You are on page 1of 115

ME 212 MECHANICS OF MATERIALS - I

CREDIT HOURS: THEORY = 3 HRS


COURSE OUTLINE

Strength of materials
Andrew Pytel
Ferdinand L. Singer
TENTATIVE COURSE SCHEDULE

WEEK No. LECTURE TOPIC


1 Simple stress: Normal stresses
2 Simple stress: Shear stress
3 Thin-walled pressure vessels
4 Simple strain: Stress-strain diagram
5 Simple strain: Hooke’s law: Axial and shearing deformations
6 Simple strain: Poisson’s ratio: Biaxial and triaxial deformations
7 Thermal stresses
8 Thermal stresses
9 Mid Term Exam
10 Torsion
11 Torsion
12 Shear and moment in beams
13 Shear and moment in beams
14 Stresses in beams
15 Stresses in beams
16 Beam deflections
STUDENT PERFORMANCE EVALUATION / GRADING

• Assignments, Quizzes, Mid Term and the Final examination will evaluate student
progress.

FINAL GRADE

The final grade will be determined by averaging each section and assigning them the
following weights:

Quizzes ............................12.5%
Assignments………………12.5%
Mid Term Examination……25%
Final Examination .............50%
---------------------------------------------------------
Total ............................. 100%
PARTICIPATION

The course consists of a three-hour lecture per week.


Attendance for all lectures is strongly advised.
Any student whose attendance is less than 75% will be dropped
from the course for insufficient participation.
Thermal Stresses
All of the members and structures that we have considered so far were assumed to
remain at the same temperature while they were being loaded. We are now going to
consider various situations involving changes in temperature.

If the temperature increases, generally a material expands, whereas if the temperature


decreases, the material contract.

Ordinarily, this expansion or contraction is linearly related to the temperature increase


or decrease that occurs.

If this is the case, and the material is homogeneous and isotropic, it has been found
from experiment that the deformation of a member having a length L can be calculated
using the formula
Coefficient of linear expansion
To understand thermal expansion, consider a simple model of a solid, the atoms of which are held together in a regular array.

As the temperature increases, the kinetic energy in


vibration of the atoms and molecules also increases. The
increased vibration makes each particle take up a little
more space, causing thermal expansion.

The forces between atoms can be compared to the forces that would be exerted by an array of springs connecting the atoms
together. At any temperature above absolute zero (–273.15°C), the atoms of the solid vibrate.

When the temperature is increased, the amplitude of the vibrations increases, and the average distance between atoms
increases. This leads to an expansion of the whole body as the temperature is increased. The change in length that arises from
a change in temperature (ΔT) is designated by δT.

Through experimentation, we find that the change in length δT is proportional to the change in temperature ΔT and the original
length L. Thus,

where α is called the coefficient of linear expansion.

It can be defined as the fractional change in length per unit change in temperature. This coefficient has different values for
different materials.

A larger value means that the material expands more over a set temperature rise than a material with a lower coefficient.
Thermal Stresses

At room temperature 

After raising the


temperature by DT

• A temperature change results in a change in length or thermal strain. There


is no stress associated with the thermal strain unless the elongation is
restrained by the supports.
Thermal Stresses
If temperature deformation is permitted to occur freely, no load or stress will be induced in
the structure. But in some cases it may not be feasible to permit these temperature
deformations; the result is that internal forces are created that resist them. The stresses
caused by these internal forces are known as thermal stresses.
For a homogeneous rod mounted between unyielding supports as shown, the thermal
stress is computed as:

Treat the additional support as redundant and apply the principle of superposition.

Deformation due to temperature changes;


Redundant :1. Exceeding what is
necessary or natural; superfluous.
2. Needlessly repetitive

Deformation due to equivalent axial stress;

The thermal deformation and the deformation from the redundant support must be
compatible.

where σ is the thermal stress in MPa, E is the modulus of elasticity of the rod in MPa.
Long structures such as railway tracks and
pipelines can fail from buckling instability by such
stresses. In the picture shown, a worker inspects
a buckled railway track that has failed from the
action of heat. The rail component - which is
welded at both ends to the next bit of track - is
constrained to the original length and so the
thermal expansion, in turn, introduces a
compressive stress along its length. When this
stress becomes too large, the rail cannot sustain
it in its original shape and it buckles sideways as
shown.
This process is highly unstable and can lead to
unsafe designs if the mechanics of this type of
system are not well understood. It can be seen in
the photograph that the buckling is localized to
one section of track, this is an important and
generic characteristic of long structures that have
some sort of support along their length.
Railway track: Railway tracks are an example where the concept of thermal expansion is used.
Space is left between railway tracks as an allowance for their expansion when temperature
increases, otherwise the rails may buckle.

Gaps between successive lengths of rails


Thermal stresses can cause the failure of machine components, buildings, and
structural members. Expansion joints and other means, including gaps between
successive lengths of rails, gaps between the blocks of a dam, and rollers on
bridge supports, are used to prevent such failures.

Bridges have expansion joints to allow for


thermal expansion of concrete.
Most bridges are built with gaps that allow the bridge to expand
without cracking the material.
Greek letters

Change in length, δ = Delta

Normal stress, σ = Sigma

Normal strain, ε = Epsilon


Sample Problem #
Additional examples

T   F
L
LDT 
E
  EDT
  (29 103 )(6.60 10 6 )(120  60)
  11.5ksi
Sample Problem #

L  2.5m
Ti  20o C
P0
T f  20o C
A  1200mm 2
 m m
  11.7 o  11.7 10 6 o
m. C m. C
E  200GPa  200 109 N / m 2

 ?
Sample Problem #

A  0.25inch 2
Ti  70o F
P  1200lb
inch
  6.5 10 6
inch o F
E  29 106 psi

 ? Tf  0

Tf  ?  0
Solution:

As the temperature is decreasing, therefore the temperature stresses are also tensile.

 Total   P   T
P
  EDT
A
1200 Upon heating (Tf > Ti ), the stress is
   (6.5 10 6 )(29 106 )(70) compressive (σ < 0), since rod expansion
0.25 has been constrained.
  4800  13195
  17995 psi  17.995ksi If the rod is cooled(Tf < Ti ), a tensile stress
will be imposed(σ > 0).
  18ksi
If the stresses become zero, then
 Total   P   T
0   P T
Upon heating (Tf  Ti ), the stress will be compressive (  0),
since rod expansion has been constrained.
0   P  ( T )
 T   P
P
EDT 
A
P
DT  Upon heating (Tf > Ti ), the stress is
AE compressive (σ < 0), since rod expansion
1200
DT  has been constrained.
0.25  6.5 10 6  29 106
1200 If the rod is cooled(Tf < Ti ), a tensile stress
DT  will be imposed(σ > 0).
47.125
DT  25.46
T f  Ti  25.46
T f  25.46  Ti
T f  25.46  70
T f  95.5o F
Ti  20o C
T f  20o C
P  5000 N
 max  130MPa
m
  11.7 10 6
m oC
E  200 109 N / m 2

d min  ?
As the temperature of the rod is decreasing, therefore the temperature stresses will be tensile.
 Total   P   T
P
  EDT
A
5000
130 106   (11.7 10 6 )(200 109 )(40)
A
5000
130 106   93.6 106 Upon heating (Tf > Ti ), the stress is
A
compressive (σ < 0), since rod expansion
5000
 130 10  93.6 10
6 6
has been constrained.
A
5000
 36 106 If the rod is cooled(Tf < Ti ), a tensile stress
A will be imposed(σ > 0).
5000  (36 10 ) A
6

5000 d 2
A 3 6
 0.1373 10  137.3 10 m  137.3mm
2 2
 A
36 106 4
d 2
137.3 
4
4(137.3)
d2   174.8

d  13.2mm
L  10m
Ti  15o C
m m
  11.7 o
 11.7 10 6
mC m oC
E  200GPa  200 109 N / m 2
a )T f  ? If  T  3mm  3 10 3 m
b) T ? If DT  25.6

Solution :
a ) T  LDT
3 10 3  11.7 10 6 10  DT
3 10 3 b) T  EDT
 DT 
11.7 10 6 10
 T  (200 109 )(11.7 10 6 )  25.6
DT  25.6
 T  59.9 106
If we increase the temperature of rail.
 T  60MPa
The rail will expand and will fill the gap.
So the final temperature may be found as
T f  Ti  DT
T f  15  25.6  40.6o C
Torsion
So far we studied how to calculate the stresses and strains in structural members subjected
to axial loading i.e. to forces directed along the axis of the member.

We are now going to consider various structural members and machine parts that are in
torsion.

Torsion is the twisting of an object due to an applied torque.


Torsion, the stress produced in a body when it is twisted. The twist in the body
is produced by torque. A body such as a cylindrical rod or a structural beam is
in torsion when one end is held stationary while the other end is rotated at right
angles to a line (the longitudinal axis) running the length of the rod or beam.
Torsion also results when the two ends of a body are twisted in opposite
directions.

A body will, within limits, tend to resist being in torsion; it will tend to snap back
to its original condition when the torque is removed. Devices such as helical
(spiral) springs, and toy airplanes powered by twisted rubber bands, are based
on this characteristic. Torsion bars, which act as springs in automobiles, are
also applications of torsion.

When torsion exceeds a body's ability to withstand it, the body will shear or
break. Structural beams, gear shafts, and other objects that are, or may be,
subjected to torque must be designed to withstand the torsion that will be
produced in them.
Members in torsion are encounter in many
engineering applications.

The most common application is provided


by transmission shafts, which is used to
transmit power from one point to another.

For example:

•Turbine exerts torque T on the shaft

•Shaft transmits the torque to the generator

•Generator creates an equal and opposite


torque T
Torsion of Circular Shafts
To better understand the behavior of a torsion
member, a grid will be superimposed on the surface
of a shaft to help illustrate the deformation of the
shaft in response to an applied torque T.

When a circular shaft is subjected to torsion, every


cross section remains plane and undistorted. In
other words, while the various cross sections along
the shaft rotate through different amount, each
cross section rotates as a solid rigid slab.
Imagine the shaft to consist of innumerable thin
slices, each of which are rigid and joined to adjacent
slices by elastic fibers.

Slice 2 will rotate past slice 1 until the elastic fibers


joining them are deformed enough to create a
resisting torque that balances the applied torque.

When this happens, slice 1 and 2 act as a rigid unit


and transmit the torque to slice 3; this slice will rotate
enough so that the elastic fibers joining it and slice 2
develop a resisting torque equal to the applied torque.

This type of deformation proceeds throughout the


length L of the shaft.
Derivation of Torsion Formulas
To examine the stresses and deformations of a torsion member, we will consider
a circular shaft of length L and radius r that is attached to a fixed support at one
end.

If a torque T is applied to the other end, the shaft will twist, with its free end
rotating through an angle  called the angle of twist.

Consider now any internal fiber located a radial distance  (rho) from the axis of
the shaft.

A fiber AB on the outside surface, which is originally straight will be twisted into a
helix A'B as the shaft is twisted through the angle  .

We consider the small square element formed by two adjacent circles and two
adjacent straight lines traced on the surface of the cylinder before any load is
applied.

As the shaft is subjected to a torsional load, the element deforms into a


parallelogram.

Since the circles defining two of the sides of the element considered here remain
unchanged, the shearing strain must be equal to the angle between lines AB and
A'B.
We observe from Figure 3.13c that, for small values of  , we can
express the arc length AA' as AA' = L
DABA'
But, on other hand we have AA' = 
P AA' AA'
It follows that tan    
B AB L
L =  '
 AA
  
L L
AA  L
'

Where  and  are both expressed in radians.

The shearing strain  at a given point of a shaft in torsion is proportional to


the angle of twist .

It also shows that  is proportional to the distance  from the axis of the
shaft to the point under consideration.

Thus, the shearing strain in a circular shaft varies linearly with the distance
from the axis of the shaft.

Recalling Hooke’s law for shearing stress and strain,   G


The equation obtained shows that, as long as the yield strength (or proportional limit) is not
exceeded in any part of a circular shaft, the shearing stress in the shaft varies linearly with the
distance  from the axis of the shaft. Figure 3.14a shows the stress distribution in a solid
circular shaft of radius r, and Fig. 3.14b in a hollow circular shaft of inner radius r and outer
radius R.

r R
Relationship between Torque and Shear stress
To determine the relationship between the torque transmitted by the shaft and the shear
stress that is developed internally in the shaft material.

Considering a shaft AB subjected at A and B to equal and opposite torques T and T',
we pass a section perpendicular to the axis of the shaft through some arbitrary point C.

The shaft is divided into two segments by the cutting plane. Consider free body
diagram of portion BC of the shaft.

Figure 3.4 shows a cross section of the shaft containing a differential element of area
dA located at the radial distance  from the axis of the shaft which carries the
differential resisting load d P.

Summing the contributions of all the differential elements across the cross-sectional
area A and equating the result to the internal torque yields
G
  G 
L
T G
 
J L
The Polar Moment of inertia is the property of an area which measures
resistance of that area to torsion.

The Polar Moment of Inertia is a geometric property of a cross section.


Physically, it is a measure of how difficult it is to turn a cross-section about an
axis perpendicular to it (the inherent rotational stiffness of the cross-section).
This means:

The greater the Polar Moment of Inertia, the more torque is require to turn the
shaft by a certain angle.

The greater the Polar Moment of Inertia, the smaller the Shear Stress required
to produce a given torque.
Sample Problem #

Alu min um Steel


Lal  6 ft  (6 12)" LSt  3 ft  (3 12)"
d al  3
" d St  2"
Gal  4 106 psi GSt  12 106 psi

T  10kip  inch
( max ) al  ?
( max ) St  ?
Lal  6 ft  (6 12)"
d al  3"
Gal  4 106 psi

LSt  3 ft  (3 12)"
d St  2"
GSt  12 106 psi
Sample Problem #

L  3'  3 12  36"


d  4"
T  15kip. ft
G  12 106 psi
 max  ?
 ?

Solution : Tr
 max 
J

d 4
J
32
 (4) 4
J  25.13inch 4
32
Tr
 max 
J

(15 103 12)  2


 max   14325.51 psi
25.13
 max  14.325 103 psi
 max  14.325ksi

TL 180  radian

JG 180
1radian 
15 103 12  (3 12) 

25.13  (12 106 ) 0.0215 180
0.0215radian 
  0.0215rad 
  1.23  0.0215radian  1.23
Sample Problem #

 max  3
L  6m
T  12kNm
N
G  83 109
m2
d min  ?
 max  ?
Solution :
TL
 d 4
JG J
TL 32
 4
d
G
32
32TL
 4
d G 1
32TL d
d4  
G
32(12 103 )(6) 180  radian
d 
4
3 
 ( )(83 109 ) 1  radian
180 180
d 4  10875m 4
d  0.1139m
d  113.9mm
d 4 d
 max 
Tr
J d  2r  r 
J 32 2

d
T
 max  24
d
32
Td 32
 max  
2 d 4
16T
 max  3
d

16  (12 103 )
 max 
 (0.1339)3
 max  41.277 Mpa
Sample Problem #

D  100mm
d  80mm
 max  60MPa
 max
 max  0.5 deg/ m   0.5
L
N
G  83 109
m2
T ?
Based on maximum allowable shearing stress :
TR
 max 
J
D
T
  2
max

(D4  d 4 )
32
16TD
 max 
 (D4  d 4 )

 max ( D 4  d 4 )
T
16 D
 (60 106 ){(0.1) 4  (0.08) 4 }
T
16(0.1)
T  6955 N  m
Based on maximum allowable angle of twist :

TL

JG

 L TL
 0.5 
L 360 JG
  0.5  L L TL

L 360 G  ( D 4  d 4 )
  ( )
2 32
L   2G ( D 4  d 4 )
 (  )rad 180  radian T
2 180 1 

radian
360  32
L 180
 2 (83 109 ){(0.1) 4  (0.08) 4 }
 ( )rad T
360 360  32
T  4198.283 N  m

Use the smaller torque, T = 4 198.28 N·m.


Sample Problem #

d  40mm  40 10 3 m  0.04m


D  60mm  60 10 3 m  0.06m
L  1.5m
 max  120MPa  120 106 N / m 2
T ?
TR
 max  d  2r  r 
d
J 2
TD
 max 
2J
2 J max
T
D

 
J (D4  d 4 )  {(0.06) 4  (0.04) 4 }  1.02 10 6 m 4
32 32

2 J max
T
D
2 1.02 10 6 120 106
T
0.06
T  4.08 103 N  m
T  4.08 KN  m
Bending
So far we studied how to calculate the stresses and strains in prismatic
members subjected to axial loads or to twisting couples.

We are now going to analyze the stresses and strains in prismatic members
subjected to bending.

Bending is a major concept used in the design of many machine and structural
components, such as beams.
The way a part is loaded determines whether it is called a tensile or compressive
member, a torsional shaft, or a beam.

If you take a ½ inch diameter steel rod and pull it lengthwise, the rod will develop a
tensile stress σ = P / A, where A is the cross-sectional area of the rod.

Loading the rod in tension parallel to its axis makes the rod a tensile member; loading
it in compression parallel to its axis makes it a compressive member.

If you twist the steel rod with torque T, then we call it a torsional shaft. If loading is
perpendicular (transverse) to its axis so that the rod bends, then the rod called a
beam. You can load a beam with point loads, uniformly distributed loads, or non
uniformly distributed loads.
Shear and Moment in Beams
Beams are important structural and mechanical elements in engineering.

Beams are structural members which offer resistance to bending due to applied
loads.

In this chapter, we will determine the stress in these members caused by bending.

Members that are slender and support loadings that are applied perpendicular to
their longitudinal axis are called beams.

In general, beams are long, straight bars having a constant cross-sectional area.

Examples include members used to support the floor of a building, the deck of a
bridge, or the wings of an aircraft.

Also, the axle of an automobile, the boom of a crane etc.

Slender: Having little


width in proportion to
height or length
Boom of a crane

Wings of an aircraft
Wing of a commercial airplane. Aircraft wings can be analyzed for stresses and
deformations by modeling them as cantilever beams.
Deck of a bridge

Axle of an automobile
Power-generating turbines on a
wind farm.

The supporting columns can be


modeled as beams subjected to
wind loading.

The determination of shear forces


and bending moments in beams
caused by various load conditions
is the topic of this chapter.
Because of the applied loadings, beams develop an internal shear force and bending
moment that, in general, vary from point to point along the axis of the beam.

In order to properly design a beam, it is first necessary to determine the maximum shear
and moment in the beam.

One way to do this is to express V and M as functions of the arbitrary position x along the
beam’s axis. These shear and moment functions can then be plotted and represented by
graphs called shear and moment diagrams. The maximum values of V and M can then
be obtained from these graphs.

Also, since the shear and moment diagrams provide detailed information about the
variation of the shear and moment along the beam’s axis, they are often used by
engineers to decide where to place reinforcement materials within the beam or how to
proportion the size of the beam at various points along its length.
TYPES OF BEAMS
Beams are classified according to their supports.

A simply supported beam, shown in Fig. 4.1(a), has a pin support at one end and a roller support at the
other end. The pin support prevents displacement of the end of the beam, but not its rotation. The term roller support
refers to a pin connection that is free to move parallel to the axis of the beam; hence, this type of support suppresses
only the transverse displacement.

A cantilever beam is built into a rigid support at one end, with the other end being free, as shown in
Fig. 4.1(b). The built-in support prevents displacements as well as rotations of the end of the beam.

An overhanging beam, illustrated in Fig. 4.1(c), is supported by a pin and a roller support, with one or both ends of the
beam extending beyond the supports. The three types of beams are statically determinate because the support
reactions can be found from the equilibrium equations.
Types of Loading
A concentrated load, such as P in Fig. 4.1(a), is an approximation of a force that acts
over a very small area.

In contrast, a distributed load is applied over a finite area. If the distributed load acts on
a very narrow area, the load may be approximated by a line load. The intensity w of this
loading is expressed as force per unit length (lb/ft, N/m, etc.).

The load distribution may be uniform, as shown in Fig. 4.1(b), or it may vary with
distance along the beam, as in Fig. 4.1(c).

The weight of the beam is an example of distributed loading, but its magnitude is usually
small compared to the loads applied to the beam.
Computing the reactions from the equilibrium equations

   Fy  0
RA  RD  14  28  0..............1)

M A  0
RD (7)  28(5)  14(2)  0
28(5)  14(2)
RD 
7
RD  24kN

Substituting value of RD in eq. 1), we get

RA  24  14  28  0
RA  18kN
An imaginary section, indicated by the dashed line,
is passed through the beam anywhere within region
AB, cutting it into two parts and analyze the portion
of the beam to the left of this section.
The analysis of the part to the left of section 2
gives
Analyzing the portion of the beam to the left
of section 3 , we obtain
To draw the Moment Diagram:

The equation M AB  18x is linear which produces a straight line between A and B, at x  0, M AB  0 and at x  2 m, M AB  36 kN·m.

The equation M BC  4x  28 is also linear. At x  2 m, M BC  36 kN·m and at x  5 m, M BC  48 kN·m.

M CD  – 24x  168 is again linear. At x  5 m, M CD  48 kNm and at x  7 m, M CD  0.


Computing the reactions from the equilibrium equations

M A  0
RC ( L)  C  0
C
RC 
L

M C  0
 RA ( L)  C  0
C
RA  
L
To draw the Moment Diagram:
C x
The equation MAB = - L is linear which
produces a straight line between A and
B.
At x = 0, MAB = 0 and at x = 3L
4
, M AB = -
3C 

4
To draw the Moment Diagram:
Co
The equation MBC = - x  Co is also linear.
L
3L 1
At x = 4 , MBC = 4
Co and

at x = L, MBC = 0.
Computing the reactions from the equilibrium equations

R B  R C - 200 - 1200  0(1)

 MB  0
R C (10) – 1200(5)  200(4)  0
 200(4)  1200(5)
RC 
10
R C  520 lb

Substituting value of R C in eq.1, we get


R B  520 - 200 - 1200  0
R B  880 lb
Figure (c) shows the FBD of the portion of the beam that lies to the left of section 1 .
(The part of the beam lying to the right of the section could also be used.)

The equilibrium equations for this part of the beam yield


The FBD of the part of the beam
that lies to the left of section 2 is
shown in Fig. (d).

(The portion of the beam lying to


the right of the section could also
be used.)

Applying the equilibrium equations


to the beam segment, we obtain
To draw the Shear Diagram:
Segment AB

For segment AB, VAB = – 200 lb; at x = 0 ft, VAB = - 200 lb and at x = 4 ft, VAB = - 200 lb.

Segment BC

For segment BC, VBC = 1160 – 120x is linear; at x= 4 ft, VBC = 680 lb, at x = 14 ft,
VBC = –520 lb.
To draw the Moment Diagram:
Segment AB

The equation MAB = –200x is linear, at x = 0, MAB = 0 and at x = 4 ft, MAB = –200(4)= -800 lb-ft.

Segment BC

The equation MBC = -60x2 +1160x – 4480 lb.ft for segment BC is second degree curve; at x = 4 ft,
MBC = -800; at x = 14 ft, MBC = 0 lb.ft.
The highest point of
a curve has zero
slope.
Centroid Location Complex Shapes
Review: Calculating area of simple shapes

Area of a
square = Area of a rectangle =

Side2 Width * Height

Area of a triangle =
Area of
a circle =
πr2 ½ (base)(height)
Centroid of a Right Triangle
The centroid of a triangle is constructed by taking any given triangle and
connecting the midpoints of each leg of the triangle to the opposite vertex.
The line segment created by connecting these points is called the median.
You see the three medians as the dashed lines in the figure below.
No matter what shape your triangle is, the centroid will always be inside the
triangle.
The centroid is the center of a triangle that can be
thought of as the center of mass. It is the balancing
point to use if you want to balance a triangle on the
tip of a pencil.

For rectangle center of gravity is at the middle but in


case of triangle its 1/3 from the height.
Centroid Location
The centroid of a right triangle is located at a
distance of 1/3 its height and 1/3 its base.

B
Computing the reactions from the equilibrium equations

1000 + RB – 2160 = 0
RB = 1160 lb

 MA = 0 MB + 1160(12) - 2160(8) = 0

MB = 3360 lb-ft
Derivation of Flexure formula
The stresses caused by the bending moment are known as bending stresses, or flexure
stresses.

The relationship between these stresses and the bending moment is called the flexure
formula.
Figure 5-1a shows two adjacent sections ab and cd separated by the distance dx.

Because of the bending caused by load P, section ab and cd rotate relative to each
other by the amount dѳ as shown in Figure-b, but remain straight and undistorted.

Fiber ac at the top is shortened, and fiber bd at the bottom is lengthened.

Somewhere between them is located fiber ef, whose length is unchanged.


Drawing the line c'd' through f parallel to ab shows that fiber ac is shortened an
amount cc' and is in compression and that fiber bd is lengthened by an amount d'd
and is in tension.

The plane containing fiber like ef is called the neutral surface because such fiber
remain unchanged in length and hence carry no stress.

It will be shown shortly that this neutral surface contains the centroids of all
transverse sections.
Consider now the deformation of a typical longitudinal fiber gh located distance y units
below the neutral surface and on the tension side of the beam.

Its elongation hk is the arc of a circle of radius y subtended by the angle dѳ and is given
by
  hk  yd

The strain is found by dividing the deformation by the original length ef of the fiber:

 hk yd
  
L gh gh
If we denote the radius of curvature of the neutral surface by , the curved length ef is
equal to d , from which the strain becomes

 hk yd yd y
    
L gh ef d 
Assuming that the material is homogeneous and obeys Hooke’s law, the stress in fiber gh is
given by
y E
  E  E  ( )y
 
y E
  E  E  ( )y ……….a)
 

The above equation indicates that the normal stress in any longitudinal fiber varies
directly with its location y from the neutral surface, since it is assumed that the modulus
of elasticity E is equal in tension and compression and the radius of curvature of the
neutral surface is independent of the location y of the fiber.

However the stress must not exceed the proportional limit, for this would invalidate
Hooke’s law on which this stress variation is based.
To complete the derivation of the flexure formula,
we apply the condition of equilibrium.

Equilibrium requires that the resultant of the normal


stress distribution over the cross section must be equal
to the bending moment M acting about the neutral axis (z-axis).

Figure 5.3 shows a typical cross section of a beam.

The normal force acting on the infinitesimal area dA


of the cross section is
dP   dA

The resisting moment about the neutral axis of


a typical element of area dA is
Moment  Force  Perpendicular distance
Moment  dP( y )
Moment   x dA( y )

Summing the contributions of all the differential elements across


the cross - sectional area A and equating the result to the resisting moment yields
Resisting moment  M r   y x dA
To satify the condition of static equlibrium, we apply  M Z  0,
or the fact that the applied moment M equals the resisting moment M r .
The resisting moment M r is the sum of the resisting moment developed by
all differential loads dP.

Applied moment  Resisting moment


M  Mr
M   y x dA

Ey  x  Ey / 
M y dA

E

M y dA 2

EI
M Moment of inertia, I   y 2 dA

EI

M

Ey EI Ey Ey
  a)    
 M  

My

I
My

I

This formula indicates that the flexure stress in any section varies directly with the distance
of the section from the neutral axis.

In a more common of the flexure formula, y is replaced by the distance c, which is defined
as the distance from the neutral axis to the remotest element.

With this change, the maximum flexure stress in any section is given by

Mc
 max 
I

If I/c is called the section modulus and denoted by S,


another common variation of the flexural formula is

M M
 max  
I c S

where S= I/c is called the section modulus of the beam.


The formulas for the section moduli of common cross
sections are given in Fig. 5.4.
I is the Moment of Inertia. For a rectangular beam .

Where S= I/C is called the section modulus of the beam.


The formulas for the section moduli of common cross sections are given
in Fig. 5.4.
Beam Deflection - Double-Integration Method

Figure 6.1(a) illustrates the bending deformation of a beam.

Recall that in a real beam, the displacements and slopes are very small if the stresses are
below the elastic limit, so that the deformation shown in the figure is greatly exaggerated.

The deformed axis of the beam is called its elastic curve.

In this section, we derive the differential equation for the elastic curve and describe a
method for its solution.
Differential equation of the elastic curve

As shown in Fig. 6.1(a), let x be the horizontal coordinate of an arbitrary point A on the
axis of the beam, measured from the fixed origin O.

As the beam deforms, its axis becomes curved and A is displaced to the position A'. The
vertical deflection of A, denoted by v, is considered to be positive if directed in the
positive direction of the y-axis—that is, upward in Fig. 6.1(a).

Because the axis of the beam lies on the neutral surface, its length does not change.
Therefore, the distance OA', measured along the elastic curve, is also x.

It follows that the horizontal deflection of A is negligible provided the slope of the elastic
curve remains small.
Consider next the deformation of an infinitesimal segment AB of the beam axis, as shown
in Fig. 6.1(b).

The elastic curve A'B' of the segment has the same length dx as the undeformed
segment.

If let v be the deflection of A, then the deflection of B is v + dv, with dv being the
infinitesimal change in the deflection over the length dx.

Similarly, the slope angles at the ends of the deformed segment are denoted by θ and θ
+ dθ.
From the geometry of the figure, we obtain
MASS MOMENT OF INERTIA (Review)
Area Moment of Inertia (Review)
lam-i-na: A thin plate, sheet, or layer.
(Review)

Ixx
Iyy
Final Exam - Paper Pattern
First attempt Q.-I (MCQ) on separate Answer Sheet which shall be taken back after 20
minutes.

Q. 1 Multiple Choice Questions (20 MCQS)

Q. 2 Thermal stresses

Q.3 Torsion

Q.4 Shear and Moment in beam

Q.5 Flexure formula

•The final exam paper will be a test not of your memory but a test of whether or not you can apply
the conceptual knowledge you have gained in the class to new situations.

•Questions in the exam will not be tougher than those taught in the class. So if you can handle
those, then you can definitely handle the exam paper.

You might also like