Strength All Chapter

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 148

Strength of Materials

GIRUM MINDAYE
ADDIS ABABA SCIENCE AND TECHNOLOGY UNIVERSITY | JULY 2021
Strength of Materials AASTU

Table of Contents
1 MECHANICAL PROPERTIES OF MATERIALS .................................................................. 1
1.1 Introduction ........................................................................................................................... 1
1.2 Stress-Strain .......................................................................................................................... 2
1.2.1 Normal Stress ................................................................................................................ 2
1.2.2 Normal Strain ................................................................................................................ 3
1.2.3 Normal Stress-Strain Diagram .................................................................................... 3
1.2.4 Changes in Lengths under Non-uniform Conditions ................................................. 6
1.3 Poisson’s Ratio ...................................................................................................................... 6
1.4 Bulk Modulus ........................................................................................................................ 9
1.5 Shear Stress-Strain ............................................................................................................... 9
1.5.1 Shear Stress ................................................................................................................... 9
1.5.2 Shear Strain ................................................................................................................. 10
1.5.3 Shear Stress-Strain Relationship ............................................................................... 11
1.6 General State of Stress........................................................................................................ 12
1.7 Allowable Stress Design ...................................................................................................... 14
1.8 Relationship between Elastic Constant ............................................................................. 15
1.9 Worked Examples ............................................................................................................... 18
2 SHEAR FORCE AND BENDING MOMENT ......................................................................... 28
2.1 Beams ................................................................................................................................... 28
2.2 Sign Convention for Shear Force and Bending Moment ................................................ 29
2.3 Relationship between Shear Force and Bending Moment .............................................. 30
2.3.1 Region of Distributed Loads ...................................................................................... 31
2.3.2 Regions of concentrated Loads and Moments .......................................................... 32
2.4 Shear Force and Bending Moment Diagrams .................................................................. 33
2.4.1 Maximum and Minimum Bending Moments ........................................................... 34
2.4.2 Rules for Shear Force and Bending Moment Diagrams.......................................... 34
2.4.3 Worked Examples ....................................................................................................... 36
2.4.4 Graphical Method for Constructing Shear and Moment Diagrams ...................... 47
3 STRESSES IN BEAMS .............................................................................................................. 53
3.1 Theory of Pure Bending ..................................................................................................... 53
3.2 Flexure Stress ...................................................................................................................... 54
3.3 Shear Stress ......................................................................................................................... 61
3.4 Worked Examples ............................................................................................................... 65
4 TORSION .................................................................................................................................... 87

GIRUM MINDAYE I
Strength of Materials AASTU

4.1 Theory of Pure Torsion ...................................................................................................... 87


4.2 Torsional Moment of Resistance ....................................................................................... 88
4.3 Shearing Strains .................................................................................................................. 89
4.4 Shear Stresses ...................................................................................................................... 90
4.5 Strength of Shaft ................................................................................................................. 92
4.6 Design of Transmission Shafts ........................................................................................... 94
4.7 Worked Examples ............................................................................................................... 95
5 COMPOUND (COMBINED) STRESSES .............................................................................. 101
5.1 Combined Stresses ............................................................................................................ 101
5.2 Plane Stress ........................................................................................................................ 101
5.3 Plane Stress Transformation............................................................................................ 102
5.4 Principle Stresses Mohr’s Circle. .................................................................................... 104
5.5 Worked Example .............................................................................................................. 108
6 DEFLECTION IN BEAMS...................................................................................................... 117
6.1 Introduction ....................................................................................................................... 117
6.2 The Double Integration Method ...................................................................................... 118
6.3 Worked Examples ............................................................................................................. 120
7 COMPRESSION MEMBER ................................................................................................... 126
7.1 Introduction ....................................................................................................................... 126
7.2 Buckling and Stability ...................................................................................................... 126
7.3 Euler's Theory of Column Buckling................................................................................ 128
7.4 Effective Length and Slenderness Ratio ......................................................................... 133
7.5 Rankine’s formula for Column ........................................................................................ 135
7.6 Worked Examples ............................................................................................................. 137
Reference ........................................................................................................................................... 145

GIRUM MINDAYE II
Strength of Materials AASTU

1 MECHANICAL PROPERTIES OF MATERIALS


1.1 Introduction

Mechanics of materials (strength of materials) is a branch of applied mechanics that deals with
the behavior of solid bodies subjected to various types of loading. The principal objective of
mechanics of materials is to determine the stresses, strains, and displacements in structures and
their components due to the loads acting on them.
An understanding of mechanical behavior is essential for the safe design of all types of
structures, whether airplanes and antennas, buildings and bridges, machines and motors, or
ships and spacecraft. That is why mechanics of materials is a basic subject in so many
engineering fields. Statics and dynamics are also essential, but those subjects deal primarily
with the forces and motions associated with particles and rigid bodies. In mechanics of
materials, we go one-step further by examining the stresses and strains inside real bodies, that
is, bodies of finite dimensions that deform under loads. To determine the stresses and strains,
we use the physical properties of the materials as well as numerous theoretical laws and
concepts.
Resistance to various applied forces is referred as mechanical properties. Some of these
properties are briefly presented below:
1. Properties related to axial loading: For designing structural elements to resist the applied
loads, the following strength properties should be known:
(a) Tensile strength: If an element is subjected to pulling forces, the resistance developed
per unit area is termed as tensile stress and the maximum tensile stress the material can
resist is termed as tensile strength.
(b) Compressive stress: Instead of pull, if push ‘P’ acts on the element, it tries to shorten
the bar and the internal resistance developed per unit area is called compressive stress.
2. Properties related to shear loading: If the applied force is trying to shear off a particular
section of the element, the resistance developed for unit area in such case along that section
is called shearing stress.
3. Properties related to torsional moment: A member is said to be in torsion when it is
subjected to a moment about its axis. The effect of a torsional member is to twist it and
hence a torsional moment is also called as a twisting moment.
4. Properties related to bending: When a member is supported at two or more points and
subjected to transverse load it bends and develops resistance to the load. The cross-sections
of the members are subjected to bending moment and shear force. Finally the load gets
transferred to the support by end shear. The shear force introduces shear stresses in the
material while bending moment introduces tension in some parts and compression in other
parts.
5. Elasticity: It is the property of the material by virtue of which it regains its original shape
and size after the removal of external load. The maximum stress level before which if the
load is removed the material regains its shape and size fully is called its elastic limit.

GIRUM MINDAYE 1
Strength of Materials AASTU

6. Plasticity: It is the property of the material to retain its changed shape and size after the
loads are removed. It is a required property when a material is to be moulded into different
shape.
7. Creep: It is the property of the material by virtue of which it undergoes changes in size with
time under the action of constant load. Concrete possesses this property.
8. Toughness: It is the property of a material whereby it absorbs energy due to straining
actions by undergoing plastic deformation
1.2 Stress-Strain

The most fundamental concepts in mechanics of materials are stress and strain. These
concepts can be illustrated in their most elementary form by considering a prismatic bar
subjected to axial forces. A prismatic bar is a straight structural member having the same
cross section throughout its length, and an axial force is a load directed along the axis of the
member, resulting in either tension or compression in the bar.

Figure 1-1: Prismatic bar in tension: Prismatic bar in tension: (a) free-body diagram of a
segment of the bar, (b) segment of the bar before loading, (c) segment of the bar after
loading, and (d) normal stresses in the bar

The letter L denotes the original length of the bar, and the Greek letter δ denotes the increase
in length due to the loads δ (delta). The internal actions in the bar are exposed if we make an
imaginary cut through the bar at section mn (Figure 1-1 c). Because this section is taken
perpendicular to the longitudinal axis of the bar, it is called a cross section.

We now isolate the part of the bar to the left of cross section mn as a free body (Figure 1-1 d).
At the right-hand end of this free body (section mn) we show the action of the removed part of
the bar (that is, the part to the right of section mn) upon the part that remains. This action
consists of continuously distributed stresses acting over the entire cross section, and the axial
force P acting at the cross section is the resultant of those stresses. (The resultant force is shown
with a dashed line in Figure 1-1 d.)
1.2.1 Normal Stress

Stress has units of force per unit area and is denoted by the Greek letter σ (sigma). In general,
the stresses s acting on a plane surface may be uniform throughout the area or may vary in
intensity from one point to another. Let us assume that the stresses acting on cross-section mn
(Figure 1-1 d) are uniformly distributed over the area. Then the resultant of those stresses must

GIRUM MINDAYE 2
Strength of Materials AASTU

be equal to the magnitude of the stress times the cross-sectional area A of the bar, that is, P=σA.
Therefore, we obtain the following expression for the magnitude of the stresses:
𝑃
𝜎=
𝐴
This equation gives the intensity of uniform stress in an axially loaded, prismatic bar of
arbitrary cross-sectional shape.
When the forces P stretch the bar, the stresses are tensile stresses; if the forces are reversed in
direction, causing the bar to be compressed, we obtain compressive stresses. Inasmuch as the
stresses act in a direction perpendicular to the cut surface, they are called normal stresses. Thus,
normal stresses may be either tensile or compressive.
When a sign convention for normal stresses is required, it is customary to define tensile stresses
as positive and compressive stresses as negative. Because the normal stress σ is obtained by
dividing the axial force by the cross-sectional area, it has units of force per unit of area.
1.2.2 Normal Strain

As already observed, a straight bar will change in length when loaded axially, becoming longer
when in tension and shorter when in compression. For instance, consider again the prismatic
bar of Figure 1-1. The elongation δ of this bar (Figure 1-1 c) is the cumulative result of the
stretching of all elements of the material throughout the volume of the bar. Let us assume that
the material is the same everywhere in the bar. Then, if we consider half of the bar (length L/2),
it will have an elongation equal to δ/2, and if we consider one-fourth of the bar, it will have an
elongation equal to δ/4.
In general, the elongation of a segment is equal to its length divided by the total length L and
multiplied by the total elongation δ. Therefore, a unit length of the bar will have an elongation
equal to 1/L times δ. This quantity is called the elongation per unit length, or strain, and is
denoted by the Greek letter ϵ (epsilon). We see that strain is given by the equation
𝛿
𝜖=
𝐿
If the bar is in tension, the strain is called a tensile strain, representing an elongation or
stretching of the material. If the bar is in compression, the strain is a compressive strain and
the bar shortens. Tensile strain is usually taken as positive and compressive strain as negative.
The strain ϵ is called a normal strain because it is associated with normal stresses.
Because normal strain is the ratio of two lengths, it is a dimensionless quantity, that is, it has
no units. Therefore, strain is expressed simply as a number, independent of any system of units.
Numerical values of strain are usually very small, because bars made of structural materials
undergo only small changes in length when loaded.
1.2.3 Normal Stress-Strain Diagram

Structural steel is one of the most widely used metals and is found in buildings, bridges, cranes,
ships, towers, vehicles, and many other types of construction. A stress-strain diagram for a
typical structural steel in tension is shown in Figure 1-2. Strains are plotted on the horizontal

GIRUM MINDAYE 3
Strength of Materials AASTU

axis and stresses on the vertical axis. (In order to display all of the important features of this
material, the strain axis in Figure 1-2 is not drawn to scale.)

Figure 1-2: Stress-strain diagram for a typical structural steel in tension


Elastic Behavior: The initial region of the curve, indicated in light orange, is referred to as the
elastic region. Here the curve is a straight line up to the point where the stress reaches the
proportional limit, σpl. When the stress slightly exceeds this value, the curve bends until the
stress reaches an elastic limit. For most materials, these points are very close, and therefore it
becomes rather difficult to distinguish their exact values. What makes the elastic region unique,
however, is that after reaching σY, if the load is removed, the specimen will recover its original
shape. In other words, no damage will be done to the material.
Because the curve is a straight line up to σpl, any increase in stress will cause a proportional
increase in strain. This fact was discovered in 1676 by Robert Hooke, using springs, and is
known as Hooke’s law. It is expressed mathematically as
𝝈 = 𝑬𝝐
Here E represents the constant of proportionality, which is called the modulus of elasticity or
Young’s modulus, named after Thomas Young, who published an account of it in 1807.
As noted in Figure 1-2, the modulus of elasticity represents the slope of the straight line portion
of the curve. Since strain is dimensionless, E will have the same units as stress, such as Pascal
(Pa), Mega-Pascal (MPa), or Giga-Pascal (GPa).
Yielding: A slight increase in stress above the elastic limit will result in a breakdown of the
material and cause it to deform permanently. This behavior is called yielding, and it is indicated
by the rectangular dark orange region in Figure 1-2. The stress that causes yielding is called
the yield stress or yield point, σY, and the deformation that occurs is called plastic deformation.

GIRUM MINDAYE 4
Strength of Materials AASTU

Although not shown in Figure 1-2, for low-carbon steels or those that are hot rolled, the yield
point is often distinguished by two values. The upper yield point occurs first, followed by a
sudden decrease in load-carrying capacity to a lower yield point. Once the yield point is
reached, then as shown in Figure 1-2, the specimen will continue to elongate (strain) without
any increase in load. When the material behaves in this manner, it is often referred to as being
perfectly plastic.
Strain Hardening: When yielding has ended, any load causing an increase in stress will be
supported by the specimen, resulting in a curve that rises continuously but becomes flatter until
it reaches a maximum stress referred to as the ultimate stress, σu. The rise in the curve in this
manner is called strain hardening, and it is identified in Figure 1-2 as the region in light green.
Necking: Up to the ultimate stress, as the specimen elongates, its cross-sectional area will
decrease in a fairly uniform manner over the specimen’s entire gage length. However, just after
reaching the ultimate stress, the cross-sectional area will then begin to decrease in a localized
region of the specimen, and so it is here where the stress begins to increase. As a result, a
constriction or “neck” tends to form with further elongation, Figure 1-3a. This region of the
curve due to necking is indicated in dark green in Fig. 3–4. Here the stress–strain diagram tends
to curve downward until the specimen breaks at the fracture stress, σf, Figure 1-3b.

Figure 1-3: Necking of a mild-steel bar in tension


True Stress–Strain Diagram: Instead of always using the original cross-sectional area A0 and
specimen length L0 to calculate the normal stress and strain, we could have used the actual
cross-sectional area A and specimen length L at the instant the load is measured. The values of
stress and strain found from these measurements are called true stress and true strain, and a
plot of their values is called the true stress–strain diagram. When this diagram is plotted, it has
a form shown by the upper blue curve in Figure 1-2. Note that the conventional and true σ-ϵ
diagrams are practically coincident when the strain is small. The differences begin to appear in
the strain-hardening range, where the magnitude of strain becomes more significant. From the
conventional σ-ϵ diagram, the specimen appears to support a decreasing stress (or load), since
A0 is constant, σ = P/A0. In fact, the true σ-ϵ diagram shows the area A within the necking
region is always decreasing until fracture, σf’, and so the material actually sustains increasing
stress, since σf' = N/A.
Because most structures are expected to function at stresses below the proportional limit, the
conventional stress-strain curve, which is based upon the original cross-sectional area of the
specimen and is easy to determine, provides satisfactory information for use in engineering
design.
Ductile Materials: Materials can be classified as either being ductile or brittle, depending on
their stress–strain characteristics. Any material that can be subjected to large strains before it

GIRUM MINDAYE 5
Strength of Materials AASTU

fractures is called a ductile material. Mild steel is a typical example. Engineers often choose
ductile materials for design because these materials are capable of absorbing shock or energy,
and if they become overloaded, they will usually exhibit large deformation before failing. One
way to specify the ductility of a material is to report its percent elongation or percent reduction
in area at the time of fracture.
The percent elongation is defined as follows:
𝐿𝑓 − 𝐿0
𝑃𝑒𝑟𝑐𝑒𝑛𝑡 𝑒𝑙𝑜𝑛𝑔𝑎𝑡𝑖𝑜𝑛 = (100)
𝐿0
Where, L0 – original length, Lf– length at fracture.
The percent reduction in area measures the amount of necking that occurs and is defined as
follows:
𝐴0 − 𝐴𝑓
𝑃𝑒𝑟𝑐𝑒𝑛𝑡 𝑟𝑒𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑖𝑛 𝑎𝑟𝑒𝑎 = (100)
𝐴0
in which A0 is the original cross-sectional area and Af is the final area at the fracture section.
For ductile steels, the reduction is about 50%.
1.2.4 Changes in Lengths under Non-uniform Conditions

Consider a homogeneous rod of length L and uniform cross section of area A subjected to a
centric axial load P. If the resulting axial stress σ=P/A does not exceed the proportional limit
of the material, Hooke’s law applies and
𝜎 = 𝐸𝜖
𝜎 𝑃 𝛿
𝑓𝑟𝑜𝑚 𝑤ℎ𝑖𝑐ℎ, 𝜖= = 𝑎𝑛𝑑 𝜖 =
𝐸 𝐴𝐸 𝐿
𝑃𝐿
𝛿=
𝐴𝐸

This equation can be used only if the rod is homogeneous (constant E), has a uniform cross
section of area A, and is loaded at its ends. If the rod is loaded at other points, or consists of
several portions of various cross sections and possibly of different materials, it must be divided
into component parts that satisfy the required conditions for the application of δ=PL/AE. Using
the internal force Pi, length Li, cross sectional area Ai, and modulus of elasticity Ei,
corresponding to part i, the deformation of the entire rod is
𝑃𝑖 𝐿𝑖
𝛿=∑
𝐴𝑖 𝐸𝑖
𝑖

1.3 Poisson’s Ratio

When a prismatic bar is loaded in tension, the axial elongation is accompanied by lateral
contraction (that is, contraction normal to the direction of the applied load). Lateral contraction
is easily seen by stretching a rubber band, but in metals, the changes in lateral dimensions (in
the linearly elastic region) are usually too small to be visible. However, they can be detected
with sensitive measuring devices.

GIRUM MINDAYE 6
Strength of Materials AASTU

When a homogeneous slender bar is axially loaded, the resulting stress and strain satisfy
Hooke’s law, as long as the elastic limit of the material is not exceeded. Assuming that the load
P is directed along the x axis (Figure 1-4a), σx=P/A, where A is the cross-sectional area of the
bar, and from Hooke’s law,
𝜎𝑥
𝜖𝑥 =
𝐸
Where E is the modulus of elasticity of the material.

Figure 1-4: (a) A bar in uniaxial tension (b) representative stress element and (c) transverse
contraction
Also, the normal stresses on faces perpendicular to the y and z axes are zero: σy= σz=0 (Figure
1-4b). It would be tempting to conclude that the corresponding strains ϵy and ϵz are also zero.
This is not the case. In all engineering materials, the elongation produced by an axial tensile
force P in the direction of the force is accompanied by a contraction in any transverse direction
(Figure 1-4c).
In this section and the following sections, all materials are assumed to be both homogeneous
and isotropic (i.e., their mechanical properties are independent of both position and direction).
It follows that the strain must have the same value for any transverse direction. Therefore, the
loading shown in Figure 1-4 must have ϵy = ϵz. This common value is the lateral strain.
The lateral strain ϵ’ at any point in a bar is proportional to the axial strain ϵ at that same point
if the material is linearly elastic. The ratio of these strains is a property of the material known
as Poisson’s ratio, named after the French mathematician Siméon Denis Poisson (1781–1840)
and denoted by the Greek letter ν (nu), can be expressed by the equation
𝑙𝑎𝑡𝑒𝑟𝑎𝑙 𝑠𝑡𝑟𝑎𝑖𝑛 𝜖′
𝜈=− =−
𝑎𝑥𝑖𝑎𝑙 𝑠𝑡𝑟𝑎𝑖𝑛 𝜖
For the loading condition represented in Figure 1-4,
𝜖𝑦 𝜖𝑧
𝜈=− =−
𝜖𝑥 𝜖𝑥
Note the use of a minus sign in these equations to obtain a positive value for ν, as the axial and
lateral strains have opposite signs for all engineering materials. For instance, the axial strain in
a bar in tension is positive and the lateral strain is negative (because the width of the bar
decreases). For compression we have the opposite situation, with the bar becoming shorter
(negative axial strain) and wider (positive lateral strain).
Solving ϵy and ϵz, which fully describe the condition of strain under an axial load applied in a
direction parallel to the x-axis:

GIRUM MINDAYE 7
Strength of Materials AASTU

𝜎𝑥 𝑣𝜎𝑥
𝜖𝑥 = , 𝜖𝑦 = 𝜖𝑧 = −
𝐸 𝐸

Consider an element of an isotropic material in the shape of a cube (Figure 1-5a). Assume the
side of the cube to be equal to unity, since it is always possible to select the side of the cube as
a unit of length. Under the given multiaxial loading, the element will deform into a rectangular
parallelepiped of sides equal to 1+ϵx, 1+ϵy, and 1+ϵz, where ϵx, ϵy, and ϵz, denote the values of
the normal strain in the directions of the three coordinate axes (Figure 1-5b). Note that, as a
result of the deformations of the other elements of the material, the element under consideration
could also undergo a translation, but the concern here is with the actual deformation of the
element, not with any possible superimposed rigid-body displacement.

Figure 1-5: Deformation of unit cube under multiaxial loading: (a) unloaded; (b) deformed.

In order to express the strain components ϵx, ϵy, ϵz in terms of the stress components σx, σy, σz,
consider the effect of each stress component and combine the results. Considering the effect of
the stress component σx, i.e., σx causes a strain equal to σx/E in the x direction and strains equal
to -νσx/E in each of the y and z directions. Similarly, the stress component σy, if applied
separately, will cause a strain σy/E in the y directions and strains -νσy/E in the other two
directions. Finally, the stress component σz causes a strain σy/E in the z direction and strains -
νσz in the x and y directions.

For the multiaxial loading condition represented in Figure 1-5b, and as long as none of the
stresses involved exceeds the corresponding proportional limit, you can apply the principle of
superposition and combine the results. The generalized Hooke’s law for a homogeneous
isotropic material under the most general stress condition is

GIRUM MINDAYE 8
Strength of Materials AASTU

𝜎𝑥 𝑣𝜎𝑦 𝑣𝜎𝑧 𝑣𝜎𝑥 𝜎𝑦 𝑣𝜎𝑧 𝑣𝜎𝑥 𝑣𝜎𝑦 𝜎𝑧


𝜖𝑥 = − − , 𝜖𝑦 = − + − , 𝜖𝑧 = − − +
𝐸 𝐸 𝐸 𝐸 𝐸 𝐸 𝐸 𝐸 𝐸
1.4 Bulk Modulus

When a member is subjected to stresses, it undergoes deformation in all directions. Hence,


there will be change in volume. The ratio of the change in volume to original volume is called
volumetric strain. It is also knows as dilation.
𝑐ℎ𝑎𝑛𝑔𝑒 𝑖𝑛 𝑣𝑜𝑙𝑢𝑚𝑒 𝛿𝑉
𝑉𝑜𝑙𝑢𝑚𝑒𝑡𝑟𝑖𝑐 𝑠𝑡𝑟𝑎𝑖𝑛 (𝜖𝑣 ) = =
𝑜𝑟𝑖𝑔𝑖𝑛𝑎𝑙 𝑣𝑜𝑙𝑢𝑚𝑒 𝑉
It can be shown that volumetric strain is sum of strains in three mutually perpendicular
directions, .i.e., ϵv = ϵx + ϵy+ ϵz

𝜎𝑥 + 𝜎𝑦 + 𝜎𝑧 2𝑣(𝜎𝑥 + 𝜎𝑦 + 𝜎𝑧 ) 1 − 2𝑣
𝜖𝑣 = − = (𝜎𝑥 + 𝜎𝑦 + 𝜎𝑧 )
𝐸 𝐸 𝐸
When a body is subjected to three mutually perpendicular stresses of equal intensity, the ratio
of direct stress to the corresponding volumetric strain is known as bulk modulus. The bulk
modulus is an extension of Young's modulus to three dimensions.
𝑣𝑜𝑙𝑢𝑚𝑒𝑡𝑟𝑖𝑐 𝑠𝑡𝑟𝑒𝑠𝑠
𝐵𝑢𝑙𝑘 𝑀𝑜𝑑𝑢𝑙𝑢𝑠 (𝐾) =
𝑣𝑜𝑙𝑢𝑚𝑒𝑡𝑟𝑖𝑐 𝑠𝑡𝑟𝑎𝑖𝑛
𝜎𝑥 = 𝜎𝑦 = 𝜎𝑧
1
𝑣𝑜𝑙𝑢𝑚𝑒𝑡𝑟𝑖𝑐 𝑠𝑡𝑟𝑒𝑠𝑠 = 𝑎𝑣𝑒𝑟𝑎𝑔𝑒 𝑠𝑡𝑟𝑒𝑠𝑠 = (𝜎 + 𝜎𝑦 + 𝜎𝑧 ) = 𝜎
3 𝑥
1 − 2𝑣 3(1 − 2𝑣)
𝑣𝑜𝑙𝑢𝑚𝑒𝑡𝑟𝑖𝑐 𝑠𝑡𝑟𝑎𝑖𝑛 = (𝜎𝑥 + 𝜎𝑦 + 𝜎𝑧 ) = 𝜎
𝐸 𝐸
𝜎
𝐾=
3(1 − 2𝑣)
𝜎
𝐸
𝑬
𝑲=
𝟑(𝟏 − 𝟐𝒗)
1.5 Shear Stress-Strain

1.5.1 Shear Stress

The intensity of force acting tangent to A is called the shear stress, τ (tau). To show how this
stress can develop, consider the effect of applying a force F to the bar in Figure 1-6a.

GIRUM MINDAYE 9
Strength of Materials AASTU

Figure 1-6
If F is large enough, it can cause the material of the bar to deform and fail along the planes
identified by AB and CD. A free-body diagram of the unsupported center segment of the bar,
Figure 1-6b, indicates that the shear force V = F/2 must be applied at each section to hold the
segment in equilibrium. The shear stress distributed over each sectioned area that develops this
shear force is defined by
𝑉
𝜏=
𝐴
Where,
τ= shear stress at the section, which is assumed to be the same at each point on the section
V = internal resultant shear force on the section determined from the equations of equilibrium
A = area of the section
Notice that τ is in the same direction as V, since the shear stress must create associated forces,
all of which contribute to the internal resultant force V. The loading case discussed here is an
example of simple or direct shear, since the shear is caused by the direct action of the applied
load F.
A shear stress acting on a positive face of an element is positive if it acts in the positive direction
of one of the coordinate axes and negative if it acts in the negative direction of an axis. A shear
stress acting on a negative face of an element is positive if it acts in the negative direction of
an axis and negative if it acts in a positive direction.
1.5.2 Shear Strain

Deformations not only cause line segments to elongate or contract, but they also cause them to
change direction. If we select two line segments that are originally perpendicular to one
another, then the change in angle that occurs between them is referred to as shear strain. This
angle is denoted by γ (gamma) and is always measured in radians (rad), which are
dimensionless.

GIRUM MINDAYE 10
Strength of Materials AASTU

Figure 1-7
For example, consider the two perpendicular line segments at a point in the block shown in
Figure 1-7a. If an applied loading causes the block to deform as shown in Figure 1-7b, so that
the angle between the line segments becomes θ, then the shear strain at the point becomes
𝜋
𝛾= −𝜃
2
Notice that if γ is smaller than π/2, Figure 1-7c, then the shear strain is positive, whereas if γ
is larger than π/2, then the shear strain is negative.
1.5.3 Shear Stress-Strain Relationship

When a small element of material is subjected to pure shear, equilibrium requires that equal
shear stresses must be developed on four faces of the element, Figure 1-8a. Furthermore, if the
material is homogeneous and isotropic, then this shear stress will distort the element uniformly,
Figure 1-8b, producing shear strain.

Figure 1-8: Small element (a) undeformed (b) deformed


In order to study the behavior of a material subjected to pure shear, engineers use a specimen
in the shape of a thin tube and subject it to a torsional loading. If measurements are made of
the applied torque and the resulting angle of twist, then by the methods to be explained in
Section 4, the data can be used to determine the shear stress and shear strain within the tube.
Plotting successive values of τ against the corresponding values of γ, the shearing stress-strain
diagram is obtained for the material as shown in Figure 1-9. This diagram is similar to the
normal stress-strain diagram from the tensile test described earlier; however, the values for the

GIRUM MINDAYE 11
Strength of Materials AASTU

yield strength, ultimate strength, etc., are about half as large in shear as they are in tension. As
for normal stresses and strains, the initial portion of the shearing stress-strain diagram is a
straight line and it will have a defined proportional limit τpl. Also, strain hardening will occur
until an ultimate shear stress τu is reached. And finally, the material will begin to lose its shear
strength until it reaches a point where it fractures, τf.

Figure 1-9: shear stress–strain diagram


For most engineering materials, like the one just described, the elastic behavior is linear, and
so Hooke’s law for shear can be written as
𝜏 = 𝐺𝛾
Here G is called the shear modulus of elasticity or the modulus of rigidity. Its value represents
the slope of the line on the τ–γ diagram, that is, G = τpl/γpl. Units of measurement for G will be
the same as those for τ (Pa), since γ is measured in radians, a dimensionless quantity.
1.6 General State of Stress

To visualize the stress condition at point Q, consider a small cube of side a centered at Q and
the stresses exerted on each of the six faces of the cube (Figure 1-10c). The stress components
shown are σx, σy and σz, which represent the normal stress on faces respectively perpendicular
to the x, y, and z axes, and the six shearing stress components τxy, τxz, etc. Recall that τxy
represents the y component of the shearing stress exerted on the face perpendicular to the x-
axis, while τyx represents the x component of the shearing stress exerted on the face
perpendicular to the y-axis. Note that only three faces of the cube are actually visible in Figure
1-10 and that equal and opposite stress components act on the hidden faces. While the stresses
acting on the faces of the cube differ slightly from the stresses at Q, the error involved is small
and vanishes as side a of the cube approaches zero.

GIRUM MINDAYE 12
Strength of Materials AASTU

Figure 1-10: Positive stress components at point Q for a general state of stress.
Consider the free-body diagram of the small cube centered at point Q (Figure 1-11). The normal
and shearing forces acting on the various faces of the cube are obtained by multiplying the
corresponding stress components by the area ΔA of each face. First, write the following three
equilibrium equations

∑ 𝐹𝑥 = 0, ∑ 𝐹𝑦 = 0, ∑ 𝐹𝑧 = 0

Figure 1-11: Positive resultant forces on a small element at point Q resulting from a state of
general stress.

Since forces equal and opposite to the forces actually shown in Figure 1-11 are acting on the
hidden faces of the cube, ∑F=0 are satisfied. Considering the moments of the forces about axes
x’, y’, and z’ drawn from Q in directions respectively parallel to the x, y, and z axes, the three
additional equations are
∑ 𝑀𝑥 = 0, ∑ 𝑀𝑦 = 0, ∑ 𝑀𝑧 = 0

GIRUM MINDAYE 13
Strength of Materials AASTU

Using a projection on the x’y’ plane (Figure 1-12), note that the only forces with moments
about the z-axis different from zero are the shearing forces. These forces form two couples: a
counterclockwise (positive) moment (τxyΔA)a and a clockwise (negative) moment -(τyxΔA)a.
∑ 𝑀𝑧 = 0, (𝜏𝑥𝑦 𝛥𝐴)𝑎 − (𝜏𝑦𝑥 𝛥𝐴)𝑎 ⟹ 𝜏𝑥𝑦 = 𝜏𝑦𝑥

Figure 1-12: Free-body diagram of small element at Q viewed on projected plane


perpendicular to z’-axis. Resultant forces on positive and negative z’ faces (not shown) act
through the z’-axis, thus do not contribute to the moment about that axis.

This relationship shows that the y component of the shearing stress exerted on a face
perpendicular to the x-axis is equal to the x component of the shearing stress exerted on a face
perpendicular to the y-axis.
𝑆𝑖𝑚𝑖𝑙𝑎𝑟𝑙𝑦 𝑓𝑜𝑟 ∑ 𝑀𝑥 = 0, ⟹ 𝜏𝑦𝑧 = 𝜏𝑧𝑦 𝑎𝑛𝑑 ∑ 𝑀𝑦 = 0, ⟹ 𝜏𝑥𝑧 = 𝜏𝑧𝑥
We conclude from this, only six stress components are required to define the condition of stress
at a given point Q, instead of nine as originally assumed. These components are σx, σy, σz, τxy,
τyz, and τzx.
For values of the shearing stress that do not exceed the proportional limit in shear, it can be
written for any homogeneous isotropic material that
𝜏𝑥𝑦 𝜏𝑦𝑧 𝜏𝑧𝑥
𝛾𝑥𝑦 = , 𝛾𝑦𝑧 = , 𝛾𝑧𝑥 =
𝐺 𝐺 𝐺
This relationship is Hooke’s law for shearing stress and strain, and the constant G is called the
modulus of rigidity or shear modulus of the material.
1.7 Allowable Stress Design

To ensure the safety of a structural or mechanical member, it is necessary to restrict the applied
load to one that is less than the load the member can fully support. There are many reasons for
doing this.
 The intended measurements of a structure or machine may not be exact, due to
errors in fabrication or in the assembly of its component parts.
 Unknown vibrations, impact, or accidental loadings can occur that may not be
accounted for in the design.

GIRUM MINDAYE 14
Strength of Materials AASTU

 Atmospheric corrosion, decay, or weathering tend to cause materials to deteriorate


during service.
 Some materials, such as wood, concrete, or fiber-reinforced composites, can show
high variability in mechanical properties.
One method of specifying the allowable load for a member is to use a number called the factor
of safety (F.S.). It is a ratio of the failure load Ffail to the allowable load Fallow,
𝐹𝑓𝑎𝑖𝑙
𝐹. 𝑆. =
𝐹𝑎𝑙𝑙𝑜𝑤
Here Ffail is found from experimental testing of the material.
If the load applied to the member is linearly related to the stress developed within the member,
as in the case of σ = N/A and τavg = V/A, then we can also express the factor of safety as a ratio
of the failure stress σfail (or τfail) to the allowable stress σallow (or τallow). Here the area A will
cancel, and so,
𝜎𝑓𝑎𝑖𝑙 𝜏𝑓𝑎𝑖𝑙
𝐹. 𝑆. = , 𝑜𝑟 𝐹. 𝑆. =
𝜎𝑎𝑙𝑙𝑜𝑤 𝜏𝑎𝑙𝑙𝑜𝑤
Specific values of F.S. depend on the types of materials to be used and the intended purpose of
the structure or machine, while accounting for the previously mentioned uncertainties. For
example, the F.S. used in the design of aircraft or space vehicle components may be close to 1
in order to reduce the weight of the vehicle. Or, in the case of a nuclear power plant, the factor
of safety for some of its components may be as high as 3 due to uncertainties in loading or
material behavior. Whatever the case, the factor of safety or the allowable stress for a specific
case can be found in design codes and engineering handbooks. Design that is based on an
allowable stress limit is called allowable stress design (ASD). Using this method will ensure a
balance between both public and environmental safety on the one hand and economic
considerations on the other.
1.8 Relationship between Elastic Constant

Young’s modulus (E), bulk modulus (K) and Rigidity modulus (G) of an elastic solid are
together called Elastic constants. When a deforming force is acting on a solid, it results in the
change in its original dimension. In such cases, we can use the relation between elastic
constants to understand the magnitude of deformation.

Consider a cubic element (Figure 1-13) subjected to only the shearing stresses τxy and τyx
applied to faces of the element respectively perpendicular to the x and y-axes. (Recall from
Sec. 1.8 that τxy= τyx=τ)

GIRUM MINDAYE 15
Strength of Materials AASTU

Figure 1-13: cubic element subjected to shearing stress.


Fig. (a) shows a face of cubic element on xy plane, which is a square element of side ‘a’. When
this face element deformed (see Fig.(b)), the length of diagonal DB increases, indicating it is
subjected to tensile stress and the length of diagonal AC decreases indicating that compressive
stress as).

Now consider the section, ADC of the element, Fig. (c). Multiplying the two shear stresses by
A0 and the normal stress by A to change into forces as shown in figure below.

Where, A0 is the area of the face, which is equal to a2 and A is the area of the diagonal cubic
element and equal to a2√2
Observing that the components along the diagonal AC of the two shear forces are equal and
opposite, the force F exerted on AC must be perpendicular to that face and is a tensile force. Its
magnitude is

∑ 𝐹𝑋 , 𝐹 = 2𝜏𝐴0 cos 450 = 𝜏𝐴0 √2

GIRUM MINDAYE 16
Strength of Materials AASTU

The corresponding stress is obtained by dividing the force F by the area A of face BD.
observing that 𝐴 = 𝐴0 √2

𝐹 𝜏𝐴0 √2
𝜎𝑛 = = =𝜏
𝐴 𝐴0 √2
Therefore, normal tensile stress developed on plane AC is numerically equal to shear stress.
Similarly, it can be proved that the compressive stress developed on plane BD is numerically
equal to shear stress
𝜏 = 𝜎 = 𝐸𝜖
Assuming that the strains are small and the angle BB’E may be taken as 450. Since angle
between DB and DB’ is very small hence EB’, is the change in the length of the diagonal DB.
𝑎
𝐷𝐵 = = 𝑎√2
sin 450

𝐵𝐵 ′ = 𝑎𝛾, 𝑤ℎ𝑒𝑟𝑒 𝛾 = 𝑠ℎ𝑒𝑎𝑟 𝑠𝑡𝑟𝑎𝑖𝑛


𝑎𝛾
𝐸𝐵 ′ = 𝐵𝐵 ′ cos 450 =
√2
𝑎𝛾
𝐷𝐵 ′ − 𝐷𝐵 𝐸𝐵′ 𝛾
𝑇ℎ𝑒 𝑠𝑡𝑟𝑎𝑖𝑛 𝑎𝑙𝑜𝑛𝑔 𝑡ℎ𝑒 𝑑𝑖𝑎𝑔𝑜𝑛𝑎𝑙 𝐵𝐷, 𝜖= = = √2 =
𝐷𝐵 𝐷𝐵 𝑎√2 2
𝜏 𝛾 𝜏
𝑓𝑜𝑟 𝛾 = ⟹𝜖= =
𝐺 2 2𝐺
Now this shear stress system is equivalent or can be replaced by a system of direct stresses at
450 as shown below. One set will be compressive, the other tensile, and both will be equal in
value to the applied shear strain.

Thus, for the direct state of stress system which applies along the diagonals:

𝜎1 𝜎2 𝜏 (−𝜏) 𝜏
𝑇ℎ𝑒 𝑠𝑡𝑟𝑎𝑖𝑛 𝑎𝑙𝑜𝑛𝑔 𝑡ℎ𝑒 𝑑𝑖𝑎𝑔𝑜𝑛𝑎𝑙 𝐵𝐷, 𝜖= −𝑣 = −𝑣 = (1 + 𝑣)
𝐸 𝐸 𝐸 𝐸 𝐸

Equating the two strains one may get


𝜏 𝜏
= (1 + 𝑣)
2𝐺 𝐸
𝑬
𝑮=
𝟐(𝟏 + 𝒗)

GIRUM MINDAYE 17
Strength of Materials AASTU

𝑬
𝐵𝑢𝑙𝑘 𝑀𝑜𝑑𝑢𝑙𝑢𝑠, 𝑲=
𝟑(𝟏 − 𝟐𝒗)

𝐸 1 𝐸
𝐾= →𝑣= −
3(1 − 2𝑣) 2 6𝐾

𝐸 𝐸
𝐺= →𝑣= −1
2(1 + 𝑣) 2𝐺

1 𝐸 𝐸 𝐸 𝐸 3
− = −1→ + =
2 6𝐾 2𝐺 2𝐺 6𝐾 2
𝐸 𝐸 3 𝟏 𝟏 𝟏
+ = → = +
2𝐺 6𝐾 2 𝑬 𝟑𝑮 𝟗𝑲
1.9 Worked Examples
Example 1.1: A circular steel rod of length L and diameter d hangs in a mineshaft and holds
an ore bucket of weight W at its lower end.
a) Obtain a formula for the maximum stress σmax in the rod, taking into account the weight
of the rod itself.
b) Calculate the maximum stress if L =40m, d =8mm, and W=1.5kN.

Figure 1-14: Steel rod supporting a weight W


Solution
(a) The maximum axial force Fmax in the rod occurs at the upper end and is equal to the
weight W of the ore bucket plus the weight Wo of the rod itself. The latter is equal to
the weight density γ of the steel times the volume V of the rod.
𝑊𝑂 = 𝛾𝑉 = 𝛾𝐴𝐿

in which A is the cross-sectional area of the rod. Therefore, the formula for the maximum stress
becomes
W + γAL W
σmax = = + γL
A A
(b) To calculate the maximum stress, we substitute numerical values into the preceding
equation. The cross-sectional area A equals πd2/4, where d =8mm, and the weight
density γ of steel is 77.0kN/m3
1.5kN
σmax = + 77.0 kN⁄𝑚3 × 40m = 32.9MPa
π(8𝑚𝑚)2⁄4

GIRUM MINDAYE 18
Strength of Materials AASTU

Example 1.2: Two solid cylindrical rods AB and BC are welded together at B and loaded as
shown. Knowing that d1=30mm and d2=50mm, find the average normal stress at the midsection
of (a) rod AB, (b) rod BC.

(a) Rod AB
𝑃 = 60𝑘𝑁 (𝑡𝑒𝑛𝑠𝑖𝑜𝑛)
𝑃 60𝑘𝑁
𝜎𝐴𝐵 = = = 84.88𝑀𝑃𝑎
𝐴𝐴𝐵 𝜋(30𝑚𝑚)2 ⁄4
(b) Rod BC
𝑃 = 60 − (2 × 125) = −190𝑘𝑁 (𝑐𝑜𝑚𝑝𝑟𝑒𝑠𝑠𝑖𝑜𝑛)
𝑃 −190𝑘𝑁
𝜎𝐵𝐶 = = = −96.77𝑀𝑃𝑎
𝐴𝐵𝐶 𝜋(50𝑚𝑚)2 ⁄4
Example 1.3: Two solid cylindrical rods AB and BC are welded together at B and loaded
as shown. Knowing that the average normal stress must not exceed 150 MPa in either
rod, determine the smallest allowable values of the diameters.

Solution
(a) Rod AB
𝑃 = 60𝑘𝑁 (𝑡𝑒𝑛𝑠𝑖𝑜𝑛)
𝑃 𝑃 𝑃 60 × 103
𝜎𝐴𝐵 = = ⟹ 𝑑 = 2√ =2 √ = 22.57𝑚𝑚
𝐴𝐴𝐵 𝜋𝑑 2 ⁄4 𝜋𝜎𝐴𝐵 𝜋 × 150
(b) Rod BC
𝑃 = 60 − (2 × 125) = −190𝑘𝑁 (𝑐𝑜𝑚𝑝𝑟𝑒𝑠𝑠𝑖𝑜𝑛)
𝑃 𝑃 𝑃 190 × 103
𝜎𝐵𝐶 = = 2 ⟹ 𝑑 = 2√ = 2√ = 40.16𝑚𝑚
𝐴𝐵𝐶 𝜋𝑑 ⁄4 𝜋𝜎𝐵𝐶 𝜋 × 150

Example 1.4: Two brass rods AB and BC, each of uniform diameter, will be brazed together
at B to form a non-uniform rod of total length 100m that will be suspended from a support
at A as shown. Knowing that the density of brass is 8470 kg/m3, determine (a) the length
of rod AB for which the maximum normal stress in ABC is minimum, (b) the corresponding
value of the maximum normal stress.

GIRUM MINDAYE 19
Strength of Materials AASTU

Solution
Areas:
𝐴𝐴𝐵 = 𝜋𝑑 2 ⁄4 = 𝜋(15)2⁄4 = 176.71𝑚𝑚2 = 176.71 × 10−6 𝑚2
𝐴𝐵𝐶 = 𝜋𝑑2 ⁄4 = 𝜋(10)2 ⁄4 = 78.54𝑚𝑚2 = 78.54 × 10−6 𝑚2
From geometry, b=100-a
Weights:
𝑊𝐴𝐵 = 𝜌𝑔𝐴𝐴𝐵 𝑙𝐴𝐵 = (8470)(9.81)(176.71 × 10−6 )𝑎 = 14.683𝑎
𝑊𝐵𝐶 = 𝜌𝑔𝐴𝐵𝐶 𝑙𝐵𝐶 = (8470)(9.81)(78.54 × 10−6 )(100 − 𝑎) = 652.594 − 6.526𝑎
Normal stress:
𝐴𝑡 𝐴, 𝑃𝐴 = 𝑊𝐴𝐵 + 𝑊𝐵𝐶 = 652.594 + 8.157𝑎
𝑃 652.594+8.157𝑎
𝜎𝐴𝐵 = 𝐴 𝐴 = 176.71×10−6 𝑚2 = 3.693 × 106 + 46.160𝑎 × 103
𝐴𝐵

𝐴𝑡 𝐵, 𝑃𝐵 = 𝑊𝐵𝐶 = 652.594 − 6.526𝑎


𝑃 652.594−6.526𝑎
𝜎𝐵𝑐 = 𝐴 𝐵 = = 8.309 × 106 − 83.091𝑎 × 103
𝐴𝐵 78.54×10−6 𝑚2

(a) Length of rod AB , The maximum normal stress in ABC is minimum when σAB=σBC
3.693 × 106 + 46.160𝑎 × 103 = 8.309 × 106 − 83.091𝑎 × 103 ⟹ 𝑎 = 𝑙𝐴𝐵
= 35.71𝑚
(b) Maximum normal stress
𝜎𝐴𝐵 = 𝜎𝐵𝑐 = 3.693 × 106 + 46.160(35.71) × 103 = 5,341,373.6𝑃𝑎 = 5.34𝑀𝑃𝑎

Example 1.5: Determine the deformation of the steel rod as shown under the given loads
(E=29×106 psi).

GIRUM MINDAYE 20
Strength of Materials AASTU

The rod is divided into three component parts in figure b, so

𝐿1 = 𝐿2 = 12 𝑖𝑛, 𝐿3 = 16 𝑖𝑛

𝐴1 = 𝐴2 = 0.9 𝑖𝑛2 , 𝐴3 = 0.3 𝑖𝑛2

To find the internal forces P1, P2, and P3, pass sections through each of the component parts,
drawing each time the free-body diagram of the portion of rod located to the right of the section
(figure c). Each of the free bodies is in equilibrium; thus

𝑃1 = 60𝑘𝑖𝑝𝑠 = 60 × 103 𝑙𝑏

𝑃2 = −15𝑘𝑖𝑝𝑠 = −15 × 103 𝑙𝑏


𝑃3 = 30𝑘𝑖𝑝𝑠 = 30 × 103 𝑙𝑏
𝑃𝑖 𝐿𝑖 1 𝑃1 𝐿1 𝑃2 𝐿2 𝑃3 𝐿3
𝛿=∑ = ( + + )
𝐴𝑖 𝐸𝑖 𝐸 𝐴1 𝐴2 𝐴3
𝑖
(60 × 10 3 )(12) (−15 × 103 )(12) (30 × 103 )(16)
1
𝛿= ( + + ) = 75.9 × 10−3 𝑖𝑛
29 × 106 0.9 0.9 0.3
Example 1.6: The bar has a constant width of 35 mm and a thickness of 10mm. Determine the
maximum average normal stress in the bar when it is subjected to the loading shown.

Solution
By inspection, the internal axial forces in regions AB, BC, and CD are all constant yet have
different magnitudes. Using the method of sections, these loadings are shown on the free-body
diagrams of the left segments as shown in figure below.

GIRUM MINDAYE 21
Strength of Materials AASTU

The largest loading is in region BC, where NBC= 30kN. Since the cross-sectional area of the
bar is constant, the largest average normal stress also occurs within this region of the bar.
𝑁𝐵𝐶 30(103 )𝑁
𝐴𝑣𝑒𝑟𝑎𝑔𝑒 𝑁𝑜𝑟𝑚𝑎𝑙 𝑆𝑡𝑟𝑒𝑠𝑠, 𝜎𝐵𝐶 = = = 85.7𝑀𝑃𝑎
𝐴 (0.035𝑚)(0.010𝑚)
Example 1.7: A 500mm long, 16mm diameter rod made of a homogenous, isotropic material
is observed to increase in length by 300μm, and to decrease in diameter by 2.4μm when
subjected to an axial 12-kN load. Determine the modulus of elasticity and Poisson’s ratio of
the material.

Solution
The cross-sectional area of the rod is
𝐴 = 𝜋𝑟 2 = 𝜋(8 × 10−3 𝑚)2 = 201 × 10−6 𝑚2
Choosing the x-axis along the axis of the rod, write
𝑃 12 × 103 𝑁
𝜎𝑥 = = = 59.7𝑀𝑃𝑎
𝐴 201 × 10−6 𝑚2
𝛿𝑥 300𝜇𝑚
𝜖𝑥 = = = 600 × 10−6
𝐿 500𝑚𝑚
𝛿𝑦 2.4𝜇𝑚
𝜖𝑦 = = = −150 × 10−6
𝑑 16𝑚𝑚
From Hooke’s law, σx=Eϵx
𝜎𝑥 59.7𝑀𝑃𝑎
𝐸= = = 99.5𝐺𝑃𝑎
𝜖𝑥 600 × 10−6
Lateral strain,

GIRUM MINDAYE 22
Strength of Materials AASTU

𝜖𝑦 −150 × 10−6
𝑣=− =− = 0.25
𝜖𝑥 600 × 10−6
Example 1.8: A specimen of steel 20mm diameter with a gauge length of 200mm was tested
to failure. It undergoes an extension of 0.20mm under a load of 60kN. Load at elastic limit is
120kN. The maximum load is 180kN. The breaking load is 160kN. Total extension is 50mm
and the diameter at fracture is 16mm. Find:
a) Stress at elastic limit
b) Young’s modulus
c) % elongation
d) % reduction in area
e) Ultimate strength
f) Nominal breaking stress
g) True breaking stress
Solution:
a) Stress at elastic limit,
𝐿𝑜𝑎𝑑 𝑎𝑡 𝑒𝑙𝑎𝑠𝑡𝑖𝑐 𝑙𝑖𝑚𝑖𝑡 120 × 103
𝜎𝑌 = = = 381.97 𝑁⁄𝑚𝑚2 = 381.97𝑀𝑃𝑎
𝐴𝑜 𝜋 × 202 ⁄4
b) Young’s Modulus, (consider a load which is within the elastic limit)
𝜎 𝑃⁄𝐴 60 × 103 ⁄𝜋 × 202 ⁄4
𝐸= = = = 190,986 𝑁⁄𝑚𝑚2 = 190.986𝐺𝑃𝑎
𝜖 𝛿 ⁄𝐿 0.2⁄200
c) % elongation
𝐿𝑓 − 𝐿𝑜 𝛿𝑓 50
% 𝑒𝑙𝑜𝑛𝑔𝑎𝑡𝑖𝑜𝑛 = × 100 = × 100 = × 100 = 20%
𝐿𝑜 𝐿𝑜 200
d) % reduction in area
𝐴𝑜 − 𝐴𝑓
% 𝑟𝑒𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑖𝑛 𝑎𝑟𝑒𝑎 = × 100
𝐴𝑜

𝜋 × 𝑑𝑜 2 ⁄4 − 𝜋 × 𝑑𝑓 2 ⁄4
% 𝑟𝑒𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑖𝑛 𝑎𝑟𝑒𝑎 = × 100
𝜋 × 𝑑0 2 ⁄4

𝑑0 2 − 𝑑𝑓 2 202 − 162
% 𝑟𝑒𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑖𝑛 𝑎𝑟𝑒𝑎 = × 100 = × 100 = 36%
𝑑0 2 202

e) Ultimate strength
𝑀𝑎𝑥𝑖𝑚𝑢𝑚 𝑙𝑜𝑎𝑑 180 × 103
𝜎𝑢 = = = 572.96 𝑁⁄𝑚𝑚2 = 572.96𝑀𝑃𝑎
𝐴𝑜 𝜋 × 202 ⁄4
f) Nominal breaking Strength

GIRUM MINDAYE 23
Strength of Materials AASTU

𝐵𝑟𝑒𝑎𝑘𝑖𝑛𝑔 𝑙𝑜𝑎𝑑 160 × 103


𝜎𝑓 = = = 509.30 𝑁⁄𝑚𝑚2 = 509.30𝑀𝑃𝑎
𝐴𝑜 𝜋 × 202 ⁄4
g) True breaking Strength

𝐵𝑟𝑒𝑎𝑘𝑖𝑛𝑔 𝑙𝑜𝑎𝑑 160 × 103


𝜎′𝑓 = = 2
= 795.77 𝑁⁄𝑚𝑚2 = 795.77𝑀𝑃𝑎
𝐴𝑓 𝜋 × 16 4⁄

Example 1.9: A composite bar consists of an aluminum section rigidly fastened between a
bronze section and a steel section as shown in figure. Axial loads are applied at the positions
indicated. Determine the change in each section and the change in total length. Given Ebr=
100GPa, Eal = 70GPa, Est = 200GPa

Solution
First, determine the forces in each section.

∑ 𝐹𝑥 = 0, 𝑃𝑏𝑟 = +4𝑘𝑁, 𝑃𝑎𝑙 = −9𝑘𝑁, 𝑃𝑠𝑡 = −7𝑘𝑁

𝑃𝑖 𝐿𝑖
𝛿=∑
𝐴𝑖 𝐸𝑖
𝑖
Change in length in each section
𝑃𝑏𝑟 𝐿𝑏𝑟 (4 × 103 )(300)
𝛿𝑏𝑟 = = = 0.1𝑚𝑚
𝐴𝑏𝑟 𝐸𝑏𝑟 (120)(100 × 103 )
𝑃𝑎𝑙 𝐿𝑎𝑙 (−9 × 103 )(400)
𝛿𝑎𝑙 = = = −0.286𝑚𝑚
𝐴𝑎𝑙 𝐸𝑎𝑙 (180)(70 × 103 )

GIRUM MINDAYE 24
Strength of Materials AASTU

𝑃𝑠𝑡 𝐿𝑠𝑡 (−7 × 103 )(500)


𝛿𝑠𝑡 = = = −0.109𝑚𝑚
𝐴𝑠𝑡 𝐸𝑠𝑡 (160)(200 × 103 )
Change in total length
𝑃𝑖 𝐿𝑖 𝑃𝑏𝑟 𝐿𝑏𝑟 𝑃𝑎𝑙 𝐿𝑎𝑙 𝑃𝑠𝑡 𝐿𝑠𝑡
𝛿=∑ = + +
𝐴𝑖 𝐸𝑖 𝐴𝑏𝑟 𝐸𝑏𝑟 𝐴𝑎𝑙 𝐸𝑎𝑙 𝐴𝑠𝑡 𝐸𝑠𝑡
𝑖
𝛿 = 𝛿𝑏𝑟 + 𝛿𝑎𝑙 + 𝛿𝑠𝑡 = 0.1 − 0.286 − 0.109 = −0.295𝑚𝑚
Example 1.10: A metallic bar 250mm×100mm×50mm is loaded as shown in the figure. Find
the change in each dimension and total volume. Take E = 200GPa, Poisson's ratio, v = 0.25

Solution:
𝑃𝑥 400 × 103
𝜎𝑥 = = = 80𝑀𝑃𝑎
𝐴𝑥 100 × 50
𝑃𝑦 4000 × 103
𝜎𝑦 = = = −160𝑀𝑃𝑎
𝐴𝑦 100 × 250

𝑃𝑧 2000 × 103
𝜎𝑧 = = = 160𝑀𝑃𝑎
𝐴𝑧 250 × 50
𝜎𝑥 𝑣𝜎𝑦 𝑣𝜎𝑧
𝜖𝑥 = − −
𝐸 𝐸 𝐸
1
𝜖𝑥 = [80 − 0.25(−160 + 160)] = 0.4 × 10−3
200 × 103
𝛿𝑥 = 𝜖𝑥 𝐿𝑥 = (0.4 × 10−3 )(250) = 0.1𝑚𝑚
𝑣𝜎𝑥 𝜎𝑦 𝑣𝜎𝑧
𝜖𝑦 = − + −
𝐸 𝐸 𝐸
1
𝜖𝑦 = [−160 − 0.25(80 + 160)] = −1.1 × 10−3
200 × 103
𝛿𝑦 = 𝜖𝑦 𝐿𝑦 = (−1.1 × 10−3 )(50) = −0.055𝑚𝑚
𝑣𝜎𝑥 𝑣𝜎𝑦 𝜎𝑧
𝜖𝑧 = − − +
𝐸 𝐸 𝐸

GIRUM MINDAYE 25
Strength of Materials AASTU

1
𝜖𝑧 = [−0.25(80 − 160) + 160] = 0.9 × 10−3
200 × 103
𝛿𝑧 = 𝜖𝑧 𝐿𝑧 = (0.9 × 10−3 )(100) = 0.09𝑚𝑚
𝜖𝑣 = 𝜖𝑥 + 𝜖𝑦 + 𝜖𝑥 = (0.4 − 1.1 + 0.9)10−3 = 0.2 × 10−3

Or
1 − 2𝑣
𝜖𝑣 = (𝜎𝑥 + 𝜎𝑦 + 𝜎𝑧 )
𝐸
1 − 2(0.25)
𝜖𝑣 = [80 − 160 + 160] = 0.2 × 10−3
200 × 103
𝛿𝑉
𝜖𝑣 = ⟹ 𝛿𝑉 = 𝜖𝑣 𝑉 = (0.2 × 10−3 )(250 × 100 × 50) = 250𝑚𝑚2
𝑉
Example 1.11: A circle of diameter d =9in is scribed on an unstressed aluminum plate of
thickness t=3/4in. Forces acting in the plane of the plate later cause normal stresses σx=12ksi
and σz = 20ksi. For E = 10x106psi and v=1/3, determine the change in (a) the length of diameter
AB, (b) the length of diameter CD, (c) the thickness of the plate, and (d) the volume of the
plate.

You can use the generalized Hooke’s Law to determine the components of strain. These strains
can then be used to evaluate the various dimensional changes to the plate, and through the
dilatation assess the volume change. Note that σy=0.
𝜎𝑥 𝑣𝜎𝑦 𝑣𝜎𝑧 1
𝜖𝑥 = − − = (𝜎𝑥 − 𝑣𝜎𝑧 )
𝐸 𝐸 𝐸 𝐸
1 1
𝜖𝑥 = 6
[(12 × 103 ) − (20 × 103 )] = 0.533 × 10−3
10 × 10 3
𝑣𝜎𝑥 𝜎𝑦 𝑣𝜎𝑧 1
𝜖𝑦 = − + − = (𝑣𝜎𝑥 − 𝑣𝜎𝑧 )
𝐸 𝐸 𝐸 𝐸
1 1 1
𝜖𝑦 = 6
[− (12 × 103 ) − (20 × 103 )] = −1.067 × 10−3
10 × 10 3 3
𝑣𝜎𝑥 𝑣𝜎𝑦 𝜎𝑧 1
𝜖𝑧 = − − + = − (𝑣𝜎𝑥 − 𝜎𝑧 )
𝐸 𝐸 𝐸 𝐸

GIRUM MINDAYE 26
Strength of Materials AASTU

1 1
𝜖𝑧 = 6
[− (12 × 103 ) + (20 × 103 )] = 1.600 × 10−3
10 × 10 3
1 − 2𝑣 1 − 2𝑣
𝜖𝑣 = (𝜎𝑥 + 𝜎𝑦 + 𝜎𝑧 ) = (𝜎𝑥 + 𝜎𝑧 )
𝐸 𝐸
1 − 2(1⁄3)
𝜖𝑣 = 6
[(12 × 103 ) + (20 × 103 )] = 1.067 × 10−3
10 × 10
a) Diameter AB.
𝑇ℎ𝑒 𝑐ℎ𝑎𝑛𝑔𝑒 𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ 𝑖𝑠, 𝛿𝐵⁄𝐴 = 𝜖𝑥 𝑑 = (0.533 × 10−3 )(9) = 4.8 × 10−3 𝑖𝑛

b) Diameter CD.
𝑇ℎ𝑒 𝑐ℎ𝑎𝑛𝑔𝑒 𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ 𝑖𝑠, 𝛿𝐶⁄𝐷 = 𝜖𝑧 𝑑 = (1.600 × 10−3 )(9) = 14.4 × 10−3 𝑖𝑛

c) Thickness. Recalling that t =3/4in


𝑇ℎ𝑒 𝑐ℎ𝑎𝑛𝑔𝑒 𝑖𝑛 𝑙𝑡ℎ𝑖𝑐𝑘𝑛𝑒𝑠𝑠 𝑖𝑠, 𝛿𝑡 = 𝜖𝑦 𝑡 = (−1.067 × 10−3 )(3⁄4) = −0.800 × 10−3 𝑖𝑛

d) Volume of the Plate.


𝛿𝑉 3
𝜖𝑣 = ⟹ 𝛿𝑉 = 𝜖𝑣 𝑉 = (1.067 × 10−3 ) (15 × 15 × ) = 0.180𝑖𝑛3
𝑉 4

GIRUM MINDAYE 27
Strength of Materials AASTU

2 SHEAR FORCE AND BENDING MOMENT


2.1 Beams

A beam may be defined as a structural element, which has one dimension considerably larger
than the other two dimensions, namely breadth and depth, and is supported at a few points. The
distance between two adjacent supports is called span. In general, beams are long, straight bars
having a constant cross-sectional area. In most cases, the loads are perpendicular to the axis of
the beam. This transverse loading causes only bending and shear in the beam. When the loads
are not at a right angle to the beam, they also produce axial forces in the beam. The system of
forces consisting of applied loads and reactions keep the beam in equilibrium. The reactions
depend upon the type of supports and type of loading.
Table 2-1: Common type of supports

Beams are classified according to the way they are supported, as shown in Figure 2-1. The
distance L is called the span. Note that the reactions at the supports of the beams in Figure
2-1a, b, and c involve only three unknowns and can be determined by the methods of statics.

GIRUM MINDAYE 28
Strength of Materials AASTU

Such beams are said to be statically determinate. On the other hand, the reactions at the
supports of the beams in Figure 2-1d, e, and f involve more than three unknowns and cannot
be determined by the methods of statics alone. The properties of the beams with regard to their
resistance to deformations must be taken into consideration. Such beams are said to be
statically indeterminate.

Figure 2-1: Common beam support configurations.

The transverse loading of a beam may consist of concentrated loads P1, P2, . . . expressed in
newtons, pounds, or their multiples of kilonewtons and kips (Figure 2-1a); of a distributed load
w expressed in N/m, kN/m, lb/ft, or kips/ft (Fig. 5.1b); or of a combination of both. When the
load per unit length w is a constant value over part of the beam (as between A and B in Figure
2-1b), the load is uniformly distributed and linearly varying value over part of the beam (as
between B and C in Figure 2-2b), the load is linearly distributed.

Figure 2-2: Transversely loaded beams.

2.2 Sign Convention for Shear Force and Bending Moment

Before presenting a method for finding the internal normal force, shear force, and bending
moment, we will need to establish a sign convention to define their “positive” and “negative”
values. Although the choice is arbitrary, the sign convention to be adopted here has been widely
accepted in structural engineering practice, and is illustrated in Figure 2-3a. On the left-hand
face of the cut member the normal force N acts to the right, the internal shear force V acts
downward, and the moment M acts counterclockwise. In accordance with Newton’s third law,
an equal but opposite normal force, shear force, and bending moment must act on the right-
hand face of the member at the section. Perhaps an easy way to remember this sign convention
is to isolate a small segment of the member and note that positive normal force tends to elongate
the segment, Figure 2-3b; positive shear tends to rotate the segment clockwise, Figure 2-3c;

GIRUM MINDAYE 29
Strength of Materials AASTU

and positive bending moment tends to bend the segment concave upward, so as to “hold water,”
Figure 2-3d.

Figure 2-3: Sign Convention.

When a beam is subjected to transverse loads, the internal forces in any section of the beam
consist of a shear force V and a bending couple M. For example, a simply supported beam AB
is carrying two concentrated loads and a uniformly distributed load (Figure 2-4a). To determine
the internal forces in a section through point C, draw the free-body diagram of the entire beam
to obtain the reactions at the supports (Figure 2-4b). Passing a section through C, then draw the
free-body diagram of AC (Figure 2-4c), from which the shear force V and the bending couple
M are found. The bending couple M creates normal stresses in the cross section, while the
shear force V creates shearing stresses. In most cases, the dominant criterion in the design of
a beam for strength is the maximum value of the normal stress in the beam.

Figure 2-4: Analysis of a simply supported beam.

2.3 Relationship between Shear Force and Bending Moment

We will now obtain some important relationships between loads, shear forces, and bending
moments in beams. These relationships are quite useful when investigating the shear forces
and bending moments throughout the entire length of a beam, and they are especially helpful
when constructing shear-force and bending-moment diagrams.

As a means of obtaining the relationships, an element of a beam is cut out between two cross
sections that are a distance dx apart (Figure 2-5). The load acting on the top surface of the
element may be a distributed load, a concentrated load, or a couple as shown in Figure 2-5a, b,
and c, respectively.. The sign conventions for these loads are as follows.

GIRUM MINDAYE 30
Strength of Materials AASTU

Distributed loads and concentrated loads are positive when they act downward on the beam
and negative when they act upward. A couple acting as a load on a beam is positive when it
is counterclockwise and negative when it is clockwise.

The shear forces and bending moments acting on the sides of the element are shown in their
positive directions in Figure 2-5. In general, the shear forces and bending moments vary along
the axis of the beam. Therefore, their values on the right-hand face of the element may be
different from their values on the left-hand face.

Figure 2-5: Element of a beam used in deriving the relationships among loads, shear forces,
and bending moments (All loads and stress resultants are shown in their positive directions)
2.3.1 Region of Distributed Loads

We consider a simply supported beam subjected to a uniformly distributed load w(x)


throughout its length (L), as shown in Figure 2-6. Consider a small element of length dx at a
distance of x form left support. Let the shear force and bending moment at a section located at
a distance of x from the left support be V and M, respectively, and at a section x + dx be V +
dV and M + dM, respectively.

Figure 2-6: Element of a beam (All loads are shown in their positive directions.)
Equilibrium of forces in the vertical direction gives

↑ + ∑ 𝐹𝑦 = 0; 𝑉 − 𝑤𝑑𝑥 − 𝑉 − 𝑑𝑉 = 0

𝒅𝑽
𝑑𝑉 = −𝑤𝑑𝑥 ⟹ = −𝒘, 𝑜𝑟 𝑽 = − ∫ 𝒘𝒅𝒙
𝒅𝒙
From this equation, the change in shear force, dV, between any two points on the beam is equal
to the area under the distributed-load curve between those same two points. In addition, the

GIRUM MINDAYE 31
Strength of Materials AASTU

rate of change of the shear force at any point on the axis of the beam is equal to the negative of
the intensity of the distributed load at that same point.
Note: If the sign convention for the distributed load is reversed so that w is positive upward
instead of downward, the minus sign is omitted in the preceding equation.

Summing moments about an axis at the left-hand side of the element and discarding products
of differentials (because they are negligible compared to the other terms), we obtain the
following relationship:
𝑑𝑥
↺ + ∑ 𝑀 = 0, −𝑀 − 𝑤𝑑𝑥 ( ) − (𝑉 + 𝑑𝑉)𝑑𝑥 + 𝑀 + 𝑑𝑀 = 0
2
𝒅𝑴
𝑑𝑀 = 𝑉𝑑𝑥, ⟹ = 𝑽, 𝑜𝑟 𝑴 = ∫ 𝑽𝒅𝒙
𝒅𝒙
This equation shows that the rate of change of the bending moment at any point on the axis of
a beam is equal to the shear force at that same point. For instance, if the shear force is zero in
a region of the beam, then the bending moment is constant in that same region. In addition, the
change in bending moment, dM, between any two points is equal to the corresponding area
under the shear force curve.
2.3.2 Regions of concentrated Loads and Moments

Consider the free-body diagram of a very thin portion of the beam (see Figure 2-5a) directly
beneath one of the concentrated loads (Figure 2-5b).

Force equilibrium for this free body can be stated as


↑ + ∑ 𝐹𝑦 = 0, 𝑉 − 𝑃 − 𝑉 − 𝑑𝑉 = 0 ⟹ 𝒅𝑽 = −𝑷
This result means that an abrupt change in the shear force occurs at any point where a
concentrated load acts. Passing from left to right through the point of the load application, the
shear force decreases by an amount equal to the magnitude of the downward load P.
Equilibrium of moments about the left-hand face of the element gives
𝑑𝑥
↺ + ∑ 𝑀 = 0, −𝑀 − 𝑃 ( ) − (𝑉 + 𝑑𝑉)𝑑𝑥 + 𝑀 + 𝑑𝑀 = 0
2
𝑑𝑥 𝒅𝑴 𝑷
𝑑𝑀 = 𝑃 ( ) + 𝑉𝑑𝑥 ⟹ =𝑽+
2 𝒅𝒙 𝟐
𝑑𝑥 𝑑𝑥
𝑑𝑀 = 𝑃 ( ) + 𝑉𝑑𝑥, 𝑠𝑖𝑛𝑐𝑒 ≈ 0 ⟹ 𝒅𝑴 = 𝑽𝒅𝒙
2 2

Equilibrium of moments about the right-hand face of the element gives

GIRUM MINDAYE 32
Strength of Materials AASTU

𝑑𝑥
↺ + ∑ 𝑀 = 0, −𝑀 − 𝑉𝑑𝑥 + 𝑃 ( ) + 𝑀 + 𝑑𝑀 = 0
2
𝑑𝑥 𝒅𝑴 𝑷
𝑑𝑀 = −𝑃 ( ) + 𝑉𝑑𝑥 ⟹ =𝑽−
2 𝒅𝒙 𝟐
𝑑𝑥 𝑑𝑥
𝑑𝑀 = −𝑃 ( ) + 𝑉𝑑𝑥, 𝑠𝑖𝑛𝑐𝑒 ≈ 0 ⟹ 𝒅𝑴 = 𝑽𝒅𝒙
2 2

Thus, the bending moment does not change when passing through the point of application of a
concentrated load.
Even though the bending moment M does not change at a concentrated load, its rate of change
dM/dx undergoes an abrupt change. At the left-hand side of the element, the rate of change of
the bending moment is dM/dx = V+ P/2. At the right-hand side, the rate of change is dM/dx
=V- P/2. Therefore, at the point of application of a concentrated load P, the rate of change
dM/dx of the bending moment decreases abruptly by an amount equal to P.
The last case to be considered is a load in the form of a couple M0 (Figure 2-5c).

Equilibrium of the element in the vertical direction

↑ + ∑ 𝐹𝑦 = 0, 𝑉 − −𝑉 − 𝑑𝑉 = 0 ⟹ 𝒅𝑽 = 𝟎

This shows that the shear force does not change at the point of application of a couple.
Equilibrium of moments about the left-hand side of the element gives

↺ + ∑ 𝑀 = 0, −𝑀 − 𝑉𝑑𝑥 + 𝑀0 + 𝑀 + 𝑑𝑀 = 0

𝑑𝑀 = −𝑀0 + 𝑉𝑑𝑥, 𝐴𝑠 𝑑𝑥 𝑎𝑝𝑝𝑟𝑜𝑎𝑐ℎ𝑒𝑠 𝑧𝑒𝑟𝑜 ⟹ 𝒅𝑴 = −𝑴𝟎


This equation shows that the bending moment decreases by M0 when moving from left to right
through the point of load application. Thus, the bending moment changes abruptly at the point
of application of a couple.
2.4 Shear Force and Bending Moment Diagrams

When designing a beam, we usually need to know how the shear forces and bending moments
vary throughout the length of the beam. Of special importance are the maximum and minimum
values of these quantities. Information of this kind is usually provided by graphs in which the
shear force and bending moment are plotted as ordinates and the distance x along the axis of

GIRUM MINDAYE 33
Strength of Materials AASTU

the beam is plotted as the abscissa. Such graphs are called shear-force and bending-moment
diagrams.
2.4.1 Maximum and Minimum Bending Moments

In mathematics, we find the maximum or minimum value of a function f(x) by first taking the
derivative of the function, setting the derivative equal to zero, and determining the location x
at which the derivative is zero. Then, once this value of x is known, it can be substituted into
f(x) and the maximum or minimum value ascertained.

In the context of shear and moment diagrams, the function of interest is the bending moment
function M(x). The derivative of this function is dM/dx, and accordingly, the maximum bending
moment will occur at locations where dM/dx= 0. Notice, dM/dx = V. If these two equations are
combined, we can conclude that the maximum or minimum bending moment occurs at
locations where V = 0. This conclusion will be true unless there is a discontinuity in the M
diagram caused by an external concentrated moment. Consequently, maximum and minimum
bending moments will occur at points where the V curve crosses the V = 0 axis, as well as at
points where external concentrated moments are applied to the beam. Bending moments
corresponding to the location of discontinuities also should be computed to check for maximum
or minimum bending moment values.
2.4.2 Rules for Shear Force and Bending Moment Diagrams

Based on section 2.3, six rules that can be used to construct shear force and bending moment
diagrams for any beam are grouped according to usage as follows:
Rules for the Shear Force Diagram
Rule 1: The shear force diagram is discontinuous at points subjected to concentrated loads P.
A downward P causes the V diagram to jump downward and an upward P causes the V diagram
to jump upward.
Rule 2: The change in internal shear force between any two locations x1 and x2 is equal to the
area under the distributed load curve. A negative area results from negative w (i.e., a downward
distributed load).
Rule 3: At any location x, the slope of the V diagram is equal to the intensity of the distributed
load -w.
Rules for the Bending Moment Diagram
Rule 4: The change in internal bending moment between any two locations x1 and x2 is equal
to the area under the shear force diagram. The area computed from negative shear force values
is considered negative.
Rule 5: At any location x, the slope of the M diagram is equal to the intensity of the internal
shear force V.
Rule 6: The bending-moment diagram is discontinuous at points subjected to external
concentrated moments. A counterclockwise external moment causes the M diagram to jump
downward and a clockwise external moment causes the M diagram to jump upward.

GIRUM MINDAYE 34
Strength of Materials AASTU

Table 2-2: construction Rules for Shear-Force and Bending-moment Diagrams

GIRUM MINDAYE 35
Strength of Materials AASTU

2.4.3 Worked Examples

Example 2.1: Draw the shear-force and bending-moment diagrams for the simply supported
beam shown.

Plan the Solution


First, determine the reaction forces at pin A and roller C. Then, consider two intervals along
the beam span: between A and B, and between B and C. Cut a section in each interval and draw
the appropriate free-body diagram (FBD), showing the unknown internal shear force V and
internal bending moment M acting on the exposed surface. Write the equilibrium equations for
each FBD, and solve them for functions describing the variation of V and M with location x
along the span. Plot these functions to complete the shear-force and bending-moment diagrams.
Solution
(a) Draw the free-body diagram of the entire beam.
(b) Determine the reactions at the supports from the free-body diagram of the entire beam.
Since this beam is symmetrically supported and symmetrically loaded, the reaction forces
must also be symmetric. Therefore, each support exerts an upward force equal to P/2.
Because no applied loads act in the x direction, the horizontal reaction force at pin support
A is zero.
𝑃𝐿 𝑃
↺ + ∑ 𝑀𝐴 = 0, 𝐶𝑦 𝐿 − = 0 ⟹ 𝐶𝑦 =
2 2
𝑃 𝑃
↑ + ∑ 𝐹𝑦 = 0, 𝐴𝑦 − 𝑃 + = 0 ⟹ 𝐴𝑦 =
2 2
(c) Determine the internal shear force and bending moment function
In general, the beam will be sectioned at an arbitrary distance x from pin support A and all
forces acting on the free body will be shown, including the unknown internal shear force V
and internal bending moment M acting on the exposed surface. Note that positive directions
are assumed for both V and M.
Section a-a (0≤x<L/2):

Since no forces act in the x direction, the equilibrium equation ∑Fx = 0 is trivial. The sum of
forces in the vertical direction yields the desired function for V:
𝑃 𝑃
↑ + ∑ 𝐹𝑦 = 0, −𝑉 =0⟹𝑉 = − − − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿⁄2
2 2
The sum of moments about section a–a gives the desired function for M:

GIRUM MINDAYE 36
Strength of Materials AASTU

𝑃 𝑃𝑥
↺ + ∑ 𝑀𝑎 = 0, − ( ) (𝑥) + 𝑀 = 0 ⟹ 𝑀 = − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿⁄2
2 2
Section b-b (L/2≤x<L):

The sum of forces in the vertical direction yields the desired function for V:
𝑃 𝑃
↑ + ∑ 𝐹𝑦 = 0, 𝑉+ =0⟹𝑉=− − − − − − −𝑓𝑜𝑟 𝐿⁄2 ≤ 𝑥 ≤ 𝐿
2 2
The equilibrium equation for the sum of moments about section b–b gives the desired
function for M:
𝑃 𝑃(𝐿 − 𝑥)
↺ + ∑ 𝑀𝑏 = 0, −𝑀 + ( ) (𝐿 − 𝑥) = 0 ⟹ 𝑀 = − − − 𝑓𝑜𝑟 𝐿⁄2 ≤ 𝑥 ≤ 𝐿
2 2

The maximum internal shear force is Vmax= ±P/2. The maximum internal bending moment is
Mmax = PL/4, and it occurs at x = L/2.
Notice that the concentrated load causes a discontinuity at its point of application. In other
words, the shear-force diagram “jumps” by an amount equal to the magnitude of the
concentrated load. The jump in this case is downward, which is the same direction as that of
the concentrated load P.

GIRUM MINDAYE 37
Strength of Materials AASTU

Example 2.2: Draw the shear-force and bending-moment diagrams for the simple beam
shown.

Solution
(a) Draw the free-body diagram of the entire beam.

(b) Determine the reactions at the supports from the free-body diagram of the entire beam.
𝑀0
↺ + ∑ 𝑀𝐴 = 0, 𝐴𝑦 𝐿 − 𝑀0 = 0 ⟹ 𝐴𝑦 =
𝐿
𝑀0 𝑀0
↑ + ∑ 𝐹𝑦 = 0, 𝐴𝑦 + = 0 ⟹ 𝐴𝑦 = −
𝐿 𝐿
Since no forces act in the x direction, the equilibrium equation ∑Fx = 0 is trivial.
(c) Determine the internal shear force and bending moment function
Section a-a (0≤x<L/2):

The sum of forces in the vertical direction yields the desired function for V:
𝑀0 𝑀0
↑ + ∑ 𝐹𝑦 = 0, − −𝑉 =0⟹𝑉 =− − − − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿⁄2
𝐿 𝐿
The sum of moments about section a–a gives the desired function for M:
𝑀0 𝑀0
↺ + ∑ 𝑀𝑎 = 0, ( ) (𝑥) + 𝑀 = 0 ⟹ 𝑀 = − 𝑥 − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿⁄2
𝐿 𝐿

The sum of forces in the vertical direction yields the desired function for V:

GIRUM MINDAYE 38
Strength of Materials AASTU

𝑀0 𝑀0
↑ + ∑ 𝐹𝑦 = 0, − −𝑉 =0⟹𝑉 =− − − − − − −𝑓𝑜𝑟 𝐿⁄2 ≤ 𝑥 ≤ 𝐿
𝐿 𝐿
The equilibrium equation for the sum of moments about section b–b gives the desired
function for M:
𝑀0 𝑀0
↺ + ∑ 𝑀𝐷 = 0, ( ) (𝑥) − 𝑀0 + 𝑀 = 0 ⟹ 𝑀 = 𝑀0 − 𝑥 − −𝑓𝑜𝑟 𝐿⁄2 ≤ 𝑥 ≤ 𝐿
𝐿 𝐿

The maximum internal shear force is Vmax = -M0/L. The maximum internal bending moment
is Mmax= ±M0/2, and it occurs at x = L/2.
Notice that the concentrated moment does not affect the shear-force diagram at B. It does,
however, create a discontinuity in the bending-moment diagram at the point of application of
the concentrated moment: The bending-moment diagram “jumps” by an amount equal to the
magnitude of the concentrated moment. The clockwise concentrated external moment M0
causes the bending-moment diagram to “jump” upward at B by an amount equal to the
magnitude of the concentrated moment.
Example 2.3: Draw the shear force and bending-moment diagrams for the simply supported
beam shown.

Solution
(a) Draw the free-body diagram of the entire beam.
(b) Determine the reactions at the supports from the free-body diagram of the entire beam.

GIRUM MINDAYE 39
Strength of Materials AASTU

Since this beam is symmetrically supported and symmetrically loaded, the reaction forces
must also be symmetric. Therefore, each support exerts an upward force equal to wL/2.
(c) Determine the internal shear force and bending moment function
Section a-a (0≤x<L):
Section the beam at an arbitrary distance x between A and B. Make sure that the original
distributed load w is shown on the FBD at the outset.

Since no forces act in the x direction, the equilibrium equation ∑Fx = 0 is trivial. The sum of
forces in the vertical direction yields the desired function for V:
𝑤𝐿 𝐿
↑ + ∑ 𝐹𝑦 = 0, − 𝑤𝑥 − 𝑉 = 0 ⟹ 𝑉 = 𝑤 ( − 𝑥) − − − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿
2 2
The sum of moments about section a–a gives the desired function for M:
𝑤𝐿 𝑥
↺ + ∑ 𝑀𝑎 = 0, − ( ) (𝑥) + (𝑤𝑥) ( ) + 𝑀 = 0
2 2
𝑤𝑥
⟹𝑀= (𝑥 − 𝐿) − − − − − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿
2

The maximum internal shear force is Vmax= ±wL/2, and it is found at A and B. The maximum
internal bending moment is Mmax= wL2/8, and it occurs at x = L/2.

GIRUM MINDAYE 40
Strength of Materials AASTU

Note that the maximum bending moment occurs at a location where the shear force V is equal
to zero.
Example 2.4: Draw the shear-force and bending-moment diagrams for the simply supported
beam shown.

Solution
(a) Draw the free-body diagram of the entire beam.

(b) Determine the reactions at the supports from the free-body diagram of the entire beam.
Because no applied loads act in the x direction, the horizontal reaction force at pin support
A is zero.
𝑤𝐿 𝐿 𝑤𝐿 3𝐿 11
↺ + ∑ 𝑀𝐴 = 0, 𝐶𝑦 𝐿 − ( ) ( ) − ( ) ( ) = 0 ⟹ 𝐶𝑦 = 𝑤𝐿
4 3 2 4 24
𝑤𝐿 𝑤𝐿 11 7
↑ + ∑ 𝐹𝑦 = 0, 𝐴𝑦 − − + 𝑤𝐿 = 0 ⟹ 𝐴𝑦 = 𝑤𝐿
4 2 24 24
(c) Determine the internal shear force and bending moment function
Section a-a (0≤x<L/2):
Section the beam at an arbitrary distance x between A and B. Make sure that the original
distributed load w is shown on the FBD at the outset.

The sum of forces in the vertical direction yields the desired function for V:
7 𝑤𝑥 2 𝑤𝑥 2 7
↑ + ∑ 𝐹𝑦 = 0, 𝑤𝐿 − −𝑉 =0⟹𝑉 =− + 𝑤𝐿 − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿/2
24 𝐿 𝐿 24
The sum of moments about section a–a gives the desired function for M:
7 𝑤𝑥 2 𝑥
↺ + ∑ 𝑀𝑎 = 0, − ( 𝑤𝐿) (𝑥) + ( )( ) + 𝑀 = 0
24 𝐿 3

GIRUM MINDAYE 41
Strength of Materials AASTU

𝑤𝑥 3 7
⟹𝑀=− + 𝑤𝐿𝑥 − − − − − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿
3𝐿 24
Section a-a (L/2≤x<L):
Section the beam at an arbitrary distance x between B and C. Make sure that you replace the
original distributed loads on the FBD before deriving the V and M functions.

Based on this FBD, the equilibrium equations can be written as follows:


7 𝑤𝐿 𝐿
↑ + ∑ 𝐹𝑦 = 0, 𝑤𝐿 − − 𝑤 (𝑥 − ) − 𝑉 = 0
24 4 2
7 𝑤𝐿 𝐿
⟹𝑉= 𝑤𝐿 − − 𝑤 (𝑥 − ) − − − −𝑓𝑜𝑟 𝐿/2 ≤ 𝑥 < 𝐿
24 4 2
7 𝑤𝐿 𝐿 𝐿 1 𝐿
↺ + ∑ 𝑀𝑏 = 0, − ( 𝑤𝐿) (𝑥) + ( ) (𝑥 − ) + 𝑤 (𝑥 − ) [ (𝑥 − )] + 𝑀 = 0
24 4 3 2 2 2
7 𝑤𝐿 𝐿 𝑤 𝐿 2
⟹𝑀= 𝑤𝐿𝑥 − ( ) (𝑥 − ) − (𝑥 − ) − − − − − − − −𝑓𝑜𝑟 𝐿/2 ≤ 𝑥 < 𝐿
24 4 3 2 2
These equations can be simplified to
13 𝑤
𝑉 = 𝑤 ( 𝐿 − 𝑥) 𝑎𝑛𝑑 𝑀= (−12𝑥 2 + 13𝐿𝑥 − 𝐿2 )
24 24

GIRUM MINDAYE 42
Strength of Materials AASTU

Notice that the maximum bending moment occurs at a location where the shear force V is equal
to zero.
Example 2.5: Draw the shear-force and bending-moment diagrams for the cantilever beam
shown.

Solution
(a) Draw the free-body diagram of the entire beam.

(b) Determine the reactions at the supports from the free-body diagram of the entire beam.
Because no applied loads act in the x direction, the horizontal reaction force at pin support
A is zero.
↺ + ∑ 𝑀𝐴 = 0, −𝑀𝐴 + 19 × 2 − 6 × 5 = 0 ⟹ 𝑀𝐴 = 8𝑘𝑁𝑚

↑ + ∑ 𝐹𝑦 = 0, 𝐴𝑦 + 19 − 6 = 0 ⟹ 𝐴𝑦 = −13𝑘𝑁
Since Ay is negative, it really acts downward. The correct direction of this reaction force will
be shown in subsequent free-body diagrams
(c) Determine the internal shear force and bending moment function
Section a-a (0≤x<2):
Section the beam at an arbitrary distance x between A and B. The FBD for this section is shown.
From the equilibrium equations for this FBD, determine the desired functions for V and M:

↑ + ∑ 𝐹𝑦 = 0, −13 − 𝑉 = 0 ⟹ 𝑉 = −13𝑘𝑁 − −𝑓𝑜𝑟 0 ≤ 𝑥 < 2

↺ + ∑ 𝑀𝑎 = 0, −8 + (13)(𝑥) + 𝑀 = 0 ⟹ 𝑀 = −13𝑥 + 8 − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 2

Section b-b (2≤x<4):

GIRUM MINDAYE 43
Strength of Materials AASTU

From a section cut between B and C, determine the desired shear and moment functions:

↑ + ∑ 𝐹𝑦 = 0, −13 + 19 − 𝑉 = 0 ⟹ 𝑉 = 6𝑘𝑁 − −𝑓𝑜𝑟 2 ≤ 𝑥 < 4

↺ + ∑ 𝑀𝑏 = 0, −8 + (13)(𝑥) − 19(𝑥 − 2) + 𝑀 = 0

⟹ 𝑀 = 6𝑥 − 30 − − − − 𝑓𝑜𝑟 2 ≤ 𝑥 < 4
Section c-c (4≤x<6):
From a section cut between C and D, determine the desired shear and moment functions:

↑ + ∑ 𝐹𝑦 = 0, −13 + 19 − 3(𝑥 − 4) − 𝑉 = 0 ⟹ 𝑉 = −3𝑥 + 18 − −𝑓𝑜𝑟 4 ≤ 𝑥 < 6


1
↺ + ∑ 𝑀𝑏 = 0, −8 + (13)(𝑥) − 19(𝑥 − 2) + 3(𝑥 − 4) [ (𝑥 − 4)] + 𝑀 = 0
2
⟹ 𝑀 = −1.5𝑥 2 + 18𝑥 − 54 − − − − 𝑓𝑜𝑟 4 ≤ 𝑥 < 6

GIRUM MINDAYE 44
Strength of Materials AASTU

Example 2.6: Draw the shear and bending-moment diagrams for a cantilever beam AB of
span L supporting a uniformly distributed load w.

Solution
Cut the beam at a point C, located between A and B, and draw the free-body diagram of AC

↑ + ∑ 𝐹𝑦 = 0, −𝑤𝑥 − 𝑉 = 0 ⟹ 𝑉 = −𝑤𝑥 − − − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿


1 1
↺ + ∑ 𝑀𝐶 = 0, (𝑤𝑥) ( 𝑥) + 𝑀 = 0 ⟹ 𝑀 = − 𝑤𝑥 2 − − − −𝑓𝑜𝑟 0 ≤ 𝑥 < 𝐿
2 2

Note that the shear diagram is represented by an oblique straight line and the bending-
moment diagram by a parabola. The maximum values of V and M both occur at B, where
𝑤𝐿2
𝑉 = −𝑤𝐿 𝑎𝑛𝑑 𝑀 = −
2

Example 2.7: A simple beam with an overhang is supported at points A and B. A uniform load
of intensity w=6kN/m acts throughout the length of the beam and a concentrated P=28kN load
acts at a point 3 m from the left-hand support. The span length is 8 m and the length of the
overhang is 2 m. Calculate the shear force V and bending moment M at cross section D located
5 m from the left-hand support.

GIRUM MINDAYE 45
Strength of Materials AASTU

Solution
(a) Reactions.
We begin by calculating the reactions RA and RB from equations of equilibrium for the entire
beam considered as a free body. Thus, taking moments about the supports at B and A,
respectively, we find
RA=40kN RB=48kN
(b) Shear force and bending moment at section D.
Now we make a cut at section D and construct a free-body diagram of the left-hand part of
the beam (Fig. b). When drawing this diagram, we assume that the unknown stress resultants
V and M are positive.

The equations of equilibrium for the free body are as follows:

+↑ ∑ 𝐹𝑦 = 0, 40 − 28 − 6 × 5 − 𝑉 = 0 ⟹ 𝑽 = −𝟏𝟖𝒌𝑵

↺ + ∑ 𝑀𝐷 = 0, −(40 × 5) + (28 × 2) + (6 × 5)(2.5) + 𝑀 = 0 ⟹ 𝑴 = 𝟔𝟗𝒌𝑵𝒎

Example 2.8: Determine the shear and moment in the beam shown in figure as a function of x.

GIRUM MINDAYE 46
Strength of Materials AASTU

Solution
(a) Support Reactions.
To determine the support reactions, the distributed load is divided into a triangular and
rectangular loading, and these loadings are then replaced by their resultant forces. These
reactions have been computed and are shown on the beam’s free body diagram, Fig. b.

(b) Shear and Moment Functions.


A free-body diagram of the cut section is shown in Fig. c. As above, the trapezoidal loading is
replaced by rectangular and triangular distributions. Note that the intensity of the triangular
load at the cut is found by proportion. The resultant force of each distributed loading and its
location are indicated. Applying the equilibrium equations, we have
1 𝑥
+↑ ∑ 𝐹𝑦 = 0, 75 − 10𝑥 − [ (20) ( ) 𝑥] − 𝑉 = 0
2 9
⟹ 𝑽 = 𝟕𝟓 − 𝟏𝟎𝒙 − 𝟏. 𝟏𝟏𝒙𝟐
𝑥 1 𝑥 𝑥
↺ + ∑ 𝑀𝑥 = 0, −75𝑥 + (10𝑥) ( ) − [ (20) ( ) 𝑥] + 𝑀 = 0
2 2 9 3
⟹ 𝑴 = 𝟕𝟓𝒙 − 𝟓𝒙𝟐 − 𝟎. 𝟑𝟕𝟎𝒙𝟑

2.4.4 Graphical Method for Constructing Shear and Moment Diagrams

The method for constructing V and M diagrams presented here is called the graphical method
because the load diagram is used to construct the shear force diagram and then the Shear force
diagram is used to construct the bending moment diagram. The six rules outlined in section
2.4.2 are used to make these constructions. The graphical method is much less time consuming
than the process of deriving V(x) and M(x) functions for the entire beam, and it provides the
information necessary to analyze and design beams. The general procedure can be summarized
by the following steps:
(1) Complete the Load Diagram: Sketch the beam, including the supports, loads, and key
dimensions. Calculate the external reaction forces, and if the beam is a cantilever, find the
external reaction moment. Show these reaction forces and moments on the load diagram,
using arrows to indicate the direction in which they act.
(2) Construct the Shear Force Diagram: The shear-force diagram will be constructed directly
beneath the load diagram. For that reason, it is convenient to draw a series of vertical lines

GIRUM MINDAYE 47
Strength of Materials AASTU

beneath significant locations on the beam in order to help align the diagrams. Begin the
shear force diagram by drawing a horizontal axis, which will serve as the x axis for the V
diagram. The shear force diagram should always start and end on the value V=0.
The idea of starting and ending at V = 0 is related to the beam equilibrium equation ΣFy= 0. A
shear force diagram that does not return to V = 0 at the rightmost end of the beam indicates
that equilibrium has not been satisfied. The most common cause of this error in the V diagram
is a mistake in the calculated beam reaction forces.
(3) Locate Key points on the Shear-Force Diagram: Special attention should be paid to
locating points where the V diagram crosses the V = 0 axis, because these points indicate
locations where the bending moment will be either a maximum or a minimum value.
(4) Construct the Bending moment Diagram: The bending moment diagram will be
constructed directly beneath the shear force diagram. Begin the bending moment diagram
by drawing a horizontal axis, which will serve as the x-axis for the M diagram. The bending
moment diagram should always start and end on the value M = 0.
The idea of starting and ending at M = 0 is related to the beam equilibrium equation ΣM = 0.
A bending moment diagram that does not return to M = 0 at the rightmost end of the beam
indicates that equilibrium has not been satisfied. The most common cause of this error in the
M diagram is a mistake in the calculated beam reaction forces. If the applied loads included
concentrated moments, another common error is “jumping” the wrong direction at the
discontinuities.
Relationships among the Diagram Shapes
If the V diagram is constant for a beam segment, then the M diagram will be linear, making the
M diagram relatively straightforward to sketch. If the V diagram is linear for a beam segment,
then the M diagram will be quadratic (i.e., a parabola). A parabola can take one of two shapes:
either concave or convex. The proper shape for the M diagram can be determined from
information found on the V diagram, since the slope of the M diagram is equal to the intensity
of the shear force V (Rule 5). Various shear-force diagram shapes and their corresponding
bending moment shapes are illustrated in Figure 7.12.

GIRUM MINDAYE 48
Strength of Materials AASTU

Table 2-3: Relationships between V and M diagram shapes.

GIRUM MINDAYE 49
Strength of Materials AASTU

Example 2.9: Draw the shear force and bending moment diagrams for the simply supported
beam shown. Determine the maximum bending moment that occurs in the span.

Plan the Solution


Complete the load diagram by calculating the reaction forces at pin A and roller D. Since only
concentrated loads act on this beam, use Rule 1 to construct the shear force diagram from the
load diagram. Construct the bending moment diagram from the shear force diagram, using Rule
4 to calculate the change in bending moments between key points.
Solution
a) Support Reactions
An FBD of the entire beam is shown. Since no loads act in the horizontal direction, the
equilibrium equation ΣFx = 0 is trivial and will not be considered further. The nontrivial
equilibrium equations are as follows:

↺ + ∑ 𝑀𝐷 = 0, −21𝐴𝑦 + (12 × 17) + (10 × 9) = 0 ⟹ 𝑨𝒚 = 𝟏𝟒𝒌𝑵𝒎

+↑ ∑ 𝐹𝑦 = 0, 14 − 12 − 10+𝐷𝑦 = 0 ⟹ 𝑫𝒚 = 𝟖𝒌𝑵

b) Construct the Shear-Force Diagram


On the load diagram, show the reaction forces acting in their proper directions. Draw a series
of vertical lines beneath key points on the beam, and draw a horizontal line that will define the
axis for the V diagram. Use the steps outlined next to construct the V diagram. (Note: The
lowercase letters on the V diagram correspond to the explanations given for each step.)

GIRUM MINDAYE 50
Strength of Materials AASTU

a V(0-) = 0 kips (zero shear at end of beam).


b V(0+) = 14 kips (Rule 1: V diagram jumps up by an amount equal to the 14 kip reaction).
c V(4-) = 14 kips (Rule 2: Since w = 0, the area under the w curve is also zero. Hence, there is
no change in the shear force diagram).
d V(4+) = 2 kips (Rule 1: V diagram jumps down by 12 kips).
e V(12-) = 2 kips (Rule 2: The area under the w curve is zero; therefore, dV = 0).
f V(12+) = –8 kips (Rule 1: V diagram jumps down by 10 kips).
g V(21-) = –8 kips (Rule 2: The area under the w curve is zero; therefore, dV = 0).
h V(21+) = 0 kips (Rule 1: V diagram jumps up by an amount equal to the 8 kip reaction force
and returns to V = 0 kips).
Notice that the V diagram started at Va = 0 and finished at Vh = 0.
c) Construct the Bending-Moment Diagram
Starting with the V diagram, use the steps that follow to construct the M diagram. (Note: The
lowercase letters on the M diagram correspond to the explanations given for each step.)

GIRUM MINDAYE 51
Strength of Materials AASTU

i M(0) = 0 (zero moment at the pinned end of


a simply supported beam).
j M(4) = 56 kip.ft (Rule 4: The change in
bending moment dM between any two points
is equal to the area under the V diagram). The
area under the V diagram between x = 0ft and
x = 4ft is simply the area of rectangle (1),
which is 4ft wide and +14kips tall. The area of
this rectangle is (+14kips)(4ft) = +56 kip.ft (a
positive value). Since M = 0kip.ft at x = 0ft and
the change in bending moment is dM = +56
kip.ft, the bending moment at x = 4 ft is M = 56
kip.ft.
k M(12) = 72 kip.ft (Rule 4: dM = area under
the V diagram). dM is equal to the area under
the V diagram between x = 8ft and x =12ft. The
area of rectangle (2) is (+2 kips) (8ft) =
+16kip.ft. Therefore, dM = +16kip.ft (a
positive value). Since M = +56 kip.ft at j and dM = +16 kip.ft, the bending moment at k is M =
56 + 16 = +72 kip.ft. Even though the shear force decreases from +14 kips to +2 kips, notice
that the bending moment continues to increase in this region
l M(21) = 0 kip.ft (Rule 4: dM = area under the V diagram). The area under the V diagram
between x = 12ft and x = 21ft is the area of rectangle (3), which is (–8 kips) (9 ft) = -72 kip.ft
(a negative value); therefore, dM = -72 kip.ft. At point k, M = +72 kip.ft. The bending moment
changes by dM = -72 kip.ft between k and l; consequently, the bending moment at x = 21ft is
M = 0 kip.ft. This result is correct, since we know that the bending moment at roller D must be
zero.
Notice that the M diagram started at Mi = 0 and finished at Ml = 0. Notice also that the M
diagram consists of linear segments. From Rule 5 (the slope of the M diagram is equal to the
intensity of the shear force V), we can observe that the slope of the M diagram must be constant
between points i, j, k, and l, because the shear force is constant in the corresponding regions.
The slope of the M diagram between points i and j is +14kips, the M slope between points j
and k is +2kips, and the M slope between points k and l is –8kips. The only type of curve that
has a constant slope is a line.
The maximum shear force is V = 14kips. The maximum bending moment is M = +72kip.ft, at
x = 12 ft. Notice that the maximum bending moment occurs where the shear force diagram
crosses the V = 0 axis (between points e and f).

GIRUM MINDAYE 52
Strength of Materials AASTU

3 STRESSES IN BEAMS
3.1 Theory of Pure Bending

Pure bending (Theory of pure bending) is a condition of stress where a bending moment is
applied to a beam without the simultaneous presence of axial, shear, or torsional forces. The
followings are the assumptions in theory of pure bending:
 The material of the beam is isotropic and homogeneous and follows Hook’s law.
 The stress induced is proportional to the strain and the stress at any point does not
exceed the elastic limit.
 Transverse sections of the beam that were plane before bending remains plane even
after bending.
 The beam is initially straight and having uniform cross section.
 The modulus of elasticity is same for the fibers of the beam under tension or
compression.
 The beam is subjected to pure bending and therefore bends in an arc of a circle.
 The radius of curvature is large compared to the dimensions of the section.
 There is no resultant pull or push on the cross section of the beam.
 The loads are applied in the plane of bending. The beam has an axial plane of
symmetry, which we take to be the xy plane (see Figure 3-1). The applied loads (such
as F1, F2 and F3 in Figure 3-1) lie in the plane of the symmetry and are perpendicular
to the axis of the beam (the x-axis).The axis of the beam bends but does not stretch.
 The transverse section of the beam is symmetrical about a line passing through the
centre of gravity in the plane of bending.
The above assumptions lead us to the following conclusion:
Each cross section of the beam rotates as a rigid entity about a line called the neutral axis of
the cross section. The neutral axis passes through the axis of the beam and is perpendicular to
the plane of symmetry, as shown in Figure 3-1. The xz-plane that contains the neutral axes of
all the cross sections is known as the neutral surface of the beam.

Figure 3-1

GIRUM MINDAYE 53
Strength of Materials AASTU

3.2 Flexure Stress

The stresses caused by the bending moment are known as bending stress, or flexure stresses.
The relationship between these stresses and the bending moment is called the flexure formula.
Consider a prismatic member AB possessing a plane of symmetry and subjected to equal and
opposite couples M acting in that plane (Figure 3-2a). If a section is passed through the member
AB at some arbitrary point C, the conditions of equilibrium of the portion AC of the member
require the internal forces in the section to be equivalent to the couple M (Figure 3-2b).

Figure 3-2: (a) A member in a state of pure bending. (b) Any intermediate portion of AB will
also be in pure bending.

Figure 3-3: Initially straight members in pure bending deform into a circular arc
We will now analyze the deformations of a prismatic member possessing a plane of symmetry.
Its ends are subjected to equal and opposite couple M acting in the plane of symmetry. The
member will bend under the action of the couples, but will remain symmetric with respect to
that plane (Figure 3-3). Moreover, since the bending moment M is the same in any cross
section, the member will bend uniformly. Thus, the line AB along the upper face of the member
intersecting the plane of the couples will have a constant curvature. In other words, the line AB
will be transformed into a circle of center C, as will the line A’B’ along the lower face of the
member. Note that the line AB will decrease in length when the member is bent (i.e., when M
> 0), while A’B’ will become longer.

GIRUM MINDAYE 54
Strength of Materials AASTU

Figure 3-4: Establishment of neutral axis. (a) Longitudinal-vertical view. (b) Transverse
section at origin.
It follows from above that a surface parallel to the upper and lower faces of the member must
exist where ϵx and σx are zero. This surface is called the neutral surface. The neutral surface
intersects the plane of symmetry along an arc of circle DE (Figure 3-4a), and it intersects a
transverse section along a straight line called the neutral axis of the section (Figure 3-4b). The
origin of coordinates is now selected on the neutral surface rather than on the lower face of the
member so that the distance from any point to the neutral surface is measured by its coordinate
y.
Denoting by ρ the radius of arc DE (Figure 3-4a), by θ the central angle corresponding to DE,
and observing that the length of DE is equal to the length L of the undeformed member, we
write
𝐿 = 𝜌𝜃
Considering the arc JK located at a distance y above the neutral surface, its length L’ is
𝐿′ = (𝜌 − 𝑦)𝜃
Since the original length of arc JK was equal to L, the deformation of JK is
𝛿 = 𝐿′ − 𝐿 = (𝜌 − 𝑦)𝜃 − 𝜌𝜃 = −𝑦𝜃
The longitudinal strain ϵx in the elements of JK is obtained by dividing δ by the original length
L of JK. Write
𝛿 𝑦𝜃
𝜖𝑥 = =−
𝐿 𝜌𝜃
𝑦
𝜖𝑥 = −
𝜌
The minus sign is due to the fact that it is assumed the bending moment is positive, and thus
the beam is concave upward. Because of the requirement that transverse sections remain plane,
identical deformations occur in all planes parallel to the plane of symmetry. Thus, the
longitudinal normal strain ϵx varies linearly with the distance y from the neutral surface.

GIRUM MINDAYE 55
Strength of Materials AASTU

Figure 3-5: Normal strain vary linearly with distance from the neutral axis.
The strain ϵx reaches its maximum absolute value when y is largest. Denoting the largest
distance from the neutral surface as c (corresponding to either the upper or the lower surface
of the member) and the maximum absolute value of the strain as σm, we have
𝑐
𝜖𝑚 =
𝜌
𝑐
𝑠𝑖𝑛𝑐𝑒, 𝜌 = −𝑦⁄𝜖𝑥 ⟹ 𝜖𝑚 =
(−𝑦⁄𝜖𝑥 )
𝑦
𝜖𝑥 = − 𝜖𝑚
𝑐
𝑓𝑟𝑜𝑚 𝐻𝑜𝑜𝑘𝑒 ′ 𝑠 𝑙𝑎𝑤, 𝜎𝑥 = 𝐸𝜖𝑥 𝑎𝑛𝑑 𝜖𝑚 = 𝜎𝑚 ⁄𝐸 :
𝑦 𝑦 𝜎𝑚 𝑦
𝜎𝑥 = 𝐸𝜖𝑥 = 𝐸 (− 𝜖𝑚 ) = 𝐸 [− ( )] = − 𝜎𝑚
𝑐 𝑐 𝐸 𝑐
𝒚
𝝈𝒙 = − 𝝈𝒎
𝒄

This equation describes the stress distribution over the cross-sectional area. The sign
convention established here is significant. For positive M, which acts in the +z direction,
positive values of y give negative values for σx, that is, a compressive stress, since it acts in the
negative x direction. Similarly, negative y values will give positive or tensile values for σx.

Figure 3-6: Bending stresses vary linearly with distance from the neutral axis.
This result shows that, in the elastic range, the normal stress varies linearly with the distance
from the neutral surface (Figure 3-6).
Internal Moment and Stress Relations: On any point on the cross section, we have σx, the
normal stress, and τxy and τxz, the components of the shearing stress. The system of these

GIRUM MINDAYE 56
Strength of Materials AASTU

elementary internal forces exerted on the cross section is equivalent to the couple M (Figure
3-7).

Figure 3-7: Stresses resulting from pure bending moment M


Selecting arbitrarily the z-axis shown in Figure 3-7, the equivalence of the elementary internal
forces and the couple M is expressed by writing that the sums of the components and moments
of the forces are equal to the corresponding components and moments of the couple M:

𝑥 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡𝑠: ∫ 𝜎𝑥 𝑑𝐴 = 0

𝑀𝑜𝑚𝑒𝑛𝑡𝑠 𝑎𝑏𝑜𝑢𝑡 𝑦 𝑎𝑥𝑖𝑠: ∫ 𝑧𝜎𝑥 𝑑𝐴 = 0

𝑀𝑜𝑚𝑒𝑛𝑡𝑠 𝑎𝑏𝑜𝑢𝑡 𝑧 𝑎𝑥𝑖𝑠: ∫(−𝑦𝜎𝑥 )𝑑𝐴 = 𝑀

The minus sign in moments about z axis is due to the fact that a tensile stress (σx> 0) leads to a
negative moment (clockwise) of the normal force σxdA about the z axis.
Three additional equations could be obtained by setting equal to zero the sums of the y
components, z components, and moments about the x-axis, but these equations would involve
only the components of the shearing stress and the components of the shearing stress are both
equal to zero.
Location of Neutral Axis: To locate the position of the neutral axis, we require the resultant
force produced by the stress distribution acting over the cross-sectional area to be equal to zero.
Noting that the force dP = σdA acts on the arbitrary element dA, we have

∑ 𝐹𝑥 = 0, ∫ 𝜎𝑥 𝑑𝐴 = 0

𝑦
∫ (− 𝜎𝑚 ) 𝑑𝐴 = 0
𝑐
𝜎𝑚
− ∫ 𝑦𝑑𝐴 = 0
𝑐
∫ 𝑦𝑑𝐴 = 0

GIRUM MINDAYE 57
Strength of Materials AASTU

This equation shows that the first moment of the cross section about its neutral axis must be
zero. It can be zero only if the neutral axis passes through centroid C of the cross-sectional
area. Thus, for a member subjected to pure bending and as long as the stresses remain in the
elastic range, the neutral axis passes through the centroid of the section. Hence, once the
centroid for the member’s cross-sectional area is determined, the location of the neutral axis is
known.
Bending Moment: We can determine the stress in the beam if we require the moment M to be
equal to the moment produced by the stress distribution about the neutral axis.

↺ + ∑ 𝑀𝑧 = 𝑀, ∫(−𝑦𝜎𝑥 )𝑑𝐴 = 𝑀
𝑦 𝜎𝑚
∫(−𝑦) (− 𝜎𝑚 ) 𝑑𝐴 = ∫ 𝑦 2 𝑑𝐴 = 𝑀
𝑐 𝑐
Recall that for pure bending the neutral axis passes through the centroid of the cross section
and ∫ 𝑦 2 𝑑𝐴 = 𝐼 is the moment of inertia or second moment of area of the cross section with
respect to a centroidal axis perpendicular to the plane of the couple M.
𝜎𝑚
𝐼=𝑀
𝑐
Solving for σm,
𝑀𝑐
𝜎𝑚 =
𝐼
We obtain the normal stress σx at any distance y from the neutral axis:
𝑦 𝑦 𝑀𝑐
𝜎𝑥 = − 𝜎𝑚 = − ( )
𝑐 𝑐 𝐼
𝑴𝒚
𝝈𝒙 = −
𝑰
This equation, called the flexure formula, shows that the stresses are directly proportional to
the bending moment M and inversely proportional to the moment of inertia I of the cross
section. Also, the stresses vary linearly with the distance y from the neutral axis, as previously
observed. Stresses calculated from the flexure formula are called bending stresses or flexural
stresses.

Figure 3-8: Relationships between signs of bending moments and directions of normal
stresses: (a) positive bending moment and (b) negative bending moment

GIRUM MINDAYE 58
Strength of Materials AASTU

If the bending moment in the beam is positive, the bending stresses will be positive (tension)
over the part of the cross section where y is negative, that is, over the lower part of the beam.
The stresses in the upper part of the beam will be negative (compression). If the bending
moment is negative, the stresses will be reversed. These relationships are shown in Figure 3-8.
Maximum Stresses at a Cross Section: The maximum tensile and compressive bending
stresses acting at any given cross section occur at points located farthest from the neutral axis.
Denote by c1 and c2 the distances from the neutral axis to the extreme elements in the positive
and negative y directions, respectively (see Figure 3-8). Then the corresponding maximum
normal stresses
𝑴𝒄𝟏 𝑴𝒄𝟐
𝝈𝟏 = − , 𝝈𝟐 = −
𝑰 𝑰
The ratio I/c depends only on the geometry of the cross section. This ratio is defined as the
elastic section modulus S, where
𝐼
𝐸𝑙𝑎𝑠𝑡𝑖𝑐 𝑠𝑒𝑐𝑡𝑖𝑜𝑛 𝑚𝑜𝑑𝑢𝑙𝑢𝑠, 𝑆=
𝑐
𝑀 𝑴 𝑀 𝑴
𝝈𝟏 = − =− , 𝝈𝟐 = − =−
𝐼 ⁄𝑐1 𝑺𝟏 𝐼 ⁄𝑐2 𝑺𝟐
𝐼 𝐼
𝑤ℎ𝑒𝑟𝑒, 𝑆1 =
𝑎𝑛𝑑 𝑆2 =
𝑐1 𝑐2
The quantities S1 and S2 are known as the section moduli of the cross-sectional area.
Doubly Symmetric Shapes: If the cross section of a beam is symmetric with respect to the z-
axis as well as the y-axis (doubly symmetric cross section), then c1=c2= c and the maximum
tensile and compressive stresses are equal numerically:
𝑴𝒄 𝑴 𝑴𝒄 𝑴
𝝈𝟏 = −𝝈𝟐 = − =− , 𝒐𝒓 𝝈𝒎𝒂𝒙 = =
𝑰 𝑺 𝑰 𝑺
Table 3-1: The section moduli of standard structural shapes

Shape Representation Moment of Inertia (I) Section Modulus (S)


𝑏ℎ3
𝐼𝑥 = 𝑏ℎ2
12 𝑆=
ℎ𝑏 3 6
𝐼𝑦 =
12
Rectangle

𝜋𝑟 4 𝜋𝑑 4 𝜋𝑟 3 𝜋𝑑3
Circle 𝐼𝑥 = 𝐼𝑦 = = 𝑆= =
4 64 4 32

GIRUM MINDAYE 59
Strength of Materials AASTU

𝜋(𝑟 4 0 − 𝑟 4 𝑖 ) 𝜋(𝑟 4 0 − 𝑟 4 𝑖 )
𝐼𝑥 = 𝐼𝑦 = 𝑆=
4 4𝑟0
Circular Or Or
Tube
𝜋(𝑑 4 0 − 𝑑 4 𝑖 ) 𝜋(𝑑 4 0 − 𝑑4 𝑖 )
𝐼𝑥 = 𝐼𝑦 = 𝑆 =
64 32𝑑0
1 ℎ3 𝑤
𝑆 = (𝑏ℎ2 − (𝑏
6 ℎ
𝑏ℎ3 (𝑏 − 𝑡𝑤 )ℎ3 𝑤
I-Beam 𝐼𝑥 = −
12 12 − 𝑡𝑤 ))

𝑏ℎ3 𝑏ℎ2
𝐼𝑥 = 𝑆=
36 24
Triangle

The following are the procedures for determination of Bending/flexure Stresses:

a) Stress at a Given Point


 Use the method of sections to determine the bending moment M at the cross section
containing the given point.
 Determine the location of the neutral axis.
 Compute the moment of inertia I of the cross- sectional area about the neutral axis.
 Determine the y-coordinate of the given point. Note that y is positive if the point lies
above the neutral axis and negative if it lies below the neutral axis.
 Compute the bending stress from σ=-My/I If correct sign are used for M and y, the stress
will also have the correct sign (tension positive, compression negative).
b) Maximum Bending Stress: Symmetric Cross Section
If the neutral axis is an axis of symmetric of the cross section, the maximum tensile and
compression, bending stresses are equal in magnitude and occur at the section of the largest
bending moment. The following procedure is recommended for determining the maximum
bending stress in a prismatic beam:
 Draw the bending moment diagram. Identify the bending moment Mmax that has the largest
magnitude (disregard the sign)
 Compute the moment of inertia I of the cross- sectional area about the neutral axis
 Calculate the maximum bending stress from σmax= [Mmax]c/I, where c is the distance from
the neutral axis to the top or bottom of the cross section
c) Maximum Tensile and Compressive Bending Stresses: Unsymmetrical Cross Section
If the neutral axis is not an axis of symmetry of the cross section, the maximum tensile and
compressive bending stresses may occur at different sections.
 Draw the bending moment diagram. Identify the largest positive and negative bending
moments.

GIRUM MINDAYE 60
Strength of Materials AASTU

 Determine the location of the neutral axis and record the distance ctop and cbot from the
neutral axis to the top and bottom of the cross section.
 Compute the moment of inertia I of the cross section about the neutral axis.
 Calculate the bending stresses at the top and bottom of the cross section where the largest
positive bending moment σ= -My/ I.
 At the top of the cross section, where y = ctop ,we obtain σtop =-M ctop/ I
 At the bottom of the cross section, where y = -cbot ,we obtain σbot =M cbot/ I
 Repeat the calculations for the cross section that carries the largest negative bending
moment.
 Inspect the four stresses thus computed to determine the largest tensile (positive) and
compressive (negative) bending stresses in the beam
Curvature: The deformation of the member caused by the bending moment M is measured by
the curvature of the neutral surface. The curvature is defined as the reciprocal of the radius of
curvature ρ and can be obtained by
1 𝜖𝑚 (𝜎𝑚 ⁄𝐸 ) 𝜎𝑚 (𝑀𝑐 ⁄𝐼 )
= = = =
𝜌 𝑐 𝑐 𝐸𝑐 𝐸𝑐
𝟏 𝑴
∴ =
𝝆 𝑬𝑰
This equation known as the moment-curvature equation, shows that the curvature is directly
proportional to the bending moment M and inversely proportional to the quantity EI, which is
called the flexural rigidity of the beam. Flexural rigidity is a measure of the resistance of a
beam to bending, that is, the larger the flexural rigidity, the smaller the curvature for a given
bending moment.
Comparing the sign convention for bending moments with that for curvature, note that a
positive bending moment produces positive curvature and a negative bending moment
produces negative curvature (see Figure 3-9).

Figure 3-9: Relationships between signs of bending moments and signs of curvatures

3.3 Shear Stress

The applied shear force will induce shear stress across transverse section of the beam. At each
point on a section, the transverse shear stress will produce a complementary horizontal shear
stress. The longitudinal shear stresses will balance the variation of bending stresses along the
beam. If the bending moment is constant, there is no shear force and hence no shear stress. If
there is no variation of bending stress between successive transverse sections, there can be no
longitudinal shear stresses.

GIRUM MINDAYE 61
Strength of Materials AASTU

Shearing stresses are important, particularly in the design of short, stubby beams. Figure 3-10
graphically expresses the elementary normal and shearing forces exerted on a transverse
section of a prismatic beam with a vertical plane of symmetry that are equivalent to the bending
couple M and the shearing force V. Six equations can be written to express this. Three of these
equations involve only the normal forces σdA and have been discussed in Sec. 3.1. These are
the sum of the normal forces is zero and that the sums of their moments about the y and z-axes
are equal to zero and M, respectively. Three more equations involving the shearing forces τxydA
and τxzdA now can be written. One equation expresses that the sum of the moments of the
shearing forces about the x-axis is zero and can be dismissed as trivial in view of the symmetry
of the beam with respect to the xy plane. The other two involve the y and z components of the
elementary forces and are

𝑥 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡𝑠: ∫ 𝜏𝑥𝑦 𝑑𝐴 = −𝑉

𝑧 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡𝑠: ∫ 𝜏𝑥𝑧 𝑑𝐴 = 0

Figure 3-10: All the stresses on elemental areas (left) sum to give the resultant shear V and
bending moment M.
Consider a prismatic beam AB with a vertical plane of symmetry that supports various
concentrated and distributed loads (Figure 3-11).

Figure 3-11: Transversely loaded beam with vertical plane of symmetry.


At a distance x from end A, we detach from the beam an element CDD’C’ with length of Δx
extending across the width of the beam from the upper surface to a horizontal plane located at
a distance y1 from the neutral axis (Figure 3-12).

GIRUM MINDAYE 62
Strength of Materials AASTU

Figure 3-12: Short segment of beam with stress element CDD’C’ defined.
The forces exerted on this element consist of vertical shearing forces V’C and V’D, and a
horizontal shearing force ΔH exerted on the lower face of the element, elementary horizontal
normal forces σCdA and σDdA, and possibly a load wΔx (Figure 3-13).

Figure 3-13: Forces exerted on element CCD‘C’.


The equilibrium equation for horizontal forces is
.
→ + ∑ 𝐹𝑥 = 0, ∆𝐻 + ∫ (𝜎𝐶 − 𝜎𝐷 )𝑑𝐴 = 0

where the integral extends over the shaded area α of the section located above the line y=y1.
Solving this equation for ΔH and using, σ=My/I, to express the normal stresses in terms of the
bending moments at C and D, provides
𝑀𝐷 − 𝑀𝐶 .
∆𝐻 = ∫ 𝑦𝑑𝐴
𝐼 ℴ
.
∫ℴ 𝑦𝑑𝐴 represents the first moment with respect to the neutral axis of the portion α of the cross
section of the beam that is located above the line y = y1 and will be denoted by Q. On the other
hand, the increment MD-MC of the bending moment is
𝑀𝐷 − 𝑀𝐶 = ∆𝑀 = (𝑑𝑀⁄𝑑𝑥)∆𝑥 = 𝑉∆𝑥
𝑀𝐷 − 𝑀𝐶 . ∆𝑀 𝑉𝑄
∆𝐻 = ∫ 𝑦𝑑𝐴 = 𝑄= ∆𝑥
𝐼 ℴ 𝐼 𝐼

The horizontal shear per unit length, which will be denoted by q, is obtained by dividing both
sides by Δx:
∆𝐻 𝑉𝑄
𝑞= =
∆𝑥 𝐼
Recall the at Q is the first moment with respect to the neutral axis of the portion of the cross
section located either above or below the point at which q is being computed and that I is the

GIRUM MINDAYE 63
Strength of Materials AASTU

centroidal moment of inertia of the entire cross-sectional area. The horizontal shear per unit
length q is also called the shear flow.
Consider again a beam with a vertical plane of symmetry that is subjected to various
concentrated or distributed loads applied in that plane. If, through two vertical cuts and one
horizontal cut, an element of length Δx is detached from the beam (Figure 3-13), the magnitude
ΔH of the shearing force exerted on the horizontal face of the element can be obtained by ∆𝐻 =
𝑉𝑄
𝐼
∆𝑥. The average shearing stress τave on that face of the element is obtained by dividing ΔH
by the area ΔA of the face. Observing that ΔA = tΔx, where t is the width of the element at the
cut, we write
𝑉𝑄 ∆𝑥 𝑉𝑄
𝜏𝑎𝑣𝑒 = =
𝐼 𝑡∆𝑥 𝐼𝑡
This equation is used in practical applications to determine the shearing stress at any point of
the cross section of a narrow rectangular beam (i.e., a beam of rectangular section of width b
and depth h with b ≤ 0.25h) because of the variation of the shearing stress τxy across the width
of the beam is less than 0.8% of τave.

Figure 3-14: Shearing stress distribution on transverse section of rectangular beam.

Observing that c=h/2, and the distance from the neutral axis to the centroid of A’ is 𝑦̅ = 𝑦 +
1 1
(𝑐 − 𝑦) = (𝑐 + 𝑦).
2 2

Recalling that 𝑄 = 𝐴′𝑦̅,


1 1
𝑄 = 𝐴′ 𝑦̅ = 𝑏(𝑐 − 𝑦) (𝑐 + 𝑦) = 𝑏(𝑐 2 − 𝑦 2 )
2 2
𝑏ℎ3 2 3
ℎ = 2𝑐, 𝑎𝑛𝑑 𝑟𝑒𝑐𝑎𝑙𝑙𝑖𝑛𝑔 𝑡ℎ𝑎𝑡 𝐼 = = 𝑏𝑐
12 3
𝑉𝑄 𝑉𝑄 3 𝑐2 − 𝑦2
𝜏𝑎𝑣𝑒 = = = 𝑉
𝐼𝑡 𝐼𝑏 4 𝑏𝑐3
or noting that the cross-sectional area of the beam is A= 2bc,
3 𝑐2 − 𝑦2 3 𝑐2 − 𝑦2 3𝑉 𝑦2
𝜏𝑎𝑣𝑒 = 𝑉 = 𝑉 = (1 − )
4 𝑏𝑐3 2 (2𝑏𝑐)𝑐2 2𝐴 𝑐2
As we can observe from this equation, the shearing stresses are zero at the top and bottom of
the cross section (y=±c). Making y=0, the value of then maximum shearing stress in a given
section of a narrow rectangular beam is

GIRUM MINDAYE 64
Strength of Materials AASTU

3𝑉
𝜏𝑎𝑣𝑒 =
2𝐴
This relationship shows that the maximum value of the shearing stress in a beam of rectangular
cross section is 50% larger than the value V/A obtained by wrongly assuming a uniform stress
distribution across the entire cross section.
The following are the procedure for analysis of shear stress:
 Use equilibrium analysis to determine the vertical shear force V acting on the cross
section containing the specified point ( the construction of a shear force diagram is
usually a goodidea).
 Locate the neutral axis and compute the moment of inertia I of the cross- sectional area
about the neutral axis (If the beam is a standard structural shape, its cross- sectional
properties are listed in steel section table.)
 Compute the first moment Q of the cross- sectional area that lies above (or below)the
specified point.
 Calculate the shear stress from τ = VQ/(Ib), where b is the width of the cross section at
the specified point.
 The maximum shear stress τmax on a given cross section occurs where Q/b is largest.
 If the width b is constant, then τmax occurs at the neutral axis because that is where Q
has its maximum value.
 If b is not constant, it is necessary to compute the shear stress at more than one point in
order to determine its maximum value.
3.4 Worked Examples

Example 3.1: The simply supported beam has a rectangular cross section 120 mm wide and
200 mm high. Compute the maximum bending stress in the beam and the bending stress at a
point on section B that is 25 mm below the top of the beam.

Solution
Preliminary Calculations
The shear force and bending moment diagrams.

GIRUM MINDAYE 65
Strength of Materials AASTU

Mmax = +16kNm, occurring at D. Because Mmax is positive, the top half of the cross section is
in compression and the bottom half is in tension.
The neutral axis (NA) is an axis of symmetry of the cross section. Due to symmetry of the cross
section about the neutral axis, the maximum tensile and compressive stresses are equal in
magnitude.

The distance c between the neutral axis and the top (or bottom) of the cross section is c =
100mm = 0.1m.
The moment of inertia of the cross section about the neutral axis is
𝑏ℎ3 0.12 × 0.23
𝐼= = = 800 × 10−6 𝑚4
12 12
The maximum bending stress in the beam on the cross section that carries the largest bending
moment, which is the section at D.
𝑀𝑚𝑎𝑥 𝑐 (16 × 103 ) × 0.1
𝜎𝑚𝑎𝑥 = = = 20 × 106 𝑃𝑎 = 20𝑀𝑃𝑎
𝐼 800 × 10−6
From Fig. (c) we see that the bending moment at section B is M = + 9.28kNm. The y-coordinate
of the point that lies 25 mm below the top of the beam is y = 100-25 = 75mm = 0.075m.
𝑀𝐵 𝑦 (9.28 × 103 ) × 0.075
𝜎𝐵 = − = = −8.70 × 106 𝑃𝑎 = −8.70𝑀𝑃𝑎
𝐼 800 × 10−6

GIRUM MINDAYE 66
Strength of Materials AASTU

Example 3.2: A cantilever beam, 50 mm wide by 150 mm high and 6 m long, carries a load
that varies uniformly from zero at the free end to 1000 N/m at the wall. (a) Compute the
magnitude and location of the maximum flexural stress. (b) Determine the type and magnitude
of the stress in a fiber 20 mm from the top of the beam at a section 2 m from the free end.
Solution

𝑦 1000 500 1 1 500 250 2


= ⟹𝑦= 𝑥, 𝐹 = (𝑥𝑦) = [𝑥 ( 𝑥)] = 𝑥
𝑥 6 3 3 2 3 3
1 250 2 1 250 3
↺ + ∑ 𝑀 = 0, 𝑀𝑥 = −𝐹 ( 𝑥) = − ( 𝑥 ) ( 𝑥) = − 𝑥
3 3 3 9
Part (a):
The maximum moment occurs at the support (the wall) or at x = 6m.
250 3
𝑀𝑚𝑎𝑥 = − (6 ) = −6,000𝑁𝑚
9

𝑀𝑚𝑎𝑥 𝑐 (−6000 × 1000) × 75


𝜎𝑚𝑎𝑥 = − =− = 32𝑀𝑃𝑎
𝐼 50 × 1503 ⁄12
Part (b):
At a section 2 m from the free end or at x = 2 m at fiber 20mm from the top of the beam:
250 3 2000
𝑀𝑥=2 = − (2 ) = − 𝑁𝑚
9 9

2000
𝑀𝑦 (− 9 × 1000) × 55
𝜎𝑥=2𝑚,𝑦=55𝑚𝑚 =− =− = 0.8691𝑀𝑃𝑎 = 869.1𝑘𝑃𝑎
𝐼 50 × 1503 ⁄12
Example 3.3: Determine the minimum height h of the beam shown in figure if the flexural
stress is not to exceed 20 MPa.

GIRUM MINDAYE 67
Strength of Materials AASTU

Solution:
1
𝑐= ℎ,
2
𝑏ℎ3 80ℎ3 20ℎ3
𝐼= = =
12 12 3
𝑀𝑚𝑎𝑥 𝑐 (5 × 106 )(ℎ⁄2)
𝜎𝑚𝑎𝑥 =− → 20 = −
𝐼 20ℎ3 ⁄3
2
ℎ = 18750 → ℎ = 137𝑚𝑚

Example 3.4: The simply supported beam has the T-shaped cross section as shown in figure.
Determine the values and locations of the maximum tensile and compressive bending stresses.

Solution
a) Find the largest positive and negative bending moment.

GIRUM MINDAYE 68
Strength of Materials AASTU

The results are shown in figure. The largest positive and negative bending moment are
3200lb·ft and 4000 lb.ft respectively.
b) Compute the moment of inertia I about the neutral axis.
The cross section to be composed of the two rectangles with areas A1= 0.8 (8) = 6.4in2 and A2=
0.8(6) = 4.8in2. The centroidal coordinates of the areas are ̅̅̅
𝑦1 = 4𝑖𝑛 and ̅̅̅
𝑦2 = 8.4𝑖𝑛, measured
from the bottom of the cross section. The coordinate 𝑦̅ of the centroid C of the cross section is
𝐴1 ̅̅̅
𝑦1 + 𝐴2 ̅̅̅
𝑦2 6.4 × 4 + 4.8 × 8.4
𝑦̅ = = = 5.886𝑖𝑛
𝐴1 + 𝐴2 6.4 + 4.8
Compute the moment of inertia I of the cross-sectional area about the neutral axis. Using the
parallel-axis theorem, 𝐼 = ∑[𝐼̅𝑖 + 𝐴𝑖 (𝑦̅𝑖 − 𝑦̅)2 ] where 𝐼̅𝑖 = 𝑏𝑖 ℎ𝑖 3 ⁄12 is the moment of inertia
of a rectangle about its own centroidal axis. Thus,

0.8 × 83 2
6 × 0.83
𝐼=[ + 6.4(4 − 5.886) ] + [ + 4.8(8.4 − 5.886)2 ] = 87.49𝑖𝑛4
12 12
c) Maximum Bending stresses
The distances from the neutral axis to the top and the bottom of the cross section are
𝑐𝑡𝑜𝑝 = 8.8 − 𝑦̅ = 8.8 − 5.886 = 2.914𝑖𝑛 and 𝑐𝑏𝑜𝑡 = 𝑦̅ = 5.886𝑖𝑛. Because these distances
are different, we must investigate stresses at two locations: at x = 4ft (where the largest positive
bending moment occurs) and at x = 10ft (where the largest negative bending moment occurs).
Stresses at x = 4ft: The bending moment at this section is M = +3200 lb.ft causing compression
above the neutral axis and tension below the axis. The resulting bending stresses at the top and
bottom of the cross section are
𝑀𝑐𝑡𝑜𝑝 (3200 × 12) × 2.914
𝜎𝑡𝑜𝑝 = − =− = −1279𝑃𝑠𝑖
𝐼 87.49
𝑀𝑐𝑏𝑜𝑡 (3200 × 12)(−5.886)
𝜎𝑏𝑜𝑡 = − =− = 2580𝑃𝑠𝑖
𝐼 87.49
Stresses at x = 10 ft: The bending moment at this section is M = -4000lb.ft, resulting in tension
above the neutral axis and compression below the neutral axis. The corresponding bending
stresses at the extremities of the cross section are
𝑀𝑐𝑡𝑜𝑝 (−4000 × 12) × 2.914
𝜎𝑡𝑜𝑝 = − =− = 1599𝑃𝑠𝑖
𝐼 87.49

GIRUM MINDAYE 69
Strength of Materials AASTU

𝑀𝑐𝑏𝑜𝑡 (−4000 × 12)(−5.886)


𝜎𝑏𝑜𝑡 = − =− = −3230𝑃𝑠𝑖
𝐼 87.49
Inspecting the above results, we conclude that the maximum tensile and compressive stresses
in the beam are
𝜎𝑏𝑜𝑡 = 2580𝑃𝑠𝑖 ( 𝑏𝑜𝑡𝑡𝑜𝑚 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑒𝑐𝑡𝑖𝑜𝑛 𝑎𝑡 𝑥 = 4 𝑓𝑡 )
𝜎𝑏𝑜𝑡 = −3230𝑃𝑠𝑖 ( 𝑏𝑜𝑡𝑡𝑜𝑚 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑒𝑐𝑡𝑖𝑜𝑛 𝑎𝑡 𝑥 = 10 𝑓𝑡 )
Example 3.5: The cantilever beam is composed of two segments with rectangular cross
sections. The width of the each section is 2in, but the depths are different, as shown in the
figure. Determine the maximum bending stress in the beam.

Solution:
Because the cross section of the beam is not constant, the maximum stress occurs either at the
section just to the left of MB= - 8000lb.ft) or at the section at D (MD= -16000lb.ft).
The section moduli of the two segments are
𝑏(ℎ𝐴𝐵 )2 2 × 42 16 3
𝑆𝐴𝐵 = = = 𝑖𝑛
6 6 3
𝑏(ℎ𝐵𝐷 )2 2 × 62
𝑆𝐵𝐷 = = = 12𝑖𝑛3
6 6
The maximum bending stresses on the two cross sections of the interest are
[𝑀𝐵 ] (8000 × 12)
(𝜎𝐵 )𝑚𝑎𝑥 = = = 18000𝑃𝑠𝑖
𝑆𝐴𝐵 16⁄3
[𝑀𝐷 ] (16000 × 12)
(𝜎𝐷 )𝑚𝑎𝑥 = = = 16000𝑃𝑠𝑖
𝑆𝐵𝐷 12
Comparing the above values, we find that the maximum bending stress in the beam is
(𝜎𝐵 )𝑚𝑎𝑥 = 18000𝑃𝑠𝑖 (𝑜𝑛 𝑡ℎ𝑒 𝑐𝑟𝑜𝑠𝑠 𝑠𝑒𝑐𝑡𝑖𝑜𝑛 𝑗𝑢𝑠𝑡 𝑡𝑜 𝑡ℎ𝑒 𝑙𝑒𝑓𝑡 𝑜𝑓 𝐵)

GIRUM MINDAYE 70
Strength of Materials AASTU

This is an example where the maximum bending stress occurs on a cross section at the bending
moment is not maximum.
Example 3.6: The simply supported beam has the cross-sectional area shown in figure.
Determine the absolute maximum bending stress in the beam and draw the stress distribution
over the cross section at this location. Also, what is the stress at point B?

Solution
(a) Maximum Internal Moment.
The maximum internal moment in the beam, M = 22.5kNm, occurs at the center, as indicated
on the moment diagram, Fig. c.

(b) Section Property.


By reasons of symmetry, the neutral axis passes through the centroid C at the mid height of the
beam, Fig. b. The area is subdivided into the three parts shown, and the moment of inertia of
each part is calculated about the neutral axis using the parallel-axis theorem. Choosing to work
in millimeters, we have

𝐼 = ∑ 𝐼 ̅ + 𝐴𝑑 2
20 × 3003 250 × 203 20 2
𝐼=[ ] + 2[ + (250 × 20) (150 + ) ] = 301 × 106 𝑚𝑚4
12 12 2

(c) Applying the flexural formula


𝑀𝑚𝑎𝑥 𝑐 (22.5 × 106 )(170)
𝜎𝑚𝑎𝑥 = = = 12.71𝑀𝑃𝑎
𝐼 301 × 106
A three-dimensional view of the stress distribution is shown in Fig. d. Specifically, at point B,
yB= 150 mm, and so as shown in Fig. d,

GIRUM MINDAYE 71
Strength of Materials AASTU

𝑀𝑚𝑎𝑥 𝑦𝐵 (22.5 × 106 )(150)


𝜎𝐵 = − =− = −11.21𝑀𝑃𝑎
𝐼 301 × 106

Example 3.7: Determine (a) the moment M that will produce a maximum stress of 70MPa on
the cross section and (b) the maximum tensile and compressive bending stress in the beam if it
is subjected to a moment of M = 6kNm.

Solution
(a) Section Property.
The coordinate 𝑦̅ of the centroid of the cross section from the bottom is
∑ 𝐴𝑖 𝑦̅𝑖
𝑦̅𝑏𝑜𝑡 =
∑ 𝐴𝑖
(12 × 250)(250⁄2) + (99 × 12)(250 + 12⁄2) + 2(12 × 75)(250 − 75⁄2)
=
(12 × 250) + (99 × 12) + 2(12 × 75)

𝑦̅𝑏𝑜𝑡 = 177.29𝑚𝑚

𝑦̅𝑡𝑜𝑝 = 250 + 12 − 𝑦̅𝑏𝑜𝑡 = 250 + 12 − 177.29 = 84.71𝑚𝑚

Compute the moment of inertia I of the cross-sectional area about the neutral axis.

GIRUM MINDAYE 72
Strength of Materials AASTU

𝐼 = ∑ 𝐼 ̅ + 𝐴𝑑 2
12 × 2503 250 2 99 × 123
𝐼=[ + (12 × 250) (177.29 − ) ]+[ + (99 × 12)(84.71 − 6)2 ]
12 2 12
12 × 753 75 2
+ 2[ + (12 × 75) (84.71 − 12 − ) ] = 34.2773 × 106 𝑚𝑚4
12 2

(b) Applying the flexural formula


 Moment M that will produce a maximum stress of 70MPa
𝑀𝑐 (𝑀)(177.29)
𝜎= ⟹ 70 = ⟹ 𝑀 = 13.53 × 106 𝑁𝑚𝑚 = 13.53𝑘𝑁𝑚
𝐼 34.2773 × 106
 Maximum tensile and compressive bending stress if M = 6kNm.
𝑀𝑦̅𝑏𝑜𝑡 (6 × 106 )(−177.29)
(𝜎𝑏𝑜𝑡 )𝑚𝑎𝑥 =− =− = 31.03𝑀𝑃𝑎 (𝑇)
𝐼 34.2773 × 106
𝑀𝑦̅𝑡𝑜𝑝 (6 × 106 )(84.71)
(𝜎𝑡𝑜𝑝 )𝑚𝑎𝑥 = − =− = −14.83𝑀𝑃𝑎 (𝐶)
𝐼 34.2773 × 106
Example 3.8: If the beam is subjected to an internal moment of M = 30kNm, determine (a) the
maximum bending stress in the beam, and (b) the resultant force caused by the bending stress
distribution acting on the top flange A.

Solution
(a) Section Property.
Because of doubly symmetry of the section the centroid 𝑦̅ = ℎ⁄2 = 75𝑚𝑚
𝑦̅𝑡𝑜𝑝 = 𝑦̅𝑏𝑜𝑡 = 75𝑚𝑚

Compute the moment of inertia I of the cross-sectional area about the neutral axis.

𝐼 = ∑ 𝐼 ̅ + 𝐴𝑑 2

100 × 1503 45 × 1203


𝐼=( ) − 2( ) = 15.165 × 106 𝑚𝑚4
12 12

GIRUM MINDAYE 73
Strength of Materials AASTU

(b) Applying the flexural formula


 Maximum bending stress
𝑀𝑐 (30 × 106 )(75)
𝜎𝑚𝑎𝑥 = = = 148.37𝑀𝑃𝑎
𝐼 15.165 × 106
 Bending stress distribution acting on the top flange A.
At y = 60 mm,
𝑀𝑦 (30 × 106 )(60)
𝜎|𝑦=60𝑚𝑚 = = = 118.69𝑀𝑃𝑎
𝐼 15.165 × 106
The bending stress distribution across the cross section is shown in Fig. a.

The resultant force acting on flange A is equal to the volume of the trapezoidal stress block
shown in Fig. a. Thus,
1
𝐹𝑅 = (148.37 + 118.69)(100 × 15) = 200,295𝑁 ≈ 200.30𝑘𝑁
2
Example 3.9: If the beam is subjected to a moment of M = 100kNm, determine the bending
stress at points A, B, and C. And determine the maximum moment M that can be applied to the
beam if the beam is made of material having an allowable tensile and compressive stress of
(σallow)t = 125 MPa and (σallow)c = 150 MPa, respectively.

GIRUM MINDAYE 74
Strength of Materials AASTU

Solution
(a) Section Property.
The coordinate 𝑦̅ of the centroid of the cross section from the bottom is
∑ 𝐴𝑖 𝑦̅𝑖 (300 × 30)(15) + (30 × 300)(30 + 300⁄2)
𝑦̅𝑏𝑜𝑡 = = = 97.5𝑚𝑚
∑ 𝐴𝑖 (300 × 30) + (30 × 300)

𝑦̅𝑡𝑜𝑝 = 30 + 300 + −𝑦̅𝑏𝑜𝑡 = 330 − 97.5 = 232.5𝑚𝑚

Compute the moment of inertia I of the cross-sectional area about the neutral axis.

𝐼 = ∑ 𝐼 ̅ + 𝐴𝑑 2
300 × 303 30 × 3003
𝐼=[ + (300 × 30)(97.5 − 15)2 ] + [ + (30 × 300)(232.5 − 150)2 ]
12 12
= 190.6875 × 106 𝑚𝑚4

(b) Applying the flexural formula


 Bending stress at points A, B, and C if M=100kNm
The distance from the neutral axis to points A, B, and C is yA= 232.5mm, yB= 97.5mm, and yC=
97.5-30=67.5mm.
𝑀𝑦𝐴 (100 × 106 )(232.5)
𝜎𝐴 = − =− = −121.93𝑀𝑃𝑎 (𝐶)
𝐼 190.6875 × 106
𝑀𝑦𝐵 (100 × 106 )(−97.5)
𝜎𝐵 = − =− = 51.13𝑀𝑃𝑎 (𝑇)
𝐼 190.6875 × 106
𝑀𝑦𝐶 (100 × 106 )(−67.5)
𝜎𝐶 = − =− = 35.40𝑀𝑃𝑎 (𝑇)
𝐼 190.6875 × 106
Using these results, the bending stress distribution across the cross section is shown in figure.

 Maximum moment M if (σallow)t = 125 MPa and (σallow)c = 150 MPa


𝑀𝑦̅𝑏𝑜𝑡 (𝑀)(−97.5)
(𝜎𝑡 )𝑎𝑙𝑙𝑜𝑤 = − ⟹ 125 = − 6
⟹ 𝑀 = 244.47 × 106 𝑁𝑚𝑚
𝐼 190.6875 × 10
∴ 𝑀 = 244.47𝑘𝑁𝑚

GIRUM MINDAYE 75
Strength of Materials AASTU

𝑀𝑦̅𝑡𝑜𝑝 (𝑀)(232.5)
(𝜎𝑐 )𝑎𝑙𝑙𝑜𝑤 = − ⟹ −150 = − ⟹ 𝑀 = 123.02 × 106 𝑁𝑚𝑚
𝐼 190.6875 × 106
∴ 𝑀 = 123.02𝑘𝑁𝑚 (𝑔𝑜𝑣𝑒𝑟𝑛𝑠)
Example 3.10: The simply supported wood beam is fabricated by gluing together three 160mm
by 80mm plans as shown in the figure below. Calculate the maximum shear stress in the glue
and the wood.

Solution
From the shear force diagram in Fig. (b), the maximum shear force in the beam is Vmax= 24kN,
occurring at the supports.
The moment of inertia of the cross-sectional area of the beam about the neutral axis is
𝑏ℎ3 0.16 ×. 243
𝐼= = = 184.32 × 10−6 𝑚4
12 12
The shear stress is the glue corresponds to the horizontal shear stress. Its maximum value can
be computed from τmax=VmaxQ/(Ib), where Q is the first moment of the area A’ shown in Fig.(c);
that is,
̅ = (0.16 × 0.08)(0.08) = 1.024 × 10−3 𝑚3
𝑄 = 𝐴′ 𝑦′

Therefore, the shear stress in the glue, which occurs over either support, is
𝑉𝑚𝑎𝑥 𝑄 (24 × 103 )(1.024 × 10−3 )
𝜏𝑚𝑎𝑥 = = = 8.33 × 103 𝑃𝑎 = 8.33𝑘𝑃𝑎
𝐼𝑏 (184.32 × 10−6 )(0.16)
Because the cross section is rectangular, the maximum shear stress in the wood can be
calculated by:
3 𝑉𝑚𝑎𝑥 3 24 × 103
𝜏𝑚𝑎𝑥 = = ( ) = 938 × 103 𝑃𝑎 = 938𝑘𝑃𝑎
2 𝐴 2 0.16 × 0.24

GIRUM MINDAYE 76
Strength of Materials AASTU

𝑉𝑚𝑎𝑥 𝑄
The same result can be obtained from𝜏𝑚𝑎𝑥 = , where now A’ is the area above the
𝐼𝑏
neutral axis, as indicated in Fig. (d).

The first moment of this area about the neutral axis is


̅ = (0.16 × 0.12)(0.6) = 1.152 × 10−3 𝑚3
𝑄 = 𝐴′ 𝑦′
𝑉𝑚𝑎𝑥 𝑄 (24 × 103 )(1.152 × 10−3 )
𝜏𝑚𝑎𝑥 = = = 938 × 103 𝑃𝑎 = 938𝑘𝑃𝑎
𝐼𝑏 (184.32 × 10−6 )(0.16)
which agrees with the previous result.
Example 3.11: The W12×40 section is used as a beam. If the vertical shear acting at a certain
section of the beam is 16 kips, determine the following at that section: (a) the minimum shear
stress in the web; (b) the maximum shear stress in the web; and (c)the percentage of the shear
force that is carried by the web.

Solution
The W12×40 section is shown in Fig.(b), where the dimensions were
obtained from the steel section tables. The drawing approximates the web
and the flanges by rectangles, thereby ignoring the small fillets and
rounded corners present in the actual section. The tables also list the
moment of inertia of the section about the neutral axis as I = 310in4.
a) The minimum shear stress in the web occurs at the junction with the
flange, where Q/b is smallest (note that b = 0.295in is constant within
the web). Q is the first moment of the area A1’ shown in Fig.(b) about the neutral axis:
11.94 − 0.515
𝑄 = 𝐴′1 ̅̅̅̅
𝑦1 ′ = (8.005 × 0.515) ( ) = 23.55𝑖𝑛3
2
The minimum shear stress in thus becomes
𝑉𝑄 (16 × 103 )(23.55)
𝜏𝑚𝑖𝑛 = = = 4120𝑃𝑠𝑖
𝐼𝑏 310 × 0.295

GIRUM MINDAYE 77
Strength of Materials AASTU

b) The maximum shear stress is located at the neutral axis, where Q/b is largest. Hence, Q is
the first moment of the area above (or below) the neutral axis.
The moment of A1’ was calculated in part (a). The moment of A2’ about the neutral axis is
where
11.94
𝐴2 ′ = ( − 0.515) (0.295) = 1.6092𝑖𝑛2
2

̅̅̅̅ 1 11.94
𝑦2 ′ = ( − 0.515) = 2.7275𝑖𝑛
2 2
𝑄 = 𝐴′1 ̅̅̅̅
𝑦1 ′ + 𝐴′2 ̅̅̅̅
𝑦2 ′ = 23.55 + (1.6092)(2.7275) = 27.94𝑖𝑛3
The maximum shear stress in the web becomes
𝑉𝑄 (16 × 103 )(27.94)
𝜏𝑚𝑎𝑥 = = = 4890𝑃𝑠𝑖
𝐼𝑏 310 × 0.295

c) Percentage of the shear force carried by the web.


The distribution of the shear stress in the web is shown in Fig.(c).The shear force carried by
the web is
Vweb= (cross section area of web) × (area of shear diagram)
The shear stress distribution is parabolic. Recalling that the area of a parabola is (2/3) (base×
height).
2
𝑉𝑤𝑒𝑏 = (10.91 × 0.295) [4120 + (4890 − 4120)] = 14910𝑙𝑏
3
Therefore the percentage of the shear force carried by the web is
𝑉𝑤𝑒𝑏 14910
× 100 = × 100 = 93.2%
𝑉 16000
The result confirms that the flanges are ineffective in resisting the vertical shear
Note that for wide flanges (W-shapes), we can use 𝜏𝑚𝑎𝑥 = 𝑉 ⁄𝐴𝑤𝑒𝑏 as a rough approximation
for the maximum shear stress.

GIRUM MINDAYE 78
Strength of Materials AASTU

𝑉 16000
𝜏𝑚𝑎𝑥 = = = 4970𝑃𝑠𝑖
𝐴𝑤𝑒𝑏 10.91 × 0.295
Which differs from τmax= 4890 psi computed in Part (b) by less than 2%.
Example 3.12: The beam shown in Fig. a is made from two boards. Determine the maximum
shear stress in the glue necessary to hold the boards together along the seam where they are
joined.

Solution
Internal Shear: The support reactions and the shear diagram for the beam are shown in
Fig.b. It is seen that the maximum shear in the beam is 19.5kN.

Section Properties: The centroid and therefore the neutral axis will be determined from the
reference axis placed at the bottom of the cross-sectional area, Fig. a. Working in units of
milimeters, we have
∑ 𝑦̃𝐴 (75)(30 × 150) + (165)(150 × 30)
𝑦̅ = = = 120𝑚𝑚
∑𝐴 (30 × 150) + (150 × 30)
The moment of inertia about the neutral axis, Fig. a, is therefore
30 × 1503 150 × 303
𝐼=[ + (30 × 150)(120 − 75)2 ] + [ + (150 × 30)(165 − 120)2 ]
12 12
= 27 × 106 𝑚𝑚4
The top board (flange) is held onto the bottom board (web) by the glue, which is applied over
the thickness t = 30mm. Consequently Q is taken from the area of the top board, Fig. a. We
have

̅ = (150 × 30) (180 − 30 − 120) = 20.25 × 104 𝑚𝑚3


𝑄 = 𝐴′ 𝑦′
2
Shear Stress: Applying the shear formula,

GIRUM MINDAYE 79
Strength of Materials AASTU

𝑉𝑚𝑎𝑥 𝑄 (19.5 × 103 )(20.25 × 104 )


𝜏𝑚𝑎𝑥 = = = 4.88𝑀𝑃𝑎
𝐼𝑡 (27 × 106 )(30)
The shear stress acting at the top of the bottom board is shown in Fig. c.

Note. It is the glue’s resistance to this longitudinal shear stress that holds the boards from
slipping at the right support.
Example 3.13: A steel wide-flange beam has the dimensions shown in figure. If it is subjected
to a shear of V = 80kN, plot the shear-stress distribution acting over the beam’s cross section.

Solution
Since the flange and web are rectangular elements, then like the previous example, the shear-
stress distribution will be parabolic and in this case it will vary in the manner shown in Fig. b.

Due to symmetry, only the shear stresses at points B’, B, and C have to be determined. To show
how these values are obtained, we must first determine the moment of inertia of the cross-
sectional area about the neutral axis. Working in milimeters, we have

GIRUM MINDAYE 80
Strength of Materials AASTU

15 × 2003 300 × 203 20 2


𝐼=[ ] + 2[ + (300 × 20) (100 + ) ] = 155.6 × 106 𝑚𝑚4
12 12 2
For point B', t'B = 300mm, and A' is the dark shaded area shown
in Fig. c. Thus,

̅ = (300 × 20) (100 + 20) = 66 × 104 𝑚𝑚3


𝑄𝐵′ = 𝐴′ 𝑦′
2
so that
𝑉𝑄𝐵′ (80 × 103 )(66 × 104 )
𝜏 =
𝐵′ = = 1.13𝑀𝑃𝑎
𝐼𝑡𝐵′ (155.6 × 106 )(300)
For point B, tB = 15mm and QB = QB', Fig. c. Hence
𝑉𝑄𝐵 (80 × 103 )(66 × 104 )
𝜏𝐵 = = = 22.62𝑀𝑃𝑎
𝐼𝑡𝐵 (155.6 × 106 )(15)
For point C, tC = 15mm and A' is the dark shaded area shown in
Fig. d. Considering this area to be composed of two rectangles,
we have
𝑄𝐶 = ∑ 𝐴′ 𝑦′̅ = (300 × 20) (100 + 20) + (15 × 100) (100)
2 2
= 73.5 × 104 𝑚𝑚3
Thus,
𝑉𝑄𝐶 (80 × 103 )(73.5 × 104 )
𝜏𝐶 = = = 25.19𝑀𝑃𝑎
𝐼𝑡𝐶 (155.6 × 106 )(15)
From Fig. b, note that the largest shear stress occurs in the web and is almost uniform
throughout its depth, varying from 22.62 MPa to 25.19 MPa. It is for this reason that for design,
some codes permit the use of calculating the average shear stress on the cross section of the
web, rather than using the shear formula; that is,
𝑉 80 × 103
𝜏𝑎𝑣𝑔 = = = 26.67𝑀𝑃𝑎
𝐴 15 × 200
Example 3.14: If the wide-flange beam is subjected to a shear of V = 20kN, determine (a) the
shear stress on the web at A, (b) the maximum shear stress in the beam and (c) the shear force
resisted by the web of the beam.

GIRUM MINDAYE 81
Strength of Materials AASTU

Solution
(a) Section Properties
Due to the doula symmetry of the section, the location of the centroid measured from the
bottom or top is h/2.
ℎ 20 + 300 + 20
𝑦̅ = = = 170𝑚𝑚
2 2
The moment of inertia of the cross-section about the neutral axis is
200 × 3403 90 × 3003
𝐼= − 2( ) = 250.1 × 106 𝑚𝑚4
12 12

(b) Shear stress on the web at A

̅ = (200 × 20) (170 − 20) = 64 × 104 𝑚𝑚3


𝑄𝐴 = 𝐴′ 𝑦′
2
Applying the shear formula,

𝑉𝑄𝐴 (20 × 103 )(64 × 104 )


𝜏𝐴 = = = 2.56𝑀𝑃𝑎
𝐼𝑡𝐴 (250.1 × 106 )(20)
The shear stress component at A is represented by the volume element shown in figure.

(c) Maximum shear stress in the beam

̅ = (200 × 20)(160) + (20 × 150)(75) = 86.5 × 104 𝑚𝑚3


𝑄𝑚𝑎𝑥 = ∑ 𝐴′ 𝑦′
The maximum shear stress occurs at the points along neutral axis since Q is maximum and
thickness t is the smallest.
𝑉𝑄𝑚𝑎𝑥 (20 × 103 )(86.5 × 104 )
𝜏𝑚𝑎𝑥 = = = 3.46𝑀𝑃𝑎
𝐼𝑡 (250.1 × 106 )(20)

(d) Shear force resisted by the web of the beam

GIRUM MINDAYE 82
Strength of Materials AASTU

For 0≤ y < 0.15m, Fig. a, Q as a function of y is

̅ = (200 × 20)(160) + [20 × (150 − 𝑦)] (𝑦 + 150)


𝑄 = ∑ 𝐴′ 𝑦′
2
(86.5 2) 4 3
= − 1000𝑦 × 10 𝑚𝑚
For 0 ≤ y < 0.15m, t = 0.02m. Thus,
𝑉𝑄 (20 × 103 )(86.5 − 1000𝑦 2 ) × 104
𝜏= = = (3.459 − 39.984𝑦 2 )𝑀𝑃𝑎
𝐼𝑡 (250.1 × 106 )(20)
The sheer force resisted by the web is
0.15 0.15

𝑉𝑤 = 2 ∫ 𝜏𝑑𝐴 = 2 ∫ [(3.459 − 39.984𝑦 2 ) × 106 ](0.02𝑑𝑦)


0 0
0.15
4 2 )𝑑𝑦 4
39.984𝑦 3 0.15
𝑉𝑤 = 4 × 10 ∫ (3.459 − 39.984𝑦 = 4 × 10 (3.459𝑦 − )|
3 0
0

𝑉𝑤 = 18,954.72𝑁 ≈ 18.95𝑘𝑁
Example 3.15: If the beam is subjected to a shear of V = 30kN, determine the web’s shear
stress at A and B. Indicate the shear-stress components on a volume element located at these
points. Set w= 200mm.

Solution

GIRUM MINDAYE 83
Strength of Materials AASTU

a) Section Properties:
The location of the centroid measured from the bottom is
∑ 𝑦̃𝐴 (10)(200 × 20) + (20 + 200)(20 × 400) + (20 + 400 + 10)(300 × 20)
𝑦̅ = =
∑𝐴 (200 × 20) + (20 × 400) + (300 × 20)
𝑦̅ = 243.33𝑚𝑚

The moment of inertia of the cross-section about the neutral axis is


200 × 203
𝐼=[ + (200 × 20)(243.33 − 10)2 ]
12
20 × 4003
+[ + (20 × 400)(243.33 − 220)2 ]
12
300 × 203
+[ + (300 × 20)(430 − 243.33)2 ] = 538.2 × 106 𝑚𝑚4
12
Referring to Fig. a,
̅ 𝐴 = (300 × 20)(440 − 243.33 − 10) = 112 × 104 𝑚𝑚3
𝑄𝐴 = 𝐴′𝐴 𝑦′

̅ 𝐵 = (200 × 20)(243.33 − 10) = 93.33 × 104 𝑚𝑚3


𝑄𝐵 = 𝐴′ 𝐵 𝑦′
b) Shear Stress: Applying the shear formula,
𝑉𝑄𝐴 (30 × 103 )(112 × 104 )
𝜏𝐴 = = = 3.12𝑀𝑃𝑎
𝐼𝑡𝐴 (538.2 × 106 )(20)
𝑉𝑄𝐵 (30 × 103 )(93.33 × 104 )
𝜏𝐵 = = = 2.60𝑀𝑃𝑎
𝐼𝑡𝐵 (538.2 × 106 )(20)
These shear stresses on the volume element at points A and B are shown in Fig. b.

GIRUM MINDAYE 84
Strength of Materials AASTU

Example 3.16: Determine the maximum shear stress in the T-beam (a) at the critical section
where the internal shear force is maximum and (b) at point C. Show the result on a volume
element at this point.

Solution
(a) Maximum shear force
The shear diagram is shown in Fig. b. As indicated, Vmax=27.5kN

(b) Section properties


The neutral axis passes through centroid c of the cross-section, Fig. c.
∑ 𝑦̃𝐴 (75)(30 × 150) + (150 + 15)(150 × 30)
𝑦̅ = = = 120𝑚𝑚
∑𝐴 (30 × 150) + (150 × 30)
30 × 1503 150 × 303
𝐼=[ + (30 × 150)(120 − 75)2 ] + [ + (150 × 30)(165 − 120)2 ]
12 12

𝐼 = 27 × 106 𝑚𝑚4
From Fig. d,
̅ = (30 × 120)(60) = 21.6 × 104 𝑚𝑚3
𝑄𝑚𝑎𝑥 = 𝐴′ 𝑦′

GIRUM MINDAYE 85
Strength of Materials AASTU

(c) Maximum shear stress at Vmax


The maximum shear stress occurs at points on the neutral axis since Q is maximum and
thickness t=30mm is the smallest.
𝑉𝑚𝑎𝑥 𝑄𝑚𝑎𝑥 (27.5 × 103 )(21.6 × 104 )
𝜏𝑚𝑎𝑥 = = = 7.33𝑀𝑃𝑎
𝐼𝑡 (27 × 106 )(30)

(d) Maximum shear stress at point C


Using the method of sections,
1
+↑ ∑ 𝐹𝑦 = 0, 𝑉𝐶 + 17.5 − (5)(1.5) = 0
2
𝑉𝐶 = −13.75𝑘𝑁
𝑉𝐶 𝑄𝑚𝑎𝑥 (13.75 × 103 )(21.6 × 104 )
𝜏𝐶 = =
𝐼𝑡 (27 × 106 )(30)
= 3.67𝑀𝑃𝑎

GIRUM MINDAYE 86
Strength of Materials AASTU

4 TORSION
4.1 Theory of Pure Torsion

A member is said to be in ‘Pure torsion’, when its cross sections are subjected to only torsional
moments (or torque) and not accompanied by axial forces and bending moment. In this section,
structural members and machine parts that are in torsion will be analyzed, where the stresses
and strains in members of circular cross section are subjected to twisting couples, or torques,
T and T’ (Figure 4-1). These couples have a common magnitude T, and opposite senses. They
are vector quantities and can be represented either by curved arrows (Figure 4-1a) or by couple
of vectors (Figure 4-1b).

Figure 4-1: Two equivalent ways to represent a torque in a free-body diagram.

Members in torsion are encountered in many engineering applications. The most common
application is provided by transmission shafts, which are used to transmit power from one point
to another (Photo 4.1). These shafts can be either solid, as shown in Figure 4-1, or hollow.
Because a circular cross section is an efficient shape for resisting torsional loads. Circular shafts
are commonly used to transmit power in rotating machinery.

Photo 4.1: In this automotive power train, the shaft transmits power from the engine to the
rear wheels.

The following are the assumptions in the theory of pure torsion for circular shaft subjected to
torsion:

GIRUM MINDAYE 87
Strength of Materials AASTU

 The materiel is homogenous i.e. of uniform elastic properties exists throughout the
material.
 The material is elastic, follows Hook's law, with shear stress proportional to shear
strain.
 The stress does not exceed the elastic limit.
 The circular section remains circular
 Cross section remain plane. This property is characteristic of circular shafts, whether
solid or hollow—but not of members with noncircular cross section. For example, when
a bar of square cross section is subjected to torsion, its various cross-sections warp and
do not remain plane (Figure 4-2b).

Figure 4-2: Comparison of deformations in (a) circular and (b) square shafts.

 Cross section rotate as if rigid i.e. every diameter rotates through the same angle.
4.2 Torsional Moment of Resistance

Consider 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 (Figure 4-3).

Figure 4-3: Shaft subject to torques and a section plane at C.

The free-body diagram of portion BC of the shaft must include the elementary shearing forces
dF, which are perpendicular to the radius of the shaft. These arise from the torque that portion
AC exerts on BC as the shaft is twisted (Figure 4-4a). The conditions of equilibrium for BC
require that the system of these forces be equivalent to an internal torque T, as well as equal
and opposite to T’ (Figure 4-4b).

GIRUM MINDAYE 88
Strength of Materials AASTU

Figure 4-4: (a) Free body diagram of section BC with torque at C represented by the
contributions of small elements of area carrying forces dF a radius ρ from the section center.
(b) Free-body diagram of section BC having all the small area elements summed resulting in
torque T.
Denoting the perpendicular distance ρ from the force dF to the axis of the shaft and
expressing that the sum of the moments of the shearing forces dF about the axis of the shaft is
equal in magnitude to the torque T, write

∫ 𝜌𝑑𝐹 = 𝑇
Since dF = τdA, where τ is the shearing stress on the element of area dA, you also can write
∫ 𝜌𝑑𝐹 = ∫ 𝜌(𝜏𝑑𝐴) = 𝑇

4.3 Shearing Strains

Now we will determine the distribution of shearing strains in a circular


shaft of length L and radius c that has been twisted through an angle ϕ
(Fig. 3.13a). Detaching from the shaft a cylinder of radius ρ, consider
the small square element formed by two adjacent circles and two
adjacent straight lines traced on the surface before any load is applied
(Fig. 3.13b). As the shaft is subjected to a torsional load, the element
deforms into a rhombus (Fig. 3.13c). Here the shearing strain γ in a
given element is measured by the change in the angles formed by the
sides of that element. Since the circles defining two of the sides remain
unchanged, the shearing strain γ must be equal to the angle between
lines AB and A’B.

Figure 3.13c shows that, for small values of γ, the arc length AA’ is
expressed as AA’= Lγ. But since AA’ =ρϕ, it follows that Lγ = ρϕ, or
𝜌𝜙
𝛾=
𝐿
Where γ and ϕ are in radians.

Figure 4-5:
Shearing strain deformation. (a) The angle of twist ϕ. (b) Undeformed portion of shaft
of radius ρ. (c) Deformed portion of shaft; angle of twist ϕ and shearing strain γ share the
same arc length AA.
Thus, the shearing strain in a circular shaft varies linearly with the distance from the axis of
the shaft. The shearing strain is maximum on the surface of the shaft, where ρ = c.

GIRUM MINDAYE 89
Strength of Materials AASTU

𝑐𝜙 𝛾𝑚𝑎𝑥 𝐿
𝛾𝑚𝑎𝑥 = ⟹𝜙=
𝐿 𝑐
Eliminating ϕ, the shearing strain γ at a distance ρ from the axis of the shaft is

𝜌𝜙 𝜌 𝛾𝑚𝑎𝑥 𝐿 𝜌
𝛾= = ( ) = 𝛾𝑚𝑎𝑥
𝐿 𝐿 𝑐 𝑐
4.4 Shear Stresses

When the torque T is such that all shearing stresses in the shaft remain below the yield strength
τy, the stresses in the shaft will remain below both the proportional limit and the elastic limit.
Thus, Hooke,s law will apply,and there will be no permanent deformation.
𝜏 = 𝐺𝛾
where G is the modulus of rigidity or shear modulus of the material.
𝜌 𝜌
𝛾 = 𝛾𝑚𝑎𝑥 , 𝑚𝑢𝑙𝑡𝑖𝑝𝑙𝑦𝑖𝑛𝑔 𝑏𝑜𝑡ℎ 𝑠𝑖𝑑𝑒𝑠 𝑏𝑦 ⟹ 𝐺𝛾 = 𝐺𝛾𝑚𝑎𝑥
𝑐 𝑐
𝜏 = 𝐺𝛾, 𝜏𝑚𝑎𝑥 = 𝐺𝛾𝑚𝑎𝑥 ,

𝜌
𝜏= 𝜏
𝑐 𝑚𝑎𝑥
This equation 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
ρ from the axis of the shaft. Figure 4-6a shows the stress distribution in a solid circular shaft of
radius c.

Figure 4-6: Distribution of shearing stresses in a torqued shaft: (a) Solid shaft, (b) Hollow
shaft.
A hollow circular shaft of inner radius c1 and outer radius c2 is shown in Figure 4-6b.

𝑐1
𝜏𝑚𝑖𝑛 = 𝜏
𝑐2 𝑚𝑎𝑥

The sum of the moments of the elementary forces exerted on any cross section of the shaft
must be equal to the magnitude T of the torque exerted on the shaft:
∫ 𝜌(𝜏𝑑𝐴) = 𝑇
𝜌 𝜏𝑚𝑎𝑥
∫ 𝜌 ( 𝜏𝑚𝑎𝑥 𝑑𝐴) = ∫ 𝜌2 𝑑𝐴 = 𝑇
𝑐 𝑐

GIRUM MINDAYE 90
Strength of Materials AASTU

∫ 𝜌2 𝑑𝐴 - represents the polar moment of inertia J of the cross section with respect to its
center O. Therefore,
𝜏𝑚𝑎𝑥 𝑇𝑐
𝑇= 𝐽 ⟹ 𝜏𝑚𝑎𝑥 =
𝑐 𝐽
The torsion formulas were derived for a shaft of uniform circular cross section subjected to
torques at its ends. However, they also can be used for a shaft of variable cross section or for a
shaft subjected to torques at locations other than its ends (Figure 4-7a). The distribution of
shearing stresses in a given cross section S of the shaft is obtained from
𝑇𝑐
𝜏𝑚𝑎𝑥 =
𝐽
Where J is the polar moment of inertia of that section and T represents the internal torque in
that section.

Figure 4-7: Shaft with variable cross section. (a) With applied torques and section S. (b) Free-
body diagram of sectioned shaft.
T is obtained by drawing the free-body diagram of the portion of shaft located on one side of
the section (Figure 4-7b) and writing that the sum of the torques applied (including the internal
torque T) is zero.
↺ + ∑ 𝑇 = 0, 𝑇𝐵 − 𝑇𝐸 − 𝑇 = 0 ⟹ 𝑇 = 𝑇𝐵 − 𝑇𝐸
Now consider two elements a and b located on the surface of a circular shaft subjected to
torsion (Fig. 3.17).

Figure 4-8: Circular shaft with stress elements at different orientations.


Since the faces of element a are respectively parallel and perpendicular to the axis of the shaft,
the only stresses on the element are the shearing stresses

𝑇𝑐
𝜏𝑚𝑎𝑥 =
𝐽
On the other hand, the faces of element b, which form arbitrary angles with the axis of the
shaft, are subjected to a combination of normal and shearing stresses. Consider the stresses and

GIRUM MINDAYE 91
Strength of Materials AASTU

resulting forces on faces that are at 450 to the axis of the shaft. The free-body diagrams of the
two triangular elements are shown in Figure 4-9.

Figure 4-9: Forces on faces at 450 to shaft axis.


From Figure 4-9a, the stresses exerted on the faces BC and BD are the shearing stresses τmax=
Tc/J. The magnitude of the corresponding shear forces is τmaxA0, where A0 is the area of the
face. Observing that the components along DC of the two shear forces are equal and opposite,
the force F exerted on DC must be perpendicular to that face and is a tensile force. Its magnitude
is

𝐹 = 2(𝜏𝑚𝑎𝑥 𝐴0 ) cos 450 = 𝜏𝑚𝑎𝑥 𝐴0 √2


The corresponding stress is obtained by dividing the force F by the area A of face DC.
Observing that 𝐴 = 𝐴0 √2

𝐹 𝜏𝑚𝑎𝑥 𝐴0 √2
𝜎= = = 𝜏𝑚𝑎𝑥
𝐴 𝐴0 √2
A similar analysis of the element of Figure 4-9b shows that the stress on the face BE is σ=-
τmax.

Figure 4-10: Shaft elements with only shearing stresses or normal stresses.
Therefore, the stresses exerted on the faces of an element c at 450 to the axis of the shaft (Figure
4-10) are normal stresses equal to ± τmax. Thus, while element a in Figure 4-10 is in pure shear,
element c in the same figure is subjected to a tensile stress on two of its faces and a compressive
stress on the other two. Also, note that all of the stresses involved have the same magnitude,
Tc/J.
4.5 Strength of Shaft

Strength of shaft is the ability of shaft to rest the action of twisting moment, T.

GIRUM MINDAYE 92
Strength of Materials AASTU

𝐽
𝑇 = 𝜏𝑚𝑎𝑥
𝑐
𝑐𝜙
𝑓𝑜𝑟 𝜏𝑚𝑎𝑥 = 𝐺𝛾𝑚𝑎𝑥 𝑎𝑛𝑑 𝛾𝑚𝑎𝑥 = ;
𝐿
𝐽 𝐽 𝑐𝜙 𝜙
𝑇 = (𝐺𝛾𝑚𝑎𝑥 ) = (𝐺 ) = 𝐺𝐽
𝑐 𝑐 𝐿 𝐿
𝜙 𝑇𝐿
𝑇 = 𝐺𝐽 ⟹𝜙=
𝐿 𝐺𝐽

Figure 4-11: Torque applied to fixed end shaft resulting in angle of twist ϕ.
𝑇𝐿
𝜙 = 𝐺𝐽 can be used for the angle of twist only if the shaft is homogeneous (constant G), has a
uniform cross section, and is loaded only at its ends (Figure 4-11). If the shaft is subjected to
torques at locations other than its ends or if it has several portions with various cross sections
and possibly of different materials, it must be divided into parts that satisfy the required
𝑇𝐿
conditions for 𝜙 = 𝐺𝐽.

Figure 4-12: Shaft with multiple cross-section dimensions and multiple loads.
For shaft AB shown in Figure 4-12, four different parts should be considered: AC, CD, DE, and
EB. The total angle of twist of the shaft (i.e., the angle through which end A rotates with respect
to end B) is obtained by algebraically adding the angles of twist of each component part. Using
the internal torque Ti, length Li, cross-sectional polar moment of inertia Ji , and modulus of
rigidity Gi, corresponding to part i, the total angle of twist of the shaft is
𝑇𝑖 𝐿𝑖
𝜙=∑
𝐺𝑖 𝐽𝑖
𝑖
Polar modulus (Zp) is the ratio of polar moment of inertia to extreme radial distance of the
fiber from the center.
𝐽
𝑇 = 𝜏𝑚𝑎𝑥 = 𝑍𝑝 𝜏𝑚𝑎𝑥
𝑐
GIRUM MINDAYE 93
Strength of Materials AASTU

𝐽
𝑤ℎ𝑒𝑟𝑒, 𝑍𝑝 = = 𝑝𝑜𝑙𝑎𝑟 𝑚𝑜𝑑𝑢𝑙𝑢𝑠
𝑐

Torsional rigidity is the product of modulus of rigidity (G) and polar moment of inertia (J).
𝑇𝑜𝑟𝑠𝑖𝑜𝑛𝑎𝑙 𝑟𝑖𝑔𝑖𝑑𝑖𝑡𝑦 = 𝐺𝐽
𝜙 𝑇𝐿
𝑓𝑟𝑜𝑚, 𝑇 = 𝐺𝐽 ⟹ 𝐺𝐽 =
𝐿 𝜙
Torsional stiffness is the amount of toque required to produce unit twist. It is also defined as
the torsional rigidity per unit length of shaft.
𝜙 1 𝐺𝐽
𝑓𝑜𝑟 𝑇 = 𝐺𝐽 , 𝑇𝑜𝑟𝑠𝑖𝑜𝑛𝑎𝑙 𝑠𝑡𝑖𝑓𝑓𝑛𝑒𝑠𝑠 (𝐾) = 𝐺𝐽 × =
𝐿 𝐿 𝐿
Table 4-1: Polar modulus of shaft

Shape Representation Moment of Inertia (I) Polar Moment of Polar modulus (Zp)
Inertia (J)
𝐽 = 𝐼𝑥 + 𝐼𝑦 𝑍𝑃 = 𝐽/𝑐

Circle 𝜋𝑟 4 𝜋𝑑 4 𝜋𝑟 4 𝜋𝑑 4 𝜋𝑟 3 𝜋𝑑 3
𝐼𝑥 = 𝐼𝑦 = = 𝐽= = 𝑍𝑃 = =
4 64 2 32 2 16

𝜋(𝑟 4 0 − 𝑟 4 𝑖 ) 𝜋(𝑟 4 0 − 𝑟 4 𝑖 ) 𝜋(𝑟 4 0 − 𝑟 4 𝑖 )


𝐼𝑥 = 𝐼𝑦 = 𝐽= 𝑍𝑃 =
4 2 2𝑟0
Circular Or Or Or
Tube
𝜋(𝑑 4 0 − 𝑑 4 𝑖 ) 𝜋(𝑑 4 0 − 𝑑 4 𝑖 ) 𝜋(𝑑 4 0 − 𝑑 4 𝑖 )
𝐼𝑥 = 𝐼𝑦 = 𝐽= 𝑍𝑃 =
64 32 16𝑑0
4.6 Design of Transmission Shafts

The principal specifications to be met in the design of a transmission shaft are the power to be
transmitted and the speed of rotation of the shaft. The role of the designer is to select the
material and the dimensions of the cross section of the shaft so that the maximum shearing
stress allowable will not be exceeded when the shaft is transmitting the required power at the
specified speed.
To determine the torque exerted on the shaft, the power P associated with the rotation of a
rigid body subjected to a torque T is
𝑃 = 𝜔𝑇
where ω is the angular velocity of the body in radians per second (rad/s). But ω =2πf, where f
is the frequency of the rotation, (i.e., the number of revolutions per second). The unit of
frequency is 1s-1 and is called a hertz (Hz).
𝑃
∴ 𝑃 = 2𝜋𝑓𝑇 ⟹ 𝑇 =
2𝜋𝑓

GIRUM MINDAYE 94
Strength of Materials AASTU

When SI units are used with f expressed in Hz and T in Nm, the power will be in N.m/s—that
is, in watts (W).
After determining the torque T to be applied to the shaft and selecting the material to be used,
the designer carries the values of T and the maximum allowable stress into 𝑇 = 𝑍𝑝 𝜏𝑚𝑎𝑥 .
𝑇
𝑍𝑝 =
𝜏𝑚𝑎𝑥
This also provides the minimum allowable parameter Zp.
4.7 Worked Examples
Example 4.1: A hollow cylindrical steel shaft is 1.5m long and has inner and outer diameters
respectively equal to 40 and 60 mm (Figure 4-13). (a) What is the largest torque that can be
applied to the shaft if the shearing stress is not to exceed 120 MPa? (b) What is the
corresponding minimum value of the shearing stress in the shaft? (c) What torque should be
applied to the end of the shaft to produce a twist of 20? Use the value G =77GPa for the modulus
of rigidity of steel. (d)What angle of twist will create a shearing stress of 70 MPa on the inner
surface of the hollow steel shaft? Use the value G =77GPa.

Figure 4-13: Hollow, fixed-end shaft having torque T applied at end.

Solution

(1) The largest torque T that can be applied to the shaft


𝑇 = 𝑍𝑝 𝜏𝑚𝑎𝑥
3
𝜏𝑚𝑎𝑥 = 120𝑀𝑃𝑎, 𝑟 𝑖 = 𝑐1 = 0.02𝑚 𝑎𝑛𝑑 𝑟 3 0 = 𝑐2 = 0.03𝑚
𝜋(𝑟 4 0 − 𝑟 4 𝑖 ) 𝜋(0.034 − 0.024 )
𝑓𝑜𝑟 𝐶𝑖𝑟𝑐𝑢𝑙𝑎𝑟 𝑇𝑢𝑏𝑒, 𝑍𝑃 = = = 34.033 × 10−6 𝑚3
2𝑟0 2 × 0.03
−6
𝑇 = (120 × 106 ) (34.033 × 10 ) = 4,083.96𝑁𝑚 ≈ 4.08𝑘𝑁
(2) The minimum shearing stress in the shaft
𝑐1 0.02
𝜏𝑚𝑖𝑛 = 𝜏𝑚𝑎𝑥 = × (120𝑀𝑃𝑎) = 80𝑀𝑃𝑎
𝑐2 0.03
(3) Torque due to a twist of 20
𝜋(𝑟 4 0 − 𝑟 4 𝑖 ) 𝜋(0.034 − 0.024 )
𝑓𝑜𝑟 𝐶𝑖𝑟𝑐𝑢𝑙𝑎𝑟 𝑇𝑢𝑏𝑒, 𝐽= = = 1.021 × 10−6 𝑚4
2 2

GIRUM MINDAYE 95
Strength of Materials AASTU

2𝜋 𝑟𝑎𝑑
𝜙 = 20 ( ) = 34.91 × 10−3 𝑟𝑎𝑑
3600
𝜙 9 −6 34.91 × 10−3
𝑇 = 𝐺𝐽 = (77 × 10 ) (1.021 × 10 ) ( ) = 1,829.68𝑘𝑁𝑚 ≈ 1.83𝑘𝑁𝑚
𝐿 1.5
(4) Angle of twist due to a shearing stress of 70MPa on the inner surface of the hollow steel
shaft
𝜌𝜙 𝐿 𝜏𝐿
𝛾= ⟹𝜙=𝛾 =
𝐿 𝜌 𝐺𝜌
(70 6 )(1.5)
𝜏𝑚𝑖𝑛 𝐿 × 10
∴𝜙= = = 68.2 × 10−3 𝑟𝑎𝑑 = 3.910
𝑐1 𝐺 (0.02)(77 × 109 )
Example 4.2: Shaft BC is hollow with inner and outer diameters of 90mm and 120mm,
respectively. Shafts AB and CD are solid and of diameter d. For the loading shown, determine
(a) the maximum and minimum shearing stress in shaft BC, (b) the required diameter d of
shafts AB and CD if the allowable shearing stress in these shafts is 65MPa.

Solution
Use free-body diagrams to determine the torque in each shaft.

The torques can then be used to find the stresses for shaft BC and the required diameters for
shafts AB and CD.
Section through shaft AB

↺ + ∑ 𝑇 = 0, 𝑇𝐴 − 𝑇𝐴𝐵 = 0 ⟹ 6 − 𝑇𝐴𝐵 = 0
𝑇𝐴𝐵 = 6𝑘𝑁𝑚
Section through shaft BC

GIRUM MINDAYE 96
Strength of Materials AASTU

↺ + ∑ 𝑇 = 0, 𝑇𝐴 + 𝑇𝐵 − 𝑇𝐵𝐶 = 0 ⟹ 6 + 14 − 𝑇𝐵𝐶 = 0
𝑇𝐵𝐶 = 20𝑘𝑁𝑚
𝜋(𝑟 0 − 𝑟 4 𝑖 ) 𝜋(0.064 − 0.0454 )
4
𝑓𝑜𝑟 𝑠ℎ𝑎𝑓𝑡 𝐵𝐶, 𝑍𝑃 = = = 231.937 × 10−6 𝑚3
2𝑟0 2 × 0.06
(a) Maximum and minimum shearing stress in shaft BC

𝑇 20 × 103
𝜏𝑚𝑎𝑥 = 𝜏2 = = = 86.230 × 106 𝑃𝑎 = 86.23𝑀𝑃𝑎
𝑍𝑝 231.937 × 10−6

𝑐1 0.045
𝜏𝑚𝑖𝑛 = 𝜏𝑚𝑎𝑥 = × (86.23𝑀𝑃𝑎) = 64.67𝑀𝑃𝑎
𝑐2 0.06
(b) Diameter d of shafts AB and CD
We note that both shafts have the same torque T = 6kNm and τmax=65MPa
𝜋𝑟 3 𝜋𝑑 3
𝑓𝑜𝑟 𝑠𝑜𝑙𝑖𝑑 𝑠ℎ𝑎𝑓𝑡, 𝑍𝑃 = =
2 16
3
𝜋𝑑 3 16 𝑇
𝑇 = 𝑍𝑝 𝜏𝑚𝑎𝑥 = 𝜏𝑚𝑎𝑥 ⟹ 𝑑 = √
16 𝜋 𝜏𝑚𝑎𝑥

3 16 6 × 103
𝑑=√ = 0.0778𝑚 = 77.8𝑚𝑚
𝜋 65 × 106

Example 4.3: The horizontal shaft AD is attached to a fixed base at D and is subjected to the
torques shown. A 44mm diameter hole has been drilled into portion CD of the shaft. Knowing
that the entire shaft is made of steel for which G= 77GPa, determine the angle of twist at end
A.

Solution
Use free-body diagrams to determine the torque in each shaft segment AB, BC, and CD. Then
𝑇𝐿
use 𝜙 = ∑𝑖 𝑖 𝑖 to determine the angle of twist at end A.
𝐺𝑖 𝐽𝑖
Passing a section through the shaft between A and B,

GIRUM MINDAYE 97
Strength of Materials AASTU

𝑊𝑒 𝑓𝑖𝑛𝑑, ↺ + ∑ 𝑀𝑥 = 0: 𝑇𝐴𝐵 − 250 = 0 ⟹ 𝑇𝐴𝐵 = 250𝑁𝑚

Passing now a section between B and C, we have

↺ + ∑ 𝑀𝑥 = 0: 𝑇𝐵𝐶 − 250 − 2000 = 0 ⟹ 𝑇𝐵𝐶 = 2250𝑁𝑚

Polar Moments of Inertia


𝜋𝑟 4 𝜋 × 0.034
𝐽𝐴𝐵 = = = 0.0795 × 10−6 𝑚4
2 2
𝜋𝑟 4 𝜋 × 0.0154
𝐽𝐵𝐶 = = = 1.272 × 10−6 𝑚4
2 2
𝜋(𝑟 4 0 − 𝑟 4 𝑖 ) 𝜋(0.034 − 0.0224 )
𝐽𝐶𝐷 = = = 0.904 × 10−6 𝑚4
2 2
Angle of Twist. Recalling that G=77GPa for the entire shaft, we have

𝑇𝑖 𝐿𝑖 1 𝑇𝐴𝐵 𝐿𝐴𝐵 𝑇𝐵𝐶 𝐿𝐵𝐶 𝑇𝐶𝐷 𝐿𝐶𝐷


𝜙𝐴 = ∑ = ( + + )
𝐺𝑖 𝐽𝑖 𝐺 𝐽𝐴𝐵 𝐽𝐵𝐶 𝐽𝐶𝐷
𝑖
1 250 × 0.4 2250 × 0.2 2250 × 0.6
𝜙𝐴 = ( + + ) = 0.0403𝑟𝑎𝑑 = 2.310
77 × 109 0.0795 × 10−6 1.272 × 10−6 0.904 × 10−6

GIRUM MINDAYE 98
Strength of Materials AASTU

Example 4.4: A steel shaft and an aluminum tube are connected to a fixed support and to a
rigid disk as shown in the cross section. Knowing that the initial stresses are zero, determine
the maximum torque T0 that can be applied to the disk if the allowable stresses are 120 MPa
in the steel shaft and 70MPa in the aluminum tube. Use G=77GPa for steel and G=27GPa for
aluminum.

Solution
We know that the applied load is resisted by both the shaft and the tube, but we do not know
the portion carried by each part. Thus, we need to look at the deformations. We know that both
the shaft and tube are connected to the rigid disk and that the angle of twist is therefore the
same for each. Once we know the portion of the torque carried by each part, we can use the
allowable stress for each to determine which one governs and use this to determine the
maximum torque.
We first draw a free-body diagram of the disk (Fig. 1) and find
𝑇0 = 𝑇1 + 𝑇2

Fig. 1 Free-body diagram of end cap.


Knowing that the angle of twist is the same for the shaft and tube, we write

𝑇1 𝐿1 𝑇2 𝐿2
𝜙1 = 𝜙2 ; =
𝐽1 𝐽2
𝑇1 (0.5) 𝑇2 (0.5)
; −6
=
2.003 × 10 0.614 × 10−6

GIRUM MINDAYE 99
Strength of Materials AASTU

𝑇2 = 0.874𝑇1
We need to determine which part reaches its allowable stress first, and so we arbitrarily assume
that the requirement τalum≤70MPa is critical. For the aluminum tube in Fig. 2, we have
𝜏𝑎𝑙𝑢𝑚 𝐽1 (70 × 106 )(2.003 × 10−6 )
𝑇1 = = = 3690𝑁𝑚
𝑐1 0.038
compute the corresponding value T2 and then find the maximum shearing stress in the steel
shaft of Fig. 3.
𝑇2 = 0.874𝑇1 == 0.874(3690) = 3225𝑁𝑚
𝑇2 𝑐2 (3225)(0.025)
𝜏𝑠𝑡𝑒𝑒𝑙 = = = 131.3 × 106 𝑁⁄𝑚2 = 131.3𝑀𝑃𝑎
𝐽2 0.614 × 10−6
Note that the allowable steel stress of 120 MPa is exceeded; the assumption was wrong. Thus,
the maximum torque T0 will be obtained

𝜏𝑠𝑡𝑒𝑒𝑙 𝐽2 (120 × 106 )(0.614 × 10−6 )


𝑇2 = = = 2950𝑁𝑚
𝑐2 0.025

𝑤𝑒 ℎ𝑎𝑣𝑒, 𝑇2 = 0.874𝑇1 → 2950 = 0.874𝑇1 ⟹ 𝑇1 = 3375𝑁𝑚


We obtain the maximum permissible torque:
𝑇0 = 𝑇1 + 𝑇2 = 3375 + 2950 = 6325𝑁𝑚 = 6.325𝑘𝑁𝑚
This example illustrates that each part must not exceed its maximum allowable stress. Since
the steel shaft reaches its allowable stress level first, the maximum stress in the aluminum shaft
is below its maximum.
Example 4.5: A shaft consisting of a steel tube of 50mm outer diameter is to transmit 100kW
of power while rotating at a frequency of 20Hz. Determine the tube thickness that should be
used if the shearing stress is not to exceed 60MPa.
𝑃 100 × 103
𝑇= = = 795.8𝑁𝑚
2𝜋𝑓 2𝜋 × 20
𝑇 795.8
𝑇 = 𝑍𝑝 𝜏𝑚𝑎𝑥 ⟹ 𝑍𝑝 = = = 13.26 × 10−6 𝑚3
𝜏𝑚𝑎𝑥 60 × 106

𝜋(𝑟 4 0 − 𝑟 4 𝑖 ) 𝜋(0.0254 − 𝑟 4 𝑖 )
𝑍𝑃 = =
2𝑟0 2 × 0.025
Equating the above two equations
𝜋(0.0254 − 𝑟 4 𝑖 )
= 13.26 × 10−6 ⟹ 𝑟𝑖 = 0.0206𝑚 = 20.6𝑚𝑚
2 × 0.025
The corresponding tube thickness is
𝑟0 − 𝑟𝑖 = 25 − 20.6 = 4.4𝑚𝑚

A tube thickness of 5mm should be used.

GIRUM MINDAYE 100


Strength of Materials AASTU

5 COMPOUND (COMBINED) STRESSES


5.1 Combined Stresses

We learned before how to determine the stress in a member subjected to either an internal
axial force, a shear force, a bending moment, or a torsional moment. Most often, however,
the cross section of a member will be subjected to several of these loadings simultaneously,
and when this occurs, then the method of superposition should be used to determine the
resultant stress. The following procedure for analysis provides a method for doing this.
In a single element of a force system, the stresses developed at that element may be written
as:
1) Normal stresses
𝑃
𝑎. 𝐷𝑢𝑒 𝑡𝑜 𝑎𝑛 𝑎𝑥𝑖𝑎𝑙 𝑓𝑜𝑟𝑐𝑒: 𝜎 = 𝐴
𝑀𝑦
𝑏. 𝐷𝑢𝑒 𝑡𝑜 𝑎 𝑏𝑒𝑛𝑑𝑖𝑛𝑔 𝑚𝑜𝑚𝑒𝑛𝑡: 𝜎 = −
𝐼
2) Shearing stresses
𝑇𝑐
𝑎. 𝐷𝑢𝑒 𝑡𝑜 𝑎 𝑡𝑜𝑟𝑞𝑢𝑒 (𝑐𝑖𝑟𝑐𝑢𝑙𝑎𝑟 𝑠ℎ𝑎𝑓𝑡𝑠): 𝜏𝑚𝑎𝑥 = 𝐽
𝑉𝑄
𝑏. 𝐷𝑢𝑒 𝑡𝑜 𝑎 𝑠ℎ𝑒𝑎𝑟 𝑓𝑜𝑟𝑐𝑒: 𝜏 =
𝐼𝑡
Combined stresses are developed when two or more different force systems simultaneously
act on a structural member.
If structural members are subjected to a simultaneous action of both bending and axial loads,
the combined stresses can be obtained by superposition of the bending stresses and the axial
stresses.

5.2 Plane Stress


It was shown in Section 1.8 that the general state of stress at a point is characterized by six
normal and shear-stress components, shown in Figure 5-1a. This state of stress, however, is
not often encountered in engineering practice. Instead, most loadings are coplanar, and so the
stress these loadings produce can be analyzed in a single plane. When this is the case, the
material is then said to be subjected to plane stress. The general state of plane stress at a point,
shown in Figure 5-1b, is therefore represented by a combination of two normal-stress

GIRUM MINDAYE 101


Strength of Materials AASTU

components, σx, σy, and one shear-stress component, τxy, which act on only four faces of the
element.

Figure 5-1: (a) General state of stress (b) Plane stress

5.3 Plane Stress Transformation

The method of transforming the normal and shear stress components from the x, y to the x', y'
coordinate axes can be developed in a general manner and expressed as a set of stress-
transformation equations.
Sign Convention. To apply these equations we must first establish a sign convention for the
stress components. As shown in Figure 5-2, the +x and +x' axes are used to define the outward
normal on the right-hand face of the element, so that σx and σx’ are when they act in the positive
x and x' directions, and τxy and τx’y’ are positive when they act in the positive y and y' directions.

Figure 5-2: Positive sign convention


The orientation of the face upon which the normal and shear stress components are to be
determined will be defined by the angle θ, which is measured from the +x axis to the +x' axis,
Figure 5-2b. Notice that the unprimed and primed sets of axes in this figure both form right-

GIRUM MINDAYE 102


Strength of Materials AASTU

handed coordinate systems; that is, the positive z (or z') axis always points out of the page. The
angle θ will be positive when it follows the curl of the right-hand fingers, i.e., counterclockwise
as shown in Figure 5-2b.
Normal and Shear Stress Components. Using this established sign convention, the element in
Figure 5-3a is sectioned along the inclined plane and the segment shown in Figure 5-3b is
isolated. Assuming the sectioned area is ∆A, then the horizontal and vertical faces of the
segment have an area of ∆A sinθ and ∆A cosθ, respectively.

Figure 5-3: Stress transformation equations are determined by considering an arbitrary


prismatic wedge element.
The resulting free-body diagram of the segment is shown in Figure 5-3c. If we apply the
equations of equilibrium along the x' and y' axes, we can obtain a direct solution for σx' and
τx’y’. We have

↗ + ∑ 𝐹𝑥′ = 0; 𝜎𝑥′ ∆𝐴 − (𝜏𝑥𝑦 ∆𝐴 sin 𝜃) cos 𝜃 − (𝜎𝑦 ∆𝐴 sin 𝜃) sin 𝜃 − (𝜏𝑥𝑦 ∆𝐴 cos 𝜃) sin 𝜃
− (𝜎𝑥 ∆𝐴 cos 𝜃) cos 𝜃 = 0
𝜎𝑥′ = 𝜎𝑥 cos2 𝜃 + 𝜎𝑦 sin2 𝜃 + 𝜏𝑥𝑦 (2 sin 𝜃 cos 𝜃)
↖ + ∑ 𝐹𝑦 ′ = 0; 𝜏𝑥 ′ 𝑦 ′ ∆𝐴 + (𝜏𝑥𝑦 ∆𝐴 sin 𝜃) sin 𝜃 − (𝜎𝑦 ∆𝐴 sin 𝜃) cos 𝜃
− (𝜏𝑥𝑦 ∆𝐴 cos 𝜃) cos 𝜃 + (𝜎𝑥 ∆𝐴 cos 𝜃) sin 𝜃 = 0
𝜏𝑥 ′ 𝑦 ′ = (𝜎𝑦 − 𝜎𝑥 ) sin 𝜃 cos 𝜃 + 𝜏𝑥𝑦 (cos2 𝜃 − sin2 𝜃)

GIRUM MINDAYE 103


Strength of Materials AASTU

To simplify these two equations, use the trigonometric identities


1 − cos 2𝜃 1 + cos 2𝜃
sin 2𝜃 = 2 sin 𝜃 cos 𝜃 , sin2 𝜃 = 𝑎𝑛𝑑 cos2 𝜃 =
2 2
Therefore,
𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
𝜎𝑥′ = + cos 2𝜃 + 𝜏𝑥𝑦 sin 2𝜃
2 2
𝜎𝑥 − 𝜎𝑦
𝜏𝑥 ′ 𝑦 ′ = − sin 2𝜃 + 𝜏𝑥𝑦 cos 2𝜃
2
If the normal stress acting in the y' direction is needed, it can be obtained by simply substituting
θ + 90° for θ into σx’ equation, Figure 5-3d. Since cos(2𝜃 + 1800 ) = − cos 2𝜃 and
sin(2𝜃 + 1800 ) = − sin 2𝜃. This yields
𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
𝜎𝑦′ = − cos 2𝜃 − 𝜏𝑥𝑦 sin 2𝜃
2 2
5.4 Principle Stresses Mohr’s Circle.

In this section, we will show how to apply the equations for plane-stress transformation using
a graphical procedure that is often convenient to use and easy to remember. Furthermore, this
approach will allow us to “visualize” how the normal and shear stress components σx’ and τx’y’
vary as the plane on which they act changes its direction.
𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
𝜎𝑥′ − = cos 2𝜃 + 𝜏𝑥𝑦 sin 2𝜃
2 2
𝜎𝑥 − 𝜎𝑦
𝜏𝑥 ′ 𝑦 ′ = − sin 2𝜃 + 𝜏𝑥𝑦 cos 2𝜃
2
then the parameter θ can be eliminated by squaring each equation and adding them together.
The result is
𝜎𝑥 + 𝜎𝑦 2 𝜎𝑥 − 𝜎𝑦 2
[𝜎𝑥′ − ( )] + 𝜏 2 𝑥 ′ 𝑦 ′ = ( ) + 𝜏 2 𝑥𝑦
2 2
Finally, since σx, σy, and τxy are known constants, then the above equation can be written in a
more compact form as
2
(𝜎𝑥′ − 𝜎𝑎𝑣𝑔 ) + 𝜏 2 𝑥 ′ 𝑦 ′ = 𝑅 2
𝜎𝑥 + 𝜎𝑦
𝑤ℎ𝑒𝑟𝑒, 𝜎𝑎𝑣𝑔 = 𝑎𝑛𝑑 𝑅
2
𝜎𝑥 − 𝜎𝑦 2
= √( ) + 𝜏 2 𝑥𝑦
2
which is the equation of a circle of radius R centered at the
point C of abscissa σavg and ordinate 0 (Figure 5-4). Due to the
symmetry of the circle about the horizontal axis, the same
result is obtained if a point N of abscissa σx’ and ordinate –τx’y’
is plotted instead of M. (Figure 5-5). This circle is called
Mohr’s circle, because it was developed by the German
engineer Otto Mohr.

GIRUM MINDAYE 104


Strength of Materials AASTU

Figure 5-4: Circular relationship of transformed stresses.

Figure 5-5: Equivalent formation of stress transformation circle.


The points A and B where the circle of Figure 5-4 intersects the horizontal axis are of special
interest: point A corresponds to the maximum value of the normal stress σx’, while point B
corresponds to its minimum value. Both points also correspond to a zero value of the shearing
stress τx’y’. Thus, the values θp of the parameter θ, which correspond to points A and B can be
obtained by setting τx’y’= 0.
𝜎𝑥 − 𝜎𝑦
𝜏𝑥 ′ 𝑦 ′ = − sin 2𝜃 + 𝜏𝑥𝑦 cos 2𝜃
2
𝜎𝑥 − 𝜎𝑦
0=− sin 2𝜃 + 𝜏𝑥𝑦 cos 2𝜃
2

GIRUM MINDAYE 105


Strength of Materials AASTU

2𝜏𝑥𝑦
tan 2𝜃𝑝 =
𝜎𝑥 − 𝜎𝑦

This equation defines two values 2θp that are 1800 apart and thus two values θ that are 900
apart. Either value can be used to determine the orientation of the corresponding element
(Figure 5-6).

Figure 5-6: Principal stresses.


The planes containing the faces of the element obtained in this way are the principal planes
of stress at point Q, and the corresponding values σmax and σmin exerted on these planes are
the principal stresses at Q. Since both values θp are obtained by setting τx’y’ =0, it is clear that
no shearing stress is exerted on the principal planes. From Figure 5-4,
𝜎𝑚𝑎𝑥 = 𝜎𝑎𝑣𝑔 + 𝑅 𝑜𝑟 𝜎𝑚𝑖𝑛 = 𝜎𝑎𝑣𝑔 − 𝑅

Substituting for σavg and R,

𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦 2
𝜎𝑚𝑎𝑥,𝑚𝑖𝑛 = ± √( ) + 𝜏 2 𝑥𝑦
2 2

Unless it is possible, to tell by inspection which of these principal planes is subjected to σmax
and which is subjected to σmin, it is necessary to substitute one of the values θp into equation
𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
𝜎𝑥′ = + cos 2𝜃 + 𝜏𝑥𝑦 sin 2𝜃
2 2
in order to determine which corresponds to the maximum value of the normal stress.
Referring again to Figure 5-4, points D and E located on the vertical diameter of the circle
correspond to the largest value of the shearing stress τx’y’. Since the abscissa of points D and E
is 𝜎𝑎𝑣𝑔 = (𝜎𝑥 + 𝜎𝑦 )⁄2 , the values θs of the parameter θ corresponding to these points are
obtained by setting 𝜎𝑥′ = (𝜎𝑥 + 𝜎𝑦 )⁄2. The sum of the last two terms in that equation must be
zero. Thus, for θ=θs,
𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
𝜎𝑥′ = + cos 2𝜃 + 𝜏𝑥𝑦 sin 2𝜃
2 2

GIRUM MINDAYE 106


Strength of Materials AASTU

𝜎𝑥 + 𝜎𝑦 𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
= + cos 2𝜃𝑠 + 𝜏𝑥𝑦 sin 2𝜃𝑠
2 2 2
𝜎𝑥 − 𝜎𝑦
cos 2𝜃𝑠 + 𝜏𝑥𝑦 sin 2𝜃𝑠 = 0
2
𝜎𝑥 − 𝜎𝑦
∴ tan 2𝜃𝑠 = −
2𝜏𝑥𝑦

This equation defines two values 2θs that are 1800 apart, and thus two values θs that are 900
apart. Either of these values can be used to determine the orientation of the element
corresponding to the maximum shearing stress (Figure 5-7). Figure 5-7 shows that the
maximum value of the shearing stress is equal to the radius R of the circle.

𝜎𝑥 − 𝜎𝑦 2
𝜏𝑚𝑎𝑥 = 𝑅 = √( ) + 𝜏 2 𝑥𝑦
2
As observed earlier, the normal stress corresponding to the condition of maximum shearing
stress is
𝜎𝑥 + 𝜎𝑦
𝜎 ′ = 𝜎𝑎𝑣𝑔 =
2

Figure 5-7: Maximum shearing stress.


tan2θs is the negative reciprocal of tan2θp . Thus, angles 2θs and 2θp are 900 apart, and therefore
angles θs and θp are 450 apart. Thus, the planes of maximum shearing stress are at 450 to the
principal planes.

𝜃𝑠 + 450 = 𝜃𝑝 ⟹ 𝜽𝒔 = 𝜽𝒑 − 𝟒𝟓𝟎

GIRUM MINDAYE 107


Strength of Materials AASTU

5.5 Worked Example

Example 5.1: A 50mmx75mm, 1.5m long beam is loaded as shown in the figure below.
Determine the maximum tensile and compressive stresses acting normal to the section through
the beam.

Solution
Axial stress:
𝑃 25,000
𝜎𝑎𝑥𝑖𝑎𝑙 = = = 6.67𝑀𝑃𝑎
𝐴 50 × 75
Bending Stress:
𝑀𝑐 (1.013 × 106 )(75⁄2)
𝜎𝑏𝑒𝑛𝑑𝑖𝑛𝑔 = = = 21.6𝑀𝑃𝑎
𝐼 50 × 753
12
Using the superposition principal:
𝑃 𝑀𝑐
𝜎𝑡𝑜𝑝 = + − = 6.67 − 21.6 = −14.93𝑀𝑃𝑎
𝐴 𝐼
𝑃 𝑀𝑐
𝜎𝑏𝑜𝑡 =+ + = 6.67 + 21.6 = 28.27𝑀𝑃𝑎
𝐴 𝐼

From the symmetry of the triangles:


𝑦 75
= ⟹ 𝑦 = 25.92𝑚𝑚
14.93 14.93 + 28.27
1
𝐶 = (50) [ (25.92 × 14.93)] × 10−3 = 9.68𝑘𝑁
2

GIRUM MINDAYE 108


Strength of Materials AASTU

1
𝑇 = (50) [ (75 − 25.92)(28.27)] × 10−3 = 34.68𝑘𝑁
2
The axial force:
𝑃 = 𝑇 − 𝐶 = 34.68 − 9.68 = 25𝑘𝑁
32.72 17.28
𝑀 = (𝑇𝑑1 ) + (𝐶𝑑2 ) = 34.68 × + 9.68 × = 1.30𝑘𝑁𝑚
1000 1000
𝑀 − 𝑃𝑒 = 1.3 − [25 × (37.5 − 25.92)] = 1.013𝑘𝑁𝑚
Example 5.2: A force of 300kN is applied to the edge of the member shown in figure. Neglect
the weight of the member and determine the state of stress at points B and C.

Solution
The member is sectioned through B and C. For equilibrium at the section, there must be an
axial force of 300kN acting through the centroid and a bending moment of 45.0kNm about the
centroidal principal axis.

Axial stress:
𝑃 300,000
𝜎𝑎𝑥𝑖𝑎𝑙 = − = = 10𝑀𝑃𝑎
𝐴 100 × 300
Bending Stress:
𝑀𝑐 (45 × 106 )(150)
𝜎𝑏𝑒𝑛𝑑𝑖𝑛𝑔 = = = 30𝑀𝑃𝑎
𝐼 100 × 3003
12
Using the superposition principal:
𝑃 𝑀𝑐
𝜎𝐶 = −− = −10 − 30 = −40𝑀𝑃𝑎
𝐴 𝐼
𝑃 𝑀𝑐
𝜎𝐵 = − + = −10 + 30 = 20𝑀𝑃𝑎
𝐴 𝐼
NOTE: The resultant stress distribution over the cross section is shown in figure,

GIRUM MINDAYE 109


Strength of Materials AASTU

where the location of the line of zero stress can be determined by proportional triangles; i.e.,
20 40
= ⟹ 𝑥 = 100𝑚𝑚
𝑥 300 − 𝑥
Example 5.3: The member shown in Figure has a rectangular cross section. Determine the state
of stress that the loading produces at point C and point D.

Solution
The support reactions on the member have been determined and are shown in figure.

If the left segment AC of the member is considered,

then the resultant internal loadings at the section consist of a normal force, a shear force, and a
bending moment. They are
𝑁 = 16.45𝑘𝑁, 𝑉 = 21.93𝑘𝑁, 𝑀 = 32.89𝑘𝑁𝑚
Stress Components at C.
a) Normal Force: The uniform normal-stress distribution acting over the cross section is
produced by the normal force. At point C,

GIRUM MINDAYE 110


Strength of Materials AASTU

𝑃 16,450
𝜎𝐶,𝑎𝑥𝑖𝑎𝑙 = − =− = −1.32𝑀𝑃𝑎
𝐴 50 × 250
b) Shear Force: Since point C is located at the top of the member (y=c).
3 𝑐2 − 𝑦2
𝜏𝐶 = 𝑉 ⟹ 𝜏𝐶 = 0
4 𝑏𝑐3
c) Bending Moment: Point C is located at y = c = 0.125 m from the neutral axis, so the
bending stress at C is
𝑀𝑐 (32.89 × 106 )(125)
𝜎𝐶,𝑏𝑒𝑛𝑑𝑖𝑛𝑔 = = = 63.16𝑀𝑃𝑎
𝐼 50 × 2503
12
d) Superposition. There is no shear-stress component. Adding the normal stresses gives a
compressive stress at C having a value of
𝑃 𝑀𝑐
𝜎𝐶 = − − = −1.32 − 63.16 = −64.5𝑀𝑃𝑎
𝐴 𝐼
Stress Components at D.
𝑃 16,450
a) Normal Force. This is the same as at C, 𝜎𝐷 = − 𝐴 = 50×250 = 1.32𝑀𝑃𝑎
b) Shear Force. Since D is at the neutral axis, and the cross section is rectangular, we can use
the special form of the shear formula,
3 𝑉 3 21.93 × 103
𝜏𝐷 = = ( ) = 2.63𝑀𝑃𝑎
2 𝐴 2 50 × 250

c) Bending Moment. Here D is on the neutral axis (y=0) and so (𝜎𝐶,𝑏𝑒𝑛𝑑𝑖𝑛𝑔 = 0).
d) Superposition. The resultant stress on the element is
𝑃 𝑀𝑐
𝜎𝐷 = − − = −1.32 + 0 = −1.32𝑀𝑃𝑎
𝐴 𝐼
𝜏𝐷 = 2.63𝑀𝑃𝑎

Example 5.4: Find the stress distribution at the section ABCD for the block shown in the
figure.

Solution

GIRUM MINDAYE 111


Strength of Materials AASTU

0.3 0.15
𝑃 = 64𝑘𝑁, 𝑀𝑦 = (64) ( ) = 9.6𝑘𝑁𝑚, 𝑀𝑧 = (64) (0.75 + ) = 9.6𝑘𝑁𝑚
2 2
𝐴 = 150 × 300 = 45,000𝑚𝑚2
150 × 3003 300 × 1503
𝐼𝑦 = = 3.375 × 108 𝑚𝑚4 , 𝐼𝑧 = = 0.844 × 108 𝑚𝑚4
12 12
𝑃 64,00
𝜎 𝑎𝑥𝑖𝑎𝑙 = − = = −1.42𝑀𝑃𝑎
𝐴 45,000
𝑀𝑦 𝑐𝑦 (9.6 × 106 )(150)
𝜎𝑦,𝑏𝑒𝑛𝑑𝑖𝑛𝑔 = = = 4.27𝑀𝑃𝑎
𝐼𝑦 3.375 × 108
𝑀𝑧 𝑐𝑧 (9.6 × 106 )(75)
𝜎𝑧,𝑏𝑒𝑛𝑑𝑖𝑛𝑔 = = = 8.53𝑀𝑃𝑎
𝐼𝑧 0.844 × 108
𝑃 𝑀𝑦 𝑐𝑦 𝑀𝑧 𝑐𝑧
𝜎𝐴 = − − − = −1.42 − 4.27 − 8.53 = −14.22𝑀𝑃𝑎
𝐴 𝐼𝑦 𝐼𝑧
𝑃 𝑀𝑦 𝑐𝑦 𝑀𝑧 𝑐𝑧
𝜎𝐵 = − + − = −1.42 + 4.27 − 8.53 = −5.68𝑀𝑃𝑎
𝐴 𝐼𝑦 𝐼𝑧
𝑃 𝑀𝑦 𝑐𝑦 𝑀𝑧 𝑐𝑧
𝜎𝐶 = − + + = −1.42 + 4.27 + 8.53 = 11.38𝑀𝑃𝑎
𝐴 𝐼𝑦 𝐼𝑧
𝑃 𝑀𝑦 𝑐𝑦 𝑀𝑧 𝑐𝑧
𝜎𝐷 = − − + = −1.42 − 4.27 + 8.53 = 2.84𝑀𝑃𝑎
𝐴 𝐼𝑦 𝐼𝑧

Example 5.5: A short 100mm square steel bar with a 50mm diameter axial hole is built at the
base and is loaded at the top as shown in the figure. Determine the value of the force P so that
the maximum normal stress at the fixed-end would not exceed 140MPa.

GIRUM MINDAYE 112


Strength of Materials AASTU

Solution
𝜋(50)2
𝐴 = 1002 − = 8,037𝑚𝑚2
4
𝑏ℎ3 𝜋(25)4
𝐼= − = 8.03 × 106 𝑚𝑚4
12 4
𝑃 cos 300 (𝑃 sin 300 × 400)𝑐 (𝑃 cos 300 × 50)𝑐
𝜎𝐴 = − − + = 140𝑀𝑃𝑎
𝐴 𝐼 𝐼
𝑃 cos 300 (𝑃 sin 300 × 400) × 50 (𝑃 cos 300 × 50) × 50
− − + = 140𝑀𝑃𝑎 ⟹ 𝑃
8,037 8.03 × 106 8.03 × 106
= 129𝑘𝑁
𝑃 cos 300 (𝑃 sin 300 × 400)𝑐 (𝑃 cos 300 × 50)𝑐
𝜎𝐵 = − + − = 140𝑀𝑃𝑎
𝐴 𝐼 𝐼

𝑃 cos 300 (𝑃 sin 300 × 400) × 50 (𝑃 cos 300 × 50) × 50


− + − = 140𝑀𝑃𝑎 ⟹ 𝑃
8,037 8.03 × 106 8.03 × 106
= 161𝑘𝑁
𝑇ℎ𝑒 𝑠𝑎𝑓𝑒 𝑓𝑜𝑟𝑐𝑒, 𝑃 = 129𝑘𝑁
Example 5.6: The state of plane stress at a point is represented on the element shown in figure.
Determine the state of stress at this point on another element oriented 30° clockwise from the
position shown.

Solution
From the established sign convention,
𝜎𝑥 = −80𝑀𝑃𝑎, 𝜎𝑦 = 50𝑀𝑃𝑎, 𝜏𝑥𝑦 = −25𝑀𝑃𝑎
Plane CD. To obtain the stress components on plane CD, the positive x' axis must be directed
outward, perpendicular to CD, and the associated y' axis is directed along CD. The angle
measured from the x to the x' axis is θ = -30° (clockwise).

GIRUM MINDAYE 113


Strength of Materials AASTU

For θ=-300
𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
𝜎𝑥′ = + cos 2𝜃 + 𝜏𝑥𝑦 sin 2𝜃
2 2
−80 + 50 −80 − 50
𝜎𝑥′ = + cos 2(−300 ) + (−25) sin 2(−300 ) = −25.8𝑀𝑃𝑎
2 2
𝜎𝑥 − 𝜎𝑦
𝜏𝑥 ′ 𝑦 ′ = − sin 2𝜃 + 𝜏𝑥𝑦 cos 2𝜃
2
−80 − 50
𝜏𝑥 ′ 𝑦 ′ = − sin 2(−300 ) + (−25) cos 2(−300 ) = −68.8𝑀𝑃𝑎
2
The negative signs indicate that σx’ and τx’y’ act in the negative x' and y' directions, respectively.

Plane BC. Establishing the y' axis outward from plane BC, then between the x and y' axes, θ =-
300+900=60° (counterclockwise).
−80 + 50 −80 − 50
𝜎𝑦′ = + cos 2(600 ) + (−25) sin 2(600 ) = −4.15𝑀𝑃𝑎
2 2
−80 − 50
𝜏𝑥 ′ 𝑦 ′ = − sin 2(600 ) + (−25) cos 2(600 ) = 68.8𝑀𝑃𝑎
2
Here τx’y’ has been calculated twice in order to provide a check. The negative sign for σy’
indicates that this stress acts in the negative y' direction.

Example 5.6: For the state of plane stress shown in figure, determine (a) the principal planes,
(b) the principal stresses, (c) the maximum shearing stress and the corresponding normal stress.

a) Principal Planes. Following the usual sign convention, the stress components are
𝜎𝑥 = 50𝑀𝑃𝑎, 𝜎𝑦 = −10𝑀𝑃𝑎, 𝜏𝑥𝑦 = 40𝑀𝑃𝑎
2𝜏𝑥𝑦 2(40) 4
tan 2𝜃𝑝 = = =
𝜎𝑥 − 𝜎𝑦 50 − (−10) 3

2𝜃𝑝 = 53.10 , 𝑎𝑛𝑑 2𝜃𝑝 = 53.10 + 1800 = 233.10

GIRUM MINDAYE 114


Strength of Materials AASTU

𝜃𝑝 = 26.60 , 𝑎𝑛𝑑 116.60

b) Principal Stresses.

𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦 2
𝜎𝑚𝑎𝑥,𝑚𝑖𝑛 = ± √( ) + 𝜏 2 𝑥𝑦
2 2

2
50 + (−10) 50 − (−10)
𝜎𝑚𝑎𝑥,𝑚𝑖𝑛 = ± √( ) + 402 = 20 ± 50
2 2

𝜎𝑚𝑎𝑥 = 20 + 50 = 70𝑀𝑃𝑎
𝜎𝑚𝑖𝑛 = 20 − 50 = −30𝑀𝑃𝑎
Making 2θ=53.10,
𝜎𝑥 + 𝜎𝑦 𝜎𝑥 − 𝜎𝑦
𝜎𝑥′ =
+ cos 2𝜃 + 𝜏𝑥𝑦 sin 2𝜃
2 2
50 + (−10) 50 − (−10)
𝜎𝑥′ = + cos 53.10 + 40 sin 53.10 = 70𝑀𝑃𝑎 = 𝜎𝑚𝑎𝑥
2 2
And 2θ=233.10
50 + (−10) 50 − (−10)
𝜎𝑦′ = + cos 233.10 + 40 sin 233.10 = −30𝑀𝑃𝑎 = 𝜎𝑚𝑖𝑛
2 2

It is confirmed that the normal stress exerted on face BC of the element is the maximum
stress.

c) Maximum Shearing Stress


2
𝜎𝑥 − 𝜎𝑦 2 50 − (−10)
𝜏𝑚𝑎𝑥 = √( ) + 𝜏 2 𝑥𝑦 = √( ) + 402 = 50𝑀𝑃𝑎
2 2
The normal stress on each of the four faces of the element is given by:
𝜎𝑥 + 𝜎𝑦 50 + (−10)
𝜎 ′ = 𝜎𝑎𝑣𝑔 = = = 20𝑀𝑃𝑎
2 2

GIRUM MINDAYE 115


Strength of Materials AASTU

Since σmax and σmin have opposite signs, τmax actually represents the maximum value of the
shearing stress at the point. The orientation of the planes of maximum shearing stress and the
sense of the shearing stresses are determined by passing a section along the diagonal plane AC
of the element. Since the faces AB and BC of the element are in the principal planes, the
diagonal plane AC must be one of the planes of maximum shearing stress. Furthermore, the
equilibrium conditions for the prismatic element ABC require that the shearing stress exerted
on AC be directed as shown.

The cubic element corresponding to the maximum shearing stress is shown.

GIRUM MINDAYE 116


Strength of Materials AASTU

6 DEFLECTION IN BEAMS

6.1 Introduction

The deflection of a beam must often be limited in order to prevent the cracking of any attached
brittle materials such as concrete or plaster. Most importantly, though, slopes and
displacements must be determined in order to find the reactions if the beam is statically
indeterminate. In this section, we will develop two important differential equations that relate
the internal moment in a beam to the displacement and slope of its elastic curve. To derive
these relationships, we will limit the analysis to the most common case of an initially straight
beam that is elastically deformed by loads applied perpendicular to the beam’s x axis and lying
in the x-v plane of symmetry for the beam’s cross-sectional area, Figure 6-1a. Due to the
loading, the deformation of the beam is caused by both the internal shear force and bending
moment. If the beam has a length that is much greater than its depth, the greatest deformation
will be caused by bending, and therefore we will direct our attention to its effects.

Figure 6-1: Beam


When the internal moment M deforms the element of the beam, each cross section remains
plane and the angle between them becomes dθ, Figure 6-1b. The arc dx that represents a portion
of the elastic curve intersects the neutral axis for each cross section. The radius of curvature
for this arc is defined as the distance ρ, which is measured from the center of curvature O’ to
dx. Any arc on the element other than dx is subjected to a normal strain. For example, the strain
in arc ds, located at a position y from the neutral axis, is 𝜖 = (𝑑𝑠 ′ − 𝑑𝑠)⁄𝑑𝑠. However,
ds=dx=ρdθ and ds’=(ρ-y)dθ, and so
𝑑𝑠 ′ − 𝑑𝑠 (𝜌 − 𝑦)𝑑𝜃 − 𝜌𝑑𝜃 1 𝜖
𝜖= = ⟹ =−
𝑑𝑠 𝑑𝑠 𝜌 𝑦
If the material is homogeneous and behaves in a linear elastic manner, then Hooke’s law
applies, ϵ=σ/E. Also, since the flexure formula applies, σ=-My/I. Combining these equations
and substituting into the above equation, we have
1 𝑀𝑦
=
𝜌 𝐸𝐼
Where,

GIRUM MINDAYE 117


Strength of Materials AASTU

ρ = the radius of curvature at a specific point on the elastic curve (1/ ρ is referred to as the
curvature)
M= the internal moment in the beam at the point where is to be determined
E= the material’s modulus of elasticity
I= the beam’s moment of inertia computed about the neutral axis
The product EI in this equation is referred to as the flexural rigidity and it is always a positive
quantity.
Sign Convention. Positive bending moment tends to bend the segment concave upward, as
shown in Figure 6-2a. Positive deflection, v, is upward, and as a result, the positive slope angle
θ will be measured counterclockwise from the x-axis. The reason for this is shown in Figure
6-3. Here, positive increases dx and dv in x and v create an increase dθ that is counterclockwise.
Also, since the slope angle θ will be very small, its value in radians can be determined directly
from 𝜃 ≈ tan 𝜃 = 𝑑𝑣⁄𝑑𝑥 .

Figure 6-2: Sign Convention for bending moment, M

Figure 6-3: Beam deflection

6.2 The Double Integration Method

The equation of the elastic curve in Figure 6-3 will be defined by the coordinates v and x.
And so to find the deflection v = f(x) we must be able to represent the curvature (1/ρ) in terms
of v and x.
1 𝑀𝑦
=
𝜌 𝐸𝐼
𝑑𝜃 𝑀
𝑆𝑖𝑛𝑐𝑒 𝑑𝑥 = 𝜌𝑑𝜃, =
𝑑𝑥 𝐸𝐼
The slope of the deflection curve is the first derivative dv/dx of the expression for the deflection
v. In geometric terms, the slope is the increment dv in the deflection (Figure 6-3) divided by
the increment dx in the distance along the x-axis. Since dv and dx are infinitesimally small, the
slope dv/dx is equal to the tangent of the angle of rotation θ. Thus,

GIRUM MINDAYE 118


Strength of Materials AASTU

𝑑𝑣
= tan 𝜃
𝑑𝑥
In a similar manner, the following relationships are obtained:
𝑑𝑥 𝑑𝑣
cos 𝜃 = , sin 𝜃 =
𝑑𝑠 𝑑𝑠
𝑑𝑣
𝐴𝑙𝑠𝑜, 𝑠𝑖𝑛𝑐𝑒 tan 𝜃 ≈ 𝜃 𝑤ℎ𝑒𝑛 𝜃 𝑖𝑠 𝑠𝑚𝑎𝑙𝑙, 𝜃 ≈ 𝑡𝑎𝑛 𝜃 =
𝑑𝑥
Thus, if the rotations of a beam are small, assume that the angle of rotation θ and the slope
dv/dx are equal. (Note that the angle of rotation must be measured in radians.)
Take the derivative of θ with respect to x
𝑑𝜃 𝑑 2 𝑣
=
𝑑𝑥 𝑑𝑥 2
𝑑𝜃 𝑑 2 𝑣 𝑀
= =
𝑑𝑥 𝑑𝑥 2 𝐸𝐼
𝑑2𝑣 𝑀
=
𝑑𝑥 2 𝐸𝐼
𝑑𝑣 𝑀
∴ 𝑆𝑙𝑜𝑝𝑒, 𝜃= = ∫ 𝑑𝑥
𝑑𝑥 𝐸𝐼
𝑀
∴ 𝐷𝑒𝑓𝑙𝑒𝑐𝑡𝑖𝑜𝑛 (𝑒𝑙𝑎𝑠𝑡𝑖𝑐 𝑐𝑢𝑟𝑣𝑒), 𝑣=∬ 𝑑𝑥
𝐸𝐼
For constant EI;
𝑑𝑣 𝑑𝑣 1
𝑆𝑙𝑜𝑝𝑒, 𝐸𝐼𝜃 = 𝐸𝐼 =𝑀 ⟹ = ∫ 𝑀𝑑𝑥
𝑑𝑥 𝑑𝑥 𝐸𝐼
2
𝑑 𝑣 1
𝐷𝑒𝑓𝑙𝑒𝑐𝑡𝑖𝑜𝑛 (𝑒𝑙𝑎𝑠𝑡𝑖𝑐 𝑐𝑢𝑟𝑣𝑒), 𝐸𝐼 2 = 𝑀 ⟹ 𝑣 = ∬ 𝑀𝑑𝑥
𝑑𝑥 𝐸𝐼
For each integration, it is necessary to introduce a “constant of integration” and then solve for
the constants to obtain a unique solution for a particular problem. If the loading on a beam is
discontinuous—that is, it consists of a series of several distributed and concentrated loads, then
several functions must be written for the internal moment, each valid within the region between
the discontinuities. For example, consider the beam shown in Fig. 8–8. The internal moment
in regions AB, BC, and CD must be written in terms of the x1, x2 and x3 coordinates.
Boundary and Continuity Conditions. The constants of integration are determined by
evaluating the functions for slope or displacement at a particular point on the beam where the
value of the function is known. These values are called boundary conditions. For example, if
the beam is supported by a roller or pin, then it is required that the displacement be zero at
these points. Also, at a fixed support the slope and displacement are both zero.

Figure 6-4: Discontinuous loading in beam

GIRUM MINDAYE 119


Strength of Materials AASTU

6.3 Worked Examples

Examples 6.1: The cantilevered beam is subjected to a couple moment M0 at its end.
Determine the equation of the elastic curve. EI is constant.

Solution
The load tends to deflect the beam as shown in figure. By inspection, the internal moment can
be represented throughout the beam using a single x coordinate.

From the free-body diagram, with M acting in the positive direction, we have

𝑀 = 𝑀0 𝑓𝑜𝑟 0 ≤ 𝑥 ≤ 𝐿
𝑑2 𝑣
𝐸𝐼 = 𝑀0
𝑑𝑥 2
𝑑𝑣
𝐸𝐼 = ∫ 𝑀0 𝑑𝑥 = 𝑀0 𝑥 + 𝐶1
𝑑𝑥
𝑑𝑣
= 0 𝑎𝑡 𝑥 = 0, 0 = 𝑀0 (0) + 𝐶1 ⟹ 𝐶1 = 0
𝑑𝑥
𝑑𝑣
𝐸𝐼 = 𝑀0 𝑥
𝑑𝑥
1
𝐸𝐼𝑣 = ∫(𝑀0 𝑥 − 𝑀0 𝐿)𝑑𝑥 = 𝑀 𝑥 2 + 𝐶2
2 0
1
𝑣 = 0 𝑎𝑡 𝑥 = 𝐿, 0 = 𝑀0 (02 ) − 𝑀0 𝐿(0) + 𝐶2 ⟹ 𝐶2 = 0
2
1
𝐸𝐼𝑣 = 𝑀0 𝑥 2
2
𝑀0 𝑥 𝑀0 𝑥 2
𝜃= , 𝑎𝑛𝑑 𝑣 =
𝐸𝐼 2𝐸𝐼
Maximum slope and displacement occur at A (x=L), for which
𝑀0 𝐿 𝑀0 𝐿2
𝜃𝐴 = , 𝑎𝑛𝑑 𝑣𝐴 =
𝐸𝐼 2𝐸𝐼
Examples 6.2: The simply supported beam is subjected to the concentrated force as shown in
figure. Determine the maximum deflection of the beam. EI is constant.

GIRUM MINDAYE 120


Strength of Materials AASTU

Elastic Curve. Two coordinates must be used, since the moment function will change at B.
Here we will take x1 and x2, having the same origin at A.

Moment Function. From the free-body diagrams


𝑀1 = 2𝑥1 𝑓𝑜𝑟 0 ≤ 𝑥1 ≤ 2𝑚
𝑀2 = 2𝑥2 − 6(𝑥2 − 2) = −4𝑥2 + 12 𝑓𝑜𝑟 2 ≤ 𝑥2 ≤ 3𝑚
Slope and Elastic Curve.
a) For M1
𝑑 2 𝑣1
𝐸𝐼 = 𝑀1
𝑑𝑥1 2
𝑑 2 𝑣1
𝐸𝐼 = 2𝑥1
𝑑𝑥1 2
𝑑𝑣1
𝐸𝐼 = 𝑥1 2 + 𝑐1
𝑑𝑥1
1
𝐸𝐼𝑣1 = 𝑥1 3 + 𝑐1 𝑥1 + 𝑐2
3
𝑓𝑟𝑜𝑚 𝑏𝑜𝑢𝑛𝑑𝑎𝑟𝑦 𝑐𝑜𝑛𝑑𝑖𝑡𝑖𝑜𝑛, 𝑣𝐴 = 0 𝑎𝑡 𝑥2 = 0
1 3
𝐸𝐼(0) = (0 ) + 𝑐1 (0) + 𝑐2 ⟹ 𝑐2 = 0
3
𝟏
∴ 𝑬𝑰𝒗𝟏 = 𝒙𝟏 𝟑 + 𝒄𝟏 𝒙𝟏
𝟑
b) For M2
𝑑 2 𝑣2
𝐸𝐼 = 𝑀2
𝑑𝑥2 2
𝑑 2 𝑣2
𝐸𝐼 = −4𝑥2 + 12
𝑑𝑥2 2
𝑑𝑣2
𝐸𝐼 = −2𝑥2 2 + 12𝑥2 + 𝑐3
𝑑𝑥2
2
𝐸𝐼𝑣 = − 𝑥 3 + 6𝑥 2 + 𝑐3 𝑥 + 𝑐4
3
𝑓𝑟𝑜𝑚 𝑏𝑜𝑢𝑛𝑑𝑎𝑟𝑦 𝑐𝑜𝑛𝑑𝑖𝑡𝑖𝑜𝑛, 𝑣𝐶 = 0 𝑎𝑡 𝑥 = 3

GIRUM MINDAYE 121


Strength of Materials AASTU

2
𝐸𝐼(0) = − (33 ) + 6(32 ) + 𝑐3 (3) + 𝑐4 ⟹ 𝒄𝟒 = −𝟑𝟔 − 𝟑𝒄𝟑
3
𝟐
∴ 𝑬𝑰𝒗𝟐 = − 𝒙𝟐 𝟑 + 𝟔𝒙𝟐 𝟐 + 𝒄𝟑 (𝒙𝟐 − 𝟑) − 𝟑𝟔
𝟑
𝑑𝑣1 𝑑𝑣2
𝒇𝒓𝒐𝒎 𝒄𝒐𝒏𝒕𝒊𝒏𝒖𝒊𝒕𝒚 𝒄𝒐𝒏𝒅𝒊𝒕𝒊𝒐𝒏 𝒂𝒕 𝑩(𝒙𝟏 = 𝒙𝟐 = 𝟐𝒎), 𝐸𝐼𝜃𝐵 = 𝐸𝐼 = 𝐸𝐼
𝑑𝑥1 𝑑𝑥2
𝑑𝑣1
𝐸𝐼 = 𝑥1 2 + 𝑐1
𝑑𝑥1
𝜃𝐵 𝑎𝑡 𝑥1 = 2𝑚
𝑬𝑰𝜽𝑩 = 𝟐𝟐 + 𝒄𝟏 = 𝟒 + 𝒄𝟏
𝑑𝑣2
𝐸𝐼 = −2𝑥2 2 + 12𝑥2 + 𝑐3
𝑑𝑥2
𝜃𝐵 𝑎𝑡 𝑥2 = 2𝑚
𝑬𝑰𝜽𝑩 = −𝟐(𝟐𝟐 ) + 𝟏𝟐(𝟐) + 𝒄𝟑 = 𝟏𝟔 + 𝒄𝟑
4 + 𝑐1 = 16 + 𝑐3
𝒄𝟏 − 𝒄𝟑 − 𝟏𝟐 = 𝟎
⟹ 𝒄𝟏 = 𝒄𝟑 + 𝟏𝟐
𝒇𝒓𝒐𝒎 𝒄𝒐𝒏𝒕𝒊𝒏𝒖𝒊𝒕𝒚 𝒄𝒐𝒏𝒅𝒊𝒕𝒊𝒐𝒏 𝒂𝒕 𝑩(𝒙𝟏 = 𝒙𝟐 = 𝟐𝒎), 𝐸𝐼𝑣𝐵 = 𝐸𝐼𝑣1 = 𝐸𝐼𝑣2
1
𝐸𝐼𝑣1 = 𝑥1 3 + 𝑐1 𝑥1
3
𝑣𝐵 𝑎𝑡 𝑥2 = 2𝑚
1
𝐸𝐼𝑣𝐵 = (23 ) + 𝑐1 (2)
3
𝟖
𝑬𝑰𝒗𝑩 = + 𝟐𝒄𝟏
𝟑
2 3
𝐸𝐼𝑣2 = 𝑥2 + 6𝑥2 2 + 𝑐3 (𝑥2 − 3) − 36
3
𝑣𝐵 𝑎𝑡 𝑥2 = 2𝑚
2
𝐸𝐼𝑣𝐵 = − (23 ) + 6(22 ) + 𝑐3 (2 − 3) − 36
3
𝟓𝟐
𝑬𝑰𝒗𝑩 = − − 𝒄𝟑
𝟑
8 52
+ 2𝑐1 = − − 𝑐3
3 3
𝟔𝒄𝟏 + 𝟑𝒄𝟑 + 𝟔𝟎 = 𝟎
6(𝑐3 + 12) + 3𝑐3 + 60 = 0
44
𝑐3 = −
3
44 8
𝑐1 = 𝑐3 + 12 = − + 12 = −
3 3
44
𝑐4 = −36 − 3𝑐3 = −36 − 3 (− ) = 8
3
Thus,
𝑑𝑣1 8
𝐸𝐼 = 𝑥1 2 + 𝑐1 = 𝑥1 2 −
𝑑𝑥1 3

GIRUM MINDAYE 122


Strength of Materials AASTU

1 1 8
𝐸𝐼𝑣1 = 𝑥1 3 + 𝑐1 𝑥1 = 𝑥1 3 − 𝑥1
3 3 3
𝑑𝑣2 44
𝐸𝐼 = −2𝑥2 2 + 12𝑥2 + 𝑐3 = −2𝑥2 2 + 12𝑥2 −
𝑑𝑥2 3
2 2 44
𝐸𝐼𝑣2 = − 𝑥2 3 + 6𝑥2 2 + 𝑐3 (𝑥2 − 3) − 36 = − 𝑥2 3 + 6𝑥2 2 − 𝑥 +8
3 3 3 2
The maximum deflection occurs at the point where slope is zero.
𝑑𝑣1 8
𝐸𝐼 = 𝑥1 2 − = 0 ⟹ 𝑥1 = √8⁄3 = 1.633𝑚
𝑑𝑥1 3
𝑑𝑣2 44
𝐸𝐼 = −2𝑥2 2 + 12𝑥2 − =0
𝑑𝑥2 3
44
−12 ± √122 − (4)(−2) (− 3 )
⟹ 𝑥2 = = 1.709𝑚 𝑜𝑟 4.291𝑚,
2(−2)
𝐴𝑛𝑑, 𝑡ℎ𝑖𝑠 𝑣𝑎𝑙𝑢𝑒𝑠 𝑎𝑟𝑒 𝑜𝑢𝑡 𝑜𝑓𝑓 𝑟𝑎𝑛𝑔𝑒 𝑓𝑜𝑟 𝑥2 , 𝑤ℎ𝑖𝑐ℎ 𝑖𝑠 2 ≤ 𝑥2 ≤ 3
Therefore, deflection is maximum at x1=1.633m
1 3 8 1 8
𝐸𝐼𝑣𝑚𝑎𝑥 = 𝑥1 − 𝑥1 = (1.6333 ) − (1.633) = −2.903𝑘𝑁𝑚3
3 3 3 3
The negative sign indicates that the deflection is downwards.
Example 6.3: The beam is subjected to a load at its end as shown in figure. Determine the
displacement at C. EI is constant.

Elastic Curve. Due to the loading, two x coordinates will be considered, namely, 0≤ x1≤2m and
0≤x2≤ 1m, where x2 is directed to the left from C, since the internal moment is easy to formulate.

Moment Functions. Using the free-body diagrams


𝑀1 = −2𝑥1 𝑓𝑜𝑟 0 ≤ 𝑥1 ≤ 2𝑚
𝑀2 = −4𝑥2 𝑓𝑜𝑟 0 ≤ 𝑥2 ≤ 1𝑚
Slope and Elastic Curve.

GIRUM MINDAYE 123


Strength of Materials AASTU

a) For M1
𝑑 2 𝑣1
𝐸𝐼 = 𝑀1
𝑑𝑥1 2
𝑑 2 𝑣1
𝐸𝐼 = −2𝑥1
𝑑𝑥1 2
𝑑𝑣1
𝐸𝐼 = −𝑥1 2 + 𝑐1
𝑑𝑥1
1
𝐸𝐼𝑣1 = − 𝑥1 3 + 𝑐1 𝑥1 + 𝑐2
3
𝑓𝑟𝑜𝑚 𝑏𝑜𝑢𝑛𝑑𝑎𝑟𝑦 𝑐𝑜𝑛𝑑𝑖𝑡𝑖𝑜𝑛, 𝑣𝐴 = 0 𝑎𝑡 𝑥2 = 0
1 3
𝐸𝐼(0) = (0 ) + 𝑐1 (0) + 𝑐2 ⟹ 𝑐2 = 0
3
𝟏
∴ 𝑬𝑰𝒗𝟏 = − 𝒙𝟏 𝟑 + 𝒄𝟏 𝒙𝟏
𝟑
b) For M2
𝑑 2 𝑣2
𝐸𝐼 = 𝑀2
𝑑𝑥2 2
𝑑 2 𝑣2
𝐸𝐼 = −4𝑥2
𝑑𝑥2 2
𝑑𝑣2
𝐸𝐼 = −2𝑥2 2 + 𝑐3
𝑑𝑥2
2
𝐸𝐼𝑣 = − 𝑥 3 + 𝑐3 𝑥 + 𝑐4
3
𝑓𝑟𝑜𝑚 𝑏𝑜𝑢𝑛𝑑𝑎𝑟𝑦 𝑐𝑜𝑛𝑑𝑖𝑡𝑖𝑜𝑛, 𝑣𝐶 = 0 𝑎𝑡 𝑥 = 1
2 3 𝟐
𝐸𝐼(0) = − (1 ) + 𝑐3 (1) + 𝑐4 ⟹ 𝒄𝟒 = − 𝒄𝟑
3 𝟑
𝟐 𝟑 𝟐
∴ 𝑬𝑰𝒗𝟐 = − 𝒙𝟐 + 𝒄𝟑 (𝒙𝟐 − 𝟏) +
𝟑 𝟑
𝒇𝒓𝒐𝒎 𝒄𝒐𝒏𝒕𝒊𝒏𝒖𝒊𝒕𝒚 𝒄𝒐𝒏𝒅𝒊𝒕𝒊𝒐𝒏 𝒂𝒕 𝑩(𝒙𝟏 = 𝟐 𝒂𝒏𝒅 𝒙𝟐 = 𝟏𝒎),
𝑑𝑣1 𝑑𝑣2
𝐸𝐼𝜃𝐵 = 𝐸𝐼 = −𝐸𝐼
𝑑𝑥1 𝑑𝑥2
There is a negative sign in this equation because the slope is measured positive
counterclockwise from the right and positive clockwise from the left.
𝑑𝑣1
𝐸𝐼 = −𝑥1 2 + 𝑐1
𝑑𝑥1
𝜃𝐵 𝑎𝑡 𝑥1 = 2𝑚
𝑬𝑰𝜽𝑩 = −𝟐𝟐 + 𝒄𝟏 = −𝟒 + 𝒄𝟏
𝑑𝑣2
𝐸𝐼 = −2𝑥2 2 + 𝑐3
𝑑𝑥2
𝜃𝐵 𝑎𝑡 𝑥2 = 1𝑚
𝑬𝑰𝜽𝑩 = −𝟐(𝟏𝟐 ) + 𝒄𝟑 = −𝟐 + 𝒄𝟑
−4 + 𝑐1 = −(−2 + 𝑐3 )
𝒄𝟏 + 𝒄𝟑 − 𝟔 = 𝟎

GIRUM MINDAYE 124


Strength of Materials AASTU

⟹ 𝒄𝟑 = −𝒄𝟏 + 𝟔
𝒇𝒓𝒐𝒎 𝒄𝒐𝒏𝒕𝒊𝒏𝒖𝒊𝒕𝒚 𝒄𝒐𝒏𝒅𝒊𝒕𝒊𝒐𝒏 𝒂𝒕 𝑩(𝒙𝟏 = 𝟐𝒎 𝒂𝒏𝒅 𝒙𝟐 = 𝟏𝒎), 𝐸𝐼𝑣𝐵 = 𝐸𝐼𝑣1 = 𝐸𝐼𝑣2
1
𝐸𝐼𝑣1 = − 𝑥1 3 + 𝑐1 𝑥1
3
𝑣𝐵 𝑎𝑡 𝑥2 = 2𝑚
1
𝐸𝐼𝑣𝐵 = − (23 ) + 𝑐1 (2)
3
𝟖
𝑬𝑰𝒗𝑩 = − + 𝟐𝒄𝟏
𝟑
𝟐 𝟑 𝟐
𝐸𝐼𝑣2 = − 𝒙𝟐 + 𝒄𝟑 (𝒙𝟐 − 𝟏) +
𝟑 𝟑
𝑣𝐵 𝑎𝑡 𝑥2 = 1𝑚
2 2
𝐸𝐼𝑣𝐵 = − (13 ) + 𝑐3 (1 − 1) +
3 3
𝑬𝑰𝒗𝑩 = 𝟎
8
− + 2𝑐1 = 0
3
4
𝑐1 =
3
4 14
𝑐3 = −𝑐3 + 6 = − + 6 =
3 3
2 2 14
𝑐4 = − 𝑐3 = − = −4
3 3 3
Thus,
𝑑𝑣1 4
𝐸𝐼 = −𝑥1 2 + 𝑐1 = −𝑥1 2 +
𝑑𝑥1 3
1 1 4
𝐸𝐼𝑣1 = − 𝑥1 3 + 𝑐1 𝑥1 = − 𝑥1 3 + 𝑥1
3 3 3
𝑑𝑣2 14
𝐸𝐼 = −2𝑥2 2 + 𝑐3 = −2𝑥2 2 +
𝑑𝑥2 3
2 3 2 2 14
𝐸𝐼𝑣2 = − 𝑥2 + 𝑐3 (𝑥2 − 1) + = − 𝑥2 3 + 𝑥2 − 4
3 3 3 3
The displacement at C is determined by setting x2= 0. We get
2 14 2 14
𝐸𝐼𝑣𝑐 = − 𝑥2 3 + 𝑥2 − 4 = − (03 ) + (0) − 4 = −4𝑘𝑁𝑚3
3 3 3 3
The negative sign indicates that the deflection is downwards.

GIRUM MINDAYE 125


Strength of Materials AASTU

7 COMPRESSION MEMBER
7.1 Introduction

A vertical member subjected to axial compressive load is called column. Load carrying
capacity of a compression member depends not only on its cross sectional area, but also on its
length and the manner in which the ends of a column are held. Columns are classified into three
according to nature of failure; short, medium and long columns.
1) Short column – whose length is so related to its cross section area that failure occurs mainly
due to direct compressive stress only and the role of bending stress is negligible
2) Medium Column - whose length is so related to its cross section area that failure occurs by
a combination of direct compressive stress and bending stress
3) Long Column - whose length is so related to its cross section area that failure occurs mainly
due to bending stress and the role of direct compressive stress is negligible
7.2 Buckling and Stability

Not only must a member satisfy specific strength and deflection requirements but it must also
be stable. Stability is particularly important if the member is long and slender, and it supports
a compressive loading that becomes large enough to cause the member to suddenly deflect
laterally or sideway. These members are called columns, and the lateral deflection that occurs
is called buckling. Quite often, the buckling of a column can lead to a sudden and dramatic
failure of a structure or mechanism, and as a result, special attention must be given to the design
of columns so that they can safely support their intended loadings without buckling.
The maximum axial load that a column can support when it is on the verge of buckling is called
the critical load, Pcr, Figure 1-1a. Any additional loading will cause the column to buckle and
therefore deflect laterally as shown in Figure 1-1b.

Figure 7-1: Critical Load


We can study the nature of this instability by considering the two-bar mechanism consisting of
weightless rigid bars that are pin connected as shown in Figure 7-2a. When the bars are in the
vertical position, the spring, having a stiffness k, is unstretched, and a small vertical force P is

GIRUM MINDAYE 126


Strength of Materials AASTU

applied at the top of one of the bars. To upset this equilibrium position the pin at A is displaced
by a small amount ∆, Figure 7-2b. As shown on the free-body diagram of the pin, Figure 7-2c,
the spring will produce a restoring force F = k∆ in order to resist the two horizontal components,
Px=Ptanθ, which tend to push the pin (and the bars) further out of equilibrium. Since θ is small,
∆ ≈θ(L/2) and tanθ≈θ. Thus the restoring spring force becomes F = kθ(L/2), and the disturbing
force is 2Px= 2Pθ.

Figure 7-2: Rigid bars


If the restoring force is greater than the disturbing force, that is, kθL/2 >2Pθ, then, noticing
that θ cancels out, we can solve for P, which gives
𝑘𝐿
𝑃< 𝑠𝑡𝑎𝑏𝑙𝑒 𝑒𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚
4
This is a condition for stable equilibrium, since the force developed by the spring would be
adequate to restore the bars back to their vertical position. However, if kθL/2 < 2Pθ, or
𝑘𝐿
𝑃> 𝑢𝑛𝑠𝑡𝑎𝑏𝑙𝑒 𝑒𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚
4
then the bars will be in unstable equilibrium. In other words, if this load is applied, and a slight
displacement occurs at A, the bars will tend to move out of equilibrium and not be restored to
their original position.
The intermediate value of P, which requires kθL/2 = 2Pθ, is the critical load. Here
𝑘𝐿
𝑃𝑐𝑟 = 𝑛𝑒𝑢𝑡𝑟𝑎𝑙 𝑒𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚
4
This loading represents a case of the bars being in neutral equilibrium. Since Pcr is independent
of the (small) displacement θ of the bars, any slight disturbance given to the mechanism will
not cause it to move further out of equilibrium, nor will it be restored to its original position.
Instead, the bars will simply remain in the deflected position.
These three different states of equilibrium are represented graphically in Figure 7-3. The
transition point where the load is equal to its criticalvalue P = Pcr is called the bifurcation

GIRUM MINDAYE 127


Strength of Materials AASTU

point. Here the bars will be in neutral equilibrium for any small value of θ. If a larger load P is
placed on the bars, then they will undergo a larger deflection, so that the spring is compressed
or elongated enough to hold them in equilibrium.
In a similar manner, if the load on an actual column exceeds its critical loading, then this
loading will also require the column to undergo a large deflection; however, this is generally
not tolerated in engineering structures or machines.

Figure 7-3: Three different states of equilibrium

7.3 Euler's Theory of Column Buckling

Once a member shows signs of buckling, it will lead to the failure of the member. This load at
which the member just buckles is called the buckling load or critical load or crippling load.
The buckling load is less than the crushing load. The value of buckling load is low for long
columns and relatively high for short columns. The value of the buckling load for a given
member depends upon the length of the member and the least lateral dimension. It also depends
upon the types of end-constraints of the column (hinged, fixed etc.). Thus, when an axially
loaded compression member just buckles, it is said to develop an elastic Instability.

Column buckling is a curious and unique subject. It is perhaps the only area of structural
mechanics in which failure is not related to the strength of the material. A column buckling
analysis consists of determining the maximum load a column can support before it collapses.
But for long columns, the collapse has nothing to do with material yield. It is instead governed
by the column's stiffness, both material and geometric.

Euler's theory of column buckling is used to estimate the critical buckling load of column since
the stress in the column remains elastic. The critical buckling load is the maximum load that a
column can withstand when it is on the verge of buckling. The buckling failure occurs when
the length of the column is greater when compared with its cross-section. The Euler's theory is
based on certain assumptions related to the point of axial load application, column material,
cross-section, stress limits, and column failure. The validity of Euler’s theory is subjected to a
condition that failure occurs due to buckling.

GIRUM MINDAYE 128


Strength of Materials AASTU

The Euler’s theory states that the stress in the column due to direct loads is small compared to
the stress due to buckling failure. Based on this statement, a formula derived to compute the
critical buckling load of column. So, the equation is based on bending stress and neglects direct
stress due to direct loads on the column. The followings are the general assumptions in Euler's
theory of column buckling:

 Initially, the column is perfectly straight.


 The cross-section of the column is uniform throughout its length.
 The load is axial and passes through the centroid of the section.
 The stresses in the column are within the elastic limit.
 The materials of the column are homogenous and isotropic.
 The self-weight of the column itself is neglected.
 The failure of the column occurs due to buckling only.
 Length of column is large compared to its cross-sectional dimensions.
 The ends of the column are frictionless.
 The shortening of column due to axial compression is negligible.

This theory does not consider the effect of direct stress in column, the crookedness in column,
which is always present, and possible shifts of axial load application point from the center of
the column cross-section. As a result, the theory may overestimate the critical buckling load.
Leonhard Euler invented the Euler theory of column buckling in 1757.

The tendency of a column to remain stable or become unstable when subjected to an axial
load actually depends upon its ability to resist bending. Hence, in order to determine the
critical load and the buckled shape of the column, we will apply equation which relates the
internal moment in the column to its deflected shape, i.e.,
𝑑2𝑣 𝑀
=
𝑑𝑥 2 𝐸𝐼
Case (i) Both Ends Pinned

Figure 7-4: Column with pinned ends: (a) ideal column, (b) buckled shape, and (c) axial force
P and bending moment M acting at a cross section

GIRUM MINDAYE 129


Strength of Materials AASTU

𝑑2 𝑣
𝐸𝐼 = 𝑀 = −𝑃𝑦
𝑑𝑥 2

The equation can be written as

𝑑2𝑣 𝑃
2
+ 𝑎2 𝑦 = 0, 𝑤ℎ𝑒𝑟𝑒 𝑎2 =
𝑑𝑥 𝐸𝐼
𝑇ℎ𝑒 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 𝑖𝑠 𝑦 = 𝐴 sin 𝑎𝑥 + 𝐵 cos 𝑎𝑥
𝐴𝑡 𝑥 = 0, 𝑦 = 0, ∴𝐵=0
𝐴𝑡 𝑥 = 𝐿, 𝑦 = 0, 𝑎𝑛𝑑 𝑡ℎ𝑢𝑠, 𝐴 sin 𝑎𝐿 = 0
If A=0, y is zero for all values of load and there is no bending, this solution is not of interest.
∴ sin 𝑎𝐿 = 0 𝑜𝑟 𝑎𝐿 = 𝑛𝜋 𝑛 = 1, 2, 3, …
𝑎𝑙 = 𝜋 (𝑐𝑜𝑛𝑠𝑖𝑑𝑒𝑟𝑖𝑛𝑔 𝑡ℎ𝑒 𝑙𝑒𝑎𝑠𝑡 𝑣𝑎𝑙𝑢𝑒)
𝜋
𝑜𝑟 𝑎=
𝐿
𝟐
𝝅𝟐 𝑬𝑰
∴ 𝑪𝒓𝒊𝒕𝒊𝒄𝒂𝒍 𝒍𝒐𝒂𝒅, 𝑷𝒄𝒓 = 𝒂 𝑬𝑰 = 𝟐
𝑳
Case (ii) One end fixed other free

Figure 7-5: Ideal column fixed at the base and free at the top

𝑑2𝑣
𝐸𝐼 2 = 𝑀 = 𝑃(𝑎 − 𝑦) = 𝑃𝑎 − 𝑃𝑦
𝑑𝑥
The equation can be written as

𝑑2𝑣 2
𝑃𝑎 𝑃
+ 𝑎 𝑦 = , 𝑤ℎ𝑒𝑟𝑒 𝑎2 =
𝑑𝑥 2 𝐸𝐼 𝐸𝐼
𝑃𝑎
𝑇ℎ𝑒 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 𝑖𝑠 𝑦 = 𝐴 sin 𝑎𝑥 + 𝐵 cos 𝑎𝑥 + = 𝐴 sin 𝑎𝑥 + 𝐵 cos 𝑎𝑥 + 𝑎
𝐸𝐼𝑎2
𝐴𝑡 𝑥 = 0, 𝑦 = 0, ∴ 𝐵 = −𝑎
𝑑𝑦
𝐴𝑡 𝑥 = 0, = 0, 𝑎𝑛𝑑 𝑡ℎ𝑢𝑠, 𝐴𝑎 cos 𝑎𝑥 − 𝐵𝑎 sin 𝑎𝑥 = 0 → 𝐴 = 0
𝑑𝑥

GIRUM MINDAYE 130


Strength of Materials AASTU

𝑦 = −𝑎 cos 𝑎𝑥 + 𝑎 = 𝑎(1 − cos 𝑎𝑥)


𝐴𝑡 𝑥 = 𝐿, 𝑦 = 𝑎,
𝑛𝜋
∴ 𝑎 = 𝑎(1 − cos 𝑎𝐿) → cos 𝑎𝐿 = 0 𝑜𝑟 𝑎𝐿 = 𝑛 = 1, 3, 5, …
2
𝜋
𝑎𝑙 = (𝑐𝑜𝑛𝑠𝑖𝑑𝑒𝑟𝑖𝑛𝑔 𝑡ℎ𝑒 𝑙𝑒𝑎𝑠𝑡 𝑣𝑎𝑙𝑢𝑒)
2
𝜋
𝑜𝑟 𝑎=
2𝐿
𝝅𝟐 𝑬𝑰
∴ 𝑪𝒓𝒊𝒕𝒊𝒄𝒂𝒍 𝒍𝒐𝒂𝒅, 𝑷𝒄𝒓 = 𝒂𝟐 𝑬𝑰 =
𝟒𝑳𝟐
Case (iii) Fixed at both ends

Figure 7-6: Buckling of a column with both ends fixed against rotation

𝑑2 𝑣
𝐸𝐼 = 𝑀 − 𝑃𝑦
𝑑𝑥 2
The equation can be written as

𝑑2 𝑣 𝑀 𝑃
2
+ 𝑎2 𝑦 = , 𝑤ℎ𝑒𝑟𝑒 𝑎2 =
𝑑𝑥 𝐸𝐼 𝐸𝐼
𝑀 𝑀
𝑇ℎ𝑒 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 𝑖𝑠 𝑦 = 𝐴 sin 𝑎𝑥 + 𝐵 cos 𝑎𝑥 + 2
= 𝐴 sin 𝑎𝑥 + 𝐵 cos 𝑎𝑥 +
𝐸𝐼𝑎 𝑃
𝑀
𝐴𝑡 𝑥 = 0, 𝑦 = 0, ∴𝐵=−
𝑃
𝑑𝑦
𝐴𝑡 𝑥 = 0, = 0, 𝑎𝑛𝑑 𝑡ℎ𝑢𝑠, 𝐴𝑎 cos 𝑎𝑥 − 𝐵𝑎 sin 𝑎𝑥 = 0 → 𝐴 = 0
𝑑𝑥
𝑀 𝑀 𝑀
𝑦=− cos 𝑎𝑥 + = (1 − cos 𝑎𝑥)
𝑃 𝑃 𝑃
𝐴𝑡 𝑥 = 𝐿, 𝑦 = 0,
𝑀
∴0= (1 − cos 𝑎𝐿) → cos 𝑎𝐿 = 1 𝑜𝑟 𝑎𝐿 = 𝑛𝜋 𝑛 = 0, 2, 4, 6, …
𝑃
𝑎𝑙 = 2𝜋 (𝑐𝑜𝑛𝑠𝑖𝑑𝑒𝑟𝑖𝑛𝑔 𝑡ℎ𝑒 𝑙𝑒𝑎𝑠𝑡 𝑝𝑟𝑎𝑐𝑡𝑖𝑐𝑎𝑙 𝑣𝑎𝑙𝑢𝑒)

GIRUM MINDAYE 131


Strength of Materials AASTU

2𝜋
𝑜𝑟 𝑎=
𝑙
𝟐
𝟒𝝅𝟐 𝑬𝑰
∴ 𝑪𝒓𝒊𝒕𝒊𝒄𝒂𝒍 𝒍𝒐𝒂𝒅, 𝑷𝒄𝒓 = 𝒂 𝑬𝑰 =
𝒍𝟐
Case (iv) One end fixed, other Pinned

Figure 7-7: Column fixed at the base and pinned at the top

𝑑2𝑣
𝐸𝐼 2 = −𝑃𝑦 + 𝑅(𝐿 − 𝑥)
𝑑𝑥
The equation can be written as

𝑑2𝑣 2
𝑅(𝐿 − 𝑥) 2
𝑃
+ 𝑎 𝑦 = , 𝑤ℎ𝑒𝑟𝑒 𝑎 =
𝑑𝑥 2 𝐸𝐼 𝐸𝐼
𝑅(𝐿 − 𝑥) 𝑅
𝑇ℎ𝑒 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 𝑖𝑠 𝑦 = 𝐴 sin 𝑎𝑥 + 𝐵 cos 𝑎𝑥 + = 𝐴 sin 𝑎𝑥 + 𝐵 cos 𝑎𝑥 + (𝐿 − 𝑥)
𝐸𝐼𝑎2 𝑃
𝑅𝐿
𝐴𝑡 𝑥 = 0, 𝑦 = 0, ∴𝐵=−
𝑃
𝑑𝑦 𝑅 𝑅
𝐴𝑡 𝑥 = 0, = 0, 𝑎𝑛𝑑 𝑡ℎ𝑢𝑠, 𝐴𝑎 cos 𝑎𝑥 − 𝐵𝑎 sin 𝑎𝑥 − = 0 → 𝐴 =
𝑑𝑥 𝑃 𝑃𝑎
𝑅 𝑅𝐿 𝑅
𝑦= sin 𝑎𝑥 − cos 𝑎𝑥 + (𝐿 − 𝑥)
𝑃𝑎 𝑃 𝑃
𝐴𝑡 𝑥 = 𝐿, 𝑦 = 0,
𝑅 𝑅𝐿
0= sin 𝑎𝑙 − cos 𝑎𝑙
𝑃𝑎 𝑃

tan 𝑎𝐿 = 𝑎𝐿
The smallest nonzero value of aL that satisfies is

GIRUM MINDAYE 132


Strength of Materials AASTU

𝑎𝐿 = 4.49𝑟𝑎𝑑 (𝑐𝑜𝑛𝑠𝑖𝑑𝑒𝑟𝑖𝑛𝑔 𝑡ℎ𝑒 𝑙𝑒𝑎𝑠𝑡 𝑣𝑎𝑙𝑢𝑒)


4.49
𝑜𝑟 𝑎=
𝐿
𝟐
𝟒. 𝟒𝟗𝟐 𝑬𝑰 𝟐𝟎. 𝟐𝑬𝑰 𝟐𝝅𝟐 𝑬𝑰
∴ 𝑪𝒓𝒊𝒕𝒊𝒄𝒂𝒍 𝒍𝒐𝒂𝒅, 𝑷𝒄𝒓 = 𝒂 𝑬𝑰 = = ≈
𝑳𝟐 𝑳𝟐 𝑳𝟐
7.4 Effective Length and Slenderness Ratio

The critical loads for columns with various support conditions can be related to the critical load
of a column through the concept of an effective length. Effective length of column (Le) is
defined as vertical height between the two points of contraflexure of the buckled column or it
can be also defined as vertical distance between to deflection caused due to buckling of column.
Another way of expressing this idea is to say that the effective length of a column is the distance
between points of inflection (that is, points of zero moment) in its deflection curve, assuming
that the curve is extended (if necessary) until points of inflection are reached. Thus, for a fixed-
free column (Figure 7-8), the effective length is

Figure 7-8: Deflection curves showing the effective length Le for a column fixed at the base
and free at the top
The general critical load for various boundary conditions can be written as follows:
𝒏𝝅𝟐 𝑬𝑰 𝝅𝟐 𝑬𝑰
𝑷𝒄𝒓 = =
𝑳𝟐 (𝒌𝑳)𝟐
Where,
k is effective length factor (𝑘 = √1⁄𝑛),
n is factor accounting for the end conditions and
L is the actual length of the column.
The effective length is often expressed in terms of an effective length factor, k:
𝐿𝑒 = 𝑘𝐿
Boundary Pinned-pinned Fixed-free Fixed-fixed Fixed-pinned
conditions column column column column
n=factor Counting 1 0.25 4 2
for End Conditions
𝑘 = √1⁄𝑛 1 2 0.5 0.7

GIRUM MINDAYE 133


Strength of Materials AASTU

Thus, the critical load is


𝝅𝟐 𝑬𝑰
𝑷𝒄𝒓 =
𝑳𝒆 𝟐
This load is termed as Euler's load and is denoted by PE and the equation is known as Euler's
formula.
It can be seen that the column will have a tendency to bend or buckle in that plane about which
flexural rigidity El is least. Therefore in the above equation minimum moment of inertia should
be used. It can be seen that critical load is proportional to flexural rigidity and inversely
proportional to length does not depend upon permissible and stress of material from which the
column is made.

Figure 7-9: Effective length of column for various end conditions.


Critical stress (σcr) is average stress over the cross section
𝑃𝑐𝑟 𝜋 2 𝐸𝐼
𝜎𝑐𝑟 = =
𝐴 𝐴𝐿𝑒 2
The moment of inertia, I, refers to the axis about which bending occurs. Putting I = Ar2,
where, r is the radius of gyration about the axis of bending,
𝐼
𝑅𝑎𝑑𝑖𝑢𝑠 𝑜𝑓 𝑔𝑦𝑟𝑎𝑡𝑖𝑜𝑛, 𝑟 = √ ⟹ 𝐼 = 𝐴𝑟 2
𝐴
𝑃𝑐𝑟 𝜋 2 𝐸𝐼 𝜋 2 𝐸𝐴𝑟 2 𝜋2𝐸 𝜋2𝐸
𝜎𝑐𝑟 = = = = = 2
𝐴 𝐴𝐿𝑒 2 𝐴𝐿𝑒 2 (𝐿𝑒 ⁄𝑟)2 𝜆
𝑠𝑙𝑒𝑛𝑑𝑒𝑟𝑛𝑒𝑠𝑠 𝑟𝑎𝑡𝑖𝑜, 𝜆 = 𝐿𝑒 ⁄𝑟

GIRUM MINDAYE 134


Strength of Materials AASTU

For a pivot-ended concentrically loaded column with no intermediate bracing to restrain lateral
motion, bending occurs about the axis of minimum moment of inertia. Therefore, r, radius of
gyration is taken as minimum.
The Euler buckling load as given by Euler’s formula agrees well with experiment only if the
slenderness ratio is large, whereas short compression members can be analyzed easily
considering direct stress σ = P/A. Many columns lie between these extremes in which neither
of these solutions is applicable. These intermediate length columns are analyzed by secant
formulae. To check the validity of Euler's formula consider that if the slenderness ratio is small,
the stress at the failure σcr will be large. Let σc be the crushing strength of column material. If
σcr > σc the failure of column will be due to crushing and not due to buckling. Hence, the Euler's
formula will not be applicable for smaller slenderness ratio. In other word, Euler’s formula is
only applicable for long columns.
For validity of Euler's formula,
𝜎𝑐𝑟 ≤ 𝜎𝑐
𝜋2𝐸 𝑬
2
≤ 𝜎𝑐 ⟹ 𝝀 ≥ 𝝅√
𝜆 𝝈𝒄
Numerous column experiments indicate that the Euler formula is reliable for designing axially
loaded columns, provided the slenderness ratio is within the range in which the eccentricity has
relatively little effect. This range is called the slender range, which ranges from 120 to 140. At
lower slenderness, the failure stress would be the compressive strength of the material. The
extent of this range is 0 - 40. In the intermediate range (i.e. slenderness ratio 40 to 120), the
only equation of rational nature that applies to the real columns is the secant formula. Since the
application of the secant formulae to centric loading requires an estimate of accidental
eccentricity ratio, the equation acquires an empirical nature. Furthermore, the application is so
involved that simpler empirical formulae have been developed which give results in reasonable
agreement with the experimental results within the intermediate range.
A column, whether short or long, is determined by the numerical values of slenderness ratios.
Smaller the slenderness ratio, lesser will be the tendency to deflect and higher will be the
buckling load.
7.5 Rankine’s formula for Column

Rankine proposed an empirical formula for columns, which cover all cases ranging from short
to long columns. He proposed the relation
1 1 1
= +
𝑃 𝑃𝑐 𝑃𝑒
Where,
𝑃𝑐 = 𝑢𝑙𝑡𝑖𝑚𝑎𝑡𝑒 𝑙𝑜𝑎𝑑 𝑓𝑜𝑟 𝑎 𝑠ℎ𝑜𝑟𝑡 𝑐𝑜𝑙𝑢𝑚𝑛,
𝜋 2 𝐸𝐼
𝑃𝑒 = = Euler′s critical load
𝐿𝑒 2
1
= 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 𝑓𝑜𝑟 𝑎 𝑚𝑎𝑡𝑒𝑟𝑖𝑎𝑙.
𝑃𝑐

GIRUM MINDAYE 135


Strength of Materials AASTU

For short columns, Pe is very large and hence 1/Pe is small in comparison to 1/Pc, thus making
the critical load P approximately equal to PC. For long columns, Pe is extremely small and
hence 1/Pe is large as compared to 1/Pc, thus making the critical load P approximately equal to
Pe. Thus, the value of P obtained from the above relation covers all cases ranging from short
to long columns.
1 1 1 𝑃𝑐 𝑃𝑒 𝑃𝑐 𝜎𝑐 𝐴 𝜎𝑐 𝐴
= + ⟹𝑃= = = =
𝑃 𝑃𝑐 𝑃𝑒 𝑃𝑐 +𝑃𝑒 1 + (𝑃𝑐 ) 1 + (𝜎𝑐 𝐴) 1 + ( 𝜎𝑐 )
𝑃𝑒 𝑃𝑒 𝑃𝑒 ⁄𝐴

𝑃𝑐𝑟 𝜋2𝐸 𝜋2𝐸


𝜎𝑐𝑟 = = = 2
𝐴 (𝐿𝑒 ⁄𝑟)2 𝜆

𝜎𝑐 𝐴 𝜎𝑐 𝐴 𝜎𝑐 𝐴 𝜎𝑐 𝐴
𝑃= = 2 = 2 =
𝜎𝑐 1 + 𝑎𝜆2
1+( 2 2 ) 1 + ( 𝜎2𝑐 ) (𝐿𝑒 ) 1 + 𝑎 (
𝐿𝑒
)
𝜋 𝐸 ⁄(𝐿𝑒 ⁄𝑟) 𝜋 𝐸 𝑟 𝑟

Where,
σc= crushing stress for the material, and
a = Rankine's constant for the material which is determined experimentally, and should not be
calculated values of σc, and E.
Material 𝝈𝒄 (𝑴𝑷𝒂) 𝒂 − 𝑹𝒂𝒏𝒌𝒊𝒏𝒆′𝒔 𝒄𝒐𝒏𝒔𝒕𝒂𝒏𝒕
Wrought iron 255 1/9000
Cast iron 550 1/1600
Mild steel 330 1/7500
Strong timber 50 1/750

It can be rearranged in terms of an average axial stress and is given as


𝑃 𝜎𝑐 𝜎𝑐
= 2 =
𝐴 𝐿 1 + 𝑎𝜆2
1 + 𝑎 ( 𝑟𝑒 )
𝜎𝑐
𝑤ℎ𝑒𝑟𝑒, 𝜎𝑐 = 𝑎𝑙𝑙𝑜𝑤𝑎𝑏𝑙𝑒 𝑠𝑡𝑟𝑒𝑠𝑠, 𝑎𝑛𝑑 𝑎 =
𝜋2𝐸

GIRUM MINDAYE 136


Strength of Materials AASTU

7.6 Worked Examples

Example 7.1: A steel column has a length of 9 m and is fixed at both ends. If the cross-sectional
area has the dimensions shown, determine the critical load. Est = 200GPa, σy = 250 MPa.

Solution
Section Properties:
𝐴 = 0.2(0.17) − 0.19(0.15) = 5.50 × 10−3 𝑚2
1 1
𝐼𝑥 = (0.2)(0.173 ) − (0.19)(0.153 ) = 28.44583 × 10−6 𝑚4
12 12
1 1
𝐼𝑦 = 2 [ (0.01)(0.23 )] + (0.15)(0.013 ) = 13.34583 × 10−6 𝑚4 (𝑔𝑜𝑣𝑒𝑟𝑛𝑠)
12 12
Effective length:
𝐹𝑜𝑟 𝑓𝑖𝑥𝑒𝑑 𝑠𝑢𝑝𝑝𝑜𝑟𝑡 𝑒𝑛𝑑𝑠 𝑐𝑜𝑙𝑢𝑚𝑛, 𝐿𝑒 = 0.5𝐿 = 0.5 × 9 = 4.5𝑚
Critical load:
𝜋 2 𝐸𝐼 𝜋 2 (200 × 109 )(13.34583 × 10−6 )
𝑃𝑐𝑟 = = = 1300.919 × 103 𝑁 = 130.919𝑘𝑁
𝐿𝑒 2 4.52
Critical Stress: Euler’s formula is only valid if σcr≤ σy.
𝑃𝑐𝑟 1300.919 × 103
𝜎𝑐𝑟 = = = 236.53 × 106 𝑁⁄𝑚2 = 236.53𝑀𝑃𝑎 ≤ 𝜎𝑦
𝐴 5.50 × 10−3
= 250𝑀𝑃𝑎 − −𝑜𝑘!

Example 7.2: A steel column, which is pinned at its top and bottom, has a length of 9 m. If the
cross-sectional area has the dimensions shown, determine the critical load. Est = 200GPa, σy =
250 MPa.

Solution
Section Properties:
𝐴 = 0.2(0.17) − 0.19(0.15) = 5.50 × 10−3 𝑚2
1 1
𝐼𝑥 = (0.2)(0.173 ) − (0.19)(0.153 ) = 28.44583 × 10−6 𝑚4
12 12
1 1
𝐼𝑦 = 2 [ (0.01)(0.23 )] + (0.15)(0.013 ) = 13.34583 × 10−6 𝑚4 (𝑔𝑜𝑣𝑒𝑟𝑛𝑠)
12 12
Effective length:

GIRUM MINDAYE 137


Strength of Materials AASTU

𝐹𝑜𝑟 𝑝𝑖𝑛 𝑠𝑢𝑝𝑝𝑜𝑟𝑡𝑒𝑑 𝑒𝑛𝑑𝑠 𝑐𝑜𝑙𝑢𝑚𝑛, 𝐿𝑒 = 𝐿 = 9𝑚


Critical load:
𝜋 2 𝐸𝐼 𝜋 2 (200 × 109 )(13.34583 × 10−6 )
𝑃𝑐𝑟 = = = 325.22987 × 103 𝑁 ≈ 325.23𝑘𝑁
𝐿𝑒 2 92
Critical Stress: Euler’s formula is only valid if σcr≤ σy.
𝑃𝑐𝑟 325.22987 × 103
𝜎𝑐𝑟 = = = 59.31 × 106 𝑁⁄𝑚2 = 59.31𝑀𝑃𝑎 ≤ 𝜎𝑦
𝐴 5.50 × 10−3
= 250𝑀𝑃𝑎 − −𝑜𝑘!

Example 7.3: A steel bar of rectangular cross section 30 x 50 mm pinned at each end is 2 m
long. Determine the buckling load when it is subjected to axial compression and calculate axial
stress using Euler's expression. Determine the minimum length for which Euler's equation may
be valid. Take σy=250MPa and Est = 200GPa
Solution
Section Properties:
𝐴 = 0.03 × 0.05 = 1.5 × 10−3 𝑚2
0.05 × 0.0312
𝐼𝑚𝑖𝑛 = = 112.5 × 10−9 𝑚4
12
𝐼𝑚𝑖𝑛 112.5 × 10−9
𝑟𝑚𝑖𝑛 =√ = √ = 8.66 × 10−3 𝑚
𝐴 1.5 × 10−3
Effective Length:
For both end pinned column, Le=L=2m
Critical Load:
𝜋 2 𝐸𝐼 𝜋 2 (200 × 109 )(112.5 × 10−9 )
𝑃𝑐𝑟 = = = 55.516 × 103 𝑁 ≈ 55.52𝑘𝑁
𝐿𝑒 2 22
Critical Stress: Euler’s formula is only valid if σcr≤ σy.
𝑃𝑐𝑟 55.516 × 103
𝜎𝑐𝑟 = = = 37.01 × 106 𝑁⁄𝑚2
𝐴 1.5 × 10−3
𝜎𝑐𝑟 = 37.01𝑀𝑃𝑎 ≤ 𝜎𝑐 = 250𝑀𝑃𝑎 − − − 𝑜𝑘!
Euler’s equation is valid.
Minimum length: for which Euler's equation is valid.
𝑳𝒆 𝑬 𝑬 −3 )√
200 × 109
𝝀= ≥ 𝝅√ ⟹ 𝑳𝒆 ≥ 𝝅𝒓√ = 𝝅(8.66 × 10 = 0.7695𝑚 ≈ 0.77𝑚
𝒓 𝝈𝒄 𝝈𝒄 250 × 106
𝐿 = 0.77𝑚 𝑤ℎ𝑖𝑐ℎ 𝑖𝑠 𝑚𝑖𝑛𝑖𝑚𝑢𝑚 𝑙𝑒𝑛𝑔𝑡ℎ
Example 7.4: A 2m long pin-ended column with a square cross section is to be made of wood.
Assuming E = 13GPa, σall = 12MPa, and using a factor of safety of 2.5 to calculate Euler’s
critical load for buckling, determine the size of the cross section if the column is to safely
support (a) a 100kN load, (b) a 200kN load.

GIRUM MINDAYE 138


Strength of Materials AASTU

Figure 7-10: Pin-ended wood column of square cross section.


Solution
𝐹𝑜𝑟 𝑝𝑖𝑛 𝑒𝑛𝑑𝑒𝑑 𝑐𝑜𝑙𝑢𝑚𝑛, 𝐿𝑒 = 𝐿 = 2𝑚
(a) For the 100kN Load.
𝐶𝑟𝑖𝑡𝑖𝑐𝑎𝑙 𝑙𝑜𝑎𝑑: 𝑃𝑐𝑟 = 2.5 × 100 = 250𝑘𝑁 𝑎𝑛𝑑 (𝑏)𝑃𝑐𝑟 = 2.5 × 200 = 500𝑘𝑁
𝜋 2 𝐸𝐼 𝑃𝑐𝑟 𝐿𝑒 2
𝑃𝑐𝑟 = ⟹𝐼= 2
𝐿𝑒 2 𝜋 𝐸
2
𝑃𝑐𝑟 𝐿𝑒 (250 × 103 )(22 )
𝐼= = = 7.794 × 10−6 𝑚4
𝐸 𝜋 2 (13 × 109 )
𝑎4
𝑅𝑒𝑐𝑎𝑙𝑙𝑖𝑛𝑔 𝑡ℎ𝑎𝑡, 𝑓𝑜𝑟 𝑎 𝑠𝑞𝑢𝑎𝑟𝑒 𝑜𝑓 𝑠𝑖𝑑𝑒 𝒂, 𝐼 =
12
𝑎4
= 7.794 × 10−6 𝑚4 ⟹ 𝑎 = 0.0983𝑚 = 98.3𝑚𝑚 ≈ 𝟏𝟎𝟎𝒎𝒎
12
Check the value of the normal stress in the column:
𝑃 100 × 103
𝜎= = = 10 × 106 𝑁⁄𝑚2 = 10𝑀𝑃𝑎 ≤ 𝜎𝑐 = 12𝑀𝑃𝑎 − − − 𝑜𝑘!
𝐴 (0.1)2
(b) For the 200kN Load.
𝐶𝑟𝑖𝑡𝑖𝑐𝑎𝑙 𝑙𝑜𝑎𝑑: 𝑃𝑐𝑟 = 2.5 × 200 = 500𝑘𝑁
𝑃𝑐𝑟 𝐿𝑒 2 (500 × 103 )(22 )
𝐼= = 2 9
= 15.588 × 10−6 𝑚4
𝐸 𝜋 (13 × 10 )
𝑎4
𝑅𝑒𝑐𝑎𝑙𝑙𝑖𝑛𝑔 𝑡ℎ𝑎𝑡, 𝑓𝑜𝑟 𝑎 𝑠𝑞𝑢𝑎𝑟𝑒 𝑜𝑓 𝑠𝑖𝑑𝑒 𝒂, 𝐼 =
12
𝑎4
= 15.588 × 10−6 𝑚4 ⟹ 𝑎 = 0.11695𝑚 = 116.95𝑚𝑚 ≈ 𝟏𝟎𝟎𝒎𝒎
12
The value of the normal stress is:
𝑃 200 × 103
𝜎= = = 14.62 × 106 𝑁⁄𝑚2 = 14.62𝑀𝑃𝑎 ≰ 𝜎𝑐 = 12𝑀𝑃𝑎 − − − 𝐼𝑛𝑣𝑎𝑙𝑖𝑑
𝐴 (0.11695)2
The dimension obtained is not acceptable, and the cross section must be selected on the basis
of its resistance to compression.
𝑃 200 × 103
𝐴= = 6
= 16.67 × 10−3 𝑚2
𝜎𝑐 12 × 10
𝑎2 = 16.67 × 10−3 𝑚2 ⟹ 𝑎 = 0.1291𝑚 = 129.1𝑚𝑚 ≈ 𝟏𝟑𝟎𝒎𝒎
A 130 × 130-mm cross section is acceptable.

GIRUM MINDAYE 139


Strength of Materials AASTU

Example 7.5: The A992 steel W200 x 46 member shown in figure is to be used as a pin-
connected column. Determine the largest axial load it can support before it begins to buckle or
yields. Use E=200GPa for steel.

Solution
For pin ended column, Le=L=3m
By inspection, buckling will occur about the y–y axis.
𝜋 2 𝐸𝐼𝜋 2 (200 × 109 )(15.3 × 10−6 )
𝑃𝑐𝑟 = = = 3.3557 × 106 𝑁 = 3,355.7𝑘𝑁
𝐿𝑒 2 32
When fully loaded, the average compressive stress in the column is
𝑃𝑐𝑟 3,3557 × 106
𝜎𝑐𝑟 = = = 569.72 × 106 𝑁⁄𝑚2 ≈ 570𝑀𝑃𝑎 ≰ 𝜎𝑐 = 345𝑀𝑃𝑎
𝐴 5890 × 10−6
Hence, the load P is determined from simple compression:
𝑃
𝜎𝑐 = ⟹ 𝑃 = 𝜎𝑐 𝐴 = (345 × 106 )(5890 × 10−6 ) = 2,032.05 × 103 𝑁 = 2,032.05𝑘𝑁
𝐴
In actual practice, a factor of safety would be placed on this loading.

Example 7.6: A W150 x 24 steel column is 8 m long and is fixed at its ends as shown in figure.
Its load-carrying capacity is increased by bracing it about the y–y (weak) axis using struts that
are assumed to be pin connected to its mid height. Determine the load it can support so that the
column does not buckle nor the material exceed the yield stress. Take Est = 200GPa and σy=
410MPa.

GIRUM MINDAYE 140


Strength of Materials AASTU

The buckling behavior of the column will be different about the x–x and y–y axes due to the
bracing. The buckled shape for each of these cases is shown in figure.

From figure, the effective length for buckling about the x–x axis is (Le)x = 0.5(8 m) = 4 m,
and for buckling about the y–y axis, (Le)y = 0.7(8 m/2) = 2.8 m.
The moments of inertia for a W150 x 24 are found from section table. We have Ix = 13.4(106)
mm4, Iy = 1.83(106) mm4.
𝜋 2 𝐸𝐼𝑥 𝜋 2 (200 × 109 )(13.4 × 10−6 )
(𝑃𝑐𝑟 )𝑥 = = = 1,653.16 × 103 𝑁 = 1,653.16𝑘𝑁
(𝐿𝑒 )𝑥 2 42

𝜋 2 𝐸𝐼𝑦 𝜋 2 (200 × 109 )(1.83 × 10−6 )


(𝑃𝑐𝑟 )𝑥 = = = 460.75 × 103 𝑁 = 460.75𝑘𝑁
(𝐿𝑒 )𝑦 2 2.82

By comparison, buckling will occur about the y–y axis.


The area of the cross section is 3060mm2, so the average compressive stress in the column is
𝑃𝑐𝑟 460.75 × 103
𝜎𝑐𝑟 = = = 150.57 × 106 𝑁⁄𝑚2 = 150.57𝑀𝑃𝑎 ≤ 𝜎𝑐
𝐴 3060 × 10−6
= 410𝑀𝑃𝑎 − − − 𝑜𝑘!
Since this stress is less than the yield stress, buckling will occur before the material yields.
Thus,
𝑷𝒄𝒓 = 𝟒𝟔𝟎. 𝟕𝟓𝒌𝑵
Example 7.7: The aluminum column is braced at its top by cables so as to prevent movement
at the top along the x axis. If it is assumed to be fixed at its base, determine the largest allowable
load P that can be applied. Use a factor of safety for buckling of F.S. = 3.0. Take Eal=70GPa,
σy= 215MPa, A = 7.5(10-3)m2, Ix = 61.3(10-6)m4, Iy= 23.2(10-6)m4.

GIRUM MINDAYE 141


Strength of Materials AASTU

Solution
Buckling about the x and y axes is shown in figure.
For x–x axis buckling, (Le)x= 2(5 m) = 10 m
and for y–y axis buckling, (Le)y= 0.7(5 m) =
3.5 m.

𝜋 2 𝐸𝐼𝑥𝜋 2 (70 × 109 )(61.3 × 10−6 )


(𝑃𝑐𝑟 )𝑥 = = = 424 × 103 𝑁 = 424𝑘𝑁
(𝐿𝑒 )𝑥 2 102
𝜋 2 𝐸𝐼𝑦 𝜋 2 (70 × 109 )(23.2 × 10−6 )
(𝑃𝑐𝑟 )𝑥 = = = 1.31 × 106 𝑁 = 1,310𝑘𝑁
(𝐿𝑒 )𝑦 2 3.52
By comparison, as P is increased the column will buckle about the x–x axis.
𝑃𝑐𝑟 424 × 103
𝜎𝑐𝑟 = = = 56.5 × 106 𝑁⁄𝑚2
𝐴 7.5 × 10−3
𝜎𝑐𝑟 = 56.7𝑀𝑃𝑎 ≤ 𝜎𝑐 = 215𝑀𝑃𝑎 − − − 𝑜𝑘!
Euler’s equation is valid.
The allowable load is therefore.
𝑷𝒄𝒓 𝟒𝟐𝟒
𝑷𝒂𝒍𝒍 = = = 𝟏𝟒𝟏𝒌𝑵
𝑭. 𝑺. 𝟑. 𝟎

GIRUM MINDAYE 142


Strength of Materials AASTU

Example 7.8: Calculate safe compressive load on a hollow cast iron column with one end
hinged and other rigidly fixed. The external and internal diameters are120mm and 90mm
respectively and length of the column is 9m. Take factor of safety as 3, σc=550MPa and E =
95GPa. Also calculate critical axial stress.
Solution
Section Properties:
𝜋(𝑑2 0 − 𝑑2 𝑖 ) 𝜋[(0.12)2 − (0.09)2 ]
𝐴= = = 4,948 × 10−6 𝑚2
4 4
𝜋(𝑑 4 0 − 𝑑4 𝑖 ) 𝜋[(0.15)4 − (0.13)4 ]
𝐼= = = 6.958 × 10−6 𝑚4
64 64
𝐼 6.958 × 10−6
𝑟= √ = √ = 37.50 × 10−3 𝑚
𝐴 4,948 × 10−6
Effective length:
𝐹𝑜𝑟 𝑜𝑛𝑒 𝑒𝑛𝑑 ℎ𝑖𝑛𝑔𝑒𝑑 𝑎𝑛𝑑 𝑜𝑡ℎ𝑒𝑟 𝑟𝑖𝑔𝑖𝑑𝑙𝑦 𝑓𝑖𝑥𝑒𝑑, 𝐿𝑒 = 0.7𝐿 = 0.7 × 9 = 6.3𝑚
Critical load:
𝜋 2 𝐸𝐼 𝜋 2 (95 × 109 )(6.958 × 10−6 )
𝑃𝑐𝑟 = = = 164.371 × 103 𝑁 ≈ 164.37𝑘𝑁
𝐿𝑒 2 6.32
Critical Stress: Euler’s formula is only valid if σcr≤ σc.
𝑃𝑐𝑟 164.371 × 103
𝜎𝑐𝑟 = = = 33.22 × 106 𝑁⁄𝑚2 = 33.22𝑀𝑃𝑎 ≤ 𝜎𝑐 = 550𝑀𝑃𝑎 − −𝑜𝑘!
𝐴 4,948 × 10−6
Allowable load:
𝑃𝑐𝑟 164.37
𝑃𝑎𝑙𝑙 = = = 54.79𝑘𝑁
𝐹. 𝑆. 3

Example 7.9: Using Rankine's fonnula find the crippling load for a mild steel strut of 500mm
long with a rectangular cross-section 50mm x 12.5mm having (a) hinged ends, aid (b) both
ends fixed. Take σc = 330MPa and a for hinged ends = 1/7500.
Solution
Section Properties:
𝐴 = 0.05(0.0125) = 625 × 10−6 𝑚2
1 1
𝐼𝑚𝑖𝑛 = (0.05)(0.01253 ) − (0.19)(0.153 ) = 8.14 × 10−9 𝑚4
12 12
𝐼𝑚𝑖𝑛 8.14 × 10−9
𝑟𝑚𝑖𝑛 = √ =√ = 11.41 × 10−3 𝑚
𝐴 625 × 10−6
(a) For hinged ends
𝐿𝑒 = 𝐿 = 0.5𝑚
𝜎𝑐 𝐴 (330 × 106 )(625 × 10−6 )
𝑃= = = 164.21 × 103 𝑁 = 164.21𝑘𝑁
𝐿 2 1 0.5 2
1 + 𝑎 ( 𝑟𝑒 ) 1+( )( )
7500 11.41 × 10−3

GIRUM MINDAYE 143


Strength of Materials AASTU

(b) For fixed ends


𝐿𝑒 = 0.6𝐿 = 0.5 × 0.5 = 0.25𝑚
𝜎𝑐 𝐴 (330 × 106 )(625 × 10−6 )
𝑃= = = 193.84 × 103 𝑁 = 193.84𝑘𝑁
𝐿 2 1 0.25 2
1 + 𝑎 ( 𝑟𝑒 ) 1+( )( )
7500 11.41 × 10−3

Example 7.10: A hollow cylindrical cast iron colulm 150mm external diameter and 20mm
thick is 6 meter in length having both ends hinged. Find the load using Rankine's formula. Take
σc = 550MPa and a = 1/1600.
Solution
Section Properties:
𝑑𝑜 = 150𝑚𝑚 𝑎𝑛𝑑 𝑑𝑖 = 𝑑𝑜 − 2𝑡 = 150 − 2(20) = 110𝑚𝑚
𝜋(𝑑2 0 − 𝑑2 𝑖 ) 𝜋[(0.15)2 − (0.11)2 ]
𝐴= = = 8,168.14 × 10−6 𝑚2
4 4
𝜋(𝑑 4 0 − 𝑑4 𝑖 ) 𝜋[(0.15)4 − (0.13)4 ]
𝐼= = = 17.663 × 10−6 𝑚4
64 64
𝐼 17.663 × 10−6
𝑟= √ = √ −6
= 46.502 × 10−3 𝑚
𝐴 8,168.14 × 10
𝐹𝑜𝑟 𝑏𝑜𝑡ℎ 𝑒𝑛𝑑𝑠 ℎ𝑖𝑛𝑔𝑒𝑑, 𝐿𝑒 = 𝐿 = 6𝑚
𝜎𝑐 𝐴 (550 × 106 )(8,168.14 × 10−6 )
𝑃= 2 = 2 = 393.907 × 103 𝑁 ≈ 393.91𝑘𝑁
𝐿 1 6
1 + 𝑎 ( 𝑟𝑒 ) 1 + (1600) ( )
46.502 × 10−3

GIRUM MINDAYE 144


Strength of Materials AASTU

Reference
1. Ferdinand Beer and E. Johnston and John DeWolf and David Mazurek, “Mechanics of
Materials, 8th Edition, McGraw-Hill, 2020”.
2. Barry J. Goodno and James M. Gere, “Mechanics of Materials 9th Edition, Cengage
Learning, 2016”.
3. Russell C. Hibbeler, “Mechanics of Materials, 10th Edition, Pearson, 2018”.
4. R.K. Rajput, ‘Strength of Materials’, 4th Edition, S. Chand Limited, 2007”.
5. R.K. Bansal, “A Text book on Strength of Materials, Lakxmi Publications, 2009”.
6. R. Subramanian, “Strength of Materials, 2nd Edition, Oxford University Press, 2010.

GIRUM MINDAYE 145

You might also like