You are on page 1of 10

MECH3361/9361 Semester 2, 2016

DISPLACEMENT AND STRAIN


2.1. Displacement
2.1.1. Two types of displacement
When a solid is subjected to external loading, the positions of materials points in solids will
change, i.e. displacement occurs. There are two types of displacement as shown in Fig.
2.1(a).

Rigid body displacement occurs when the relative position between any points
within the solid body remain unchanged. This is a topic of Engineering Mechanics
(Statics), Engineering Dynamics I, II, and III.
Deformation occurs when the relative position between any points within the solid
body change – a topic of Mechanics of Solids I, and II.
Rigid body motion
Q

P 90°

R
Q

90°
R
P Q y
y

90°- y
R
P
0 x

Deformation 0 x

(a) (b)

Fig. 2.1 Rigid body motion and deformation

2.1.2. Deformation of a cantilever bar


If we look at the displacement in a cantilever bar shown in Fig. 2.2, the x-displacement is not
uniform, though the stress is uniform. The elongation of the bar can be calculated as:

𝑃𝑥
𝛿= = 𝑢(𝑥)
𝐸𝐴

y A B C
0 x
y

A B C
0 x

uA uB uC

Fig. 2.2 Deformation of a cantilever bar

1
MECH3361/9361 Semester 2, 2016

𝑃
Let 𝑐 = 𝐸𝐴 denote a constant. Thus:

𝑃
𝑢𝐴 = 𝑢(𝑥 = 𝑥𝐴 ) = ( ) 𝑥 = 𝑐𝑥𝐴
𝐸𝐴 𝐴
𝑃
𝑢𝐵 = 𝑢(𝑥 = 𝑥𝐵 ) = ( ) 𝑥𝐵 = 𝑐𝑥𝐵
𝐸𝐴
𝑃
𝑢𝐶 = 𝑢(𝑥 = 𝑥𝐶 ) = ( ) 𝑥𝐶 = 𝑐𝑥𝐶
𝐸𝐴

In general, the displacements in a solid subjected to deformation are non-uniform and would
be some function of co-ordinates. In a general 2D case (x-y case), 𝑢 = 𝑢(𝑥, 𝑦) and 𝑣 =
𝑣(𝑥, 𝑦), where u and v are the displacements in x and y directions, respectively. In 3D, 𝑢 =
𝑢(𝑥, 𝑦, 𝑧), 𝑣 = 𝑣(𝑥, 𝑦, 𝑧), and 𝑤 = 𝑤(𝑥, 𝑦, 𝑧), where w is the displacement in the z direction.
Note that we ignore rigid body displacement in Mechanics of Solids I and II.

2.2. Strain
2.2.1. Normal and shear strain
Similar to stress, strains can be classified as either normal or shear strains. Normal strain is
the relative change in distance between two ends of a segment of original length l0.

𝑙 − 𝑙0
𝜀=
𝑙0

(2.1)

If the segment is in tension, then 𝑙 > 𝑙0 and the normal strain is positive. In compression, 𝑙 <
𝑙0 , resulting in a negative normal strain.

Shear strain is defined by the change in the included angle of two segments.

1 1
𝛾 = tan 𝜓 ≈ 𝜓
2 2

(2.2)

If shear causes a reduction in the included angle, then shear strain is positive. If shear causes
an increase in the included angle, then shear strain is negative.

Note that the displacement at a point is easy to measure, however, strain is more difficult to
measure, as we require at least two measurement points in the solid. The accuracy of a point
strain measurement will therefore be dependent on the distance between these two points.

2
MECH3361/9361 Semester 2, 2016

2.3. Strain-displacement relationship


Two short bars (PQ and PR) undergo in-plane deformation in the xy-plane as shown in Fig.
2.3. After deformation, the displacements at P are assumed to be 𝑢 = 𝑢(𝑥, 𝑦) and 𝑣 =
𝑣(𝑥, 𝑦). The corresponding x and y displacements at points R and Q are:

𝜕𝑢(𝑥, 𝑦)
𝑢 + 𝛥𝑢 = 𝑢 + 𝛥𝑥
𝜕𝑥

𝜕𝑣(𝑥, 𝑦)
𝑣 + 𝛥𝑣 = 𝑣 + 𝛥𝑦
𝜕𝑦

Normal strains can be expressed as:

𝜕𝑢
𝑃′ 𝑅 ′ − 𝑃𝑅 (𝛥𝑥 + 𝑢 + 𝛥𝑥 − 𝑢) − 𝛥𝑥 𝜕𝑢
𝜀𝑥𝑥 = lim [ ] = lim [ 𝜕𝑥 ]=
𝑥→0 𝑃𝑅 𝑥→0 𝛥𝑥 𝜕𝑥

(2.3)

𝜕𝑣
𝑃′ 𝑄 − 𝑃𝑄 (𝛥𝑦 + 𝑣 + 𝛥𝑦 − 𝑣) − 𝛥𝑥 𝜕𝑣
𝜕𝑦
𝜀𝑦𝑦 = lim [ ] = lim [ ]=
𝑦→0 𝑃𝑄 𝑦→0 𝛥𝑦 𝜕𝑦

(2.4)

y Q’

v+ v
Q R’

y u
P’ u+ u
v
P x
x R

Fig. 2.3 Displacement and normal strain

3
MECH3361/9361 Semester 2, 2016

y
u+ u
u Q’
Qy’

Q y R’

v
y u x
P’ Rx’ v+ v
v
P x
x R

Fig. 2.4 Displacement and shear strain

Shear strain can be expressed as (Fig. 2.4):

𝛥𝑣 𝛥𝑢
𝜓 = 𝜓𝑥 + 𝜓𝑦 ≈ tan 𝜓𝑥 + tan 𝜓𝑦 = +
when 𝜓𝑥 and 𝜓𝑦 are small 𝛥𝑥 𝛥𝑦

The average angle change is:

1 1 𝛥𝑣 𝛥𝑢 1 𝜕𝑣 𝜕𝑢
𝛾= lim tan 𝜓 = lim ( + ) = ( + )
2 𝛥𝑥→0 𝛥𝑥→0 2 𝛥𝑥 𝛥𝑦 2 𝜕𝑥 𝜕𝑦
𝛥𝑦→0 𝛥𝑦→0

(2.5)

Thus the strain-displacement relationship in 2D is:

𝜕𝑢
𝜀𝑥𝑥 =
𝜕𝑥
𝜕𝑣
𝜀𝑦𝑦 =
𝜕𝑦
1 𝜕𝑣 𝜕𝑢
𝜀𝑥𝑦 = ( + )
2 𝜕𝑥 𝜕𝑦

(2.6)

For a 3D case, the strain-displacement relationship is:

𝜕𝑢 𝜕𝑣 𝜕𝑤
𝜀𝑥𝑥 =
, 𝜀𝑦𝑦 = , 𝜀𝑧𝑧 =
𝜕𝑥 𝜕𝑦 𝜕𝑧
1 𝜕𝑣 𝜕𝑢 1 𝜕𝑤 𝜕𝑣 1 𝜕𝑢 𝜕𝑤
𝜀𝑥𝑦 = ( + ) , 𝜀𝑦𝑧 = ( + ) , 𝜀𝑥𝑧 = ( + )
2 𝜕𝑥 𝜕𝑦 2 𝜕𝑦 𝜕𝑧 2 𝜕𝑧 𝜕𝑥
(2.7)

4
MECH3361/9361 Semester 2, 2016

Example 2.1
A rectangle parallelepiped of infinitesimal dimension lx, ly, lz is subjected to two different
states of small strain expressed by the strain tensors (matrices). Determine the relative
volume changes of the element during these deformations.

𝜀𝑥𝑥 𝜀𝑥𝑦 𝜀𝑥𝑧 10 2 4


𝜀
𝜀1 = [ 𝑦𝑥 𝜀𝑦𝑦 𝜀𝑦𝑧 ] = [ 2 −6 8 ] × 10−5
𝜀𝑧𝑥 𝜀𝑧𝑦 𝜀𝑧𝑧 4 8 15

−10 2 4
𝜀2 = [ 2 −6 8 ] × 10−5
4 8 15

Solution

Initial volume before deformation is: 𝑉0 = 𝐼𝑥 𝐼𝑦 𝐼𝑧 .

After deformation: 𝑉 = [𝐼𝑥 (1 + 𝜀𝑥𝑥 )][𝐼𝑦 (1 + 𝜀𝑦𝑦 )][𝐼𝑧 (1 + 𝜀𝑧𝑧 )]

V=[lx(1+ x)] [ly(1+ y)cos


cos 1

ly(1+ y) l
y

lx
lx(1+ x)

Since the deformations are small, the relative angle change can be ignored.

𝑉 = 𝑉0 (1 + 𝜀𝑥𝑥 )(1 + 𝜀𝑦𝑦 )(1 + 𝜀𝑧𝑧 )


= 𝑉0 (1 + 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧 + 𝜀𝑥𝑥 𝜀𝑦𝑦 )(1 + 𝜀𝑥𝑥 )

= 𝑉0 (1 + 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧 + 𝜀⏟𝑥𝑥 𝜀𝑦𝑦 + 𝜀𝑦𝑦 𝜀𝑧𝑧 + 𝜀𝑥𝑥 𝜀𝑧𝑧 + 𝜀𝑥𝑥 𝜀𝑦𝑦 𝜀𝑧𝑧 )
ignore higher order results

𝑉 = 𝑉0 (1 + 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧 )

Hence the relative volume change can be expressed as a volume strain or dilatation.

𝑉 − 𝑉0
𝛥=
𝑉0
𝑉0 (1 + 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧 ) − 𝑉0
=
𝑉0
𝛥 = 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧

(2.8)

5
MECH3361/9361 Semester 2, 2016

For strain states ε1 and ε2, the volume changes Δ1 and Δ2 can therefore be calculated.

𝛥1 = 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧 = (10 − 6 + 15) × 10−5 = 1.9 × 10−5


𝛥2 = 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧 = (−10 − 6 + 15) × 10−5 = −1 × 10−5

Under strain state ε1, a positive dilatation occurs, while under strain state ε2, there is a
negative dilatation.

2.4. Strain transformation


2.4.1. Strains in any direction
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. 2.5:

𝜀𝑥 ′𝑥 ′ = 𝜀𝑥𝑥 cos 2 𝜃 + 𝜀𝑦𝑦 sin2 𝜃 + 2𝜀𝑥𝑦 cos 𝜃 sin 𝜃

(2.9)

2.4.2. Strain rosettes


A special application of the strain transformation is the strain rosette. Using Eq. (2.10), the
strain at any angle can be determined. Inversely, if the strain at any angle θ has been
measured, Eq. (2.10) 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 to measure strain using a
strain gauge rosette. The 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. 2.6. From such an
experiment, we can acquire three sets of data: εθ1, εθ2, εθ3.

y 3
2

2 1
3
x

Fig. 2.5 Strain transformation Fig. 2.6 Strain gauge rosettes

Based on these three sets of data, we want to determine the normal and shear strains about the
x and y axes. Since we have three unknown terms and aim to find, ε xx, εyy, εxy, we can use Eq.
(2.10) three times, once for each angle, simultaneously solving for the three unknown strain
terms, εxx, εyy, εxy, as shown in Eq. (2.11).

6
MECH3361/9361 Semester 2, 2016

𝜀𝜃1 = 𝜀𝑥𝑥 cos 2 𝜃1 + 𝜀𝑦𝑦 sin2 𝜃1 + 2𝜀𝑥𝑦 cos 𝜃1 sin 𝜃1


𝜀𝜃2 = 𝜀𝑥𝑥 cos 2 𝜃2 + 𝜀𝑦𝑦 sin2 𝜃2 + 2𝜀𝑥𝑦 cos 𝜃2 sin 𝜃2
𝜀𝜃3 = 𝜀𝑥𝑥 cos 2 𝜃3 + 𝜀𝑦𝑦 sin2 𝜃3 + 2𝜀𝑥𝑦 cos 𝜃3 sin 𝜃3

(2.10)

Example 2.2
A “thick bar” is subjected to a set of external loads within the xy-plane as illustrated in Fig.
2.7. Using the strain gauge rosette shown in the figure, the direct strains at a point are
measured to be 𝜀𝐴 = −1 × 10−4, 𝜀𝐵 = 1 × 10−4, and 𝜀𝐶 = 1.8 × 10−4. Determine εxx, εyy,
and εxy.

60°
C B 120°
60° 60°
A

Solution

𝜀𝐴 = 𝜀𝑛𝑛 (𝜃 = 0°) = 𝜀𝑥𝑥 cos 2 0° + 𝜀𝑦𝑦 sin2 0° + 2𝜀𝑥𝑦 cos 0° sin 0°


= 𝜀𝑥𝑥
𝜀𝐵 = 𝜀𝑛𝑛 (𝜃 = 120°) = 𝜀𝑥𝑥 cos 2 120° + 𝜀𝑦𝑦 sin2 120° + 2𝜀𝑥𝑦 cos 120° sin 120°
2
1 2 √3 1 √3
= 𝜀𝑥𝑥 (− ) + 𝜀𝑦𝑦 (− ) + 2𝜀𝑥𝑦 (− ) ( )
2 2 2 2
𝜀𝑥𝑥 3𝜀𝑦𝑦 √3𝜀𝑥𝑦
= + −
4 4 4
𝜀𝐶 = 𝜀𝑛𝑛 (𝜃 = 60°) = 𝜀𝑥𝑥 cos 2 60° + 𝜀𝑦𝑦 sin2 60° + 2𝜀𝑥𝑦 cos 60° sin 60°
2
1 2 √3 1 √3
= 𝜀𝑥𝑥 ( ) + 𝜀𝑦𝑦 ( ) + 2𝜀𝑥𝑦 ( ) ( )
2 2 2 2
𝜀𝑥𝑥 3𝜀𝑦𝑦 √3𝜀𝑥𝑦
= + +
4 4 4

Thus the three equations are: 𝜀𝑥𝑥 = −1 × 10−4


𝜀𝑥𝑥 3𝜀𝑦𝑦 √3𝜀𝑥𝑦
+ − = 1 × 10−4
4 4 4
𝜀𝑥𝑥 3𝜀𝑦𝑦 √3𝜀𝑥𝑦
+ + = 1.8 × 10−4
4 4 4

7
MECH3361/9361 Semester 2, 2016

Solving simultaneously gives the following results.

𝜀𝑥𝑥 = −1 × 10−4
𝜀𝑦𝑦 = 2.6 × 10−4
𝜀𝑥𝑦 = 0.46 × 10−4

𝜀𝑥𝑥 𝜀𝑥𝑦 𝜀𝑥𝑧 −1 0.46 0


The strain tensor can be written as: 𝜀 = [𝜀𝑦𝑥 𝜀𝑦𝑦 𝜀𝑦𝑧 ] = [0.46 2.6 0] × 10−4
𝜀𝑧𝑥 𝜀𝑧𝑦 𝜀𝑧𝑧 0 0 0

2.5. Principal strains


Similar to stress, the principal strains and their direction cosines can be determined directly
from the strain tensor using the eigenvalue method for a 3D case. The equations used for
calculating principal stresses and directions have analogous forms for strain.

2.5.1. Strain invariants


The first, second, and third strain invariants (I1,ε, I2,ε, and I3,ε) are:

𝐼1,𝜀 = 𝛥 = 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧

2 2 2
𝐼2,𝜀 = 𝜀𝑥𝑥 𝜀𝑦𝑦 + 𝜀𝑦𝑦 𝜀𝑧𝑧 + 𝜀𝑧𝑧 𝜀𝑥𝑥 − 𝜀𝑥𝑦 − 𝜀𝑦𝑧 − 𝜀𝑧𝑥

2 2 2
𝐼3,𝜀 = 𝜀𝑥𝑥 𝜀𝑦𝑦 𝜀𝑧𝑧 + 2𝜀𝑥𝑦 𝜀𝑦𝑥 𝜀𝑧𝑥 − 𝜀𝑥𝑥 𝜀𝑦𝑧 − 𝜀𝑦𝑦 𝜀𝑧𝑥 − 𝜀𝑧𝑧 𝜀𝑥𝑦

(2.11)

2.5.2. The Eigen equation


These invariants are used to solve the characteristic Eigen equation (Eqn. 2.12) for the
principal strains in 3D.

𝜀 3 − 𝐼1,𝜀 𝜀 2 + 𝐼2,𝜀 𝜀 − 𝐼3,𝜀 = 0

(2.12)

There are three roots to this equation for a 3D case. As with stress, various methods can be
used to solve for the principal strains. MATLAB can be used to determine the roots
numerically using the coefficients of the Eigen equation (which are the strain invariants
adjusted for the signs in the equation).

>> a = [1 –I1,ε I2,ε –I3,ε];


>> x = roots(a)

Alternatively, the principal strains and principal strain direction cosines can be calculated as
eigenvalues and eigenvectors directly from the strain tensor.

8
MECH3361/9361 Semester 2, 2016

>> A = [εxx εxy εxz; εyx εyy εyz; εzx εzy εzz];
>> [cosines,principals] = eig(A)

Example 2.3
Given the following strain state, determine the principal strains.

𝜀𝑥𝑥 𝜀𝑥𝑦 𝜀𝑥𝑧 −1 0.46 0


𝜀
𝜀 = [ 𝑦𝑥 𝜀𝑦𝑦 𝜀𝑦𝑧 ] = [0.46 2.6 0] × 10−4
𝜀𝑧𝑥 𝜀𝑧𝑦 𝜀𝑧𝑧 0 0 0

Solution

𝐼1,𝜀 = 𝜀𝑥𝑥 + 𝜀𝑦𝑦 + 𝜀𝑧𝑧


= (−1 + 2.6 + 0) × 10−4
= 1.6 × 10−4
2 2 2
𝐼2,𝜀 = 𝜀𝑥𝑥 𝜀𝑦𝑦 + 𝜀𝑦𝑦 𝜀𝑧𝑧 + 𝜀𝑧𝑧 𝜀𝑥𝑥 − 𝜀𝑥𝑦 − 𝜀𝑦𝑧 − 𝜀𝑧𝑥
= (−1)(2.6) + 0 + 0 − (0.46)2 − 0 − 0
= −2.8116 × 10−4
2 2 2
𝐼3,𝜀 = 𝜀𝑥𝑥 𝜀𝑦𝑦 𝜀𝑧𝑧 + 2𝜀𝑥𝑦 𝜀𝑦𝑥 𝜀𝑧𝑥 − 𝜀𝑥𝑥 𝜀𝑦𝑧 − 𝜀𝑦𝑦 𝜀𝑧𝑥 − 𝜀𝑧𝑧 𝜀𝑥𝑦
=0

Substituting into the Eigen equation.

𝜀 3 − 1.6𝜀 2 − 2.8116𝜀 = 0
𝜀(𝜀 2 − 1.6𝜀 − 2.8116) = 0
1.6 ± √(−1.6)2 − 4(1)(−2.8116)
𝜀𝑝 = 0, and 𝜀𝑝 = × 10−4
2(1)

Ranking the principal strain solutions according to 𝜀1 > 𝜀2 > 𝜀3, we get:

𝜀1 = 2.658 × 10−4
𝜀2 = 0
𝜀3 = −1.058 × 10−4

Note that similar to principal stress, the principal strain planes have no shear strain.

9
MECH3361/9361 Semester 2, 2016

2.6. Compatibility of strains

Incompatible
Compatible

Fig. 2.7 Strain compatibility

Using the strain displacement relationships developed in Section 2.3, strain components are
easily determined by known displacements. Conversely, to determine the displacements from
the known strain components, a strain compatibility condition is required. This is because:

Mathematically, there will not be a unique solution to the displacement field (three
components: u,v,w) from six known strain components (εxx, εyy, εzz, εxy, εyz, εzx)
Physically, displacement fields should be continuous and cannot be arbitrary, as
illustrated in Fig. 2.7.

In a 2D case, we have the following strain-displacement relationships.

𝜕𝑢 𝜕𝑣 1 𝜕𝑣 𝜕𝑢
𝜀𝑥𝑥 = , 𝜀𝑦𝑦 = , 𝜀𝑥𝑦 = ( + )
𝜕𝑥 𝜕𝑦 2 𝜕𝑥 𝜕𝑦

To combine these relationships, differentiate twice with respect to x or y to obtain:

𝜕 2 𝜀𝑥𝑥 𝜕3𝑢
=
𝜕𝑦 2 𝜕𝑥𝜕𝑦 2
𝜕 2 𝜀𝑦𝑦 𝜕3𝑣
=
𝜕𝑥 2 𝜕𝑦𝜕𝑥 2
𝜕 2 𝜀𝑥𝑦 1 𝜕 3 𝑣 𝜕3𝑢
= ( + )
𝜕𝑥𝜕𝑦 2 𝜕𝑥 2 𝜕𝑦 𝜕𝑦 2 𝜕𝑥

𝜕2 𝜀𝑥𝑥 𝜕2 𝜀𝑦𝑦 𝜕2 𝜀𝑥𝑦


Substituting and into , a 2D strain compatibility equation can be derived.
𝜕𝑦 2 𝜕𝑥 2 𝜕𝑥𝜕𝑦

𝜕 2 𝜀𝑥𝑥 𝜕 2 𝜀𝑦𝑦 𝜕 2 𝜀𝑥𝑦


+ =2
𝜕𝑦 2 𝜕𝑥 2 𝜕𝑥𝜕𝑦
(2.13)

This compatibility condition applies to 2D deformation in the xy-plane, and guarantees a


continuous deformation field in the xy-plane.

10

You might also like