0% found this document useful (0 votes)
46 views

SM Lab Manual BUET

The Structural Mechanics and Materials Testing Laboratory Manual provides comprehensive guidelines on material testing, including background theory and practical workbooks for various tests. It covers essential topics such as static tension and compression tests, shear tests, hardness tests, impact tests, and the mechanical properties of engineering materials. The manual also emphasizes the importance of standardized testing procedures and report writing for accurate results and documentation.

Uploaded by

nirob133611
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
46 views

SM Lab Manual BUET

The Structural Mechanics and Materials Testing Laboratory Manual provides comprehensive guidelines on material testing, including background theory and practical workbooks for various tests. It covers essential topics such as static tension and compression tests, shear tests, hardness tests, impact tests, and the mechanical properties of engineering materials. The manual also emphasizes the importance of standardized testing procedures and report writing for accurate results and documentation.

Uploaded by

nirob133611
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 97

STRUCTURAL MECHANICS

AND MATERIALS
TESTING LABORATORY
MANUAL

DEPARTMENT OF CIVIL ENGINEERING


DHAKA INTERNATIONAL UNIVERSITY (DIU)
CONTENTS

PART I BACKGROUND THEORY


CHAPTER 1 INTRODUCTION TO MATERIAL TESTING
1.1 Introduction 1
1.2 Layout of the Book 1
1.3 Testing of Materials 1
1.3.1 Experiment and Test 1
1.3.2 Field test and Laboratory Test 1
1.3.3 Destructive and Non-Destructive Tests 2
1.3.4 Material Test and structural model Test 2
1.4 Standard Specification 2
1.5 Types of Loading 3
1.6 Report Writing 3
1.6.1 Structure of the Report 3
1.6.2 Tips on Graph Plotting 4
1.6.3 Tips on Writing Discussions on the Test Results 6
1.7 Recommended Laboratory Practice 6

CHAPTER 2 STATIC TENSION AND COMPRESSION TEST


2.1 Introduction 7
2.2 Definition of the Terms 7
2.3 Stress-Strain Diagram of Mild Steel 11
2.4 Measurement of Ductility 13
2.5 Engineering Stress and True Stress 14
2.6 Speed of Loading 14
2.7 Yield Strength by Offset Method 15
2.8 Proof Strength 16
2.9 The Standard Tension Test 17
2.9.1 Test for Conventional Stress and Strain 17
2.9.2 Construction of the Stress-Strain Diagram 19
2.9.3 Example: Stress-Strain Diagram 20

CHAPTER 3 SHEAR TEST OF METAL SPECIMEN


3.1 Introduction 23
3.2 Shear Stress Distribution in Beams 23
3.3 Torsion 26
3.4 Behavior of a Closely Coiled Helical Spring 27
3.5 Shear Tests 29

CHAPTER 4 HARDNESS TEST OF METAL


4.1 Introduction 31
4.2 Hardness Measurement 31
4.2.1 Pyramid Hardness 31
4.2.2 Brinell Hardness 32
4.2.3 Rockwell Hardness 33
4.3 Measurement of Rockwell Hardness 34
4.3.1 Operating principle of Rockwell Hardness Tester 34
4.3.2 The Minor Load Feature of the Rockwell Machine 35
4.3.3 Dial Reading Technique 35

ii
4.3.4 Location and Spacing of Rockwell Test Impression 36
4.4 Relationship between Hardness Numbers 36
4.5 Relation between the Hardness Number and Other Properties 40

CHAPTER 5 IMPACT TEST: FLEXURE AND TENSION


5.1 General 41
5.2 Impact Testing Machine 41
5.3 Specimens for Impact Testing 42
5.4 Impact Testing 43
5.5 Usefulness of Impact Tests 43

CHAPTER 6 BUCKLING OF SLENDER COLUMNS


6.1 Introduction 44
6.2 The Buckling Phenomenon 44
6.3 Euler’s Theory for Slender Columns 45
6.4 End Restrained Column: The Effective Length Concept 49

CHAPTER 7 BENDING OF BEAMS


7.1 Theory of Bending 50
7.2 Two Point Loading 53

CHAPTER 8 MECHANICAL PROPERTIES OF


ENGINEERING MATERIAL
8.1 Introduction 54
8.2 Materials for Engineering Use 54
8.2.1 Ferrous Metals 54
8.2.2 Non-Ferrous Metals 56
8.3 Desirable Mechanical Properties in Particular Situation 57

PART II WORKBOOK FOR LABORATORY PRACTICE


Work Book 1 Tension Test of Mild Steel 61
Work Book 2 Compression Test on Timber Specimen 66
Work Book 3 Direct Shear Test of Metal Specimens 70
Work Book 4 Test of Helical Spring 74
Work Book 5 Hardness Test of Metal Specimens 78
Work Book 6 Impact Test of Metal Specimens 82
Work Book 7 Slender Column Test for Different 86
End Conditions
Work Book 8 Static Bending Test of Timber Beam 90

iii
PART I
BACKGROUND THEORY

iv
CHAPTER 1
INTRODUCTION TO MATERIAL TESTING
1.1 Introduction
This introductory chapter is intended to provide the reader a guideline in using this book.
The layout of the book is described first, followed by a general overview on the material
testing. Also included are tips for preparing laboratory reports that the students are
expected to submit upon completion of a test. The student should, therefore, carefully read
this chapter and refer to it for a general guidance in the preparation of his report in a
standard form and also for handy tips on methods of preparing graphs to present test
information which requires curve plotting.

1.2 Layout of the Book


In Part I of this book, there are seven more chapters apart from this introductory one. Part
II of this book contains eight workbooks on the eight different tests covered. The
workbooks have been designed to enable the students to independently carry out the tests,
to record the observations, and to aid reporting by the student. The tabular format of results
provided may also serve as model for reports of practical testing. For each workbook there
is a chapter of tests, which includes the background materials about the particular test.

Chapter 2 covers the two fundamental tests, namely the tension test and the compression
test. Chapter 3 deals with shear test. Both direct shear and torsional shear are included.
Chapter 4 elucidates the various aspects of hardness measurement of metals. Chapter 5 is
about measuring the impact strength of metals. Chapter 6 introduces the buckling
phenomenon of slender columns and chapter 7 overviews the bending stresses in beams.

A final chapter has been included which provides a brief description of the properties of
common engineering materials. It is common and at the same time quite important for an
engineer to appreciate the limits of these properties within which they must fall for any
given use of a material.

1.3 Testing of Materials


The purpose of testing is mainly to obtain information about material properties. The term
material testing, in general, refers to the resistance to failure of an entire piece of a
material, a small part of it, or even the surface. Since material testing is concerned with the
strength of materials, the term ‘testing machine’ refers to a machine for applying known
loads.

1.3.1 Experiment and Test


The terms experiment and test are often loosely used although a distinction may be made
between them. Experimentation involves the idea that the outcome may be uncertain i.e.,
unknown results may be forthcoming. Testing involves the idea of a more or less
established procedure and that the limits of the results are generally defined.

1.3.2 Field test and Laboratory Test


Unlike the fine distinction of the two terms presented in the preceding paragraph, the
distinction between the terms field test and laboratory test is quite easily apparent. But one
should be wary of the fact that because of difficult or hazardous working conditions,
interference, time limitations and variable weather conditions tests conducted in the field
usually lack the precision of similar tests conducted in the laboratory. However,
performance of work in the laboratory does not necessarily ensure precision. Again there

Structural Mechanics and Materials Testing Lab Page 1


are certain tests which cannot be performed in the laboratory. So that the question of field
test versus laboratory test does not arise.

1.3.3 Destructive and Non-Destructive Tests


In respect of the usability of a tested material or any part thereof, tests may be classified as
destructive or non destructive. Tests to determine ultimate strength of a given material
naturally mean destruction of the sample. For finished products it is desirable to use non-
destructive test (NDT), if possible. Some hardness tests are of this type. Proof tests, applied
to fabricated parts of structural elements, are of the non-destructive type; e.g., a proof test
of a crane hook involves the application of a load some-what in excess of the working load
but less than any damaging load, in order to give assurance that no harmful defects, which
might cause failure in service, are present. Other types of NDT includes: (i) magnetic
analysis- by noting changes in the magnetic characteristics of the material from point to
point, (ii) radiographic tests- which make use of the radiation of short electromagnetic
waves such as X rays and gamma rays.

1.3.4 Material Test and structural model Test


A material test involves the determination of the properties of the material itself, whereas a
structural model test is done to study elastic behavior of structural elements. This later type
of test will essentially give similar behavior for all elastic material as opposed to the
destructive or non-destructive tests.

1.4 Standard Specification


A specification is the attempt on the part of the consumer to tell the producer what he
wants. To ensure proper quality control, a specification is intended to be a statement of a
standard of quality. A standard specification, prepared by standardizing agency, sets
standard material properties and also method of testing. Material testing should not be
arbitrarily performed. In order to make the test result representative, to enable comparison
of results and improve their reproducibility, it is imperative to use standardized procedure
of testing. Since material behavior often varies with the size and shape of the specimen, the
rate of loading and the temperature, standardization has been found necessary for all types
of tests. Based on research results and requirement of particular use of the material,
standard procedures for testing have been devised by the various standardizing agencies.
Different countries have their own standardizing agencies. Table 1.1 lists few standardizing
agencies.

Table 1.1: List of some Standardizing Agencies


Standardizing Agency Name of the Standard
ASTM-American Society for ASTM standard
Testing Materials
BSTI- Bangladesh Standard and Bangladesh Standard (BDS)
Testing Institution
BSI- British Standards Institution British Standard (BS)
ISI- Indian Standards Institution Indian Standard (IS)

For ease in making reference to a particular standard procedure each standard is uniquely
numbered which number is preceded by the appropriate name of the standard. For example,
ASTM A370, BDS 101 etc.

1.5 Types of Loading


Method of loading is the most common basis for designating or classifying mechanical
tests. This classification can be based on:

Structural Mechanics and Materials Testing Lab Page 2


(i) The kind of stress induced,
(ii) The-rate at which the load is applied, and
(iii) The number of times the load is applied.

According to the basis of the kind of stress induced, loading can be of the following types:
tension, compression, direct shear, torsion, and flexure.

With respect to the rate at which load is applied, loading may be classified into three
groups. If the load is applied over a relatively short time and yet slowly enough so that the
speed of testing can be considered to have a practically negligible effect on the results, the
loading is termed as static loading and the corresponding test is referred to as static test.
Such tests may be conducted over periods ranging from several minutes to several hours.
By far the majority of tests fall in this category. If the load is applied very rapidly so that
the effect of inertia and the time element are involved, the tests are called dynamic tests; in
the special case where the load is applied suddenly as by striking a blow, the load is
referred as impact load and the test is designated as impact test. If the load is sustained over
a long period, say months or even years, the test is a long-time test, of which creep tests are
a special type.

With respect to the number of times load is applied, tests may be classified into two groups.
In the first group, which includes the greatest number of all tests made, a single application
of load constitutes the test. In the second group, the test load is repeated many times,
millions if necessary; the most important category of tests in this group are the endurance
or fatigue tests, whose purpose is to determine the endurance or fatigue limit of a material
(of which the specimens are made) or of an actual part.

1.6 Report Writing


This section is intended to provide the student a guide-line for the preparation of the report
on the sessional on Strength of Materials. Each student is expected to submit a separate
report on the experiment performed, preferably on the same day of his performance.

1.6.1 Structure of the Report


A ready to use workbook on each experiment is provided for the students. The 'workbooks
are so designed that the students can independently carry out the experiments and fill-in the
relevant sections of the workbook. A typical format of a workbook is given here:

(i) Objective
(ii) Apparatus
(iii) Specimen
(iv) Procedure
(v) Recording and reporting of the results
(vi) Sample calculations
(vii) Graphs
(viii) Discussions
(ix) Assignments

Items 1 through 4 are descriptive and the students are required only to carefully read these
items. Item no. 5 is about recording the observations in a tabular format. It is not necessary
to show every calculation, but one calculation which is typical should be shown under the
heading Sample Calculation. Graphs should be plotted following the recorded data,
wherever applicable.

Structural Mechanics and Materials Testing Lab Page 3


1.6.2 Tips on Graph Plotting
Graphs are the largest single cause of re-submission for the sessional reports. This section
presents few tips for drawing graphs.

I. Labeling
(a) Provide a title for the graph.
(b) Label the axes in words and symbols. Do not forget to provide 'units'.
(c) Label curves to distinguish one from another or points of interest.
II. Plotting
(a) Theoretical curves should be plotted as continuous curves, but in such a way as
to distinguish them from the experimental curves. Use French or flexi-curves to
draw these; not in freehand.
(b) Experimental points should be clearly marked and ringed.
(c) Experimental curves should be a 'best fit' to the experimental points. They need
not pass through all points and must not be drawn freehand. Again, use French
or flexi-curves to draw these.
(d) If a plot is defined as linear, fit the best straight line to the experimental point.
Do not force this line to pass through the origin; mark the zero-error.
(e) If the slope of a line is required, don't simply divide the two absolute values of
one set of co-ordinates. Take a genuine tangent slope.
III. General
(a) Use the space available on a sheet of graph paper:
If slopes have to be measured, the greatest accuracy is achievable by having the
line as near as possible to 45°; this can be done by suitably choosing the axes of
the graph.
(b) Plot graphs so that they can be distinctly seen i.e. use a H.B pencil.

These features of graph plotting are illustrated in the example plots in Fig. 1.1

Structural Mechanics and Materials Testing Lab Page 4


250

200 Theoretical
Curve

150

Experimental
Curve
100

50

0
0 50 100 150 200 250 300 350

40

35

30

25

20 Slope

15

10

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4

Fig 1.1: Examples showing different features of graph plotting

Structural Mechanics and Materials Testing Lab Page 5


1.6.3 Tips on Writing Discussions on the Test Results
In the discussion section of your report you are expected to show your grasp over the test
performed. So it is advisable to think for a while before you write your discussion. Here are
few tips to guide you. Try to address the following questions:

(i) What does the test result signify?


(ii) Can the numerical results be applied directly to design and similar uses, or are they
of value only for comparison with other results?
(iii)Considering the methods of testing and the kind of material, what are the limitations
of the test results, or how reliable are the test data'?
(iv) Are the assumed conditions satisfied?
(v) What is most likely to go wrong or produce incorrect results?

1.7 Recommended Laboratory Practice


This book and the workbooks contained in it are designed for a laboratory course on
Mechanics of Materials. While the teacher of the course may adopt any suitable scheme for
the working on the workbooks by the students, the following scheme is recommended:

(i) Each student is issued a complete book - text plus workbook. The student retains
this and studies it before coming to the class.
(ii) During the class hour, the student fills up the relevant workbook, draws the
necessary graph(s).
(iii)The last 15 minutes of the student's work will be utilized for an objective test. The
objective questions will be supplied to the students. When completed, the student
should detach the workbook from the bound volume along the perforation and
submit the workbook along with the graph and the answer script.

Structural Mechanics and Materials Testing Lab Page 6


CHAPTER 2
STATIC TENSION AND COMPRESSION TESTS
2.1 Introduction
In tension and compression test, attempt is made to apply an axial load to a test specimen
so that uniform stress distribution can be ensured over the critical section. In such tests the
specimen is subjected to a gradually increasing (i.e. static) uniaxial load until failure
occurs. The static tension and compression tests are the most commonly made and are
among the simplest of all the mechanical tests. These tests provide almost all the
fundamental mechanical properties for use in design. The use of the tension as against the
compression test in all probability is largely determined by the type of service to which a
material is to be subjected. Metals, for example, generally exhibit relatively high tenacity
and are therefore better suited to and are more efficient for resisting tensile loads than
materials of relatively low tensile strength. For brittle materials such as mortar, concrete,
brick, and ceramic products, whose tensile strengths are low compared with their
compressive strengths, and which are principally employed to resist compressive forces,
the compression test is more significant and finds greater use.

2.2 Definition of the Terms


Fundamental to the understanding of the tests in this chapter and also in subsequent
chapters is a clear concept about the various terms used. This section, therefore, describes
the important terms used in connection with the physical testing of material.

Stress (σ) is defined as the intensity of the internal distributed forces or components of
forces that resist a change in the form of a body. Stress is measured in terms of force per
unit area, e.g. psi, kg/cm2, N/mm2 etc.

Strain is defined as the change per unit of length in a linear dimension of a body. It is a
ratio, or dimensionless number, and is therefore the same whether measured in inches per.
inch of length or centimeters per centimeter, etc. (Strictly, this definition of strain is limited
and is applicable to axial strain only. However, this definition is provided for easy
conceptualization by the students; a more general definition is that strain is the intensity of
the deformation whatever be the nature of deformation. See the next paragraph).

Deformation is the term used to indicate the change in the form of a body. Under axial
force it is usually taken to be a linear change (a shortening or a lengthening of a member)
and is measured in length units. In the case of torsion, it is customary to measure the
deformation as an angle of twist between two specified sections; whereas in the tension test
it is the elongation of the gage length that is measured as deformation. A shearing
deformation, on the other hand, is angular as shown in Fig.2.1. The shearing force F causes
the material to deform into the position shown by dotted line. The total deformation is δS,
and this deformation occurs over the length L. The deformation per unit of length is δS/L=
tanφ. For very small angle φ, tanφ ≈ φ. Thus the shearing unit strain is the angle φ (Greek
letter phi) measured in radians.

Structural Mechanics and Materials Testing Lab Page 7


F δS

φ L

Fig 2.1: Shearing deformation

Poisson's ratio is the ratio of unrestrained unit lateral contraction (or expansion) to the unit
longitudinal elongation (or contraction). For example, if a rubber band is stretched, its
width and thickness would decrease. This is called the Poisson effect.

The Poisson's ratio is generally denoted by μ (Greek letter mu). From Fig. 2.2, Poisson's
'
ratio,   b ; where δ′=|b-b'| and b is the original lateral dimension. Numerical values of

L
Poisson's ratio generally vary between 0.25 and 0.35.

P
δ
b
b
L
b'
b'

δ
P
Fig 2.2: Concept of Poisson’s ratio

Gage length is the length over which deformation measurement is made.


Elasticity is that property of a material by virtue of which deformations caused by stress
disappear upon the removal of stress. A perfectly elastic body is the one which completely
recovers its original shape and dimensions after release of stress. No materials are known
which are perfectly elastic throughout the entire range of stress up to rupture, although
some materials, such as steel, appear to be elastic over a considerable range of stress.

Plasticity is the property by virtue of which a material can undergo permanent deformation
without rupture.

Permanent Set or Set is defined as the amount of plastic deformation that remains even
after the removal of load (see Fig.:2.3)

Elastic Limit is defined as the maximum stress that a material is capable of developing
without a permanent set remaining upon complete release of stress.

Structural Mechanics and Materials Testing Lab Page 8


Proportional Limit (σpl) is defined as the maximum stress a material is capable of
developing without deviation from straight-line proportionality between stress and strain.

Total Strain

Loading

Stress
Unloading

Parallel
lines

Strain
Permanent Elastic
Set recover
Fig 2.3: Permanent set on unloading

Modulus of Elasticity (E) is defined as the ratio of stress to corresponding strain under
normal (axial) stress within the proportional limit. It is a measure of stiffness - the greater
the stress required to produce a given strain, the stiffer the material is said to be. For
material whose stress-strain relationship is not linear, the above definition of modulus of
elasticity is clearly inadequate. However, in such case the concept of initial tangent
stiffness (slope of the stress-strain curve at origin) or secant stiffness (slope of a line
connecting the origin and a point on the stress-strain curve at any specified stress level) or
tangent stiffness (slope of the tangent to the stress-strain curve at the average allowable
stress) is utilized.

The term modulus of rigidity (also termed as shearing modulus of elasticity) is the ratio of
stress to corresponding strain under shearing stress. Modulus of rigidity (G) and modulus
of elasticity (E) of a material is related by the following relationship:

E
G (2.1)
2(1   )
Where μ is the Poisson’s ratio.

Energy Capacity: The amount of energy that can be recovered on removal of stress form
the elastic limit is called elastic resilience. The energy stored per unit volume at the elastic
limit is the modulus of resilience and it is the area under the stress-strain diagram up to the
elastic limit (Fig.2.4). It is a measure of elastic energy strength of the material and is of
importance in the selection of materials for service where the stresses must be kept within
the elastic limit.

Structural Mechanics and Materials Testing Lab Page 9


σpl

Stress, σ
εpl Strain, ε

Fig. 2.4: Elastic Resilience or Modulus of Resilience.

Toughness is the energy required to rupture a material. It is measured by the amount of


work per unit volume of a material required to carry the material to failure under static
loading, called the modulus of toughness and represented by the area under the complete
stress-strain diagram.

Yield Point (σy) exists only for ductile materials such as mild steel. Yield point is defined
as the stress at which there occurs a marked increase in strain without increase in stress.
Only materials that exhibit this phenomenon have a yield point and the term yield point
should not be used in connection with a material whose stress-strain diagram above the
proportional limit is a line of gradual curvature with continually increasing stress.

Yield Strength (σy) is the practical and most commonly used measure of elastic strength. It
is the value of stress at which inelastic action begins.

Yielding is a very localized action and yielding becomes measurable only after many local
internal adjustments have occurred and after a considerable portion of the piece is affected
by yielding. It is, therefore, necessary to specify a certain measurable amount of yielding as
the beginning of yielding. This will be discussed in further details in section 2.8.
Stress

(b)
(c)

(a)
(a) Brittle material
(b) Material showing low ductility
(c) Ductile material

Strain
Fig 2.5: Typical stress-strain diagram for different types of material.

Structural Mechanics and Materials Testing Lab Page 10


Ductility is a special way of expressing plasticity (the general term to indicate the ability of
a material to deform in the inelastic or plastic range without rupture) of a material.
Ductility is that property of a material which enables it to be drawn-out to a considerable
extent before rupture and at the same time to sustain appreciable load. Mild steel is an
example of ductile materials. A non-ductile material is called brittle material which
fractures at relatively little or no elongation. Cast iron and concrete are examples of brittle
materials, which characteristically have low tensile strength (Fig.2.5).

2.3 Stress-Strain Diagram of Mild Steel


The relation between the stress and strain is commonly shown by means of a stress-strain
diagram, with the values of stress as ordinates and values of strain as abscissa. The usual
procedure in obtaining a stress-strain diagram is to plot data from a series of load readings
against corresponding data from the readings of a strainometer. Figure 2.6 shows one such
diagram which introduces the various terminologies defined in the previous section.

D
Stress

Upper YP
B E
A Fracture

C
Breaking strength
Lower YP

Ultimate strength

O
Strain
Fig 2.6: Stress-strain diagram of mild steel

The initial OA segment of the stress-strain curve is linear, indicating that stress and strain
are proportional up to point A, the proportional limit. Stresses not exceeding the
proportional limit results in practically elastic deformation only i.e. the strain that
disappears when the load is removed. For this reason the proportional limit is frequently
identified with the elastic limit. This is not correct although this should be accurate enough
for most practical purposes. Usually the elastic limit is defined as the stress at which the
permanent set is of the order of 0.005 per cent strain. The proportional limit and the elastic
limit are essential characteristics of a metal. For most structural steels the elastic limit has
nearly the same numerical value as the proportional limit. The initial straight line part
stress
indicates that stress is proportional to strain. It thus obeys Hook’s law i.e. = a
strain
constant, known as the Young’s modulus of elasticity.

If the load is increased beyond the elastic limit a point is reached where sudden extension
takes place with no increase in load. This is known as the 'yield point' (YP) and can be
detected easily by using a coating of brittle lacquer. In some materials the onset of plastic
deformation is denoted by a sudden drop in load indicating both an upper and a lower yield
point. In Fig.2.6, the upper YP is the highest stress before sudden extension occurs and its
value is affected by surface finish, shape of test piece and rate of loading. The lower YP,
which is normally measured in commercial testing, is the stress producing the large
elongation. The stress-strain diagram for this type of behavior is a discontinuous curve and
it is of great importance because it is characteristic of mild steel, one of the most common
structural materials. Other materials which sometimes yield discontinuously include

Structural Mechanics and Materials Testing Lab Page 11


molybdenum, cadmium, brass, aluminium and zinc. The reason of such discontinuous
yielding is attributed to the presence of small amount of impurities e.g. foreign atoms in the
material, usually carbon and nitrogen (as little as 0.003 percent nitrogen or carbon is
sufficient for this effect to take place). The foreign atoms, in solution, occupy more or less
unstable interstitial positions in the lattice. These foreign atoms naturally tend to migrate to
vacancies in the lattice or to regions of higher energy such as dislocations. Many of them
find stable positions in the space just below the extra planes of atoms in edge dislocations.
Therefore, they line up in these spaces and effectively anchor the dislocations against
movement. Such a line of foreign atoms is called an atmosphere (Fig.2.8). When a shear
stress is applied on a slip plane in mild steel, it must pull the dislocations away from their
atmospheres before slip can take place. Figure 2.9 illustrates the variation of stress with the
displacement, where the stress is largest when the dislocation has been displaced ro/√3,
where ro is a constant of the order of one atomic spacing, τ is the applied stress and x is the
displacement. As the dislocation moves further away, it does so at a decreased stress. Thus
a high stress is required to initiate yielding but a low stress will continue it and we have the
upper and lower yield points. After this point, the specimen stretches plastically and the
stress to produce continued plastic deformation increases with increasing strain i.e. the
metal strain-hardens. This strengthening mechanism is related to an increase in
deformation. The increase in external stress is due to the interaction of the moving
dislocation with existing dislocations.

Extra plane of atoms

Atmosphere of foreign atoms

Fig 2.7: Cutaway view of dislocation, showing row, or atmosphere, of foreign atoms

τ
r0
3

x
Fig 2.8: Variation of stress on an anchored dislocation

As the specimen elongates, it decreases uniformly along the gage length in cross-sectional
area. Initially, the strain-hardening more than compensates for this decrease in area and the
engineering stress continues to rise with increasing strain. Eventually a point is reached
where the decrease in specimen cross-sectional area is greater than the increase in

Structural Mechanics and Materials Testing Lab Page 12


deformation load arising from strain hardening. This condition will be reached first at some
point in the specimen that is slightly weaker than the rest. All further plastic deformation is
concentrated in this range and the specimen begins to neck or thin down locally (Fig.2.9).
Because the cross-sectional area now is decreasing far more rapidly than the deformation
load is increased by strain-hardening, the actual load required to, deform the specimen falls
off and the engineering stress continues to decrease until fracture occurs.

Undeformed bar

Necking begins with a reduction in cross-sectional area

Fig 2.9: Necking of mild steel bar in tension

2.4 Measurement of Ductility


When a tension test is performed, ductility is measured by Percentage elongation (note
that percentage elongation depends on the gage length used) and/or percentage reduction
in the x-sectional area. The percentage elongation is usually based on a 2 in. gage length,
but other lengths are also used. Since the largest part of the plastic strain is localized in the
neck with relatively little in other parts of the specimen, the percentage elongation varies
with the length used. This is illustrated with the following example. A steel bar is tested to
failure. Segment A of the bar, located at the neck, was originally 1/2 in. long but is 1 in.
long after fracture -- i.e. a 100% elongation. The other parts of the bar are elongated
uniformly by 10 percent (say). This means that each inch becomes 0.1 in. longer after
fracture. In a 2 in. length including the neck, the total elongation is 0.5 in for segment A,
plus 0.15 in. for the one and one-half inches (1.5*0.1).The percentage elongation in 2 in.
is, therefore, (0.5+0.15)*100/2 or 32.5 percent. On the other hand, if an 8 in gage length is
used (which also includes segment A), the elongation is 0.5 in plus 7.5*0.1 in., and the
percentage elongation in 8 in. amounts to be (0.5+0.75)* 100/8 or 15.6 percent. This later
percentage elongation based on an 8 in gage length is less than half as much as that in 2 in.
gage length. This example demonstrates that the importance of specifying the length when
reporting the percentage elongation. It may, however, be noted that percentage elongation
is used only to compare ductility’s of various material and it has no other significance.

The percentage reduction in area is found by measuring the cross-sectional area of the
member before loading (Ao) and that of the neck after fracture (A). The differences
between these two areas expressed as percentage of the original area, is the reduction of
area.

Ao  A
Percentage reduction in area =  100 (2.2)
Ao

For specimens of the same size and shape this quantity provides a good ductility rating,
since this is not subjected to any ambiguity as in percentage elongation.

2.5 Engineering Stress and True Stress


It must be emphasized at this stage that the stress that is normally reported is the
engineering stress (also known as conventional stress or nominal stress) obtained by
dividing the applied force by original (undeformed) cross-sectional area of the specimen.
The true stress on the other hand, is obtained as the ration of the applied load to the

Structural Mechanics and Materials Testing Lab Page 13


corresponding reduced area (instantaneous area) of the cross-section of the specimen. In
the elastic range the reduction of area is very small (for steel, with Poisson’s ratio = 0.3 and
50,000 psi the reduction is only 0.1%) and original area of cross-section may be used
without losing any significant accuracy.

Similarly engineering (also known as conventional stress or nominal) strain is the term
given when original length of the specimen is used for computing strain. The quantities
engineering stress and strain are used throughout the usual courses in the mechanics of
materials.
In the plastic range, however, it must be recognized that both the reduction in area and the
strain can become very large and the use of instantaneous area and instantaneous gage
length becomes more important in calculating the stress and strain. These are called true
stress and true strain (also called natural strain) and they are the proper measures of the
mechanical behavior of materials in all ranges, even in the elastic range. Figure 2.10
compares true stress with conventional stress, both plotted against conventional strain. The
monotonic increase of true stress up to the breaking point illustrates that strain-hardening
continues throughout the plastic range.
True stress

Stress

Conventional stress

Strain
Fig 2.10: Comparison of true stress and conventional stress

2.6 Speed of Loading


The speed of loading has a definite effect on the strength either in tension or in
compression. Over the range of speed used with ordinary testing machines, the effects of
variation of speed are fairly small. In general, strengths tend to increase and ductility to
decrease with increased speeds. For example, in tension tests of standard specimens of
structural steel it was found that an eight fold increase in rate of strain increased the yield
point by about 4 percent, the tensile strength by about 2 percent and decreased the
elongation by about 5 percent.

In compression tests on concrete the relation between strength and rate of loading is
approximately logarithmic the more rapid the rate the higher the indicated strength. The
strength of a specimen loaded say, at 6000 psi per min. would be about 15 percent greater
than the strength of a specimen loaded at 100 psi per min.

Structural Mechanics and Materials Testing Lab Page 14


2.7 Yield Strength by Offset Method
Yielding in a material usually begins very gradually; the very first plastic strain is
extremely minute and cannot actually be measured by any known method. Most materials,
unlike mild steel (Workbook No.1); do not have a very well defined yield point.
Consequently it is necessary to specify a certain measurable amount of yielding as the
beginning of yielding. The actual value of the yield strength, then, will depend upon this
arbitrarily selected amount of yielding. A value of yield strength given without this
qualifying amount is ambiguous and can be used only as a rough approximation.

Figure 2.11 (a) shows a Hypothetical stress-strain diagram for a material loaded to a stress
(YS) somewhat above the proportional limit (PL) and then unloaded. The distance CA
represents a deviation or offset from Hook's law at the stress (YS). The set after release of
load is indicated as the strain 'a' on the diagram. For most materials the unloading path AB
has the same slope as the elastic loading path OC. That is, the line AB (Fig.2.11 (a)) is
parallel to the line OC and the offset CA approximates the permanent set 'a'. The offset thus
approximates the inelastic deformation at a given stress. This concept is the basis for the
determination of yield strength by the offset method, in accordance with the definition of
yield strength given by the ASTM, namely the stress at which a material exhibits a
specified limiting permanent set (ASTM E6).

In figure 2.11 (b) OX represents a portion of the stress-strain diagram for a material that
does not exhibit a marked yield at any particular stress but yields gradually after the
proportional limit is exceeded. At a specified permanent set 'a' point B is marked on the
strain axis. A line BA is drawn parallel to the initial straight portion of the stress-strain
diagram to intersect the curve at A, thus determining the yield strength (YS) as defined by
the offset method.

Stress Stress
YS C A YS C A

PL PL

B B
O O Strain
a (a) Strain a (b)

Fig 2.11: Determination of yield strength by offset method

The magnitude of the offset 'a' is chosen as some measurable value that is considered from
experience to be of practical significance in defining a limit to the elastic range. Values of
'a' in common use are shown in Table 2.1. In reporting values of yield strength obtained by
the offset method, the specified value of set used should be stated. For example, yield
strength (offset= 0.1 percent)= 52,000 psi indicates that at a stress of 52,000 psi the
permanent set of the material tested was 0.1 percent strain.

Structural Mechanics and Materials Testing Lab Page 15


Table 2.1: Recommended values of offset for yield strength.
Material Offset strain (%)
Cast Iron 0.02 to 0.05
Wood (parallel to the grain) 0.05
Concrete 0.01 to 0.02

2.8 Proof Strength


The offset method described in the preceding section is less widely used in the routine
commercial tests for ascertaining the yield strength. This is because the elaborate procedure
of the offset method is more suitable to an academic environment rather than to every day
routine tests situation. In commercial practice a nomenclature "proof strength" is used
which is synonymous to the term yield strength for materials not exhibiting the actual yield
phenomenon. Proof strength is more often used in the British specification whilst yield
strength is specified in the North American practice. For certain materials like prestressing
wires and strands, American Codes specify stress values at 1% elongation which is
equivalent to proof strength at 1% total strain as per British terminology.

Determination of proof strength differs from that of offset method in a way that for the
latter case a fixed value of permanent set is specified while for the former the total strain
amount is specified. Figure 2.12 illustrates the concept of determining proof strength for
specified total strain value of 'b'.

Stress

Proof
strength

b=specified total strain

O Strain
b
Fig 2.12: Determination of proof strength

From Fig 2.12, it is quite clear that one does not need to draw the complete stress-strain
diagram for obtaining the proof strength. Generally, standard codes of practice specify the
amount of total strain to be used for different materials. For example, the specification for
cold twisted ribbed bar requires the tensile proof strength to be measured at the stress
corresponding to a 0.5% strain in 200 mm gage length. This means that when such a bar is
tested in tension the elongation measurement is made over the 200 mm gage length as the
loading progresses. When the gage marks reach to 201 mm (i.e. 1mm elongation in 200
mm) the corresponding load is read from the machine dial as the proof load. Dividing proof
load by the cross-sectional area of the specimen gives the proof strength. This method has
found wide use in the industry because of its simplicity. Reporting of proof strength should
be accompanied by the amount of total strain on which it is based. For example, 0.5% (total
strain) proof strength = 50,000 psi.

Structural Mechanics and Materials Testing Lab Page 16


2.9 The Standard Tension Test
The tension test is the most widely used experiment for the study of mechanical behavior
of a material. It is often called the basic mechanical test. In the idealized state of static
tension it is assumed that (a) the load is applied axially, (b) the stress is constant at cross-
sections along the length of the member, and (c) the load changes so slowly that no
dynamic effects are present. In the standard tension test these conditions are approached as
closely as possible within the practical limitations of the equipment available. A standard
specimen is loaded slowly with an axial load, and a series of observations of load and
deformation are made.

Deformation may be axial elongation or lateral contraction (or both). In the tension test for
conventional stress and strain, it is necessary only to measures the elongation with an
extensometer. The true stress-strain tension test, on the other hand, requires the lateral
contraction to be measured, in addition to the measurement of the elongation. This makes
the conventional stress-strain test of widespread use. In fact all standardized tension tests
arc for conventional stress and strain and this is why the procedure for this test only will be
described here.

2.9.1 Tests for Conventional Stress and Strain


The test methods for obtaining information about stress and strain differ from material to
material. The ASTM Standard E8 covers such test for metallic materials. The method used
for metallic materials, however, applies also to others, and consequently this method will
be discussed in details.

Specimens: Specimens are almost always cylindrical or prismatic, with substantially


constant cross-sectional area for uniformity of stress. The ends arc usually enlarged for
added strength so that rupture will not take place near the grips, where the stress
distribution is complicated. The experimental measurements are all made on the central
portion (i.e. the reduced section). In cylindrical specimens this portion is commonly ½ inch
in diameter and 21/4 inches long. The ends are 3/4 inch in diameter and are threaded to
screw into specimen holders on the testing machine. Since abrupt changes in cross-section
cause stress concentrations, the transition from the central portion to the larger ends must
be made by fillet of large radius (Fig.2.l3). The 21/4 inch central portion allows .the use of
a 2-inch gage length.

13 " 2 1 " reduced section 13 "
8 4 8
3/4"

0.505˝ diameter
3/4"-10 no. thread
1/2" radius

Fig 2.13: Tension test specimen

It is often tapered very slightly (0.003-0.005 inch) toward the centre to help ensure rupture
near the centre. Rectangular specimens are used for plate or sheet materials. Their
proportions are similar to those of the cylindrical type. The ends are gripped by jaws in the
testing machine. A number of other types and sizes are in common use for various
purposes. Details are given in the appropriate ASTM Standards.

Structural Mechanics and Materials Testing Lab Page 17


Loading: Before the specimen is placed in the testing machine its dimensions must be
measured with care. The diameter of a ½ inch specimen is usually measured to the nearest
0.001 inch (for more detailed instructions see ASTM E8). The gage length is fixed by the
extensometer, the gage points or knife edges of which are at the proper spacing. The range
of most extensometers is much less than the total extension of a ductile specimen, and
consequently some other means must be used to measure the elongation far into the plastic
range. A pair of gage marks is usually punched on the surface of the specimen so that
changes in their spacing can be measured with dividers for large elongation. These gage
marks are also necessary if the final elongation is to be obtained. The specimen is placed
in the testing machine by using the appropriate grips care being taken to align it carefully
with respect to the axis of loading.

After the specimen is properly aligned, a small initial load is usually applied to hold it
firmly while the extensometer is being attached. The type of extensometer used depends
on the gage length of the specimen, the accuracy and range desired, and whether the
strains are to be recorded autographically. As mentioned before, extensometers have a
limited range.

The load is increased at a rate slow enough so that dynamic effects may be assumed to be
negligible, and the specimen can always be considered to be in a state of equilibrium. This
is not always possible. For certain materials the rate of increase to guarantee these
conditions would be much too slow to be practical. Some materials continue to flow for
long periods of time after the load has been increased and held at a certain value. To wait
for equilibrium conditions would require almost unlimited time. Consequently a certain
specified rate must be maintained at all times for these materials to ensure comparable
results. The ASTM standard test methods give rates for various materials. This rate may be
expressed in terms of increase in load (dP/dt), stress (dσ/dt), or strain (dε/dt), or in terms of
cross-head speed.

For metals the rate of strain should be held steady in the plastic range. This can usually be
done by keeping the testing machine controls at approximately the same setting throughout
this part of the test. Testing machines are sometimes equipped with electronic controls
which automatically hold one of the various rates constant.

If a stress-strain diagram is to be plotted, the load and extensometer readings must be


obtained at regular intervals. Intervals of load or extensometer reading are chosen before
the start of the test to provide the desired number of readings. At each interval of the
chosen variable the corresponding value of the other variable is recorded. As soon as it
nears the end of its operating range the extensometers should be removed (before this if
there is any reason to believe that the specimen might rupture prematurely). Observations
of strain can then be continued with the dividers and gage points, using a magnifying glass
if necessary.

Particular loads, such as yield point load (if any), ultimate (maximum) load, and breaking
load, are recorded during the progress of the experiment. If it is carried all the way to
rupture, the final elongation and cross-sectional area can be measured.

The character of the fracture is often a· revealing piece of information and should be
described. The extensometer readings must be converted to strain by multiplying the gage
factor and dividing by the gage length.

Structural Mechanics and Materials Testing Lab Page 18


2.9.2 Construction of the Stress-Strain Diagram
Variation of stress and strain in the tension test can best be presented in the form of a
stress-strain diagram. The conventional diagram is plotted from conventional stresses and
strains computed from the loads and extensometer readings. Sometimes the load is plotted
directly, without converting to stress, since the only difference between the two is a
constant factor the original area. The same applies to the extensometer reading, which is
sometimes plotted directly, without converting it to strain.

It is common practice to plot the stress on the vertical axis and the strain on the horizontal
axis to give curves of the type shown in various illustrations throughout the text. The
construction of the diagram involves selecting the proper type of graph paper, choice of
scale, location of the axes on the sheet, plotting points, drawing the curve, and indicating
pertinent facts about the experiment.

The choice of scale should be carefully made. There are two considerations in choosing
scale: the size of the resulting graph, and the ease of interpolation. It is usually desirable to
make the graph as large as possible, allowing for suitable margins on all sides. It is also
important that the values of both variables be readable at any point on the curve - both at
and between the plotted points. For this reason scale chosen should involve factors like
2,5,10, or sometimes 4. The following are examples of good scale: I major division = 0.001
in. /in.; 5 major divisions = 1000 psi. Scales involving the factor 3 should be avoided,
because this causes interpolation of values very inconvenient. (Examples of bad scales are
1 major division = 3000 psi; or 3 major divisions = 0.001 in/in.).

Another important consideration is the resultant size of the elastic range on the graph. If the
material is steel and has a yield point of 30,000 psi, the elastic range covers only 0.001 in.
/in. strain. Use of the scales above would confine the elastic range to one-tenth of a major
division, too small a size if any information is desired from the elastic range. A separate
graph would have to be drawn to show only the elastic range at a larger scale. A possible
choice of scales would be 1 major division = 0.0002. This larger-scale curve could be
drawn on the same sheet with the overall curve and the same stress scale used for both.
This larger-scale curve should be used for determining the elastic properties like modulus
of elasticity and modulus of resilience.

When the scales have been chosen and the extent of the graph is known, the axes should be
drawn so that the graph will be approximately centered on the sheet. The top of the graph
should always be either at the top of the sheet or next to the binding holes, so that it can be
read from either the lower or the right-hand side of the sheet. Each axis should be drawn in
pencil on one of the major division lines (thick lines on the graph paper). The scales are
shown alongside the axes. Examples are: stress, psi (or ksi); Strain, in. /in. Notations such
as psi x 103 or in. /in. x 10-3 should not be used, as they are apt to be confusing. The
intersection of the two axes should represent the actual origin of co-ordinates so that each
scale is shown completely, from zero to the maximum value.

In plotting the stress-strain diagram, follow the average trend of the experimental points.
The points representing the elastic range in mild steel, for example, should be average by a
straight line, while those in the yielding region may sometimes be followed more closely,
using an irregular curve connecting most of the points. Generally, points near or at the
origin are least reliable. The effects of friction and lost motion in the instruments are most
pronounced in the low ranges, and these reading are thus subject to greater errors than
those at higher loads. As a result the fitted curve often does not pass through the origin of
coordinates even though the first point may have been observed as zero stress and zero
strain. Provided the load indicator was properly zeroed before the test, it can be assumed

Structural Mechanics and Materials Testing Lab Page 19


that the zero axis for stress (the horizontal axis) is correct. The true origin for the curve is
therefore at the point where the curve crosses the horizontal axis. For computations from
the graph, all strains should be measured from this point. It is not necessary to change the
graph but merely to apply a correction equal to the strain reading at the true origin to all
other strain readings.

These principles and rules for constructing the stress-strain diagram are demonstrated in the
example that follows. When the diagram is complete, various properties may be computed
from it and pertinent facts identified.

2.9.3 Example: Stress-Strain Diagram


For materials whose stress-strain diagram does not show a pronounced yield point there are
two common ways of determining yield strength from the stress-strain curve. These are
explained with the help of Fig. 2.14.

(i) The Offset Method


The offset is the horizontal distance between the modulus line and any line running parallel
to it. The value of the offset for a given material is usually expressed this way: Yield
Strength, 0.1 % or 0.2% Offset. "0.2% Offset" means 0.2% of the fundamental extension
units of inches per inch, or 0.002 in. /in. Starting at the origin of the curve, measure off a
distance equal to 0.002 in. /in. along the X-axis. Now using that as the origin, draw a line
(C-D) parallel to the modulus line. Note that the line C-D intersects the stress-strain curve
at a certain point (Y in Fig.2.14). The ordinate of this point (the amount of stress in psi) is
the yield strength at 0.2% Offset.

(ii) The Proof Strength method


This method involves drawing an ordinate line (that is, a completely vertical line) from the
point on the X-axis where the elongation equals the specified extension, e.g.: Yield
Strength=0.5% Extension. To make this determination, locate the point on the abscissa
which is equal to 0.005 in. /in. of extension from the origin of the curve (E in Fig.2.14).
Draw an ordinate line E-F from this point up through the curve. The stress value on the
curve is the Proof Strength or Yield Strength at 0.5% Extension under Load.

Young's Modulus of Elasticity


The modulus of elasticity (Young's modulus) is the ratio of stress in pounds per square
inch to strain in inches per inch as computed from the modulus line.

Stress ( psi )
Young’s Modulus (psi) = (2.3)
Strain(in. / in.)

To find modulus, take any two points (K and L) on the modulus line (A-B), and divide the
differential between their stress values in psi form the strain differential in in. /in. The
result of this division is the modulus of the material tested.
For example:

Stress (lbs/sq. in) Strain (in. /in.)


Point K 87,000 0.003
Point L 29,000 0.001
Difference (∆) 58,000 0.002

Structural Mechanics and Materials Testing Lab Page 20


58000 psi
The modulus of elasticity =  29,000,000 psi
0.002in. / in.

Calculation of Tensile Properties of Materials:


The calculation of tensile properties from the stress-strain diagram shown in Fig. 2.14 is
outlined below:
Line A-B is Modulus Line:
(Young's) Modulus of Elasticity = slope of initial straight line portion of curve expressed
Stress( psi )
as ratio of
Strain(in. / in.)
Line C-D is 0.2% (0.002 in. /in.) Offset Line.
Line E-F is 0.5% (0.005 in. /in.) Extension Line.
Curve A-R is complete Stress-Strain Curve to specimen failure.
Point V = Proportional Limit.
Point X = Yield Strength, 0.1 % Offset.
Point Y = Yield Strength, 0.2% Offset.
Point Z = Yield Strength, 0.5% Extension under Load.

Material which produces a "sharp kneed" stress-strain curve, yield points are easily
identified as P1, the upper yield point and P2, the lower yield point (Fig.2.15).

Structural Mechanics and Materials Testing Lab Page 21


Structural Mechanics and Materials Testing Lab Page 22
CHAPTER 3
SHEAR TEST OF METAL SPECIMEN
3.1 Introduction
A shearing stress is one that acts parallel to a plane as distinguished from the tensile and
compressive stresses that act normal to a plane. It resists the tendency of the part of the
body on one side of the plane to slide relative to the part of the body on the other side of
the same plane. Following types of loading produce shear in material:

(i) The resultants of parallel but opposed forces act through the centroid of sections
that are spaced infinitesimal distances apart. In such cases the shearing stresses
over the sections should be uniform and the state of pure direct shear would exist.
This condition may be approached but is never realized practically. An
approximation of this condition is the case of a rivet in shear as shown in (Fig
3.1(a)); here, for practical purposes, direct shear may be considered to exist within
the rivet on planes xx and yy.

(ii) The applied opposed forces are parallel, act normal to a longitudinal axis of the
body, but are spaced finite distances apart. In such case, in addition to the shearing
stresses produced, bending stresses are also set up. In the case of a rectangular beam
subjected to transverse loads [Fig 3.1(b)], the intensity of shearing stresses on any
cross-sections vary parabolically from zero at the upper and lower surfaces of the
beam to a maximum at the neutral axis.

(iii) The applied forces are parallel and opposite but do not lie in a plane containing the
longitudinal axis of the body; here a couple is set up which produces a twist about
a longitudinal axis. This twisting action of one section of a body with respect a
contiguous section is termed as torsion. Fig 3.1(c) represents a piece of circular
rod subjected to a torque. Torsional shearing stresses on circular cross-sections
vary linearly from zero at axis of twist to a maximum at the extreme fibers. If no
bending is present, 'pure shear' exists.

3.2 Shear Stress Distribution in Beams


The shear force and the bending moment acting on a cross-section of a beam are closely
interrelated. Let us consider an element of a beam of length dx (Fig. 3.2 The bending
moment increases over the length dx from M to M+dM, and the direct stresses increases
proportionately.. Let us now cut the beam horizontally at a distance y above the neutral
axis, and consider the equilibrium of the element of width b (at right angles to the sketch),
of length dx, and of depth [(d/2)-y] (Fig.3.3). Since the compressive stresses due to the
bending moment M+dM are greater than those due to the moment M, there is a resultant
force tending to push the element to the left. This is resisted by the horizontal shear force
V, acting along the cut which we made at a distance y above the neutral axis. There can be
no shear at the top face of the beam, because at a free horizontal surface there is nothing to
offer resistance to a horizontal shear force. Since the direct stresses increase over the length
dx (fig 3.2), a shear force V is required to balance the forces acting on the element of the
beam dx long and b wide. The shear stress acting on this area, υ, increases parabolically
from zero at the free surface to a maximum at the neutral axis (Fig 3.3).

Structural Mechanics and Materials Testing Lab Page 23


P/2 x x
P
y y
P/2

(a) Direct shear in a rivet

CL

Shear diagram

(b) Shear in a homogeneous beam of rectangular section.

e
Torque=Pe P

(c) Shear produced by torsional loading

Fig 3.1: Loading producing shear

Structural Mechanics and Materials Testing Lab Page 24


Fig 3.2: Stress distribution due to bending on a infinitesimal section of a beam

d
y
2 3V
Vmax=
2bd
V=vbdx

Fig 3.3: Horizontal shear force and stress distribution in a beam.

V
v
bW bw d
d

Fig 3.4: Shear stress distribution in I-sections.

For an I-section, the shear stress is invariably proportional to the width, b (Fig 3.4). Since
the width changes abruptly in an I-section at the junction of the web and the flange, the
shear stress changes in the proportion of b to bw. Since the flange width, b is much greater
than the thickness of the web, bw, the shear force is almost entirely resisted by the web.

The shear force V = υbdx; where v is the shear stress, b is the width of the section, and dx
the length of the element. The shear stress builds up with depth from the free surface, as the
difference between the direct compressive forces increases. Since this shear stress is the
sum (or integral) of the difference in compressive stress over the depth (which is a linear or
first order function), the shear stress varies parabolically with depth. The total shear force
V equals the area of this diagram, and a parabola has an area two-thirds that of the
enclosing rectangle. In a rectangular section of width b and depth d, the maximum shear
stress (which occurs at the neutral axis) is therefore,

V
v max  (3.1)
2 3 bd

A rigorous proof of this formula is given in many textbooks on the strength of materials is
therefore is not included here. In the case of an I-beam, there is a sharp change in shear
stress at the junction of the flanges with the web, since v is inversely proportional to b.
The shear stress distribution diagrams thus almost rectangular (Fig. 3.4), and the
maximum shear stress in a steel I-section is approximately

V
v (3.2)
bw d

where, bw is the width of the web.

Structural Mechanics and Materials Testing Lab Page 25


Evidently an I-section divides itself into the flanges, which resist practically the whole of
the bending moment, and the web, which resists practically the whole of the shear.
However, it should be noted that in most practical cases the shear stress is computed by
dividing the shear force acting on the cross-section by the area of the cross-section. This
represents the average shear stress in the cross-section and is satisfactory for practical
purposes.

3.3 Torsion
Torsion is a common consideration in the design of engines; for example, in the shafts of
rotary motors (Fig.3.5). It causes one part of the shaft to shear relative to the other, and
torsion is thus a form of shear (see Fig.3.6). If beams are subjected to torsion, then the
torsional shear stresses are additive to the shear stresses resulting from the shear force
caused by bending. The solution of torsion problems is simple for circular sections, but
complicated for all other. Since circular sections are rarely used in buildings, torsion is
discussed only in advanced structural texts.

The distribution of the shearing strain due to torsional moment (or torque) in circular
sections are assumed to vary linearly from the centre of the section (see Fig.3.7). This
assumption backed by the following two concepts is used to derive the torsion formula:

(i) Equilibrium of the internal or resisting torque and external torque.


(ii) Material properties in the form of Hook's law are used to relate the assumed strain
variation to stress.

T
T

Fig 3.5: A circular shaft subjected to torsion

Fig 3.6: Failure of a concrete beam in diagonal tension due to torsion

c τ
τmax

ρ
T

Fig 3.7: Torsional shear stress in a circular cross-section under an applied torque T.

Structural Mechanics and Materials Testing Lab Page 26


These assumptions eventually lead to the torsional formula for circular section as follows:

T
 (3.3)
J

where τ is the shearing stress at a radial distance ρ from the centre of the shaft. T = applied
torque (i.e. torsional moment) J = polar moment of inertia of the section and for circular
r 4 d 4
section, J   ; with r (radius of the section) or d (diameter of the section) in
2 32
inch, J will have a unit of inch4. Thus with T in lb-in, ρ in inch and J in inch4, from
equation 3.3, it follows that τ will have a unit of psi. Evidently from fig 3.7 τmax will occur
at ρ = r; i.e.
Tr
 max  (3.4)
J
Note that this formula is valid only for solid or hollow circular section by using the
appropriate value of J.

In the following civil engineering structures, torsional shear stress is an important design
consideration: (i) spandrel beams loaded eccentrically [Fig.3.8 (a)] (ii) balcony girders
[Fig.3.8 (b)], (iii) spandrels of a building without corner columns [Fig.3.8(c)], and (iv)
spiral stair slabs.

(a) (b)

(c)

Fig 3.8 (a) Torsion in spandrel beam caused by eccentric loading of outer wall.
(b) Balcony girder. A beam projecting beyond its support—e.g., a balcony beam-is
subjected to combined bending and torsion. (c) When two spandrels meet at the corner
of a building without a supporting column, both are subjected to torsion.

3.4 Behavior of a Closely Coiled Helical Spring


Helical springs made of rods or wires of circular cross section may be analyzed in the
elastic range by superposition of the shearing stresses. One important assumption need to
be made here is that anyone coil of such a spring will be assumed to lie in a plane which is
nearly perpendicular to the axis of the spring. This assumption can be made if the adjoining
coils are close enough. With this assumption, a section taken perpendicular to the spring's
rod may be taken to be vertical. Therefore, the forces acting on any section of the spring

Structural Mechanics and Materials Testing Lab Page 27


are (Fig.3.9) shearing force (V=P) and torque (T=PR) where, P=axial force applied on the
spring and R= distance of the axis of the spring to the centroid of the rod's cross section.

Fig 3.9: Forces acting on a closely coiled helical spring

The maximum shearing stress at an arbitrary section of the spring can be found by
superposing the direct and torsional shear stresses. Direct shearing stress in spring is
customarily taken as the average shearing stress uniformly distributed over the cross
section. Torsional shearing stress having a linear stress distribution as discussed in the
preceding section will produce the maximum shear stress in the same direction as that of
the direct· shear stress at the inner edge of the coil at point E (Fig.3.9 )

Therefore, by superposing one obtains,

P Tr P PRr P  2 R 
 max     4  1   (3.5)
A J A r A r 
2
2
where, A=πr =area of the spring’s rod.

Structural Mechanics and Materials Testing Lab Page 28


The deflection of a helical spring can be obtained (neglecting the effect of direct shearing
stress, which is normally small) by using the following relationship:

PR 2 L
 (3.6)
GJ

where, L = length of the spring’s rod, and G = shearing modulus of elasticity (also known
as modulus of rigidity). For a rigorous derivation of the above relationship reference can be
made to any text of Strength of Material, e.g. E.P.Popov's 'Introduction to Mechanics of
Solids'; Chapter 8.

For a closely coiled spring the length L of the wire may be taken with sufficient accuracy
as 2πRN, where N is the number of live or active coils of the spring.

2PR 3 N 64 PR 3 N
Therefore,   (3.7)
JG Gd 4

This equation can be used to obtain the deflection of a closely coiled helical spring along
its axis when such aspring is subjected to either tensile or compressive force P.

The stiffness of a spring, commonly referred to as spring constant, is frequently used to


define the spring's behavior. The spring constant is defined as the force required to produce
unit deflection. Thus, spring constant, k, defined as

P Gd 4
k  (3.8)
 64 R 3 N

is a useful quantity. The expression of k in equation 3.8 is subject to the limitation arising
due to neglect of the deformation due to the direct shearing stresses.

3.5 Shear Tests


Two types of shear test are in common use; they are:

(i) The direct shear test


(ii) The torsion test

The torsion test enables a more precise determination of shearing properties (proportional
limit, yield strength in shear, shearing resilience, and stiffness i.e. modulus of rigidity or
modulus of elasticity in shear) by measuring the angle of twist. Helical springs tested in
axial compression or tension constitute one type of shear test, since the stresses developed
are largely those of torsional and direct shear but principally the former (note that helical
spring test is an indirect form of torsion test and direct torsion test of metals is carried out
in a special testing machine designed for the purpose). The modulus of rigidity of the
material of the helical spring can be determined from the applied loads and the deflections
and using the relationship of equation 3.8. Note that such determination of the modulus of
rigidity is subjected to the same limitation as equation 3.8 above.

The direct shear test (also called transverse shear test) gives an approximation to the correct
values of shearing strength. This test is usually done in a Johnson type of shear tool by
clamping a portion of the material so that bending stresses are minimized across the plane
along which the shearing load is applied. Because of inevitable bending and friction

Structural Mechanics and Materials Testing Lab Page 29


between parts of the tool it gives an indication of the shearing resistance of the material.
The transverse-shear test has the further limitation of being useless for the determination of
the elastic strength or of the modulus of rigidity, owing to the impossibility of measuring
strains.

In direct shear test, the shearing stress is considered to be uniformly distributed over a
cross- section. The two different cases may arise:

For single shear [see Fig.3.l0 (a)] the shearing stress,

P
SS  (3.9)
A

and for double shear [see Fig 3.10(b)]

P (3.10)
SD 
2A

P
P

(a) Single shear

P/2
P
P/2

(b) Double shear

Fig 3.10: Rivets under shear

Structural Mechanics and Materials Testing Lab Page 30


CHAPTER 4
HARDNESS TEST OF METAL
4.1 Introduction
The term 'hardness' refers to the resistance of a metal to permanent deformation to its
surface. This deformation may be in the form of scratching, mechanical wears, indentation
or cutting. Depending on the particular deformation type (or for that matter particular type
of stressing) hardness may be of the following types:
(i) Indentation hardness test (by indenting)
(ii) Scratch hardness test (by scratching)
(iii) Dynamic hardness test (by impact)
(iv) Rebound hardness test (by the rebound of a falling ball)
Hardness has correlation with other mechanical properties of the material; thus hardness
tests have wide application as a non-destructive test of metals.

4.2 Hardness Measurement


One way in which a surface may be deformed is by indentation, in which a permanent
deformation is produced by pressing an indentor of some kind into the surface of the
material. The depth of penetration and the force required are measured which provides an
indication of hardness (indentation hardness). This hardness is nothing but the resistance to
permanent deformation. Indentation hardness is the most commonly used hardness test and
different forms of this hardness test are in use. A general description of some of the
common form of indentation hardness tests are given here:

4.2.1 Pyramid Hardness


The most versatile hardness measurement makes use of diamond points, ground in the
shape of a pyramid, as indentor. The indentor is forced into the surface, leaving an
impression, the size of which is measured by the length of the diagonal, ‘d’. The hardness
number (H) is then defined as the ratio of the load (P) to the area of the impression (A):

P
H (4.1)
A

It has been found experimentally that for pyramid indentors, the load varies as the square of
the diagonal, ‘d’. Therefore, P= λd2; where λ is a constant and A= βd2; depending on the
material and shape of pyramid. Constant ‘β’ depends on the shape of the pyramid.
Substituting these expressions in equation (4.1) we have,


H (4.2)

It can be seen that H is independent of both load and. size of indentation. This equation has
a very important meaning. Within certain limits, the hardness number for a given shape of
pyramid is the same regardless of the load used. The hardness measured by a square based
pyramid is the ‘Vicker's hardness' and is standardized by ASTM under the more general
name 'Diamond Pyramid Hardness' (DPH).

Structural Mechanics and Materials Testing Lab Page 31


F
F = Load (kgf)
d d
d  1 2 (mm)
2
F
HV  1.845 2
d
HV = Vicker’s hardness
d1 d2

Fig 4.1: Vicker’s test

4.2.2 Brinell Hardness


Brinell developed one hardness measurement using a hardened steel ball. The indentation
is a circular depression and the Brinell hardness number is defined as:

P
HB = A (4.3)

where, A = area of contact in sq. mm between the ball and the indentation and P = applied
load in kgf.

F F = Load (kgf)
D = ball diameter (mm)
D d = mean diameter of indentation (mm)
HB = Brinell hardness
2F
HB =
 
D D D2  d 2 
d

Fig 4.2: Brinell Test

In Brinell hardness test, the selection of the load (P) and the diameter (D) of the ball
(penetrator) is quite important. Injudicious selection of these parameters may lead to
meaningless results. Fig 4.3 shows the importance of the use of appropriate load and
penetrator. Fig.4.3 (b), for example, shows a case where too great a load compared to the
diameter of the ball has been applied on a relatively soft metal. As a result the ball has sunk
to its full diameter and the result is obviously meaningless. For different materials,
therefore, the ratio of P/D2 has been standardized as shown in Table 4.1. The Brinell
hardness number is measured as the load in kilograms per square millimeter of spherical
impression made in the test.

(a) (b) (c)

(a) Depth of impression too shallow which may lead to inaccurate determination of diameter and hence of hardness. (b) The
ball has sunk to its full diameter and hence determination of hardness from this is meaningless. (c) Depth of impression is
satisfactory.

Structural Mechanics and Materials Testing Lab Page 32


Fig 4.3: Depth of impression and accuracy of hardness determination

Table 4.1: Brinell hardness test conditions for different materials


Specimen Standard conditions for P/D2 for other Range of
Brinell Test penetrator HB
Material Min. Load (P) Ball dia. (D) size
thickness kg mm
Steel ------ 3,000 10 30
Grey cast >15 mm 1,000 10 10 <140
iron 5 - 15 mm 750 5 30 140 – 500
<5 mm 120 2 ------
Copper & its ------ 10 25 – 200
alloys ------ 5 <40
------ 30 >190
Light metals ------ 5 <55
and alloys ------ 5 – 15 55 – 80
------ 15 80

4.2.3 Rockwell Hardness


The Rockwell hardness is measured by use of a steel ball or a cone shaped diamond point.
It differs from measurements already discussed in that the depth of the impression is
measured instead of its diameter. However, since the two are always geometrically related,
the hardness measurement is the same in principle. Generally a minor load (FO) is applied
first to properly position the indentor to prevent any slippage due to the applied major load
(F). The hardness value is defined as (see Fig. 4.4):

F
FO FO FO

E E-e

Rockwell hardness HR = E-e


Where, E = 100 (diamond brale) for C scale.
E = 130 (steel ball) for B scale
e = permanent increase in depth of impression due to
load F (kgf), measured in units of 0.002 mm

Fig 4.4 Rockwell hardness

The Rockwell hardness value (HR), which is a dimensionless index, is read directly from a
specially graduated dial indicator. The readings taken after the major load is applied and
removed while the minor load is still in position. Two types of indentor are used in the
Rockwell test; (i) a steel ball of 1/16 inch in diameter (other diameters are also used) and
(ii) a diamond brale (a diamond cone with 120° apex angle). The steel ball indentor is used
with softer metals whereas the diamond brale is used for harder metals. The major and
minor load to be used in either case appears in Table 4.2.

Structural Mechanics and Materials Testing Lab Page 33


Table 4.2: Relevant details for using with a steel ball (1/16") or brale indentor in a Rockwell tester
Type of indentor Scale to be used Minor load (FO) kg Major load (F) kg
Steel ball 1/16" dia. B 10 90
Diamond brale C 10 140

The dial of the Rockwell hardness tester has two sets of figures, one red and the other
black. In the designation of scales, it should be noted that the red figures are used for
readings obtained with ball indentor regardless of the size of the ball or magnitude of major
load and that black figures are used only when the brale indentor is in use. These two scales
are separated by 30 Rockwell units. This is done to avoid a negative reading for the
relatively softer materials. The red and black figures correspond to the B and C scales
respectively when used with the load and penetrator size as specified in Table 4.2.

Other scales are also used with the Rockwell machine. Each of these scale is indicated by a.
symbol (A,D,E,F,G,H,K,L,M,P,R,S and V), which denotes the sizes of the penetrator and
the load. However, Rockwell B and C are the most commonly used scales.

In reporting the Rockwell hardness values, it is important to mention the scale used. For
example, a hardness measurement reported merely as a number as read form the instrument
dial, say 50, has no meaning whatsoever. In other words the hardness scale is not defined.
Thus the hardness numbers must be prefixed by the letter B or C as appropriate.

4.3 Measurement of Rockwell Hardness

4.3.1 Operating principle of Rockwell Hardness Tester


The principle of operation of Rockwell hardness tester is described in Fig.4.5. The various
steps involved in the test are:

(i) The specimen of interest is placed on the anvil at the upper end of the elevating
screw.
(ii) The elevating screw is rotated so as to bring the specimen surface into contact with
the penetrator. By further elevating the specimen, the minor load of 10 kg is applied
and is fully effective when the small pointer is coincident with its index mark .This
has forced the indentor into the specimen to a depth corresponding to A –B.
(iii) The major load is applied by means of a release handle. Through an oil-dashpot
arrangement the major load is applied at a definite rate. The application of the
major load has forced the ball penetrator to an additional depth corresponding to B-
C.
(iv) Without removing the minor load of 10 kg, the major load is withdrawn allowing
the impression to recover elastically by an amount C-D.
(v) With removal of major load the hardness test is complete. The hardness number is
directly read from the dial.

Structural Mechanics and Materials Testing Lab Page 34


Fig 4.5: Illustrating the principle of operation of the Rockwell hardness tester.

4.3.2 The Minor Load Feature of the Rockwell Machine


Application of the minor load is always made in order to sink the penetrator below the
surface of the test specimen. The remaining portion of the loading force (the major load) is
then used to make a hole whose depth will be the inverse measure of the hardness. This
feature of first applying a small part of the force, as a means of embedding the penetrator
below the surface of the test specimen, gives much more reliable and reproducible results
.It almost entirely eliminates the effect of surface finish differences by smoothening out
irregularities .The depth of penetration is measured from this smooth surface created by the
minor load .This method of hardness testing consists of measuring the increment of depth
of the penetrator that was forced into the metal by a primary and a secondary load and is
also known as " differential depth measurement method".

4.3.3 Dial Reading Technique


Correct hardness measurement requires a thorough understanding of the revolution of the
indicator dial gauge On a Rockwell machine when the testing load is applied to the
specimen. The force applied by the loading weights through the leverage system, the
sharpness of the penetrator and the resistance to penetration exerted by the materiel are the
variables that have a combined effect on the amount of the revolution of the large pointer
.The amount may be more or less than one revolution of the dial. Consequently, it is not
sufficient to read the hardness value by observing the final position of the pointer when the
test is completed .The entire movement of the pointer from the moment the additional
major load is applied until the pointer comes to a rest with the additional load removed at
the conclusion of the testing procedure must be recorded.

For correct determination of hardness values one has to understand the circular
arrangement of the dial scale which is capable of indicating not only numbers from 0 to
100 which the dial shows, but also numbers less than zero and over 100, even though such
numbers do not appear on the dial. The dial scale should be thought of as representing a

Structural Mechanics and Materials Testing Lab Page 35


scale curled into a circle with negative numbers and numbers over 100 overlapping the 0 to
100 portion. This is illustrated in Fig. 4.6. Figure 4.6 (a) represents the black numbered
scale and Fig. 4.6 (b) the red numbered scale. These two scales are out of phase with each
other by 30 Rockwell units.

Fig 4.6: Illustration of dial reading

As the major load is applied, the pointer moves counter clockwise from the set position of
100 to 90, 80, 70, etc. If it continues down beyond 0, its position will then be negative with
the 90 actually being (-10), 80 being (-20) etc. Subsequently when the additional load is
removed the pointer returns partway upscale to its final reading position. Similar
movement occurs on the red-scale. Thus merely reading the dial does not necessarily assure
valid hardness reading. A negative reading is invalid because the penetrator has sunk into
the metal tested, beyond the extent of its calibrated surface zone. Pointer movement
diagrams are useful in understanding the final reading and are shown in Fig. 4.7 for the RB
and RC scales.

4.3.4 Location and Spacing of Rockwell Test Impression


In order to ensure an accurate test, the centre to centre distance of adjacent indentations
must be at least 3 diameters of impression, closer spacing will give invalidly high results.
Similarly, the minimum permissible distance from the edge of a test specimen to the centre
of an impression must be at least 2.5 times the diameter of the impression.

4.4 Relationship between Hardness Numbers


Truly speaking there is no precise relationship exists between the several types of hardness
number. Approximate relationships have been developed by carrying out tests on the same
material using various devices. It should be borne in mind that these relationships are
material using various devices. It should be borne in mind that these relationships are
affected by many factors (materials, heat treatments etc.) and for this reason too much
reliance on them must be avoided. However, the, relationship provided Tables 4.3 to 4.5
can be used to obtain Brinell hardness number from the known Rockwell hardness number.

Structural Mechanics and Materials Testing Lab Page 36


Brinell (3,000 kg load) and Rockwell hardness numbers may also be converted
interchangeably with an accuracy of about 10 per cent according to the following
relationships:

For RB 35 to RB 100
7300
BHN  (4.4)
130  RB
For RC 20 to RC 40
20000
BHN  (4.5)
100  RC
For RC 41 or greater
25000
BHN  (4.6)
100  RC

Structural Mechanics and Materials Testing Lab Page 37


Table 4.3: Approximate hardness conversion numbers for nonaustenitic steel (ASTM
A370) (Rockwell C to Brinell hardness)
Rockwell C scale Brinell hardness no. Approximate tensile
150 kgf load and Diamond penetrator 3000 kgf load and 10 mm ball strength (ksi)
68 --- ---
67 --- ---
66 --- ---
65 739 ---
64 722 ---
63 706 ---
62 688 ---
61 670 ---
60 654 ---
59 634 351
58 615 338
57 595 325
56 577 313
55 560 301
54 543 292
53 525 283
52 512 273
51 496 264
50 482 255
49 468 246
48 455 238
47 442 229
46 432 221
45 421 215
44 409 208
43 400 201
42 390 194
41 381 188
40 371 182
39 362 177
38 353 171
37 344 166
36 336 161
35 327 156
34 319 152
33 311 149
32 301 146
31 294 141
30 286 138
29 279 135
28 271 131
27 264 128
26 258 125
25 253 123
24 247 119
23 243 117
22 237 115
21 231 112
20 226 110

Structural Mechanics and Materials Testing Lab Page 38


Table 4.4: Approximate hardness conversion numbers for nonaustenitic steel (ASTM
A370) (Rockwell B to Brinell hardness)
Rockwell B scale Brinell hardness no. Approximate tensile
100 kgf load and 1/16” steel ball 3000 kgf load and 10 mm ball strength (ksi)
penetrator
100 240 116
99 234 114
98 228 109
97 222 104
96 216 102
95 210 100
94 205 98
93 200 94
92 195 92
91 190 90
90 185 89
89 180 88
88 176 86
87 172 84
86 169 83
85 165 82
84 162 81
83 159 80
82 156 77
81 153 73
80 150 72
79 147 70
78 144 69
77 141 68
76 139 67
75 137 66
74 135 65
73 132 64
72 130 63
71 127 62
70 125 61
69 123 60
68 121 59
67 119 58
66 117 57
65 116 56
64 114 ---
63 112 ---
62 110 ---
61 108 ---
60 107 ---
59 106 ---
58 104 ---
57 103 ---
56 101 ---
55 100 ---
Note: Table 4.3 and 4.4 give approximate relationships of hardness values and approximate tensile
strength of steels. It is possible that steels of various compositions and processing histories will deviate in
hardness-tensile strength relationship from the data presented in these Tables.

Structural Mechanics and Materials Testing Lab Page 39


Table 4.5: Approximate hardness conversion numbers for austenitic steels
(ASTM A370) (Rockwell B to Brinell hardness)

Rockwell B scale Brinell hardness no.


100 kgf load and 1/16” steel ball 3000 kgf load and 10 mm ball
penetrator
100 256
99 248
98 240
97 233
96 226
95 219
94 213
93 207
92 202
91 197
90 192
89 187
88 183
87 178
86 174
85 170
84 167
83 163
82 160
81 156
80 153

4.5 Relation between the Hardness Number and Other Properties


In general, no correlation exists between any indentation hardness and the yield strength
determined in a tension test since the amount of inelastic strain involved in the hardness
test is much greater than in the test for yield strength. However, because of greater
similarity in inelastic strain involved in the test for ultimate tensile strength and indentation
hardness, empirical relations have been developed between these two properties. The
approximate tensile strength of steel for several types of material is indicated in Tables 4.3
and 4.4. The empirical relationships of Table 4.6 also give approximate tensile strength
from Brinell hardness number.

Table 4.6: Tensile strength from Brinell number.


Material Tensile strength (ksi)
For heat treated alloy steels with Brinell No.
Brinell No. X 0.42
of 250 to 400
For heat treated carbon steels as rolled
Brinell No. X 0.43
normalized or annealed
For medium carbon steels as rolled
Brinell No. X 0.44
normalized or annealed
For mild steels normalized or annealed Brinell No. X 0.46
For non-ferrous rough alloy such as duralumin  Brinell .Hardness 
  1  2.0
 4 

Structural Mechanics and Materials Testing Lab Page 40


Chapter 5
IMPACT TEST: FLEXURE AND TENSION
5.1 General
The behavior of materials under dynamic loading may sometimes differ markedly from
their behavior under static or slowly applied loads. Impact loading is an important type of
dynamic loading in which the load is applied suddenly, as from the impact of a moving
mass. The velocity of a striking body is changed, there must occur a transfer of energy,
work is done on the parts receiving the blow. The mechanics of impact involves not only
the question of stresses induced but also a consideration of energy transfer and energy
absorption and dissipation. The effect of an impact load in producing stress depends on the
extent to which the energy is expended in causing deformation. In the design of many types
of structures and machines that must take impact loading, the aim is to provide for the
absorption of as much energy as possible through elastic action and then to rely upon some
kind of damping to dissipate it. In such structures the resilience (i.e. the elastic energy
capacity) of the material is the significant property. In most cases the resilience data
derived from static loading may be adequate. Examples of impact loading include rapidly
moving loads such as those caused by a train passing over a bridge, or direct impact caused
by the drop of a hammer. In machine service, impact loads are due to gradually increasing
clearances which develop between parts with progressive wear.

5.2 Impact Testing Machine


The standard notched bar impact testing machine is of the pendulum type (Fig. 5.1). The
specimen is held in an anvil and is broken by a single blow of the pendulum or hammer,
which falls from a fixed starting point. In this condition it has a potential energy equal to
the "WH" where W is the weight of the pendulum and H is the height of the centre of
gravity above its lowest point. Upon release, and during its downward swing the energy of
the pendulum is transformed from potential to kinetic energy is equal to WH. A certain
portion of this kinetic energy goes into breaking the specimen. The remainder carries the
pendulum through the lowest point and is then transformed back into potential energy WH'
by the time the pendulum comes to rest where H' is the height in position B. The energy
delivered to the specimen is (WH-WH'), the impact value. The values of H and H' are
indicated by a pointer moving on a scale. The scale is usually calibrated to read directly in
feet-pound.
O Point of support

Angle
of fall R
Angle
of rise A

W
B H

W H'

Fig 5.1: Space relation of pendulum machine.

Structural Mechanics and Materials Testing Lab Page 41


The impact testing machine generally has arrangement for two different initial position of
the hammer block. The one in the higher position is for specimens which are likely to
withstand higher energy. The energy-scale for this case is graduated over arrange of 0-240
ft-lb, whereas for the specimens which are likely to absorb less energy, are tested with the
hammer block positioned at a lower initial height. This second position of the hammer
block is to be used with a corresponding energy scale over a range of 0-100 ft-lb.

5.3 Specimens for Impact Testing


Following three types of specimen are used for impact testing of metals:
(i) Charpy simple beam
(ii) Izod cantilever beam
(iii) Charpy tension rod

The standard flexure test specimen is a piece 10X10X55 mm notched as shown in Fig.5.2
(a) (ASTM A370). The specimen which is loaded as a simple beam is placed horizontally
between two anvils as shown in Fig.5.2 (b), so that the knife strikes opposite the notch at
the mid-span. For impact-tension tests a specimen is secured to the back edge of the
pendulum. As the pendulum falls, a hammer block secured to the outstanding end of the
specimen strikes against two extended anvils, the specimen being ruptured as the pendulum
passes between the two anvils. Tension specimens may be plain or with circumferential
notch.

Fig 5.2: Details of Charpy simple beam Fig 5.3: Cantilever beam specimen and
mounting for the Izod Test

One type of plain specimen has a diameter of 6 mm; a corresponding notched specimen has
a diameter of 6 mm as for the first type. The tension test has not been standardized and is
not used to any great extent in commercial practice.

The cantilever specimen is a 10X10 mm in section and 75 mm long having a standard 45
notch 2 mm deep. The specimen is clamped to act as a vertical cantilever. The mounting of
the specimen and the relative position of the striking edge are shown in Fig. 5.3.

Structural Mechanics and Materials Testing Lab Page 42


5.4 Impact Testing
Impact tests are performed by applying a sudden load or impulse to a standardized test
piece held in a vise in specially designed testing machines (section 5.2). Notched bar test
specimens of different designs are commonly used for impact test. Two types of specimens
are standardized for notched impact testing and the impact tests for metals and alloys are
generally classified as the Charpy test (popular in U.S.A.) and the Izod test (commonly
used in U. K.). Usually the same impact machine is designed to conduct both the Charpy
and the Izod tests, with provisions for inter-changing the specimen supports.

In the Charpy test, the pendulum consists of an I-section with heavy disc at its end. The
pendulum is suspended from a shaft that rotates in ball bearings and swings midway
between two upright stands, at the base of which is located the specimen support. The
specimen which is loaded as a simple beam, is placed horizontally; between two anvils, so
that the knife strikes opposite the notch at the mid-span. The essential difference between
the Izod and the Charpy test is in the positioning of the specimen (Fig.5.2 and Fig.5.3).

5.5 Usefulness of Impact Tests


The results obtained from notched bar tests are not readily expressed in terms of design
requirements. Furthermore, there is no general agreement on the interpretation or
significance of results obtained with this type of test. The impact test measures the energy
absorbed in fracturing the specimen.
.
The sharpness of the notch significantly affects this value, but geometrically similar
notches do not produce the same results on large parts as they do on small test pieces.
Consequently true behavior of a metal in service can only be obtained by testing full-size
components, in a manner as they will be used in service. The figures, therefore, have no
design value. A material of 40 ft-lb. is not twice as strong as one of 20 ft-lb.

Structural Mechanics and Materials Testing Lab Page 43


CHAPTER 6

BUCKLINGOF SLENDER COLUMNS


6.1 Introduction
It has been discussed in the earlier chapters that the ultimate (axial) load carried by a
structural member is
PU   U A (6.1)

where,  U is the appropriate ultimate strength (tensile or compressive as the case may be)
of the material, and A is the cross-sectional area of the member. While the above equation
for determining the ultimate capacity of an axially loaded member is always true in the
case of tension member, for the compression member (i.e. column), however, this is true
only in the certain range of slenderness of the member concerned. For example, in the case
of stocky column (short column) the above equation will hold good and the ultimate load
will be obtained through complete plastification of the cross-section of the member. On the
other hand for relatively slender column (long column), the column will fail at a load lower
than that given by equation 6.1. The cross-section of the column may be entirely or at least
partly elastic, when this type of failure takes place. This is known as the buckling effect,
often an important consideration in the design of compression members. It should,
therefore, be clear that compression members may have two different modes of failure.
One is by exhaustion of the capacity (strength) of the section (eq. 6.1) and the other is due
to lack of stiffness (instability effect). The later mode of failure is considered for an
elaborate discussion in the remainder of this chapter.

6.2 The Buckling Phenomenon


Let us consider three simple examples with long, slender steel sections. In the first, we will
use two pieces of identical length and cross-sectional area, but of different shape: one piece
is round, 1 4 in inch diameter, and the other is thin 1Z0.05 inch (Fig 6.1). The thin piece
carries 1 18 th of the load of the round one, even though it has the same weight and cross-
sectional area. P P
1 2

P1=18 P2

0.25 inch dia


1.0 inch

0.05 inch

Fig 6.1: A thin strut has a lower buckling load. The cross-sectional area, the material, the
length, and the support conditions are the same for both struts.

Structural Mechanics and Materials Testing Lab Page 44


In the next example, we will take two pieces of identical cross-section, but one twice as
long as the other (Fig.6.2). The long piece carries only a quarter of the load of the shorter
piece.

P1 P2

L P1=4P2

2L

Fig 6.2: A long strut has a lower buckling load. The cross-sectional dimension, the
material, and the support conditions are the same for both the struts.

In the third example we will take three pieces of steel of identical length and cross-section.
We will load one as a cantilever; i.e., one end is firmly fixed and the other is allowed to
move freely. The second one loaded as a pin-ended column; i.e., the ends are allowed to
rotate freely, but they are restrained to remain in line with the load. Another is built-in; i.e.,
its ends are not allowed to rotate, and it can move only in the direction of the load (fig 6.3).
The built-in column carries four times as much as the pin-ended column, and sixteen times
as much as the cantilever column. Buckling evidently depends on the end restraint, the
stiffness, and the length and the cross-sectional geometry (which determine the
slenderness) of the column. It is an elastic phenomenon, and failure occurs whenever the
energy required to recover the original shape is greater than the energy required to continue
the buckling deformation.

P1 P2 P3

Built-in column Pin-ended column Cantilever column

P1 = 4P2 = 16P3

Fig 6.3: Effect of end restraints on the buckling load. The cross-sectional dimensions, the
material, and the length are the same for all three columns.

6.3 Euler’s Theory for Slender Columns


Leonard Euler derived the formula for the strength of a slender pin-ended column in 1757.
It is an important part of the theory of structures and is the oldest structural formula still in
use. Consider an initially straight elastic strut under compressive end-loads (Fig 6.4).

Structural Mechanics and Materials Testing Lab Page 45


Suppose that this strut can have equilibrium positions not only when it is straight, but also
when flexed elastically (Fig 6.5). If we use the co-ordinates marked on Fig 6.5, x to give
the distance from the rigidly pined end and y to give the deflection corresponding to x, and
if the strut’s bending stiffness in the y direction is EI, we can use the usual beam bending
formula (see equation 7.6),

M E
 (6.2)
I R

where, M is the moment at a given section,


I is the moment of inertia of the section,
E is Young’s modulus and
R is the local radius of curvature at the section considered.

It is easily shown that the radius of curvature at any section is given to a first
approximation by
1 d2y
 2 (6.3)
R dx
d2y
Hence, M   EI 2
dx

The bending moment M is in this case given by M=Py, so that

d2y
Py   EI (6.4)
dx 2

which is the differential equation for equilibrium configuration of the strut.

l
P P

Fig 6.4: Pin-ended column under compressive end loads.


y
C
A B
P P
x

Fig 6.5: Pin-ended column buckled elastically under compressive loads.

The general solution of this equation is standard and is


P P
y  A sin x  B cos x
EI EI

We know that at x=0, y=0 and at x=l, y=0. Applying these boundary conditions:

(i) B=0, and


P
(ii) A sin l 0
EI

Structural Mechanics and Materials Testing Lab Page 46


Hence, for equilibrium to exist we have either A=0 (i.e., the strut does not deflect from its
straight configuration)
P
or sin l  0;
EI

P
the later case implies that, l  0,  , 2 ,......... .
EI
P
we can eliminate l  0 , for it merely considers the unloading case. The next solution is
EI
P
l   , which leads to the critical value of P (i.e., the buckling load):
EI
 2 EI
P  Pcr  (6.5)
l2
P
Since we know now that y  A sin x then:
EI
either i) A=0
 2 EI 4 2 EI
or ii) A is indeterminate if P  , , etc.
l2 l2
Thus, there is no deflected equilibrium position of the strut until a critical load given by
 2 EI
Pcr  2 is reached, when the magnitude of the possible deflection becomes
l
indeterminate. This is seen as an abrupt buckling of the strut when the critical (Euler) load
is reached. No further information is given by the present analysis, but a more exact
analysis would show that the undeflected equilibrium position loses its stability (though it
still exists) above the critical load, and that the deflected equilibrium positions are not
indeterminate but have very large amplitudes compared with the range of accuracy of the
1 d2y
initial assumption   2
R dx
In practice all struts will have very small shape imperfections which mean that small
deflections do in fact occur before the critical load (see Fig. 6.6). However, these are of a
 2 EI
much lower order of magnitude than the buckling deflections. The formula Pcr  2 can
l
be rewritten if we write I as Ar2 (i.e. the area of cross-section times the square of the radius
of gyration). We can then divide through by A and obtain:
Pcr  2 Er 2
  cr  , or
A l2
 2E
 cr  2
(6.6)
l
 
r

l
Where  cr is the critical stress and is defined as the slenderness ration. The equation
r
l
6.6 is plotted as the graph of  cr against and this is known as the Euler column curve of
r
a given material (fig 6.7).

Structural Mechanics and Materials Testing Lab Page 47


If the material is elasto-plastic with a yield stress  y , it is obvious that if the elastic  cr is
greater than  y , plastic squashing rather than elastic buckling takes place. This gives the
sharp cut-out to column strength curve at the lower slenderness ratios (Fig.6.7). In practice
the cut-off is not sharp as indicated but curves gradually, due to inelastic buckling.

Fig 6.6 Buckling of a Perfect and Imperfect Column

Fig 6.7 Column strength curve for pin ended column

Structural Mechanics and Materials Testing Lab Page 48


6.4 End Restrained Column: The Effective Length Concept
The preceding section describes the derivation of the critical load for a pin-ended column.
But in real situation columns may have a range of support condition giving rise to the
concept of end restrained column. The presence of end restraint affect the column strength
and it is desirable to make proper allowance for this in both the analysis and the design of
column. Use of the Euler type of approach to study the elastic buckling of end restrained
columns enables the elastic critical load to be related directly to the stiffness of the restraint
at either end, through the concept of effective length. This requires that the critical load
formula
 2 EI
Pcr  (6.7)
kl 2
where k is the effective length factor and kl (= ) is the effective length, which is defined
as the length of the equivalent pin-ended column that would have the same elastic critical
load as the actual end-restrained column. The effective length factor (k) for several
standard cases of rigid end restraint are shown in Table 6.1.

Table 6.1: Effective length for various end conditions


End condition of the column Effective Length ,
Both ends hinged 1.0 l
One end hinged and one end fixed 0.707 l
Both ends fixed 0.50 l
One end fixed and one end free ( i.e. cantilever) 2.0 l

Structural Mechanics and Materials Testing Lab Page 49


CHAPTER 7
BENDING OF BEAMS

7.1 Theory of Bending


When beams are subjected to loads, bending stresses are set. The computation of this stress
at a given section of the beam is facilitated by means of bending theory due to M. H.
Navier. He postulated that under a uniform bending moment, initially plane and parallel
cross-sections remain plane during bending and converge on a common centre of
curvature. This can be visually demonstrated by drawing a square grid on a rubber beam
(Fig.7.l (a)) and then bending the beam (Fig.7.1 (b)). Nevier's assumption has since been
proved correct for all structural materials by strain measurements on test beams under load
(Fig. 7 .2); strain gages are fixed at, say, each tenth of the depth of the beam, and the
variation of strain throughout the section is plotted as the beam is loaded.
Navier's theorem follows from the assumption that plane sections remain plane.
Consequently, under the action of a uniform bending moment M, two originally straight
and parallel lines AC and EG (Fig.7.3) remain straight, but converge on the common centre
of curvature, O. Evidently the beam is strained .in compression at the top and in tension at
the bottom, and somewhere between, there is a line at which the strain is neither tensile nor
compressive. This line of zero strain is called the neutral axis, BF.
Let us now draw a line through F parallel to AC. Since the planes started parallel, and are
now converging on the point O, the distance DE represents the tensile change of length at
the bottom of the beam, and the distance GH represents the compressive change of length
at the top of the beam. There is no change in length at the neutral axis, F. Let us call the
tensile change of length at the bottom of the beam DE=dx, and the original length
AD=BF=x, then the tensile strain at the bottom of the beam

dx
e
x

Let us call the radius of curvature, measured to the neutral axis, FO=R, and the distance of
the bottom form the neutral axis EF=y. Since the triangles DEF and BFO have all their
sides parallel, they are similar, and consequently
DE EF
 ; which gives the strain
BF FO
dx y
e  (7.1)
x R
We now determine the corresponding stress by Hook’s law:
Ey
f  Ee  (7.2)
R
The stress varies proportionately to the distance y from the neutral axis, and it is tensile
below and compressive above for a positive bending moment. The force acting on an
infinitesimally small area dA at a distance y from the neutral axis is and the moment of the
force about the neutral axis
Ey 2
dM  y.dP  y. f .dA  dA (7.3)
R

Structural Mechanics and Materials Testing Lab Page 50


Fig 7.1: Navier’s assumption. Originally plane and parallel section
(a) remains plane after bending (b) but converge onto a common center of curvature

Fig 7.2: Proof of Navier’s assumption with strain gauges. The strain is measured at
several points on the same cross section and the result for each load is plotted. So long as
the beam is elastic, the neutral axis remains in the same position, and the strains vary
proportionately with the distance from neutral axis.

The total resistance moment M of the section is the sum, or integral, of all the
infinitesimally small elements dM.

Ey 2
M  dA (7.4)
R

Since the modulus of elasticity is a constant, and the radius of curvature does not vary with
the depth y, we can take them outside the integral (or summation) sign.
E E
M   y 2 dA  I (7.5)
R R

Structural Mechanics and Materials Testing Lab Page 51


Fig 7.3: The theory of bending, the beam under action of a uniform bending moment M,
bends into a circular arc, whose radius of curvature is R=OF. Originally plane and
parallel sections ABC and EFG thus converge onto a center of curvature O. This causes
compressive strains on top and tensile strains at the bottom. The natural axis BF is the
line of zero strain. The maximum tensile strain, at a distance y below the neutral axis, is
DE/AD=dx/x. The corresponding stresses are shown on the left hand side of the diagram.

Where I   y 2 dA is called the second moment of area (also called as moment of inertia),
which is a purely geometric property of the section. Now from equation 7.2,
E f

R y
Substituting this in Eq. 7.5, one obtains,
M f E
  (7.6)
I y R
which is Navier's Theorem. In this equation, M is the bending moment at that particular
section of the beam, R is the radius of curvature to which the beam is bent by the moment
M, f is the stress at a distance y from the neutral axis, I is the second moment of area
(moment of inertia) of the section, and E is the modulus of elasticity of the material. In
general, we are interested only in the maximum stress (which must be kept within the
permissible range). This occurs at the greatest distance y form the neutral axis; in an
unsymmetrical section there are two different values- one for the bottom (Yb) and one for
the top (Yt), as shown in Fig.7.4. It is therefore convenient to introduce a further geometric
section property, the section modulus
I
s (7.7)
y
In the case of an unsymmetrical section there are different section moduli for bottom and
I I
top (Fig 7.4): sb  and st 
yb yt

Structural Mechanics and Materials Testing Lab Page 52


Navier’s theorem then becomes,
M  fs (7.8)

t
yt

yb
b

Fig 7.4: T-Section


7.2 Two Point Loading
In two point loading a simply supported beam is loaded at points L/3 from either support,
where L is the span of the beam as shown in Fig 7.5. The portion of the beam between the
loads is free from any shear and is subjected to purely flexural stress. The maximum
deflection (in the elastic range) at the centre of the beam for such a loading is given by:
23PL3
 max  (7.9)
1296 EI
where, P is the total applied load on the beam and EI is the flexural rigidity of the section.
Under a known value of load P, the deflection at the centre span can be measured
experimentally and use of the above relationship can be made to compute the value of ‘E’,
the modulus of elasticity of the material of the beam. For the purpose of examining the
Navier’s hypothesis two point loading scheme is adopted in the laboratory ensuring a
region of pure bending between the point loads. Strain measurements are made at various
distances from the neutral axis in the central portion of the beam.
P/2 P/2

Flexural rigidity = EI

L/3 L/3 L/3

P/2
P/2 Shear force diagram

PL/6

Bending moment diagram

δ
23PL3
 
Beam’s elastic deflection 1296 EI

Fig 7.5: Two point loading: pure moment in the central portion of the beam.

Structural Mechanics and Materials Testing Lab Page 53


CHAPTER 8
MECHANICAL PROPERTIES OF ENGINEERING
MATERIAL
8.1 Introduction
The efficient use of materials for engineering purposes requires an in-depth knowledge of
their mechanical properties. The most important of these properties are strength, stiffness,
ductility, hardness, resilience, toughness etc. The standardized methods for determining
these properties are described in earlier chapters. It is, however, important to appreciate the
limits of these properties within which they must fall for any given use of a material. This
final chapter, therefore, provides a brief description of the properties of common
engineering materials. Numerical values of these properties which are often useful are
included in tabular format for a handy reference. These values are indicative only and
should not be used for design purposes.

8.2 Materials for Engineering Use


A verity of materials is used for engineering purposes. Some of them are obtained
naturally, while the others are manmade, manufactured to give desirable properties. Figure
8.1 summarizes the strength (tensile and compressive), stiffness and density of various
materials used in buildings. These materials are classed in three groups depending on the
time of their availability in the history. This figure is more or less self explanatory. It is
interesting to note that stone, timber and some cast and wrought iron were the principal
structural materials available prior to nineteenth century. Most of these materials were
capable of withstanding high compressive load but were weak in tension. It was not until
the mid of twentieth century when the structural steel capable of resisting high tensile stress
was available. A brief description of some of the common materials is included in the next
section.

8.2.1 Ferrous Metals


The source of iron is the iron ore found in the earth's crust. Iron ore occurs in the form of
iron oxide containing small amounts of silica and other impurities. The iron is separated
from the oxide in a blast furnace where the molten iron is drawn off at the bottom of the
furnace and is cast in the form of small bars called pig iron. This pig iron cannot be used
directly but requires additional processing and refining to finally become cast iron, wrought
iron or steel.

(a) Cast Iron


Cast iron represents materials of varying mechanical and physical properties. A
considerable variation of properties of cast iron is achieved by differing the compositions
of the materials. Grey cast iron, the most widely used type, contains 2 to 4 percent carbon
and 1 to 3.5 percent silicon. The structure of grey cast iron contains free graphite plates
which impart properties peculiar to this material such as resistance to wear and corrosion,
dampening ability, and machinability. Other types of cast irons are chilled or white cast
iron, malleable cast iron and nodular cast iron. White cast iron is highly resistant to
abrasion; malleable cast iron has strength and ductility; and nodular cast iron, the structure
of which contains spheroidal graphite, tends to develop properties similar to those of
carbon-steel castings.

Structural Mechanics and Materials Testing Lab Page 54


(b) Wrought Iron
Wrought iron is obtained by refining pig iron in a furnace to reduce the carbon content to
less than 0.15 percent. Wrought iron possesses the important properties of ductility,
malleability, and toughness. It is, therefore, especially suitable for machine parts to be
shaped by forging.

Structural Mechanics and Materials Testing Lab Page 55


Also, its range of usefulness is considerably increased because of the ease with which it can
be welded. Its ultimate strength is about three-quarters of that of structural steel.

(c) Steel
The pig iron is refined by suitable process to remove excess carbon and other impurities
such as silicon, manganese, and sulfur, and then adding carbon and/or any of the alloying
elements required to produce steel having the desired qualities and properties. Steel is thus
an iron product containing carbon up to about 1.7 percent and is often technically referred
to as carbon steel. An alloy steel on the other hand is a carbon steel to which one or more
additional elements have been added to give specific properties to it.
(i) Carbon Steel: Pure iron (ferrite) is relatively soft and is both ductile and malleable but it
does not possess great strength. It (pure iron) has an ultimate tensile strength of about
40,000 psi. Carbon steel is produced by the addition of carbon to pure iron, in amounts
ranging from 0.05 to 1.7 percent. Addition of carbon has the effect of increasing strength
and hardness but at the cost of ductility and toughness. A carbon content of about 0.1
percent produces a so-called mild-steel and has an ultimate tensile strength of about 50,000
psi. Structural steel, which is commonly used for rolling into structural shapes (such as
angles, beams and columns of I or H shape), contains carbon of about 0.25 percent. This
gives a fair amount of ductility and toughness to the material with an ultimate strength of
about 64,000 to 72,000 psi. A carbon content of 0.40 percent produces machine steel
having an ultimate tensile strength of about 80,000 psi. When the carbon content is
increased to 0.75 percent, hard spring steel is produced which have an ultimate tensile
strength of 100,000 psi. A very hard tool steel is produced with 0.90 to 1 percent carbon
and has an ultimate tensile strength of 120,000 to 130,000 psi.
(ii) Alloy Steel: Alloy steel is produced by adding small amounts of one or more metals to
carbon steel in sufficient quantities so as to impart definite new properties in the steel. The
most commonly used alloying metals are nickel, manganese, chromium, vanadium and
molybdenum. The addition of these metals singly or in combinations has the effect of
increasing its strength and hardness, sometimes without sacrificing other desirable
properties already possessed by the carbon steel.

8.2.2 Non-Ferrous Metals

(a) Aluminium
Aluminium is obtained from an ore called bauxite, chemically hydrated oxide, found in
abundance on the earth's surface. Aluminium is very widely used in the manufacture of
item which need to be light weight e.g. in aircraft. Pure aluminium has an ultimate tensile
strength of less than 10,000 psi but when alloyed with 4 to 5 percents copper and smaller
percentage of manganese, magnesium, and/or other metals, various grades of wrought
aluminium can be obtained having ultimate tensile strengths between 20.000 to 60,000 psi
or more. Pure aluminium is resistant to corrosive actions but it is much less resistant in its
alloyed forms, especially when alloyed with copper.

(b) Brass
Brass is a Copper alloy. Ordinary brasses are those alloyed with Zinc only. Pure Copper
has strength in the range of 35,000 to 40,000 psi whereas brasses may have strength
ranging from 40,000 to 78,000 psi. Brass has elastic modulus in the range of 13x 106 to 17x
106 psi, which is much lower than steel. The main advantages of brass are its low
coefficient of friction, high corrosion resistance and high electric conductivity. These
properties make brass an ideal material for application such as sliding parts, radiator pipes,
wires, etc.

Structural Mechanics and Materials Testing Lab Page 56


8.3 Desirable Mechanical Properties in Particular Situation
Strength: Ductile materials, such as steel and aluminium alloys, have approximately the
same strength in compression as they do in tension. However, ductile materials do not
fracture under compression; they simply bulge and flatten. For this reason, the ultimate
compressive strength cannot be determined for ductile materials. Generally, it is assumed
to be the same as the ultimate tensile strength. Table 8.1 includes ultimate strengths of
commonly used structural materials and Table 8.2 shows typical yield strengths.
Ductile materials have shear strength of about two-thirds of that in tension. While in other
materials such as grey cast iron, brick, and concrete, which are brittle in nature have tensile
and shear strength which is only a fraction of their strength in compression.

Table 8.1: Typical Tensile Strength of Engineering Materials


Material Tensile Strength  U , (psi)
Alloy steela 230,000
Alluminium alloyb 80,000
Structural steelc 70,000

Table 8.2: Typical Yield Strength of Engineering Materials


Material Yield Strength at (Lower) Yield
0.2% offset  y , (psi) Point  y , (psi)
Alloy steela 215,000 --------
Alluminium alloyb 70,000 --------
c
Structural steel ---------- 40,000
a
Manganese-silicon steel; bAlloy 7075 (heat treated); cC1020 hot-rolled;

Elasticity: This property of a material is important in all structures subjected to varying


loads and is especially important in precision tools and machines. Conveniently the
proportional limit is regarded as the end of elastic action, though this is not true. Typical
values of proportional limit are provided in Table 8.3. Some materials such as cast iron,
concrete, timber has elastic properties somewhat less well defined than those of steel. The
former materials do not follow the Hook's law to a considerable extent (Le. their stress-
strain relationship is linear at only very early stages of loading). This lack of linearity in
behavior has not been found to be objectionable. However, this feature makes the
determination of the modulus of elasticity (stiffness) for these materials rather difficult. It is
common practice to take the slope of a tangent to the curve at the average allowable stress.

Stiffness: This property is desirable in materials used in beams, columns, machines, and
machine tools. Modulus of elasticity, which is obtained by dividing the unit stress by the
unit deformation caused by that stress; is the measure of stiffness of a material. Typical
values of stiffness for common engineering materials are included in Table 8.4.

Structural Mechanics and Materials Testing Lab Page 57


Table 8.3: Typical Elastic Properties of Engineering Materials
Material Proportional limit Modulus of
 pl , (psi) Resilience, in-
lb/in3
Alloy steela 210,000 735
Alluminium alloyb 60,000 180
c
Structural steel 35,000 20
a
Manganese-silicon steel; bAlloy 7075 (heat treated); cC1020 hot-rolled;

Table 8.4: Typical Stiffness and Other Properties of Engineering Materials


Material Modulus of Modulus of Poisson’s Unit weight,
Elasticity, E Rigidity, G Ratio, μ lb/ft3
(psi) (psi)
Aluminium (all Alloys) 10.3 E+06 3.80 E+06 0.334 169
Brass 15.4 E+06 5.82 E+06 0.324 534
Carbon steel 30.0 E+06 11.5 E+06 0.292 487
Cast iron grey 14.5 E+06 6.0 E+06 0.211 450
Concrete 3.0 E+06 1.4 E+06 0.100 150
Stainless steel 27.6 E+07 10.6 E+06 0.305 484
Copper 17.2 E+06 6.49 E+06 0.326 556

Ductility: Copper, aluminium, wrought iron, mild steel are the examples of ductile
materials. The measures of ductility are the percentage elongation and/or percentage
reduction of area before rupture of a test specimen. Table 8.5 compares typical ductility
values for common engineering materials.

Table 8.5: Typical Ductilities of Engineering Materials


Material Elongation in 2 in. (%) Reduction in area (%)
Alloy steela 10 20
b
Alluminium alloy 10 30
Structural steelc 35 60
Aluminium 25 80
a b c
Manganese-silicon steel; Alloy 7075 (heat treated); C1020 hot-rolled;

In a structural-steel bar, the average percentage of elongation in the 2 in. gage length (note
that percentage elongation should always be reported with the gage length, otherwise it
wouldn't carry any sense. Can you explain, why?) might average around 35 percent. The
reduction in area at the ruptured section would be about 60 percent. These percentages
reveal that structural-steel is a highly ductile material. This property of ductility is
extremely valuable in steel used for buildings, bridges and other structures, since noticeable
deformation often would give ample warning of impending failure.

A similar test on a hard steel specimen (carbon content 0.6 percent) would give a higher
strength, higher elastic limit, and no yield point. It would show low ductility; consequently
this material is brittle and would give little warning of impending failure.

Energy Absorption Capacity: When the materials are required to resist the action of
dynamic and impact loads, the energy absorption capacity becomes a useful tool in

Structural Mechanics and Materials Testing Lab Page 58


comparing materials for dynamic and impact applications. Two such energy quantities are
defined as: (i) Resilience and (ii) Toughness.

(i) Resilience: This property of a material is essential for all types of springs, as in an
automobile, on railroad cars, in watches, etc., where energy must be absorbed quickly
without causing permanent deformation. Resilience is sometimes referred to as elastic
toughness, since the energy must be absorbed without stressing the material beyond its
elastic limit. Typical values of modulus of resilience of some materials can be found in
Table 8.3.

(ii) Toughness: Toughness is the property which is measured by the amount of work done
to break a specimen. It often gives an indication of the materials resistance to sudden load.
The modulus of toughness is the amount of energy per unit volume which the material can
absorb prior to fracture (i.e. area under the stress-strain diagram up to failure). It should be
emphasized that a material can be stronger but still have a lower modulus of toughness than
another material (try to construct hypothetical stress-strain diagram of two such materials
for yourself).

Structural Mechanics and Materials Testing Lab Page 59


PART II
WORKBOOKS FOR LABORATORY PRACTICE

Structural Mechanics and Materials Testing Lab Page 60


CE 232 INDEX #ID: _______________

Experiment Experiment Name Date of Date of Remark


No Performance Submission

Tension test of mild steel


01
specimen

Compression test on
02 timber specimen

Direct shear test of metal


03 specimen

Test of helical spring


04

Hardness test of metal


05 specimen

Impact test of metal


06
specimens

Slender column test for


07 different end conditions

Static bending test of


08 timber beam

Structural Mechanics and Materials Testing Lab Page 61


WORK BOOK: 01 ID #
Date: Group #
EXPERIMENT NO: 01
TENSION TEST OF MILD STEEL SPECIMEN
1. OBJECTIVE:
i. To test a mild steel specimen to failure under tensile load.
ii. To draw stress-strain diagram.
iii. To study the failure characteristics of mild steel.
iv. To determine different properties of mild steel specimen.

2. APPARATUS:
i. Extensometer ii. Punch gauge iii. Slide calipers iv. Elongation scale

3. MACHINE:
i. Universal Testing Machine (UTM).

4. SPECIMEN:
Standard round specimen (ASTM E8)

5. PROCEDURE:
i. Measure the diameter of the specimen by slide calipers and draw a neat sketch of the
specimen.
ii. Record extensometer constant and gauge length.
iii. Fix the specimen in position.
iv. Fix the extensometer with the specimen.
v. Apply load and read the extensometer reading at a convenient interval of load (say
500lb).
vi. When yield point is reached, stop loading and remove the extensometer (since
extensometer would have reached its maximum dial reading by now; elongation
beyond this point (YP) is taken by means of an elongation scale.)
vii. Start loading again and take reading by the elongation scale at regular interval. In this
way, increase load gradually till a waist is formed in between the marked points and the
specimen breaks.
viii. Record the maximum and the breaking load.
ix. Remove the broken specimen and measure smallest X-sectional area and the final length
(L) between the gauge marks (see figure below) by fitting the two ends of the broken test
piece together.
x. Note the characteristics of the fractured surfaces and draw a sketch of the failed
specimen.

Lf

Fig: Final elongation of the specimen is obtained by putting the fractured ends of the tensile specimen.

Structural Mechanics and Materials Testing Lab Page 62


6. REPORT OF TENSILE TEST FOR MILD STEEL SPECIMEN:
Gauge length =___________________; Measured diameter =_______________________;
Original cross-sectional area =_____________________________.
Observation Load Extensometer Elongation Elongation Stress Strain Remarks
No. (lbs) reading (%) (inches) (psi) (in/in)

From your test results,


(i) Draw complete stress-strain diagram
(ii) Draw the stress-strain diagram up to the yield point in an elongated scale and
(iii)Fill the following (with appropriate units):
Proportional limit = ____________________Yield stress =__________________________
Ultimate strength =____________________Breaking stress =_______________________
Modulus of elasticity =_________________Modulus of resilience =__________________
Ductility: (a) % elongation =____________% reduction in area =____________________

Structural Mechanics and Materials Testing Lab Page 63


7. SAMPLE CALCULATIONS:

8. GRAPHS:

9. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 64


ANSWER THE FOLLOWING QUESTIONS
Mark the correct answer(s) by a tick (√) and fill in the blanks as appropriate:
1.Yield point exists in the case of :
(a) Mild steel (b) Cast iron (c) Concrete (d) Aluminium (e) Timber
2. The gauge length in our tension specimen was:
(a) 2.5 inch (b) 10 cm (c) 2 inch (d) 5 inch (e) 8 inch
3. The percentage of carbon in mild steel is typically:
(a) Below 0.1% (b) Above 0.3% (c) Around 0.2% (d) None of these
4. The stresses in the stress-strain diagram that you have drawn represents:
(a) True stress (b) Apparent stress (c) Engineering stress (d) None of these
5. Mild steel fractures in a characteristic _________________type of failure.
6. Some metals show discontinuous yielding due to the presence of
___________________.
7. The ratio of gauge length to the dia. of the specimen was __________________.
8. Ductile materials show a well defined _______________________.
9. A tensile specimen is tested using an extensometer that is inaccurate. Indicate by tick
marks which of the following properties will be in error:
(a) Proportional limit (b) Ultimate strength (c) Yield point (d) Yield strength

(e) Modulus of elasticity (f) Modulus of resilience (g) % elongation

Problem No.1
A specimen of 1 inch diameter has a modulus of resilience of 3600 in-lb/in3. Find the
elastic energy capacity of the specimen which is 4 inch in length.

Problem No. 2
A steel specimen (area=0.2 sq. inch) is to be tested. Its elastic strength is estimated to be
30,000 psi, corresponding to a strain of 0.001 in/in. It is desired to obtain ten points in the
elastic range. Determine convenient load increment to be used for observation.

Load
(lb)

Structural Mechanics and Materials Testing Lab Page 65


WORK BOOK: 02 ID #
Date: Group #
EXPERIMENT NO: 02
COMPRESSION TEST ON TIMBER SPECIMEN
1. OBJECTIVE:
i. To test a timber specimen under compressive loading parallel to the grain.
ii. To draw stress-strain diagram.
iii. To study the failure characteristics of the timber specimen.
iv. To determine different properties of timber specimen.

2. APPARATUS:
i. Compressometer ii. Steel scale iii. Slide calipers

3. MACHINE:
i. Universal Testing Machine (UTM).

4. SPECIMEN:
i. 2”X2”X8” wooden block (ASTM D143)

5. PROCEDURE:
i. Measure the size of the specimen with a slide caliper.
ii. Record gauge length and multiplying factor of the Compressometer.
iii. Attach Compressometer with the specimen and set the specimen in proper position of
the testing machine.
iv. Adjust the Compressometer dial to read zero.
v. Apply load continuously at a uniform speed until failure and read the Compressometer
at
1000 lbs interval.
vi. Take off the Compressometer when breaking starts.
vii. Record the maximum load and the final length between the gauge marks.
viii. Note the characteristics of the fractured surface and draw the extended surfaces of the
failed specimen and show the failure plane.

Structural Mechanics and Materials Testing Lab Page 66


6. REPORT OF COMPRESSION TEST FOR TIMBER SPECIMEN:
Gauge length =____________; Original cross-sectional area =_______________________

Observation Load Deformation Stress Strain Remarks


No. (lbs) (inch) (psi) (in/in)

From your test results,


(i) Draw complete stress-strain diagram and
(ii) Fill the followings (with appropriate units):
Proportional limit = _________________ Yield stress =_______________________

Ultimate strength =___________________ Breaking stress =_____________________

Modulus of elasticity =________________ Modulus of resilience =________________

Structural Mechanics and Materials Testing Lab Page 67


7. SAMPLE CALCULATIONS:

8. GRAPHS:

9. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 68


ANSWER THE FOLLOWING QUESTIONS
Mark the correct answer(s) by a tick (√) and fill in the blanks as appropriate:
1. The failure took place in a plane inclined by _______degree with the axis of the
specimen.
2. The timber specimen failed due to:
(a) Compression (b) Bending (c) Shear
3. ________________method was used to estimate the ________________ strength of the
specimen identified as the onset of the inelastic action.
4. For timber specimen the compressive strength parallel to the grain would be
_________________ than that of perpendicular to the grain.
5. Modulus of elasticity is a measure of:
(a) Strength (b) Stiffness (c) Both.

Structural Mechanics and Materials Testing Lab Page 69


WORK BOOK: 03 ID #
Date: Group #
EXPERIMENT NO: 03
DIRECT SHEAR TEST OF METAL SPECIMEN
1. OBJECTIVE:
i. To test metal specimen under shear
ii. To determine average strength in double and in single shear.

2. APPARATUS:
i. Johnson’s shear tool ii. Slide calipers

3. MACHINE:
i. Universal Testing Machine (UTM).

4. SPECIMEN:
i) Brass rod ii) Mild steel rod iii) High carbon steel (H.C.S) rod

5. PROCEDURE:
i. Measure the diameter of the specimen with a slide calipers.
ii. Fix the specimen in the shear tool such that it is in single shear and apply load until
rupture takes place.
iii. In the same way test the specimen in double shear.
iv. Repeat the experiment for two more specimens.

Structural Mechanics and Materials Testing Lab Page 70


6. REPORT OF DIRECT SHEAR TEST FOR METAL SPECIMENS:

Specimen Diameter Area, Single Shear Average Double Shear Average


designation (in) A shear stress, shear shear stress, shear
(in2) force, F1/A stress force, F2 F2/2A stress
F1 (lb) (psi) (psi) (lb) (psi) (psi)

Brass

Mild steel

H.C.S

From your test results, fill in the following (with appropriate unit):

Shear strength of Brass in single shear_____________________________.


Shear strength of Brass in double shear_____________________________.

Shear strength of M.S in single shear_______________________________.


Shear strength of M.S in double shear_______________________________.

Shear strength of H.C.S in single shear______________________________.


Shear strength of H.C.S in double shear_____________________________.

7. SAMPLE CALCULATIONS:

Structural Mechanics and Materials Testing Lab Page 71


8. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 72


ANSWER THE FOLLOWING QUESTIONS
1. Give three examples (with necessary sketches) of structural members subjected to shear
stress.

2. Fill in the blanks or tick(√) the correct choice as appropriate:


i) The special attachment used for direct shear test
was____________________________________.
ii) The single shear strength of a specimen should always be ________________than its
double shear strength, given that all other variables are same for both the cases.
iii) Were the test specimens subjected to any stress other than that of transverse shear?
Choose which of the following this stress may be:
a) torsional shear b) compression c) bending d) none of the mentioned.
iv) The distribution of shearing stress on a cross-section of a beam is
_______________________.
v) In direct shear test it is not possible to obtain the modulus of rigidity due to the fact that:
a) strains are small b) Hook’s law is not valid c) strain are immeasurable d) none
of these.
vi) Johnson shear tool is not very accurate for measuring shear strength, because of
a)______________________ and b)_____________________.
vii) Which part of the I-section mainly carries the shear stress?
a) Top flange b) Both flange c) Web d) none of these.

3. If a given round steel bar having proportional limit in shear and tension of 36,000 and
60,000 psi respectively, is used to absorb energy without undergoing plastic deformation,
would it absorb more energy if used as torsion member or as an axially loaded tension
member? Explain.

Structural Mechanics and Materials Testing Lab Page 73


WORK BOOK: 04 ID #
Date: Group #
EXPERIMENT NO: 04
TEST OF HELICAL SPRING
1. OBJECTIVE:
i. To find the stiffness of the spring.
ii. To compare the theoretical stiffness with the observed value.
iii. To draw a curve by plotting load against deflection and find the modulus of rigidity of
the spring material.

2. APPARATUS:
i. Slide calipers

3. MACHINE:
i. Testing machine

4. SPECIMEN:
i. Helical spring

5. PROCEDURE:
i. Measure the diameter of the spring wire at several locations to find the mean diameter.
Record the number of turn of the spring as N.
ii. Measure the outer and inner diameter of the helical spring and hence find the mean
radius of the spring.
iii. Place the spring between the table and the compression block of the testing machine.
iv. Before applying any load record the vernier scale reading. Apply load at an interval of
50 kg and record the vernier reading for deflection.
v. Repeat the same operation up to 500 kg.
vi. Repeat the whole process for decreasing load at the same interval.

Structural Mechanics and Materials Testing Lab Page 74


6. REPORT OF HELICAL SPRING TEST:
Height of the spring =_______________; Dia. of the wire (d) =______________________
No. of turns (N) =__________________; Mean radius of the helix(R) =_______________

Load increasing Load decreasing Average Actual Theoretical Average


Applied
Scale Deflection Scale Deflection deflection stiffness stiffness stiffness
Load, P
reading δ1 reading δ2 Δδ=( δ1+ δ2)/2 ΔP/Δδ (Eqn: 3.7) (graph)
(kg)
(mm) (mm) (mm) (mm) (mm) (kg/mm) (kg/mm) (kg/mm)

From your test results, fill in the following (with appropriate unit):
Stiffness of the spring =______________________________
Modulus of rigidity =________________________________
Maximum shear stress resisted =________________________
Ratio of max. torsional shear stress to the max. total shear stress (%) =________________

7. SAMPLE CALCULATIONS:

Structural Mechanics and Materials Testing Lab Page 75


8. GRAPHS:

9. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 76


ANSWER THE FOLLOWING QUESTIONS
1. Does the diameter of the wire and the diameter of the helix need to be measured with the
same precision? Explain.

2. One turn of the helix is assumed to be exactly 2πR in length. On what condition this
assumption is valid?

3. Is there any difference between the stiffness of a material and the stiffness of the spring?

4. Comment on the apparent difference between the theoretical and actual stiffness.

5. By sketch show the distribution of torsional shearing stress and direct shearing stress in a
circular section.

6. Give two example (with sketch) of structural members subjected to torsion.

Structural Mechanics and Materials Testing Lab Page 77


WORK BOOK: 05 ID #
Date: Group #
EXPERIMENT NO: 05
HARDNESS TEST OF METAL SPECIMEN
1. OBJECTIVE:
i. To determine the Rockwell Hardness Number of metal specimen.
ii. To find Brinell’s Hardness Number from that of Rockwell’s.
iii. To find the tensile strength of the metal from Brinell Hardness Number by using
empirical relationships.

2. MACHINE:
i. Rockwell Hardness Tester

3. SPECIMEN:
Plates of i) Brass, ii) Mild steel, iii) High carbon steel (H.C.S)

4. PROCEDURE:

a) Determination of Rockwell Hardness Number


i. Examine the machine and get it ready for testing (note that the weight marked 100 kg is
in position)
ii. Place the specimen upon the anvil of the machine.
iii. Raise the anvil and the test piece by elevating screw until the specimen came in contact
with the ball.
iv. Apply 10 kg minor load on the specimen by moving the pointer of the dial by three
times.
v. Apply the major load (140 kg for harder specimen and 90 kg for softer material) on the
specimen by pressing the appropriate handle.
vi. After few seconds of the application of the major load, return the handle to its original
position (i.e. the major load is now withdrawn).
vii. Read the position of the pointer on the scale of the dial as the Rockwell Hardness
Number (C-scale is used for harder specimen and B-scale for softer material).
viii. Take three readings for each specimen by changing the location of indentation.
ix. Repeat the above steps for each of the different specimen supplied.
b) Determination of Brinell Hardness Number
Use the chart provided in chapter 3 of your laboratory manual to obtain Brinell Hardness
Number from the Rockwell Numbers determined above (note the scale used in each case
when the Rockwell Hardness Number was determined). If the values are not available in
the chart use the approximate equations of Section 4.3.

c) Determination of Tensile Strength


i. Use the empirical relationships (Sec 4.4) to obtain tensile strength from the known
Brinell Hardness Number (determined in ‘b’ above).
ii. Another set of tensile strength should be obtained using the chart (Sec 4.3). Use the
already obtained Brinell Number for this purpose.

Structural Mechanics and Materials Testing Lab Page 78


5. REPORT OF HARDNESS TEST FOR METAL SPECIMENS:

Load Scale Rockwell Mean Brinell Tensile Strength

indentator used
applied used Hardness R.N. Hardness (ksi)

observation

designation
Specimen
(kg) Number (B.N.)

Type of
No. of

From chart
Empirical
(R.N.) (3000 kg
load. 10
mm ball)

1
2 Brass
3
1
2 M.S.
3
1
2 H.C.S
3
1
2 C.I
3

6. SAMPLE CALCULATIONS:

Structural Mechanics and Materials Testing Lab Page 79


7. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 80


ANSWER THE FOLLOWING QUESTIONS
1. Indicate which of the following statements are true or false?
a) Rockwell hardness is an indentation hardness test.
b) Brinell hardness has unit Rockwell hardness is non-dimensional.
c) Rockwell hardness is in fact a reverse measurement of the depth of penetration.
d) It is possible to have a negative value in a Rockwell test.
e) Rockwell’s ‘B’ & ‘C’ scales are separated by 30 hardness number.
f) Rockwell hardness has only two scales (B and C) depending on the load and penetration
used.

2. Indentation hardness will give indication about ultimate tensile strength than yield
strength, why?

3. What is the advantage of hardness test in compare to the tension test, if tensile strength
of a material is desired?

4. What is the significance of applying a minor load in Rockwell test?

Structural Mechanics and Materials Testing Lab Page 81


WORK BOOK: 06 ID #
Date: Group #
EXPERIMENT NO: 06
IMPACT TEST OF METAL SPECIMENS
1. OBJECTIVE:
i. To find energy absorbed in fracturing mild steel and cast iron specimens under impact
load.

2. APPARATUS:
i. Slide calipers
ii. Impact testing machine

3. SPECIMEN:
Mild steel and cast iron specimen of the following types:
i) Charpy simple beam ii) Izod cantilever beam iii) Charpy tension rod.

4. PROCEDURE:
i. Measure the lateral dimensions of the specimen at full section and at the notch.
ii. Set the pointer to read minimum on the graduated disc and note down the initial errors
as explained in (iii).
iii. Set the hammer block in the position ‘A’ and the pointer along with the carrier, in the
position ‘a’ (see figure below) and then release it. When the hammer block stops
swinging, the pointer should be in position ‘b’. If not read the initial error bb’= I (this is
to be used for high scale).
iv. In the same way as described in (iii) determine the initial error for low scale.
v. Place and position the sample appropriately in the vice.
vi. Raise the pendulum and fix it in proper position. Release the pendulum and record the
energy absorbed.
vii. The correct energy absorbed by the specimen is then found by taking into consideration
of the initial error determined earlier. Note the condition of the failed specimen
(whether broken or not).
viii. Repeat steps (v) to (vii) for each of the specimen supplied.

A
(0)
b'
240
(100) 0
a b

Figure: Impact testing machine

Structural Mechanics and Materials Testing Lab Page 82


5. REPORT OF IMPACT TEST FOR METAL SPECIMENS:

Type of Material Cross Scale Initial Energy Corrected Remarks


specimen of the sectional used error, i absorbed, E energy, E-i
specimen area (in2) (ft-lb) (ft-lb) (ft-lb) (ft-lb)
Charpy M.S.
simple
beam C.I.
Izod M.S.
cantilever
beam C.I.
Charpy M.S.
tension
rod C.I.

6. SAMPLE CALCULATIONS:

Structural Mechanics and Materials Testing Lab Page 83


7. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 84


ANSWER THE FOLLOWING QUESTIONS
1. Give two examples, where the structure (or member) is to be designed for impact load.

2. Can you use the results of your impact test directly in design? Give reasons in support of
your answer.

3. Doubling the length of a bar of given diameter, doubles its resistance to axial impact,
why?

4. What type of material is suitable for resisting impact? Explain briefly.

Structural Mechanics and Materials Testing Lab Page 85


WORK BOOK: 07 ID #
Date: Group #
EXPERIMENT NO: 07
SLENDER COLUMN TEST FOR DIFFERENT END CONDITIONS
1. OBJECTIVE:
i. To determine the critical load or buckling load of slender columns.
ii. To compare the experimental and theoretical critical loads.
iii. To draw the column strength curves.

2. APPARATUS:
i. Slide calipers ii. Column testing apparatus iii. Steel scale iv.
Weights

3. SPECIMEN:
i. High tensile steel columns of different length

4. PROCEDURE:
i. Measure the length and the mean diameter of the column to be tested.
ii. Place the column in the column testing apparatus with the both end hinged condition so
that column ends can rotate easily but held in position.
iii. Apply load by placing equal weights on both pans. Give a small side force to laterally
deflect the column. Increase the load if the column straightens back upon removal of
the side force.
iv. Continue the process until the applied load is just sufficient to hold the column in a
bend condition. Record this load as the critical load.
v. Perform the experiment for four such columns with different lengths.
vi. Repeat the above steps (i) to (v) for two more end conditions: a) one end hinged and
one end fixed, b) both end fixed.
vii. Record the results in the table provided in the next section and draw the column
strength curves including the theoretical buckling curves.

Structural Mechanics and Materials Testing Lab Page 86


4
3
2
1
4
3
2
1
4
3
2
1
No. of Observation

End condition

one

end
end

One

Both
Both

fixed
fixed

hinged
hinged
Dia. of
the column, d (in)

Cross Sectional
Area, A
(in2)

6. SAMPLE CALCULATIONS:
Radius of Gyration,
r (in)
Length of the

Structural Mechanics and Materials Testing Lab


column, L (in)
5. REPORT OF SLENDER COLUMN TEST:

Effective Length,
Le (in)
Slenderness Ratio,
Le/r
Critical load, P
(lb)

Critical stress,
P/A (psi)

Page 87
 cr 

(psi)
 r 
 
 Le 
 2E
Theoretical

2
critical stress
7. GRAPHS:

8. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 88


ANSWER THE FOLLOWING QUESTIONS
1. Determine the slenderness ratio
a) Of a square column having a height breadth ratio of 10 (h/b=10)
b) Of a circular column having a height diameter ratio of 10 (h/d=10)
2. The resistance offered by a slender column is a function of what property of the
material?
3. A column has the cross-sections shown below. If the proportional limit is 30,000 psi,
find out the length at which it can be just called a slender column. Find out the same for
a material whose proportional limit is 40,000 psi. For both cases the modulus of
elasticity of the material is 30X106 psi.

0.25”

1.0”

0.5”

Structural Mechanics and Materials Testing Lab Page 89


WORK BOOK: 08 ID #
Date: Group #
EXPERIMENT NO: 08
STATIC BENDING TEST OF TIMBER BEAM
1. OBJECTIVE:
i. To study the behavior of a timber beam under load.
ii. To verify Navier’s theorem.
iii. To find the value of ‘f’ and to see that it is same for both the top and bottom fibers’.

2. APPARATUS:
i. Beam testing machine ii. Whittmore strain gauge iii. Deflectometer
iv. Steel scale or tape v. Bearing devices vi. Rollers

3. SPECIMEN:
i. Timber beam

4. PROCEDURE:
i. Place the beam over the platform of beam testing machine with supports at two ends.
ii. Measure distance of each of the gauge lines from the bottom of the beam.
iii. Set the deflectometer at the bottom centre of the beam.
iv. Prior to the application of the load (self wt. of beam may be ignored) take initial
reading of the strain gauge at different holes.
v. Arrange two point loading and increase the load at regular intervals and record strain
gauge and deflectometer readings. (Loading should be kept within such limit so that the
beam stresses remain elastic. This is important because Navier’s theorem as well as the
deflection formula we are using holds good for elastic cases only.)
vi. Calculate theoretical stresses along the gauge lines by Navier’s theorem.
My
f 
I
vii. Draw the load-deflection curve (average straight line of the experimental points) from
which obtain the value of ‘E’ using the relationship
23PL3

1296 EI
viii. Plot stress-strain diagram for the top holes and graphically find out E.

Structural Mechanics and Materials Testing Lab Page 90


5. REPORT OF FLEXURE TEST FOR TIMBER BEAM:
Gauge length: ________________ Span of the beam: ___________________
Depth of the beam: ________________ Width of the beam: __________________
Loa 1st hole 2nd hole 3rd hole 4th hole 5th hole

Deflectometer

Deflection
reading
d Unit Unit Unit Unit Unit

(in)
S.G.R1 Strain S.G.R Strain S.G.R Strain S.G.R Strain S.G.R Strain
(lb) (in/in) (in/in) (in/in) (in/in) (in/in)

Draw the following curves from the results obtained:


a) Load vs. Deflection
b) Stress-Strain diagram
c) Strain-Distance diagram
Results:
a) Maximum tensile stress = __________________________
b) Maximum compressive stress = ______________________
c) Modulus of elasticity (E) = __________________________
d) Flexural rigidity of the section (EI) = __________________

6. SAMPLE CALCULATIONS:

____________________________________
1
S.G.R=Strain gauge reading

Structural Mechanics and Materials Testing Lab Page 91


7. GRAPHS:

8. DISCUSSIONS:

Structural Mechanics and Materials Testing Lab Page 92


ANSWER THE FOLLOWING QUESTIONS

1. State briefly how the modulus of elasticity of a beam material can be obtained in the
field?
2. For the beam sections shown, draw qualitative strain (i.e. strain-distance) diagram in
each case.

Structural Mechanics and Materials Testing Lab Page 93

You might also like