SM Lab Manual BUET
SM Lab Manual BUET
AND MATERIALS
TESTING LABORATORY
MANUAL
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
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.
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.
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.
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.
(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.
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
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
(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.
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.
φ L
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
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.
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.
Stress, σ
εpl Strain, ε
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.
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
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
Undeformed bar
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.
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
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.
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)
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.
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
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.
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.
0.505˝ diameter
3/4"-10 no. thread
1/2" radius
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.
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.
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.
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
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.
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:
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).
(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.
CL
Shear diagram
e
Torque=Pe P
d
y
2 3V
Vmax=
2bd
V=vbdx
V
v
bW bw d
d
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
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:
T
T
c τ
τmax
ρ
T
Fig 3.7: Torsional shear stress in a circular cross-section under an applied torque T.
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.
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 )
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.
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.
2PR 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.
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.
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
In direct shear test, the shearing stress is considered to be uniformly distributed over a
cross- section. The two different cases may arise:
P
SS (3.9)
A
P (3.10)
SD
2A
P
P
P/2
P
P/2
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).
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
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) 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.
F
FO FO FO
E E-e
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.
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.
(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.
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
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.
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
Angle
of fall R
Angle
of rise A
W
B H
W H'
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.
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).
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.
P1=18 P2
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.
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
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.
M E
(6.2)
I R
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
d2y
Py EI (6.4)
dx 2
l
P P
We know that at x=0, y=0 and at x=l, y=0. Applying these boundary conditions:
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).
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
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
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
t
yt
yb
b
Flexural rigidity = EI
P/2
P/2 Shear force diagram
PL/6
δ
23PL3
Beam’s elastic deflection 1296 EI
Fig 7.5: Two point loading: pure moment in the central portion of the beam.
(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.
(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.
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.
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.
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
(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).
Compression test on
02 timber 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.
8. GRAPHS:
9. DISCUSSIONS:
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)
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.
8. GRAPHS:
9. DISCUSSIONS:
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.
Brass
Mild steel
H.C.S
From your test results, fill in the following (with appropriate unit):
7. SAMPLE CALCULATIONS:
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.
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.
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:
9. DISCUSSIONS:
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.
2. MACHINE:
i. Rockwell Hardness Tester
3. SPECIMEN:
Plates of i) Brass, ii) Mild steel, iii) High carbon steel (H.C.S)
4. PROCEDURE:
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:
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?
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
6. SAMPLE CALCULATIONS:
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?
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.
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
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:
0.25”
1.0”
0.5”
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.
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)
6. SAMPLE CALCULATIONS:
____________________________________
1
S.G.R=Strain gauge reading
8. DISCUSSIONS:
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.