Theory of Elasticity Supplementary Notes: J. N. Reddy
Theory of Elasticity Supplementary Notes: J. N. Reddy
Supplementary Notes
J. N. Reddy
Distinguished Professor
Department of Mechanical Engineering
Texas A&M University, College Station
Texas, USA 77843–3123
1. Introduction
Consider a deformable body of known geometry, constitution, and loading.
For a given geometry and loading, the body will undergo deformation (i.e.,
macroscopic geometric changes within the body). If the applied loads are
time-dependent, the deformation of the body will be a function of time, i.e.,
the geometry of the body will change continuously with time. If the loads are
applied slowly so that the deformation is only dependent on the loads, the
body will take a definitive shape at the end of each load application. Whether
the deformation is time dependent or not, the forces in the deformed body
will be in equilibrium at all times.
Suppose that the body initially occupies a configuration C0 , in which a
particle X of the body occupies the position X, referred to a rectangular
Cartesian system (X1 , X2 , X3 ). Note that X is the name of the particle that
occupies the location X in configuration C0 , and therefore (X1 , X2 , X3 ) are
called the material coordinates. After the application of the loads, the body
moves and deforms and assumes a new configuration C. The particle X now
occupies the position x in the deformed configuration C (see Figure 1.1).
An analytical description of the deformation of a continuous body follows
one of the two approaches. In the first approach, called the material or
Lagrangian description, the motion of the body is referred to a reference
configuration CR , which is often chosen to be the undeformed configuration,
CR = C0 . Thus, in the Lagrangian description, the current coordinates
(x1 , x2 , x3 ) are expressed in terms of the reference coordinates (X1 , X2 , X3 ):
x = x(X, t) (1.1)
and the variation of a typical variable φ over the body is described with respect
to the material coordinates (X1 , X2 , X3 ):
φ = φ(X, t) (1.2)
In the spatial or Eulerian description, the motion is referred to the current
configuration C occupied by the body, and φ is described with respect to the
position (x1 , x2 , x3 ) in space, currently occupied by material particle X:
φ = φ(x, t), X = X(x, t) (1.3)
The coordinates (x1 , x2 , x3 )are termed the spatial coordinates.
Equations (1.2) and (1.3) each convey a different information. In Eq. (1.2),
a change in time t implies that the same material particle X, occupying
position X in C0 , has a different value φ. Thus the attention is focused on
the material particle X. In Eq. (1.3), a change in time t implies that a
different value φ is observed at the same spatial location x, now probably
occupied by a different material particle X. Hence, attention is focused on a
spatial position x.
Example 1.1:
To illustrate the difference between the two descriptions further, consider the 1-D
mapping x = X(1 + 0.5t) defining the motion of a rod of initial length two units. The
rod experiences a temperature distribution T given by the material description T = 2Xt2 or
by the spatial description T = xt2 /(1 + 0.5t) (see Figure 1.2).
It is clear from Figure 1.2 that the particle material coordinate (label) X remains
associated with the particle while its spatial position x changes. The temperature at a given
time can be found in one of the two ways: for example, at time t = 3 the temperature
of the particle labeled X = 2 is T = 2 × 2(3)2 = 36; alternatively, the temperature
of the same particle which at t = 3 is at a spatial position x = 2(1 + 0.5 × 3) = 5 is
T = 2 × 5(3)2 /(1 + 0.5 × 3) = 36.
In the study of solid bodies, the Eulerian description is less useful since
the configuration C is unknown. On the other hand, it is the preferred
description for the study of the motion of fluids because the configuration is
known and remains unchanged, and we wish to determine the changes in the
fluid velocities, pressure, density and so on. Thus, in the Eulerian description,
attention is focussed on a given region of space instead of on a given body of
matter. In much of the current study we focus our attention on the Lagrangian
description of the motion of solid bodies undergoing geometric changes.
2. Kinematics
2.1 Deformation Gradient Tensor
Consider two material particles P and Q in the neighborhood of each other in
the reference configuration C0 (see Figure 2.1). The positions of P and Q are
denoted by XP and XQ , respectively. The position of Q relative to P is given
by the elemental vector dX in C0 :
dX = XQ − XP
After deformation the material particles P and Q occupy spatial positions xP
and xQ , respectively in C; they are now labeled as P̄ and Q̄. The position of
Q̄ relative to P̄ is denoted by dx and it is given by
dx = xQ − xP
uP = xP − XP , uQ = xQ − XQ (2.1)
x=A·X+c
J. N. REDDY 5
Example 2.1:
Consider the uniform deformation of a square of side two units and initially centered at
X = (0, 0). The deformation is defined by the mapping
which produces the deformed shape shown in Figure 9.2.2. The components of the
deformation gradient tensor and its inverse can be expressed in matrix form as
´ ∂x1 ∂x1 µ ´ µ ´ ∂X1 ∂X1 µ ´ µ
1.0 0.5 1.0 −0.5
[F ] = ∂X1
∂x2
∂X2
∂x2 = ; [F ]−1 = ∂x1
∂X2
∂x2
∂X2 =
∂X1 ∂X2
0.0 1.0 ∂x1 ∂x2
0.0 1.0
Ý1 and e
The unit vectors e Ý2 in the current configuration deformed from the vectors
´ µº » º » ´ µº » º »
1.0 −0.5 1.0 1.0 1.0 −0.5 0.0 −0.5
= ; =
0.0 1.0 0.0 0.0 0.0 1.0 1.0 1.0
(dS)2 = dX · dX (2.6a)
(ds)2 = dx · dx = (F · dX) · (F · dX) = dX · (FT · F) · dX
≡ dX · C · X (2.6b)
C = FT · F (2.7)
The change in the squared lengths that occurs as the body deforms from
the initial to the current configuration can be expressed relative to the original
length as
(ds)2 − (dS)2 = 2 dX · E · dX (2.8)
1° T ± 1
E= F · F − I = (C − I) (2.9)
2 2
1h i
= (I + ∇o u)T · (I + ∇o u) − I
2
1h i
= (∇o u)T + ∇o u + (∇o u)T · (∇o u) (2.10)
2
Clearly, the Green strain tensor is symmetric. Also, the change in the squared
lengths is zero if and only if E = 0.
Alternatively, the change in the squared lengths that occurs as the body
deforms from the initial to the current configuration can be expressed relative
to the current length as
1° ± 1
e= I − F−T · F−1 = (I − c) (2.12)
2 2
1h T
i
= I − (I − ∇u) · (I − ∇u)
2
1h i
= (∇u)T + ∇u − (∇u)T · (∇u) (2.13)
2
where c is the Cauchy strain tensor, and its inverse is called the left Cauchy-
Green or Finger tensor.
In the Cartesian component form, we can write
with components
In expanded notation, the Green strain components, for example, are given
by
"² ³2 ³2 ³2 #
∂u1 1 ∂u1 ∂u2 ∂u3
² ²
E11 = + + +
∂X1 2 ∂X1 ∂X1 ∂X1
"² ³2 ³2 ³2 #
∂u2 1 ∂u1 ∂u2 ∂u3
² ²
E22 = + + +
∂X2 2 ∂X2 ∂X2 ∂X2
"² ³ #
∂u3 1 ∂u1 2 ∂u2 2 ∂u3 2
³ ² ³ ²
E33 = + + +
∂X3 2 ∂X3 ∂X3 ∂X3
1 ∂u1 ∂u2 ∂u1 ∂u1 ∂u2 ∂u2 ∂u3 ∂u3
² ³
E12 = + + + +
2 ∂X2 ∂X1 ∂X1 ∂X2 ∂X1 ∂X2 ∂X1 ∂X2
1 ∂u1 ∂u3 ∂u1 ∂u1 ∂u2 ∂u2 ∂u3 ∂u3
² ³
E13 = + + + +
2 ∂X3 ∂X1 ∂X1 ∂X3 ∂X1 ∂X3 ∂X1 ∂X3
1 ∂u2 ∂u3 ∂u1 ∂u1 ∂u2 ∂u2 ∂u3 ∂u3
² ³
E23 = + + + + (2.18)
2 ∂X3 ∂X2 ∂X2 ∂X3 ∂X2 ∂X3 ∂X2 ∂X3
8 THEORY OF ELASTICITY
Example 2.2:
Consider a rectangular block of dimensions a × b × h, where h is very small compared to
a and b. Suppose that the block is deformed into the diamond shape shown in Figure
2.3. By inspection, the geometry of the deformed body can be described as follows: let
(X1 , X2 , X3 ) denote the coordinates of a material point in the undeformed configuration.
Thus the coordinate mapping and its inverse are given by
e ab ae0
x1 = X1 + 0 X2 , X1 = x1 − x2
b ab − e20 ab − e20
e0 be0 ab
x2 = X2 + X , X2 = − x1 + x2
a 1 ab − e20 ab − e20
x3 = X3 , X3 = x3 .
Thus, the displacement components of a material point in the Lagrangian description are
e e
u1 = x1 − X1 = 0 X2 , u2 = x2 − X2 = 0 X1 , u3 = x3 − X3 = 0.
b a
and the spatial description they are
e20 ae0
u1 = x1 − X1 = − x1 + x2
ab − e20 ab − e20
be0 e20
u2 = x2 − X2 = 2
x1 − x2
ab − e0 ab − e20
u3 = x3 − X3 = 0.
The only nonzero Green strain tensor components are given by
1 e0 2
±2
e e 1 e0
° ± °
E11 = , E12 = 0 + 0 , E22 = .
2 a 2b 2a 2 b
The Almansi strain tensor components are
e20 e20 (e20 + b2 )
´ µ
1
e11 = − − ,
ab − e20 2 (ab − e20 )2
e0 (a + b) e3 (a + b)
e12 = 2
+ 0 ,
(ab − e0 ) (ab − e20 )2
e20 e20 (e20 + a2 )
´ µ
1
e22 = − − .
ab − e20 2 (ab − e20 )2
The same results can be obtained using the elementary mechanics of materials approach,
where the strains are defined to be the ratio of the difference between the final length and
original length to the original length. For example, a line element AB in the undeformed
body moves to position ĀB̄. Then the strain in the line AB is given by
r ±2
ĀB̄ − AB 1 e0
q °
E11 = εAB = = a2 + e20 − 1 = 1+ −1
AB a a
´ ±2 µ ±2
1 e0 1 e0
° °
= 1+ + ··· − 1 ≈ .
2 a 2 a
Similarly, ´ µ
±2 ±2
1 e0 1 e0
° °
E22 = 1+ + ··· − 1 ≈ .
2 b 2 b
Example 2.3:
For the deformation given in Example 2.1 the right Cauchy–Green deformation tensor and
the Cauchy strain tensors are respectively
´ µ´ µ ´ µ
1.0 0.0 1.0 0.5 1.0 0.5
[C] = [F ]T [F ] = =
0.5 1.0 0.0 1.0 0.5 1.25
´ µ´ µ ´ µ
−T −1 1.0 0.0 1.0 −0.5 1.0 −0.5
[c] = [F ] [F ] = =
−0.5 1.0 0.0 1.0 −0.5 1.25
The Green and Almansi strain tensor components in matrix form are given by
´ µ ´ µ
1 0.0 0.5 1 0.0 0.5
[E] = ; [e] =
2 0.5 0.25 2 0.5 −0.25
Example 2.4:
Consider the uniform deformation of a square of side 2 units initially centered at X = (0, 0).
The deformation is given by the mapping
1 1
x1 = (18 + 4X1 + 6X2 ), x2 = (14 + 6X2 )
4 4
We wish to
(a) sketch the deformed configuration of the body,
(b) compute the components of the deformation gradient tensor F and its inverse (display
them in matrix form),
(c) compute the components of the right and left Cauchy-Green deformation tensors (display
them in matrix form), and
(d) compute Green’s and Almansi’s strain tensor components (display them in matrix form).
Solution:
(a) Sketches of the undeformed and deformed bodies are shown in Figure 2.4.
10 THEORY OF ELASTICITY
The matrix form of the deformation gradient tensor and its inverse are
´ ∂x1 ∂x1 µ ´ µ ´ ∂X1 ∂X1 µ ´ µ
∂X1 ∂X2 1 2 3 −1 ∂x1 ∂x2 1 3 −3
[F ] = ∂x2 ∂x2 = ; [F ] = ∂X2 ∂X2 =
∂X1 ∂X2
2 0 3 ∂x1 ∂x2
3 0 2
(c) The right and left Cauchy–Green deformation tensors are, respectively,
´ µ ´ µ
1 2 3 1 13 9
[C] = [F ]T [F ] = , [c] = [F ][F ]T =
2 3 9 4 9 9
(d) The Green and Almansi strain tensor components in matrix form are respectively
´ µ
1 T ¡ 1 0 3
[E] = [F ] [F ] − [I] =
2 2 3 7
´ µ
1 1 0 9
[I] − [F ]−T [F ]−1 =
¡
[e] =
2 18 9 −4
J. N. REDDY 11
1h i
ε= (∇u)T + ∇u (2.19)
2
In the case of infintesimal strains, no distinction is made between the material
coordinates X and the spatial coordinates x. Therefore, the linear Eulerian
strains and linear Lagrangian strains become the same. The linear strain-
displacement relations in the Cartesian system are given by
1
εij = (ui,j + uj,i ) (2.20a)
2
or, in expanded form,
The strains defined by Eqs. (2.10), (2.13) and (2.19) are valid in any
coordinate system. Hence, they can be expressed in component form in any
given coordinate system by exanding the strain tensors in the dyadic form
and the operator ∇ in that coordinate system. For example, in the cylindrical
coordinate system we have
∂êr ∂êθ
= êθ , = −êr . (2.23)
∂θ ∂θ
Using u and ∇ from Eqs. (2.21) and (2.22), we obtain
∂ 1 ∂ ∂
² ³
∇u = êr + êθ + êz (ur êr + uθ êθ + uz êz )
∂r r ∂θ ∂z
12 THEORY OF ELASTICITY
−ω12 −ω13
0
1
[ω] = ω12 0 −ω23 (2.28)
2
ω13 ω23 0
1
ωij = −eijk θk , or θk = − ekij ωij (2.29)
2
It follows that
1 1
θk = ekij uj,i or θ = curl u = ∇ × u (2.30)
2 2
Certain motions do not produce infinitesimal strains but they may produce
finite strains. For example, the following displacement field
u1 = b1 + c2 X3 − c3 X2
u2 = b2 + c3 X1 − c1 X3
u3 = b3 + c1 X2 − c2 X1
for arbitrary constants (bi , ci ) produces no infinitesimal strains, but the fnite
strain components due to the displacement field are not zero. The deformation
gradient tensor components associated with the dispacement field are
1 −c3
c2
[F ] = c3 1 −c1
−c2 c1 1
Thus, [F ] = [I] if and only if all ci = 0, and for any sufficiently large ci 6= 0,
the mapping does not represent a rigid body motion. Figure 2.5 depicts the
deformation for the two-dimensional case, with b1 = 2, b2 = 3, and c3 = 1.
Clearly, the motion is not a rigid body motion, although the infinitesimal
strains produced by the motion are zero. The finite strains and deformation
gradient tensor give true measures of the deformation.
14 THEORY OF ELASTICITY
Figure 2.5: A mapping that produces zero infinitesimal strains but nonzero
finite strains.
F=R·U (2.31)
C = FT · F = UT · RT · R · U = UT · U (2.32)
3
X
C= λ2α N̂α N̂α (2.33)
α=1
J. N. REDDY 15
where λ2α are the eigenvalues and N̂α are the eigenvectors of C. Then it follows
that
3
X
U= λα N̂α N̂α (2.34)
α=1
Once the stretch tensor is known, the rotation tensor R can be obatined from
Eq. (2.31) as R = F · U−1 .
Example 2.5:
Consider the deformation given by the mapping
1 1
x1 = [4X1 + (9 − 3X1 − 5X2 − X1 X2 ) t] , x2 = [4X2 + (16 + 8X1 ) t]
4 4
(a) For X = (0, 0) and t = 1 determine the deformation gradient tensor F and right Cauchy-
Green tensor C.
(b) Find the eigenvalues (stretches) λ1 and λ2 and the associated eigenvectors N1 and N2 .
(c) Use the polar decomposition to determine the symmetric stretch tensor U and rotation
tensor R.
Solution:
(a) For X = (0, 0) and time t = 1 the components of the deformation gradient tensor F and
right Cauchy–Green tensor C are
´ µ ´ µ
1 1 −5 1 65 27
[F ] = , [C] =
4 8 4 16 27 41
(b) The eigenvalues λ21 and λ22 of matrix [C] are determined by setting
so that λ1 = 2.2714 and λ2 = 1.2107. The eigenvectors are (in vector component form)
º » º »
0.8385 −0.5449
{N (1) } = , {N (2) } =
0.5449 0.8385
Ý (1) N
U = λ1 N Ý (1) + λ2 N
Ý (2) N
Ý (2)
° ±° ±
(1)
= λ1 N1 e Ý1 + N2(1) e
Ý2 (1)
N1 e Ý1 + N2(1) e
Ý2
° ±° ±
(2)
+ λ2 N1 e Ý1 + N2(2) e
Ý2 (2)
N1 e Ý1 + N2(2) e
Ý2
° ± ° ±
(1) (2)
= λ1 [N1 ]2 + λ2 [N1 ]2 e Ý1 + λ1 [N2(1) ]2 + λ2 [N2(2) ]2 e
Ý1 e Ý2 e
Ý2
° ±
+ λ1 N1(1) N2(1) + λ2 N1(2) N2(2) (Ý Ý2 + e
e1 e Ý2 e
Ý1 )
16 THEORY OF ELASTICITY
or in matrix form ´ µ
1.9564 0.4846
[U ] =
0.4846 1.5257
3. Compatibility Equations
The task of computing strains (infinitesimal or finite) from a given
displacement field is a straightforward exercise. However, the determination
of displacements from a given strain field is not as straightforward due to
the fact there are six independent partial differential equations (i.e., strain-
displacement relations) for only three unknown displacements, which would
in general overdetermine the solution. We will find some conditions, known
as St. Venant’s compatibility equations. The derivation is presented for
infinitesimal strains. For finite strains the same steps may be followed but
the process is so difficult that it is never attempted (although some general
compatibility conditions may be stated to insure integrability of the six
nonlinear partial differential equations).
To understand the meaning of strain compatibility, imagine that a material
body is cut up into pieces before it is strained, and then each piece is
given a certain strain. The strained pieces cannot be fitted back into a
single continuous body without further deformation. On the other hand if
the strain in each piece is related to or compatible with the strains in the
neighboring pieces, then they can be fitted together to form a continuous
body. Mathematically, the six relations that connect six strain components to
the three displacement components should be consistent. To make this point
clear, consider the two-dimensional case. We have three strain-displacement
relations in two displacements:
∂u1
= ε11 (3.1)
∂x1
∂u2
= ε22 (3.2)
∂x2
∂u1 ∂u2
+ = 2ε12 (3.3)
∂x2 ∂x1
If the given data (ε11 , ε22 , ε12 ) is compatible (or consistent), any two of
the three equations should yield the same displacement components. The
compatibility of the data can be established as follows. Differentiate the first
J. N. REDDY 17
∂ 3 u1 ∂ 2 ε11
= (3.10 )
∂x1 ∂x22 ∂x22
∂ 3 u2 ∂ 2 ε22
2 = (3.20 )
∂x2 ∂x1 ∂x21
∂ 3 u1 ∂ 3 u2 ∂ 2 ε12
+ = 2 (3.30 )
∂x22 ∂x1 ∂x21 ∂x2 ∂x1 ∂x2
Using Eqs. (3.1’) and (3.2’) in (3.3’), we arrive at the following equation
between the three strains:
Equation (2.35) is called the strain compatibility condition among the three
strains for a 2D case.
Similar procedure can be followed to obtain the strain compatibility
conditions for the 3D case. In addition to Eq. (3.4) five more such conditions
can be derived:
∂ 2 ε11 ∂ 2 ε33 ∂ 2 ε13
+ = 2 (3.5)
∂x23 ∂x21 ∂x1 ∂x3
1 ∂ 2 uj ∂ 1 ∂uj ∂θk
² ³
∇ × ε = eijk êk êr = eijk êk êr = êk êr
2 ∂xi ∂xr ∂xr 2 ∂xi ∂xr
or
∂θk ∂
(∇ × ε)T = êr êk = êr (θk êk ) = ∇θ
∂xr ∂xr
Since curl of a the gradient of a vector (or tensor) is zero, we take curl of the
above equation and arrive at the compatibility equation in vector/tensor form:
Example 3.1:
Consider the problem of an isotropic cantilever beam bent by a load F0 at the free end (see
Figure 3.1). From the elementary beam theory, we have the following infinitesimal strains
F0 x1 x2 F0 x1 x2 (1 + ν)F0 2
ε11 = − , ε22 = −νε11 = ν , ε12 = − (h − x22 ), (3.13)
EI3 EI3 2EI3
where ν is the Poisson ratio, EI3 is the bending stiffness (E is Young’s modulus and I3 is
the second moment of area about the x3 −axis), and 2h is the height of the beam. We wish
to determine the two-dimensional displacement field (u1 , u2 ) in the beam.
First, we determine if the given strains are compatible. Substituting εij into the single
compatibility equation for 2D, we obtain 0 + 0 = 0. Thus the strains satisfy the compatibility
equation in two dimensions. Although the two-dimensional strains are compatible, the
three-dimensional strains are not compatible. For example, using the additional strains,
ε33 = −νε11 , ε13 = ε23 = 0, one can show that all of the equations except the one in Eq.
(3.9) are satisfied.
Since the strains are compatible in two-dimensional theory of elasticity, we wish to
compute the displacement field u = e Ý2 u2 of the beam using the known strains.
Ý1 u1 + e
Integrating the strains in Eq. (3.13), we obtain
∂u1 F x x F0 x21 x2
ε11 = =− 0 1 2 or u1 = + f1 (x2 ), (3.14a)
∂x1 EI3 2EI3
∂u2 νF0 x1 x2 νF0 x1 x22
ε22 = = or u2 = + f2 (x1 ), (3.14b)
∂x2 EI3 2EI3
where (f1 , f2 ) are functions of integration. Substituting u1 and u2 into the definition of the
shear strain 2ε12 , we obtain
But this must be equal to the strain value given in Eq. (3.13):
Since the left side depends only on x1 and the right side depends only on x2 , and yet the
equality must hold, it follows that both sides should be equal to a constant, say c0 :
(2 + ν)F0 3
f1 (x2 ) = x2 + c0 x2 + c1
6EI3
F0 3 (1 + ν)F0 h2
f2 (x1 ) = x − x1 − c0 x1 + c2 ,
6EI3 1 EI3
where c1 and c2 are constants of integration that are to be determined. The displacements
(u1 , u2 ) are now given by
F0 2 (2 + ν)F0 3
u1 (x1 , x2 ) = − x x + x2 + c0 x2 + c1 , (3.16a)
2EI3 1 2 6EI3
(1 + ν)F0 h2 νF0 F0 3
u2 (x1 , x2 ) = − x1 + x x2 + x − c0 x1 + c2 . (3.16b)
EI3 2EI3 1 2 6EI3 1
20 THEORY OF ELASTICITY
The constants c0 , c1 , and c2 can be evaluated using the boundary conditions on the
displacements. We impose the following boundary conditions:
¬
∂u1 ¬
° ±¬
u1 (L, 0) = 0, u2 (L, 0) = 0, = 0. (3.17)
∂x2 ¬
¬
(L,0)
Substituting the expressions for u1 and u2 into the boundary conditions (3.17), we obtain
u1 (L, 0) = 0 → c1 = 0
F0 L3 (1 + ν)F0 h2 L
u2 (L, 0) = 0 → − c0 L + c2 = ,
6EI3 EI3
¬
∂u1 ¬ F0 L2 F0 L2
° ±¬
=0 → − +c=0 or c= .
∂x2 ¬ 2EI3 2EI3
¬
(L,0)
F0 (2 + ν)F0 3
L2 − x21 x2 +
¡
u1 (x1 , x2 ) = x2 , (3.18a)
2EI3 6EI3
F0 ¡ F (1 + ν)h2
x1 − 3L2 x1 + 2L3 + 0
3
u2 (x1 , x2 ) = (L − x1 )
6EI3 EI3
νF0
+ x x2 . (3.18b)
2EI3 1 2
The values of the constants c, c1 , and c2 depend on the boundary conditions used. For
example, if one uses (∂u2 /∂x1 ) = 0 in place of (∂u1 /∂x2 ) = 0 at x1 = L and x2 = 0, the
resulting displacements would be
F0 (2 + ν)F0 2
L2 − x21 x2 − x2 3h − x22 ,
¡ ¡
u1 (x1 , x2 ) = (3.19a)
2EI3 6EI3
F0 3 ¡ νF0
u2 (x1 , x2 ) = x1 − 3L2 x1 + 2L3 + x x2 . (3.19b)
6EI3 2EI3 1 2
The displacement field in the beam according to the Euler–Bernoulli beam theory is
given by
F0
L2 − x21 x2 ,
¡
u1 (x1 , x2 ) = (3.20a)
2EI3
F0
x31 − 3L2 x1 + 2L3 ,
¡
u2 (x1 , x2 ) = (3.20b)
6EI3
F0
L2 − x21 x2 ,
¡
u1 (x1 , x2 ) = (3.21a)
2EI3
F0 F0
x31 − 3L2 x1 + 2L3 +
¡
u2 (x1 , x2 ) = (L − x1 ), (3.21b)
6EI3 Ks GA
where Ks is the shear correction factor, and A is the area of cross section. The transverse
deflection u2 of the Timoshenko beam theory has an extra term due to the inclusion of the
J. N. REDDY 21
transverse shear force/strain effect. For a rectangular beam of height 2h and width b, we
have A = 2bh and I3 = 2bh3 /3.
A comparison of the elasticity solution (3.19a,b) against the Timoshenko beam solution
(3.21a,b) indicates that the elasticity solution has some higher-order terms that are negligible
(x2 ∼ h). Also one can see that (GA = 3EI3 /2[(1 + ν)h2 ]) the shear correction factor is
Ks = 2/3.
4. Measures of Stress
4.1 Cauchy Stress Tensor and Cauchy’s Formula
The equations of motion or equilibrium of a material body must be derived
for the deformed configuration of the body at time t. However, since the
geometry of the deformed configuration is not known, the equations must
be written in terms of the known reference configuration. In doing so we
introduce various measures of stress. They emerge in a natural way as we
transform volumes and areas from the deformed configuration to undeformed
(or reference) configuration.
First we introduce the true stress, i.e., stress in the deformed configuration.
Consider a deformed body at its current position. The surface force acting
on a small element of area in a continuous medium depends not only on the
magnitude of the area but also upon the orientation of the area. It is customary
to denote the direction of a plane area by means of a unit vector drawn normal
to that plane. To fix the direction of the normal, we assign a sense of travel
along the contour of the boundary of the plane area in question. The direction
of the normal is taken by convention as that in which a right-handed screw
advances as it is rotated according to the sense of travel along the boundary
curve or contour. Let the unit normal vector be given by n Ý . Then the area
can be denoted by s = sÝ n.
If we denote by df(n̂) the force on a small area n̂da located at the position
x, the stress vector can be defined as (see Figure 4.1)
∆f(n̂)
t(n̂) = lim . (4.1)
∆a→0 ∆a
We see that the stress vector is a point function of the unit normal n̂ which
denotes the orientation of the surface ∆a. The component of t that is in the
direction of n̂ is called the normal stress. The component of t that is normal
to n̂ is called a shear stress. Because of Newton’s third law for action and
reaction, we see that t(−n̂) = −t(n̂).
At a fixed point r for each given unit vector n̂ there is a stress vector t(n̂)
acting on the plane normal to n̂. Note that t(n̂) is, in general, not in the
direction of n̂. It is fruitful to establish a relationship between t and n̂.
22 THEORY OF ELASTICITY
It follows that
∆a1 = (n̂ · ê1 )∆a, ∆a2 = (n̂ · ê2 )∆a, ∆a3 = (n̂ · ê3 )∆a. (4.4)
∆h
∆v = ∆a, (4.5)
3
where ∆h is the perpendicular distance from the origin to the slant face.
Substitution of Eqs. (4.3) and (4.4) in (4.2) and dividing throughout by
∆S reduces it to
∆h
t = (n̂ · ê1 )t1 + (n̂ · ê2 )t2 + (n̂ · ê3 )t3 + ρ (a − f). (4.6)
3
In the limit when the tetrahedron shrinks to a point, ∆h → 0, we are left with
The terms in the parenthesis are to be treated as a dyadic, called stress dyadic
↔
or stress tensor σ :
↔
σ ≡ ê1 t1 + ê2 t2 + ê3 t3 . (4.9)
The stress tensor is a property of the medium that is independent of the n̂.
Thus, we have
↔ ↔T
t(n̂) = n̂ · σ = σ · n̂ (4.10)
and the dependence of t on n̂ has been explicitly displayed. In Cartesian
component form, we have ti = nj σji .
It is useful to resolve the stress vectors t1 , t2 , and t3 into their orthogonal
components. We have
The component σij represents the stress (force per unit area) on an area
perpendicular to the ith coordinate and in the jth coordinate direction (see
Figure 4.3). The stress vector t represents the vectorial stress on an area
perpendicular to the direction n̂. Equation (4.10) is known as the Cauchy
↔
stress formula, and σ is termed the Cauchy stress tensor. Thus, the Cauchy
stress tensor σ is defined to be the current force per unit deformed area
df = tda = σ · da, where Cauchy’s formula, t = n̂ · σ = σ T · n̂, is used.
The matrix form of the Cauchy’s formula (for computing purposes) is given
by
t1 σ11 σ21 σ31 n1
t = σ12 σ22 σ32 n2 (4.13)
2
t3 σ13 σ23 σ33 n3
Example 4.1:
With reference to a rectangular Cartesian system (x1 , x2 , x3 ), the components of the stress
dyadic at a certain point of a continuous medium are given by
"
200 400 300
#
[σ] = 400 0 0 psi.
300 0 −100
We wish to determine the stress vector t at the point and normal to the plane, x1 + 2x2 +
2x3 − 6 = 0, and then compute the normal and tangential components of the stress vector
at the point.
First we should find the unit normal to the plane on which we are required to find the
stress vector. The unit normal to a surface defined by the equation φ(x1 , x2 , x3 ) =constant
is given by
∇φ
Ý=
n , φ(x1 , x2 , x3 ) = x1 + 2x2 + 2x3 − 6,
|∇φ|
1
Ý = (Ý
n e + 2Ý e2 + 2Ý
e3 ).
3 1
The components of the stress vector are
(
t1
) "
200 400 300
# ( )
1
(
1600
)
1 1
t2 = 400 0 0 2 = 400 psi,
3 3
t3 300 0 −100 2 100
or
1
t(Ý
n) = (1600Ýe1 + 400Ý
e2 + 100Ýe3 ) psi.
3
The normal component tnn of the stress vector t on the plane is given by
2600
tnn = t(Ý Ý=
n) · n psi,
9
and the tangential component is given by (the Pythagorean theorem)
p 102 p
tns = |t|2 − t2nn = (256 + 16 + 1)9 − 26 × 26 psi
9
√
1781
= 100 = 468.9 psi.
9
df = t da = n̂ · σ da = da · σ where da = da n̂ (4.16)
= dA · P, where dA = dA N̂ (4.17)
where N̂ is the unit normal to the undeformed area dA. The stress tensor P is
called the first Piola–Kirchhoff stress tensor, and it gives the current force
per unit undeformed area. The tensor Cartesian component representation
of P is given by
P = PIi ÊI êi (4.18)
Thus the first Piola–Kirchhoff stress tensor is a mixed tensor and it is
unsymmetric.
26 THEORY OF ELASTICITY
da · σ = dA · P
Thus, the second Piola–Kirchhoff stress tensor gives the transformed current
force per unit undeformed area. Cartesian component representation of S is
P = J F−1 · σ = S · FT (4.23)
1 1
σ = F · P = F · S · FT (4.24)
J J
S = J F−T · σ · F−1 = P · F−T (4.25)
J. N. REDDY 27
Dφ ∂φ ∂φ
² ³
≡ = (5.1)
Dt ∂t X=const ∂t
In particular we have
∂x ∂x
² ³ ² ³
= ≡v (5.2)
∂t X=const ∂t X=const
∂v ∂v
² ³ ² ³
= ≡a (5.3)
∂t X=const ∂t X=const
Dφ ∂φ ∂φ ∂xi ∂φ
² ³ ² ³ ² ³
≡ = +
Dt ∂t X=const ∂t x=const ∂t X=const ∂xi
∂φ ∂φ
² ³
= + vi
∂t x=const ∂xi
∂φ
² ³
= + v · ∇φ
∂t x=const
28 THEORY OF ELASTICITY
where Eq. (5.2) is used in the second line. Thus, the material derivative
operator is given by
D ∂
² ³
= +v·∇ (5.4)
Dt ∂t x=const
Example 5.1:
This example illustrates the calculation of material time derivative based on material spatial
descriptions. Suppose that the motion is described by the mapping, x = (1 + t)X. Then the
velocity in the material and spatial descriptions respectively are
∂x x
v(X) = =X or v(x) = (a)
∂t 1+t
Now consider a function φ(X, t) = Xt2 in material description. The material time
derivative of φ in the material description is simply
∂φ
= 2Xt
∂t
Dφ ∂φ ∂φ
= +v
Dt ∂t
² ∂x ³
±² 2 ³
2xt + xt2 x t
°
= +
(1 + t)2 1+t 1+t
2xt
= (b)
1+t
which is the same as that in Eq. (a), except that it is expressed in terms of the current
coordinate, x.
D Dφ
Z Z
ρφ dv = ρ dv (5.5)
Dt v v Dt
C at time t > 0, and it has a density ρ and vlume v. As per the principle of
conservation of mass, we have
Z Z
ρ0 dV = ρdv (5.6)
V v
This is the global form of the continuity equation. Since the material volume
V we selected is arbitrarily small, as we shrink the volume to a point, we obtain
the local form of the continuity equation
ρ0 = Jρ (5.8)
where the divergence theorem is used in arriving at the last line. Equating
Eqs. (5.9) and (5.10), we obtain
∂ρ
Z Z
dv = − ∇ · (ρv)dv
v ∂t v
30 THEORY OF ELASTICITY
or
∂ρ
Z ´ µ
+ ∇ · (ρv) dv = 0 (5.11)
v ∂t
which is the global form of the continuity equation in spatial description. The
local form is given by
∂ρ
+ ∇ · (ρv) = 0 (5.12)
∂t
Alternate form of Eq. (5.12) is obtained as follows:
∂ρ
0= + ∇ · (ρv)
∂t
∂ρ
= + v · ∇ρ + ρ∇ · v
∂t
Dρ
= + ρ∇ · v (5.13)
Dt
where the definition of material time derivative, Eq. (5.4), is used in arriving
at the last line.
d
(mv) = F, (6.1)
dt
where m is the total mass, v the velocity, and F the resultant force on the
collection particles. For constant mass, Eq. (6.1) becomes
dv
F=m = ma, (6.2)
dt
which is the familiar form of Newton’s second law. To derive the equation of
motion applied to a fixed region in space through which material flows, we
must identify the forces acting on it.
Forces acting on a volume element can be classified as internal and
external. The internal forces resist the tendency of one part of the region/body
to be separated from another part. The internal force per unit area is termed
stress, as defined in Section 4. The external forces are those transmitted by
J. N. REDDY 31
the body. The external forces can be further classified as body (or volume)
forces and surface forces.
Body forces act on the distribution of mass inside the body. Examples
of body forces are provided by the gravitational and electromagnetic forces.
Body forces are usually measured per unit mass or unit volume of the body.
Let b denote the body force per unit mass. Consider an elemental volume dv
inside v. The body force of the elemental volume is equal to ρ dv b. Hence,
the total body force of the control volume is
Z
ρ b dv. (6.3)
v
Surface forces are contact forces acting on the boundary surface of the
body. Examples of surface forces are provided by applied forces on the surface
of the body. Surface forces are reckoned per unit area. Let t denote the surface
force per unit area (or surface stress vector). The surface force on an elemental
surface ds of the volume is tds. The total surface force acting on the closed
surface of the volume v is I
t ds. (6.4)
s
Since the stress vector t on the surface is related to the (internal) stress tensor
↔
σ by Cauchy’s formula [see Eq. (4.10)]
↔
t = n̂ · σ , (6.4)
where n̂ denotes the unit normal to the surface, we can express the surface
force as I
↔
n̂ · σ ds.
s
which is the global form of the equation of motion. The local form is given by
↔ Dv
∇ · σ + ρb = ρ (6.8)
Dt
In Cartesian rectangular system, we have
∂σji ∂vi
+ ρbi = ρ (i = 1, 2, 3). (6.9)
∂xj ∂t
↔ ∂σji
∇ · σ + ρb = 0 or + ρbi = 0 (6.10)
∂xj
∂σ11 ∂σ21
² ³ ² ³
σ11 + dx1 dx2 dx3 − σ11 dx2 dx3 + σ21 + dx2 dx1 dx3
∂x1 ∂x2
∂σ31
² ³
− σ21 dx1 dx3 + σ31 + dx3 dx1 dx2 − σ31 dx1 dx2 + ρf1 dx1 dx2 dx3
∂x3
∂σ11 ∂σ21 ∂σ31
² ³
= + + − ρb1 dx1 dx2 dx3 .
∂x1 ∂x2 ∂x3
By Newton’s second law of motion, the sum of the forces is equal to the product
of mass and acceleration in the x1 -direction
∂v1
(ρdx1 dx2 dx3 ) ,
∂t
where ρ is the density. Thus, upon dividing throughout by dx1 dx2 dx3 , we
obtain
∂σ11 ∂σ21 ∂σ31 ∂v1
+ + + ρb1 = ρ . (6.11)
∂x1 ∂x2 ∂x3 ∂t
J. N. REDDY 33
∂σji ∂vi
+ ρbi = ρ (i = 1, 2, 3), (6.12)
∂xj ∂t
which is the same as that in Eq. (6.11). For static equilibrium, we set the
time derivative terms to zero and obtain the equilibrium equations
∂σji
+ ρbi = 0 (i = 1, 2, 3). (6.13)
∂xj
D
I Z Z
(x × t)ds + (x × ρb)dv = (x × ρv)dv (6.14)
s v Dt v
D
I Z Z
eijk xi tj ds + ρeijk xi bj dv = ρeijk xi vj dv (6.15)
s v Dt v
We use several steps to simply the expression. First replace tj with tj = np σpj .
Then transform the surface integral to a volume integral and use the Reynold’s
transport theorem for the material time derivative of a volume integral to
obtain
D
Z Z Z
eijk (xi σpj ),p dv + ρeijk xi bj dv = ρeijk (xi vj ) dv
v v v Dt
Carrying out the indicated differentiations and noting Dxi /Dt = vi , we obtain
Dvj
Z Z ² ³
eijk (xi σpj,p + δip σpj + ρxi bj ) dv = ρeijk vi vj + xi dv
v v Dt
Dvj
Z ´ ² ³ µ
eijk xi σpj,p + ρbj − ρ + σij dv = 0
v Dt
or
eijk σij = 0 (6.16)
which implies that σij = σji . That is, the matrix of the stress components is
symmetric:
σ23 = σ32 , σ31 = σ13 , σ12 = σ21 .
Thus, there are only six stress components that are independent.
The symmetry of the stress tensor can also be established using Newton’s
second law for moments. Consider the moment of all forces acting on the
parallelepiped about the x3 −axis (see Figure 6.1). Using the right-handed
screw rule for positive moment, we obtain
1
Dividing throughout by 2 dx1 dx2 dx3 and taking the limit dx1 → 0 and
dx2 → 0, we obtain
σ12 − σ21 = 0.
Similar considerations of moments about the x1 -axis and x2 -axis give,
respectively, the relations
Example 6.1:
Given the following state of stress (σij = σji ),
determine the body force components for which the stress field describes a state of static
equilibrium.
Thus the body is in static equilibrium for the body force components ρb1 = 0, ρb2 = 0, and
ρb3 = −4.
Example 6.2:
We wish to determine the principal stresses and principal strains associated with the stress
tensor "
2 1 0
#
[σ] = 1 4 1 psi.
0 1 2
The characteristic equation is obtained by setting det (σij − λ δij ) to zero:
¬ ¬
¬2 − λ 1 0 ¬
¬ 1 4−λ 1 ¬ = (2 − λ)[(4 − λ)(2 − λ) − 1] − 1 · (2 − λ) = 0,
¬ ¬
¬ 0 1 2 − λ¬
or
(2 − λ)[(4 − λ)(2 − λ) − 2] = 0.
Hence, the eigenevalues (or principal stresses) are
√ √
λ1 = 3 + 3 = 4.7321 psi, λ2 = 3 − 3 = 1.2679 psi, λ3 = 2 psi.
36 THEORY OF ELASTICITY
This gives
A2 = 0, A1 = −A3 .
Using A21 + A22 + A23 = 1, we obtain
Ý 3 = ± √1 (1, 0, −1),
A for λ3 = 2.
2
√
Similarly, the eigenvectors corresponding to λ1,2 = 3 ± 3 are given by
√
√ ¡ ¡ √
Ý 1 = ± (3 − 3) 1, 1 + 3 , 1 ,
A for λ1 = 3 + 3,
12
√
(3 + 3) √ ¡ ¡ √
Ý
A2 = ± 1, 1 − 3 , 1 , for λ2 = 3 − 3.
12
7. Constitutive Equations
7.1 General Introduction
The kinematic relations and the mechanical (e.g., conservation of mass and
momenta) and thermodynamic principles (e.g., conservation of energy) are
applicable to any continuum irrespective of its physical constitution. The
kinematic variables such as the strains and the kinetic variables such as the
stresses were introduced independent of each other. In this section we shall
consider equations characterizing the individual material and its reaction to
applied loads, and these equations relate the kinetic variables to the kinematic
variables. They are called the constitutive equations. The formulation of the
constitutive equations for a given material is guided by certain rules (i.e.,
constitutive axioms). We will not discuss them here but will review the linear
constitutive relations for solids.
Materials for which the constitutive behavior is only a function of the
current state of deformation are known as elastic. In the special case in which
the work done by the stresses during a deformation is dependent only on the
initial state and the current configuration, the material is called hyperelastic.
A material body is said to be homogeneous if the material properties
are the same throughout the body (i.e., independent of position). In a
heterogeneous body, the material properties are a function of position. An
anisotropic body is one that has different values of a material property in
different directions at a point, i.e., material properties are direction-dependent.
J. N. REDDY 37
An isotropic body is one for which every material property in all directions at a
point is the same. An isotropic or anisotropic material can be nonhomogeneous
or homogeneous.
A material body is said to be ideally elastic when, under isothermal
conditions, the body recovers its original form completely upon removal of
the forces causing deformation, and there is a one-to-one relationship between
the state of stress and the state of strain in the current configuration. The
constitutive equations described here do not include creep at constant stress
and stress relaxation at constant strain. Thus, the material coefficients
that specify the constitutive relationship between the stress and strain
components are assumed to be constant during the deformation. This does
not automatically imply that we neglect temperature effects on deformation.
We account for the thermal expansion of the material, which can produce
strains or stresses as large as those produced by the applied mechanical forces.
Here, we discuss the constitutive equations of linear elasticity (i.e., relations
between stress and strain are linear) for the case of infinitesimal deformation
(i.e., |∇u| << 1). Hence, we will not distinguish between various measures of
stress and strain, and use S ≈ σ for the stress tensor and E ≈ ε for strain tensor
in the material description used in solid mechanics. The linear constitutive
model for infinitesimal deformation is referred to as the generalized Hooke’s
law.
Before we discuss the generalized Hooke’s law, a comment is in order about
the so-called principle of material frame-indifference. Functions and fields
whose values are scalars, vectors, or tensors are called frame-indifferent or
objective if both the dependent and independent vector and tensor variables
transfor according to the following equations:
σ = C : ε + σ0, 0
σij = Cijk` εk` + σij , (7.1)
∂U0 0
σij = = Cijk` εk` + σij , (7.2)
∂εij
we have
∂ 2 U0
= Cijk` .
∂εij ∂εk`
Since the order of differentiation is arbitrary, ∂ 2 U0 /∂εij ∂εk` = ∂ 2 U0 /∂εk` ∂εij ,
it follows that Cijk` = Ck`ij . This reduces the number of independent material
stiffness components to 21. To show this we express Eq. (7.1) in an alternate
form using single subscript notation for stresses and strains and two subscript
notation for the material stiffness coefficients:
11 → 1 22 → 2 33 → 3 23 → 4 13 → 5 12 → 6. (7.3b)
It should be cautioned that the single subscript notation used for stresses
and strains and the two-subscript components Cij render them non-tensor
J. N. REDDY 39
Now the coefficients Cij must be symmetric (Cij = Cji ) by virtue of the
assumption that the material is hyperelastic. Hence we have 6 + 5 + 4 + 3 + 2 +
1 = 21 independent stiffness coefficients for the most general elastic material.
We assume that the stress-strain relations (7.4a,b) are invertible. Thus,
the components of strain are related to the components of stress by
where Sij are the material compliance parameters with [S] = [C]−1 (the
compliance tensor is the inverse of the stiffness tensor: S = C−1 ). In matrix
form Eq. (7.5a) becomes
coordinate system (x, y, z) used to write the equatins of motion and strain-
displacement equations will be called the problem coordinates to distinguish
them from the material cordinate system. Note that the phrase ‘material
cordinates’ used in connection with the material decription should not be
confused with the present term. In the remaining discussion, we will use the
material desciption for everything, but we may use one material coordinate
system, say (x, y, z), to describe the kinematics as well as stress state in
the body and another material coordinate system (x1 , x2 , x3 ) to describe the
stress-strain behavior. Both are fixed in the body, and the two systems are
oriented with respect to each other. When elastic material parameters at a
point have the same values for every pair of coordinate systems that are mirror
images of each other in a certain plane, that plane is called a material plane
of symmetry (e.g., symmetry of internal structure due to crystallographic
form, regular arrangement of of fibers or molecules, etc.). We note that
the symmetry under discussion is a directional property and not a positional
property. Thus, a material may have certain elastic symmetry at every point
of a material body the properties may vary from point to point. Positional
dependence of material properties is what we called the inhomogeneity of the
material.
In the following we discuss various planes of symmetry and forms of
associated stress-strain relations. Note that use of the tensor components
of stress and strain is necessary because the transformations are valid only
for the tensor components. The fourth-order tensor, for example, transforms
according to the formula
0
Cijkl = `ip `jq `kr `ls Cpqrs (7.6)
where `ij are the direction cosines associated with the coordinate systems
(x1 , x2 , x3 ) and (x01 , x20 , x03 ), and Cijkl
0 and Cpqrs are the components of the
fourth-order tensor C in the primed and unprimed coordinates systems,
respectively.
while all ther independent stress and strain components remain unchanged
in value by the change from one coordinate system to the other. Using the
stress-strain relations of the form in (7.4b), we can write
σ10 = C11 ε01 + C12 ε20 + C13 ε03 + C14 ε40 + C15 ε05 + C16 ε06
σ1 = C11 ε1 + C12 ε2 + C13 ε3 − C14 ε4 − C15 ε5 + C16 ε6
Note that the elastic parameters Cij are the same for the two coordinate
systems because they are the mirror images in the plane of symmetry. From
the above two equations (subtract ne from the other) we arrive at
The above equation holds only if C14 = 0 and C15 = 0. Similar discussion
with the two alternative expressions of the remaining stress components yield
C24 = 0 and C25 = 0; C34 = 0 and C35 = 0; and C46 = 0 and C56 = 0.
Thus out of 21 material parameters, we only have 21 − 8 = 13 independent
parameters, as indicated below
and so on. These constants are measured using simple tests like uniaxial
tension test or pure shear test. Because of their direct and obvious physical
meaning, engineering constants are used in place of the more abstract stiffness
coefficients Cij and compliance coefficients Sij . Next we discuss how to relate
Sij to the engineering constants.
One of the consequences of linearity (both kinematic and material
linearizations) is that the principle of superposition applies. That is, if the
applied loads and geometric constraints are independent of deformation, the
sum of the displacements (and hence strains) produced by two sets of loads
is equal to the displacements (and strains) produced by the sum of the two
sets of loads. In particular, the strains of the same type produced by the
application of individual stress components can be superposed. For example,
(1)
the extensional strain ε11 in the material coordinate direction x1 due to the
stress σ11 in the same direction is σ11 /E1 , where E1 denotes Young’s modulus
(2)
of the material in the x1 direction. The extensional strain ε11 due to the stress
σ22 applied in the x2 direction is −ν21 σ22 /E2 , where ν21 is the Poisson ratio
ε11
ν21 = −
ε22
moduli in the xi -xj plane. Writing Eqs. (a)-(d) in matrix form, we obtain
1
E1 − νE212 − νE313 0 0 0
ε1 σ
− νE121 1
− νE323 0 0 0 1
ε2
E2 σ2
− νE131 − νE232
1
0 0 0
ε3 E3
σ3
= σ4 , (7.13)
ε4 1
0 0 0 G23 0 0
ε 1 σ
5
0 0 0 0 G13 0
5
ε6 0 0 0 0 0 1 σ6
G12
or, in short
νij νji
= (no sum on i, j) (7.14)
Ei Ej
for i, j = 1, 2, 3. The 9 independent material coefficients for an orthotropic
material are
E1 , E2 , E3 , G23 , G13 , G12 , ν12 , ν13 , ν23 . (7.15)
It is important to note the difference, for example, between νij and νji for
i 6= j for an orthotropic material. For example the difference between ν12 and
ν21 for an orthotropic material is illustrated in Figure 7.1 with two cases of
uniaxial stress for a square element of length a. First a stress σ is applied in
the x1 -direction as shown in Figure 7.1a. The resulting strains are
(1) σ (1) ν12
ε11 = ε22 = − σ (7.16)
E1 E1
where the direction of loading is denoted by the superscript and negative sign
indicates compression. Next, the same value of stress is applied in the x2 -
direction as shown in Figure 7.1b. The strains are
(2) ν21 (2) σ
ε11 = − σ ε22 = (7.17)
E2 E2
J. N. REDDY 45
(1) (2)
While it is obvious that ε11 < ε22 if E1 > E2 , we have no clue about the
(1) (2)
relative magnitudes of ε22 and ε11 . However, the displacements associated
with the two loadings are
(1) σ (1) ν12
u1 = a u2 = −a σ (7.18a)
E1 E1
(2) ν21 (2) σ
u1 = −a σ u2 = a σ (7.18b)
E2 E2
(1) (2)
and the reciprocal relation (7.14) gives u2 = u1 , which is the statement of
Betti’s reciprocity theorem (we shall study it in the later chapters).
Comparing Eqs. (7.11) and (7.13), we note that
1 ν12 ν13
S11 = , S12 = − , S13 = −
E1 E1 E1
1 ν23 1
S22 = , S23 = − , S33 =
E2 E2 E3
1 1 1
S44 = , S55 = S66 = (7.19)
G23 G13 G12
46 THEORY OF ELASTICITY
Consequently, Eqs. (7.10) and (7.13), in view of the relations (7.19), (7.20)
and (7.21) [also note that 1 − 3ν 2 − 2ν 3 = (1 + ν)2 (1 − 2ν)], take the form
σ 1−ν ν ν 0 0 0 ε1
1
1−ν
σ2
ν ν 0 0 0
ε2
σ3 ν ν 1−ν 0 0 0 ε3
= Λ ,
σ4
0
0 0 1 − 2ν 0 0 ε4
1 − 2ν
σ5 ε5
0 0 0 0 0
σ6 0 0 0 0 0 1 − 2ν ε6
(7.22)
ε1 1 −ν −ν 0 0 0 σ1
−ν −ν
ε2
1 0 0 0
σ 2
1
ε3
−ν −ν 1 0 0 0
σ 3
= , (7.23)
ε4
E 0 0 0 1+ν 0 0 σ4
ε
5
0 0 0 0 1+ν 0 σ5
ε6 0 0 0 0 0 1+ν σ6
where
E
Λ= (7.24)
(1 + ν)(1 − 2ν)
J. N. REDDY 47
where λ and µ are called Lamé constants. Therefore, the stress-strain relation
for the isotropic case takes the form
We note the following relations between the Lamé constants λ and µ and
engineering constants E, ν and G for an isotropic material:
µ(3λ + 2µ) λ 2
E= , ν= , K = λ + µ, G=µ (7.28)
λ+µ 2(µ + λ) 3
The following definitions and constitutive relations are of interest in the sequel:
1
mean stress, σ̃ ≡ σii , dilatation, e ≡ εii (7.29)
3
1
deviatoric stress, σ 0 = σ − σ̃I ,deviatoric strain, ε0 = ε − tr(ε) (7.30)
3
2
σii = (3λ + 2µ)εii , σ̃ = Ke, K = λ + µ (7.31)
3
where K is the bulk modulus and µ = G is the shear modulus.
In view of the relations between the Lamé constants and engineering
constants, Eqs. (7.26) and (7.27) can be written in terms of engineering
constants:
E νE E νE
σij = εij + εkk δij , σ = ε+ tr(ε)I
1+ν (1 + ν)(1 − 2ν) 1+ν (1 + ν)(1 − 2ν)
(7.32)
1 1
εij = [(1 + ν)σij − νσkk δij ] , ε = [(1 + ν)σ − ν tr(σ)I] (7.33)
E E
The strain energy density for a linear isotropic material is given by
1 1
U0 = Cijk` εij εk` = σij εij
2 2
1
= (σ11 ε11 + σ22 ε22 + σ33 ε33 + 2σ12 ε12 + 2σ13 ε13 + 2σ23 ε23 ) (7.34)
2
48 THEORY OF ELASTICITY
where the Qij , called the plane stress-reduced stiffnesses, are given by
S22 E1
Q11 = 2 = ,
S11 S22 − S12 1 − ν12 ν21
S12 ν12 E2
Q12 = 2 = ,
S11 S22 − S12 1 − ν12 ν21
S11 E2
Q22 = 2 = ,
S11 S22 − S12 1 − ν12 ν21
1
Q66 = = G12 . (7.37)
S66
where Q44 = C44 = G23 and Q55 = C55 = G13 . For an isotropic material, Eqs.
(7.36) and (7.38) take the form
σ1 1 ν 0 0 0 ε1
σ ν 1 0 0 0
ε
2
E 2
0 1−ν
σ4 = 0 2 0 0
ε4 ,
(7.39)
− 2
1 ν 1−ν
σ 0 0 0 0 ε5
5 2
1−ν
σ6 0 0 0 0 2
ε6
J. N. REDDY 49
σ1 Q11 Q12 0 0 0 ε1 − α1 ∆T
ε2 − α2 ∆T
σ
2
Q12 Q22 0 0 0
σ4 =
0 0 Q44 0 0 ε4 , (7.42)
σ 0 0 0 Q55 0 ε5
5
σ6 0 0 0 0 Q66 ε6
Example 7.1:
For an isotropic material with E = 71.0 GPa and G = 26.6 GPa, determine the strain tensor
and the strain energy density at a point in a body if the stress tensor at that point is given
Ý} by the following matrix of scalar components:
with respect to the vector basis {e
"
20 −4 5
#
[σ] = −4 0 10 MPa
5 10 15
1
U0 = Cijk` εij εk`
2
1
= [λδij δk` + µ (δik δj` + δi` δjk )] εij εk`
2
1 1
= (λδij εkk + 2µεij ) εij = σij εij
2 2
1
= (σ11 ε11 + σ22 ε22 + σ33 ε33 + 2σ12 ε12 + 2σ13 ε13 + 2σ23 ε23 )
2
1
= [(20)(211) + (0)(−165) + (15)(117) + 2(−4)(−75) + 2(5)(94) + 2(10)(188)] 103
2
= 5637.5 Pa