You are on page 1of 25

Chapter 1 States of stress and strain at a point

1.0 Introduction
In the first Chapter we discuss analysis of stress under conditions of plane stress and
plane strain; strain rosettes; generalized stress-strain relationships for linearly elastic and
isotropic materials.

1.1 Stress – Strain Relations


Deformations (bending, twisting, compression, torsion and shear) in a material occur for
a number of reasons, such as external applied loads, change in temperature, tightening of
bolts, etc.

In some materials (rubber, plastic, wood) the deformations may be large for relatively
small loads and readily observable by eye. In metals, small load will produce relatively
small deformation. Stress values do not always provide the limiting factor in design, for
although a component may be safe and employ material economically wuith regard to
stress, the deformation with reghard to stress need to be consided.

For example, a large deflection of a aeroplane wing may affect the aerodynamic
characteristics.

1.2 Definition of Stress


External forces on a body
Consider an element of continuous (no voids) and cohesive (no cracks, breaks and
defects) material subjected to a number of externally applied loads as shown in Fig. 1.1a.
It is supposed that the member is in equilibrium.

F4 Free Body Diagram


n
F3 F5 ∆F
∆ Fn
∆ Ft t
Cross section: A ∆A

F2 F2

F1 F1
(a) (b)
Fig. 1.1 External and internal forces in a structural member

If we now cut this body, the applied forces can be thought of as being distributed over the
cut area A as in Fig. 1.1b. Now if we look at infinitesimal regions ∆A, we assume the
resultant force in this infinitesimal area is ∆F. In fact, ∆F is also a distributed force.

1
When ∆A is extremely small, we can say that the distributed force ∆F is nearly uniform.
In other words, if we look at the whole sectioned area, we can say that the entire area A is
subject to an infinite number of forces, where each one (of magnitude ∆F) acts over a
small area of size ∆A. Now, we can define stress.

Definition: Stress is the intensity of the internal force on a specific plane


passing through a point.

Mathematically, stress at a point can be expressed as


∆F
σ = lim (1.1)
∆A→0 ∆A

Dividing the magnitude of internal force ∆F by the acting area ∆A, we obtain the stress.
If we let ∆A approach zero, we obtain the stress at a point. In general, the stress could
vary in the body, which depends on the position that we are concerning.

Normal and Shear Stress


As we known, force is a vector that has both magnitude and direction. But in the stress
definition, we only consider the magnitude of the force so far. Obviously, this may easily
confuse us. Let’s still take patch ∆A as an example. As we can see, force ∆F is not
perpendicular to the sectioned infinitesimal area ∆A. If we only take the magnitude of the
force into account, apparently, the stress may not reflect the real mechanical status at this
point. In other words, we need to consider both magnitude and direction of the force.
Now let’s resolve the force ∆F in normal (∆Fn) and tangential (∆Ft) directions of the
acting area as Fig. 1.1b. The intensity of the force or force per unit area acting normally
to section ∆A is called Normal Stress, σnn, and it is expressed as:
∆F
σ nn = lim n (1.2)
∆A→0 ∆A

If this stress “pulls” on the area it is referred as Tensile Stress and defined as Positive. If
it “pushes” on the area it is called Compressive Stress and defined as Negative.
The intensity or force per unit area acting tangentially to ∆A is called Shear Stress, τnt
(tau), and it is expressed as:
∆Ft
τ nt = lim (1.3)
∆A→0 ∆A

Unit of stress: N/m2 or Pa (Pascal). In engineering practice: KPa=103Pa, MPa=106Pa,


GPa=109Pa are used generally.

1.3 Stress Sign Convertion


Obviously, the elementary notation described above is not sufficiently flexible and
convenient for use in general, because (1) the direction of surface ∆A can change, and (2)
there are infinite tangential directions on a specific surface. i.e. the normal stress σnn can
vary with the direction change of n and shear stress τnt can be in any tangential direction
of the surface.

2
Cartesian coordinate system
However, we can recall that we often solve the engineering problems under a reference
coordinate system, for instance, a Cartesian coordinate system xyz as show in Fig. 1.2.
Clearly it will be convenient to discuss stress at a point of interest P on an infinitesimal
plane through P with its external normal n in one of the direction along a reference
coordinate.

For example, we can consider an infinitesimal sectional plane through P in coordinate z,


where n is coincident with z. Then the traction can be resolved along these three axes as
On a z-sectional plane:
 ∆F   dF   dFx + dF y + dFz   dFx dF y dFz 
σ = lim   =  =   =  + +  = (σ x i + σ y j + σ z k )z
∆A→0 ∆A
  z  dA  z  dA  z  dA dA dA  z

Resolution direction
(i.e.coordinate directions)

= σ zx i + σ zy j + σ zz k
z - sectional plane (1.4)

The first suffix of a stress component indicates the direction of sectional plane. The
second denotes the direction of the stress component along the coordinate.
Positive/negative planes:
If normal is the same as coordinate direction, this plane is positive. Otherwise negative
Similarly, we can have other infinitesimal planes through P in the direction of other
coordinates. Positive/negative z; Positive/negative y; Positive/negative x; totally six
planes.
F4

F3
z
F5 z
+ z

σ zz n
Positive
Z-plane P z-plane
P
P σ zy Negative
σ zx y z-plane
y −n y
o o o
x x x
F1
Normal in the same Normal in the opposite
F2 direction of coordinate (z) direction to coordinate (z)
Fig. 1.2 Stress in Cartesian coordinate

1.4 Representation of infinitesimal cube

3
For the sake of convenient presentation, we often use an infinitesimal cube formed by the
six infinitesimal planes mentioned above as shown in Fig. 1.3. On each plane, we have
one normal stress component and two shear stress components.
z σ zz
σ zy
σ zx σ yz
σ xz σ yy
σ xy
σ yx
σ xx
y
o
x
Fig. 1.3 Stress in infinitesimal cube

We can arrange the stress component in a form of matrix (or tensor)


Direction of stress component
(coordinate direction)
x y z
Plane normal to x σ xx σ xy σ xz
Plane normal to y σ yx σ yy σ yz
Plane normal to z σ zx σ zy σ zz

Thus from above, we know that the stress state at Point P should be expressed by nine
stress components. In engineering, we denote them as “stress matrix” or “stress tensor”:
σ xx σ xy σ xz  σ xx τ xy τ xz 
   
σ yx σ yy σ yz  or τ yx σ yy τ yz 
σ zx σ zy σ zz   τ zx τ zy σ zz 
   

Fig. 1.4 Resultant force on area ∆A

4
x plane y plane z plane

Fig. 1.5 Stresses acting on an ‘x, y, ans z’ plane at point Q

Example 1.1
Draw the stress states at two different points in a machine component measured as
16 18 0  − 19 20 0
A = 18 17 − 15 B =  20 − 25 0 
   
 0 − 15 19   0 0 20
Soln: To reflect the stress matrix to an infinitesimal cube, you can first determine the
corresponding notation.
16 18 0  σ xx σ xy σ xz 
 
A = 18 17 − 15 = σ yx σ yy σ yz 
 
 0 − 15 19  σ zx σ zy σ zz 
Thus σ xx = 16, σ xy = 18, σ xz = 0 , which means in x-section (ie. front face), the three
components long x, y, and z directions can be determined. Thus, in the x-section, draw
positive 16 in x-direction, positive 18 in y-direction, and 0 in z-direction as shown in the
front face of the left infinitesimal element. Similarly, y and z plane stresses can be drawn
as shown:
19 z-section 20 z-section
z z
15

18 18 17 20 25
15 20
x-section 16 x-section 19
y-section y-section
y y
o o
x x

1.6 Sign of Stresses


It is necessary to define positive and negative sense of stresses for convenience.

5
Positive direction of normal stress.
Consider normal stress σnn on an infinitesimal plane ∆An, whose external normal is n, We
define that σnn is positive if its direction is in the normal direction of section, as in Fig.
1.7 (left). Otherwise, negative (right).

Tensile = positive Compressive= negative



z z
+
y y
o o
n n
x x
σ xx positive σ xx negative
Fig. 1.7 Sign of normal stress (left – the same direction as the normal direction of section
plane; right – opposite to the normal direction of section plane)
Positive direction of shear stress.
For a shear stress σnt in the infinitesimal plane ∆An where t is a tangential direction of
∆An.
In the positive plane (where the external normal n of ∆An has the same direction as a
coordinate axis n), the positive σnt should have the same direction as coordinate axis t
(i.e. y in Fig. 1.8a) In the negative plane, the positive σnt should have the opposite
direction to coordinate axis t (i.e. y in Fig. 1.8d).


z z
(a)
+ (b)

y y
o σ xy positive o σ xy negative
x n x n
(Positive-section) (Positive-section)

(c)
z
− −n (d)
z
+ −n
negative positive
σ xy σ xy

y y
o o
x (Negative-section)
x (Negative-section)
Fig. 1.8 Sign of shear stress

6
1.5 Symmetry of the stress matrix

Are they all independent? Or Can we use a smaller number of stress component to
facilitate the description of stress state?

Let’s check the infinitesimal element shown before. If we cut the infinitesimal element in
the middle, i.e. a z-section as shown in dashed line. We can have the sectional model on
the right hand side, a 2D version of infinitesimal element.

Fig. 1.9 Equilibrium of element

In the 2D element, the moment equation can be written as:


 1   1   1   1 
τ xy (∆y )(∆z ) ∆x  + τ xy (∆y )(∆z ) ∆x  − τ yx (∆x )(∆z ) ∆y  − τ yx (∆x )(∆z ) ∆y  = 0
 2   2   2   2 
∴τ xy − τ yx = 0
Thus ∴τ xy = τ yx (Or: σ xy = σ yx )
Similarly by checking equilibrium conditions in the yz- and xz planes, we can have
τ xz = τ zx (Or: σ xz = σ zx )
τ yz = τ zy (Or: σ yz = σ zy )
Thus we have shown that the stress matrix or stress tensor is symmetrical, i.e.
τ ij = τ ji (where i, j = x , y , z ) (1.5)
Only SIX independent stress components are needed to describe the stress state at a
point.

σ xx 
 
τ yx σ yy 
τ zx τ zy σ zz 
 

If a shear stress exists on any plane, there must also be a shear stress of the same
magnitude acting on the orthogonal plane (i.e. a perpendicular plane)

Example 1.2
(1) Draw an infinitesimal cube to show the stress tensor σA

7
 10 0 − 40
σ A = 0 − 30 0 

 
 − 40 0 10 
Soln:
10
+z-plane
-40

-30

z -40
10 +y-plane

o y +x -plane
x

(2) Write the stress tensor from the stressed infinitesimal cube (note the signs of the shear
stress are not given in the figure and you need to decide them):
60 60
+x-plane
40 +z-plane 40

60 60
20 20 40 20 20 40
z z
y 60 y 60
- y-plane
o x o x

Soln:
Look at the planes in the infinitesimal cube. Obviously, the front y-section is a negative
plane (its normal direction is opposite to y-positive). Other two faces shown are positive
planes:
x-plane (positive):
σ xx = −60(compression), σ xy = −20(opposite to y ), σ xz = −40 (opp to z )
y-plane (negative): σ yx = −20( same as x ), σ yy = 60(tension or same as y ), σ yz = 0
z-plane (positive): σ zx = −40(opposite to x ), σ zy = 0, σ zz = 60 (tension)
So the stress tensor can be written as:
 − 60 − 20 − 40
σ B = − 20 60 0 
 
− 40 0 60 

Plane stress (2D)


The stress are only present in the x and y direction here, σ z = τ zx = τ zy = 0 and from Eq
1.5 we have τ xz = τ yz = 0 . Therefore, only σ xx , σ yy , τ yx and τ xy are present as shown in
Fig. 1.10.

8
Fig. 1.10

1.6 Stress Transformation (General Equations)


Stresses in any direction (2D)
Cut a triangular section, leaving the left and bottom sides and a third side inclined at an
angle θ from the vertical. Two of its surfaces have the normals in the x and y directions;
the third has a normal at an angle θ from the x axis, as shown in Fig. 1.11 (right).
σyy t n
τtn θ
τxy σnn
y
θ y’ σy’y’ σyy
σx’y’
σxx σxx σx’x’
Acosθ

σxx θ A x’
θ θ σxy
τxy Asinθ
y y σxx
τyx τ yx θ θ θ+90° x
σyy θ σ o
x x yy

Fig. 1.11 Stress in different direction Fig. 1.12 Stresses with coordinate rotation

It is now necessary to apply the equilibrium equations about the Normal n & Tangent t
axes.
∑ Fn = 0 = Aσ nn − (σ1yy42
A sin θ ) sin θ − (σ xx A cosθ ) cosθ − (τ yx A sin θ ) cosθ − (τ xy A cosθ ) sin θ
43 14243 14243 14243
Fy Fx Vx Vy

Since τ xy = τ yx , which simplifies to give:


σ nn = σ xx cos 2 θ + σ yy sin 2 θ + 2τ xy cos θ sin θ (1.6a)
Using the following trigonometric functions:
cos 2 θ = (1 + cos 2θ ) sin 2 θ = (1 − cos 2θ)
1 1
sin 2θ = 2 cos θ sin θ
2 2
we can obtain:
(
σ xx + σ yy ) (
σ xx − σ yy )
σ nn = + cos 2θ + τ xy sin 2θ (1.6b)
2 2
And in a similar way, by applying equilibrium in tangential (t) axis and using the
trigonometric functions we can get:

9
∑ Ft = 0 = Aσ nt − (σ1yy42
A sin θ ) cosθ + (σ xx A cosθ ) sin θ + (τ yx A sin θ ) sin θ − (τ xy A cosθ ) cosθ
43 14243 14243 14243
Fy Fx Vx Vy

σ nt =
1
2
[ 1
2
]
2σ yy sin θ cosθ − [2σ xx sin θ cos θ ] − τ yx sin 2 θ + τ xy cos2 θ

Use trigonometrics, we can have: cos2 θ − sin 2 θ = (1 + cos 2θ ) − (1 − cos 2θ ) = cos 2θ


1 1
2 2
τ tn =
(
σ yy − σ xx )
sin 2θ + τ xy cos 2θ (1.7)
2
Based on these above two equations, we can determine the stress in any plane.

Stresses with coordinate axis rotation


Let’s consider 2D stress state undergoing coordinate rotation, from xoy to x’oy’ (Fig.
1.12).
σ x 'x ' can be determined from the previous section where the coordinate x’ is coincide
with normal n. Thus the stress in an inclined plane of θ can be calculated by
σ x ' x ' = σ xx cos2 θ + σ yy sin 2 θ + 2σ xy cosθ sin θ (1.8)
σ y ' y ' can be determined by viewing the inclined plane with angle of (θ+90°)
σ y ' y ' = σ xx cos2 (θ + 90°) + σ yy sin 2 (θ + 90°) + 2σ xy cos(θ + 90°) sin(θ + 90°)
= σ xx sin 2 θ + σ yy cos2 θ + 2σ xy (− sin θ ) cos θ
= σ xx sin 2 θ + σ yy cos2 θ − 2σ xy sin θ cos θ
Thus σ y ' y ' = σ xx sin 2 θ + σ yy cos2 θ − 2σ xy sin θ cosθ (1.9)

Similarly τ x' y' =


(σ yy − σ xx ) sin 2θ + σ cos 2θ (1.10)
xy
2

Remarks:
Let’s add (1.8) to (1.9)
( ) ( )
σ x ' x ' + σ y ' y ' = σ xx sin 2 θ + cos2 θ + σ yy sin 2 θ + cos2 θ + 2σ xy sin θ cosθ − 2σ xy sin θ cosθ

Thus: σ x ' x ' + σ y ' y ' = σ xx + σ yy


which means that the summation of two normal stress components is independent on
the rotation of coordinate system.

Example 1.3
Rotate the following stress tensors about z-axis for θ=90o (units MPa).

10
 1 − 1 0
 
[σ ] =  − 1 2 0  .
 0 0 3
 
Soln:
Since the rotation is about z-axis, we do not need to change z-directional stress
σ x ' x ' = σ xx cos2 θ + σ yy sin 2 θ + 2σ xy cosθ sin θ
= (1) × cos2 90° + (2) × sin 2 90° + 2 × ( −1) × cos 90° sin 90°
= (1) × 0 + ( 2) × 1 + 2 × ( −1) × 0 × 1 = 2
σ y ' y ' = σ xx sin 2 θ + σ yy cos2 θ − 2σ xy sin θ cosθ
= (1) × sin 2 90° + ( 2) × cos2 90° − 2 × ( −1) × sin 90° × cos 90° = 1
( σ − σ xx
τ x ' y ' = yy
) sin 2θ + σ xy cos 2θ
2
=
(2 − 1) sin(2 × 90°) + (− 1)cos(2 × 90°) = 1 sin(180°) + (− 1)cos(180°) = 1
2 2
σ z ' z ' = σ zz = 3 (rotates about z-axis)
Thus
 2 1 0
 
[σ ' ] =  1 1 0  MPa .
 0 0 3
 

Example 1.4
At a point on a structural member subjected to plane stress, normal and shear stresses (50
MPa) exist on horizontal and vertical planes through the point as shown. Use the stress
transformation equations to determine the normal and shear stress on the indicated plane
surface.

σ xx = +16MPa since the normal stress acting on the x face creates tension in the element
σ yy = +42MPa
τ xy = −50MPa since the shear stress on the positive x face acts on the negative y
direction
τ yx = −50MPa
Tan θ = 8/5; θ = +58o. Since x & y axis are rotated CCW to the n & t axis.

11
(σ xx + σ yy ) (σ xx − σ yy )
σ nn = + cos 2θ + τ xy sin 2θ
2 2

σ nn =
(16 + 42) + (16 − 42) cos 2(58) + −50 sin 2(58)
2 2

= -10.24Mpa

τ tn =
(σ yy − σ xx ) sin 2θ + τ cos 2θ
xy
2

τ tn =
(42 − 16) sin 2(58) + −50 cos 2(58)
2

= 33.6 MPa

Example 1.5
The grain of a wooden member forms an angle of 150 with the vertical. For the state of
stress shown, determine (a) the in-plane shearing stress parallel to the grain, (b) the
normal stress perpendicular to the grain.

σ xx = 0
σ yy = 0
τ xy = 4.2MPa
τ yx = 4.2MPa
θ = - 15o. Since x & y axis are rotated CW to the n & t axis.

12
(σ xx + σ yy ) (σ xx − σ yy )
σ nn = + cos 2θ + τ xy sin 2θ
2 2

σ nn =
(0 + 0) + (0 − 0) cos 2(−15) + 4.2 sin 2(−15)
2 2

= -2.1 Mpa

τ tn =
(σ yy − σ xx ) sin 2θ + τ cos 2θ
xy
2

τ tn =
(0 − 0) sin 2(−15) + 4.2 cos 2(−15)
2

= 3.64 MPa

1.7 Principal stresses and maximum shear stress


The stress transformation equations 1.6 to 1.10 gives the normal stress σ nn and the shear
stress τ tn acting on any plane through a point on a stressed body. For design purposes,
the critical stresses are important that is, the maximum and minimum normal stress and
the maximum shear stress. Equations 1.6 to 1.10 can be modified to obtain there results.

The derivations are not important here, however, the following single equations gives the
two in-plane principal stresses σ p1 and σ p 2 :

σ p1, p 2 =
(σ xx + σ yy )  (σ − σ yy ) 
±  xx
2

 + (τ xy )2
2  2 

And the value of θ for which the normal stress is maximum or minimum can be
determined by the equation:

τ xy
tan 2θ p =
(σ x − σ y ) / 2

Note:
(1) There will be only two planes where either a maximum or a minimum normal
stress occurs;
(2) That there two planes will be 90o apart (ie orthogonal to each other)
(3) If σ xx − σ yy is +ve, θp indicates the orientation of σ p1
(4) If σ xx − σ yy is -ve, θp indicates the orientation of σ p 2

13
The maximum inplane shear stress is given by

 (σ − σ yy ) 
2

τ max = ±  xx  + (τ xy )2
 2 

And the value of θ for which the shear stress is maximum or minimum can be determined
by the equation:
(σ x − σ y ) / 2
tan 2θ s =
τ xy

This means that θp and θs are 45o apart. The plane on which the maximum in plane shear
stresses occur are rotated 45o from the principal planess.

Example 1.6
Consider a point in a structural member that is subjected to plane stress. Normal and
shear stresses acting on horiziontal and vertical planes at the point are shown.

(a) Determine the principal stresses and the maximum in plane shear stress acting at a
point

In plane principal stresses

σ p1, p 2 =
(σ xx + σ yy )  (σ − σ yy ) 
±  xx
2

 + (τ xy )2
2  2 

(70 + 150) ±  (70 − 150) 


2

σ p1, p 2 =  + (− 55)
2

2  2 

= 178.0 MPa and 42.0 MPa


Maximum in plane shear stress

14
 (σ − σ yy ) 
2

τ max = ±  xx  + (τ xy )2
 2 

 (70 − 150) 
2

τ max = ±   + (− 55)
2

 2 

= ± 68.0 MPa

Average normal stress


σ + σ yy 70 + 150
σ avg = xx = = 110 MPa (T)
2 2

(b) Show these stresses in an appropriate sketches


τ xy − 55 − 55
tan 2θ p = = =
(σ x − σ y ) / 2 (70 − 150) / 2 − 40

θp = 27o.

- The angle is +ve hence it is turned CCW from x axis;


- Since σ xx − σ yy is –ve, θp indicates the orientation of σ p 2
- The other principal stress σ p1 (178 MPa) acts on a perpendicular plane
- The maximum in-plane shear stress is shown on the sloped face of the wedge
which is rotated 45o.

1.8 Strain at a Point


Two strains: There are two types of deformations.
Normal strain Change relative distance of two ends of a segment Direct strain:
l − l0
ε=
l0

15
If l increases (tensile) strain is positive, if l reduces (compressive) strain is negative.
Shear strain Change the included angle of two segments: Shear strain:
1 1
γ = tanψ ≈ ψ
2 2
Note that if shear causes the angle reduction – shear strain is positive; If shear causes the
angle increase – shear strain is negative.

Fig. 1.13

The complete state of strain at a point in a body under load can be determined by
considering the deformation associated with a small volume of material surrounding the
point (Fig. 1.13). For this case deformations are assumed to be small, i.e,
a) Planes initiall parallel to each other will remain parallel after deformation
b) Straight lines before deformation will remain straight after deformation
In Cartesian components the strain at a point can be expressed as

d x/ − d x π
εx = γ xy = − θ xy/
dx 2
d − dy
/
π
εy = y
γ yz = − θ yz/
dy 2
d z/ − d z π
εz = γ zx = − θ zx/
dz 2

In a similar manner the normal strain and shear strain component in n, t component
system can be expressed as

d n/ − d n π
εn = γ nt = − θ nt/
dn 2

Plain Strain (2D) (Fig. 1.14)


The condition of plane strain is that ε z , γ xz and γ yz = 0.

16
Fig. 1.14

1.9 Strain Transformation (General Equations)


In a similar way to stress, strain can be transformed to determine the maximum direct and
shear strains experienced by the structure. What you find when you do the strain
transformation analysis is that the equations for strain transformation are very similar to
the equations for stress transformation. Taking 2D as an example as in Fig. 1.15:

y
y’ Deformed Element

dy Original Element
x
dx x’
x’ v’ u’
y’
y
θ x
Fig. 1.15 Strain transformation

ε nn = ε xx cos 2 θ + ε yy sin 2 θ + γ xy cos θ sin θ (1.11)

ε nn =
(ε xx + ε yy )
+
(ε xx − ε yy )
cos 2θ +
γ xy
sin 2θ (1.12)
2 2 2

γ tn
=
(ε yy − ε xx )
sin 2θ +
γ xy
cos 2θ (1.13)
2 2 2

1.10 Principal strains and maximum shear strain


The stress transformation equations 1.6 to 1.10 gives the normal stress σ nn and the shear
stress τ tn acting on any plane through a point on a stressed body. Similarily, equations
1.11 to 1.13 gives the normal strain and the shear strain acting on any plane through a
point. All of the equations and derivations carried out for stress transformation can be
applied here as well.

17
The derivations are not important here, however, the following single equations gives the
two in-plane principal strain ε p1 and ε p 2 :

ε p1, p 2 =
(ε xx + ε yy )  (ε − ε yy )   γ xy 
±  xx
2

 +  
2

2  2   2 

And the value of θ for which the normal strain is maximum or minimum can be
determined by the equation:

γ xy
tan 2θ p =
(ε x − ε y )

Note:
(1) There will be only two planes where either a maximum or a minimum normal
strain occurs;
(2) That there two planes will be 90o apart (ie orthogonal to each other)
(3) If ε xx − ε yy is +ve, θp indicates the orientation of ε p1
(4) If ε xx − ε yy is -ve, θp indicates the orientation of ε p 2

The maximum inplane shear strain is given by

 (ε − ε yy )   γ xy 
2 2
γ max
= ±  xx  +  
2  2   2 

Example 1.7
At point p of the bearing’s housing is subjected to a state of plane stress, and the strain
components εx = −0.0024, εy = 0.0044, and γxy = −0.0030. If θ = 20◦, what are the strains
εx/, εy/ and γxy/ at p?

18
1.10 Strain Rosette
A special application of strain transformation is “strain rosette”. Using Eq. (1.11), the
strain at any angle can be determined. Inversely, if the strain at any angle θ has been
measured, Eq. (1.11) can then be used to determine the direct and shear strains in
the structure about the x and y axes. One of the typical approaches is that these
measurements are done using a Strain Gauge Rosette. A normal arrangement is to have
three strain gauges oriented at three different angles w.r.t the horizontal axis of the
structure as in Fig. 1.16. From experiment, we can acquire three sets of data as εθ1, εθ2,
εθ3.
Based on these three sets of data, we want to determine the normal and shear strains
about the x and y axes. Because we have three unknown terms and aim to find, εxx, εyy γxy,
we can use Eq. (1.11) three times, once for each angle. Then simultaneously solve for the
three unknown strain terms, εxx, εyy γxy. The procedure is as follows.

y 3
2

θ3 θ2 θ1
x

19
Fig. 1.16 Strain Gauge rosettes

Solve for the three unknown strains, εxx, εyy, γxy from the three measurements of εθ1, εθ2,
εθ3.
 ε θ 1 = ε xx cos 2 θ1 + ε yy sin 2 θ1 + γ xy cos θ1 sin θ1

ε θ 2 = ε xx cos θ 2 + ε yy sin θ 2 + γ xy cos θ 2 sin θ 2
2 2

 ε = ε cos 2 θ + ε sin 2 θ + γ cos θ sin θ


 θ3 xx 3 yy 3 xy 3 3

Example 1.8
A “thick bar” is subjected to a set of external loads within the xy-plane as illustrated in
Fig. 1.14. Using the strain gauge rosette shown in the figure, the direct strains at a point
are measured to be ε A = −1 × 10−4 , ε B = 1 × 10−4 and ε C = 1.8 × 10 −4 . Determine ε xx , ε yy ,
ε zz .
y

60°
C B 120°
60° 60°
A

x
Solution:
ε A = ε nn (θ = 0) = ε xx cos 2 θ + ε yy sin 2 θ + γ xy cos θ sin θ
= ε xx cos 2 0° + ε yy sin 2 0° + γ xy cos 0° sin 0° = ε xx
ε B = ε nn (θ = 120°) = ε xx cos 2 120° + ε yy sin 2 120° + γ xy cos120° sin 120°
2
 1
2
 3 1  3 1
 + γ xy  − 
3 3
= ε xx  −  + ε yy  
 = ε xx + ε yy −
 2  4 γ xy
 2  2   2   4 4
ε C = ε nn (θ = 60°) = ε xx cos 60° + ε yy sin 60° + γ xy cos 60° sin 60°
2 2

2
1
2
 3 1  3 1
 + γ xy  
3 3
= ε xx   + ε yy   
 = ε xx + ε yy +
 γ xy
2  2   2  2  4 4 4

ε xx = −1
 ε xx = −1 × 10 −4
1 3 3 
Thus:  ε xx + ε yy − γ xy = 1 (Unit × 10−4 ) ε yy = 2.2 × 10
−4

 4 4 4 
γ xy = 0.924 × 10
−4
1 3 3
 ε xx + ε yy + γ xy = 1.8
4 4 4

20
ε xx γ xy γ xz   − 1 0.92 0
 
Strain tensor thus is: ε = γ yx ε yy γ xz  = 0.92 2.2 0 × 10 −4
γ zx γ zy ε zz   0 0 0

Example 1.9
A long bar section shown in figure below is under plane strain state in the xy-plane.
Using the strain gauge technique, the normal strains at a surface point in the directions as
described in the figure are obtained as ε0=8×10-6, ε60=11×10-6 and ε90=4×10-6. Find the
expressions of principal strains at the point, ε1 and ε2.

Solution:
Step 1 Determine strain ε xx , ε yy , ε xy
ε 0° = ε nn (θ = 0) = ε xx cos 2 θ + ε yy sin 2 θ + γ xy cos θ sin θ
= ε xx cos 2 0° + ε yy sin 2 0° + γ xy cos 0° sin 0° = ε xx = 8 × 10 -6

ε 90° = ε nn (θ = 90°) = ε xx cos 2 90° + ε yy sin 2 90° + γ xy cos 90° sin 90°
= ε xx (0 ) + ε yy (1) + γ xy (0 )(1) = ε yy = 11 × 10 -6
2 2

ε 60° = ε nn (θ = 60°) = ε xx cos 2 60° + ε yy sin 2 60° + γ xy cos 60° sin 60°
2
1
2
 3 1  3 1
 + γ xy  
3 3
= ε xx   + ε yy  
 = ε xx + ε yy +
 2  4 γ xy
2  2   2   4 4
Solve for the equation we have:
4  1 3 
∴ γ xy =  ε 60° − ε 0° − ε 90°  = 13.86 × 10
-6

3 4 4 

Step 2: calculate the principal strains:

21
2
ε xx + ε yy  ε − ε yy   γ xy 
2

ε 1,3 = ±  xx  +  
 
2  2   2 
2
 4  1 3 
  ε 60° − ε 0° − ε 90°  
ε 0° + ε 90° ε −ε 
2

±  0° 90°  +  3  
4 4
=
2  2   2 
 

8+4  8 − 4   13.86 
2 2

= ±   + 
2  2   2 
13.21
= 6 ± 52 = 
− 1.211

Step 3: Rank them in 3D, thus


ε1 = 13.21 × 10−6
ε2 = 0
ε 3 = −1.211 × 10−6

1.11 Isotropic Material:


Stress-strain relations are the same in all directions. Only two constants are
independent.
σ 3 ε3
σ 1 ε1

σ2 ε2 σ 2 ε2

σ 1 ε1
σ 3 ε3
Fig Fig.
1.173.3 Infinitesimalcube
Infinitesimal in thein
principal directions
principal directions

Consider an infinitesimal element of an isotropic material (Fig. 1.17). Let’s assume its
surfaces are the principal stress and strain directions, where all shear stress/strain are
zero. We can simplify by presenting only non-zero components in the principal planes, as
σ 1  C11 C12 C13   ε1 
    
σ 2  = C21 C22 C23  ε 2 
σ  C  
 3   31 C32 C33  ε 3 

22
σ 1 = C11ε1 + C12ε 2 + C13ε 3

i.e. σ 2 = C21ε1 + C22ε 2 + C23ε 3
σ = C ε + C ε + C ε
 3 31 1 32 2 33 3
Since the material is isotropic (which means that material behaves the same in different
directions), the magnitude of σ 1 should not change if we interchange ε 2 and ε 3 .
σ 1 = C11ε1 + C12ε 2 + C13ε 3 = C11ε1 + C12ε 3 + C13ε 2 ,
C12ε 2 + C13ε 3 = C12ε 3 + C13ε 2
C12 (ε 2 − ε 3 ) + C13 (ε 3 − ε 2 ) = 0
(C12 − C13 )(ε 2 − ε 3 ) = 0
In general, ε 2 ≠ ε 3 , thus C12 = C13
Similarly, we can derive C12 = C23 and C13 = C23 which gives
C12 = C21 = C23 = C32 = C13 = C31
Since isotropic behaviours, the Young’s modulus should be the same in all directions.
Thus
C11 = C22 = C33
We only need two independent parameters to determine isotropic material
properties.

Generalised Hooke’s Law using E and ν


For isotropic material elastic modulus E and v are same in every direction.

Fig 1.18

In Fig. 1.18 the differential element is subjected to three different normal stresses
( σ xx , σ yy and σ zz ). Although if only σ xx is applied to the cube, normal strains are
produced in the y and z directions because of the Poisson’s effect. Note that these strains
in the transverse direction are negative. Hence, for all three forces acting alone we have;

σ xx σ xx σ xx
Strains produce by normal stress, σ xx ,  ε xx = , ε yy = −v , ε zz = −v
E E E

σ yy σ yy σ yy
Strains produce by normal stress, σ yy ,  ε yy = , ε xx = −v , ε zz = −v
E E E

23
σ zz σ zz σ zz
Strains produce by normal stress, σ xx ,  ε zz = , ε xx = −v , ε yy = −v
E E E
If all three normal stresses are present together then we have

ε xx =
1
E
[ ]
σ xx − v(σ yy + σ zz )

ε yy =
1
E
[
σ yy − v(σ xx + σ zz ) ] (1.14)

ε zz =
1
E
[ ]
σ zz − v(σ xx + σ yy )

Further, the deformations by shear stresses τ xy , τ yz and τ xz are shown in Fig. 1.19 and
can be determined by

1 1 1
γ xy = τ xy γ yz = τ yz γ zx = τ zx
G G G

Where G is the shear modulus given by

E
G=
2(1 + v )

Fig. 1.19

Equations 1.14 can be solved for stresses in terms of the strains as

24
[ ]
σ xx = (1 + ν )(1 − 2ν ) (1 − ν )ε xx + ν (ε yy + ε zz )
 E


σ yy =
(1 + ν
E
)(1 − 2ν )
[
(1 −ν )ε yy + ν (ε zz + ε xx )] (1.15)


σ zz =
E
(1 + ν )(1 − 2ν )
[
(1 −ν )ε zz + ν (ε xx + ε yy )]

Example 1.10
At a point of a material subjected to a state of plane stress, the strains measured by the
strain rosette are εa = 0.006 (-68o), εb = −0.003 (0o), and εc = −0.002 (56o). The modulus
of elasticity and Poisson’s ratio of the material are E = 30GPa and ν = 0.33. What is the
state of stress at the point?

ε xx =
1
E
[
σ xx − v(σ yy + σ zz )]

ε yy =
1
E
[
σ yy − v(σ xx + σ zz ) ]

E
where G is calculated using G = = 11.28 x 109 N/m2.
2(1 + v)

25

You might also like