You are on page 1of 71

SUNYANI TECHNICAL UNIVERSITY, SUNYANI

DEPARTMENT OF MATERIALS ENGINEERING


FACULTY OF ENGINEERING

LECTURE NOTES ON
STRENGTH OF MATERIALS II
COMPILED BY
ASUMADU TABIRI KWAYIE
MSc. MATERIALS ENGINEERING
CVE 104 STRENGTH OF MATERIALS II (1-2-2)
GENERAL OBJECTIVE
Fundamental concepts of Load Analysis and relationship between loads and Factors important in
Civil engineering Design.
0.0 Analysis of statistically determinate beams
0.1 Relation between load, shear force and bending moment
0.2 Calculation of bending moments and shear forces.
0.3 Bending moment and shear force diagrams
1.0 Theory of pure bending
1.1 Definition of pure bending
1.2 Types of stresses and strains resulting from pure binding.
1.3 Equation of bending stress.
1.4 Stress distribution in a section of a beam
1.5 Section modulus and calculation of bending stress.

3.0 Shear Stress


3.1 Equation of shear stress in a beam
3. 2 Shear stress distribution in a beam of rectangular cross-section
3.3 Shear stress distribution in an l- section beam
3.4 Solve examples of bending stress and shear stress, introduction to design principles
of a section.
4.0 Torsion
4.1 Define torque and angle of twist
4.2 Expression of torsion for a solid circular shaft, a hollow circular shaft and a
rectangular section in the elastic stages.
4.3 Torsion shear distribution for a solid shaft, a hollow shaft and a rectangular section in
the elastic stages.
4.4 Calculation of torsion shears for a solid and a hollow shaft.

i
Contents
CVE 104 STRENGTH OF MATERIALS II (1-2-2) ......................................................................i
GENERAL OBJECTIVE ............................................................................................................i
Contents ..........................................................................................................................................ii
CHAPTER 1: MECHANICAL PROPERTIES OF MATERIALS ................................................ 1
INTRODUCTION ...................................................................................................................... 1
Tensile Test ................................................................................................................................. 2
Ductile materials ......................................................................................................................... 4
Brittle materials .......................................................................................................................... 4
Allowable working stress-factor of safety .................................................................................. 5
Load factor .................................................................................................................................. 5
Toughness ................................................................................................................................... 6
FATIGUE AND CREEP ............................................................................................................ 7
FAILURE ................................................................................................................................... 9
INTRODUCTION .................................................................................................................. 9
FRACTURE ............................................................................................................................... 9
FUNDAMENTALS OF FRACTURE .................................................................................... 9
DUCTILE FRACTURE ........................................................................................................... 10
BRITTLE FRACTURE ............................................................................................................ 12
PRINCIPLES OF FRACTURE MECHANICS ....................................................................... 13
STRESS CONCENTRATION ................................................................................................. 14
FRACTURE OF POLYMERS ................................................................................................. 15
IMPACT FRACTURE TESTING ............................................................................................ 16
IMPACT TESTING TECHNIQUES ....................................................................................... 17
DUCTILE-TO-BRITTLE TRANSITION ................................................................................ 18
CHAPTER 2: BENDING MOMENT AND SHEAR FORCE..................................................... 19
Concept of a Beam.................................................................................................................... 19
TYPES OF LOADS AND BEAMS ......................................................................................... 19
Bending Moment ...................................................................................................................... 20
Shear Force ............................................................................................................................... 20
Sign conventions ....................................................................................................................... 21
B.M. AND S.F. DIAGRAMS ................................................................................................... 21
B.M. Diagram ........................................................................................................................... 21
S.F. Diagram ............................................................................................................................. 21
B.M AND S.F DIAGRAMS FOR A SIMPLY SUPPORTED BEAM .................................... 21
ii
Uniformly Distributed Load ................................................................................................. 22
Uniformly Varying Loads......................................................................................................... 23
SIMPLY SUPPORTED BEAM SUBJECTED TO A COUPLE ............................................. 27
CANTILEVER BEAM ............................................................................................................. 28
Concentrated Load at Free End ................................................................................................ 28
Uniformly Varying Load .......................................................................................................... 28
OVERHANGING BEAMS ...................................................................................................... 29
Concentrated Loads .................................................................................................................. 29
UnIformly Varying Load .......................................................................................................... 32
RELATIONSHIP BETWEEN LOAD, SHEAR FORCE AND BENDING MOMENT ......... 35
CHAPTER 3: PURE BENDING .................................................................................................. 39
Kinematics of pure bending ...................................................................................................... 39
Assumptions made in the theory of Pure Bending ................................................................... 39
The axial strain in a line element a distance y above the neutral surface is given by............... 40
Stress distribution in pure bending: .......................................................................................... 40
Axial load and the location of the neutral axis: ........................................................................ 41
Bending moment and its relation to radius of curvature: ......................................................... 43
CHAPTER 4: SHEAR STRESS ................................................................................................... 32
Complementary Shear Stress. ................................................................................................... 32
The Modulus of Rigidity .......................................................................................................... 35
Strain Energy ............................................................................................................................ 35
CHAPTER 5 TORSION ............................................................................................................. 36
Circular Shafts .......................................................................................................................... 36
Hollow, Thin-Walled Shafts ..................................................................................................... 38

iii
CHAPTER 1: MECHANICAL PROPERTIES OF MATERIALS
INTRODUCTION
Many materials, when in service, are subjected to forces or loads; examples include the aluminum
alloy from which an airplane wing is constructed and the steel in an automobile axle. In such
situations it is necessary to know the characteristics of the material and to design the member from
which it is made such that any resulting deformation will not be excessive and fracture will not
occur. The mechanical behavior of a material reflects the relationship between its response or
deformation to an applied load or force. Important mechanical properties are strength, hardness,
ductility, and stiffness.
The mechanical properties of materials are ascertained by performing carefully designed
laboratory experiments that replicate as nearly as possible the service conditions. Factors to be
considered include the nature of the applied load and its duration, as well as the environmental
conditions. It is possible for the load to be tensile, compressive, or shear, and its magnitude may
be constant with time, or it may fluctuate continuously. Application time may be only a fraction
of a second, or it may extend over a period of many years. Service temperature may be an
important factor.
Mechanical properties are of concern to a variety of parties (e.g., producers and consumers of
materials, research organizations, government agencies) that have differing interests.
Consequently, it is imperative that there be some consistency in the manner in which tests are
conducted, and in the interpretation of their results. This consistency is accomplished by using
standardized testing techniques. Establishment and publication of these standards are often
coordinated by professional societies. In the United States the most active organization is the
American Society for Testing and Materials (ASTM). Its Annual Book of ASTM Standards
comprises numerous volumes, which are issued and updated yearly; a large number of these
standards relate to mechanical testing techniques. Several of these are referenced by footnote in
this and subsequent chapters.
The role of structural engineers is to determine stresses and stress distributions within members
that are subjected to well-defined loads. This may be accomplished by experimental testing
techniques and/or by theoretical and mathematical stress analyses. These topics are treated in
traditional stress analysis and strength of materials texts.
Materials and metallurgical engineers, on the other hand, are concerned with producing and
fabricating materials to meet service requirements as predicted by these stress analyses. This
necessarily involves an understanding of the relationships between the microstructure (i.e.,
internal features) of materials and their mechanical properties.
Materials are frequently chosen for structural applications because they have desirable
combinations of mechanical characteristics. This chapter discusses the stress–strain behaviors
of metals, ceramics, and polymers and the related mechanical properties; it also examines
their other important mechanical characteristics.

If a load is static or changes relatively slowly with time and is applied uniformly over a cross
section or surface of a member, the mechanical behavior may be ascertained by a simple
stress–strain test; these are most commonly conducted for metals at room temperature. There
are three principal ways in which a load may be applied: namely, tension, compression, and
shear (Figures 7.1a, b, c). In engineering practice many loads are torsional rather than pure
shear; this type of loading is illustrated in Figure 7.1d.

1
Tensile Test
In order to compare the strengths of various materials it is necessary to carry out some standard
form of test to establish their relative properties. One such test is the standard tensile test in which
a circular bar of uniform cross-section is subjected to a gradually increasing tensile load until
failure occurs. Measurements of the change in length of a selected gauge length of the bar are
recorded throughout the loading operation by means of extensometers and a graph of load against
extension or stress against strain is produced as shown in Fig. 1.3; this shows a typical result for
a test on a mild (low carbon) steel bar.

Fig. 1.3 Typical tensile test curve for mild steel.


For the first part of the test, it will be observed that Hooke’s law is obeyed, i.e., the material
behaves elastically and stress is proportional to strain, giving the straight-line graph indicated.
Some point A is eventually reached, however, when the linear nature of the graph ceases and this
point is termed the limit of proportionality. For a short period beyond this point the material may
still be elastic in the sense that deformations are completely recovered when load is removed (i.e.,
strain returns to zero) but Hooke’s law does not apply. The limiting point B for this condition is
termed the elastic limit. For most practical purposes it can often be assumed that points A and B
are coincident. Beyond the elastic limit plastic deformation occurs and strains are not totally
recoverable.
There will thus be some permanent deformation or permanent set when load is removed. After
the points C, termed the upper yield point, and D, the lower yield point, relatively rapid increases
in strain occur without correspondingly high increases in load or stress. The graph thus becomes
much shallower and covers a much greater portion of the strain axis than does the elastic range of
the material. The capacity of a material to allow these large plastic deformations is a measure of
the so-called ductility of the material, and this will be discussed in greater detail below. For certain
materials, for example, high carbon steels and non-ferrous metals, it is not possible to detect any
difference between the upper and lower yield points and in some cases no yield point exists at all.
2
In such cases a proof stress is used to indicate the onset of plastic strain or as a comparison of the
relative properties with another similar material. This involves a measure of the permanent
deformation produced by a loading cycle; the 0.1 % proof stress, for example, is that stress which,
when removed, produces a permanent strain or “set” of 0.1% of the original gauge length-see Fig.
1.4(a).

Fig. 1.4. (a) Determination of 0.1% proof stress (b) Permanent deformation or “set” after straining
beyond the yield point.
The 0.1%proof stress value may be determined from the tensile test curve for the material
in question as follows: Mark the point P on the strain axis which is equivalent to 0.1%strain. From
P draw a line parallel with the initial straight-line portion of the tensile test curve to cut the curve
in N. The stress corresponding to Nis then the 0.1%proof stress. A material is considered to
satisfy its specification if the permanent set is no more than 0.1%afterthe proof stress has been
applied for 15 seconds and removed. Beyond the yield point some increase in load is required to
take the strain to point E on the graph. Between D and E the material is said to be in the elastic-
plastic state, some of the section remaining elastic and hence contributing to recovery of the
original dimensions if load is removed, the remainder being plastic. Beyond E the cross-sectional
area of the bar begins to reduce rapidly over a relatively small length of the bar and the bar is said
to neck. This necking takes place whilst the load reduces, and fracture of the bar finally occurs at
point F. The nominal stress at failure, termed the maximum or ultimate tensile stress, is given by
the load at E divided by the original cross-sectional area of the bar. (This is also known as the
tensile strength of the material of the bar.) Owing to the large reduction in area produced by
the necking process the actual stress at fracture is often greater than the above value. Since,
however, designers are interested in maximum loads which can be carried by the complete cross-
section, the stress at fracture is seldom of any practical value. If load is removed from the test
specimen after the yield point C has been passed, e.g., to some position S, Fig. 1.4(b), the
unloading line ST can, for most practical purposes, be taken to be linear. Thus, despite the fact
that loading to S comprises both elastic (OC)and partially plastic (CS) portions, the unloading

3
procedure is totally elastic. A second load cycle, commencing with the permanent elongation
associated with the strain OT, would then follow the line TS and continue along the previous curve
to failure at F. It will be observed, however, that the repeated load cycle has the effect of increasing
the elastic range of the material, i.e., raising the effective yield point from C to S, while the tensile
strength is unaltered. The procedure could be repeated along the line PQ, etc., and the material is
said to have been work hardened. In fact, careful observation shows that the material will no
longer exhibit true elasticity since the unloading and reloading lines will form a small hysteresis
loop, neither being precisely linear. Repeated loading and unloading will produce a yield point
approaching the ultimate stress value but the elongation or strain to failure will be much reduced.

Ductile materials

It has been observed above that the partially plastic range of the graph of Fig. 1.3coversa much
wider part of the strain axis than does the elastic range. Thus, the extension of the material over
this range is considerably in excess of that associated with elastic loading. The capacity of a
material to allow these large extensions, i.e. the ability to be drawn out plastically, is termed its
ductility. Materials with high ductility are termed ductile materials, members with low ductility
are termed brittle materials. A quantitative value of the ductility is obtained by measurements of
the percentage elongation or percentage reduction in area, both being defined below.

𝑖𝑛𝑐𝑟𝑒𝑎𝑠𝑒 𝑖𝑛 𝑔𝑎𝑢𝑔𝑒 𝑙𝑒𝑛𝑔𝑡ℎ 𝑡𝑜 𝑓𝑟𝑎𝑐𝑡𝑢𝑟𝑒


𝑃𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒 𝑒𝑙𝑜𝑛𝑔𝑎𝑡𝑖𝑜𝑛 = 𝑥 100
𝑜𝑟𝑖𝑔𝑖𝑛𝑎𝑙 𝑔𝑎𝑢𝑔𝑒 𝑙𝑒𝑛𝑔𝑡ℎ

𝑟𝑒𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑖𝑛 𝑐𝑟𝑜𝑠𝑠 − 𝑠𝑒𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑎𝑟𝑒𝑎 𝑜𝑓 𝑛𝑒𝑐𝑘𝑒𝑑 𝑝𝑜𝑟𝑡𝑖𝑜𝑛


𝑃𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒 𝑟𝑒𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑖𝑛 𝑎𝑟𝑒𝑎 = 𝑥 100
𝑜𝑟𝑖𝑔𝑖𝑛𝑎𝑙 𝑎𝑟𝑒𝑎

The latter value, being independent of any selected gauge length, is generally taken to be the more
useful measure of ductility for reference purposes. A property closely related to ductility is
malleability, which defines a material's ability to be hammered out into thin sheets. A typical
example of a malleable material is lead. This is used extensively in the plumbing trade where it is
hammered or beaten into corners or joints to provide a weatherproof seal. Malleability thus
represents the ability of a material to allow permanent extensions in all lateral directions under
compressive loadings.

Brittle materials
A brittle material is one which exhibits relatively small extensions to fracture so that the partially
plastic region of the tensile test graph is much reduced (Fig. 1.8). Whilst Fig. 1.3 referred to a low
carbon steel, Fig. 1.8could well refer to a much higher strength steel with a higher carbon content.
There is little or no necking at fracture for brittle materials.

4
Fig. 1.8. Typical tensile test curve for a brittle material.

Allowable working stress-factor of safety

The most suitable strength or stiffness criterion for any structural element or component is
normally some maximum stress or deformation which must not be exceeded. In the case of stresses
the value is generally known as the maximum allowable working stress. Because of uncertainties
of loading conditions, design procedures, production methods, etc., designers generally introduce
a factor of safety into their designs, defined as follows:

𝑚𝑎𝑥𝑖𝑚𝑢𝑚 𝑠𝑡𝑟𝑒𝑠𝑠
𝑓𝑎𝑐𝑡𝑜𝑟 𝑜𝑓 𝑠𝑎𝑓𝑒𝑡𝑦 =
𝑎𝑙𝑙𝑜𝑤𝑎𝑏𝑙𝑒 𝑤𝑜𝑟𝑘𝑖𝑛𝑔 𝑠𝑡𝑟𝑒𝑠𝑠

However, in view of the fact that plastic deformations are seldom accepted this definition is
sometimes modified to
𝑦𝑖𝑒𝑙𝑑 𝑠𝑡𝑟𝑒𝑠𝑠 (𝑜𝑟 𝑝𝑟𝑜𝑜𝑓 𝑠𝑡𝑟𝑒𝑠𝑠)
𝑓𝑎𝑐𝑡𝑜𝑟 𝑜𝑓 𝑠𝑎𝑓𝑒𝑡𝑦 =
𝑎𝑙𝑙𝑜𝑤𝑎𝑏𝑙𝑒 𝑤𝑜𝑟𝑘𝑖𝑛𝑔 𝑠𝑡𝑟𝑒𝑠𝑠

In the absence of any information as to which definition has been used for any quoted value of
safety factor the former definition must be assumed. In this case a factor of safety of 3 implies that
the design is capable of carrying three times the maximum stress to which it is expected the
structure will be subjected in any normal loading condition. There is seldom any realistic basis for
the selection of a particular safety factor and values vary significantly from one branch of
engineering to another. Values are normally selected on the basis of a consideration of the social,
human safety and economic consequences of failure.

Load factor

In some loading cases, e.g., buckling of struts, neither the yield stress nor the ultimate strength is
a realistic criterion for failure of components. In such cases it is convenient to replace the safety
factor, based on stresses, with a different factor based on loads. The load
factor is therefore defined as:

5
𝑙𝑜𝑎𝑑 𝑎𝑡 𝑓𝑎𝑖𝑙𝑢𝑟𝑒
𝑙𝑜𝑎𝑑 𝑓𝑎𝑐𝑡𝑜𝑟 =
𝑎𝑙𝑙𝑜𝑤𝑎𝑏𝑙𝑒 𝑤𝑜𝑟𝑘𝑖𝑛𝑔 𝑙𝑜𝑎𝑑

This is particularly useful in applications of the so-called plastic limit design procedures.

Stress concentrations- stress concentration factor

If a bar of uniform cross-section is subjected to an axial tensile or compressive load the stress is
assumed to be uniform across the section. However, in the presence of any sudden change of
section, hole, sharp corner, notch, keyway, material flaw, etc., the local stress will rise
significantly. The ratio of this stress to the nominal stress at the section in the absence of any of
these so-called stress concentrations is termed the stress concentration factor.

Toughness

Toughness is defined as the ability of a material to withstand cracks, i.e. to prevent the transfer
or propagation of cracks across its section hence causing failure. Two distinct types of toughness
mechanism exist and, in each case, it is appropriate to consider the crack as a very high local stress
concentration.
The first type of mechanism relates particularly to ductile materials which are generally regarded
as tough. This arises because the very high stresses at the end of the crack produce local yielding
of the material and local plastic flow at the crack tip. This has the action of blunting the sharp tip
of the crack and hence reduces its stress concentration effect considerably (Fig. 1.15).

Fig. 1.15. Toughness mechanism-type.

The second mechanism refers to fibrous, reinforced or resin-based materials which have weak
interfaces. Typical examples are glass-fiber reinforced materials and wood. It can be shown that
a region of local tensile stress always exists at the front of a propagating crack and provided that
the adhesive strength of the fiber/resin interface is relatively low (one-fifth the cohesive strength
of the complete material) this tensile stress opens up the interface and
produces a crack sink, i.e., it blunts the crack by effectively increasing the radius at the crack tip,
thereby reducing the stress-concentration effect (Fig. 1.16). This principle is used on occasions to

6
stop, or at least delay, crack propagation in engineering components when a temporary "repair" is
carried out by drilling a hole at the end of a crack, again reducing its stress-concentration effect.

Fig. 1.16. Toughness mechanism-type 2.

FATIGUE AND CREEP


In the preceding paragraphs it has been suggested that failure of materials occurs when the ultimate
strengths have been exceeded. Reference has also been made in §1.15 to cases where excessive
deformation, as caused by plastic deformation beyond the yield point, can be considered as a criterion
for effective failure of components. This chapter would not be complete, therefore, without reference
to certain loading conditions under which materials can fail at stresses much less than the yield stress,
namely creep and fatigue.
Creep is the gradual increase of plastic strain in a material with time at constant load. Particularly at
elevated temperatures some materials are susceptible to this phenomenon and even under the constant
load mentioned strains can increase continually until fracture. This form of fracture is particularly
relevant to turbine blades, nuclear reactors, furnaces, rocket motors, etc.

7
The general form of the strain versus time graph or creep curve is shown in Fig. 1.17 for two typical
operating conditions. In each case the curve can be considered to exhibit four principal features.
(a) An initial strain, due to the initial application of load. In most cases this would be an
(b) A primary creep region, during which the creep rate (slope of the graph) diminishes.
(c) A secondary creep region, when the creep rate is sensibly constant.
(d) A tertiary creep region, during which the creep rate accelerates to final fracture.
It is clearly imperative that a material which is susceptible to creep effects should only be subjected
to stresses which keep it in the secondary (straight line) region throughout its service life. This
enables the amount of creep extension to be estimated and allowed for in design. Fatigue is the failure
of a material under fluctuating stresses each of which is believed to produce minute amounts of
plastic strain.
Fatigue is particularly important in components subjected to repeated and often rapid load
fluctuations, e.g., aircraft components, turbine blades, vehicle suspensions, etc. Fatigue behavior of
materials is usually described by a fatigue life or S-N curve in which the number of stress cycles N
to produce failure with a stress peak of S is plotted against S. A typical S-N curve for mild steel is
shown in Fig. 1.18. The particularly relevant feature of this curve is the limiting stress S, since it is
assumed that stresses below this value will not produce fatigue failure however many cycles are
applied, i.e., there is infinite life. In the simplest design cases, therefore, there is an aim to keep all
stresses below this limiting level. However, this often implies an over-design in terms of physical
size and material usage, particularly in cases where the stress may only occasionally exceed the
limiting value noted above. This is, of course, particularly important in applications such as aerospace
structures where component weight is a premium. Additionally, the situation is complicated by the
many materials which do not show a defined limit, and modern design procedures therefore
rationalize the situation by aiming at a prescribed, long, but finite life, and accept that service stresses
will occasionally exceed the value S, it is clear that the number of occasions on which the stress
exceeds S, and by how much, will have an important bearing on the prescribed life and considerable
specimen, and often full-scale, testing is required before sufficient statistics are available to allow
realistic life assessment. The importance of the creep and fatigue phenomena cannot be over
emphasized and the comments above are only an introduction to the concepts and design
philosophies involved.

8
FAILURE
INTRODUCTION
The failure of engineering materials is almost always an undesirable event for several reasons;
these include human lives that are put in jeopardy, economic losses, and the interference with the
availability of products and services. Even though the causes of failure and the behavior of
materials may be known, prevention of failures is difficult to guarantee. The usual causes are
improper materials selection and processing and inadequate design of the component or its misuse.
It is the responsibility of the engineer to anticipate and plan for possible failure and, in the event
that failure does occur, to assess its cause and then take appropriate preventive measures against
future incidents.
Topics to be addressed in this chapter are the following: simple fracture (both ductile and brittle
modes), impact fracture testing, the ductile-to-brittle transition, fatigue, and creep. These
discussions include failure mechanisms, testing techniques, and methods by which failure may be
prevented or controlled.

FRACTURE
FUNDAMENTALS OF FRACTURE
Simple fracture is the separation of a body into two or more pieces in response to an imposed
stress that is static (i.e., constant or slowly changing with time) and at temperatures that are low
relative to the melting temperature of the material. The applied stress may be tensile,
compressive, shear, or torsional; the present discussion will be confined to fractures that result
from uniaxial tensile loads. For engineering materials, two fracture modes are possible: ductile
and brittle. Classification is based on the ability of a material to experience plastic deformation.
Ductile materials typically exhibit substantial plastic deformation with high energy absorption
before fracture. On the other hand, there is normally little or no plastic deformation with low
energy absorption accompanying a brittle fracture. The tensile stress–strain behaviors of both
fractures.
‘‘Ductile’’ and ‘‘brittle’’ are relative terms; whether a particular fracture is one mode or the other
depends on the situation. Ductility may be quantified in terms of percent elongation and percent
reduction in area Furthermore, ductility is a function of temperature of the material, the strain rate,
and the stress state.

9
Any fracture process involves two steps—crack formation and propagation—in response to an
imposed stress. The mode of fracture is highly dependent on the mechanism of crack propagation.
Ductile fracture is characterized by extensive plastic deformation in the vicinity of an advancing
crack. Furthermore, the process proceeds relatively slowly as the crack length is extended. Such
a crack is often said to be stable. That is, it resists any further extension unless there is an increase
in the applied stress. In addition, there will ordinarily be evidence of appreciable gross
deformation at the fracture surfaces (e.g., twisting and tearing). On the other hand, for brittle
fracture, cracks may spread extremely rapidly, with very little accompanying plastic deformation.
Such cracks may be said to be unstable, and crack propagation, once started, will continue
spontaneously without an increase in magnitude of the applied stress. Ductile fracture is almost
always preferred for two reasons. First, brittle fracture occurs suddenly and catastrophically
without any warning; this is a consequence of the spontaneous and rapid crack propagation. On
the other hand, for ductile fracture, the presence of plastic deformation gives warning that fracture
is imminent, allowing preventive measures to be taken. Second, more strain energy is required to
induce ductile fracture inasmuch as ductile materials are generally tougher. Under the action of
an applied tensile stress, most metal alloys are ductile, whereas ceramics are notably brittle, and
polymers may exhibit both types of fracture.

DUCTILE FRACTURE
Ductile fracture surfaces will have their own distinctive features on both macroscopic and
microscopic levels. Figure 9.1 shows schematic representations for two characteristic
macroscopic fracture profiles. The configuration shown in Figure 9.1a is found for extremely soft
metals, such as pure gold and lead at room temperature, and other metals, polymers, and inorganic
glasses at elevated temperatures. These highly ductile materials neck down to a point fracture,
showing virtually 100% reduction in area.
The most common type of tensile fracture profile for ductile metals is that represented in Figure
9.1b, which fracture is preceded by only a moderate amount of necking. The fracture process
normally occurs in several stages (Figure 9.2). First, after necking begins, small cavities, or
microvoids, form in the interior of the cross section, as indicated in Figure 9.2b. Next, as
deformation continues, these microvoids enlarge, come together, and coalesce to form an elliptical
crack, which has its long axis perpendicular to the stress direction. The crack continues to grow
in a direction parallel to its major axis by this microvoid coalescence process (Figure 9.2c).
Finally, fracture ensues by the rapid propagation of a crack around the outer perimeter of the neck
(Figure 9.2d), by shear deformation at an angle of about 45 with the tensile axis—this is the angle
at which the shear stress is a maximum. Sometimes a fracture having this characteristic surface
contour is termed a cupand-cone fracture because one of the mating surfaces is in the form of a
cup, the other like a cone. In this type of fractured specimen (Figure 9.3a), the central

10
FIGURE 9.1 (a) Highly ductile fracture in which the specimen necks down to a point. (b)
Moderately ductile fracture after some necking. (c) Brittle fracture without any plastic deformation.

(a ) (b) (c)

Shear
Fibrous

(d ) (e)

FIGURE 9.2 Stages in the cup-and-cone fracture. (a) Initial necking. (b) Small cavity
formation. (c) Coalescence of cavities to form a crack. (d) Crack propagation. (e) Final shear
fracture at a 45-angle relative to the tensile direction.

11
BRITTLE FRACTURE
Brittle fracture takes place without any appreciable deformation, and by rapid crack propagation.
The direction of crack motion is very nearly perpendicular to the direction of the applied tensile
stress and yields a relatively flat fracture surface, as indicated in Figure 9.1c.
Fracture surfaces of materials that failed in a brittle manner will have their own distinctive patterns;
any signs of gross plastic deformation will be absent. For example, in some steel pieces, a series
of V-shaped ‘‘chevron’’ markings may form near the center of the fracture cross section that point
back toward the crack initiation site (Figure 9.5a). Other brittle fracture surfaces contain lines or
ridges that radiate from the origin of the crack in a fanlike pattern (Figure 9.5b). Often, both of
these marking patterns will be sufficiently coarse to be discerned with the naked eye. For very hard
and fine-grained metals, there will be no discernible fracture pattern. Brittle fracture in amorphous
materials, such as ceramic glasses, yields a relatively shiny and smooth surface.
For most brittle crystalline materials, crack propagation corresponds to the successive and repeated
breaking of atomic bonds along specific crystallographic planes; such a process is termed
cleavage. This type of fracture is said to be trans granular (or transcrystalline), because the
fracture cracks pass through the grains. Macroscopically, the fracture surface may have a grainy
or faceted texture (Figure 9.3b), as a result of changes in orientation of the cleavage planes from
grain to grain. This feature is more evident in the scanning electron micrograph shown in Figure
9.6a.
In some alloys, crack propagation is along grain boundaries; this fracture is termed intergranular.
Figure 9.6b is a scanning electron micrograph showing a typical intergranular fracture, in which
the three-dimensional nature of the grains may be seen. This type of fracture normally results
subsequent to the occurrence of processes that weaken or embrittle grain boundary regions.

Figure 9.3 (a) Cup-and-cone fracture in aluminum. (b) Brittle fracture in a gray cast iron.
Figure 9.6 (a) Schematic cross-section profile showing crack propagation through the interior of
grains for transgranular fracture. (b) Scanning electron fractograph of ductile cast iron showing a
transgranular fracture surface. Magnification unknown.

PRINCIPLES OF FRACTURE MECHANICS


Brittle fracture of normally ductile materials, such as that shown in the chapter opening photograph
of this chapter, has demonstrated the need for a better understanding of the mechanisms of fracture.
Extensive research endeavors over the past several decades have led to the evolution of the field
of fracture mechanics. This subject allows quantification of the relationships between material
properties, stress level, the presence of crack-producing flaws, and crack propagation mechanisms.
Design engineers are now better equipped to anticipate, and thus prevent, structural failures. The
present discussion centers on some of the fundamental principles of the mechanics of fracture.
(a)

(b)

FIGURE 9.5 (a) Photograph showing V-shaped ‘‘chevron’’ markings characteristic of


brittle fracture. Arrows indicate origin of crack. Approximately actual size. (From R. W.
Hertzberg, Deformation and Fracture Mechanics of Engineering Materials, 3rd edition.
Copyright 1989 by John Wiley & Sons, New York.

STRESS CONCENTRATION
The measured fracture strengths for most brittle materials are significantly lower than those
predicted by theoretical calculations based on atomic bonding energies. This discrepancy is
explained by the presence of very small, microscopic flaws or cracks that always exist under normal
conditions at the surface and within the interior of a body of material. These flaws are a detriment
to the fracture strength because an applied stress may be amplified or concentrated at the tip, the
magnitude
Figure 9.7 (a) Schematic cross-section profile showing crack propagation along grain boundaries for
intergranular fracture. (b) Scanning electron fractograph showing an intergranular fracture surface.
50×.

FIGURE 9.7 (a) The geometry of surface and internal cracks. (b) Schematic stress
profile along the line X –X in (a), demonstrating stress amplification at crack tip positions.

FRACTURE OF POLYMERS
The fracture strengths of polymeric materials are low relative to those of metals and ceramics. As
a general rule, the mode of fracture in thermosetting polymers is brittle. In simple terms, associated
with the fracture process is the formation of cracks at regions where there is a localized stress
concentration (i.e., scratches, notches, and sharp flaws). Covalent bonds in the network or
crosslinked structure are severed during fracture.
For thermoplastic polymers, both ductile and brittle modes are possible, and many of these
materials are capable of experiencing a ductile-to-brittle transition. Factors that favor brittle
fracture are a reduction in temperature, an increase in strain rate, the presence of a sharp notch,
increased specimen thickness, and, in addition, a modification of the polymer structure (chemical,
molecular, and/or

(a) (b)

Fibrillar bridges Microvoids Crack

FIGURE 9.17 Schematic drawings of (a) a craze showing microvoids and fibrillar bridges,
and (b) a craze followed by a crack. (From J. W. S. Hearle, Polymers and Their Properties,
Vol. 1, Fundamentals of Structure and Mechanics, Ellis Horwood, Ltd., Chichester, West
Sussex, England, 1982.)

microstructural). Glassy thermoplastics are brittle at relatively low temperatures; as the temperature
is raised, they become ductile in the vicinity of their glass transition temperatures and experience
plastic yielding prior to fracture. This behavior is demonstrated by the stress–strain characteristics of
polymethyl methacrylate in Figure 7.24. At 4C, PMMA is totally brittle, whereas at 60C it becomes
extremely ductile.
One phenomenon that frequently precedes fracture in some glassy thermoplastic polymers is
crazing. Associated with crazes are regions of very localized yielding, which lead to the formation of
small and interconnected microvoids (Figure 9.17a). Fibrillar bridges form between these microvoids
wherein molecular chains become oriented. If the applied tensile load is sufficient, these bridges
elongate and break, causing the microvoids to grow and coalesce; as the microvoids coalesce, cracks
begin to form, as demonstrated in Figure 9.17b. A craze is different from a crack in that it can support
a load across its face. Furthermore, this process of craze growth prior to cracking absorbs fracture
energy and effectively increases the fracture toughness of the polymer. Crazes form at highly stressed
regions associated with scratches, flaws, and molecular inhomogeneities; in addition, they propagate
perpendicular to the applied tensile stress, and typically are 5 m or less thick. Figure 9.18 is a
photomicrograph in which a craze is shown.

IMPACT FRACTURE TESTING


Prior to the advent of fracture mechanics as a scientific discipline, impact testing techniques were
established so as to ascertain the fracture characteristics of materials. It was realized that the results
of laboratory tensile tests could not be extrapolated to predict fracture behavior; for example, under
some circumstances normally
ductile metals fracture abruptly and with very little plastic deformation. Impact test conditions were
chosen to represent those most severe relative to the potential for fracture, namely, (1) deformation
at a relatively low temperature, (2) a high strain rate (i.e., rate of deformation), and (3) a triaxial
stress state (which may be introduced by the presence of a notch).

IMPACT TESTING TECHNIQUES


Two standardized tests, the Charpy and Izod, were designed and are still used to measure the
impact energy, sometimes also termed notch toughness. The Charpy V-notch (CVN) technique is
most commonly used in the United States. For both Charpy and Izod, the specimen is in the shape
of a bar of square cross section, into which a V-notch is machined (Figure 9.19a). The apparatus
for making V-notch impact tests is illustrated schematically in Figure 9.19b. The load is applied as
an impact blow from a weighted pendulum hammer that is released from a cocked position at a
fixed height h. The specimen is positioned at the base as shown. Upon release, a knife edge
mounted on the pendulum strikes and fractures the specimen at the notch, which acts as a point of
stress concentration for this high velocity impact blow. The pendulum continues its swing, rising
to a maximum height h, which is lower than h. The energy absorption, computed from the
difference between h and h, is a measure of the impact energy. The primary difference between the
Charpy and Izod techniques lies in the manner of specimen support, as illustrated in Figure 9.19b.
Furthermore, these are termed impact tests in light of the manner of load application. Variables
including specimen size and shape as well as notch configuration and depth influence the test
results.
Both plane strain fracture toughness and these impact tests determine the fracture properties of
materials. The former is quantitative in nature, in that a specific property of the material is
determined (i.e., K Ic). The results of the impact tests, on the other hand, are more qualitative and
are of little use for design purposes. Impact energies are of interest mainly in a relative sense and
for making comparisons—absolute values are of little significance. Attempts have been made to
correlate plane strain fracture toughness’s and CVN energies, with only limited success. Plane
strain fracture toughness tests are not as simple to perform as impact tests; furthermore, equipment
and specimens are more expensive.
FIGURE 9.19 (a) Specimen used for Charpy and Izod impact tests. (b) A schematic drawing of
an impact testing apparatus. The hammer is released from fixed height h and strikes the
specimen; the energy expended in fracture is reflected in the difference between h and the swing
height h. Specimen placements for both Charpy and Izod tests are also shown.
DUCTILE-TO-BRITTLE TRANSITION
One of the primary functions of Charpy and Izod tests is to determine whether or not a material
experiences a ductile-to-brittle transition with decreasing temperature and, if so, the range of
temperatures over which it occurs. The ductile-to-brittle transition is related to the temperature
dependence of the measured impact energy absorption. This transition is represented for a steel by
curve. At higher temperatures the CVN energy is relatively large, in correlation with a ductile mode
of fracture. As the temperature is lowered, the impact energy drops suddenly over a relatively
narrow temperature range, below which the energy has a constant but small value; that is, the mode
of fracture is brittle.

CHAPTER 2: BENDING MOMENT AND SHEAR FORCE

Concept of a Beam
Any member of a machine or structure whose one dimension (length) is very large as cothared to
the other two dimensions (width and thickness) and which can carry lateral or transverse loads in
the axial plane is called a beam. A beam may be of rectangular, square, triangular, hexagonal and
circular, etc. cross-sections. A beam is a very important member in structural mechanics to
withstand the transverse loads. A beam may be made of timber, flitched beam, i.e, timber
reinforced with mild steel strips, steel, and reinforced concrete. The reinforced concrete beams
are mostly used in building construction, bridges and flyovers.

TYPES OF LOADS AND BEAMS


The transverse loading of a beam may consist of
1. Concentrated Loads

2. Distributed Loads
a) Uniformly distributed loads

b) Uniformly varying loads

Beams are classified according to the way they are supported at their ends
Bending Moment
The bending moment (B.M.) at any point along a loaded beam is the algebraic sum of the
moments of all the vertical forces acting to one side of the point about the point. Consider a
simply supported beam AB carrying concentrated loads as shown in Fig. 3.1 Let RA and RB be
the vertical reactions at supports A and B respectively. Consider a section x—x at a distance x
from end A. The clockwise moment at this section due to all the loads acting on the beam to the
left of the section is:

If we consider the forces to the right of the section x—x, then anticlockwise moment is:

For equilibrium of the beam, the B.M. given by equations (1) and (2) are equal. The SI units of
B.M. are N.m or kN.n.

Shear Force
The shear force (S F) at any point along a loaded beam is the algebraic sum of all the vertical
forces acting to one side of the point. Thus, for the beam AB shown in Fig. 3.1, the shear force at
cross-section x—x as measured from the left-hand side is:

The shear force as measured from the right-hand side is:

Since the beam is in equilibrium, the S.F. given by equations (1) and (2) are equal.
Sign conventions

B.M. AND S.F. DIAGRAMS

B.M. Diagram
To draw the B.M.D, the following procedure may be followed:
1. Take a sheet of graph paper. Draw the beam along with loading to an appropriate scale.
2. Calculate the reactions at the supports by applying the equations of equilibrium, i.e., =0, EF, 0
and EM0 =0.
3. Choose a section x-x at a distance x from the left-hand support. The section may be chosen
either after every concentrated load or before the right-hand support. For udi, the section may be
taken within the load.
4. Calculate the B.M. beneath every concentrated load. For udi, the B.M. may be
calculated along the length of the load. Ignore that term which becomes negative
on substituting the value of x.
5. Draw the B.M.D. for the beam on a convenient scale. Of course, sign convention has to be
followed for B.M.
Point of Inflexion. It is the point on the beam in the B.M.P. where the bending moment becomes
zero.
Point of Contraflexure. It is the point in the B.M.D where the B.M. changes slope from an
increasing one to a decreasing one. Contraflexure means opposite and flexure means bending.
Some authors consider point of inflexion and point of contraflexure to be synonymous.

S.F. Diagram
The first three steps for the S.F.D. are the same as for the B.M.D., and need not be repeated.
4. Calculate the S.F. beneath every concentrated load just to the left and to the right. For u.d.I.,
the S.F. has to be calculated along the length of the load.
5. Draw the S.F.D. to a convenient scale using the sign convention.

B.M AND S.F DIAGRAMS FOR A SIMPLY SUPPORTED BEAM


Uniformly Distributed Load
Consider a beam AB of span 1 simply supported at the ends and carrying a uniformly distributed
load of intensity w per unit length as shown in Fig. 3.5 (a)
Uniformly Varying Loads
Consider a simply supported beam AB of span 1 carrying udi which varies from w1 per unit
length at end A to w2 per unit length at end B as shown in Fig. 3.6 (a). The varying load can be
considered as the sum of two loads, one of uniform intensity w1 and the other triangle variation
from zero to (w2 — w1).
SIMPLY SUPPORTED BEAM SUBJECTED TO A COUPLE
CANTILEVER BEAM
Concentrated Load at Free End
Consider a cantilever beam AB of span 1 carrying a concentrated load P at the free end

It represents a straight line giving linear variation. The S.F.D is shown in Fig. 3.10 (c).

Uniformly Varying Load


Consider a cantilever beam AB of spand 1 carrying a uniformly varying load of intensity zero at
the free end to O at the fixed end as shown in Fig. 3.11(a). Consider a section x-x at.
OVERHANGING BEAMS
Concentrated Loads

This is a negative B.M. as it produces tension on the top fibres. It is a linearly varying
B.M.
Span AB : At a distance x from C near to B,

3.7.2 Uniformly Distributed Load


UnIformly Varying Load
Consider a beam A13CD with equal overhangs on both sides of the supports and carrying
uniformly varying load from zero at end A too per unit length at end D as shown
It represents a cubic curve. The B.M. is negative as it produces tension on the top fibres of the
beam.

Span CD:
At a distance x from D, we have
The S.F. is parabolic in nature. The S.F.D. is shown in Fig. 3.14 (c).

RELATIONSHIP BETWEEN LOAD, SHEAR FORCE AND BENDING MOMENT


Consider a simply supported beam carrying audi of intensity w per unit length as shown in Fig.
3.15. Consider an elementary length of the beam of length ox between cross-
Sections

Therefore, the first derivative of shear force with respect to x at a point gives the
intensity of loading at the point.

Therefore, the rte of change of bending moment with respect to x is equai to the shear force.
Whenever, bending moment is maximum or minimum, the shear force is zero.
Taking the derivative again, we get

Example 3.1 Draw the bending moment and shear force diagrams for the simply
supported beam loaded as shown in Fig. 3.16(a).
Example 3.2 Draw the bending moment and shear force diagrams for the simply supported
beam shown in Fig. 3.17(a).
.
The S.F.D. is shown in Fig. 3.18 (c) to a scale of 1 mm = 0.5 kN.

CHAPTER 3: PURE BENDING

Pure bending is a condition of stress where a bending moment is applied to a beam without the
simultaneous application of axial, shear, or torsional forces. Pure bending is the flexure
(bending) of a beam under a constant bending moment (M) therefore pure bending only occurs
when the shear force (V) is equal to zero since dM/dx= V. In reality, this state of pure bending
does not practically exist, because such a state needs an absolutely weightless member. The state
of pure bending is an approximation made to derive formulas.

Kinematics of pure bending

1. In pure bending the axial lines bend to form circumferential lines and transverse lines
remain straight and become radial lines.
2. Axial lines that do not extend or contract form a neutral surface.

Assumptions made in the theory of Pure Bending

1. The material of the beam is homogeneous and isotropic.


2. The value of Young's Modulus of Elasticity is same in tension and compression.
3. The transverse sections which were plane before bending, remain plane after bending also.
4. The beam is initially straight and all longitudinal filaments bend into circular arcs with a
common centre of curvature.
5. The radius of curvature is large as compared to the dimensions of the cross-section.
6. Each layer of the beam is free to expand or contract, independently of the layer, above or
below it.
When a bar is subjected to a pure bending moment as shown in the figure it is observed that axial
lines bend to form circumferential lines and transverse lines remain straight and become radial
lines.

In the process of bending there are axial line that do not extend or contract. The surface
described by the set of lines that do not extend or contract is called the neutral surface. Lines on
one side of the neutral surface extend and on the other contract since the arc length is smaller on
one side and larger on the other side of the neutral surface. The figure shows the neutral surface
in both the initial and the bent configuration.

The axial strain in a line element a distance y above the neutral surface is given by

where is the radius to the neutral surface.

Stress distribution in pure bending:

By Hooke’s law, the axial stress is given in terms of the axial strain by the relation
Therefore, the axial stress is zero on the neutral surface and increases linearly as one moves
away from the neutral axis.

Axial load and the location of the neutral axis:

There is zero axial load in a member under pure bending. Therefore, the axial load generated by
the stress should be zero. The axial load generated by the stress applied on the
area of the cross section is given by the approximate relation

The total load on the cross section can be calculated by integrating this relation over the cross
section. This yields
Since the axial load is zero during pure bending, one concludes that for pure bending

The reader recalls that the location of the centroid of an area is calculated from the relation

Therefore, for the axial load to be zero, the neutral axis must pass through the centroid of the
cross section (i.e., yc=0). In the event that the axial load is not zero, the location of the neutral
axis relative to the centroid of the cross section can be calculated from the relation
Bending moment and its relation to radius of curvature:

The bending moment about the neutral surface that is created by the normal
load resulting from the normal stress acting on the area of the cross section can be
calculated by

Integrating over the cross section to get the total moment transmitted through the cross section
gives

Recalling that the integral in this relation is the area moment of inertial I about the neutral axis
(the line resulting from the intersection of the cross section and the neutral surface), the relation
between the bending moment M and radius of curvature of the neutral axis of the
beam becomes
From this relation one can calculate the expression for stress as a function of the bending
moment by substituting in the expression for axial stress this relation for the radius of curvature.
This gives

As can be seen in the figure, the maximum and minimum normal stresses occur in the material
that is furthest away from the neutral surface (either at the top or bottom of the bar depending on
the actual direction of the moment).
1) Center of gravity and neutral axis.
2) Moment of inertia.
3) Theory of simple bending.
4) Bending of composite or flitched beam.
5) Combined bending and direct stress.
Centre of Gravity
Centroid of an area is the point about which the area could be balanced if it was supported
from that point.

a) Simple shapes

b) Area has an axis of symmetry

c) Two axes of symmetry do not occur


Where AT = total area of the composite shape ȳ = distance to the centroid of the composite
shape measured form some reference axis OX.
Ai = area of one component part the shape.
yi = distance the centroid at the component part the reference axis.

Or

Where x = distance to the centroid of the composite shape measured form some reference axis
OY.
xi = distance the centroid at the component part the reference axis.

2) Moment of Inertia
Moment of inertia is a measure of the resistance of the section to applied moment or load that
tends to bend it.
Moment of inertia depends on the shape and not material. It is a derived property.

I XX = ∑da.y2 = ∫ y2.dA is called the moment of inertia or the second moment of area about the
axis XX.

IYY =∑da.x2 =∫ x2.dA is called the moment of inertia or the second moment of area about the
axis YY.

Parallel axis Theorem

This theorem is mainly used to transfer the moment of inertia of a body from its individual axis
to another reference line.
3) Assumption for the Simple Bending Theory
1. The beam is initially straight and unstressed.
2. The material is homogeneous, same density and elastic properties.
3. The elastic limit is nowhere exceeded.
4. Young’s modulus for the materials is the same in tension and compression.
5. Plane cross–section remains plane before and after bending.
6. No resulted force perpendicular to any cross – section.

Bending theory

• If we apply a constant B.M.


• If will bend to radius R.
• The top fibers of the beam will be subjected to tension.
• The bottom fibers of the beam will be subjected to compression.
• Somewhere between the two there are points at which the stress is zero.
Locus of all points is termed the neutral axis.
The radius of curvature R is then measured to this axis.
The neutral axis will always pass through the centre of area of centroid.
• The maximum tension and compression will be at the outer surface or the farthest distance
from the neutral axis.

Consider fiber AB distance y from the N.A. when the beam is bent this will
stretch to A′B′.

Strain in AB =Extension = A′B′ − AB


Original Length AB

AB = CD
CD = C′D′
Strain ε= A′B′-C′D′
C′D′
C′D′ = Rθ

A′B′ = (R + y)θ

∴ Strain ε= (R + y)θ− Rθ = y
Rθ R
Within the elastic region:

If the strip of area δA

This force has a moment about the N.A of

The total moment for the whole cross section is:

∑ y .δA is the second moment of area of the cross-section.


2
If the beam is of uniform section, the material of the beam is homogeneous and the
applied moment is constant, I, E, M remain constant and hence the radius of curvature of
the beam will also be constant.

EI is known as the flexural rigidity.

• For larger value of R smaller deflection greater the rigidity.

Maximum Stress σmax = M y max


I

is termed the section modulus Z.

M = Zσmax for symmetrical beam

for unsymmetrical sections

4) Bending of Composite or Flitched Beams


Composite beam is one which is constructed from a combination of material (flitched beam).

Forcesteel = (σ.dy.t)steel = (σ′.dy.t′)equivalent


Since the moment at any section must be the same in the equivalent section as in the original, so
that the force at given dy in the equivalent beam must be equal to that at the strip it replaces.
σ.t =σ′.t′

ε.E.t =ε′.E′.t′
Strain must be equal
ε=ε′
E.t = E′.t′

Thus, to replace the steel strip by an equivalent wooden strip the thickness must be
multiplied by the modular ratio .

Where: σ= the stress in the steel section. σ′= the stress in


the equivalent section. t = the thickness of the
steel section.
t′= the thickness of the equivalent

5) Combined Bending and Direct Stress – Eccentric Loading


a) Eccentric loading on one axis
The equation of the N.A can be obtained by setting σ equal to zero.

I
y N =  xx
A.e

The larger the eccentricity of the load the closer the N.A. will be to the axis of
symmetry.
yN is the distance between the N.A. and the axis of symmetry.

b) Eccentric loading on two axis

If the applied load will not be applied on either of the axis of symmetry, there will be a direct
stress effect plus simultaneous bending about both axis.
The total stress at any point (x, y):

Equation of the N.A. is obtained by equating to zero:


This is linear equation in x and y. The line may or may not cut the section.

For rectangular and circular cross sections, provided that the load is applied within certain
defined areas, no tension will be produced whatever the magnitude of the applied compressive
load.
● 27
● 28
● 29
● 30
● 31
● 32
CHAPTER 4: SHEAR STRESS

If the applied load consists of two equal and opposite parallel forces which do not share the same line of action,
then there will be a tendency for one part of the body to slide over, or shear from the other part.
In the figure below, if the section LM is parallel to the forces and has an area A, then the average Shear Stress

. If the Shear Force varies then at a point

Note : The Shear Stress is tangential to the area over which it acts, and is expressed in the same units as Direct
Stress, i.e. Load per unit Area.
See Also the section on Shear Force and Bending Moment

Complementary Shear Stress.


A,B,C,D is a small rectangular element whose sides are x, y and z which are perpendicular to the figure. A Shear
Stress S acts on the planes A,B and C,D
33 ●

It can be seen that these Stresses form a Couple which can only be balanced by tangential Stresses on the
planes A,D and B,C. These are known as Complementary Shear Stresses. Let S' be the Complementary Shear
Stress induced on planes A,D and B,C. Then as the element is in equilibrium:

(1)

(2)
This shows that every Shear Stress is accompanied by an equal Complementary Shear Stress on planes at
right angles. To produce balancing Couples, the direction of the Shear Stress on the element are either both
towards, or both away from a corner.
The existence of Complementary Shear Forces may be an important factor in the failure of anisotropic materials
such as timber, which is weak in Shear along the grain.
Near to a Free Boundary there are no externally applied Forces and it can be seen that The Shear Stress on any
cross section must act in a direction parallel to the Boundary. This is because if there were a component in a
direction at right angles to boundary it would require a Complementary Shear Stress on the Boundary wall. For
Example, the Shear Stress distribution over a section of a rivet must be as in the left-hand diagram and NOT as on
the right.

This produces complications in that the Shear Stress varies in both magnitude and direction, although in the
particular case of rivets, it is not normal to make any allowance for this in the design.
Example: 1

Example - Shear stress on shaft bolts


Problem
A Flange Coupling which joins two sections of a Shaft is required to transmit 250 h.p. at 1000 r.p.m. If six bolts
are used on a pitch circle of 6 ins. find the diameter of the bolts. The allowable mean Shear Stress is 5 tons/in2.
Workings
● 34
Torque to be transmitted s given by:

(1)

(2)
If d is the diameter of a bolt the load carried by one bolt is:

(3)
Multiplying by the number of bolts and their radius arm, the torque carried is:

(4)

(5)

(6)
Solution
Diameter of bolt,

The distortion produced by Shear Stress on an element or Rectangular Block is shown in the following diagram.
● 35
The Shear Strain or "Slide" is and can be defined as the change in the right angle. It is measured in Radians
and is dimensionless.

The Modulus of Rigidity


For Elastic materials it is found that within certain limits, Shear Strain is proportional to the Shear Stress
producing it.

The Ratio is called the Modulus of Rigidity and is denoted by C.

and in the Imperial system is in

Strain Energy
Within the limits of proportionality Strain is proportional to Stress and:
Strain Energy (U) = Work done in Straining.

(3)
For a gradually applied stress the work done is shown on the diagram as the shaded area.

(4)

(5)
● 36
CHAPTER 5 TORSION

Torsion refers to the twisting of long members. Torsion can occur with members of any cross-sectional shape, but
the most common is the circular shaft. Another fairly common shaft configuration, which has a simple solution, is
the hollow, thin-walled shaft.

Circular Shafts
Fig. 9.10(a) shows a circular shaft before loading; the r-q-z cylindrical coordinate system is also shown. In
addition to the outline of the shaft, two longitudinal lines, two circumferential lines, and two diametral lines are
shown scribed on the shaft. These lines are drawn to show the deformed shape loading. Fig. 9.10(b) shows the
shaft after loading with a torque T. The double arrow notation on T indicates a moment about the z axis in a
right-handed direction. The effect of the torsion is that each cross-section remains plane and simply rotates with
respect to other cross-sections. The angle f is the twist of the shaft at any position z. The rotation f(z) is in the q
direction.
The distance b shown in Fig. 9.10(b) can be expressed as b = fr or as b = g z. The shear strain for this special
case can be expressed as

For the general case where f is not a linear function of z the shear strain can be expressed as

df/dz is the twist per unit length or the rate of twist.

Figure 9.10 Torsion in a circular shaft

The application of Hooke’s Law gives

The torque at the distance z along the shaft is found by summing the contributions of the shear stress at each
point in the cross-section by means of an integration
● 37
where J is the polar moment of inertia of the circular cross-section. For a solid shaft with an outer radius of ro
the polar moment of inertia is

For a hollow circular shaft with outer radius ro and inner radius ri, the polar moment of inertia is

Note that the J that appears in Eq. (9.43) is the polar moment of inertia only for the special case of circular
shafts (either solid or hollow). For any other crosssection shape, Eq. (9.43) is valid only if J is redefined as
a torsional constant not equal to the polar moment of inertia. Eq. (9.42) can be combined with Eq. (9.43) to
give

The maximum shear stress occurs at the outer radius of the shaft and at the location along the shaft where the
torque is maximum.

The angle of twist of the shaft can be found by integrating Eq. (9.43)

For a uniform circular shaft with a constant torque along its length, this equation becomes

Example 9.10
The hollow circular steel shaft shown in Exhibit 12 has an inner diameter of 25 mm, an outer diameter of
50 mm, and a length of 600 mm. It is fixed at the left end and subjected to a torque of 1400 N • m as
shown in Exhibit 12. Find the maximum shear stress in the shaft and the angle of twist at the right end.
Take G = 84 GPa.
● 38

Exhibit 12

Solution

Hollow, Thin-Walled Shafts


In hollow, thin-walled shafts, the assumption is made that the shear stress τsz is constant
throughout the wall thickness t. The shear flow q is defined as the product of τsz and t. From a
summation of forces in the z direction, it can be shown that q is constant—even with varying
thickness. The torque is found by summing the contributions of the shear flow. Fig. 9.11 shows
the cross-section of the thin-walled tube of nonconstant thickness. The z coordinate is
perpendicular to the plane of the paper. The shear flow q is taken in a counter-clockwise sense.
The torque produced by q over the element ds is

Figure 9.11 Cross-section of thin-walled tube

The total torque is, therefore,


● 39

The area dA is the area of the triangle of base ds and height r,

So that

where Am is the area enclosed by the wall (including the hole). It is best to use the
centerline of the wall to define the boundary of the area, hence Am is the mean area.
The expression for the torque is

and from the definition of q the shear stress can be expressed as

EXAMPLE
● 40

REFERENCE
1. William D. Callister, Jr. Fundamentals of Materials Science, 2006
2. Mehrdad Negahban and the University of Nebraska, 1996-2000.
All rights reserved Copy and distribute freely for personal use only.
Department of Engineering Mechanics, University of Nebraska, Lincoln, NE 68588-
0526
3. Loremate.com – Strength of Materials - Part 1Chapter 3 : Bending Moment and Shear
Force Diagrams (Part 1 )
4. James R. Hutchinson Strength of MaterialsFundEng_Index.book Page 273
Wednesday, November 28, 2007
5. CodeCogs © 2004-2013 Zyba Ltd Last Modified: 2011-11-21 14:55:09 Page
Rendered: 2013-01-02 05:05:12

You might also like