You are on page 1of 23

LECTURE 10

Plane Stress Transformation

10.1 Introduction

As discussed in Lecture 2, the general three-dimensional state of stress is obtained, as


was discussed in Lecture 2, by passing a set of three orthogonal planes though a point
of the solid body and isolating the infinitesimal volume located around the point. The
three-dimensional state of stress is illustrated in Figure 10.1 where, for clarity, only
the stresses drawn on the faces with positive normal are represented.

Figure 10.1 Three-Dimensional State of Stress

Mathematically the three-dimensional state of stress is represented by a generalized


stress tensor T ( x, y, z) defined by nine distinct stress components:

 x  xy  xz
T ( x, y, z )   yx  y  yz (10.1)
 zx  zy  z

Applying the shear stress duality principle the number of independent tensorial
components is reduced from nine to six and, consequently, the symmetric shear stress
components are related as:

 xy   yx (10.2)

- 181 -
Lecture 10 – Plane Stress Transformation

 xz   zx (10.3)

 yz   zy (10.4)

There are instances when some of the stress tensor components vanish and the general
three-dimensional stress tensor T ( x, y, z) degenerates into a tensor characterized by
only three independent components. This condition is called plane state of stress. For
the present theoretical development, it is assumed that all stress components pertinent
to the planes having normal vector parallel to the z axis are zero:

z  0 (10.5)

 zx   xz  0 (10.6)

 zy   yz  0 (10.7)

This type of plane stress tensor is illustrated in Figure 10.2.

Figure 10.2 Plane Stress Tensor

Additionally, the remaining non-zero stress tensor components contained in equation


(10.1) are assumed to be independent of the variable z .

T ( x, y, z)  T ( x, y) (10.8)

Consequently, the plane stress tensor T ( x, y ) can be represented in plane oxy as


depicted in Figure 10.3:

Figure 10.3 Plane State of Stress

- 182 -
Lecture 10 – Plane Stress Transformation

The plane stress tensor T ( x, y ) is defined by three non-zero components:

 x  xy
T ( x, y )  (10.9)
 yx  y

Several of the conditions encountered in previous lectures, including the study of the
axial and torsional deformation and pure and non-uniform bending, are characterized
by different states of plane stress. In reality, the plane stress tensor is a direct result of
the assumptions imposed on the deformation.

Examples of plane stress tensors, such as uni-axial, pure shear and bi-axial, are
illustrated in Figure 10.4.

Figure 10.4 Examples of Plane State of Stress


(a) Uni-Axial, (b) Pure Shear and (c) Bi-Axial

10.2 Plane Stress Transformation Equations

Suppose that the components of the plane stress tensor T ( x, y ) , as expressed by


equation (10.9), are defined at any point P( x, y) of the vertical plane oxy . The
variation of stress components when the reference system attached to point P( x, y) is
rotated with a counterclockwise angle  , as illustrated in Figure 10.5, is the subject of
this section.

- 183 -
Lecture 10 – Plane Stress Transformation

Figure 10.5 Representation of the Plane State of Stress


(a) Normal Planes and (b) Rotated Planes

The transformation relations between stresses  n ,  t and  nt pertinent to a rotated


plane and the stresses  x ,  y and  xy are obtained by writing the equilibrium
equations for the infinitesimal triangular element depicted in Figure 10.6. The
inclined plane is defined by its positive normal n which is rotated counterclockwise
with an angle  from the horizontal direction x .

Figure 10.6 Equilibrium of the Infinitesimal Triangular Element


(a) Stresses and (b) Forces

Using the notation shown in Figure 10.6.b the equilibrium equations are written as:

F n 0

( n * A)  ( x * Ax ) * cos  ( y * Ay ) * sin  


(10.10)
 ( xy * Ax ) * sin   ( yx * Ay ) * cos  0

F t 0

( nt * A)  ( x * Ax ) * sin   ( y * A y ) * cos  


(10.11)
 ( xy * Ax ) * cos   ( yx * A y ) * sin   0

- 184 -
Lecture 10 – Plane Stress Transformation

From Figure 10.6 the following geometrical relations can be derived:

Ax  A * cos (10.12)

Ay  A * sin  (10.13)

Substituting equations (10.12) and (10.13) into equations (10.10) and (10.11) and
using the shear stress duality principle the following expressions are obtained for
normal  n and shear  nt stresses:

 n   x * cos 2    y * sin 2   2 * xy * sin  * cos  (10.14)

 nt  ( x   y ) * sin  * cos    xy * (cos 2   sin 2  ) (10.15)

Equations (10.14) and (10.15) can be re-written using the trigonometric relations
between the angle  and the double angle (2 * ) :

 x  y  x  y
n   * cos(2 *  )   xy * sin( 2 *  ) (10.16)
2 2

x  y
 nt   * sin( 2 * )   xy * cos(2 * ) (10.17)
2

Equation (10.16) and (10.17) are called the plane stress transformation equations.

In general, two faces are needed to express the plane stress tensor around a point
P( x, y) . Consequently, the formulae (10.16) and (10.17) are applied twice: first,
considering the rotation angle    xx ' and, secondly, for the complementary
angle    xy ' . The notation is illustrated in Figure 10.7. The rotation angles  xx ' and
 xy are related as:
'


 xy   xx 
' ' (10.18)
2

The relation between the double angles necessary in equations (10.16) and (10.17) is
obtained as:

2 *  xy '  2 *  xx '   (10.19)

Consequently, the following trigonometric relations can be established:

sin( 2 *  xx ' )   sin( 2 *  xy ' ) (10.20)

- 185 -
Lecture 10 – Plane Stress Transformation

cos( 2 *  xx ' )   cos( 2 *  xy ' ) (10.21)

Figure 10.7 Stresses on Orthogonal Rotated Faces

The stresses on two orthogonal rotated faces are expressed as:

 x  y  x  y
 x  n(   xx ) 
' '  * cos(2 *  xx ' ) 
2 2 (10.22)
  xy * sin( 2 *  xx ' )

 x  y
 x y   nt (   xx )  
' ' ' * sin( 2 *  xx' ) 
2 (10.23)
  xy * cos(2 *  xx' )

 x  y  x  y
 y  n(   xx  90  ) 
' '  * cos(2 *  xx ' ) 
2 2 (10.24)
  xy * sin( 2 *  xx ' )

 x  y
 y x   nt (   xx  90  ) 
' ' ' * sin( 2 *  xx ' ) 
2 (10.25)
  xy * cos(2 *  xx ' )

If equations (10.22) and (10.24) are summed, the invariance of the summation of
normal stresses is established:

 x  y   x  y
' ' (10.26)

- 186 -
Lecture 10 – Plane Stress Transformation

10.3 Principal Stresses

The maximum and minimum normal stresses are called principal stresses and
mathematically represent the extreme values of the normal stress function  n ( ) . The
extreme values are obtained by imposing the condition that the first derivative of the
normal stress  n ( ) relative to the rotation angle  is zero:

d n ( )
0 (10.27)
d

The explicit expression for equation (10.27) is obtained by differentiating equation


(10.16):

 ( x   y ) * sin( 2 *  )  2 *  xy * cos( 2 *  )  0 (10.28)

Dividing by 2* cos(2 * ) the trigonometric equation (10.28) is transformed into


equation (10.29) relating the tangent of twice the principal directions angle,  p , to the
stresses ion the orthogonal planes x and y :

 xy
tan(2 *  p )  (10.29)
 x  y
2

The angle  p represents the angle for which the normal stress  n ( ) reaches its
extreme value. The geometrical illustration of the equation (10.29) is presented in
Figure 10.8.a where the distance R is calculated as:

( x   y ) 2
R   xy
2
(10.30)
4

Figure 10.8 Geometrical Representation of Equation (10.29)

Solution of the trigonometric equation (10.29) yields two solutions, ( 2 *  p1 )


and ( 2 *  p 2 ) , where the two angles are related as:

- 187 -
Lecture 10 – Plane Stress Transformation

2 *  p 2  2 *  p1   (10.31)

From equation (10.31), the orthogonality of the two principal directions is established:


 p 2   p1  (10.32)
2

To calculate the values of the normal stress  n ( ) corresponding to the angles  p1 and
 p 2 it is necessary to evaluate the trigonometric functions sin( 2 *  p ) and cos( 2 *  p )
contained in equation (10.16). Using the notation shown in Figure 10.8.a and the
trigonometric relations (10.20) and (10.21) these functions can be expressed as:

 xy
sin( 2 *  p1 )  (10.33)
R

 x  y
cos(2 *  p1 )  2 (10.34)
R

 xy
sin( 2 *  p 2 )   sin( 2 *  p1 )   (10.35)
R

 x  y
cos(2 *  p 2 )   cos(2 *  p1 )   2 (10.36)
R

Substituting first the trigonometric expressions (10.33) and (10.34) into equation
(10.16) and then (10.35) and (10.36) the principal stresses  1 and  2 are obtained:

 x  y ( x   y ) 2
 1   n (   p1 )     xy   av  R
2
(10.37)
2 4

 x  y ( x   y ) 2
 2   n (   p 2 )     xy   av  R
2
(10.38)
2 4

The average normal stress is calculated as:

 x  y
 av  (10.39)
2

The principal stresses are schematically depicted in Figure 10.9.

- 188 -
Lecture 10 – Plane Stress Transformation

Figure 10.9 Principal Stresses and Directions

The right-hand expression of equation (10.28) being identical with the expression
(10.17), representing the shear stress  nt , allows equation (10.28) to be re-written as:

d n ( )
  nt  0 (10.40)
d

Equation (10.40) indicates that the principal normal stresses are obtained for a
rotated plane where the shear stress is zero.

The invariance of the sum of the normal stresses is again shown to be valid for the
case of the principal stresses. Summation of equations (10.37) and (10.38) yields:

 1   2  2 *  av   x   y (10.41)

To identify which of the two angles,  p1 or  p 2 , corresponds to the maximum


principal stress  1 the second derivative of the function  n ( ) relative to the rotation
angle  is employed. The condition for the point to be a maximum is:

d 2 n ( )
(   p )  0 (10.42)
d 2

The condition (10.42) is explicitly written as:

 x  y
 * cos(2 *  p )  xy* sin( 2 *  p )  0 (10.43)
2

The inequality (10.43) can be manipulated and cast in a new form:

 x  y 1
[ *   xy ] * sin( 2 *  p )  0 (10.44)
2 tan(2 *  p )

If the sin( 2 * p ) is expanded the following trigonometric expression is established:

- 189 -
Lecture 10 – Plane Stress Transformation

sin( 2 *  p )  2 * sin  p * cos  p  2 * tan  p * (cos  p ) 2 (10.45)

Substituting equations (10.29) and (10.45), representing the tan( 2 *  p ) and


sin( 2 * p ) , into the inequality (10.44) the following expression is obtained:

( x   y ) 2 tan p
  xy ] * cos 2  p * 0
2
2 *[ (10.46)
4  xy

The condition for the inequality (10.46) to hold true is:

tan p
0 (10.47)
 xy

Note: It is important to note that for the inequality (10.47) to hold true, the signs of
the tangent of the angle  p and shear stress  xy must be identical.

The angle corresponding to the direction of the maximum normal stress can also be
obtained by successively assigning to the angle  in equation (10.16) the values  p1
and  p 2 and observing which angle produces the maximum principal stress.

10.4 Maximum Shear Stresses

The maximum shear stresses are determined in a similar manner as the principal
stresses. The extreme condition for the shear stress function  nt ( ) contained in
equation (10.17) is written as:

d nt ( )
0 (10.48)
d ( )

The explicit format of equation (10.48) is obtained by differentiating the expression


(10.17):

 ( x   y ) * cos( 2 *  )  2 *  xy * sin( 2 *  )  0 (10.49)

Dividing by 2 * cos(2 * ) the trigonometric equation (10.49), the tangent of the


principal directions angle  s is obtained:

 x  y

2 1
tan(2 *  s )   (10.50)
 xy tan(2 *  p )

- 190 -
Lecture 10 – Plane Stress Transformation

The angle  s represents the angle for which the shear stress  nt ( ) reaches its extreme
value. Equation (10.50) is illustrated in Figure 10.10.

Figure 10.10 Angular Relation between (2 * s ) and ( 2 *  p )

Solving the trigonometric equation (10.50), two solutions (2 * s1 ) and ( 2 *  s 2 ) are


obtained. They are related as:

2 * s 2  2 * s1   (10.51)

Dividing equation (10.50) by two (2), the orthogonality of the two angles  s1 and  s 2 is
obtained:


 s 2   s1  (10.52)
2

Equation (10.50) indicates that a relation between the angles ( 2 *  p ) and ( 2 *  s ) can
be established. With this intent, equation (10.50) is recast into a new format as
follows:

sin( 2 *  s ) cos(2 *  p )
 0 (10.53)
cos(2 *  s ) sin( 2 *  p )

First, multiplying by the terms in the denominator, equation (10.53) becomes:

sin( 2 *  s ) * sin( 2 *  p )  cos( 2 *  p ) * cos( 2 *  s )  0 (10.54)

Then, equation (10.54) is simplified as:

cos[( 2 *  s )  (2 *  p )]  0 (10.55)

Therefore,


(2 *  s )  (2 *  p )]   (10.56)
2

- 191 -
Lecture 10 – Plane Stress Transformation

The relationship between angles  s and  p is calculated from equation (10.56):


s   p  (10.57)
4

Still, equation (10.57) does not indicate how to identify the direction of the maximum
shear stress. By examination of Figure 10.10, the following angular relations can be
established:

 3* 
2 *  s1  2 *  p1     2 *  p1   2 *  p1  (10.58)
2 2 2

The relationship between the angles of the maxim principal and shear stress
directions,  s and  p , is obtained from (10.58) as:


 s1   p1  (10.59)
4

Again using the notation shown in Figure 10.10, the following trigonometric relations
are obtained:

 x y

sin( 2 *  s1 )  2 (10.60)
R

 xy
cos(2 *  s1 )  (10.61)
R

 x y
sin( 2 *  s 2 )  2 (10.62)
R

  xy
cos(2 *  s 2 )  (10.63)
R

Successively substituting the two groups of expressions, (10.60) and (10.61), and,
(10.62) and (10.63), into equation (10.17) the maximum and minimum shear stresses
are calculated as:

( x   y ) 2
 s1   nt (   s1 )    xy  R   max
2
(10.64)
4

( x   y ) 2
 s 2   nt (   s 2 )     xy   R   min
2
(10.65)
4

- 192 -
Lecture 10 – Plane Stress Transformation

The normal stresses corresponding to the maximum and minimum shear stresses are
calculated by substituting the expressions (10.60) through (10.63) into equation
(10.16):

 x  y
 s1   n (   s1 )  (10.66)
2

 x  y
 s 2   n (   s 2 )  (10.67)
2

The maximum and minimum shear stresses and the corresponding normal stresses are
illustrated in Figure 10.11.

Figure 10.11 Relationships between Principal Planes and Maximum Shear Stress
Planes

Note: From Figure 10.12.b it can be concluded that, in contrast, to the principal
planes which are free of shear stress, the planes on which the shear stress
achieves extreme values are not necessarily free of normal stresses.

10.5 Mohr’s Circle for Plane Stresses

Mohr’s circle is a graphical construction reflecting the variation of the plane state of
stress around a particular point, including information pertinent to the principal and
maximum shear stresses.

From equations (10.16) and (10.17), first squared and then summed, the following
relation is obtained:

( n   av ) 2   nt  R 2
2
(10.68)

- 193 -
Lecture 10 – Plane Stress Transformation

Equation (10.68) represents the equation of a circle of radius R defined in the  n , nt


plane. The center of the circle is located at point C ( av ,0) . The values of circle
radius R and average normal stress  av are calculated employing equations (10.30) and
(10.39), respectively.

Intersecting the equation of the circle (10.68) with the horizontal axis  nt  0 the
intersection points P 1 ( n1 ,0) and P2 ( n 2 ,0) are obtained:

 n1   av  R   1 (10.69)

 n 2   av  R   2 (10.70)

It can be concluded that these intersection points represent the principal stresses
 1 and  2 .

The Mohr’s circle for plane stress condition is drawn relative to a Cartesian system
with the abscissa and the ordinate axis representing the normal stresses  and the
shear stress  , respectively. The following sign convention is employed as illustrated
in Figure 10.12:

(a) the positive shear stress  axis is downward;

(b) the positive angle is measured counterclockwise;

(c) the shear stress on a face plots as positive shear if tends to rotate the face
counterclockwise.

Figure 10.12 Mohr’s Circle Notation

- 194 -
Lecture 10 – Plane Stress Transformation

Note: The positive direction of the vertical axis, representing the shear stress  ,
pointing downward (sign convention (a)) is elected in order to be able to
enforce the positive measurement of the angle (sign convention (b)).
Examining the notations shown in Figure 10.12 clarifies that all the angles are
measured from the line XY (   0 ) in the anticlockwise direction.

Morh’s circle for plane stress is constructed in the following steps:

(a) The coordinate system is drawn as shown in Figure 10.12. The horizontal axis
represents the normal stress  , while the vertical axis represents the shear
stress  . To gain full advantage of the graphical benefits of the method it is
necessary that the drawing to be made on scale. However, the method is also
helpful as a conceptual tool in combination with the governing equations
wherein it may be drawn more roughly. The representation considers that the
following conditions are met:  x   y and  xy  0 ;

(b) Using the calculated values of the normal stresses  x and  y and the shear
stress  xy two points noted as X ( x , xy ) and Y ( y , xy ) are placed on the
drawing. The line XY intersects the horizontal axis at point C which
represents the center of the Mohr’s circle;

(c) The distance CX represents the radius of the circle. Using the radius CX and
the position of the center C ( avg ,0) the Mohr’s circle is constructed. The
intersection points, P1 and P2 , between the circle and the horizontal axis
represent the maximum and the minimum principal stresses;

(d) The value of the tan( 2 *  p ) can be calculated from the graph. The angle
( 2 *  p ) is identified on the graph by the XCP1 angle and is measured from
the   0 to the principal directions line in the counterclockwise direction;

(e) The lines XP1 and XP2 represent the principal direction1 (associated with the
maximum principal stress) and 2 (associated with the minimum principal
stress), respectively.

Every point on the Mohr’s circle corresponds to a pair of stresses  and  on a


particular face. To emphasize the face involved the point is labeled identically with
the face where it belongs. For example, the face x , y and n are represented on the
Mohr’s circle by the points X , Y and N . To reinforce the shear stress sign convention
(c) two icons indicating the rotation sense induced by the shear stress are shown in
Figure 10.12. The angle measured from   0 to the radius line CN in the
counterclockwise direction is equal to twice the angle of the plane rotation (2 * ) .

Note: The points X and Y represent the case of orthogonal planes having normals
parallel to axes x and y , respectively. The line XY corresponds to angle

- 195 -
Lecture 10 – Plane Stress Transformation

  0 . The double angle (2 *  p1 ) of the maximum principal direction is


measured from the line CX to the line CP1 . By these conventions, the double
angle sense is established as being positive in the counterclockwise direction.
The angle  p1 is the angle XP2 P1 and has the same direction as the double
angle ( 2 *  p1 ) . The angle associated with the minimum direction  p 2 is
perpendicular to the angle  p1 . The stresses corresponding to a plane rotated
with an angle  are obtained by placing on the circle the radius CN located by
measuring in the counterclockwise direction an angle of (2 * ) from the
line CX . The corresponding stresses  n and  nt are a function of the location
of the point N ( n , nt ) position in the    coordinate system and may be
obtained by scaling them from the figure or by the use of the analytical
equations (10.30), (10.39) and (10.68). The opposite point
T ( n , nt ) represents the stresses on the orthogonal rotated face.

In comparison with the technique used to show the principal axes in the oxy
representation (Figures 10.7 through 10.11) the plot obtained from the Mohr’s circle
appears to be misleading. The cause is that the Mohr’s circle is drawn in the   
coordinate system. In the oxy representation the principal directions are correctly
plotted by artificially rotating the principal directions obtained from the Mohr’s
circle around the point C with an angle ( 2 *  p1 ) measured in the counterclockwise
direction.

10.6 Principal Stresses Distribution in Beams

One of the most important applications of the plane state of stress theory described
above is found in the study of variation of the stresses in beams under non-uniform
bending. Recall from Lecture 7 that under some imposed kinematic assumptions, a
beam subjected to transversal loading is in a state of plane stress. With the exception
of some areas (around the supports or the application points of concentrated loads) the
beam theory characterizes the existence of only two types of stresses: normal stress
 x and shear stress  xy . The normal stress  x is calculated using Navier’s formula
expressed by equation (10.71), while the shear stress  xy is obtained employing
Jurawski’s formula contained in equation (10.72):

M z ( x)
 x ( x, y )   *y (10.71)
Iz

'
V y ( x) * S zA ( y )
 xy ( x, y )  (10.72)
Iz *t

- 196 -
Lecture 10 – Plane Stress Transformation

The notation used in the formulae (10.71) and (10.72) is explained in Lecture 7 and is
not repeated herein.

The plane stress tensor previously expressed in equation (10.9) is written for the case
of the beam in nonuniform bending as:

 x  xy
T ( x, y )  (10.73)
 yx 0

The entire theoretical development described in the previous sections can be without
restriction applied to the study of the particular plane stress tensor (10.73).
Consequently, the variation of the stresses around any point in a beam subjected to
nonuniform bending can be calculated. Figure 10.13 represents an example of the
application of plane stress theory for the case of a simply supported beam.

In the example, the beam has a rectangular cross-section and is subjected to a single
concentrated force 2 * P located at the mid-span. It is evident that the ratios of the
beam dimensions and the loading do not violate any of the assumptions related to the
applications of the formulae (10.71) and (10.72). The shear force diagram V (x) and
the bending diagram M (x) , where the axis identification indices were dropped for
clarity, are plotted.

Figure 10.13 Simple Supported Beam

The geometrical characteristics of the rectangular cross-section involved in the


evaluation of the formulae (10.71) and (10.72) are:

b * (2 * c) 3 2
Iz   * b * c3 (10.74)
12 3

- 197 -
Lecture 10 – Plane Stress Transformation

t b (10.75)

yc 1 y2
S z ( y )  b * (c  y ) *  * b * c 2 * (1  2 ) (10.76)
2 2 c

For the left half of the beam 0  x  5c the shear force V (x) and the bending moment
M (x) are expressed as:

V ( x)  P (10.77)

M ( x)  P * x (10.78)

Substituting equations (10.74) through (10.78) into equations (10.71) and (10.72), the
expressions for normal and shear stresses are obtained as:

P*x 3 P
 x ( x, y )   *y * *x* y (10.79)
2 2 b * c3
* b * c3
3

1 y2
'
V y ( x) * S zA ( y ) P * * b * c * (1  2 )
2

 xy ( x, y )    2 c 
Iz *t 2
* b * c3 * b (10.80)
3
2
3 P y
 * * (1  2 )
4 b*c c

Note: The minus (-) sign appearing in formula (10.81) has been inserted in order to
comply with the shear sign convention (c).

To obtain an illustrative variation of the principal stresses, the rectangular domain of


the beam is divided by superimposing a rectangular mesh. For the case under study,
the mesh has five spaces in the longitudinal x direction and eight spaces in the
vertical y direction. Using Mathcad programming capabilities the principal stresses
and corresponding angular directions can be easily calculated for every point of the
mesh. The principal stresses calculated for two cross-sections x  c and x  4 * c and
all nine points vertically describing the cross-sections are contained in Table 10.1.
 max  min
The ratios and are tabulated instead of the  max and  min ,
0 0
P
where  0  .
2*b*c

A review of the results presented in Table 10.1 shows that at the extreme fibers the
principal stresses correspond with the normal stresses and reach the maximum values.
At the extreme fiber locations the shear stress  xy is zero. The situation is different for

- 198 -
Lecture 10 – Plane Stress Transformation

the case of wide-flange beams where both  x and  xy have significant values at the
junction between the web and the flange.

Table 10.1

Today, with the help of modern computer codes, the formulae involved in the
calculation of the principal stresses and directions can be computed using a very
refined mesh. The graph containing the curves tangent to the principal directions in
every point of the mesh is called the stress trajectory. Two sets of curves are drawn
and they are orthogonal at every point. The stress trajectory graph pertinent to the
simply supported beam investigated above is pictured in Figure 10.14. A typical
example of practical usage of the stress trajectory curves is the placement of the
reinforcement in reinforced concrete beam. Because the stress trajectory graph does
not give any indication about the magnitude of the principal stresses another type of
graph is also used. This is called a stress contour plot and contains curves of equal
principal stress magnitudes. The commercial codes employed today can provide these
plots.

- 199 -
Lecture 10 – Plane Stress Transformation

Figure 10.14 Stress Trajectory Plot

10.7 Example

The theoretical formulation derived above is used to investigate the following


practical case:

 x 20 MPa

 y 10 MPa

 xy 10 MPa

The corresponding stress tensor is written as:

20  10
T ( x, y )  * MPa
 10  10

The state of stress for the case above is shown in Figure 10.13.

Figure 10.13 Example Plane State of Stress

The following values illustrated in Figure 10.13 are calculated as:

x  y 20  (10)
 avg    5MPa
2 2

- 200 -
Lecture 10 – Plane Stress Transformation

 x  y 20  (10)
  15MPa
2 2

( x   y ) 2
R   xy  15 2  (10) 2  18.028MPa
2

 xy  10
tan(2 *  p )    0.667
 x  y 15
2

2 *  p  33 .69

Figure 10.14 Geometrical Relations

From equation (10.47) it is established that the angle related to the maximum
principal direction must have a negative tangent. Consequently, the angles of the
principal direction are:

 p1  16 .84 

 p 2  16 .84  90  73 .16 

The principal stresses, shown in Figure 10.15, are obtained as:

 1   avg  R  5  18 .028  23 .028 MPa

 2   avg  R  5  18 .028  13 .028 MPa

- 201 -
Lecture 10 – Plane Stress Transformation

Figure 10.15 Principal Stresses

The angle of the maximum shear stresses is calculated as:

 s1   p1  45   16 .84   45   61 .84 

The maximum shear stresses are calculated as:

( x   y ) 2
 s1   nt (   s1 )    xy  18.028MPa
2

( x   y ) 2
 s 2   nt (   s 2 )     xy  18.028MPa
2

The normal stresses acting on the maximum shear planes are calculated as:

 x  y
 s1   n (   s1 )   5MPa
2

x  y
 s 2   n (   s 2 )   5MPa
2

Figure 10.16 Maximum Shear Stresses

- 202 -
Lecture 10 – Plane Stress Transformation

The maximum shear stresses and the corresponding normal stresses are illustrated
Figure 10.16

Morh’s circle pertinent to the problem is illustrated in Figure 10.17.

Figure 10.17 Mohr’s Circle

Note: The points X ' and Y ' represent the case of orthogonal planes having the
normal directions rotated with angles of 30 and 120 , respectively, from the
x axis. Successively substituting the above angular values in equations
(10.16) and (10.18) the following stresses pertinent to points X ' and Y ' are
obtained:

 for   30

 X  3.84 MPa'

 X Y  17 .99 MPa


' '

 for   120

 Y  6.16 MPa
'

 '
Y 'X '
 17 .99 MPa

- 203 -

You might also like