Sie sind auf Seite 1von 21

An Introduction to Mechanics of

Materials

©
Vijay Gupta

Lovely Professional University, Punjab

1
shear force which is statically equivalent to the sectional shear
distribution must be applied.
 The strain energy in slender beams is dominated by the energy
due to bending stresses and is given by ∫ . / .

Problems

6 Combined Stresses and Strains

6.1 Introduction
We have in the previous chapters discussed cases of structures where the
loading was simple. We started in Chapter 2 with simple axially-loaded
members which resisted the loads through axial stresses. We dealt in
Chapter 3 with shafts subjected to torsional loads such that there were
only shear stresses on the members. In Chapter 5 beams were considered.
Of course in these there were tensile as well as shear stresses but, as was
shown, the shear stresses were much smaller than the bending stresses for
slender beams. In the cases where longitudinal stresses were dominant we
designed structures such that the stresses within the members did not
exceed the safe level of such stresses. This safe level was determined by
subjecting a specimen of the material to tensile test, and was termed as
the tensile strength of the material. Similarly, a shaft was designed so that
the maximum shear stress did not exceed the safe level for the material of
the shaft.
Real structures are not that simple. Consider
the structure shown in Fig. 6.1. The structure
near the root is subjected to both torsion and
bending (in addition to shear) loading. At a
point near the root, there will be all types of
stresses: tensile and shear, and along many Fig. 6.1
directions. How does one analyse this case to
92
determine if the structure is safe? We had further seen in Sec. 1.7 that the shear stresses on complementary
planes are equal in magnitude, so that instead of nine stress components
We had earlier in Sec. 1.6 introduced a double-index notation for the
only six are required. This reduces the number of distinct stress
stress components recognizing three mutually-perpendicular planes and
components to six.
three components of
stresses on each plane. We, in this chapter, present a framework for analysing the structure with
This means that there combined stresses (and strains). The strategy begins by realizing that it is
are a total of nine possible to transform the stresses (and strains) obtained in one set of axes
components of stresses to any arbitrary set of axes, then to determine stresses in those axes where
(Fig. 6.2). the stresses picture is simpler. Once that is done, we can compare these
stresses with the maximum prescribed stresses for the material of the
It will be shown in this
structure.
chapter that given the
nine components of Example 6.1 Wire over a pulley
stress at a point in a
material with three Fig 6.3 shows a steel wire of diameter 1
arbitrary mutually- mm that goes over a pulley of diameter
perpendicular planes, it 500 mm and carries a load of 100 N as
is possible to determine Fig. 6.2 Stress components shown. Determine the stress distribution
the stress components in the wire cross-section at AA.
on any other three mutually-perpendicular planes. This is termed as
transformation of stresses and is a very useful tool in the analysis of Solution:
structures. The section at AA is subjected to bending
Thus, the state of stress at any given location in a structure can be fully as well as direct pull by the load of 100
Fig. 6.3 Wire over a pulley
described by specifying the three vectors (with three components each) on N. The bending stresses are evaluated
any arbitrary three mutually perpendicular faces (specified by the by noting that the radius of curvature of the bent wire is (500 + 1)/2 mm.
directions of their outward normals). We conventionally express this in Using this value of ρ in Eq. 5.6, we get
the form of a matrix. This matrix73 has the three stress vectors as its three
rows.
The maximum value of this stress is at y = 0.0005 mm. Thus, the
| | maximum bending stress in the wire is 399.2 MPa, compressive on the
inside, and tensile on the outside as shown.
The stress due to the direct load is tensile everywhere equal to 100 N
divided by the re of the wire, or 100 N/(π 2
m2) = 127.3 kN.

73
The maximum compressive stress on the inside, thus, is (399.2 – 127.3)
This matrix is known as stress tensor. Just as a vector has a direction or 171.9 MPa, and the maximum tensile stress on the outside is (399.2 +
associated with it, a tensor has two directions associated with it. Whereas each
element in a vector has an associated coordinate direction, each element of a
tensor has two associated directions, one, the direction of the normal to the plane
on which it acts, and the other, the direction of the stress component itself.
93
127.3) or 426.5 MPa. This level of tensile strength is quite high and Solution:
probably exceeds the tensile strength of most steel wires74.
The hoop stress ζθθ for a thin closed cylinder was evaluated in Sec. 1.8 as
Example 6.2 A spirally-wound cylindrical vessel75 pr/2t ( ), whereas the longitudinal stress ζzz was
evaluated as pr/t ( ).
Consider a spirally wound cylindrical vessel of diameter 40 mm (Fig.
6.4a), much in the fashion of the cardboard tube discussed in Example As was discussed in Example 1.6, the angle θ of the spiral can be
1.6. The vessel is obtained by imagining that as the brass strip is wound up, the strip should
made from a thin advance in the axial direction a distance equal to the width of the strip as
copper strip of width we go around the tube through one circumference, πD as shown in Fig.
30 mm wound 6.4c. This angle θ is, therefore,
spirally. The edges
of the strip are brazed .
together. The
diameter of the Following the strategy adopted in Example 1.6 let us consider an
cylinder is 40 mm infinitesimal element ABC at the seam of the cylinder where the side AB
and the wall is along the circumferential (θ) direction, BC is along the z-direction, and
thickness is 2 mm. AC coincides with the seam (Fig. 6.4b). Clearly, is the angle θ
What is the shear Fig. 6.4 Stresses in a closed cylinder evaluated above. Let the area of the face δA be taken as unity. The area
strength of the brazed of face AC is then δAtanθ, and that of face AC is δAcosecθ. Equilibrium
joint if the internal pressure that this cylinder can withstand is 3 MPa. of the element requires:
( )( ) ( )( )
( )( ) ( )( )
On simplification we get
74
Suppose we increase the wire diameter to 2 mm. Would the stresses and
decrease? Let us find out. Since the radius of curvature of the wire over
the pulley has changed little (from 0.2505 m to 0.251 m) the bending ( )
stresses increases due to change in the value of y (from 0.0005 m to 0.001 For the values of ζzz, ζθθ and θ calculated above, the values of ζnn and ηnt
m). Thus, the maximum value of bending stress changes to 798.4 MPa. are obtained as 23.2 MPa and = 7.5 MPa, respectively. Thus, the strength
Since the area of cross-section of the wire increases by a factor of 4, the of the brazed material in shear is just 7.5 MPa.
direct stress becomes one-fourth of the earlier value, or 31.8 MPa. Thus,
the maximum stress in the wire now is 830.2 MPa tensile! A very
interesting result, indeed. Increasing the wire diameter result in the 6.2 Plain stresses
increase in stress level, and the wire is sure to fail now. As mentioned earlier, we require a total of six stress components to
We need to decrease the wire diameter to make it safer (if we cannot specify the state of stress at a location in a structure. But in all the
increase the pulley diameter). examples we have dealt with so far not all components are present. This
fortunately is the situation in many problems of interest. It is seen that
75
This problem can be solved in a straight forward manner by using the Mohr
circle approach introduced in Sec 6.4
94
many of the stress components are either absent or are so small (in Consider the material element shown under plane stress in Fig. 6.6a. The
comparison to the other components) that they can be neglected. stresses acting on the x- and y- planes are shown. The equivalence of
shear stresses on complementary planes requires ηxy = ηyx. As the material
Let us first consider the simpler case of plane stresses. A structure (or a
is under plane stress there are no z-components of stresses. Let us now
structural element) is said to be under the condition of plane stress if all
attempt to determine the stresses on planes x’ and y’, rot ted through an
the stress vectors are within
arbitrary angle θ, counter-clockwise as shown in Fig. 6.6b. Let us, for this
one plane. This implies that
purpose isolate an infinitesimal wedge of material as shown in Fig. 6.6c.
there are no stress
Here the faces AB (of length δx) and BC (of length δy) are aligned along
components in one the
the x- and y- directions. The face AC is aligned along the x’-direction at
coordinate direction
angle θ, and hence is a y’-plane. Its length δl is clearly equal to δx/cosθ or
perpendicular to that plane.
δy/sinθ. Fig. 6.6c shows the forces on each face (of a unit length in the z-
The lamina of Fig. 6.5 is
direction). The equilibrium of the element gives:
under the state of plane
stress since there are no ∑ ( )
components in the z- Fig. 6.5 Plane stress ( ) ( ) ( )
76
coordinate direction . (a)
Further, the use of equivalence of shear stresses on complementary planes ∑ ( )
(i.e., ηxy = ηyx) reduces to number of distinct stress components in the case
( ) ( ) ( )
of plane stress to three: ζxx, ζyy and ηxy.
(b)
Replacing δl with δx/cosθ or δy/sinθ where appropriate, we can simplify
6.3 Transformation of plane stresses these two equations to:
We will now show that given the three stress components ζxx, ζyy and ηxy at ( )
( )
a point on two mutually perpendicular planes x and y in the case of plane
(c)
stresses, it is possible to determine the stresses on any other plane at that
point. We have already done such an exercise for the simpler cases of
uniaxial stress (in Sec. 1.6) and for the case of plane stress without the (d)
shear stress component ηxy ( in example 6.2). We follow exactly the same
approach here. We could, in a similar fashion, draw up a wedge of material which
exposes the x’-face, and obtain a similar equation for the stress
component ζx’x’:

(e)
Eqs. a-c are conventionally cast in the following form77 which lead
directly to a graphical construction known as Mohr circle:

76
It is interesting to note that the shear stress components on the z-plane, ηzx and
ηzy, which are not in the z-direction, must also vanish in the case of plane stress.
Why? (a) (b) (c) 77
We have used sin2θ and cos2θ relations in Eqs. a-c for the purpose.
Fig. 6.6 Transformation of plane stress 95
. / . / 1. The plane X is plotted with coordinates (ζxx, −ηxy) and plane Y is
plotted with coordinates (ζyy, +ηxy). Note the difference in the
(6.1) sign of the shear stress for the two planes.
. /
(6.2) For the purpose of understanding this sign convention, let us first
introduce the concept of x-like axes and y-like axes. The convention lays
. / . / (6.3) down that a positive y-like direction is 90o clock-wise from the positive x-
like direction. Thus, all axes marked as A in Fig. 6.8 are x-like, while all
The significance of this result is that it shows that it is possible to axes marked as B are y-like.
calculate the stress on any arbitrary plane x’ or y’ given the stresses on
two mutually perpendicular planes x and y.
In 1882, Otto Mohr, a French engineer devised a graphical procedure to
implement the above transformation which is quite useful. It is described
below.

6.4 Mohr circle Fig. 6.8 A is x-like while B is y-like in each set

A Mohr circle is a special construction which helps in determination of The convention for plotting of shear stresses is the following:
stress vectors on variously oriented planes. It is a plot of the components
A positive shear stress on an x-like face78 is plotted downward
of stress vectors on differently oriented planes at a location. The tensile
while a positive stress on a y-like face is plotted upwards.
stresses are plotted along the horizontal axis while the shear stresses are
Reverse for the negative shear stresses.
plotted along the vertical axis.
2. Join the line XY. Let it intersect the ζ axis at point C. With C as
Consider the state of stress given in Fig. 6.7.
centre and CX or CY as radius draw a circle as shown. This is the
Mohr circle. It should be easy to see from the construction of Fig.
6.8 that the co-ordinates of the centre C are [(ζxx + ζyy)/2, 0], and

the radius of the circle is √. /

78
Fig. 6.7 Mohr circle An x-like face is a face with its outward normal along the x-like axis.
96
Thus we see that the coordinates of the points X’ and Y’ give the
components of the stresses on the rotated planes following the
sign convention stated above79.
Let us summarize the graphical construction of Mohr circle to determine
the stress components on arbitrary planes:
 Plot the stress components on x- and y- planes as points X and Y on the
ζ-η plane using the convention that a positive shear stress on an x-like
face is plotted downward while a positive stress on a y-like face is
plotted upwards. Reverse for the negative shear stresses.
 Join points X and Y to locate the centre C of the circle on the ζ-
axis, and draw the Mohr circle with CX (or CY) as radius.
 To determine the stress components on planes which are at a
(a) (b)
clockwise angle θ from the original faces, rotate the diameter
Fig. 6.9 Using Mohr circle to obtain stress components in arbitrary directions YCX through an angle 2θ to obtain points X’ and Y’.
 The coordinates of points X’ and Y’ give the stress components on
3. To obtain stress components on a set of planes represented by
the new planes with normals in the X’ and Y’ directions. The
outward normals in x’- and y’ directions obtained by rotating the
signs of the shear stress components are to be interpreted using
x’- and y’- axes clockwise through an angle θ, rotate the shaded
the convention given above.
triangles clockwise through an angle 2θ (twice the angle through
which the axes are rotated. The points X’ and Y’ so obtained
represent the stress components on x’- and y’- planes. This is 6.5 Principal planes, principal stresses and
shown in Fig. 6.10.
Maximum shear stresses
The horizontal coordinate of point X’ is seen to be
Consider a general plane stress situation as depicted in Fig. 6.10a. The
. Noting that . /, corresponding Mohr circle is drawn as Fig. 6.10b. The stress components
. /, and , we see that this is equal to ζx’x’ as on the x-plane (ζxx, ηxy) are plotted as point X with coordinates (ζxx, −ηxy).
The negative sign with the shear stress is because f the fact that this is an
given by Eq. 6.1. x-like plane. The y-plane is plotted as point Y with coordinates (ζyy, +ηxy)
Similarly, the horizontal coordinate of point Y’ is seen to be with the positive sign with the shear stress because this is a y-like plane.
. Noting that . /, The centre is located, as before, at [(ζxx + ζyy)/2, 0], and the radius of the

. /, and , we see that this is equal to ζy’y’ circle is √. / .


as given by Eq. 6.3.
The vertical coordinate of point X’ is seen as
= . / , which is the 79
Namely, a positive shear stress on an x-like face79 is plotted downward while a
negative of as given by Eq. 6.2. positive stress on a y-like face is plotted upwards. Reverse for the negative shear
stresses.

97
As we move along the circle from point X and arrive at another point X’, It is easy to show that the sum of principal stresses ζ1 and ζ2 is equal to
say angle θ clockwise from point X, the magnitude of the vertical the sum of the two original tensile stresses81 ζxx and ζyy. One can verify
coordinate decreases till we reach the point P1. The point X represents a from Fig. 6.10b that
plane whose outwards normal is inclined at an angle θ/2 (half the angle
,
on the Mohr circle). Thus, the (positive) shear stress on this x-like plane
decreases till it vanishes at point P1. This point P1 represents a plane or
which has a special significance and is termed as a principal plane which
is defined as a plane that has no shear stress component acting on it. This . / √. / (6.4)
point is at an angle θP from the point X on the Mohr circle, and represents
a plane which is at angle θP/2 from the plane X on the physical diagram and . / √. / (6.5)
shown as Fig. 6.10c. The point P2 on the Mohr circle is also a principal
plane. There are two such planes, 90o apart. Note that the tensile stress ζ1 Planes S1 and S2 represent to other significant planes. These are the
planes with maximum values of shear stress. The maximum value of the
shear stress is

√. /
(6.6)
These planes of maximum shear also carry tensile stress equal to
. / in each case.

Given below are a few examples on the transformation of stresses which


illustrate some significant results.

Example 6.3 Uniaxial tension


Consider a two-force member such as the cross-link ab of the step ladder
shown in Fig. 6.11a. The member experiences only an axial tension.
Away from the two ends, the material is under uniform tensile stress, say
ζxx (St. Ven nt’s principle: Section 1.4). Draw the Mohr circle for the
state of stress within this link, and determine the principal stresses and the
maximum shear stress (and the planes they act on).
Fig. 6.10 Principal planes and planes of maximum shear
is the maximum possible tensile stress and ζ2 is the minimum possible
tensile stress in any plane at the given point80. These stresses are termed
solids. We shall see that brittle materials fail when the tensile stress at any point
as the two principal stresses at this point. in any direction exceeds the ultimate tensile strength of the material. Thus, to see
if a brittle structural element is safe, we just need to ensure that the maximum
principal stress is less than the ultimate tensile strength.
80 81
The fact that principal stress ζ1 is the maximum tensile stress on any plane is In fact, one can establish easily from the Mohr circle that the sum of tensile
the reason why principal stresses play such an important role in mechanics of stresses on any two mutually perpendicular planes is the same.
98
Solution: Figs. 6.11e and f illustrate a very interesting point. The first one shows a
rod made of ductile material after failure when subjected to a pure tensile
Fig. 6.11b shows a small element of this link with the x- and y- faces
load, while the second one shows the fracture of a brittle material under
exposed, and all the stresses acting on it.
tensile load. The cup-and-cone failure of the ductile material indicates
The stress components on the x-plane are (ζo, 0). This is plotted as the that this material has failed in shear, because shear is the maximum on a
point X in Fig. 6.11c. Point Y has the coordinates (0, 0), the stress plane inclined at 45o to the axis. It is quite a general property of the
components on the y-plane. The centre C is midway point on the ζ-axis ductile materials: these are weaker in shear than in tension.
with coordinates (ζo/2, 0). The Mohr circle is drawn with the centre at C
The failure of the brittle material (Fig. 6.11f) perpendicular to the axis
and radius equal to length CX, that is, ζo/2.
indicates it has failed along the principal plane P1, i.e., it has failed in
We note from the Mohr circle that x- and y-planes are the principal planes tension. Brittle materials are weaker in tension than in shear.
themselves (no shear stress), and the maximum shear points S1 and S2 are
90o clockwise from the X and Y points on the Mohr circle. This implies Example 6.4 Simple axial compression load
that physically these are 45o from the x- and y-planes. The stress Two-force compressively loaded structures are very commonly
components on these planes are shown in Fig. 6.11d. encountered. The vertical columns of the elevated water tank of Fig.
6.12a have compressive stresses at section ab. Draw the Mohr circle for
the state of stress within this column, and determine the principal stresses
and the maximum shear stress (and the planes they act on).

Fig. 6.11 Uniaxial stresses

Each of these planes has a tensile stress as well as a shear stress82, all of
magnitude equal to ζo/2.

82
It is interesting to note that the plane S1 is an x-like plane. Positive shear stress plotted upwards, it must be negative, or in the negative S2-direction, as the
on an x-like plane is plotted downwards. Since the shear stress on this plane is outward normal to this plane is in the positive-direction.
99
lying in the negative ζ space), the material fails in shear. The failure lines
are largely inclined to the vertical axis.

Example 6.5 Hydrostatic Loading


Recall from the elementary fluid statics that in a fluid at rest there are no
shear stresses and the pressure at a point is invariant with the direction of
the plane. Let us consider a structural element subject to compressive
stress ζo in both x- and y- directions and with no shear stress ηxy. Draw the
Mohr circle and discuss the state of stress.

Fig. 6.12 Simple uniaxial compressive loading

Solution:
Fig. 6.12b shows a small element of this column with the x- and y- faces
exposed, and all the stresses acting on it.
Fig.6.13 Hydrostatic loading
The stress components on the y-pl ne re (−ζo, 0). This is plotted as point
Y in Fig. 6.11c. Point X has the coordinates (0, 0), the stress components Solution:
on the x-plane. The centre C is the midway point on the ζ-axis with
In the absence of shear stress ηxy, both the planes X and Y have the
coordin tes (−ζo/2, 0). The Mohr circle is drawn with the centre at C and
coordin tes (− ζo, 0) and are plotted at the same location on the stress
radius equal to length CX, that is, ζo/2.
graph. Clearly, the Mohr circle here is reduced to a single point.
The x- and y-planes are the principal planes themselves (no shear stress), Consequently, there are no shear stresses on any plane, and the
and the maximum shear points S1 and S2 are 90o clockwise from the X compressive stress is ζo in all directions.
and Y points on the Mohr circle. This again implies that, physically, these
are 45o from the x- and y-planes. The stress components on these planes Example 6.6 A shaft under pure torsion
are shown in Fig. 6.12d. Each of these planes has a tensile stress as well Consider a circular shaft subjected to pure torsion as shown in Fig. 6.14a.
as a shear stress, all of magnitude equal to ζo/2. Note that the plane S1 is It was shown in Chapter 3 that it is convenient to work in the polar
an x-like plane. Positive shear stress on an x-like plane is plotted coordinates r, θ and z, and that the only component of stress in these
downwards. Since the shear stress on this plane is plotted upwards, it coordinate planes is ηzθ (or ηθz). Determine the principal axes and the
must be negative, or in the negative S2-direction, as the outward normal principal stresses in the shaft.
to this plane is in the positive-direction
Figs. 6.12e illustrates the mode of failure in a brittle material. Since most
brittle material are weak in tension (but not in compression), and since
there are no tensile stresses in this case (the whole of the Mohr circle
100
weaker in tension than in shear. Therefore it breaks83 along the principal
plane P1. A ductile material, on the other hand is weaker in shear, and
therefore it fractures along the plane carrying the maximum shear, i.e., the
z-plane itself.

Example 6.7
Fig. 6.15 shows the stress components on an aircraft panel loaded in its
own plane. Determine the orientation of principal planes and the
principal stresses. Determine also the maximum shear stresses acting at
this point of the panel.

Fig. 6.14 Stresses in a circular shaft

Solution:
As was seen in Chapter 3, the maximum shear stress occurs in an element
at the surface of the shaft where ( ) , where R is the radius of the
shaft, T is the twisting torque, and Izz is the polar moment of the shaft
( ) Let this be denoted by ηo as shown in Fig.
6.14b.
We note that the stress components on the z-pl ne re (0, −ηo). Since this
is an x-like plane, the shear stress will be plotted upwards. Point Z in Fig.
6.14c represents this plane. Similarly, the θ- plane which is a y-like plane Fig. 6.15
nd h s the s me stress components (0, −ηo) is plotted downwards as point
Θ. The Mohr circle is drawn with centre at origin and with radius equal to Solution:
ηo. Clearly, points P1 and P2 represent the principal planes, there being The coordinates of point X representing the x-plane on the stress diagram
no shear stress on them. These are 90o away clockwise from the Z- and are (100, -40), this being an x-like plane. Similarly, the coordinates of
Θ- points, respectively on the Mohr circle, or 45o clockwise fron z- and θ- point Y are (30,+40). The coordinates of the centre C are [(110 + 50)/2,
axes in the physical plane as shown in Fig. 6.14c.
Figs. 6.14 d and e show the failure of brittle and ductile materials,
respectively. As was discussed in Example 6.4 above, a brittle material is
83
Looking at the fracture line of the brittle shaft in Fig. 6.14d, can you determine
if the torque this shaft was subjected to was positive or negative?
101
0], or (80, 0). The radius of the Mohr circle is √. / , or

√. / 50, all in MPa. The principal stresses are:

- and
- - -
For the first principal plane, angle 2θ = .
Thus, the normal to the first principal plane inclined at an angle of
53.13o/2, or 26.6o clockwise to the x-axis. The normal to the second (a) (b) (c)
principal plane is likewise inclined at 26.6o clockwise to the y-axis. Fig. 6.17 Stresses in the wire of torsional pendulum

The maximum shear stress at this location is equal to the radius of the maximum shear stress in the material, we need to construct a Mohr circle.
Mohr circle, i.e., 50 MPa. The Mohr circle for this is as shown in Fig. 6.17b. The point Z with
coordinates (63.7 MPa, 41.8MPa) represents the stresses on the z-plane.
Example 6.8 A torsional pendulum As this is an x-like plane, the positive shear stress is plotted downwards.
The point Θ with coordinates (0, 41.8 MPa) represents the θ-plane. The
Consider the torsional pendulum of
centre of the circle is at ,( ) -, or at (31.8 MPa, 0). The
length 1 m shown in Fig. 6.16. The steel
wire has a diameter of 1 mm. Can the radius of the circle is √(( ) ) , or 52.5 MPa. The
steel wire withstand a torsional
amplitude of 60o, if the weight of the disc maximum shear on the wire, therefore, is 52.5 MPa and occurs on a plane
is 50 N. Assume that the wire fails when that makes an angle clockwise from the z-axis.
the shear stress exceeds 100 MPa, and a
This results in a factor of safety of 100 MPa/52.5 MPa = 1.9, well below
factor of safety of 1.6 is desired.
the prescribed factor. Hence the wire is safe.
Fig. 6.16 Torsional Pendulum
Solution:
Example 6.9 Combined stresses in a cylindrical pressure vessel
There are two sources of stresses in the wire: the direct tension, and the
Consider a thin-walled cylindrical pressure vessel of radius R with wall
shear stresses due to torsion. The direct tension is uniform across the
thickness t as shown in Fig. 6.18a. We had determined in Sec. 1.8 that in
section of the wire:
the polar coordinate directions r, θ and z, there are no shear stresses84,
, ( ) - and that the tensile stresses were given by Eqs. 1.8-1.10 as ,
The shear stresses due to twisting given by Eq. 3.2 vary with the radius , and
with the maximum value occurring at the outer surface.
( )
( ) ( )
( )

Fig. 6.16a shows the stress components on an element at the surface of


the wire. Clearly, this is a case of combined stresses. To determine the 84
This was established by using symmetry arguments
102
Since there are no writing the equilibrium equations. It can be easily verified that ζx’x’ and
shear stresses on ηx’y’ are given by the same expressions as for the case of plane stress, i.e.,
these planes, these Eqs. 6.1 and Eq. 6.2. This happens because in the two equilibrium
are the principal equations the contributions of ηzx and ηzy on the +z- and –z- face cancel
planes. Determine each other. Thus, the construction of Mohr circle for the case of plane
the magnitude and stress can be used even in this general 3-D case for all planes with
direction(s) of the normals in the x-y plane. However, the determination of the principal
maximum shear planes would not be correct, since the usual Mohr circle would ensure that
stress. there is no shear stress ηx’y’ o r ηy’x’ on these planes. ηx’z will, of course,
occur. This will be zero on these planes only if ηxz and ηyz vanish. In can
Solution: be shown that the z-component of shear stress is given by
Fig. 6.18 b shows a (6.7)
two dimensional
element of the thin- Elaborate schemes for transforming 3-D stress have been evolved. Suffice
walled cylindrical Fig. 6.18 Cylindrical pressure vessel it to say here that:
vessel with stresses  The results given by Eqs. 6.1 and 6.3 and the construction of the
shown. Fig. 6.18c shows the corresponding Mohr circle with centre at Mohr circle described above are valid for any orientation of x’-
,( ) - or (3pr/4t, 0), and the radius is ( ) or pr/4t. and y’- axes within the x-y plane even if ζyz, ηyz and ηxz are non-
Thus the maximum stresses are equal to pr/4t, positive on the S2 plane zero.
located 45o anti-clockwise from the z- plane, and negative on the S1 plane  However, if ηyz and ηxz are non-zero, the shear stress on the x’-
located 45o clockwise from the z- plane. plane is given by Eq. 6.7 above, and the Mohr circle cannot be
used for determining the principal planes and principal stresses.

6.6 General 3-D stress 6.7 Displacement and strain


Let us now consider the general case of stress where the stresses are
Consider the deformation of a small element ABCD (Fig. 6.20) which
confined to a plane. This means
undergoes deformation under the action of a generalized load to acquire a
that at a point there are z-
displaced location A’B’C’D’ as shown. Its sides AB and AD of original
components of the stresses as
lengths δx and δy are originally parallel to the coordinate axes. Let us
well. Before considering the
denote the x- and y- components of displacements of point A of the
general transformation, let us
element by u and v, respectively. Then the displacements of points B and
consider the case when the new
plane in which we seek to D can be written as . / and . /,
determine the stresses has a respectively as shown. It is easy to verify that the length of the lines A’B’
normal within the x-y plane as and A’D’, that were originally δx and δy, respectively are now
shown in Fig. 6.19. and , respectively. It is easy to see that the linear strain
We can determine the stress in the x- and y- directions are given by:
component on the inclined plane
Fig. 6.19 General 3-D stress
(with normal in x’-direction) by
103
and (6.8) Consider a plane strain situation depicted in Fig. 6.21. Let u and v and
represent, as before, the displacements in x- and y- directions. Consider a
It can also be line AB of length δl along the x’-
verified, that the axis inclined at an angle θ to the
line A’B’ is x- axis within the material. Let
inclined to the this line move the location A’B’
vertical at an angle after displacement. For
equal to , the convenience of analysis we move
the line A’B’ parallel to itself to
minus sign the location AB”. This is
indicating that this equivalent to subtracting from the
rotation is displacement of the point B the
clockwise. displacement (u, v) of the point A.
Similarly, the line We drop a perpendicular BD from Fig. 6.21 Strain in an arbitrary direction.
A’D’ is inclined to point B to line AB”. For small
the horizontal at displacements, length AD ≈ length AB, so that DB” represents the
an angle equal to elongation of the line AB. To determine this elongation we resort to the
. Thus, the construction of Fig. 6.21 wherein BC is parallel to the x-axis and CB” is
angle between the Fig. 6.20 Deformation of a small element parallel to the y-axis. The coordinates of point B are δlcosθ nd δlsinθ.
two lines is Clearly, the length BC, the additional x-displacement of point B (over that
reduced from a value of π/2 by an amount equal to the sum of the two. of point A) is , and the length CB”, the additional
This reduction in the angle between the two lines perpendicular to each
other to begin with is termed as the shear strain . Thus, y-displacement of point B (over that of point A) is
. The elongation DB” of line AB then is
(6.9)
. The longitudinal strain εx’x’ at this point is obtained
We can, similarly, define the rotation of the element (about the z-axis by dividing this length by the original length δl of line AB. Using the
values of BC and CB” evaluated above, we determine the strain as
perpendicular to the paper) as . /.
. /
This determination of stress component assumes that the strains are an
order less than one. Recognizing the partial derivatives as the strain components εxx, γxy and εyy
we can write this as
6.8 Transformation of plane strains , or on
We have related above the strains in the x- and y- directions to the rearranging,
gradients of the displacements in a material body. We turn next to the
. / . / (6.10)
question if we can relate the strains in any other direction to the strains in
the x- and y- directions, just as we related the stresses on an arbitrarily We can, in a similar fashion, obtain the following equations:
directed plane to the stresses given in the x- and y- directions.
104
. / , and  Plot the strain components on x- and y- planes as points X and Y on the
ε-γ/2 plane using the convention that a positive shear strain on an x-like
(6.11) face is plotted downward while a positive shear strain on a y-like face is
plotted upwards. Reverse for the negative shear strains.
. / . / (6.12)
 Join points X and Y to locate the centre C of the circle on the ε-
Eqs. 6.10 – 6.12 are seen to be identical to the Mohr circle equations, Eqs. axis, and draw the Mohr circle with CX (or CY) as radius.
6.1 – 6.3 except that the stresses and are replaced by the strains  To determine the strain components on planes which are at a
and , and the shear stress is replaced by the shear strain . clockwise angle θ from the original faces, rotate the diameter
YCX through an angle 2θ to obtain points X’ and Y’.
The significance of these results is that they show that it is possible to  The coordinates of points X’ and Y’ give the strain components on
calculate the strains on any arbitrary plane x’ or y’ given the strains on the new planes with normals in the X’ and Y’ directions. The
two mutually perpendicular planes x and y, just as it was possible to do so signs of the shear strain components are to be interpreted using
for the stresses. Thus, the three strain components , the convention given above.
describe completely the state of strain at any point. The identical nature
of the two sets of transformation equations suggests that it should be Example 6.10 Distortion in computer graphics
possible to have a graphical construction for strains as the Mohr circle for An error in a computer-graphics software distorts the images appearing
stress described above. The only significant change will be that the axes on the monitor. The horizontal lines are stretched 1% and the vertical
for such a construction should be ε and in place of ζ and η, the axes for lines are compressed 5% (See Fig. 6.23). Show that little rectangles on
screens may be distorted
into parallelograms.
Determine the maximum
angular distortions of any
rectangles, and also the
orientation of these.

Solution:
Even though the image
on the screen is not a Fig. 6.23 Distortion in computer graphics
material object, there is
nothing in our discussion of the previous section which should restrict the
(a) (b) consideration of deformations and strain from being applied to it. We can
Fig. 6.22 Using Mohr circle to obtain strain components in arbitrary directions thus, look at the problem as one of transformation of strains. In the
the Mohr circle for stresses. This construction is conveniently termed as coordinate system of Fig. 6.23, the given data can be interpreted as: εxx =
the Mohr circle for strain and is shown in Fig. 6.22. 0.01, εyy = − 0.05 nd ηxx = 0.

We summarize below the procedure for the construction of Mohr circle The direction of maximum angular distortion is the direction of maximum
to determine the strain components on arbitrary planes: shear strain and that can be determined by the construction of the Mohr
circle of strain, as shown in Fig. 6.23. Point X on the diagram with
coordinates (+ 0.01, 0) represent the strains on the x-plane, while point Y
105
with coordin tes (− 0.05, 0) represent the str ins on the y-plane. We tensile stress in the principal direction 1, and a compressive stress in the
determine the centre of the circle on ε-axis midway between the two principal direction 2, both of magnitude ηo. We next determine the
points and draw the Mohr circle with radius [0.01 – (−0.05)]/2 = 0.03. principal strains ε1 and ε2 as
The maximum shear strain occurs at planes S1 and S2 located 45o ( ) and ( ). We next draw the Mohr
clockwise from the x- and y- planes, respectively. As is clear from the circle of strain for this state of strain as shown in Fig. 6.24e. Clearly, the
Mohr circle, the maximum angular strain γS1S2 equal to two times the planes inclined at 45o to the principal directions 1 and 2. These are the
radius of the circle, i.e., 0.06. This is the maximum distortion in the original x- and y- directions. The shear strain γxy is seen from this Mohr
image, appearing as a decrease in angle between two lines inclined at 45 o circle is equal to twice the radius, i.e., 2ε1, or ( ). But this is
to the x- and y- directions.
the same shear strain γxy as obtained in Fig. 6.24a. Thus, ( )
. This leads to our desired result:
6.9 Relation among elastic properties of a material
We had in Chapter 2 introduced three elastic properties of materials: ( )
(6.13)
elastic modulus E, Poisson ratio ν, and shear modulus G. We can now
show that the three are related, so that there are only two independent Thus, there are only two independent elastic properties.
elastic properties. Consider a two-dimensional state of stress shown in
6.10 Strain gauges
Resistance strain gauges are based on the principle that when a wire is
stretched its resistance increases. Fig. 6.25 shows the layout of a common
type of strain gage. The gauge is laid out
on a backing sheet which is bonded to the
structural member whose strain is to be
measured. The gauge consists of a long
wire been folded such that its entire
length is aligned in one direction, but is
confined to a small area. The long length
of the resistance wire increases the
change in resistance for a given strain
increasing the sensitivity of the gauge.
The compact size of the gauge ensures Fig. 6.25 Resistance strain gauge
that the gauge responds to the strain in a
small region making it possible to measure the variation of strains with
location. The two leads are soldered to the two tabs shown and
Fig. 6.24 connected to a resistance measuring device.85 The gauge measures the
Fig. 6.24a where an element is subjected to pure shear stress ηo. The shear
strain is seen as . We draw its Mohr circle as Fig. 6.24b. This
shows that the principal planes are inclined at 45o to the x- and y- axes.
85
The orientation of this element is shown as Fig. 6.24c. The element has Such as a Wheatstone bridge circuit.
106
linear strain on the member in the direction of its axis, which is the We determine next the principal directions and the principal strains
direction along which the length of the wire is aligned. through the Mohr circle construction of Fig. 6.27c. We note that these
directions are at 90o in the Mohr circle plane from the axial direction, or
Strain gauges find wide applications in measurement of forces and
inclined at 45o to the axis, and the principle strains are equal to γo/2,
acceleration. Many sensors for vibrations and aerodynamic forces use
tensile on P1 plane and compressive on P2 plane. Thus, the strain gauge
strain gauges to pick
should be fixed such that it is inclined at 45o to the axis of the shaft.
up linear strains and
Notice that the strain gauge shown in Fig. 6.27a is inclined 45o counter-
then convert them into
clockwise, and therefore, it will measure linear strain on the principal
desired quantities. Fig.
plane P1. Thus, the linear strain measured by this strain gauge will be
6.26 shows one such
γo/2, or the maximum shear strain in the shaft will be twice the linear
application. This is the
strain measured by the strain gauge mounted as shown.
schematic of an
experimental set up to Fig. 6.26 Schematic of a set up to pickup vibrations of Example 3.11 Measuring pressure in a soft-drink can
determine the model vocal chords using strain gauge
aerodynamic forces We attempt to measure the pressure inside a soft-drink can by pasting a
that excite a model of human vocal chords. Here a strain gauge is applied strain gauge along the axial direction when it is sealed. The can is opened
near the base of cantilevered reed. The oscillating aerodynamic forces on to release the pressure. The resulting strain (as compared to when the can
the model vocal chord result in the vibrations of the reed. The resulting was pressurized) can be related to the inside pressure of the can.
oscillating strains are picked up by the strain gauge which is calibrated to However, there was an error in fixing the strain gauge and we end up
give the variations of displacement with time. The nature and magnitude having its axis off by an angle of 10o from the axial direction. If the
of the forces can be inferred from the measured rate of growth of the aluminium can has a radius of 3 cm and a wall thickness of 0.07 mm,
amplitude of what was the pressure inside the can if the measured strain was 202 μ,
vibrations. compressive?
Though a strain Solution:
gauge measures only
the linear strains, it This method of measuring the internal pressure of a can is quite
can be used innovative. When we stick a gauge to a can with pressure (and hence,
judiciously with with the wall under stresses and tensile strain), the gauge will read zero to
Mohr circle to begin with. As the can is popped open the strains on the walls will be
measure shear strain
as well. Fig. 6.27a
shows a shaft
subjected to a pure
torque. This will
result in shear strains
as shown for an
Fig. 6.27 Use of strain gauge to measure shear strain.
infinitesimal element (a) Fixing the strain gauge at 45o to a shaft, (b) Shear (a) (b)
at the surface of the stresses on an element at the surface, (c) Mohr circle,
shaft in Fig. 6.27b. and (d) Principal directions and stresses Fig. 6.28 Mohr circle for strains for a can

107
relieved, but in the process, the gauge will be experience a compressive combinations to determine the complete state of stress at a point. A strain
strain, exactly of the same magnitude as the tensile strain in the can along rosette (Fig. 6.29) has three strain gauges on the same backing paper. If
that axis of the strain when the can was pressurized. We will, therefore, the required strain components are εxx, εyy, and γxy, the rosette is cemented
determine the tensile strain (in the can material) along the axis of the to the structure such that gauge a is aligned
strain gauge when the gage is subjected to an internal excess pressure p, with the x-axis. The linear strains measured
and equate it to the (negative) of the gauge reading when the pressure has by gauges b and c (aligned at angles α and β,
been relieved. respectively to the gauge a) are then given in
terms of the desired strains εxx, εyy, and γxy by
Fig. 6.28a shows the schematic of the strain gauge stuck on the outside of
Eq. 6.10 as
the can, while Fig. 6.28b shows the corresponding Mohr circle for strains
with internal excess pressure p. For a can of radius r and wall thickness t, . / . /
and , and Fig. 6.29 Strain rosette

With these stresses, . / . /


0 1, and

0 1 These together with the fact that measured stress εa is equal to εxx give
three equations for the three desired strains components.
( )
The centre C of the Mohr circle is located at , or at . The To give better accuracy in the determination of stresses, there are two
radius of the circle is , or
( )
The axis of the gauge is common types of rosettes available: the 45o-rosette with α = 45o and β =
o 90o, and the the 60o-rosette with α = 60o and β = 120o as shown in Fig.
inclined 10 clockwise to the z-axis. Thus, point G representing this axis 6.30.
on Mohr circle is located at 20o clockwise from the point Z as shown.
The linear stress at the gauge location is, therefore, OC – Rcos20o, or The relevant equations for the 45o-rosette are:
, ( ) ( ) -.
. / . / ,
The gauge is cemented on the can when the material has this strain. As (6.14a)
the pressure is relieved, this strain in the material disappears but the gauge
is subjected to a strain negative of this value. The measured v lue is −202 . / , and (6.14b)
μ. We can, thus, evaluate the can pressure p by equating , (
−6
. / . /
) ( ) - to +202×10 . Using the value of ν as 1/3 and of E
as 70 GPa for aluminium, this excess pressure is determined as 1.77×105 (6.14c)
Pa, or about 1.7 times the atmospheric pressure. Solving the three equations, we get:
, , and ( ) (6.15)
6.10 Strain Rosettes
Even though the strain gauges respond only the axial strain along the
length of its sensing element, we can use a number of them in

108
shear, while a brittle specimen fails at an angle (see Fig. 6.14e) indicating
a failure in tension.87
A number of theories have been proposed to establish failure criteria for
structural elements in a general state of stress, i.e., when both tensile
stresses and shear stresses are present.
One simple theory known as maximum normal stress criterion postulates
that the failure occurs when the
maximum tensile (or compressive) stress
in any direction exceeds the tensile
strength of the material. Since the
maximum tensile stresses, positive or
(a) (b) negative, are the principal stresses, the
o o
Fig. 6.30 (a) A 45 -rosette (b) A 60 -rosette failure is predicted when any of the
numerical value of the principal stresses
We can similarly obtain the strain components from the measurement on exceeds the yield strength of the material.
a 60o-rosette as In Fig. 6.31, thus, if the state of stress Fig. 6.31 Maximum principal
falls within the shaded region the stress criterion
, , and (6.16) structure is safe. This theory worries only
about the maximum value of the principal
Example 3.12 stress and neglects the effect of the other
two principal stresses as long as they are
below the yield strength. Nevertheless,
6.11 Criteria for Failure the results on brittle materials with about
We had seen in Sec. 2.6 that when a specimen of a material is subjected to equal strength in tensile and compressive
a tension test it fails. A brittle material fails at a higher level of stress in loadings agree well with this theory.
compression than in tension. We had also seen that in a tension test while Note that the yield strength is replaced by
a ductile material showed a cup-and-cone failure, a brittle material the failure strength for brittle materials. Fig. 6.32 Mohr criterion for
showed no such failure. As explained in Example 6.3, the cup-and-cone brittle materials strong in
For brittle materials with vastly different compression
failure of a ductile specimen suggested that the material failed in shear,86
tensile and compressive strengths, the
and the flat failure of a ductile material was suggestive of a failure in
Mohr criterion of failure gives fairly good prediction. Here, two Mohr
tension.
circles are drawn: one for the uniaxial tension test with the failure
But all this is only for a tensile loading. What about failure in a pure strength (ζY)T in tension as one principal stress and 0 as the other, and the
torsion condition? When we subject a bar to a pure torsion, like the other circle for the uniaxial compression test with the failure strength
loadings discussed in Chapter 3 and in Example 6.6, a ductile specimen (ζY)C in compression as one principal stress and 0 as the other, as in Fig.
fractures perpendicular to the axis (see Fig. 6.14f) indicating a failure in

87
Because in this case of pure torsion the principal planes where maximum
86 o
Because in a uniaxial loading the maximum shear occurs at 45 to the axis. tensile stresses occurs are inclined at ±45o to the axis.
109
6.32. The shaded area of the diagram is the safe zone. A structure ζ2). Thus, the line CD in Fig. 6.33 represents the combination of principal
ruptures if its state of stress falls outside the shaded region. stresses ζ1 and ζ2 for which the maximum shear stress is equal to ζY/2.
The Maximum shear stress criterion or The boundary shaded area of Fig. 6.34 represents the combinations of ζ1
Tresca criterion gives fairly good (but and ζ2 for which the maximum shear stress (on any plane) is ζY/2. The
conservative) results for ductile shaded area, thus, represents the safe zone. The material is expected to
materials. It postulates that the failure of yield if the state of its stress falls outside this area. We can, thus, write
a ductile material occurs when the the maximum shear stress criterion as:
maximum shear stress at a point exceeds
A structure fails whenever either | | | | or | |
the yield value of the shear stress.
Let us next consider the specimen subject to pure torsion, a case that was
Value of the maximum shear stress at a
Fig. 6.33 Mohr circle for plane discussed as Example 6.6. We had seen there that the two principal
point is related to the values of the two
stress stresses ζ1 and ζ2 were of equal magnitude but of opposite signs. This is
principal stresses ζ1 and ζ2 at that point.
the case represented by point S of Fig. 6.34. Clearly, then, Tresca criteria
There are two distinct cases: one, when
suggests that the maximum shear stress that a specimen may be subjected
the sign of the two principal stresses are the same, i.e., both of them are
to before yielding is that which results in ζ1 = − ζ2 = ζY/2. The Mohr
either tensile or compressive, and the second, when one of them is
circle for this case, drawn as Fig. 6.14c shows that these values of
compressive and the other tensile.
principal stresses are obtained when the maximum shear stress in the
Mohr circle for the first case is shown as torsion test is itself equal to ζY/2. But in actual torsion test on most
Fig. 6.33. Both ζ1 and ζ2 are positive in ductile materials, the measured yield shear strength ηY is about 0.58ζY, a
this figure. Fig. shows the Mohr circle good 15% higher than predicted by this theory.
for such a case. Here the maximum shear
Based on this and the fact that most ductile materials are able to sustain
in the material does not occur in the
very large hydrostatic88 loads, Von Mises proposed that it is the distortion
plane within which ζ1 and ζ2 lie, but in a
of a material, as opposed to its volumetric strain that causes it to yield.
plane formed by the direction of
The proposed measure is the energy of distortion. We had seen in Eq.
principal stress ζ1 and the third direction
Fig. 6.34 Maximum shear stress 2.17 that the strain energy per unit volume of the material is measured by:
(where the principal stress is 0). The
criterion
maximum shear stress thus, is ζ1/2. In a ( )
uniaxial test the material yields at ζ1 =
ζY. Thus, ζ1 (or ζ2 in case it is larger than From this total strain energy density we subtract the strain energy due to
ζ1) should be equal to or greater than ζY change in volume to obtain the distortion energy.
for yielding. Thus, lines AB and BC in If we deal in principal planes and principal stresses and strains, this
Fig. 6.34 represent the limiting case. equation reduces to:
Any state of stress within the shaded
square OABC is safe from yielding, ( ) (6.17)
while the points outside represent
yielding.
Mohr circle for the other case with the Fig. 6.35
two principal stresses having opposite signs is shown in Fig. 6.35. The 88
A hydrostatic load is one where a material is subjected to a uniform load in all
maximum shear stress in this case is given by (| |) | |) or (ζ1 − directions. See Example 6.5.
110
Using Eqs. 2.2-2.4, we can write the principal strains in terms of principal Von Mises yield criterion, thus, predicts yielding when
stresses: , ( )-, etc. Using these strains, we ,( ) ( ) ( ) - 2 (6.21)
get the total energy density as:
The ellipse of Fig. 6.37 is a plot of this for the plane stress condition. A
, ( )- point inside the shaded portion is safe from yielding. Note that it is less
(6.18) conservative that the maximum shear stress criterion, and is seen to give
good results in most cases.
Let us next consider the energy of the volumetric strain. For this purpose
we resolve the given stresses into hydrostatic stresses and deformation Let us consider the torsion test again. At the yield point when η = ηY, the
stresses as in Fig. 6.36. Here we replace the average of the three principal stresses ζ1 and ζ2 are both
equal to ηY (by Mohr circle of Example
principal stresses, ( ) , by representing the hydrostatic
6.6), and therefore, by Eq. 6.21:
stress. With the three principal stresses equal to , Eq. 6.19 gives the
strain energy of volumetric change as , or , almost
exactly the result obtained for most
ductile material (point S in Fig. 6.36).

Fig. 6.37 Von Mises yield criterion

Fig. 6.36 Resolving stresses into hydrostatic stresses and deformation stresses

( )
(6.19)
The energy of distortion (per unit volume) is obtained by subtracting uh
from the total strain energy:
( )
This can be recast as;
( )
,( ) ( ) ( ) -
(6.20)
The theory proposes that a material yields when this distortion energy
(per unit volume) exceeds the distortion energy per unit volume in a
uniaxial tension specimen stressed to its yield strength.
In uniaxial tension test, ζ1 = ζY, and ζ2 = ζ3 = 0.
( )
Thus, the distortion energy at yield point is .

111

Das könnte Ihnen auch gefallen