL/EI at B. The free-body
diagram for the conjugate beam is shown in Figure 3.2(c). The reactions at A of the conjugate
beam are given by
PL PL2
1
VA   L   anti  clock wise
2 EI 2 EI
1
MA    L
PL  2 L PL3
  
2 EI  3 3EI
The slope at A ,  A and the deflection  A at the free end A of the actual beam in Figure 3.2(d)

76
Slope and Deflection: Conjugate Beam Method

are respectively, equal to the “shearing force” V A and the “bending moment” M A at the fixed
end A of the conjugate beam in Figure 3.2(d).

A 
PL2
2 EI
anti  clock wise direction A 
PL3
3EI

and
Note that  A points downward because M A causes tension in bottom fiber of the beam at A (i.e.
sagging moment) .

Worked Example 3.2 Determine the slope at A and deflection of B of the beam shown in Figure
3.3(a) using the conjugate beam method.
w/unit length
V=0
A M=0
B
L

(a) Beam
2
wL
2EI

A
B
x dx
(b) Conjugate Beam

Figure 3.3

Solution:
The vertical reaction at A in the real beam is given by V A  wL
1 2
M x  wLx  wx
The bending moment at any point X at a distance x from A is given by 2

The corresponding conjugate beam and loading acting on it are shown in Figure 3.3(b). The
wL2
loading on the beam varies parabolically with maximum value as 2 EI .
The slope at A ,  A in the original beam will be equal to the shear force at A in the conjugate
beam, thus,
2 wL2 wL3
A    L 
3 2 EI 3EI (clockwise direction)
The deflection of B in the real beam will be equal to the bending moment at B in conjugate beam
i.e.
2 wL2 2 wL2 3L 5wL4
B   L  L   L  
3 2 EI 3 2 EI 8 24EI (downward direction)

77
Slope and Deflection: Principle of Virtual Work

SECTION IV: Principle of Virtual Work


Background
Consider a structural system subjected to a set of forces ( P1 , P2 , P3 ... referred as P force) under
stable equilibrium condition as shown in Figure 4.1(a). Further, consider a small element within
the structural system and stresses on the surfaces caused by the P forces are shown in Figure

4.1(b) and referred as p .
P1
P2

P3

(a) (b)

p
Fig.4.1 Deformable body subjected to external force. (b)-Stress on small element,

Let the body undergoes to a set of compatible virtual displacement D . These displacements
are imaginary and fictitious as shown by dotted line. While the body is displaced, the real forces
acting on the body move through these displacements. These forces and virtual displacements
must satisfy the principle of conservation of energy i.e.
We  Wi (4.1)
n

 P (D)   
i 1
i i p ( )dV
V (4.2)
This is the principle of virtual work.
If a system in equilibrium under a system of forces undergoes a deformation, the work
done by the external forces ( P ) equals the work done by the internal stresses due to those

forces, ( p ).

In order to use the above principle for practical applications, we have to interchange the role of
the forces and displacement. Let the structure acted upon by a virtual force is subjected to real
displacements then the Eq. (4.10) can be written as
n

 D (P)    (
i 1
i i p )dV
V (4.3)
This is the principle of complimentary virtual work and used for computing displacements.
Consider a structure shown in Figure 4.2(a) and subjected to P force and it is required to find
the displacement of point C in the direction specified. First apply a virtual force F at C in the
required direction. Next apply the external (real) loads acting on the structures as shown in
Figure 4.2(a) with the virtual force remain in the position. The displacement of C in the required

78
Slope and Deflection: Principle of Virtual Work

direction be  and the internal elements deform by an amount L . Using Eq. (4.3)
F     f  L
(4.4)
The left hand side of Eq. (4.4) denotes the external work done by the virtual force F moving
through the real displacement  . On the other hand, the right hand side of Eq. (4.4) represents
the internal work done by the virtual internal element forces f moving through the
displacement L . Since F is arbitrary and for convenience let F  1 (i.e. unit load). The Eq.
(4.4) can be re-written as
1     f  L
(4.5)
where f denotes the internal force in the members due to virtual unit load.
The right hand side of Eq. (4.5) will directly provide the displacement of point C due to applied
external forces. This method is also known as unit load method.
Similarly for finding out a rotation,  at any point of a loaded structure, the corresponding Eq.
(4.5) will take place as
1     f  L
(4.6)
where f denotes the internal force in the members due to virtual unit moment applied in the
direction of interested  .
P1
P2

f L

P3
f C C
C'
F 1 1
(a) (b) (c)

Fig.4.2

APPLICATION TO PIN-JOINTED STRUCTURES


C

A B
P1 P2 P3

1
C

A B

Fig.4.3

79
Slope and Deflection: Principle of Virtual Work

Consider a pin-jointed structure as shown in Figure 4.13 and subjected to external force P1 , P2
and P3 . Let the vertical displacement of point C,  CV is required. Under the action of the real
external load, let the axial force in typical member be FP and therefore, the deformation of the
member L  FP L / AE ( L and AE are the length and axial rigidity of typical member).
Apply a unit vertical load at C and substituting in Eq. (4.4) leads to
FP L
1   CV   Fu  L  CV   Fu
thus AE
(4.7)
The basic steps to be followed for finding the displacements of the pin-jointed structure are
(i) Compute the axial force in various members (i.e. FP ) due to applied external forces.
(ii) Compute the axial force in various members (i.e. Fu ) due to unit load applied in the
direction of required displacement of the point.
FP L
F u
AE for all members.
(iii) Compute the product
FP L
F u
AE will provide the desired displacement.
(iv) The summation
(v) The axial force shall be taken as positive if tensile and negative if compressive.
FP L
F u
AE implies that the desired displacement is in the direction of applied
(vi) The positive
unit load and negative quantity will indicate that the desired displacement is in the opposite
direction of the applied unit load.

Worked Example 4.1 For the pin-jointed structure shown in the Figure 4.4, find the horizontal
and vertical displacement of the joint D. The area of cross-section, A  500mm and
2

E  200,000N / mm 2 for all the members.


D 180KN

60°
C
2m

B
2m

A 30° 30° F

Fig.4.4

Solution:
The axial rigidity of the members, AE  500  2  10  10  10 KN . The computation of the
5 3 5

desired displacements is presented in Table 4.1.

80
Slope and Deflection: Principle of Virtual Work

Table 4.1
Length For  DH For  DV
Member FP
L(m) Fu FP Fu L Fu Fu FP L
AB 2 720 / 3 4/ 3 1920 0 0
BC 2 360 / 3 2/ 3 480 0 0
CD 2 360 / 3 2/ 3 480 0 0
DE 2  180 / 3 1/ 3 120 1  360 / 3
EF 2  180 / 3  3 1080 1  360 / 3
CE 2  360 / 3 2/ 3 480 0 0
BF 2 360 / 3 2/ 3 480 0 0
 5520
 1440 / 3
The horizontal deflection of
5520
D  103  55.2mm()
105 .
The vertical deflection of
1440 3
D 5
 103  8.31mm()  8.31mm()
10

APPLICATION TO BEAMS AND FRAMES


In order to find out the vertical displacement of C of the beam shown in Figure 4.5(a), apply a
unit load as shown in Figure 4.5(b).

A B
CV

(a)

1
C B
A
(b)

Fig.4.5

The internal virtual work is considered mainly due to bending and caused due to internal
moments M u under going the rotation d due to the applied loading. (internal virtual work done
by shearing forces and axial forces is small in comparison to the bending moments and hence
d  M P dx
ignored). Since the EI where M P is the moment due to applied loading, the Eq. (4.7)

81
Slope and Deflection: Principle of Virtual Work

for the displacement of C will take a shape of

L L
M P dx
1   CV   M u  d  CV   M u
EI
0 (4.8)
0

The basic steps to be followed for finding the displacement or slope of a beams and frames are
summarized as
(i) Compute the bending moment (i.e. M P ) due to applied external forces.
(ii) Compute the bending moment (i.e. M u ) due to unit load applied in the direction of required
displacement or slope.
L
M P dx
M u
EI
(iii) Compute the integral 0 over the entire members of the beam or frame which will
provide the desired displacement.
(iv) The bending moment shall be taken as positive if sagging and negative if hogging (in case
of beams).
L
M P dx
M u
EI
(v) The positive 0 implies that the desired displacement is in the direction of applied
unit load and negative quantity will indicate that the desired displacement is in the opposite
direction of the applied unit load.

Worked Examples for Beams


Question One: Determine the slope and deflection of point A of the cantilever beam AB with
length L and constant flexural rigidity EI.
P 1
1 X
X A
A B A B B
x x
L dx dx
(b) (c)
(a)

Solution
Deflection under the Load:
Apply a vertical unit load at point A of the beam as shown in Figure 4.6(b). Consider any point
X at a distance of x from A.
M P   P  x   Px and M u  1  x   x
The vertical deflection of point A is given by
( Px )dx
L L
M dx
 AV   M u P   ( x)
0
EI 0
EI
PL3
 AV 
3EI
Slope at the free end:
Apply a unit couple at point A of the beam as shown in Figure 4.6(c). Consider any point X at a
distance of x from A.

82
Slope and Deflection: Principle of Virtual Work

M P  P  x M u  1
and
The slope at A is given by
( Px )dx
L L
M P dx
A   Mu   (1)
0
EI 0
EI
PL2
A 
2 EI

Question Two: Determine mid-span deflection and end slopes of a simply supported beam of
span L carrying a uniformly distributed load, w per unit length.
w/unit length 1

A B A B A B
C C 1
(a) (b) (c)
Solution:
 Mid-span deflection: Apply a unit load at mid span as shown in Figure (b) above.
Consider any point X at a distance of x from A
wL wx 2 wx ( L  x)
MP  x  (0  x  L )
2 2 2
1 x
Mu   x  ( 0  x  L / 2)
2 2
1  L L x
M u   x  1  x    ( L / 2  x  L)
2  2 2
The vertical deflection of point C is given by
 wx ( L  x)   wx ( L  x) 
 dx L  dx
 x   L x
L L/2
 CV   M u
M P dx 2 2 
0
EI
 
0
 
2 EI  
L/2 

2  EI
4
5wL
 CV 
384EI
 End slopes : Applying a unit couple at A as shown in Figure (c) above. Consider any
point X at a distance of x from A
wL wx 2 wx ( L  x)
MP  x  (0  x  L )
2 2 2
1 Lx
Mu  1  x  (0  x  L / 2)
L L
The slope at A is given by
 wx ( L  x) 
 dx
L
A   Mu
M P dx

 L  x 
L
2 
0
EI 0
L EI
wL3
A 
24EI

83
Slope and Deflection: Principle of Virtual Work

wL3
C   A 
Due to symmetry 24EI (anti-clockwise direction)

Exercise: (Assignment)
Determine vertical deflection and rotation of point B of the beam shown in Figure (a) below.
The beam is subjected to a couple M 0 at C.
Mo
B
A C

a b
(a)

Worked Examples for Frames


Determine horizontal deflection of C and slope at A of a rigid-jointed plane frame as shown in
Figure (a) below. Both members of the frame have same flexural rigidity, EI.
w/unit length
C C
B C B 1 B 1
1

A A
A 1 (c)
L (b) 1
(a)
Solution:
 Horizontal deflection of C : Apply a unit load C as shown in Figure 4.9(b).
MP  0
Consider AB : ( x measured A ); M u  1 x  x
wL wx 2 wx ( L  x)
MP  x  and
2 2 2
Consider BC : ( x measured C ) M u  1 x  x
The horizontal deflection of point C is given by
M P dx M dx
 CH  M
AB
u
EI
  Mu P
BC
EI
 wx ( L  x) 
L L  dx
  ( x)
(0)dx
  ( x)  2 
0
EI 0
EI
wL4
 CH 
24EI
 Note: THE READER SHOULD DETERMINE THE SLOPE AT POINT

84
Slope and Deflection: Strain Energy Method

SECTION V. Strain Energy Method

5.1 Deflection by Strain Energy Method


The concepts of strain, strain-displacement relationships are very useful in computing energy-
related quantities such as work and strain energy. These can then be used in the computation of
deflections. In the special case, when the structure is linear elastic and the deformations are
caused by external forces only, (the complementary energy U * is equal to the strain energy U )
Pj
the displacement of structure in the direction of force is expressed by
U
j 
Pj
(5.1)
This equation is known as Castigliano's theorem. It must be remembered that its use is limited
to the calculation of displacement in linear elastic structures caused by applied loads. The use of
this theorem is equivalent to the virtual work transformation by the unit-load theorem.

Calculation of Strain Energy


When external loads are applied on an elastic body they deform. The work done is transformed
into elastic strain energy U that is stored in the body. We will develop expressions for the strain
energy for different types of loads.

Consider an elastic spring as shown in the Fig.5.1. When the spring is slowly pulled, it deflects
by a small amount u1 . When the load is removed from the spring, it goes back to the original
position. When the spring is pulled by a force, it does some work and this can be calculated once
the load-displacement relationship is known. It may be noted that, the spring is a mathematical
idealization of the rod being pulled by a force P axially. It is assumed here that the force is
applied gradually so that it slowly increases from zero to a maximum value P. Such a load is
called static loading, as there are no inertial effects due to motion. Let the load-displacement
relationship be as shown in Fig. 5.2. Now, work done by the external force may be calculated
as,
1 1
Wext  P1u1  ( force  displacement )
2 2 (5.2)

A B
P

u
Fig.5.1 linear Spring
The area enclosed by force-displacement curve gives the total work done by the externally
applied load. Here it is assumed that the energy is conserved i.e. the work done by gradually
applied loads is equal to energy stored in the structure. This internal energy is known as strain
energy. Now strain energy stored in a spring is

85
Slope and Deflection: Strain Energy Method

1
U P1u1
2 (5.3)
Work and energy are expressed in the same units. In SI system, the unit of work and energy is
the joule (J), which is equal to one Newton metre (N.m).

F
P
p+dp
p
Strain energy

u X, displacement

Fig.5.2 Force-Displacement relation

Strain energy under axial load


Consider a member of constant cross sectional area A, subjected to axial force N as shown in
Fig. 5.3. Let E be the Young’s modulus of the material. Let the member be under equilibrium
under the action of this force, which is applied through the centroid of the cross section. Now,
the applied force N is resisted by uniformly distributed internal stresses given by average stress
N

A as shown by the free body diagram (vide Fig. 5.3). Under the action of axial load N
applied at one end gradually, the beam gets elongated by (say) u . This may be calculated as
follows. The incremental elongation du of small element of length dx of beam is given by,
 N
du  dx  dx  dx
E AE (5.4)
Now the total elongation of the member of length L may be obtained by integration
L
N
u dx
AE
0 (5.5)
y

N x N N
dx du
x dx u
L

Fig.5.3

86
Slope and Deflection: Strain Energy Method

1
W Nu
Now the work done by external loads 2 (5.6)
In a conservative system, the external work is stored as the internal strain energy. Hence, the
strain energy stored in the bar in axial deformation is,
1
U Nu
2 (5.7)
Substituting equation (5.5) in (5.6) we get,
L
N2
U  dx
2 AE
0 (5.8)

Strain energy due to bending


A positive bending moment applied to a length of beam causes the upper longitudinal fibres to
be compressed and the lower ones to stretch as shown in Fig. 5.4(a). The bending moment
therefore does work on the length of beam and this work is absorbed by the beam as strain energy.

Fig. 5.4 Deflected shape of a symmetrical section beam subjected to a pure


bending moment

Suppose that the bending moment, M, in Fig. 5.4(a) is gradually applied so that when it reaches
its final value the angle subtended at the centre of curvature by the element z is  . From Fig.
5.4(a) we see that
R  z
EI
M 
Substituting in Eq. R for R we obtain
EI
M 
s (5.9)

87
Slope and Deflection: Strain Energy Method

so that  is a linear function of M. It follows that the work done by the gradually applied
moment M is M / 2 subject to the condition that the limit of proportionality is not exceeded.
The strain energy, U , of the elemental length of beam is therefore given by
1
U  M
2 (5.10)
or, substituting for  from Eq. (4.25) in Eq. (5.10),
1 M2
U  s
2 EI x
The total strain energy, U, due to bending in a beam of length L is therefore
M2
U  ds
L 2 EI
x (5.11)
CASTIGLIANO’S THEOREM FOR TRUSSES
L
N2
U  dx
2 AE
The strain energy for a member of a truss is given by Eq. (5.8), 0 . Substituting this
equation into Eq. (5.1) and omitting the subscript j , we have
 L N2
P 0 2 AE
 dx
(5.12)
L  N  dx
   N 
0
 P  AE (5.12)
Where:
  external joint displacement of the truss.
P  external force applied to the truss joint in the direction of 
N  internal force in a member caused by both the force P and the loads on the truss.
L  length of a member.
A  cross-sectional area of a member.
E  modulus of elasticity of a member.

Worked Example 5.1 Find the horizontal deflection at joint C of the pin-jointed frame as shown
in Figure 5.5(a). AE is constant for all members.
C
B P

A D

L
(a)

Fig.5.5(a)

88
Slope and Deflection: Strain Energy Method

Solution:
The force in various members of the frame is shown in Figure 5.5(b). Calculation of strain energy
of the frame is shown in Table 5.1.
P C
B P
- 2P
P 0

A D

(b)

Fig.5.5(b) Force due to applied load


Table 5.1
P2L
Member Length Force( P ) U
2 AE
AB L P P 2 L / 2 AE
BC L P P 2 L / 2 AE
BD 2L P 2 2 2P 2 L / 2 AE
CD L 0 0
 ( 2  1) P 2 L / AE

U   ( 2  1) P 2 L 
 CH    
P P  AE 

Horizontal displacement of joint C ,
 2( 2  1) PL 
   ()

 AE 

CASTIGLIANO’S THEOREM FOR BEAMS AND FRAMES


M2
U 
L
dx
The internal bending strain energy for a beam or frame is given by Eq. (5.11),
0 2 EI x .
Substituting this equation into Eq. (5.1) and omitting the subscript j , we have
 L M2
P 0 2 EI
 dx
(5.14)
L  M  dx
   M 
0
 P  EI (5.15)
Where:
  external displacement of the point caused by the real loads acting on the beam and frame.
P  external force applied to the beam or frame in the direction of  .
M  internal moment in the beam or frame, expressed as a function of x and caused by both the
force P and the real loads on the beam.

89
Slope and Deflection: Strain Energy Method

E  modulus of elasticity of beam material.


I  moment of inertia of cross-sectional area computed about the neutral axis.
If the slope  at a point is to be determined, we must find the partial derivative of the internal
moment M with respect to an external couple moment M  acting at the point, i.e.,
L  M  dx
   M 
0
 M   EI (5.16)

Worked Example 5.2 Determine the displacement of point B of the beam shown in Fig.5.7(a).
Take E  200Gpa, I  50010 mm .
6 4

12 KN/m 12 KN/m P

A B A B
x
10m 10m
(a) (b)

Fig.5.7
Solution:
External force P. A vertical force P is placed on the beam at B as shown in Fig.5.7(b).
Internal Moment M. A single x coordinate is needed for the solution, since there are no
discontinuities of loading between A and B. Using the method of sections, we have
 x
M   Px  12x    Px  6 x 2
2
M
 x
P
Setting P  0 yields
M  6x 2
Castigliano’s Therorem. Appling Eq. (5.15), we have
 M  dx 10 (6 x )( x)dx
2
L
B   M      15 103 KN .m 3 / EI
0
 P  EI 0 EI
or
15  103  103 N  109 mm 3
B   150mm
200  103` N / mm 2  500  106 mm 4

Worked Example 5.3 Determine the slope at point B of the beam shown in Fig.5.8(a). Take
E  200Gpa, I  60  106 mm 4 .

3 KN
B
A C

5m 5m

(a)

90
Slope and Deflection: Strain Energy Method

Solution:
External Couple Moment M  . Since the slope at point B is to be determined, an external couple
M  is placed at this point. Fig.5.8(b)
3 KN
M'
A C
B
x1 x2
(b)

Fig.5.8(b)
Internal moment M. Two coordinates, x1 and x 2 must be used to determine the internal
moments within the beam since there is a discontinuity, M  , at B . As shown in Fig.5.8(b), x1
ranges from A to B and x 2 ranges from B to C. Using the method of sections, the internal
moments and partial derivatives are computed as follows:
For x1 :
M 1  3x1
M 1
0
M 
For x2 :
M 2  3(5  x2 )  M 
M 2
1
M 

Castigliano’s Theorem:
Setting M   0 and Appling Eq. (5.16), we have
L  M  dx
B   M  
0
 M   EI
5 ( 3 x )(0) dx 5  3(5  x )(1) dx
 1 1
 2 2
 112.5KN .m 2 / EI
0 EI 0 EI
 112.5 103 106 Nmm 2
B   0.009375 rad
or 200103 N / mm 2  60 106 mm 4

The negative sign indicates that  B is opposite to the direction of the couple moment M  .

91
Slope and Deflection: Strain Energy Method

Worked Example 5.4 Determine the vertical displacement of point C of the beam shown in Fig.
5.9(a). Take E  200Gpa, I  15010 mm .
6 4

8 KN/m 20 KN

A C B
4m 4m

(a)

Fig.5.9(a)
Solution:
External force P. A vertical force P is applied at point C, Fig.5.9(b). Later this force will be set
equal to a fixed value of 20KN .
8 KN/m 20 KN

A C B
x1 x2
24+0.5P (b) 8+0.5P

Fig.5.9(b)
Internal Moment M. In this case two x coordinates are needed for the integration, Fig.5.9(b),
since the load is discontinuous at C. using the method of sections, we have
For x1 :
M 1  (24  0.5P) x1  4 x12
M 1
 0.5 x1
P
For x2 :
M 2  (8  0.5P) x2
M 2
 0.5 x 2
P

Castigliano’s Theorem.
Setting P  20KN and Appling Eq. (5.15), we have
L  M  dx
 CV   M  
0
 P  EI
4 (34 x  4 x )(0.5 x ) dx
2
4 (18x )(0.5 x ) dx
 1 1 1 1
 2 2 2
0 EI 0 EI
234.7 KN .m 3 192KN .m 3 426.7 KN .m 3
  
EI EI EI
426.7 10 10 N .mm
3 9 3
 CV   14.22mm
or 200103 N / mm 2 150106 mm 4

92
Principle Stresses

CHAPTER 7: PRINCIPAL STRESSES AND MAXIMUM SHEAR STRESS

INTRODUCTION
1. Direct stress (  )
When a body in equilibrium is subjected to external forces, it is reacted by internal forces set up
within the material. If, therefore, a bar is subjected to a uniform tension or compression, i.e. a direct
force, which is uniformly or equally applied across the cross-section, then the internal forces set up
are also distributed uniformly and the bar is said to be subjected to a uniform direct stress, the stress
being defined as
stress (  ) = load  P
area A
Stress σ may thus be compressive or tensile depending on the nature of the load and will be measured
in units of Newton’s per square metre (N/m2) or multiples of this.

2. Direct strain (  )
If a bar is subjected to a direct load, and hence a stress, the bar will change in length. If the bar has
an original length L and changes in length by an amount L , the strain produced is defined as follows:
strain (  )= strain in length  δL
originallength L
Strain is thus a measure of the deformation of the material and is non-dimensional, i.e. it has no units;
it is simply a ratio of two quantities with the same unit.

In many engineering problems both direct (tensile or compressive stress) and shear stresses are
acting at the same time. In such situation the resultant stress across any section will be neither
normal nor tangential to the plane. In this chapter, the stresses acting on an inclined plane (or
oblique section) will be analyzed.

PRINCIPAL PLANES AND PRINCIPAL STRESSES


The planes, which have no shear stress, are known as principal planes. Hence principal planes are
the planes of zero shear stress. These planes carry only normal stresses.
The normal stresses, acting on a principal plane, are known as principal stresses.

METHODS FOR DETERMINING STRESSES ON OBLIQUE SECTION


The stresses on oblique section are determined by the following methods:
1. Analytical method and
2. Graphical method.

ANALYTICAL METHOD FOR DETERMINING STRESSES ON OBLIQUE SECTION


The following two cases will be considered:
1. A member subjected to a direct stress in one plane.
2. The member is subjected to like direct stresses in two mutually perpendicular directions.

93
Principle Stresses

Case I. A member subjected to a direct stress in one plane,


Fig. 7.1 (a) shows a rectangular member of uniform cross-sectional area A and of unit thickness.

Let, P= Axial force acting on the member.


A= Area of cross-section
Consider a cross-section on EF which is perpendicular to the line
of action of the force P.
Then area of section, E = EF  1
= A.
The stress on the section EF is given by
Force P
  … (i)
Aera of EF A
The stress on the section EF is entirely normal stress. There is no shear stress (or tangential stress)
on the section EF.

Now consider a section FG at an angle  with the normal cross-section EF as shown in Fig.7.1
(a).
Area of section FG = FG  1 (member is having unit thickness)
EF EF EF
= 1 ( In EFG ,  cos FG  )
cos FG cos
A
= ( EF  1  A)
cos
Force
 Stress on the section, FG =
Area of section FG
P
= cos   cos … (7.1)
A
The stress, on the section FG, is parallel to the axis of the member. This stress is not normal to the
section FG. The normal stress and tangential stress (i.e., shear stress) on the section FG are obtained
as given below [Refer to Fig. 7.1 [b)].
Let Pn =The component of the force P, normal to section FG =Pcos 
Pt =The component of force P, along the surface of the section FG (or tangential to the
surface FG) = P sin 
 n =Normal stress across the section FG
 t =Tangential stress (i.e., shear stress) across the section FG.
 Normal stress and tangential stress across the section FG are obtained as,

Force normal to section FG


Normal stress,  n 
Area of section FG
pn p cos p
   cos2    cos2  … (7.2)
A A A
cos cos

94
Principle Stresses

Tangential force across section FG


Tangential stress (i.e., shear stress),  t 
Area of section FG
pt p sin p
   sin . cos   sin . cos
A A A
cos cos … (7.3)
 
  2 sin . cos  sin 2
2 2

From equation (7.2), it is seen that the normal stress (  n ) on the section FG will be maximum,
when cos2  or cos is maximum. And cos will be maximum when   0 as cos0  1 . But
when   0 , the section FG will coincide with section EF. But the section EF is normal to the line
of action of the loading. This means the plane normal to the axis of loading will carry the maximum
normal stress.
 Maximum normal stress,
  cos2    cos2 0   … (7.4)

From equation (7.3), it is observed that the tangential stress (i.e. shear stress) across the section FG
will be maximum when sin 2 is maximum. And sin 2 will be maximum when sin 2 =1 or 2
=90° or 270° or  =45° or 135°. This means the shear stress will be maximum on two planes
inclined at 45° and 135° to the normal section EF.
 Max. value of shear stress,
  
 sin 2  sin 90  … (7.5)
2 2 2
Note: From equations (7.4) and (7.5) it is seen that maximum normal stress is equal to  whereas
the maximum shear stress is equal to  / 2 or equal to half the value of greatest normal stress.

Worked Example 7.1: A rectangular bar of cross-sectional area 10000mm 2 is subjected to an


axial load of 20 KN. Determine the normal and shear stresses on a section which is inclined at an
angle of 30º with normal cross-section of the bar
Solution
Given Data:
Cross-sectional area of the rectangular bar, A  10000mm 2
Axial load, P=20 KN =20,000 N
Angle of oblique plane with the normal cross-section of the bar,   30

p 20000
Now direct stress,     2 N / mm 2
A 10000
Let  n = Normal stress on the oblique plane
 t = Shear stress on the oblique plane.
Using equation (7.2) for normal stress, we get
 n   cos2   2  cos2 30  2  0.8662  1.5N / mm2

95
Principle Stresses

Using equation (7.3) for shear stress, we get


 2
 t  sin 2  sin(2  30)  1 sin 60  0.866N / mm 2
2 2

Worked Example 7.2: Find the diameter of a circular bar which is subjected to an axial pull of
160 KN, if the maximum allowable shear stress on any section is 65N/mm2
Solution
Given Data:
Axial pull, P=160 kN= 160000 N
Maximum shear stress = 65 N/mm2


Let D=Diameter of the bar, such that; Area of the bar  D2
4
p 160000 640000
 Direct stress,     N / mm 2
A  D 2
D2
4
Maximum shear stress is given by equation (7.5)
 640000
 Maximum shear stress =  . But maximum shear stress is given as=65
2 2  D 2
N/mm2.
Hence equating the two values of maximum shear, we get
640000 640000
 65  .  D2   1567.  D  39.58mm.
2  D 2 2    65

Worked Example 7.3: A rectangular bar of cross-sectional area of 11000 mm2 is subjected to a
tensile load P as shown in Fig. 7.2. The permissible normal and shear stresses on the oblique
plane BC are given as 7N/mm2 and 3.5 N/mm2 respectively. Determine the safe value of P.
Solution.
Given Data:
Area of cross-section A = 11000 mm2
Normal stress,  n =7 N /mm2
Shear stress,  t =3.5 N/mm2
Angle of oblique plane with the axis of bar =60°.
 Angle of oblique plane BC with the normal cross-section of the bar,   90  60  30

Let P =Safe value of axial pull


 =Safe stress in the member.
Using equation (7.2),  n   cos2 
7
7   cos2 30   2
 9.334N / mm 2
0.866

96
Principle Stresses


Using equation (7.3), t  sin 2
2
 3.5  2 7
3 .5  sin 2  30     8.083N / mm 2
2 sin 60 0.866
The safe stress is the least of the two, i.e., 8.083 N/mm2.
 Safe value of axial pull, P =Safe stress  Area of cross section
=8.083  11000
=88913 N =88.913 kN.

Case II. A member subjected to like direct stresses in two mutually perpendicular directions.

Fig. 7.3 (a) shows a rectangular bar ABCD of uniform cross-sectional area A and of unit thickness.
The bar is subjected to two direct tensile stresses (two-principal tensile stresses) as shown in the
Fig.7.3 (a).

Let FC be the oblique section on which stresses are to be calculated. Consider the forces acting on
wedge FBC.
Let  =Angle made by oblique section FC with normal cross-section BC
1 =Major tensile stress on face AD and BC
2
2 =Minor tensile stress on face AB and CD
P1 =Tensile force on face BC D C
P2 =Tensile force on face FB. Pt 1
1
The tensile force on face BC, P1 = 1 BC 1
F Pn
P1=  1  Area of cross section=  1  BC  1 A B
The tensile force on face FB,
P2=Stress on FB  Area of FB =  2  FB  1. 2 P2= 2 BF 1

Fig.7.3(a)
The tensile forces P1 and P2 are also acting on the oblique section FC. The force P1 is acting in the
axial direction, whereas the force P2 is acting downwards as shown in Fig. 7.3 (a). The two forces
P1 and P2 each can be resolved into two components i.e. one normal to the plane FC and other along
the plane FC.
Let Pn=Total force normal to section FC
= Component of force P1 normal to section FC + Component of force P2 normal to section
FC
= P1 cos  P2 sin
=  1  BC  cos   2  BF  sin
Pt = Total force along the section FC
= Component of force P1 along the section FC + Component of force P2 along the section FC
= P1 sin  (P2 cos ) (-ve sign is taken due to opposite direction)

97
Principle Stresses

= P1 sin  P2 cos
=  1  BC  sin   2  BF  cos (Substituting the values P1 and P2)

 n = Normal stress across the section FC


Total force normal to the section FC
=
Aera of section FC
Pn

FC  1
 1  BC  cos   2  BF  sin

FC
BC BF
 1   cos   2  sin
FC FC
  1  cos  cos   2  sin  sin
BC BF
(  In triangle FBC,  cos ,  sin )
FC FC
  1  cos2    2  sin 2  ...[ 7.5( A)]
1  cos 2 1  cos 2
 1( ) 2( )
2 2
   2 1  2
 1  cos 2 ...( 7.6)
2 2

 t =Tangential stress (or shear stress) along section FC


Total force along the section FC

Area of section FC
Pt

FC  1
  BC  sin   2  BF  cos
 1
FC
BC BF
 1   sin   2  cos
FC FC
  1  cos  sin   2  sin  cos
BC BF
(  In triangle FBC,  cos ,  sin )
FC FC
 ( 1   2 )  cos sin
( 1   2 )
( )  2 cos sin
2
(   2 )
 1 sin 2 ...( 7.7)
2

The resultant stress on the section FC will be given as


 R   n2   t2 … (7.8)

98
Principle Stresses

Obliquity [Refer to Fig. 7.3 (b).]


The angle made by the resultant stress with the normal of the oblique plane, is known as obliquity.
It is denoted by  . Mathematically,
Pt
tan   …[ 7.8(A)]
Pn
D C

Pt
PR
Pt
Pn
A F B
Fig.7.3(b)
Maximum shear stress
The shear stress is by equation (7.7). The stress (  t ) will be maximum when sin 2  =1 or 2  =90°
or 270° ( sin 90°=1 and also sin 270°=1) or  =45° or 135°

And maximum shear stress,


  2
( t ) max  1 … (7.9)
2
Hence the planes, which are at an angle of 45" or 135° with the normal cross-section BC (see Fig.
7.3 (b)), carry the maximum shear stresses.

Principal planes
Principal planes are the planes on which shear stress is zero. To locate the position of principal
planes, the shear stress given by equation (7.7) should be equated to zero.
For principal planes,

99
Principle Stresses

( 1   2 )
sin 2  0
2
or sin 2  0 [ ( 1   2 ) cannot be equal to zero]
or 2  0 or 180
  0 or 90
   2 1  2
when   0 ,  n  1  cos 2
2 2
   2 1  2
 1  cos 0
2 2
   2 1  2
 1  1   1
2 2
   2 1  2
when   90 ,  n  1  cos 2  90
2 2
   2 1  2
 1  cos180
2 2
   2 1  2
 1   (1)   2
2 2

Worked Example 7.4: The principal tensile stresses at a point across two mutually perpendicular
planes are 120 N/mm2 and 60 N/mm2. Determine the normal, tangential and resultant stresses on
a plane inclined at 30° to the axis of the minor principal stress.
2
2 =60N/mm
=120N/mm 2

=120N/mm 2
Axis of minor
principal stress

30°
Axis of major
principal stress
1

2
2 =60N/mm

Fig.7.4
Solution
Given Data:
Major Principal Stress,  1 = 120 N/mm2
Minor principal,  2 =60 N/mm2
Angle of oblique plane with the axis of minor principal stress,  =30°.

Normal stress
The normal stress (  n ) is given by equation (7.6),

100
Principle Stresses

1   2 1   2
n   cos 2
2 2
120  60 120  60
  cos(2  30)  105N / mm 2 . Ans
2 2

Tangential stress
The tangential (or shear stress)  t is given by equation (7.7).

( 1   2 ) (120  60)
t  sin 2  sin(2  30)  25.98N / mm 2 Ans
2 2
Resultant stress
The resultant stress (  R ) is given by equation (7.8)
  R   n2   t2  1052  25.952  108.16 M / mm 2 Ans

The End!

101

You might also like