You are on page 1of 32

DMV 4343

JAN ~ JUN ‘07

INFORMATION SHEET

DEPARTMENT MANUFACTURING / PRODUCT DESIGN / SEMESTER 4/6


MOULD / TOOL AND DIE
COURSE MECHANICS OF MATERIALS DURATION 8 hrs
COURSE CODE DMV 4343 / DMV 5343 REF. NO.
VTO’S NAME MISS AFZAN BINTI ROZALI PAGE 29

TOPIC
INTRODUCTION TO STRESS AND STRAIN

SUB TOPIC
1.1 Normal Stress and Strain
1.2 Stress – Strain Relationship and Diagram
1.3 Elasticity and Plasticity
1.4 Linear Elasticity and Hooke’s Law
1.5 Allowable Stress and Allowable Load
1.6 Deformation of Axially Loaded Member

REF NO. :
PAGE :
29

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p1


DMV 4343
JAN ~ JUN `07

INTRODUCTION TO STRESS AND STRAIN

Figure 1.0a shows an L-shaped bracket loaded as a two-force member, and Figure 1.0b
shows the internal resultants, F(x), V(x), and M(x), that are required to maintain the
equilibrium of the two sectioned parts of the bracket. Although we could compute the
internal resultants shown on Figure 1.0b by using static equilibrium procedures (free-body
diagrams and equations of equilibrium), those procedures are clearly insufficient for
determining the complex internal force distribution making up those resultants. The concept
of stress is introduced in this chapter to enable us to quantify internal force distributions.
The shape of the bracket also changes due to the applied loads; that is, the member
deforms. The concept of strain is introduced to permit us to give a detailed analytical
description of such deformation. Finally, stress and strain are related to each other. This
relationship, which depends on the material(s) used in the fabrication of the member, must
be determined by performing certain stress-strain tests, which are described in this chapter.

(a) A two-force member (b) Internal resultants: F, V, and M


FIGURE 1.0 An illustration of internal resultants.

1.1 Normal Stress and strain

Equal and opposite forces of magnitude P acting on a straight bar cause it to elongate, and
also to get narrower, as can be seen by comparing Figure 1.1a and 1.1b. The bar is said to
be in tension. If the external forces had been applied in the opposite sense, that is, pointing
toward each other, the bar would have shortened and would then be said to be in
compression.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p2


DMV 4343
JAN ~ JUN `07

FIGURE 1.1
A straight bar
undergoing axial
loading, (a) The
undeformed bar, with
vertical lines indicating
cross sections. (b) The
deformed bar. (c) The
distribution of internal
force at section A. (d)
The distribution of
internal force at section
B.

Definition of Normal Stress. The thin red arrows in Figure 1.1c and 1.1d represent the
distribution of force on cross sections at A and B, respectively. (A cross section is a plane
that is perpendicular to the axis of the bar.) Near the ends of the bar, for example at section
A, the resultant normal force, FA, is not uniformly distributed over the cross section; but at
section B, farther from the point of application of force P, the force distribution is uniform. In
mechanics, the term stress is used to describe the distribution of a force over the area
on which it acts and is expressed as force intensity, that is, as force per unit area.
Force
Stress =
Area
The units of stress are units of force divided by units of area. In the U.S. Customary
System of units (USCS), stress is normally expressed in pounds per square inch (psi) or in
kips per square inch, that is, kilopounds per square inch (ksi). In the International System of
units (SI), stress is specified using the basic units of force (newton) and length (meter) as
newtons per meter squared (N/m2). This unit, called the pascal (1 Pa = 1 N/m2), is quite
small, so in engineering work stress is normally expressed in kilopascals (1 kPa = 10 3 N/
m2), megapascals (1 MPa = 106 N/ m2), or gigapascals (1 GPa = 109 N/ m2). For example, 1
psi = 6895 Pa = 6.895 kPa.
There are two types of stress, called normal stress and shear stress. In this section we will
consider only normal stress; shear stress is introduced in Chapter 2. In words, normal stress
is defined by
Force normal (i.e., perpendicular) to an area
Normal stress =
Area on which the force acts

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p3


DMV 4343
JAN ~ JUN `07

The symbol used for normal stress is the lowercase Greek letter sigma (σ). The
normal stress at a point is defined by the equation
lim Normal
σ(x, y,
= ΔA
( ΔF ) Stress 1.1
z) ΔA
→0

where, as shown in Figure 1.2a, ΔF is the normal force (assumed positive in tension) acting
on an elemental area ΔA containing the point (x, y, z) where the stress is to be determined.
The sign convention for normal stress is as follows:
• A positive value for σ indicates tensile stress, that is, the stress due to a force
ΔF that pulls on the area on which it acts.
• A negative value for σ indicates compressive stress.

(a) Distributed normal (b) Resultant of distributed


stress on a cross section normal stress in (a)
FIGURE 1.2 Normal force on a cross section

Thus, the equation σ = 6.50 MPa signifies that σ is a tensile stress of magnitude 6.50 MPa,
or 6.50 MN/m2, and the equation σ = -32.6 ksi indicates a compressive stress of magnitude
32.6 kips/in2.
Average Normal Stress. Even when the normal stress varies over a cross section, as it
does in Figure 1.1c, we can compute the average normal stress on the cross section by
letting
Average
F
σavg = Normal 1.2
A
Stress
Thus, for Figures 1.1c and 1.1d we get
FA P FB P
(σavg)A = = , (σavg)B = =
A A A A
Much of the rest of this textbook is devoted to determining how stress is distributed on cross
sections of structural members under various loading conditions. However, in many
situations the normal stress on a cross section is either constant or very nearly constant, as
in the next two examples.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p4


DMV 4343
JAN ~ JUN `07

EXAMPLE 1.1
In the evening, a contractor attaches a steel wire to an eyebolt
at point A on his air compressor (Figure 1), and, with the boom of
his construction crane, he raises the compressor a safe distance
above the ground to prevent mischief from being done to the
compressor overnight. While the compressor is above the
ground, the wire will be in tension, and the normal stress on any
cross section of the wire, except near ends A and B, can be
assumed to be constant. If the compressor weighs 600 N and
the diameter of the wire is d = 10 cm, what is the average FIGURE 1

tensile stress in the wire?

Solution From the free-body diagram in Figure 2, the tensile


force in the wire is equal to the weight of the compressor,

so
∑F=0 F = 600 N
and from Eq. 1.2,
σavg = F/A = F/(πd2/4) = 4(600N)/π(0.1m)2 = 76,394 N/m2
Rounded to three significant figures, the average normal stress FIGURE 2 Free body

on typical cross section of the wire is diagram

σavg = 76.4 kPa (T)

where the (T) stands for tension. Actually, this is quite a high
stress, as you will discover in Section 1.4. Figure 3 illustrates
how this average normal stress would be distributed over a
cross section of the wire.

FIGURE 3 Tensile stress


on a typical cross section

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p5


DMV 4343
JAN ~ JUN `07

EXAMPLE 1.2
The Washington Monument (Figure 1a) stands 555 m
high and weighs 181,700 kN. The monument was made
from over 36,000 blocks of marble and granite. As shown
in Figure 1b, the base of the monument is a square that
is 665.5 m long on each side, and the stone walls at the
base are 180 m thick.
Determine the compressive stress that the foundation
exerts over the cross section at the base of the monument,
assuming that this normal stress is uniform.
FIGURE 1

Solution From the free-body diagram in Figure 2, the


total normal force on the base of the monument is equal
to negative of the weight of the monument,
so
∑F=0 F = -181,700 kN
(Note: In accordance with the sign convention for normal
stress, the normal force F is taken positive in tension.
The negative value for F indicates that it is a compressive
force, as is clearly evident in this case.) The cross-
sectional area of the base is
(665.5 m)2 - (665.5 m - 360
A = FIGURE 2 Free body diagram
2
m)
= 349,600 m2

Therefore, from Eq. 1.2,


F -181,700kN
σavg = = = - 519.8 Pa
A 349,600 m2
Rounded to three significant figures, the average normal
stress on the cross section of the monument at its base is
σavg = 520 Pa (C)

where the (C) stands for compression. This is a very low FIGURE 3 Compressive stress
stress, even for stone. Figure 3 illustrates how this uniform at the base
compressive stress would be distributed over the
foundation at the base of the monument.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p6


DMV 4343
JAN ~ JUN `07

The compressive stress that results when one object bears on another, like the stress that
the monument exerts on the foundation in the above example problem, is frequently called
bearing stress. Bearing stress is just a special case of compressive normal stress.

Stress Resultant.
Given the distribution of normal stress on a cross section, σ = σ (x, y, z), we can integrate
over the cross section to determine the magnitude and point of application of the resultant
normal force:

∑ Fx : F(x) = ∫ A
σ dA
∑ My
:
zR F(x) = ∫ A
zσ dA
1.3
∑ Mz
:
-yR F(x) = -∫ A
yσ dA

The two moment equations are used to locate the line of action of the force F(x). Note that
the sign convention for σ implies that the force F in Eq. 1.3 is to be taken positive in tension.
This is the reason that we will consistently, take normal force resultants to be positive in
tension.
Resultant of Constant Normal Stress on a Cross Section: Let us determine the resultant of
normal stress on the cross section at x (Figure 1.2a) if the normal stress is constant over the
cross section. We will prove that normal stress that is constant on a cross section
corresponds to an axial force F(x) = A σ (x) acting through the centroid of the cross section
at x.
Let the resultant be assumed to be a force F(x) acting parallel to the x axis and passing
through point (yR, ZR), as in Figure 1.2b. We must show that
F(x = A(x) σ (x) yR = ZR =
) , y, Z

For this we can use Eqs. 1.3. Substituting the condition σ (x, y, z) = σ (x) into Eqs. 1.3, we get

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p7


DMV 4343
JAN ~ JUN `07

∑ Fx : F(x) = σ (x)∫ dA
A
σ (x)dA
∑ My
:
zR F(x) = σ (x) ∫ zdA
A
σ (x)z A

∑ Mz
:
-yR F(x) = - σ (x) ∫ ydA A
-σ(x) yA

For generality, the normal force has been permitted to be a function of x in Figure 1.2b and in
Eqs. 1.3. Of course, F(x) = P = const in the axial-loading case illustrated in Figure 1.1.

Therefore, if the normal stress is uniform over a cross section, the normal stress on the
cross section, also called the axial stress, is given by

F(x
Axial-
σ )
= Stress 1.4
(x) A(x
Equations
)

and corresponds to a force F(x) (tension positive) acting at the centroid of the cross section,
that is, at Z = z, yR = y.
R

We would certainly expect uniform stress on a circular rod to correspond to a force acting
along the axis of the rod, and similarly for a square or rectangular bar. Hence, it is
"reasonable" that a uniform normal stress distribution acting over a cross section of general
shape produces a resultant force acting through the centroid of the cross section. In most
cases, the cross-sectional area is constant throughout the length of the member, but Eq. 1.4
may also be used if the cross-sectional area varies slowly with x.

Uniform Normal Stress in an Axially Loaded Bar: Under certain assumptions, an axially
loaded bar will have the same uniform normal stress on every cross section; that is, σ (x, y,
z) = σ = constant. These assumptions are:

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p8


DMV 4343
JAN ~ JUN `07

• The bar is prismatic; that is, the bar is straight and it has the same cross section
throughout its length.
• The bar is homogeneous; that is, the bar is made of the same material
throughout.
• The load is applied as equal and opposite uniform stress distributions over
the two end cross sections of the bar.

So long as the resultant force at each end of the bar is applied at the centroid of the end
cross section, the last assumption—that the loads are applied as uniform normal stress
distributions on the end cross sections—can be relaxed. As illustrated in Figure 1.1 a-d, the
stress is uniform on every cross section, except on cross sections that are very near the
points of application of load. This is an application of Saint-Venant's Principle.
The uniform, prismatic bar in Figure 1.3a is labeled as member "i" and is subjected to equal
and opposite axial forces Fi, acting through the centroids at its ends. Its cross-sectional area
is Ai.
The normal stress on cross sections of an axially loaded member, like the one in Figure 1.3,
is called the axial stress. Since, from the free-body diagram in Figure 1.3b, the resultant
force, F(x), on every cross section of the bar is equal to the applied load Fi, and since the
cross-sectional area is constant, from Eq. 1.4 we get the following formula for the uniform
axial stress:
F
Axial-
i
σi = = constant Stress 1.5
A
Equations
i

FIGURE 1.3 Uniform stress in an axially loaded prismatic bar

Example 1.3 shows one application of the axial-stress equation.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p9


DMV 4343
JAN ~ JUN `07

EXAMPLE 1.3
Two solid circular rods are welded to a plate at B
to form a single rod, as shown in Figure 1.
Consider the 30-kN force at B to be uniformly
distributed around the circumference of the collar
at B and the 10 kN load at C to be applied at the
centroid of the end cross section. Determine the
axial stress in each portion of the rod. FIGURE 1
Plan the Solution
Since each segment of the rod satisfies the
conditions for uniform axial stress, we can use
Eq. 1.5 to calculate the two required axial
stresses. First, however, we need to compute the
force in each rod by using an appropriate free-
body diagram and equation of equilibrium.
FIGURE 2
Solution
Free-body Diagrams: First we draw free-body diagrams that expose the rod forces F1 (or
FAB) and F2 (or FBC). We show F1 and F2 positive in tension.
Equations of Equilibrium: From free-body diagram 1 (Figure 2a),

+
∑ (F)1 = 0: - F1 - 30kN + 10kN = 0, F1 = - 20 kN

And, from free-body diagram 2 (Figure 2b),

+
∑ (F)2 = 0: - F2 + 10kN = 0, F2 = 10 kN

Π d12 Π (20 mm)2
A1 = = = 314.2 mm2
4 4
Π d22 Π (15 mm)2
A2 = = = 176.7 mm2
4 4

Axial Stresses: Using Eq. 1.4, we obtain the axial stresses


F1 - 20 kN
σ1 = = 2
= - 63.7 MN / m2 @ 63.7 MPa (C)
A1 341.2 mm
F2 10 kN
σ2 = = 2
= 56.6 MN / m2 @ 56.6 MPa (T)
A2 176.7 mm

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p10


DMV 4343
JAN ~ JUN `07

The free-body diagrams in Figure 2 of Example 1.3 illustrate the application of the method of
sections to structures that have axially loaded segments that are collinear. For example, the
free-body diagram in Figure 2a makes it possible with a single free-body diagram to relate
the internal force F1 to the external forces applied at the two nodes (joints) B and C; this is
the most efficient way to determine F1.

STRAIN
Figure 1.4a illustrates an underformed tension specimen with two points on the specimen
marking the original gage length, L0. The notation L0 is used here to emphasize that this is
the original gage length, not the total length of the specimen. An axial load P causes the
portion of the specimen between the gage marks to elongate, as indicated in Figure 1.4b.

(a) Undeformed
specimen.

(b) Deformed
specimen.

FIGURE 1.4 Tension and compression test specimens.

As the specimen is pulled, the load P is measured by the testing machine and recorded. The
extensometer provides a simultaneous measurement of the corresponding length. L* =
L*(P), of the test section, or else it directly measures the elongation

Δ = L - L0 1.6
L *

In a static tension test the length of the specimen is increased very slowly, in which case the
loading rate need not be measured. In some situations, however, a dynamic test must be
performed. Then, the rate of loading must be measured and recorded, since material
properties are affected by high rates of loading.

L* - L0 Norma
Є =
L0 l Strain

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p11


DMV 4343
JAN ~ JUN `07

1.2 Stress – Strain Relationship and Diagram

In order to relate the loads on engineering structures to the deformation produced by the
loads, experiments must be performed to determine the load-deformation behavior of the
materials (e.g., aluminum, steel, and concrete) used in fabricating the structures. Many
useful mechanical properties of materials are obtained from tension tests or from
compression tests, and these properties are listed in tables like those in Appendix F. This
section describes how a tension test is performed and discusses the material properties that
are obtained from this type of test.

Tension Tests and Compression Tests.


Figure 1.5 shows a computer-controlled, hydraulically actuated testing machine that may be
used to apply a tensile load or a compressive load to a test specimen, like the steel tension
specimen in Figure 1.6a or the concrete compression specimen in Figure 1.6c. Figure 1.6b
shows a close-up view of a ceramic tension specimen mounted in special testing-machine
grips. Electromechanical extensometers are mounted on the specimens in Figure 1.6a and
1.6b to measure the extension (i.e., the elongation) that occurs over the gage length of the
test section.

FIGURE 1.5 A computer-controlled hydraulically actuated testing machine, (MTS Systems


Corp. photo.)

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p12


DMV 4343
JAN ~ JUN `07

(a) A metal tension (b) A ceramic tension (c) A concrete cylinder


specimen with specimen with before and after
extensometer attached extensometer attached. compression testing.
(MTS Systems Corp.
photo.)

FIGURE 1.6 Tension and compression test specimens.

Stress-Strain Diagrams.
A plot of stress versus strain is called a stress-strain diagram, and from such stress-strain
diagrams we can deduce a number of significant mechanical properties of materials.1" The
values of normal stress and extensional strain that are used in plotting a conventional
stress-strain diagram are the engineering stress (load divided by original cross-sectional
area of the test section) and engineering strain (elongation divided by original gage length),
that is,
P L* - L0
σ = , Є = 1.7
A0 L0

Mechanical Properties of Materials.


Figures 1.7a and 1.7b are stress-strain diagrams for structural steel (also called mild steel,
or low-carbon steel), which is the metal commonly used in fabricating bridges, buildings,
automotive and construction vehicles, and many other machines and structures. A number
of important mechanical properties of materials that can be deduced from stress-strain
diagrams are illustrated in Figure 1.7. In Figure 1.7a the stress is plotted accurately, but the
strain is plotted to a variable scale so that all important features can be shown and
discussed. In Figure 1.7b, which gives typical numerical values of stress and strain for
structural steel, one stress-strain curve, the lower one is plotted against a strain scale that

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p13


DMV 4343
JAN ~ JUN `07

emphasizes the low-strain region; the upper curve is plotted against a strain scale that
emphasizes the high-strain region and puts the entire stress-strain history into perspective.
Starting at the origin A in Figure 1.7a and continuing to point B, there is a linear relationship
between stress and strain. The stress at point B is called the proportional limit, σPL.

(a) Strain not plotted to scale

(b) Strain plotted to two different scales


FIGURE 1.7 Stress-strain diagrams for structural steel in tension.

The ratio of stress to strain in this linear region of the stress-strain diagram is called Young's
modulus, or the modulus of elasticity, and is given by

Δσ Young’s
E = , σ < σPL 1.12
ΔЄ Modulus
Typical units for E are ksi or GPa.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p14


DMV 4343
JAN ~ JUN `07

At B the specimen begins yielding, that is, smaller and smaller increments at load are
required to produce a given increment of elongation. The stress at C is called the upper
yield point, (σYP)u, while the stress at D is called the lower yield point, (σYP)l. The upper
yield point has little practical importance, so the lower yield point is usually referred to simply
as the yield point, σYP. From D to E the specimen continues to elongate without any increase in
stress. The region DE is referred to as the perfectly plastic zone. The stress begins to
increase at E, and the region from E to F is referred to as the zone of strain hardening. The
stress at F is called the ultimate stress, or ultimate strength, σF. At F the load begins to drop,
and the specimen begins to "neck down." This neck-down behavior continues until, at G, fracture
occurs at the fracture stress, σF. Figure 1.8a, shows a hot-rolled steel specimen at three stages of
tensile testing: (1) before testing, (2) as removed from the testing machine at a point between F and
G with pronounced reduction in area referred to as necking or neck-down, and (3) after fracture.
Figure 1.8 b shows the typical cup and cone fracture of a hot-rolled steel tensile specimen.

(a) (b)
FIGURE 1.8 A hot-rolled steel tensile specimen

The true stress, σtrue, is the load at some instant during the test divided by the actual
minimum cross-sectional area of the specimen at that instant. Thus, when a specimen
starts to neck down, the true stress is taken as the load divided by the minimum cross-
sectional area in the neck-down region. The true strain, Єtrue, is the instantaneous change
in length of a test section divided by the instantaneous length of that test section.
True stress and true strain are given by the formulas
P
σtrue = , Єtrue = ln (1 + Є) 1.13
Amin

True strain can also be expressed in terms of area change

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p15


DMV 4343
JAN ~ JUN `07

A0
Єtrue = ln ( )
Amin
The solid-line curve in Figure 1.7a is a conventional stress-strain diagram of engineering
stress versus engineering strain, while the dashed curve is a sketch of true stress versus
true strain. The curves differ only when strain is large and when the cross-sectional area is
decreasing significantly.

Design Properties.
Now that you have some idea of the stress-strain behavior of several common metallic
materials, let us note the material properties that are of primary interest to an engineer
designing some structure or machine. From the design standpoint the most significant
stress-strain properties can be categorized under the three headings—strength, stiffness,
and ductility.
 Strength—There are three strength values of interest,
1. The yield strength, σY, is the highest stress that the material can withstand
without undergoing significant yielding. The yield-point stress or the offset
yield stress, whichever is appropriate for the particular material, is taken as
the yield strength (i.e., σY = σYP or σY = σYS as appropriate).
2. The ultimate strength, σU is the maximum value of stress (i.e., the maximum
value of engineering stress) that the material can withstand. Finally,
3. The fracture stress, σF, if different from the ultimate stress, may be of
interest. It is the value of the stress at fracture.
 Stiffness—The stiffness of a material is basically the ratio of stress to strain.
Stiffness is of interest primarily in the linearly elastic region; therefore, Young's
modulus, E, is the value used to represent the stiffness of a material.
 Ductility—Materials that can undergo large strain before fracture are classified as
ductile material; those that fail at small values of strain are classified as brittle
materials. Strictly speaking, the terms ductile and brittle refer to modes of fracture,
and a material like structural steel, which behaves in a ductile manner at room
temperature, may exhibit brittle behavior at very low temperatures. Therefore, when
we speak of a "brittle material" or a "ductile material," we are referring to the normal
(room temperature) behavior of the material.

Ductility. The two commonly used measures of ductility are the percent elongation (the final
elongation expressed as a percentage of the original gage length) and the percent reduction

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p16


DMV 4343
JAN ~ JUN `07

in area at the section where fracture occurs (the area reduction expressed as a percentage
of the original area). The percent elongation is given by the formula

LF - L0
Percent elongation = (100%)
L0

where L0 is the original gage length and LF is the length of the gage section at fracture.
When percent elongation is stated, the gage length should also be stated, since the
elongation at fracture is not uniform over the entire gage length but is concentrated in the
necked-down region. The percent reduction in area is given by the formula

A0 – AF
Percent reduction in area = (100%)
A0

FIGURE 1.9 A superplastically deformed steel-alloy specimen tested over 1100%


elongation. (Courtesy of Prof. Oleg D. Sherby, Standford University).

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p17


DMV 4343
JAN ~ JUN `07

1.3 Elasticity and Plasticity

In the previous section we considered the stress-strain behavior of tension specimens under
loading. Specifically, the tensile tests were assumed to be performed in a relatively short
time with monotonically increasing tensile strain. We consider now what happens when the
loading is reversed, that is, when the strain is allowed to decrease.

Elastic Behavior and Plastic Behavior.


Consider the loading and unloading behavior of a material, as illustrated in Figure 1.10. The
upward-pointing arrows in Figure 1.10a and Figure 1.10b indicate the loading curves, that is,
the curves that would be followed for monotonically increasing initial loading. The stress-
strain behavior of a material is said to be elastic if the unloading path retraces the
loading path. The stress at C is called the elastic limit, σEL. Since the stress-strain curve
from the origin A in Figure 1.10a to the elastic limit at C is not a straight line, material
behavior in this range is termed nonlinear elastic behavior. Unloading from a point below
point C (e.g., point B) retraces the loading path indicated in Figure 1.10a, but unloading from
a point beyond C (e.g., point D) follows a path DE that is different than the loading path.
Unloading from a point on the stress-strain curve beyond the elastic limit typically follows a
straight-line path whose slope is parallel to the tangent of the stress-strain curve at the
origin. The strain that remains at point E when the stress returns to zero is called the
permanent set, or residual strain, as illustrated in Figure 1.10b.

FIGURE 1.10 Illustrations of elastic and plastic stress-strain behaviour.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p18


DMV 4343
JAN ~ JUN `07

1.4 Linear Elasticity and Hooke’s Law

For structural steel the elastic limit is very close to the proportional limit and also to the yield
point. Therefore, for structural steel the proportional limit stress and the elastic limit stress
are usually assumed to be equal to the yield-point stress. It is sometimes convenient to
approximate the stress-strain behavior of mild steel and similar materials by the linearly
elastic, perfectly plastic representation of Figure 1.11.

FIGURE 1.11 Linearly elastic, perfectly plastic material model.

In order to keep deformations small and stresses at safe levels, most structures and
machine parts are designed so that stresses remain well below the yield stress. Fortunately,
most engineering materials exhibit a linear stress-strain behavior at these lower stress
levels.

Hooke's Law.
Let us, for the present, restrict our discussion to the case of uniaxial stress applied to a
homogeneous (same properties throughout), isotropic (same properties in every direction)
member oriented along the x axis (Figure 1.12). The linear relationship between stress and
strain, given by Eq. 2.12, applies for 0 ≤ σ ≤ σY. Therefore,
σ
= EЄx Hooke’s Law 1.11
x

This equation is called Hooke's Law." The subscripts on σ and Є identify the axis of the
particular stress and strain.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p19


DMV 4343
JAN ~ JUN `07

FIGURE 1.12 The deformation (much exaggerated)


of a homogeneous, isotropic specimen under uniaxial
stress.

Hooke's Law is valid for uniaxial tension or compression within the linear portion of the
stress-strain diagram. As noted earlier, E is called the modulus of elasticity, or Young's
modulus. It has the units of stress, typically ksi (or psi) in U.S. Customary units, and GPa (or
MPa) in SI units. Representative approximate values of E are: 30 X 103 ksi (200 GPa) for
steel, 10 X 103 ksi (70 GPa) for aluminum, and 300 ksi (2 GPa) for nylon. Values of the
modulus of elasticity for selected materials are listed in Table F.2.

Poisson's Ratio.
Associated with the elongation of a member in axial tension, there is a transverse
contraction, which is illustrated in Figure 1.12. The transverse contraction during a tensile
test is related to the longitudinal elongation by
Єtranvs = - vЄlongit Poisson’s Ratio 1.12

where v (Greek symbol nu) is Poisson's ratio. This expression also holds when the
longitudinal strain is compressive; then, the lateral strain results in an expansion of the
transverse dimensions. Poisson's ratio is dimensionless, with typical values in the 0.25-0.35
range. For the orientation of axes in Figure 1.12 the transverse strains are related to the
longitudinal strain by
Є
= Єz = - vЄx 1.13
y

Equations 1.11 and 1.13 apply to the simple case of uniaxial stress.
All the examples considered so far in this chapter have dealt with slender members subjected
to axial loads, i.e., to forces directed along a single axis. Choosing this axis as the X- axis, and
denoting by P the internal force at a given location, the corresponding stress components
were found to be σx = P/A, σy = 0, and σz = 0.
Let us now consider structural elements subjected to loads acting in the directions of the
three coordinate axes and producing normal stresses σx, σy, and σz which are all different from
zero (Figure 1.13). This condition is referred to as a multiaxial loading. Note that there is no
shearing stresses are included among the stresses shown in Figure 1.13.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p20


DMV 4343
JAN ~ JUN `07

FIGURE 1.13 A body subjected to a multiaxial loadings


The approach we propose here will be used repeatedly in this text, and is based on the
principle of superposition. This principle states that the effect of a given combined loading on a
structure can be obtained by determining separately the effects of the various loads and
combining the results obtained, provided that the following conditions are satisfied:
1. Each effect is linearly related to the load that produces it.
2. The deformation resulting from any given load is small and does not affect the
conditions of application of the other loads.
In the case of a multiaxial loading, the first condition will be satisfied if the stresses do not
exceed the proportional limit of the material, and the second condition will also be satisfied if the
stress on any given face does not cause deformations of the other faces that are large enough
to affect the computation of the stresses on those faces.
Considering first the effect of the stress component σx, recall that Єx = σx/E in the x direction,
and strains equal to Єx = -νσx/E in each of the y and z directions. Similarly, the stress
component σy if applied separately, will cause a strain Єy = σy/E in the y direction and strains Єy
= -νσy/E in the other two directions. Finally, the stress component σz causes a strain σz /E in the
z direction and strains Єy = -νσy/E in the x and y directions. Combining the results obtained, we
conclude that the components of strain corresponding to the given multiaxial loading are
νσ νσ
σx
Єx = + - y
- z
E E E
νσ νσ
σy Generalized Hooke’s
Єy = - x
+ - z
Law
E E E
νσ νσ
σz
Єz = - x
- y
+
E E E
The relations (2.28) are referred to as the generalized Hooke's law for the multiaxial loading of
a homogeneous isotropic material. As we indicated earlier, the results obtained are valid only
as long as the stresses do not exceed the proportional limit, and as long as the deformations
involved remain small. We also recall that a positive value for stress component signifies

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p21


DMV 4343
JAN ~ JUN `07

tension, and a negative value compression. Similarly, a positive value for a strain component
indicates expansion in the corresponding direction, and a negative value contraction.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p22


DMV 4343
JAN ~ JUN `07

1.5 Allowable Stress and Allowable Load

An engineer in charge of the design of a structural member or mechanical clement must


restrict the stress in the material to a level that will be safe. Furthermore, a structure or
machine that is currently in use may, on occasion, have to be analyzed to see what additional
loadings its members or parts can support. So again it becomes necessary to perform the
calculations using a safe or allowable stress.
To ensure safety, it is necessary to choose an allowable stress that restricts the applied load to
one that is less than the load the member can fully support. There are several reasons for
this. For example, the load for which the member is designed may be different from actual
loadings placed on it. 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.
Atmospheric corrosion, decay, or weather ing tends to cause materials to deteriorate during
service. And lastly, 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 the design or analysis of a member is to
use a number called the factor of safety. The factor of safety (F.S.) is a ratio of the failure
load Ffail divided by the allowable load, F allow. Here Ffail is found from experimental testing of the
material, and the factor of safety is selected based on experience so that the above mentioned
uncertainties are accounted for when the member is used under similar conditions of loading
and geometry. Stated mathematically,

Ffail
F.S =
Fallow

If the load applied to the member is linearly related to the stress developed within the
member, as in the case of using σ = P/ A then, we can express the factor of safety as a ratio
of the failure stress σ fail to the allowable stress σ allow that is,

σ fail
F.S =
σ allow

Where F.S must be greater than 1. Normally: 1< F.S < 3.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p23


DMV 4343
JAN ~ JUN `07

In any of these equations, the factor of safety is chosen to be greater than 1 in order to avoid
the potential for failure. Specific values depend on the types of materials to be used and the
intended purpose of the structure or machine. 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. On the other hand, in the case of a nuclear power plant, the factor of safety for some
of its components may be as high as 3 since there may be uncertainties in loading or
material behavior. In general, however, factors of safety and therefore the allowable loads or
stresses for both structural and mechanical elements have become well standardized, since
their design uncertainties have been reasonably evaluated. Their values, which can be
found in design codes and engineering handbooks, are intended to form a balance of
insuring public and environmental safety and providing a reasonable economic solution to
design.

DESIGN OF SIMPLE CONNECTIONS

By making simplifying assumptions regarding the behavior of the material, the equations σ =
P/ A can often be used to analyze or design a simple connection or a mechanical element.
In particular, if a member is subjected to a normal force at a section, its required area at the
section is determined from
P
A =
σ allow

Cross-Sectional Area of a Tension Member.


The cross-sectional area of a prismatic member subjected to a tension force can be
determined provided the force has a line of action that passes through the centroid of the
cross section. For example, consider the "eye bar" shown in Fig. 1-27a. At the intermediate
section a-a, the stress distribution is uniform over the cross section and the required area A
is determined, as shown in Fig. 1-27b

Required Area to Resist Bearing.


A normal stress that is produced by the compression of one surface against another is called
a bearing stress. If this stress becomes large enough, it may crush or locally deform one or
both of the surfaces. Hence, in order to prevent failure it is necessary to determine the
proper bearing area for the material using an allowable bearing stress. For example, the
area A of the column base plate B shown in Fig. 1-29 is determined from the allowable
bearing stress of the concrete using A = P/(σb)allow. This assumes, of course, that the
allowable bearing stress for the concrete is smaller than that of the base plate material, and

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p24


DMV 4343
JAN ~ JUN `07

furthermore the bearing stress is uniformly distributed between the plate and the concrete as
shown in the figure.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p25


1.6 Deformation of Axially Loaded Member

Axial deformation is characterized by two fundamental kinematic assumptions:


1. The axis of the member remains straight.
2. Cross sections, which are plane and are perpendicular to the axis before
deformation, remain plane and remain perpendicular to the axis after deformation.
And, the cross sections do not rotate about the axis.

These assumptions are illustrated in Figure 1.13, where A and B designate cross sections at
x and (x + Δx) prior to deformation, and where A* and B* designate these same cross
sections after deformation.
The distance that a cross section moves in the axial direction is called its axial displacement.
The displacement of cross section A is labeled u(x), while the neighboring section B
displaces an amount u(x + Δx). The displacement u(x) is taken to be positive in the +x
direction. We can derive a strain-displacement expression that relates the axial strain Є to
this axial displacement u by considering the fundamental definition of extensional strain:

Final length – Initial length


Є =
Initial length

FIGURE 1.13 The geometry of axial deformation.

The axial strain of any fiber1 of infinitesimal length Δx that is parallel to the x axis and
extends from section A to section B of the undeformed member may be determined from the
fundamental definition of extensional strain at a point. By letting the initial length of a typical
fiber be Δx, and then letting Δx approach zero, we can write the following expression for the
axial strain:
Єx(x = lim ( Δx* - ) = lim [ u(x + Δx)- ] = d
DMV 4343 / DMV 5343
JAN~JUN 2007

Δx ux u
) Δx Δx d
Δx→0 Δx→0 x

Therefore, the axial strain at section x is the derivative (with respect to x) of the axial
displacement, or
Strain
Є(x du(x Displacemen
=
) ) t
dx
Equation

This equation relating axial strain to axial displacement is called the strain-displacement
equation for axial deformation. The two fundamental kinematic assumptions stated above
imply that the axial strain Є may be a function of x, but that it is not a function of position in
the cross section, that is, of y or z. To emphasize this point, a plot of the strain distribution at
an arbitrary cross section at x is shown in Figure 1.14 superimposed on a sketch of a portion
of the member. To reiterate, axial deformation is characterized by extensional strain that is
not a function of position in the cross section.
Figure 1.14 Extensional stress
distribution for a member undergoing
axial deformation.

Figure 1.15 Definition


of the total elongation
e.

As indicated in Figure 1.15, the total elongation of the member is the difference between the
displacements of its two ends, that is,
u
e = - u (0)
(L)

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p27


DMV 4343 / DMV 5343
JAN~JUN 2007

By summing up the changes in length of increments dx over the entire length of the
member, we get the following equation for the elongation of an axial-deformation member of
initial length L:
L
Elongation
e = ∫ 0
Є(x) dx
Formula
How the strain Є varies with x depends on the loads that are applied to the member,
whether is has a constant cross section or the cross section varies with x, and the behavior
of the material (or materials) from which the member is constructed. Next, we will consider
this material behavior.

Material Behavior.
Having the strain distribution given by Eq. 1.14 and illustrated in Figure 1.14, we can now
employ the uniaxial stress-strain behavior of materials to determine the stress distribution in
a member undergoing axial deformation. Let us first take the simplest case—linearly elastic
behavior—and let us assume that the temperature remains constant (i.e., ΔT = 0) and that
σY = σ = 0. Then, from Eq. 1.14 we get the uniaxial stress-strain equation
Z

σ = σ
= EЄx Hooke’s Law 1.14
x

Equation 1.14 gives the distribution of the axial stress σ on the cross section at x. As
indicated in Eq 1.14, Є may only vary with x, but not with position in the cross section. Most
axial-deformation members are homogeneous, so that Young's modulus, E, is constant
throughout the member.
Stress Resultant. Deformable-body mechanics problems are simplified by making
assumptions that reduce a basically three-dimensional problem to a one-dimensional
problem, like the axial-deformation kinematic assumptions discussed earlier in this section.
Given the distribution of axial stress σ (Eq. 1.14), we can replace the distributed stress by a
single resultant axial force and relate that force to the axial stress on the cross section.
Figure 1.15 shows the three stress resultants that are related to the axial stress σ, and
defines the sign convention for these stress resultants. The resultants in Figure 1.15 are
shown for the +x face: equal and opposite resultants act on the -x face at the cross section.
The axial force, F, on the cross section will always be taken to be positive in tension. The
bending moments My and Mz are taken positive according to the right-hand rule. By
summing up the contributions to these resultants of the forces dF on infinitesimal areas, dA,
we get

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p28


DMV 4343 / DMV 5343

F(x) = ∫ A
σ dA
JAN~JUN 2007

Stress
My(x) = ∫ A
zσ dA
Resultants
Mz(x) = -∫ A
yσ dA

Equations 1.15 define the three stress resultants associated with the axial stress σ.
However, when the distribution of stress on the cross section is known, we can combine
these to form a single resultant force on the cross section. Consider the most common case,
a homogeneous linearly elastic member. Then, Young's modulus is constant throughout
the entire member, so Eq. 1.14 takes the form
Stress
σ(x EЄ(x
= Distribution
) )
(E = const)

(a) Stress resultants on the cross section at x.

(b) Axial stress at a point in the cross section at x.

FIGURE 1.15

In previous section we have cover that strain, Є is


L* - L0 Norma
Є =
L0 l Strain

Where L* - L0 is the change in length in the axis of applied load. If we let L* - L0 = δ and L0 =
L, then

δ
Є =
L

Stress,
σ=P/A

From Hooke’s Law

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p29


DMV 4343 / DMV 5343
JAN~JUN 2007

σx
E =
Єx

Substituting in Hooke’s Law we get,

P/A PL
E = =
δL Aδ

Since E = constant, we can manipulate this equation

PL
δ =
AE

EXAMPLE 1.4
The cylindrical rod in Figure 1 is made of steel with E = 30 MPa, v = 0.3, and σY = 50 MPa. If
the initial length of the rod is L = 4 m and its original diameter is d = 10 cm, what is the
change in length, ΔL, and what is the change in diameter, Δd, due to the application of an
axial load P = 10 kN?

FIGURE 1
Plan the Solution
From the load P and the cross-sectional area A we can determine the stress σ. Then, if σ ≤
σY, we can use Hooke's Law to relate the stress σ to the strain Є. Then we can relate the
uniform strain, Є, to the elongation ΔL. The change in diameter is due to the Poisson’s ratio
effect.
Solution

FIGURE 2 Free-body diagram

Equilibrium: From the free-body diagram in Figure 2,


∑Fx =0 F(x) = P = const

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p30


DMV 4343 / DMV 5343
JAN~JUN 2007

From Eq. 1.5,


σx = P/A = 10 kN / π (0.05 m)2 = 1.273 MPa
Therefore, linearly elastic behaviour occurs when the load P = 10 kN is applied to the bar.
Material Behavior: From Eq 1.14

Є σx PL
= =
x E AE

And, from Eq 1.7


-v P
Єradial = -v Єx =
AE

Strain-Displacement: From Eq. 2.7,


PL
ΔL = L Єx =
AE
And
- v Pd
Δd = dЄradial =
AE

Substituting numerical values into Eqs (3), we get

PL (10 kN)(4 m)
ΔL = L Єx = = = 0.17 m.
AE Π (0.05 m)2 (30 MPa)
And
- v Pd - 0.3(10 kN)(0.1 m)
Δd = dЄradial = = = -1.273 mm.
AE Π (0.05 m)2 (30 MPa)

Review the Solution


The changes in length and diameter are quite small in comparison with the original length
and the original diameter, respectively, as they should be. Δd should definitely be smaller
than ΔL, and the signs should be different, which is the case. The extensometer in Figure
1.5a is capable of measuring both Δd and ΔL.

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p31


DMV 4343 / DMV 5343
JAN~JUN 2007

Chapter 1 INTRODUCTION TO STRESS AND STRAIN p32

You might also like