8XTjICF2ER Elements of Continuum Mechanics PDF
8XTjICF2ER Elements of Continuum Mechanics PDF
8XTjICF2ER Elements of Continuum Mechanics PDF
This text is intended to provide a modern and integrated treatment of the founda-
tions and applications of continuum mechanics. There is a significant increase in
interest in continuum mechanics because of its relevance to microscale phenomena.
In addition to being tailored for advanced undergraduate students and including
numerous examples and exercises, this text also features a chapter on continuum
thermodynamics, including entropy production in Newtonian viscous fluid flow and
thermoelasticity. Computer solutions and examples are emphasized through the
use of the symbolic mathematical computing program MathematicaÒ.
Joanne L. Wegner
University of Victoria
James B. Haddow
University of Victoria
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo
www.cambridge.org
Information on this title: www.cambridge.org/9780521866323
© Joanne L. Wegner and James B. Haddow 2009
Preface Page ix
v
vi Contents
3. Stress ...................................................................................................... 84
3.1 Contact Forces and Body Forces 84
3.2 Cauchy or True Stress 87
3.3 Spatial Form of Equations of Motion 90
3.4 Principal Stresses and Maximum Shearing Stress 91
3.5 Pure Shear Stress and Decomposition of the Stress Tensor 94
3.6 Octahedral Shearing Stress 95
3.7 Nominal Stress Tensor 98
3.8 Second Piola-Kirchhoff Stress Tensor 100
3.9 Other Stress Tensors 101
EXERCISES 102
REFERENCES 104
EXERCISES 147
REFERENCES 149
Index 272
Preface
ix
x Preface
1.1 Introduction
1
2 Cartesian Tensor Analysis
x ¼ x1 e1 þ x2 e2 þ x3 e3 ,
x3
e3
0 e2
x2
e1
x1
u ¼ u1 e1 þ u2 e2 þ u3 e3 :
3
u ¼ + ui e i : ð1:1Þ
i¼1
u ¼ ui ei ¼ u1 e1 þ u2 e2 þ u3 e3 ð1:2Þ
without the summation sign. The sum of two vectors is commutative and is
given by
u þ v ¼ v þ u ¼ ðui þ vi Þei ,
u v ¼ ui vi ¼ u1 v1 þ u2 v2 þ u3 v3 : ð1:3Þ
Repeated suffixes are often called dummy suffixes since any letter that does
not appear elsewhere in the expression may be used, for example,
ui vi ¼ uj vj :
Equation (1.3) indicates that the scalar product obeys the commutative law
of algebra, that is,
u v ¼ v u:
6 Cartesian Tensor Analysis
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
juj ¼ u u ¼ ui ui :
Other examples of the use of suffix notation and the summation convention
are
A suffix that appears once in a term is known as a free suffix and is un-
derstood to take in turn the values 1, 2, 3 unless otherwise indicated. If
a free suffix appears in any term of an equation or expression, it must
appear in all the terms.
ei ej ¼ dij , ð1:4Þ
1 for i ¼ j
dij ¼ : ð1:5Þ
0 for i 6¼ j
The base vectors ei are orthonormal, that is, of unit magnitude and mutu-
ally perpendicular to each other. The Kronecker delta is sometimes called
the substitution operator because
x2
e2
x2′
e1′
x1′
e2′
e1
0 x1 Figure 1.2. Change of axes.
x3 e3
e3′
x3′
The coordinates of a point P are xi with respect to 0xi and x9i with
respect to 0xi . Consequently,
where the e9i are the unit base vectors for the system 0xi . Forming the inner
product of each side of equation (1.7) with e9k and using equation (1.4) and
the substitution operator property equation (1.6) gives
where
Similarly
It is evident that the direction of each axis 0x9k can be specified by giving its
direction cosines aki ¼ e9k ei ¼ cosðx9k 0xi Þ referred to the original axes
0xi . The direction cosines, aki ¼ e9k ei , defining this change of axes are
tabulated in Table 1.1.
The matrix ½a with elements aij is known as the transformation matrix;
it is not a tensor.
8 Cartesian Tensor Analysis
e1 9 e2 9 e3 9
e1 a11 a21 a31
e2 a12 a22 a32
e3 a13 a23 a33
@x9k @xi
aki ¼ ¼ , ð1:11Þ
@xi @x9k
@xi @x9j
ei ¼ ej9 ¼ e9k , ð1:12Þ
@x9k @x9k
since @xj =@xk ¼ @jk , and from equations (1.11) and (1.12) that
and
Equations (1.13) and (1.14) are the transformation rules for base vectors.
The nine elements of aij are not all independent, and in general,
aki 6¼ aik :
2
det aij ¼ 1, ð1:19Þ
where det aij denotes the determinant of aij . A detailed discussion of
determinants is given in section 10 of this chapter. The negative root of
equation (1.19) is not considered unless the transformation of axes involves
a change of orientation since, for the identity transformation xi ¼ x9i ,
aik ¼ dik and det½dik ¼ 1. Consequently, det½aik ¼ 1, provided the trans-
formations involve only right-handed systems (or left-handed systems).
The transformations (1.8), (1.10), and (1.15) subject to equation (1.17)
or (1.18) are known as orthogonal transformations. Three quantities ui are
the components of a vector if, under orthogonal transformation, they trans-
form according to equation (1.15). This may be taken as a definition of
a vector. According to this definition, equations (1.8) and (1.10) imply that
the representation x of a point is a bound vector since its origin coincides
with the origin of the coordinate system.
10 Cartesian Tensor Analysis
/9 ¼ det aij /, ð1:20Þ
u9i ¼ det aij aij uj : ð1:21Þ
aij ¼ dij þ 2ni nj , det aij ¼ 1:
B c A
so that
where c is any vector. The tensor product has nine components obtained as
follows:
ðu vÞij ¼ ui vj , ð1:23Þ
12 Cartesian Tensor Analysis
2 3 2 3
u1 ½v1 v2 v3 u1 v1 u1 v 2 u1 v3
½u vij ¼ 4 u2 5 ¼ 4 u2 v1 u2 v 2 u2 v3 5, ð1:24Þ
u3 u3 v1 u3 v 2 u3 v3
u v ¼ ui vj ei ej : ð1:25Þ
The basis vectors ei ej are not vectors in the sense of elementary vector
analysis but belong to a nine-dimensional vector space.
A second-order tensor S is a linear transformation from V3 into V3 that
assigns to each vector t a vector
s ¼ St: ð1:26Þ
2 3 2 32 3
s1 S11 S12 S13 t1
4 s2 5 ¼ 4 S21 S22 S23 54 t2 5, ð1:27Þ
s3 S31 S32 S33 t3
where the square matrix on the right-hand side of equation (1.27) is the matrix
of the nine components Sij, of S referred to the rectangular Cartesian coordi-
nate axes 0xi . Equation (1.26) or (1.27) can be written in suffix notation as
si ¼ Sij tj , ð1:28Þ
S ¼ Sij ei ej ; ð1:29Þ
and this shows that a second-order tensor is the weighted sum of tensor
products of the base vectors. The left-hand side of equation (1.29) is in
coordinate free symbolic form; however, the right-hand side involves the
components Sij , which are with respect to a particular coordinate system.
Since a tensor is invariant under coordinate transformation,
and
for the componts of S are obtained. It follows from equation (1.29) that the
Cartesian components of S are given by
that is, Sij is the scalar product of the base vector ei with the vector Sej .
It is convenient at this stage to introduce the transpose of a vector and
second-order tensor. The components of the transpose uT of a vector u
form a row matrix ðu1 , u2 , u3 Þ. The transpose ST of a second-order tensor S
has components STij ¼ Sji so that
S T ¼ Sji ei ej :
14 Cartesian Tensor Analysis
u S T v ¼ v ðSuÞ:
ðsÞ ð Þ
C ¼ C þC a ,
where
ðsÞ 1
C ¼ C þ CT
2
ðaÞ 1
C ¼ C CT
2
2 3
1 2 3
Cij ¼ 4 4 2 3 5:
5 2 1
1.5 Concept of a Second-Order Tensor 15
2 3
h i 1 4 5
CTij ¼ 4 2 2 2 5:
3 3 1
Let the tensors CðsÞ and C ðaÞ denote the symmetric and antisymmetric parts
of C, respectively; then
2 3 2 3
h i h i 2 2 8 1 1 4
ðsÞ 1 14 4
C ¼ CþC T
¼ 2 4 5
5 ¼ 1 2 2:5 5
2 2
8 5 2 4 2:5 1
and
2 3 2 3
h i h i 0 6 2 0 3 1
ð Þ 1 1
Ca ¼ C CT ¼ 4 6 0 1 5 ¼ 4 3 0 0:5 5:
2 2
2 1 0 1 0:5 0
The inner or scalar product of the vectors w and Su obeys the com-
mutative rule and is denoted by
w Su or Su w: ð1:33Þ
since the left-hand side is a scalar and the right-hand side is a tensor. Also
we note that
w Su ¼ Sij wi uj : ð1:35Þ
The product, not to be confused with the dyadic product, of two second-
order tensors S and T is a second-order tensor with components
ST 6¼ TS,
that is, the product of two second-order tensors does not, in general, satisfy
the commutative law.
1.5 Concept of a Second-Order Tensor 17
The components of the unit tensor I are given by the Kronecker delta
so that
I ¼ dij ei ej : ð1:37Þ
The unit tensor is an isotropic tensor, that is, it has the same components
referred any coordinate system since
where the substitution property of the Kronecker delta has been employed.
It may also be shown that
I ¼ ei ei : ð1:38Þ
u ¼ Iu, ð1:39Þ
and
S ¼ IS ¼ SI: ð1:40Þ
T
tr S T ¼ Sij T ij , ð1:41Þ
where tr denotes the trace and has the same significance as in linear algebra
so that trS ¼ Sii , that is, it is the sum of the diagonal elements of the
component matrix. The magnitude of a second-order tensor is a scalar de-
fined by,
1=2 n o1=2
jS j ¼ Sij Sij ¼ tr ST S : ð1:42Þ
18 Cartesian Tensor Analysis
ðST ÞT ¼ T T S T : ð1:43Þ
C1 C ¼ CC 1 ¼ I, ð1:44Þ
h
T i
tr ei ej ðek et Þ ¼ dik djt , ð1:45Þ
and a simple application of the tensor transformation rule shows that the
sum of second-order tensors is a second-order tensor. Consequently, the set
of all second-order tensors is a nine-dimensional linear vector space with
orthonormal basis ei ej , where the term vector is taken in its more general
sense as belonging to a linear vector space.
It is often convenient to decompose a second-order tensor T into two
parts, a deviatoric part and an isotropic part, as follows
T ij ¼ T ij þ Td
^ ij , ð1:46Þ
where T^ ¼ Tkk =3 ¼ trT=3. The deviatoric part with components T ij has
^ ij has the same
the property, trT ¼ T ¼ 0, and the isotropic part Td
ii
components in all coordinate systems. The decomposition equation
(1.46) is unique and is important in linear elasticity and plasticity theory.
The concept of tensor multiplication can be extended to more than
two vectors or to higher-order tensors. For example, the tensor product
T S of two second-order tensors T and S is a fourth-order tensor with
1.6 Tensor Algebra 19
The following results are easily verified by using the transformation rule.
are the components of a vector. The reader should prove this. If Cij is a set
of nine quantities possessing the property that Cij uj is a first-order tensor
for an arbitrary vector u, then Cij are the components of a second-order
tensor. Extension to consider higher-order quantities is straightforward.
det½A kI ¼ 0, ð1:51Þ
k3 J1 k3 þ J2 k J3 ¼ 0, ð1:52Þ
where
1h i 1h i
J 1 ¼ Aii ¼ tr A, J2 ¼ ðAii Þ2 Aik Aki ¼ ðtrAÞ2 trðA2 Þ ; J 3
2 2
¼ det½ A:
1.7 Invariants of Second-Order Tensors 21
ðÞ
Aij dij k1 nj 1 ¼ 0, ð1:53Þ
and the condition nð1Þ nð1Þ ¼ 1. If the x91 axis of a new coordinate system is
in this direction, the tensor A has the primed components
2 3
k1 k 0 0
det4 0 A922 k A923 5 ¼ 0
0 A923 A933 k
or
h i
ðk1 kÞ k2 ðA922 þ A933 Þk þ A922 A933 A9232 ¼ 0: ð1:54Þ
22 Cartesian Tensor Analysis
h 2
i
k2 ðA922 þ A933 Þk þ A922 A933 A923 ¼ 0, ð1:55Þ
2
ð A922 þ A933 Þ2 4 A922 A933 A9232 ¼ ðA922 A933 Þ2 þ4 A923 > 0: ð1:56Þ
ð Þ ð Þ
Aij nj 1 ¼ k1 ni 1 . ð1:57Þ
Similarly
ð Þ ð Þ
Aij nj 2 ¼ k2 ni 2 : ð1:58Þ
ð2Þ ð1Þ
Scalar multiplication of equation (1.57) by ni and equation (1.58) by nj
and subtracting gives
ðiÞ ðiÞ
a9ij ¼ nj ¼ n ej ,
1.7 Invariants of Second-Order Tensors 23
ðiÞ ð jÞ
A0i ¼ nk nt Akt , ð1:59Þ
2 3
k1 0 0
A0ij ¼ 4 0 k2 0 5.
0 0 k3
This is known as the canonical form of the matrix of components. For many
problems and proofs, it is desirable to refer components of second-order
symmetric tensors to the principal axes. Equation (1.59) can be expressed
in symbolic form
ðÞ
Nij ¼ nj i :
3
ðÞ ðÞ
A ¼ + ki n i n i ,
i¼1
which is known as the spectral form. If two eigenvalues are equal, that is, if
k1 ¼ k2 ¼ k3 , then
ð Þ ð Þ ð Þ ð Þ
A ¼ k1 n 1 n 1 þ k2 I n 1 n 1 :
A ¼ kI,
where ½I is the unit ð3 3 3Þ matrix and ½k is the diagonal matrix of the
components of C referred to the principal axes. Since a tensor equation
1.9 Higher-Order Tensors 25
C 3 J1 C2 þ J2 C J3 I ¼ 0, ð1:61Þ
Dijk ei ej ek ,
D9ijk ¼ air ajs akt Drst , Dijk ¼ air ajs akt D9rst :
ei 3 ej ¼ eijk ek ,
u 3 v ¼ ui vj ei 3 ej ¼ eijk ui vj ek : ð1:65Þ
A useful identity is
ðu 3 vÞ ðs 3 t Þ ¼ eijk eipq uj vk sp tq ,
and
u 3 v ¼ ðu1 e1 þ u2 e2 þ u3 e3 Þ 3 ðv1 e1 þ v2 e2 þ v3 e3 Þ
¼ e1 ðu2 v3 v2 u3 Þ þ e2 ðu3 v1 v3 u1 Þ þ e3 ðu1 v2 v1 u2 Þ:
Similarly
1.10 Determinants
2 3
A1 A2 A3
det4 B1 B2 B3 5 ¼ eijk Ai Bj Ck , ð1:68Þ
C1 C2 C3
2 3
e1 e2 e3
4
u 3 v ¼ det u1 u2 u3 5,
v1 v2 v3
2 3
u1 u2 u3
4
u v 3 w ¼ det v1 v2 v3 5 ¼ eijk ui vj wk : ð1:69Þ
w1 w2 w3
2 7
C11 C12 C13 7
7
6 7
det 4 C21 C22 C23 5 ¼ eijk C1i C2j C3k expanded by rows
ð1:70Þ
C31 C32 C33
¼ eijk Ci1 Cj2 Ck3 expanded by columns:
1
eijk erst Cir Cjs Ckt ¼ erst det½C: ð1:73Þ
6
In equations (1.70) to (1.73) we assume that Cij are the components of the
second-order tensor C, although the results are valid for any ð3 3 3Þ ma-
trix. The determinant of a second-order tensor is a scalar and is known as
the third invariant of the tensor, as already mentioned in section 7. If
det½C ¼ 0, the tensor is said to be singular. The coefficient of Cij in the
expansion of the determinant is called the cofactor of Cij and is denoted by
Cijc , so that expansion by rows gives
c c c
det½C ¼ C1j C1j ¼ C2j C2j ¼ C3j C3j ,
c
It may be deduced from equation (1.74) that Ckj are the components of
a second-order tensor; consequently, we can express equation (1.74) in
symbolic notation as,
where the matrix of components of CcT is the transpose of the matrix of the
cofactors of C. It then follows from equation (1.74) that
c 1
Cmi ¼ eijk eimst Csj Ctk :
2
c
Compare with Cri Cmi c
¼ drm det½C to obtain Cmi ¼ 12 eijk emst Csj Ctk .
The inverse of C is a second-order tensor denoted by C 1 and
If the tensor C is singular, the inverse does not exist. The term inverse
is meaningful for scalars and second-order tensors but not for tensors of
any other order. The relation ðCBÞðCBÞ1 ¼ I follows from equation
(1.76), and premultipying each side first by C1 and then by B1 gives
the the result
c
Cji
C1
ij ¼ : ð1:77Þ
det½C
h i
T
det½C ¼ det C : ð1:78Þ
Au Av 3 Aw ¼ det½Aðu v 3 wÞ:
SOLUTION.
det½aC ¼ a3 det½C,
h h ii1
det½C ¼ det C 1 , ð1:80Þ
QQT ¼ QT Q ¼ I, ð1:81Þ
QT ¼ Q1 ,
det½Q ¼ 61,
u v ¼ Qu Qv, ð1:82Þ
ui vi ¼ Qij Qik uj vk :
and positive definite if the equality sign holds if and only if u is the null
vector. The antisymmetric part of S does not contribute to u Su since, if W
is an antisymmetric second-order tensor,
u Wu ¼ Wij ui uj ¼ 0:
T
u Su ¼ u A A u ¼ ðAuÞ ðAuÞ > 0: ð1:85Þ
u AT A u ¼ uj Aij Aik uk ¼ Aij uj ðAik uk Þ:
where k1 , k2 , k3 are the eigenvalues of S and u^1 , u^2 , and u^3 are the compo-
nents of u referred to the principal axes of S. When all the eigenvalues of
S are negative, it follows from equation (1.84) that u Su < 0 with the
equality sign holding only when u is the null vector. The tensor S is then
said to be negative definite.
The square root S1=2 of a positive definite tensor S is defined as the
symmetric second-order tensor whose spectral form is
3
ð Þ ð Þ
S 1=2 ¼ + ki1=2 n 1 n 1 ð1:86Þ
i¼1
1.11 Polar Decomposition Theorem 35
det½A > 0
A ¼ QU, ð1:87Þ
A ¼ VQ, ð1:88Þ
36 Cartesian Tensor Analysis
1=2 1=2
T 1
T T
A A A A A A ¼ A AT A AT
1
¼ AA1 AT AT ¼ I,
1=2
and det½P ¼ det½Q ¼ 1, so that A ¼ QU ¼ VP, where U ¼ AT A
1=2
and V ¼ AAT .
The square root of a positive definite tensor is unique, when the pos-
itive roots are taken in equation (1.83), so that the decomposition A ¼ QU
is unique. We now have to show that Q ¼ P to complete the proof of the
theorem. This follows since
A ¼ VP ¼ P P T VP ,
ðiÞ T ðiÞ
l ¼ Q m : ð1:89Þ
C ¼ AT A
is given by
2 32 3
1:1 0:4 0:2 1:1 0:5 0:3
6 76 7
½C ¼ 4 0:5 1:2 0:3 54 0:4 1:2 0:5 5
0:3 0:5 0:9 0:2 0:3 0:9
2 3
1:41 0:97 0:35
6 7
¼ 4 0:97 1:78 1:02 5
0:35 1:02 1:15
C1=2 ¼ U
by
2 3
1:11294 0:406455 0:0785375
6 7
½U ¼ 4 0:406455 1:18478 0:459452 5,
0:0785375 0:459452 0:965783
2 3
h i 1:036 0:3956 0:104
1 6 0:532 7
U ¼ 4 0:3956 1:186 5
0:104 0:532 1:280
Q ¼ AU 1
is
2 32 3
1:1 0:5 0:3 1:036 0:3956 0:104
6 76 7
½Q ¼ 4 0:4 1:2 0:5 54 0:3956 1:186 0:532 5
0:2 0:3 0:9 0:104 0:532 1:280
2 3
0:9726 0:018 0:2324
6 7
¼ 4 0:0085 1:0 0:04314 5:
0:2322 0:4393 0:9717
@/ @/ @/
0 ¼ 0 ,
@xj @xi @x
j
1.12 Tensor Fields 39
@/ @/
0 ¼ aji : ð1:90Þ
@xj @xi
Equation (1.90) is the transformation rule for the vector ð@/=@xi Þei 2 V 3 ,
which is known as the gradient of / and is often denoted by =/ or grad/. At
any point =/ is normal to the surface /ðx1 , x2 , x3 Þ ¼ constant passing
through the point and is in the direction of / increasing. The scalar
=/ n, where n is a unit vector, is the rate of change of / with respect to
the distance in the direction n.
We use the symbol = preceding an operand to denote either the dif-
ferential operator
@ðÞ
= [ ei ð1:91Þ
@xi
or
@ðÞ
=[ ei , ð1:92Þ
@xi
@vi
=v ¼ :
@xi
2
@
ð= =Þ [ =2 [ :
@xi @xi
40 Cartesian Tensor Analysis
@vk
curl v ¼ eijk ei ; ð1:93Þ
@xj
so that
curl v [ = 3 v,
=3v
0
@vt @vn
0 ¼ atn amj :
@xm @xj
@vk
=v ¼ ek ei : ð1:94Þ
@xi
2 3
@v1 @v1 @v1
6 @x1 @x2 @x3 7
6 7
@vi 6 7
6 @v @v2 @v2 7
¼ 6 2 7: ð1:95Þ
@xj 6 @x1 @x2 @x3 7
6 7
4 @v3 @v3 @v3 5
@x1 @x2 @x3
1.12 Tensor Fields 41
It follows from equation (1.95) that = v ¼ trð= vÞ. The notation vk;i is
often used for @vk =@xi , and this is compatible with equations (1.94) and
(1.95). When definition (1.92) is used we obtain the transpose of the right-
hand side of equation (1.95). With definition (1.92),
dv ¼ ð= vÞdx;
df
df ðxÞ ¼ dx
dx
dv ¼ dx ð= vÞ:
In spite of the difficulty with the curl it is usually more convenient to adopt
definition (1.92). The gradient of the vector field vðxÞ may also be defined as
the linear transformation that assigns to a vector a the vector given by the
following rule:
@Tij
divT ¼ ej ð1:96Þ
@xi
42 Cartesian Tensor Analysis
@T ij
divT ¼ ei ð1:97Þ
@xj
@T ij
=T ¼ ei ej ek ,
@xk
1 T
D ¼ = v þ ð= vÞ Þ
2
has components
1
Dij ¼ vi;j þ vj;i ,
2
where we have adopted the notation vi;j for @vi =@xj , and the antisymmetric
part
1
V ¼ = v ð= vÞT
2
1.12 Tensor Fields 43
has components
1
Vij ¼ vi;j vj;i :
2
1
w ¼ e1 v3;2 v2;3 þ e2 v1;3 v3;1 þ e3 v2;1 v1;2 : ð1:98Þ
2
We note that the independent components of Vij are related to the com-
ponents of the vector w as follows,
1
wk ¼ eijk Xij : ð1:100Þ
2
ðsÞ
eijk Cjk ¼ 0, ð1:101Þ
Z Z
@/
dv ¼ /ni da, ð1:102Þ
@xi
V @V
where v denotes the region, @v is the boundary with area a, and n is the
outward unit normal to the boundary. The divergence theorem
Z Z
divudv ¼ u nda,
v @v
Z Z
@ui
dv ¼ ui ni da ð1:103Þ
@xi
v @v
Z Z
divTdv ¼ T ij nj da
@V
V
Z Z
n 3 uda ¼ curludv,
@R R
SOLUTION.
Z Z Z Z
@uk
n 3 uda ¼ eijk nj uk da ei ¼ eijk dv ei ¼ curludv:
@R @R R @xj R
H
where denotes the integral around the curve C, dx is a vector element
of C, and n is the unit normal to A taken in the sense of the right-hand
rule.
46 Cartesian Tensor Analysis
EXERCISES
(a) curl = a ¼ 0:
(b) divcurl u ¼ 0:
(c) divðauÞ ¼ adivu þ u grada:
(d) curlðauÞ ¼ acurlu þ =a 3 u:
(e) curlcurlu ¼ =divu =2 u:
(f) a=2 b ¼ divða=bÞ =a =b:
(g) u 3 curlu ¼ 12 =ðu uÞ u =u:
1.2. Consider the tensor I n n ¼ dij ni nj ei ej , where n is a unit
vector. Show that ðI n nÞu is the projection of u onto a plane
normal to n and that
ðI n nÞ2 ¼ ðI n nÞ:
h i
det S 1=2 ¼ fdet½S g1=2 :
1.4. Let Tij be the components of the deviatoric part of the second-order
symmetric tensor T : Prove that
1
det½T ¼ trðT 3Þ:
3
1.5. Consider the nine quantities Aij referred to a given coordinate system.
If Aij Bij transforms as a scalar under change of coordinate system for
an arbritrary second-order symmetric tensor B, prove that Aij þ Aij
are components of a second-order tensor.
1.6. Show that for a proper orthogonal tensor Q, there exists a vector u
such that Qu ¼ u.
2 3
3 2 1:5
½B ¼ 4 2 4 1 5:
1:5 1 2
2 3
h i 1:57443 0:542316 0:476527
B1=2 ¼ 4 0:542316 1:91127 0:230106 5
0:476527 0:230106 1:31148
det½Q I ¼ 0:
Z Z
u snda ¼ fu divs þ ð= uÞsgdv:
@V V
Z Z
ð= uÞdv ¼ ðn uÞda,
v @V
Z Z
/=2 w þ =/ =w dV ¼ /=w ndA;
V @V
SUGGESTED READING
50
2.1 Description of Motion 51
1 ¼ X t ðx, t Þ,
Descriptions (2) and (3) are emphasized in the applications that follow and
(1) and (4) will play only a minor role in the topics considered in this book.
We will take the reference configuration for description (2) to be the con-
figuration at time t ¼ 0, usually the undeformed state of the body, and will
refer the material and spatial coordinates to the same fixed rectangular
Cartesian coordinate system.
It is convenient to use lowercase suffixes for the spatial configuration,
for example, the position vectors of a particle in the spatial and reference
configurations are given by
x ¼ xi ei ,
and
X ¼ XK EK ,
52 Kinematics and Continuity Equation
ðe1 , e2 , e3 Þ ¼ ðE1 , E2 , E3 Þ:
^
x ¼ xðX, tÞ: ð2:1Þ
^
v ¼ vðx, tÞ: ð2:2Þ
Equation (2.2) represents the velocity at time t of the particle occupying the
point with position vector x. The stream lines for a fixed time t0 are
given by the solution of the three, in general nonlinear, ordinary
differential equations,
^
dxi vi
¼ , i 2 f1, 2, 3g,
ds jvj
where s is the distance measured along the curve so that dx=ds is the unit
vector tangential to the curve. A stream
. line and particle path coincide
^
in steady flow, that is, if @vðx, tÞ @t [ 0, but the converse is not
necessarily true.
2.2 Spatial and Referential Descriptions 53
Equation (2.3) represents the velocity at time t of the particle that occupied
the point X in the reference configuration.
Similarly it follows from equation (2.1) that any quantity f that may be
expressed as a function
^
f ¼ f ðx, tÞ
^
f^ðX, tÞ ¼ f ðxðX, tÞ, tÞ
1 n^ o
^
df ^
¼ lim f ðx þ vDt, t þ DtÞ f ðx, tÞ : ð2:4Þ
dt t!0 Dt
Using suffix notation and expanding the right-hand side of equation (2.4) as
a Taylor series gives1
" ^ #
^
df 1 @f
^
@f ^ ^
2 ^ @f @f
¼ lim vi Dt þ Dt þ O Dt , ¼ vi þ , ð2:5Þ
dt Dt @xi @t @xi @t
1
The meaning of the O symbol is as follows: f ðxÞ ¼ OðgðxÞÞasx ! 0 implies that
limjf ðxÞj=jgðxÞj ¼ K where K is bounded. When K = 0, we use the small o symbol,
x!0
that is, f ðxÞ ¼ ogðxÞ implies that limjf ðxÞj=jgðxÞj ¼ 0.
x!0
54 Kinematics and Continuity Equation
^ ^
df ^ @f
¼ v =f þ : ð2:6Þ
dt @t
^ dx @ ðxðX, tÞÞ
v¼ ¼ ¼ x:
_ ð2:7Þ
dt @t
The motion is called steady flow if the velocity is independent of time, that
is, if @v=@t ¼ 0.
d/ @/ @/
¼ þ vi ¼ 0, ð2:8Þ
dt @t @xi
@/ @/
þ ui ¼ 0 ð2:9Þ
@t @xi
2.3 Material Surface 55
since
/ðx; tÞ ¼ 0:
us ð@/=@xs Þ
uðnÞ ¼ u n ¼ ,
@/ @/ 1=2
@xk @xk
ð@/=@tÞ
uðnÞ ¼ : ð2:10Þ
@/ @/ 1=2
@xk @xk
vs ð@/=@xs Þ
vðnÞ ¼ : ð2:11Þ
@/ @/ 1=2
@xk @xk
Substituting for @/=@t from equation (2.10) and for vs ð@/=@xs Þ from equa-
tion (2.11) in
d/ @/ @/
¼ þ vs
dt @t @xs
gives
d/ @/ @/ 1=2
¼ vðnÞ uðnÞ : ð2:12Þ
dt @xk @xk
If the surface always consists of the same particles, uðnÞ ¼ vðnÞ , and it follows
from equation (2.12) that d/=dt ¼ 0, evaluated on / ¼ 0, is a necessary
condition for / ¼ 0 to be a material surface. This proof of the necessity of
the condition can be given in words. If the surface is a material surface, the
velocity, relative to the surface, of a particle lying in it must be wholly
56 Kinematics and Continuity Equation
tangential (or zero); otherwise we would have flow across the surface.
It follows that the instantaneous rate of variation of / for a surface parti-
cle must be zero, if /ðx, tÞ ¼ 0 is a material surface. To prove the sufficiency
of the condition we put
@UðX, tÞ
¼ 0:
@t
ð Þ ð Þ ð Þ
dv ¼ dx 1 3dx 2 dx 3 ¼ eijk dx1i dx2j dx3k : ð2:15Þ
2.4 Jacobian of Transformation and Deformation Gradient F 57
dX(3) dx(3)
dX(2)
dx(2) dx(1)
dX(1)
Figure 2.1. Deformation of infinitesimal parallelepiped.
Substituting
ðsÞ @xi ðsÞ ðsÞ 2
dxi ¼ dXL þ O dX , s 2 f1, 2, 3g ð2:16Þ
@XL
dv ¼ JdV, ð2:18Þ
where
@xi
J¼det , ð2:19Þ
@XK
@ ðx1 , x2 , x3 Þ
J¼ :
@ ðX1 , X2 , X3 Þ
58 Kinematics and Continuity Equation
0 < J < ‘,
and this also follows from the invertibility of equation (2.1). A motion for
which there is no volume change for any element of the body is described as
isochoric. If a body is assumed to be incompressible, it can only undergo
isochoric motions, otherwise thermodymamic anomalies arise. It follows
from equation (2.18) that J ¼ 1 for an isochoric motion.
The elements Fik ¼ @xi =@XK of the matrix in equation (2.19) are the
components of a second-order two-point tensor,
F ¼ FiK ei EK , ð2:20Þ
2
O @xi O
O
xi xi ¼ XK XK þ O XK XK : ð2:21Þ
@XK
Equation (2.21) should be compared with equation (2.16). The term two-
point tensor indicates the relationship to both the spatial and reference
configurations.
2.5 Reynolds Transport Theorem 59
F ¼ Gradx
where
@ðÞ
Grad [ EK :
@XK
1 @XK 1 @XK
FKi ¼ ,F ¼ E K ei ð2:23Þ
@xi @xi
dX ¼ F 1 dx:
Z Z Z
d @/
/ðx, tÞdv ¼ dv þ /v ndS, ð2:24Þ
dt @t
vðtÞ vðtÞ @vðtÞ
60 Kinematics and Continuity Equation
where v is the velocity field, n is the outward unit normal to the surface S of
the material volume vðtÞ. According to equation (2.24), the rate of change
of the integral of f over the material volume vðtÞ is equal to the integral of
@/=@t over the fixed volume that coincides instantaneously with vðtÞ at time
t plus the flux /v through the surface S of this fixed volume.
In order to prove equation (2.24) we need the result
dJ
¼ J= v, ð2:25Þ
dt
@xi
Jdij ¼ AjK , ð2:26Þ
@XK
where AiK is the cofactor of FiK ¼ @xi =@XK in the determinant J. Since the
derivative of J with respect to t is the sum of the three determinants formed
by replacing the elements of a different row (or column) of each determi-
nant by their time derivatives,
dJ d @xi @vi @vi @xj
¼ AiK ¼ AiK ¼ AiK ,
dt dt @XK @XK @xj @XK
dJ @vi
¼J ¼ J= v: ð2:27Þ
dt @xi
Z Z
d @
/ðx, tÞdv ¼ J /^ðX, tÞ dV, ð2:28Þ
dt @t
vðt Þ V
2.5 Reynolds Transport Theorem 61
^ ðX, tÞ ¼ /ðxðX, tÞ, tÞ. Since the integral on the right-hand side of
where /
equation (2.28) is over a fixed volume we can differentiate inside the in-
tegral sign to obtain
Z Z !
d @ /^ ^ dJ
/ðx, tÞdv ¼ J þ/ dV,
dt @t dt
vðtÞ V
Equation (2.24) can be obtained from equation (2.29) with the use of
the identity
= ð/vÞ ¼ /= v þ v =/,
^
relation (2.6), with f replaced by /, and the divergence theorem.
Equations (2.24) and (2.29) are valid when /ðx, tÞ is continuous. Sup-
pose /ðx, tÞ has a discontinuity across a surface +ðtÞ, that moves with
velocity wðx, tÞ, not necessarily equal to the field velocity v, and divides
the volume v into two parts, vþ and v , as indicated in Figure 2.2. The unit
normal m to +ðtÞ points from the –side to the +side.
Then, equations (2.24) and (2.29) must be replaced by equivalent forms
Z Z Z Z
d @/
/ðx, tÞdv ¼ dv þ /v ndS ½½ðv wÞ m/da,
dt @t
vðtÞ vðtÞ @vðtÞ RðtÞ
ð2:30aÞ
Z Z Z
d d/
/ðx, tÞdv ¼ þ /divv dv ½½ðv wÞ m/da, ð2:30bÞ
dt dt
vðtÞ vðt Þ RðtÞ
a+(t)
v (x,t)
m
w(x,t)
v+
(t)
v-
a-(t)
Figure 2.2. Discontinuity across surface RðtÞ:
change of the integral of / over the volumes vþ ðtÞ and v ðtÞare equal to the
integral of @/=@t over the fixed volumes that coincide instantaneously with
vþ ðtÞ and v ðtÞat time t plus the flux /v through the surfaces of the fixed
volumes. It then follows that
Z Z Z
d @/
/ðx, tÞdv ¼ dvþ /v nda
dt @t
þ þ þ
v ðt Þ v ðt Þ a ðt Þ
Z
ðv wÞ m/þ da
RðsÞ
Z Z Z
d @/
/ðx, tÞdv ¼ dvþ /v ndaþ
dt @t
v ðt Þ v ðtÞ a ðt Þ
Z
ððv wÞ m/ Þda,
RðsÞ
Z
d
qðx, tÞdv ¼ 0,
dt
vðtÞ
Z
dq
þ q= v dv ¼ 0:
dt
v ðt Þ
Since this result must hold for an arbitrary material volume and the in-
tegrand is continuous,
dq
þ q= v ¼ 0, ð2:31Þ
dt
which is the spatial form of continuity equation and which can also be
expressed in the form
@q
þ = ðqvÞ ¼ 0,
@t
by using the expression for the material derivative of q and the relation
= ðqvÞ ¼ q= v þ v =q. The material form of the continuity equation is
qO ¼ qJ, ð2:32Þ
Z Z
d du
qudv ¼ q dv ð2:33Þ
dt dt
vðtÞ vðt Þ
@vi
=v¼ ei ej :
@xj
A common notation is
L ¼ = v,
d @xi @vi @vi @xj
¼ ¼ ,
dt @XK @XK @xj @XK
F_ ¼ LF ð2:34Þ
or
_ 1 ,
L ¼ FF ð2:35Þ
2.8 Rate of Deformation and Spin Tensors 65
which gives the relation between the velocity gradient and the deformation
gradient.
and let the velocities of P and Q
Consider a material line element PQ,
be vO and v, respectively. It then follows from a Taylor series expansion
that the velocity of Q relative to P is
v vO ¼ = vðx xO Þ þ O jx xO j2 ,
dv ¼ = vdx, ð2:36Þ
@vi
dvi ¼ dxj :
@xj
We must note the meanings of dx and dv, namely, that dx is the vector of
and dv is the velocity of Q
the infinitesimal directed line segment PQ
relative to P. The material time derivative of dx is dv, since
d d @xi @vi @vi @XK @vi
ðdxi Þ ¼ dXK ¼ dXK ¼ dxj ¼ dxj ¼ dvi :
dt dt @XK @XK @XK @xj @xj
ð2:37Þ
1 T
D¼ LþL ð2:38Þ
2
66 Kinematics and Continuity Equation
and
1
W¼ L LT : ð2:39Þ
2
We use notations dij and wij for the components of D and W, respectively,
so that
1 @vi @vj
dij ¼ þ
2 @xj @xi
and
1 @vi @vj
wij ¼ :
2 @xj @xi
d 2 d d @vi
ds ¼ ðdxi dxi Þ ¼ 2dxi ðdxi Þ ¼ 2 dxj dxi ¼ 2 dij þ xij dxi dxj
dt dt dt @xj
¼ 2dij dxi dxj ,
ð2:40Þ
since
d 2 d
ds ¼ 2ds ðdsÞ,
dt dt
1 d dxi dxj
ðdsÞ ¼ dij ¼ dij ni nj : ð2:42Þ
ds dt ds ds
2.8 Rate of Deformation and Spin Tensors 67
Equation (2.42) gives an insight into the physical significance of the rate of
deformation tensor D, which is often called the strain rate tensor especially
in plasticity theory.
It follows from equation (2.40) or (2.42) that a necessary and sufficient
condition for a rigid body motion, that is, a motion for which the distance
between any two particles remains invariant, is dij ¼ 0. This is known as
Killing’s theorem. It was shown in the previous section that the velocity of
Q relative to P is dv ¼ = vdx for the material line element PQ ¼ dx,
and it may be deduced from Killing’s theorem that the part W dx of
= vdx ¼ Ldx corresponds to a rigid body motion.
In order to see the significance of the spin tensor W , it is desirable to
introduce the vector
1 1 @v3 @v2 @v1 @v3 @v2 @v1
v ¼ curlv ¼ e1 þ e2 þ e3 :
2 2 @x2 @x3 @x3 @x1 @x1 @x2
1
xi ¼ eijk wjk ,wij ¼ eijk xk : ð2:43Þ
2
The significance of the vector x and the tensor W then becomes evi-
dent from the following theorem. The angular velocity of the triad of
material line elements that instantaneously coincides with the principal
axes of the rate of deformation tensor D is given by v ¼ 12 curlv. In order
to prove this theorem we need to show that
d ðaÞ ð Þ
n ¼ x3n a , ð2:44Þ
dt
68 Kinematics and Continuity Equation
1 d ðaÞ
ðaÞ
dnk d dxk 1 d
¼ ¼ ðdxk Þ ds dxk , ð2:45Þ
dt dt dsðaÞ ð Þ
ds a dt ð Þ2
ds a dt
ða Þ ðaÞ
dik d dki ni ¼ 0,
ðaÞ
dnk @vk ðaÞ ð Þ ðaÞ ð Þ ð Þ
¼ n d a nk ¼ wki ni a ¼ ekji xj ni a ¼ ðv 3 nÞk ,
dt @xi i
@vi
vi ¼ rs þ O jr j2 ;
@xs
Z Z
@vk
Hi ¼ eijk qrj rs dv ¼ eijk ðdks þ wks Þ qrj rs dv: ð2:46Þ
@xs
v v
Z
Iij ¼ q r2 dij ri rj dv, ð2:48Þ
v
where Iij are the components of the inertia tensor of the element. Also at
the instant immediately after the element becomes rigid, the rate of de-
formation tensor is zero, so that
Z
Hi ¼ eijk x9ks qrs rj dv, ð2:49Þ
v
70 Kinematics and Continuity Equation
0
where xks are the components of the spin tensor. Equation (2.49) is equiv-
alent to equation (2.47), since substituting equation (2.38)2 in equation
(2.43) gives
Z Z
Hi ¼ eijk estk x9t qrs rj dv ¼ dij djt dit djs x9t qrs rj dv
Z v v
R
Referring to equation (2.46), dks rs rj dv are the components of a second-
v R
order symmetric tensor if and only if drs and qrs rj dv are coaxial, that is, if
v
and only if they have the same principal axes or eigenvectors. Also the
R
principal axes of Iij and qrs rj dv coincide. Comparison of equations (2.46)
v
and (2.49) shows that if and only if dij and Iij are coaxial,
wij ¼ w9ij
since
Z
eijk dks qrs rj dv¼0,
v
R
if and only if drs qrs rj dv is symmetric.
v
T T
U ¼ R VR,V ¼ RUR : ð2:51Þ
3 3
ða Þ ðaÞ ð Þ ð Þ
U ¼ + ka N N , V ¼ + ka n a n a , ð2:52aÞ
a¼1 a¼1
if the ka are distinct, where N ðaÞ and nðaÞ are the eigenvectors of U and V,
respectively, and
ð3Þ ð3 Þ
ð3Þ ð3Þ
U ¼ k3 N N
þk IN N ,
ð2:52bÞ
ð Þ ð Þ ð Þ ð Þ
V ¼ k3 n 3 n 3 þ k I n 3 n 3 ,
if k1 ¼ k2 ¼ k 6¼ k3 , and
ð Þ ðaÞ ðaÞ ð Þ
n a ¼ RN ,N ¼ RT n a :
dx dx ¼ dX dX
72 Kinematics and Continuity Equation
for all material line elements at a point, and using equation (2.20) and the
substitution operator property of the Kronecker Delta we obtain
x1 ¼ X1 þ KX2 , x2 ¼ X2 , x3 ¼ X3 , ð2:53Þ
2 3
1 K 0
½F ¼ 4 0 1 0 5: ð2:54Þ
0 0 1
This is a problem of plane motion where all the particles move parallel to
the 0x1 x2 plane and the motion does not depend on x3 . Consequently, the
rotation tensor R is of the form
2 3
cos h sin h 0
½R ¼ 4 sin h cos h 0 5, ð2:55Þ
0 0 1
2.9 Polar Decomposition of F 73
X2
0
X1
V ¼ FRT
2 3 2 32 3
V11 V12 V13 1 K 0 cos h sin h 0
4 V21 V22 V23 5 ¼ 4 0 1 0 54 sin h cos h 0 5
V31 V32 V33 20 0 1 0 0 1 3
cos h þ K sin h sin h þ K cos h 0
¼4 sin h cos h 0 5:
0 0 1
Consequently,
K
tan h ¼ :
2
1=2
V11 V12
¼ 4þK
2 2 þ K2 K þ 2K
V21 V22 K 2
74 Kinematics and Continuity Equation
and
F ¼ J 1=3 I F : ð2:56Þ
U ¼ J 1=3 I U and C ¼ J 2=3 I C : ð2:57Þ
It may be deduced from equation ð2:57Þ1 that the stretches ka , a 2 f1, 2, 3g,
that are the principal components of U can be similarly decomposed as
1=3
ka ¼ J ka ,
where k1 k2 k3 ¼ 1 and the ka are known as the modified stretches.
2.11 Strain 75
2.11 Strain
The right and left stretch tensors U and V, respectively, which result from
the polar decomposition of the deformation gradient tensor F, may be
expressed in the form
1=2 1=2
U ¼ F TF , V ¼ FF T , ð2:58Þ
so that the squares of U and V are given by the positive definite symmetric
tensors
C ¼ U 2 ¼ F T F and B ¼ V 2 ¼ FF T , ð2:59Þ
Tensors C and B are known as the right and left Cauchy strain tensors,
respectively, and are a measure of the deformation at a material point in
a body since the square of the length of a material line element in the spatial
configuration is given by
1 1
dXK dXK ¼ Bij dxi dxj ¼ dx B dx: ð2:61Þ
dx dx dX dX ¼ dX EdX ð2:62Þ
1 1 T 1
2
E ¼ ðC I Þ ¼ F FI ¼ U I , ð2:63Þ
2 2 2
j ¼ x X: ð2:64Þ
ni ¼ xi XK dKi , ð2:65Þ
where
EK ei ¼ dKi :
It follows from equations ð2:63Þ2 and (2.65) that E may be expressed in the
component form
" #
1 @ n^K @ n^L @ n^J @ n^J
EKL ¼ þ þ , ð2:66Þ
2 @XL @XK @XK @XL
where n^K ¼ dKi ni , and in this form E is sometimes called the Lagrangian
strain.
A further measure of deformation is given by the tensor h defined by
dxi dxi dXK dXK ¼ 2glm dxl dxm ¼ 2dx hdx, ð2:67Þ
2.11 Strain 77
1 1
h¼ I B1 ¼ I F T 1 F , ð2:68Þ
2 2
1 @XK @XK
gij ¼ dij ,
2 @xi @xj
and with the use of equation (2.65) this can be put in the form
1 @ni @nj @ns @ns
gij ¼ þ :
2 @xj @xi @xi @xj
T 1 1
h¼F EF :
E_ ¼ F T DF, ð2:69Þ
@xi @xj
E_ KL ¼ dij :
@XK @XL
78 Kinematics and Continuity Equation
This result is obtained from equations (2.34) and (2.63) as follows, using
symbolic notation,
1 T
1 d T
_
E¼ F FI ¼ F_ F þ FT F_
2 dt 2
1 T T
¼ F L F þ F T LF ¼ F T DF:
2
Eb ¼ U I,
and the logarithmic strain tensors lnU and lnV whose principal components
are lnki .
where
@ n^
=X j^ ¼ K EK EL
@XL
Consequently, if =X j^is, in some sense, small, it is reasonable to make the
approximation E ’ e since
E ¼ e þ O e2 ð2:71Þ
where
e ¼ =X n^,
F ¼ Q ¼ I þ a, jaj << 1
a ¼ aT þ O e2 ,
where e ¼ jaj. Green’s strain tensor vanishes for a rigid displacement of the
spatial configuration so that from equation (2.67) we have, for a rigid body
rotation,
e ¼ O e2 :
R ¼ ðI þ aÞ and U ¼ ðI þ bÞ
noting that a and b are both of OðeÞ. It follows from the orthogonality of R
and the symmetry of U that
a ¼ aT þ O e2 , b ¼ bT :
Then we have
2
=X j^ ¼ F I ¼ a þ b þ O e ,
1 @ui @uj 1 @ui @uj
eij ¼ þ , aij ¼ : ð2:72Þ
2 @xj @xi 2 @xj @xi
The theory for the infinitesimal strain and rotation tensors is analogous to
the theory for the rate of deformation tensor D and the spin tensor W with
the infinitesimal displacement analogous to the velocity.
When the use of infinitesimal strain approximation (2.69) is justified it
is evident that
e_ ij dij :
This is perhaps the reason why D is often called the strain rate tensor.
Exercises 2.1–2.9 81
EXERCISES
q0 @XK
dai ¼ dAK ,
q @xi
q0 1T
da ¼ F dA:
q
(b)
@ 1
J FiK ¼ 0:
@xi
(c)
@UK @ 1
¼J J FiK UK :
@XK @xi
2.3. Let dxð1Þ and dxð2Þ be material line elements emanating from a particle
P. They specify a material surface element given by
ð Þ ð Þ
da ¼ dx 1 3dx 2 :
82 Kinematics and Continuity Equation
Show that
d @vj @vj
ðdai Þ ¼ dai daj :
dt @xj @xi
T
_ Tþ 1 _ 1 RT RUU
_ 1 RT
W ¼ RR RUU
2
and
1 n _ 1 o
D ¼ R UU þ U 1 U_ RT :
2
2.5. If e is the isotropic part of the infinitesimal strain tensor show that
J ¼ 1 þ 3e þ 0 e2 ,
where e ¼ j=X U j.
2.6. Show that the acceleration vector can be expressed in the form
@v 1
a¼ þ ð= vÞv ð= vÞT v þ =v v:
@t 2
@v @ ðqvÞ
q þ v gradv ¼ þ divðqv vÞ:
@t @t
2.8. Prove that, for each particle of a deforming continuum there is at least
one material line element whose direction is instantaneously
stationary.
Exercises 2.1–2.9 83
DF
t ðn, x, tÞ ¼ lim :
Da!0 Da
t ðnÞ ¼ t ðnÞ:
This condition is not satisfied when we have singular surfaces across which
discontinuities of certain field variables can occur, for example, for shock
waves.
84
3.1 Contact Forces and Body Forces 85
m+
P
a
m_
Figure 3.1. Force on area element at point Pð xÞ:
DM
lim ¼ 0,
Da!0 Da
so that contact forces are present but not contact torques. In certain media,
known as polar media, contact torques can exist; however, in this book we
will not be concerned with this aspect of continuum mechanics.
The stress vector at points on the surface @v of a body, with n the
outward unit normal to the surface, is known as the surface traction.
The stress vector can be resolved into a normal component
n o1=2
SðnÞ ¼ jt ðnÞj2 N ðnÞ2 . ð3:2Þ
Z Z Z
d
tda þ qf dv ¼ qmdv,
dt
@v v v
Z Z Z ð3:3Þ
d
ti da þ qfi dv ¼ qvi dv,
dt
@v v v
Z Z
d dv
qmdv ¼ q dv:
dt dt
v v
Z Z Z
d
x 3 tda þ x 3 qf dv ¼ x 3 qvdv,
dt
@v v v
Z Z Z ð3:4Þ
d
eijk xj tk da þ eijk xj qfk dv ¼ eijk xj qvk dv,
dt
@v v v
3.2 Cauchy or True Stress 87
where x is the position vector with respect to the fixed point or the mass
center.
where r11 , r12 , and r13 are the x1 , x2 , and x3 components, respectively, of the
stress vector acting on the plane element; r11 is the normal component, and
r12 and r13 are the shear components. It follows that the resulting shearing
stress on the element is
1=2
Sðe1 Þ ¼ r212 þ r213 :
The matrix rij specifies the state of stress at P. Component rij is defined as
the stress component in the xj direction of the stress vector acting on the plane
element whose normal is in the xi direction. Some texts define this as rji rather
than rij . The signs of the components rij are obtained as follows. If the normal
to a plane element is in the 6 xi direction and the stress vector component is
88 Stress
dv dvi
where v_ ¼ ¼ v_ i ei ¼ ei is the particle acceleration.
dt dt
Substituting
T
ti ¼ rji nj , t ¼ r n: ð3:10Þ
The elements of the matrix rji are the components of a second-order
tensor rT since rji nj are the components of a vector for arbitrary unit vector
x2 σ22
σ21
σ12
σ11 σ11
σ12
x1
σ21
σ22
x3
t
C
n
P
B x2
x1
Figure 3.3. Equilibrium of infinitesimal. Tetrahedron PABC.
n. This follows from the test for tensor character given in chapter 1 since rT
is a tensor so is r. If rij and rij9 are the components of stress referred to the
rectangular Cartesian coordinate systems Oxi and Ox9,
i respectively, it fol-
lows from the tensor transformation rule
where aik are the elements of the transformation matrix ½a. Equations
(3.11) and (3.12) can be expressed in symbolic matrix notation as
½r9 ¼ ½a½r½aT ,
respectively.
Equations (3.1) and (3.10) give the expression
for the normal stress acting on a plane element with unit normal n and the
resultant shearing stress is given by equation (3.2).
90 Stress
r ¼ rij ei ej
Since equation (3.14) must hold for an arbitrary volume and it is assumed
that the integrand is continuous,
@rji dvi
þ qfi ¼ q ,
@xj dt
ð3:15Þ
dv
divrT þ qf ¼ q :
dt
@vi =@t þ vj @vi @xj , is the material derivative of v. For linearized problems
dv=dt is replaced by @v=@t.
Similarly the angular momentum equation with (3.10) substituted in
equation (3.4) is
Z Z
eijk xj rrk nr da þ eijk xj qðfk v_ k Þdv ¼ 0 ð3:16Þ
@v v
3.4 Principal Stresses and Maximum Shearing Stress 91
in Cartesian tensor suffix notation. With the use of the divergence theorem,
the substitution property of the Kronecker delta and equation (3.15), equa-
tion (3.16) becomes
Z
eijk rjk dv ¼ 0. ð3:17Þ
v
Since equation (3.17) must hold for an arbitrary volume and it is assumed
that the integrand is continuous, eijk rjk ¼ 0. It then follows that
rjk ¼ rkj , r ¼ rT ,
which shows that the stress tensor r is symmetric in the absence of body
moments.
and it follows that rðaÞ , a 2 f1, 2, 3g are the roots of the cubic equation
det rij kdij ¼ 0: ð3:20aÞ
k3 I1 k2 þ I2 k I3 ¼ 0,
where
1n o
I1 ¼ rii , I2 ¼ ðrii Þ2 rij rij , I3 ¼ det rij ,
2
1n 2
o
I1 ¼ trr, I2 ¼ ðtrrÞ2 trr , I3 ¼ det½r
2
are independent of the choice of coordinate system and are known as the
first, second, and third basic invariants, respectively, of the stress tensor r.
If the third basic invariant I3 ¼ 0, then the tensor r is singular and one of
the principal stresses is zero.
The components of nðaÞ , a 2 f1, 2, 3g, corresponding to ra , can be
obtained from any two of equations (3.19) and
nðaÞ
¼ 1 when the ra have
been obtained.
When the ra are distinct, the spectral representation of the Cauchy
stress tensor is
3
r= R ra nðaÞ na , a 2 f1, 2, 3g,
a¼1
@
rij ni nj kni ni ¼ 0,
@nk
2
2 2 2 2 2 2 2
S ðnÞ ¼ ðr1 n1 Þ þðr2 n2 Þ þðr3 n3 Þ r1 n1 þ r2 n2 þ r3 n3 : ð3:21Þ
@ 2
S kni ni ¼ 0,
@nk
94 Stress
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi 1
n1 ¼ 0, n2 ¼ 1=2, n3 ¼ 1=2, jSj ¼ jr2 r3 j,
2
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi 1
n1 ¼ 1=2, n2 ¼ 0, n3 ¼ 1=2, jSj ¼ jr3 r1 j, ð3:22Þ
2
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi 1
n1 ¼ 1=2, n2 ¼ 1=2, n3 ¼ 0, jSj ¼ jr1 r2 j:
2
It follows from equation (3.22) that the maximum value of jSj is given by
1
jSjmax ¼ ðrmax rmin Þ, ð3:23Þ
2
where rmax and rmin are the algebraically greatest and least principal
stresses, respectively. The result (3.23) has an important application in
the theory of plasticity since the Tresca yield condition states that the onset
of plastic deformation of a metal occurs when jSjmax reaches a certain
critical value, which is equal to the yield point in simple shear.
where the ei are the unit base vectors of a coordinate system for which
r11 ¼ r22 ¼ r33 ¼ 0. This coordinate system is not unique since the form
(3.24) is valid for an infinite number of coordinate systems. The proof of
3.6 Octahedral Shearing Stress 95
this is given by Pearson [1]. It is clear that rii ¼ trr is a necessary condition
for a state of pure shear stress and the proof of this is also a sufficient
condition is also given by Pearson [1].
We now introduce the decomposition of the stress tensor into its iso-
tropic or hydrostatic part and deviatoric parts
rkk
rij ¼ sij þ dij ,
3
r ¼ s þ trr , ð3:25Þ
I,
3
trr
where I and s are the isotropic and deviatoric parts, respectively, of the
3
stress tensor. It is easily verified that sii ¼ trs ¼ 0 and that the basic invari-
ants of the tensor s are
1
J1 ¼ 0, J2 ¼ sij sij , J3 ¼ det sij ,
2
r ¼ se1 e2 ,
x2
τ
x1
τ
1
n ¼ pffiffiffi ð6e1 6e2 6e3 Þ: ð3:26Þ
3
1 I1
Noct ¼ ðr1 þ r2 þ r3 Þ ¼ ,
3 3
that is, the isotropic part of the stress tensor, acts on each octahedral plane.
The shearing stress jSoct j acting on a plane element with a normal given by
equation (3.26) is known as an octahedral shearing stress, and it follows from
equations (3.21) and (3.26) that it is the same for all eight octahedral planes
and is given by
1 2 1
S2oct ¼ r1 þ r22 þ r23 ðr1 þ r2 þ r3 Þ2 , ð3:27Þ
3 9
2 1n o
Soct ¼ ðr1 r2 Þ2 þðr2 r3 Þ2 þðr3 r1 Þ2 : ð3:28Þ
9
3.6 Octahedral Shearing Stress 97
e3
e1 e2
1 2
S2oct ¼ s1 þ s22 þ s23 ,
3
1 2
S2oct ¼ sij sij ¼ J2 : ð3:29Þ
3 3
2
sij sij ¼ 2k ,
where k is the yield stress in simple shear. The reader should show that
pffiffiffi
Y ¼ 3k, where Y is the yield stress in simple tension.
The Tresca yield criterion states that yielding of a metal occurs when the
maximum shearing stress reaches a critical value, that is, when
where rmax and rmin are the algebraically greatest and least principal
stresses.
T
Ti ¼ RKi NK , T ¼R N: ð3:30Þ
that
q0 @XK
ni da ¼ NK dA,
q @xi
where q0 and q are the densities in the reference and spatial configurations,
respectively, gives
q0 @XK q0 1
RKi ¼ rji , R¼ F r: ð3:31Þ
q @xj q
The nominal stress tensor may be expressed in terms of its Cartesian com-
ponents as
R ¼ RKi EK ei :
The form of equation (3.3) in terms of integrals over the reference volume
V and surface @V is
Z Z Z
@v0i ðXK , tÞ
Ti dA þ q0 f0i dV ¼ q0 dV,
@t
@V V V
Z Z Z ð3:32Þ
@vo ðX, tÞ
TdA þ q0 f 0 dV ¼ q0 dV,
@t
@V V V
noting that equation (3.32) applies to any part of the body gives the equa-
tion of motion:
where the notation Div is used to denote the divergence operation based on
the operator,
@ð Þ
=X [ EK ,
@XK
S ¼ SKL EK EL ,
@XK @XL T
SKL ¼ J rij , S ¼ JF 1 rF 1 ð3:34Þ
@xi @xj
@XK T
SKL ¼ RLi , S ¼ F 1 RT ¼ RF 1 : ð3:35Þ
@xi
The tensor S unlike the tensor R is symmetric, and this may be advanta-
geous in some applications, but it cannot be directly related to the stress
vector acting on a plane element. It is a particular case of the class of
Lagrangian tensors.
3.9 Other Stress Tensors 101
The following stress tensor is obtained from the nominal stress tensor as
follows:
^
S ¼ RR ¼ RKi RiL EK EL , ð3:36Þ
1 1
SB ¼ RR þ RT RT , SB
KL ¼ ðRKi RiL þ RiK RLi Þ ð3:37Þ
2 2
is known as the Biot stress tensor. Equation (3.36) gives rise to the polar
^ T , which is analogous to the polar decomposition of
decomposition R ¼SR
the deformation gradient tensor F. However, ^
S is not unique since the
deformation due to a nominal stress R may not be unique [3]. Using equa-
tion (3.35) and the polar decomposition F ¼ RU, equation (3.37) can be
put in the form
1 1
S B ¼ ðSU þ US Þ, SB
KL ¼ ðSKL ULJ þ UKL SLJ Þ: ð3:38Þ
2 2
The Biot stress tensor is useful when certain problems of isotropic elasticity
are considered. For isotropic elasticity the tensors S and U are coaxial so
that the product is commutative. It then follows that
S B ¼ US, ð3:39Þ
pffiffiffiffiffiffiffiffiffiffi
SB ¼ RRT : ð3:40Þ
102 Stress
It should be noted that equations (3.39) and (3.40) are valid only
for isotropic elastic media. It may be shown that RRT is positive definite
or positive semidefinite if R is nonsingular or singular, respectively. If
RRT is positive definite its principal values zð1Þ , zð2Þ , and zð3Þ are positive.
If RRT is positive semidefinite, its principal values are non-negative.
pffiffiffiffiffiffiffiffiffiffi
The principal values of S B ¼ RRT are the positive or negative
square roots of zð1Þ , zð2Þ , and zð3Þ . It is evident that a given R, does not,
in general, determine a unique S B , and this is discussed in detail by
Ogden [3].
It is easily shown that det½R ¼ det SB and that RKi diL ¼ SKL if R ¼ I.
EXERCISES
3.1. Given the state of true stress with the following components
2 3
3 1 3
4 1 2 2 5
3 2 5
Find
(a) the principal Cauchy stresses,
(b) the octahedral shearing stress,
(c) the direction cosines of the normals to the octahedral planes,
(d) the normal stress, shearing stress, and direction of the shearing
stress acting on the plane with unit normal
1 1 1
n¼ e1 þ e2 þ pffiffiffi e3 :
2 2 2
3.2. Show that if the Cauchy stress tensor is singular, that is, det½rij ¼ 0,
one of the principal stresses is zero.
3.3. Prove that if the Cauchy stress tensor r and left Cauchy-Green strain
tensor B are coaxial, the second Piola-Kirchhoff stress tensor S and
the right Cauchy-Green strain tensor C are also coaxial.
Exercises 3.1–3.10 103
x1 ¼ X1 þ KX2 , x2 ¼ X2 , x3 ¼ X3 :
Z
1
rm ¼ t Xda:
3V
@V
3.6. Let m and n be unit vectors normal to surface elements Dam and Dan
at a point, and let t ðmÞ and t ðnÞ be the stress vectors acting on these
plane elements. Show that
n tðmÞ ¼ m t ðnÞ:
3.7. Consider a state of plane stress with r33 ¼ r32 ¼ r31 ¼ 0. Let
be the normal stress on plane elements that are normal to the x3 plane
and form angles of 120° with each other. Find the principal stress and
the matrix of components of the stress tensor if the normal stress N(1)
is in the x1 direction and the direction of Nð2Þ is in the second quadrant
of the 0x1 x2 plane.
104 Stress
where r^ij and rij are the components of the Cauchy stress with respect
xi and Oxi , respectively.
to O^
3.9. A body of current volume V is in equilibrium under the action of
surface traction rn and body force f. Show that the mean stress
R
r ¼ V1 rdV is given by
V
8 9
Z Z
1< T =
r ¼ r n x da þ f xdv :
v: ;
@v v
SC_ ¼ 2RF:
_
REFERENCES
This chapter is concerned with some aspects of work, energy, and contiuum
thermodynamics. The treatment is general and reference is not made to any
particular material. Notation for deformation and stress variables is the
same as in chapters 2 and 3, where the lower- and uppercase suffixes refer
to the spatial and reference configurations, respectively. Most of the equa-
tions in this chapter are given in both suffix and symbolic notation.
Stress power is the rate of work done by the stresses and is a purely mech-
anical quantity, that is, obtained from the rate of work of the body forces
and surface tractions acting on a material body and does not involve any
other forms of energy transfer such as heating. It is a specific quantity per
unit volume or per unit mass and can be expressed in terms of either the
Cauchy, nominal, or second Piola-Kirchhoff stress and the respective con-
jugate rate variable.
The rate of work of the body forces and surface tractions acting on
a material body of volume v and surface a in the spatial configuration is
given by
Z Z
W_ ¼ ti vi da þ qfi vi dv,
@v v
Z Z ð4:1Þ
W_ ¼ t vda þ qf vdv,
@v v
105
106 Work, Energy, and Entropy Considerations
where r is the Cauchy stress tensor and n is the outward unit normal to the
surface, in equation (4.1) and applying the divergence theorem gives
Z
@rij
W_ ¼ vi þ qfi þ rij Lij dv
@xj
Zv ð4:3Þ
W_ ¼ fv ½divs þ qf þ trðsLÞgdv
v
where Lij ¼ @vi @xj are the components of the velocity gradient tensor
= v. With the use of the equation of motion
@rij dvi
þ qfi ¼ q ,
@xj dt
ð4:4Þ
dv
divs þ qf ¼ q ,
dt
Z Z
d vi vi
W_ ¼ q dv þ rij Lij dv,
dt 2
v v
Z Z ð4:5Þ
d vv
W_ ¼ q dv þ trðsLÞdv:
dt 2
v v
The first term on the right-hand side of equation (4.5) is the rate of change
of the kinetic energy of the body, and the second term is integral over the
volume of the specific stress power per unit volume of the spatial configu-
ration of the body. It follows that the specific stress power per unit mass is
1 1
P ¼ rij Lij ¼ trðsLÞ: ð4:6Þ
q q
4.1 Stress Power 107
1 1
P ¼ rij Dij ¼ trðsDÞ, ð4:8Þ
q q
Z Z
W_ ¼ T i m0i dA þ q0 f0i m0i dV,
@V V
Z Z ð4:9Þ
W_ ¼ T m 0 dA þ q0 f 0 m 0 dV,
@V V
where
and
T i ¼ RKi NK , T ¼ RT N ð4:10Þ
108 Work, Energy, and Entropy Considerations
Z
@RKi
W_ ¼ vi _
þ q0 f0i þ RKi FiK dV,
@XK
ZV n n o ð4:11Þ
T o
W_ ¼ v: DivR þ q0 f 0 þ tr RF_ dV,
V
where FiK ¼ @xi =@XK , are the components of the deformation gradient
tensor and F_iK ¼ Lij FjK , F_ ¼ LF. With the use of the equations of motion
@RKi @v0i
þ q0 f0i ¼ q0 ,
@XK @t
@m 0
DivRT þ q0 f 0 ¼ q0 , ð4:12Þ
@t
Z Z
@ v0i v0i
W_ ¼ q0 dV þ RKi F_iK dV,
@t 2
V V
Z Z ð4:13Þ
@ v0 v0
W_ ¼ q0 dV þ tr RF_ dV:
@t 2
V V
It follows from equation (4.13) that the specific stress power (per unit mass)
is given by
1 1
P¼ _
ðRKi F_iK Þ ¼ trðRFÞ: ð4:14Þ
q0 q0
4.1 Stress Power 109
_
The stress power in terms of the second Piola-Kirchhoff stress S and E,
the Lagrangian strain rate, cannot be determined directly in the same
manner as equations (4.8) and (4.14) but can be determined from equation
(4.8) or (4.14). Substitution of
q @xi @xj q T
rij ¼ SKL ; s¼ FSF ,
q0 @XK @XL q0
@XM @XN
dij ¼ E_ MN ; D ¼ F T EF
_ 1 ,
@xi @xj
1 1 1 _
P¼ SKL E_ KL ¼ tr S E_ ¼ tr S C : ð4:15Þ
q0 q0 2q0
1 T _
P¼ tr SF F : ð4:16Þ
q0
.
E_ KL ¼ ðF_iK FiL þ FiK F_iL Þ 2, E_ ¼ ðF_ T FþF T FÞ
_ 2: ð4:17Þ
tr SF_ T F ¼ tr S F_ T F þ F T F_ ¼ tr S E_ :
2
110 Work, Energy, and Entropy Considerations
The principle of virtual work applied to static problems involves the spatial
form of the equations of equilibrium,
@rij
þ qfi ¼ 0,
dxj ð4:18Þ
divs þ qf ¼ 0,
@RKi
þ q0 f0i ¼ 0,
@XK ð4:19Þ
DivRT þ q0 f0 ¼ 0,
where du ¼ 0 on the part @vu of the surface. With the use of equation (3.10)
and the divergence theorem (4.20) becomes
Z
@ rij duj
dW ¼ þ qfi dui dv
@xi
v
Z
@rij @duj
¼ duj þ rij þ qfi dui dv: ð4:21Þ
@xi @xi
v
Then, with the use of the equation of equilibrium (4.18), (4.20), (4.21), and
rij @dui @xj ¼ rij deij ;
Z Z Z
ti dui da þ qfi dui dv ¼ rij deii dv,
@vs v v
Z Z Z ð4:22Þ
t dui da þ qf dudv ¼ trðsdeÞdv,
@vs v v
112 Work, Energy, and Entropy Considerations
where
1 @dui @duj 1
de ¼ d ð= uÞ þ ð= uÞT , ð4:24Þ
2
Z Z Z
_ _
Ti dui dA þ q0 f0i dui dV ¼ RKi dFiK dV,
@VT V V
Z Z Z ð4:25Þ
_ _
T dudA þ q0 f 0 dudV ¼ trðRdF ÞdV,
@VT V V
forces, is equal to virtual work of the internal forces, that is, the virtual work
of the internal stresses.
Another form of the principle of virtual work for static or quasi-
static processes arises when we consider a statically admissible stress field
rðxÞ along with a kinematically admissible velocity field vðxÞ in a region v
with surface @v of a material body. A kinematically admissible velocity
field satisfies the velocity boundary conditions on the part of the bound-
ary upon which the velocity is prescribed and any kinematical constraints
such as incompressibility. We then obtain the result for quasi-static
processes
Z Z Z
ti vi da þ qvi fi dv ¼ rij dij dv,
dv v v
Zt Z Z ð4:26Þ
t:vda þ qv f dv ¼ trðsDÞdv,
dvt v v
Z Z Z Z
ti dui da þ qfi dui dv qv_ i dui dv ¼ rij deij dv,
@vt v v v
Z Z Z Z ð4:27Þ
t duda þ qf dudv qv_ dudv ¼ trðsdeÞdv,
@vt v v v
114 Work, Energy, and Entropy Considerations
where n is the outward unit normal to the surface, u ¼ uðx, tÞ is the specific
internal energy, r is rate of heat supply per unit mass, and q is the heat flux
vector. Fourier’s law for the spatial configuration,
@H
qi ¼ kij , q ¼ k=H, ð4:29Þ
@xj
this text. Modifications of equation (4.29) in order to remove the defect have
been extensively researched and [5] is a useful reference on this topic.
Heat production can arise from various external agencies such as
radiation, or from Joule heating in electrical conductors. Potential energy
is neglected in equation (4.28). Substituting for ti , from equation (4.2), in
equation (4.28), and using the divergence theorem and the equation of
motion (4.4) gives
Z
@qi
qu_ qr rij dij þ dv ¼ 0,
@xi
v
Z ð4:30Þ
ðqu_ qr trðsDÞ þ divqÞdv ¼ 0:
v
@qi
qu_ qr rij dij þ ¼ 0,
@xi ð4:31Þ
qu_ qr trðsDÞ þ divq ¼ 0:
Equation (4.31) is the local form of the energy equation referred to the spatial
configuration. The local form of the energy equation can also be referred to
the reference configuration. In order to obtain this form we note that
where Q is the heat flux vector referred to the reference configuration and
dA and da are the vector areas of a plane element in the reference and
spatial configurations, respectively. Using equation (4.32) along with
@XK T
dai ¼ J dAK , da ¼ JF dA,
@xi
gives
1 1
qi ¼ J FiK QK , q ¼ J FQ: ð4:33Þ
4.3 Energy Equation and Entropy Inequality 117
@ 1 1
ðJ FiK Þ ¼ 0, divðJ FÞ ¼ 0,
@xi
give
@qi @QK
¼ J 1 , divq ¼ J 1 DivQ, ð4:34Þ
@xi @XK
where, as before,
q0
J¼ ¼ det½F:
q
It then follows from equations (4.8), (4.14), (4.26), and (4.29) that the local
form of the energy equation (4.31), referred to the reference configuration, is
@QK
q0 u_ 0 q0 r0 RKi F_iK þ¼ 0,
@XK ð4:35Þ
_ þ DivQ ¼ 0,
q0 u_ 0 q0 r0 trðRFÞ
where u0 ðX, tÞ ¼ uðxðX, tÞ, tÞ and r0 ðX, tÞ ¼ rðxðX, tÞ, tÞ. Equation (4.35)
can also be obtained from the reference form of equation (4.28),
Z Z Z
@ v0i v0i
q0 ðu0 þ ÞdV ¼ q0 ðr0 þ foi v0i ÞdV þ ðTi v0i Qk NK ÞdA:
@t 2
V V @V
Z Z Z
d qr qi ni
qsdv dv þ da > 0,
dt H H
v v @v
Z Z Z ð4:36Þ
d qr qn
qsdv dv þ da > 0,
dt H H
v v @v
118 Work, Energy, and Entropy Considerations
Z Z
d
qsdv ¼ _ > 0,
qsdv
dt
v v
@qi qi @H
qHs_ qr þ > 0,
@xi H @xi
q
qHs_ qr þ divq :=H > 0,
H
qi @H
_ þ rij dij
qðHs_ uÞ > 0,
H @xi ð4:37Þ
q
qðHs_ uÞ
_ þ trðsDÞ :=H > 0:
H
_
_ __
__ @vi QK @H
q0 ðH s uÞ þ RKi _ > 0,
@XK H @XK ð4:38Þ
_ __ _ Q =X H > 0,
__ _
q0 ðH s uÞ þ trðRFÞ _
H
_
where s ðX, tÞ ¼ sðxðX, tÞ, tÞ, HðX, tÞ ¼ HðxðX, tÞ, tÞ. In equation (4.38) RF_
_
_
can be replaced by SE_ or S C=2.
Exercises 4.1–4.8 119
qi @H
_ þ f_Þ þ rij dij
qðsH > 0,
H @xi ð4:39Þ
_ þ f_Þ þ trðrDÞ q :=H > 0:
qðsH
H
The equality signs in equations (4.37) to (4.39) hold only for an ideal thermo-
dynamically reversible process. The left-hand sides of equations (4.37) and
(4.38) are equal to q0 Hc_ and qHc,
_ respectively, where c>0
_ is the rate of
entropy production per unit mass. It is interesting to note that the rate of
entropy production per unit mass due to heat conduction alone is given by
qi @H q
c_ ðhcÞ ¼ ¼ :=H, ð4:40Þ
qH2 @xi qH2
EXERCISES
4.1. Obtain the spatial form of the stress power in terms of the deviatoric
and isotropic parts of the Cauchy stress tensor s and the rate of de-
formation tensor D.
4.2. Show that for an isotropic linear elastic solid the specific stress power is
equal to the rate of change of specific strain energy.
4.3. Obtain the following virtual relations,
1 1
dF ¼ ð= duÞF and dF ¼ F ð= duÞ:
120 Work, Energy, and Entropy Considerations
@duk @duk
deij ¼ dgij þ gkj þ gik ,
@xi @xj
T
de ¼ dg þ ð= duÞ h þ hð= duÞ:
4.5. Show that the virtual variation of the Lagrangian strain tensor is given
by
1n T o
dE ¼ F ð=X duÞ þ ð=X duÞF :
2
qðHs_ uÞ
_ þ trðsDÞ > 0
REFERENCES
5.1 Introduction
The response of material bodies to applied forces and heating has not been
considered in previous chapters, except that it is assumed that the materials are
nonpolar. In order to obtain relations between stress, deformation, and tem-
perature fields, it is necessary to have constitutive equations. These equations
give the response of idealized models of real materials, to the application of
forces and heating. In this book electromagnetic effects are not considered.
Admissible constitutive equations must not violate the laws of thermo-
dynamics and should satisfy certain additional principles. In this chapter some
classical constitutive equations and other relations for models of actual mate-
rials are considered, before discussing these additional principles in detail.
As in previous chapters, certain relations are presented in both suffix
and symbolic notation; however, relations that involve tensors of higher
order than two are given in suffix notation only.
T
FiK ¼ QiK , det½F ¼ 1, C ¼ F F ¼ I and D ¼ 0:
121
122 Material Models and Constitutive Equations
A rigid body may be defined as a body for which the distance between any
two particles of the body remains invariant, and the motion given by equa-
tion (5.1) satisfies this definition. To show this, take any two particles of the
body with position vectors X 1 and X 2 and corresponding position vectors
x1 and x2 after a rigid body motion that involves translation and rotation. It
then follows from equation (5.1) that
ðx1 x2 Þ ðx1 x2 Þ ¼ QðX 1 X 2 Þ QðX 1 X 2 Þ
¼ ðX 1 X 2 Þ ðX 1 X 2 Þ,
@qi
qu_ qr þ ¼ 0,
@xi ð5:4Þ
qu_ qr þ divq ¼ 0,
5.3 Ideal Inviscid Fluid 123
where q is the heat flux vector and r is rate of heat supply per unit mass, for
example, Joule heating due the passage of an electric current. A semi-
empirical law, Fourier’s Law of heat conduction, given by
@H
qi ¼ kij ðHÞ , q ¼ kðHÞgradH ð5:5Þ
@xj
@H @ @H
qcðHÞ qr ðkij Þ ¼ 0,
@t @xi @xj
ð5:6Þ
@H
qcðHÞ qr divðk gradHÞ ¼ 0:
@t
For many problems the temperature range is small enough that the tem-
perature dependence of k and c can be neglected. If in addition the rigid
body is isotropic, k ¼ Ik, and equation (5.6) becomes
@H
qc qr k=2 H ¼ 0,
@t
where
.
=2 [ divgrad [ @ 2 @xi @xi :
p
¼ RH, ð5:8Þ
q
where R is the gas constant for the particular gas, H is the absolute tem-
perature, p is the absolute pressure, and q is the density. Since the internal
energy of a perfect gas can be expressed as a function of temperature only,
equation (5.3) also holds for a perfect gas if cðHÞ is replaced by cv ðHÞ, the
specific heat at constant volume. The stress power of a compressible in-
viscid fluid is P ¼ pv_ where v ¼1=q.
If we consider a viscous fluid, equation (5.7) holds only for equilibrium
states and is not a constitutive relation.
This is an inviscid fluid model that can only undergo isochoric flow. The
constitutive equation chosen to model a given material may be influenced
by the problem considered. For example, for some problems involving
a liquid, such as gravity water waves, it may be realistic to assume an in-
compressible inviscid fluid, whereas for other problems, such as water
hammer in a pipe, compressibility effects must be considered, and for
boundary layer problems, viscous effects are dominant.
If a fluid is assumed to be mechanically incompressible the coefficient
of volume thermal expansion must be zero so that only isochoric deforma-
tion is possible. It has been shown by Müller [1] that a fluid model that is
mechanically incompressible cannot have a nonzero coefficient of volume
thermal expansion, otherwise the second law of thermodynamics is
violated. Since an incompressible fluid can undergo only isochoric
deformation,
dii ¼ 0, trD ¼ 0:
5.5 Newtonian Viscous Fluid 125
The stress power for an incompressible inviscid fluid is zero and equation
(5.3) holds.
where g and 1 are viscosity coefficients. For isochoric flow it follows from
the constitutive equation (5.9) that
rij þ pdij ¼ 2gdij , dii ¼ 0,
which indicates that g is the viscosity component for simple shear. For pure
dilatation it follows from equation (5.9) that
rkk
þ p ¼ ndkk , ð5:10Þ
3
@rij dvi
þ qfi ¼ q ,
@xj dt
126 Material Models and Constitutive Equations
2
@p @ 2 vi 1 @ vj dvi
þg þ nþ g þ qfi ¼ q : ð5:11Þ
@xi @xj @xj 3 @xi @xj dt
An elastic solid is a material for which the stress depends only on the
deformation, when the deformation is isothermal. Classical elasticity is
a linear elastic theory based on the assumption that the displacement gra-
dient tensor is infinitesimal so that the strain and rotation tensors are
linearized and the distinction between the spatial and referential forms
5.6 Classical Elasticity 127
eij ¼ ð1=2Þ @ui @xj þ @uj @xi :
Since rij and eij are the components of second-order symmetric tensors,
where the cijkl are the components of a fourth-order tensor known as the
stiffness tensor, with symmetries.
since rij ¼ rji and ekl ¼ elk . Equation (5.13) is sometimes known as the
generalized Hooke’s law. It can be shown that there is a further symmetry,
cijkl ¼ cklij if a strain energy function,
1
wðeÞ ¼ cijkl eij ekl , ð5:14Þ
2
@wðeÞ
rij ¼ : ð5:15Þ
@eij
It then follows that a classical linear elastic solid can have at most twenty-
one independent elastic constants. It should be noted that, in equation
128 Material Models and Constitutive Equations
(5.14), the symmetry of the infinitesimal strain tensor e is ignored for the
differentiation (5.15). Sokolnikov [5] has given more extensive discussion
of the symmetries of the stiffness tensor. Equation (5.13) can be inverted to
give
eij ¼ aijkl rkl , ð5:16Þ
where aijkl are the components of a fourth-order tensor known as the com-
pliance tensor, which has the same symmetries as the stiffness tensor. The
strain energy function in terms of the stress components, sometimes known
as the complementary energy per unit volume, is
1
wc ¼ aijkl rij rkl ,
2
and equation (5.16) can be obtained from
@wc ðrÞ
eij ¼ :
@rij
where a, b, and c are scalars. It then follows that for an isotropic linear
elastic solid, equation (5.13) is of the form,
lð3k þ 2lÞ 2 k
E¼ , K ¼ k þ l, and m¼ ,
kþl 3 2ðk þ lÞ
1 kdij
eij ¼ rij þ rkk : ð5:19Þ
2l 2lð3k þ 2lÞ
where ½r and ½e are ð631Þ column matrices related to the Cauchy stress
and linear infinitesimal strain tensors by
2 3
C11 ¼ c1111 C12 ¼ c1122
C13 ¼ c1133 C14 ¼ c1123 C15 ¼ c1113 C16 ¼ c1112
6 C22 ¼ c2222
C23 ¼ c2233 C24 ¼ c2223 C25 ¼ c2231 C26 ¼ c2212 7
6 7
6 C33 ¼ c3333 C34 ¼ c3323 C35 ¼ c3331 C36 ¼ c3312 7
6 c2323 þc2332 7
6 7
½C¼ 6 Symmetric C44 ¼ C45 ¼ c2331 C46 ¼ c2312 7:
6 2 7
6 c3131 þc3113 7
6 C55 ¼ C56 ¼ c3112 7
6 2 7
4 c1212 þc1221 5
C66 ¼
2
ð5:21Þ
130 Material Models and Constitutive Equations
The matrix ½C is known as the Voigt matrix. It follows from equation (5.21)
that for an isotropic solid,
2 3
2l þ k k k 0 0 0
6 2l þ k k 0 0 07
6 7
6 2l þ k 0 0 07
½C ¼ 6
6
7:
6 Symmetric l 0 077
4 l 05
l
6
rp ¼ + Cpq eq , p 2 f1; 2: . . . 6g;
q¼1
½e ¼ ½ A½r,
av
etij ¼ ðH H0 Þdij , ð5:22Þ
3
and
_
f ¼ f eij9, e, H , ð5:26Þ
_
respectively, where uð0, 0, s0 Þ ¼ 0 and f ð0, 0, H0 Þ ¼ 0. The strain tenor e is
replaced by e9 þ ðe=3ÞI in equations (5.25) and (5.26) in order to indicate
the decoupling of the distortional and dilatational effects.
A Gibbs relation for a linear isotropic thermoelastic solid is
_ _ _
@u @u @u
du ¼ deij9 þ de þ ds
@eij9 @e @s
that
_ _ _
@u @u @u
sij ¼ q , r¼q , H¼q : ð5:28Þ
@eij9 @e @s
Equation (5.28) indicates that the fundamental equation of state for the
_
internal energy is of the form u ¼ u eij9, e, s .
A Gibbs relation
@f @f @f
df ¼ deij þ de þ dH
@eij9 @e @H
5.7 Linear Thermoelasticity 133
that
_ _ _
@f @f @f
sij ¼ q , r¼q , s¼ : ð5:30Þ
@eij9 @e @H
Equations (5.30) indicate that the fundamental equation of state for the
_
Helmholtz free energy is of the form f ¼ f eij9, e, s .
Substituting equation (5.24) in equation (5.29) and integrating gives
_ 1 1 2
f eij9, e, H ¼ leij9eij9 þ Ke Kav eðH H0 Þ þ F ðHÞ, ð5:31Þ
q 2
Kav e
s¼ F9ðHÞ: ð5:32Þ
q
Since ðHdsÞe¼0 ¼ ce dH, where ce is the specific heat at constant strain e, and
se¼0 ¼ F 0 ðhÞ,
ZH _
dH h
F9ðHÞ ¼ ce _ ¼ ln : ð5:33Þ
H h0
H0
H H0
F9ðHÞ ¼ ce , ð5:34Þ
H0
Kav e H H0
s¼ þ ce : ð5:35Þ
q H0
134 Material Models and Constitutive Equations
_ 1 1 2 1 ðH H0 Þ2
f eij9, e, H ¼ leij9 eij9 þ Ke Kav eðH H0 Þ ce :
q 2 2 H0
ð5:36Þ
1
wðe9, eÞ ¼ qf ðe9, e, H0 Þ ¼ leij9 eij9 þ Ke2 ,
2
Kav eH0
H H0 ¼
qcv
!
a2v K 2 H0
r¼ Kþ e:
qce
a2v K 2 H0
Ka ¼ K þ ,
qce
5.7 Linear Thermoelasticity 135
and the isothermal and isentropic shear moduli are equal. It can be deduced
from the Gibbs relation (5.27) that the following expression for the specific
internal energy
1 1 2
u^ eij9, e, H ¼ leij9 eij9 þ Ke þ Kav eH0
q 2
ð5:37Þ
1 ðH H0 Þ2
þ ce þ ce ðH H0 Þ
2 H0
@ u^
ce ¼ :
@H H¼H0
1 1 2
u~ eij9, e, s ¼ leij9 eij9 þ Ke þ Kav eH0 þ
q 2
( )
s Kav eHo 1 s Kav eHo 2
ce H0 þ H0 ð5:38Þ
ce qce 2H0 ce qce
eij9 dsij þ edr
dh ¼ Hds : ð5:39Þ
q
136 Material Models and Constitutive Equations
@h @h @h
dh ¼ dsij þ dr þ ds
@sij @r @s
and
@h @h @h
eij9 ¼ q , e¼q , H¼ : ð5:40Þ
@sij @r @s
Equation (5.40) indicates that the fundamental equation of state for the
enthalpy is of the form h ¼ h sij , r, s . Before determining the fundamental
form h sij , r, s it useful to consider the form
^ sij sij 1 K r 2
h sij ,r,H ¼ þ þav ðHH0 Þ þ
2ql 2 q K
Kav H0 r r 1 ðHH0 Þ2
þav ðHH0 Þ þce ðHH0 Þþ ce , ð5:41Þ
q q K 2 H0
^ !
@h
cr ¼ :
@H
H¼H0
Ka2v H0
cr ¼ ce þ :
q
, !
av r Ka2v ce
H H0 ¼ s þ ,
q q h0
5.7 Linear Thermoelasticity 137
@g @g @g
dg ¼ dsij þ dr þ dH
@sij @r @H
that
@g @g @g
eij9 ¼ q , e ¼ q , s¼ : ð5:43Þ
@sij @r @H
sij sij r2
ws sij , r ¼ , ð5:45Þ
4l 2K
138 Material Models and Constitutive Equations
which is equal to the negative of wðe9, eÞ, the strain energy per unit volume
for isothermal deformation. Often the complementary energy per unit
volume is described as the strain energy in terms of stress with the negative
signs in equation (5.45) replaced by positive signs. However, the form
(5.45) with negative.signs is obtained from the Legendre transformation,
g ¼ f sij eij9 þ re q.
Constitutive relations for nonlinear elastic solids are considered in
later chapters.
It has already been noted that constitutive equations should not violate the
laws of thermodynamics. In addition they should satisfy the principles of
determinism, local action, and frame indifference. These principles are
sometimes called axioms, and it is possible to formulate more than three.
For example, Eringen [6] lists eight axioms of constitutive theory; however,
the three principles discussed here appear to be sufficient to develop a con-
stitutive theory.
The principle of determinism states that the stress and specific internal
energy at a material point at time t are determined uniquely by the past
history of the motion and temperature field up to and including time t. The
heat flux q is also so determined; however, this is usually of less importance
in our considerations than the stress. It is reasonable to assume that the
history of the motion and temperature fields in the recent past has more
influence on the state of the medium than that of the distant past, that is,
the material has a fading memory. A perfectly elastic solid has a perfect
memory of an undeformed reference configuration and an inviscid fluid or
a Newtonian viscous fluid has no memory of a reference configuration.
The principle of local action states that the stress, specific internal en-
ergy, and heat flux at a material point X are independent of the history of
motion and temperature outside an arbitrary small neighborhood of X.
Materials that obey this principle are known as simple materials. In a simple
material, the stress at time t at a material point depends only on the history of
the deformation gradient and temperature. This implies that all relations
5.8 Determinism, Local Action, and Material Frame Indifference 139
between stress and past history of motion can be obtained from experiments
that involve homogeneous deformation and temperature. These relations
are constitutive equations. In this book we consider only simple materials,
since all the usual material models considered in civil and mechanical engi-
neering refer to simple materials.
Before discussing the principle of frame indifference it is necessary to
introduce the concepts of a frame of reference and an observer that are
often regarded as synonymous. A frame of reference may be regarded as
a hypothetical rigid body to which an infinite number of coordinate systems
may be attached. Right-handed, rectangular Cartesian coordinate systems
are considered for the following discussion.
The concept of a frame of reference involves an observer that can
measure time and the relative position of a point in E3 . Two different
observers are said to be equivalent if they observe the same distances
between pairs of points in E3 and the same interval of time between events.
Transformation relations between coordinate systems attached to the
same frame of reference are time independent. However, transformation
relations between coordinate systems attached to different frames of ref-
erence are, in general, time dependent unless the frames are moving in
relative translation. Equations or physical quantities, which are invariant
under coordinate transformation in the same frame of reference, must be
tensors or tensor equations; however, this does not necessarily ensure
invariance for transformations between coordinate systems attached to
different frames of reference. Physical quantities are said to be objective
if observers on different frames, in relative motion, observe the same phys-
ical quantities, that is, if the quantities are invariant under all changes of
frame. Temperature is an example of an objective scalar, velocity is an
example of a vector that is, intuitively, not objective. Applied to constitu-
tive equations the principle of material frame indifference, sometimes
known as the principle of material objectivity, requires that equivalent
observers observe the same material properties or constitutive equations.
An alternative statement is, constitutive equations are frame indifferent if
they are invariant under all changes of frame. A simple example is the
extension of an elastic spring due to an axial force. Equivalent observers
observe the same extension of the spring and, since the axial force is related
140 Material Models and Constitutive Equations
to the extension, the various observers observe the same force. We can
assert that contact forces are objective; however, body forces may not be.
For example, an observer attached to a body in free fall, under the action of
the earth’s gravitational field, observes no body force since no acceleration
is observed. However, an observer fixed in space observes a body force.
A tensor is said to be objective if equivalent observers on different frames
of reference observe the same tensor, so that the tensor is invariant under
change of frame. The precise meaning of this statement is discussed later in
this chapter.
Consider the ordered pair fx, tg, where x is the position vector of
a point in E3 . A change of frame or observer is a one-to-one mapping of
E3 3 t onto itself so that the same distance between any two points in E3 is
observed and time intervals between events are preserved. Let x and t be
the position vector, of a point P in E3 , relative to an origin O, and time,
respectively, in frame of reference F, and let x and t ¼ t a, where a is
a constant, be the position vector of P relative to an origin O , and time,
respectively, in frame of reference F . It should be noted that x is observed
by observer F and x by observer F . If F is moving in translation, relative
to F , we have the trivial relation
x ¼ cðt Þ þ x,
x2 x1 ¼ Qðx2 x1 Þ: ð5:47Þ
x x
¼ jx2 x1 j,
2 1
for all F and F , as does the directed line segment ðx2 x1 Þ, so that different
observers observe the same vector. In order to clarify this, let fe1 , e2 , e3 g be
the unit base vectors for coordinate system 0xi , and e1 , e2 , e3 the unit
base vectors for 0 xi , given by
ei ðtÞ ¼ QðtÞei : ð5:49Þ
The relation
ei ei ¼ I
T
T u ¼ QTQ Qu ¼ QTu:
a ¼ Qa þ c€ þ X_ X2 ðx cÞ þ 2Xðv c_ Þ: ð5:55Þ
5.8 Determinism, Local Action, and Material Frame Indifference 143
In equation (5.55) the terms X_ ðx cÞ, X2 ðx cÞ, and 2Xðv c_ Þ are
analogous to the terms, which appear in the expressions ar ¼ r€ rh_2 ,
ah ¼ rh€ þ 2r_h,
_ for acceleration components in terms of plane polar coordi-
nates ðr, hÞ. It is clear that the acceleration vector is objective only if
2
c€ þ X_ X ðx cÞ þ 2Xðv c_ Þ ¼ 0,
and transformations that satisfy this condition are known as Galilean trans-
formations. A frame of reference, with respect to which Newton’s laws are
valid, is known as an inertial frame of reference, and all frames obtained
from an inertial frame of reference by a Galilean transformation are also
inertial frames of reference.
Some further examples of physical quantities and aspects of objectivity
of interest in continuum mechanics are now considered. In what follows the
objective is taken to mean objective for all observer transformations.
The material time derivative of any objective vector is not objective
since differentiation of equation (5.48), with respect to time, gives
u_ ¼ Qu_ þ Qu:
_ ð5:56Þ
@T ðx, tÞ
T_ ¼ þ v =T ðx, tÞ ð5:57Þ
@t
or
@T ðxðX ÞtÞ
T_ ¼ ,
@t
_ T þ QTQ
_ T T
T_ ¼ QTQ þ QT Q_ : ð5:58Þ
144 Material Models and Constitutive Equations
It is evident that the material rate equation (5.57) is not suitable for use in
constitutive equations that involve rates of objective second-order tensors.
In order to obtain an objective rate of an objective second-order tensor it is
desirable to consider first the two-point deformation gradient tensor,
F ¼ Gradx, ðFiK ¼ @xi =XK Þ, and then the velocity gradient tensor,
L¼ gradv, Lij ¼ @vi @xj , and its symmetric and antisymmetric parts.
Since no change in the reference configuration of a deforming body is
observed by observers on F and F , Grad [ Grad . It then follows that
F ¼ Gradx and substitution of x ¼ cðt Þ þ Qðt ÞxðX, tÞ gives
F ¼ QF: ð5:59Þ
which shows that L is not objective. It follows from equation (5.60) that, for
the rate of deformation and spin tensors,
1
D ¼ L þ LT ¼ QDQT ð5:61Þ
2
and
1 T
T _ T,
W ¼ L L ¼ QW Q þ QQ ð5:62Þ
2
respectively, which show that the rate of deformation is objective and the
spin is not objective.
It is interesting to consider the left and right Cauchy-Green tensors,
T T
B ¼ FF and C ¼ F F, ð5:63Þ
5.8 Determinism, Local Action, and Material Frame Indifference 145
B ¼ QBQT ð5:64Þ
and
C ¼ C: ð5:65Þ
Equations (5.64) and (5.65) indicate that B is objective but C does not
satisfy condition (5.52). However, like F, it may be regarded as quasi-
objective since it does not change under an observer transformation. We
now consider the material rate of change of a second-order objective ten-
sor. Elimination of Q_ and Q_ T from equation (5.58),
Q_ ¼ W Q QW
and
Q_ T ¼ QT W þ W QT ,
Consequently,
o
T ¼ T_ WT þ TW ð5:67Þ
D
T ¼ T_ þ LT T þ TL ð5:68Þ
146 Material Models and Constitutive Equations
d drij
r_ ðtÞ ¼ rij ui uj ¼ ui uj þ rij u_ i uj þ rij ui u_ j , ð5:69Þ
dt dt
and the rij are the stress components with respect to triad ui , i 2 f1, 2, 3g.
The dependence of r on x has been suppressed. The rate of change of stress
dr
relative to the system of base vectors xi ðtÞ is dt ij ui uj . It follows from
equation (2.39) that
drij d
ui uj ¼ rij ui uj Wip rpj ui uj þ sip Wpj ui uj
dt dt ð5:71Þ
o dr
or r ¼ W r þ rW :
dt
o drij
r¼ ui uj :
dt
Exercises 5.1–5.8 147
1
g¼ I F T1 F 1
2
D
g ¼ g þ LT g þ gL ¼ D:
EXERCISES
5.1. Decompose the specific strain energy function W ðeÞ for a linear iso-
tropic elastic solid into a dilatational part and a distortional part. Obtain
the specific strain energy as a function Wc ðrÞ, known as the complemen-
tary energy. Decompose Wc ðrÞ into a dilatational part and a distortional
part and show that the distortional part is related to the octahedral
shearing stress and the second invariant of the deviatoric Cauchy stress.
5.2. The potential energy of a linear elastic body is given by
Z Z Z
V¼ wðeÞdv ti ui da qfi ui dv,
v @vt v
where @vt is the part of the surface @v upon which the surface traction
is prescribed.
Use the result that the strain energy function, wðeÞ ¼ 12 cijkl eij ekl , is
positive definite to show that, V is a minimum when ui ðxÞ is the solu-
tion displacement field in competition with all other members of the
set of kinematically admissible displacement fields.
148 Material Models and Constitutive Equations
Hint, show that V ui V ðui Þ > 0, where ui is a kinematically admis-
sible displacement field and ui is the solution displacement field, with
the equality holding if and only if ui ¼ ui
5.3. The complementary energy of a linear elastic body is given by
Z Z
Vc ¼ wc ðrÞdv ti ui da,
v @vu
where @vu is the part of the surface @v upon which the displacement is
prescribed.
Use the result that the complementary energy function,
1
wc ðrÞ ¼ 2 aijkl rij rkl , is positive definite, to show that V c is a minimum
when rij ðxÞ is the solution stress field in competition with all other
members of the setof statically
admissible stress fields.
Hint, show that V c rij V c rij > 0, where rij ðxÞ is a statically ad-
r_ ¼ 2lD þ kItrD,
o
r ¼ 2lD þ kItrD
is materially objective.
5.7. Show that a constitutive equation of the form r ¼ f ðDÞ is materially
objective.
5.8. Show that the time-dependent basic invariants of the stress tensor r
are stationary when the Jaumann stress rate vanishes.
References 149
REFERENCES
6.1 Introduction
The constitutive equations of the linear theory of elasticity, which have been
discussed in chapter 5, can be obtained as alimitingcase of those of finite
deformation elasticity by neglecting terms O j=X uj2 . However, the consti-
tutive equations of linear elasticity are not objective, although those of finite
deformation elasticity are objective. In this chapter we consider the constitu-
tive equations of finite deformation elasticity, henceforth described as non-
linear elasticity. The emphasis is on isotropic solids and it is evident where the
specialization to isotropy occurs. Symbolic notation is mainly used, although
some equations are given in both symbolic and suffix notation.
An elastic solid is a solid that is rate independent and the components of the
Cauchy stress tensor are single-valued functions of the deformation gradi-
ent tensor, for either isothermal or adiabatic deformation, so that
In equation (6.1) any dependence on X and t has been suppressed, that is,
the solid is assumed to be homogeneous in the undeformed reference
configuration. The form of the tensor function q, of the tensor argument
F, depends on whether the deformation is isothermal or adiabatic and also
depends on the reference configuration unless the solid is isotropic. The
difference between the forms of q for isothermal and adiabatic deforma-
tion, of materials for which the nonlinear elastic model is realistic, is very
150
6.2 Cauchy Elasticity 151
small and is negligible for most applications. In this chapter and chapter 7,
isothermal deformation is assumed. Equation (6.1) is an example of a con-
stitutive model for a simple material, and the function q can be obtained
from experiments involving homogeneous static deformations.
The function in equation (6.1) must be objective. An objective func-
tion is obtained by noting that the stress depends only on the change of
shape and is not influenced by rigid body rotations of a deformed config-
uration. Suppose a deformed configuration of an elastic body is given a rigid
rotation, described by a proper orthogonal tensor Q, about a fixed material
point taken as the origin. The position vector x of a particle before the
rotation becomes
x ¼ Qx
after the rotation. Then the deformation gradient and Cauchy stress tensors
after rotation are given by,
F ¼ QF ð6:2Þ
and
T
s ¼ QsQ , ð6:3Þ
s ¼ qðF Þ: ð6:4Þ
or
for any Q 2 Orthþ where Orthþ is the set of all proper orthogonal tensors.
Substitution of the polar decomposition of the deformation gradient,
F ¼ RU, in equation (6.5) gives
can be put in a different form, which may be convenient since the decom-
position of F is not required. Using the polar decomposition of F, we obtain
s ¼ FU 1 qðU ÞU 1 F T ,
T
s ¼ FfðCÞF , ð6:8Þ
6.2 Cauchy Elasticity 153
where
fðC Þ ¼ C 1=2 q C1=2 C1=2 :
S ¼ JF 1 sF T ,
S ¼ JfðC Þ, ð6:9Þ
pffiffiffiffi
where J ¼ det½U ¼ det C ¼ det½F . The Cauchy stress and the second
Piola-Kirchhoff stress can be obtained from equations (6.8) and (6.9),
respectively, for a given deformation if the function f is known. The func-
tion f is a symmetric second-order tensor function of a symmetric tensor
argument that depends on what is known as the material symmetry and must
give rise to a physically reasonable response to the application of stress.
For example, application of a hydrostatic pressure should result in a decrease
in volume, and an axial tension force applied to a bar should result in an
increase in length for either isothermal or isentropic deformation.
According to equations (6.1) and (6.9), the stress in an elastic solid is
determined uniquely from the deformation from the undeformed reference
configuration and does not depend on the deformation path. A solid that
satisfies this condition is described as Cauchy elastic. If the stress cannot be
obtained from a scalar potential function known as the strain energy func-
tion, the elastic solid is not a conservative system and the work done by the
stress could be dependent on the deformation path.
The simplest type of material symmetry is isotropy and the theory for
an isotropic elastic solid is given in this chapter. An isotropic material has
properties that are directionally independent and this is made more precise
in what follows.
If the stress due to an applied deformation is unchanged by a rigid
body rotation of the natural reference configuration, the rotation is called
154 Finite Deformation of an Elastic Solid
q9ðF9Þ ¼ q FH 1 ,
2
f ¼ /0 I þ /1 C þ /2 C , ð6:11Þ
1n o
I1 ¼ trC, I2 ¼ ðtrCÞ2 trC2 , I3 ¼ det C:
2
6.3 Hyperelasticity
P ¼ tr SE_ ¼ SKL E_ KL , ð6:12Þ
and P ¼ w_ ðEÞ. It is convenient to use suffix notation for the next develop-
ment. Then
@wðEÞ _
w_ ðEÞ ¼ EKL : ð6:13Þ
@EKL
It should be noted that in equation (6.13) all components of EKL are taken
as independent, for the differentiation, that is, the symmetry of EKL is
ignored. Then it follows from equations (6.12) and (6.13) that
@wðEÞ
SLK ¼ : ð6:14Þ
@EKL
the right Green deformation tensor C. It then follows from equation (6.12)
and E ¼ ð1=2ÞðC I Þ that
1 1
P ¼ trS C_ ¼ SKL C_ KL ð6:15Þ
2 2
and
@wðCÞ
SLK ¼ 2 : ð6:16Þ
@CKL
@wðCÞ
2 ¼ J / 0 þ /1 C þ /2 C 2 ,
@CKL
and this determines wðCÞ. For isotropy, the strain energy function wðCÞ is
a scalar isotropic function of C, that is,
w QCQT ¼ wðCÞ:
^ ðI1 , I2 , I3 Þ,
w¼w ð6:17Þ
of the invariants of C. Since the strain energy must be zero for no defor-
mation w must satisfy the equivalent conditions,
w ðI Þ ¼ 0 ^ ð3, 3, 1Þ ¼ 0:
or w
The form (6.17) has been used extensively in the literature of hyperelasticity,
and constitutive equations based on this form are now developed as follows:
@wðC Þ 3 @w
^ @Ii
¼+ , ð6:18Þ
@CKL i¼1 i @CKL
@I
158 Finite Deformation of an Elastic Solid
where
@w^ @w^ @w^ ^
@w @w^ @w^
SKL ¼2 þ I1 þ I2 dKL þ I1 CKL þ CKM CMK ,
@I1 @I2 @I3 @I2 @I3 @I3
ð6:20Þ
^
@w @w^ @w^ ^
@w @w^ @w^ 2
S¼2 þ I1 þ I2 I þ I1 Cþ C :
@I1 @I2 @I3 @I2 @I3 @I3
The relation between the Cauchy stress and the second Piola-Kirchhoff
stress
1 T
s ¼ J FSF
has already been obtained in chapter 5. Using this relation and FCF T ¼ B2
and FC 2 F T ¼ B3 , gives
@w^ @w^ @w^ ^
@w @w^ @w^ 3
s ¼ 2J 1 þ I1 þ I2 B þ I1 B2 þ B ,
@I1 @I2 @I3 @I2 @I3 @I3
@w^ @w^ @w^ @w^ 2
s ¼ 2J 1 I3 Iþ þ I1 B B : ð6:21Þ
@I3 @I1 @I2 @I2
1 @w^ @w^ @w^ @w^ 1
s ¼ 2J I2 þ I3 Iþ B I3 B , ð6:21aÞ
@I2 @I3 @I1 @I2
6.3 Hyperelasticity 159
_
w_ ðF Þ ¼ tr RFÞ: ð6:22Þ
@wðF Þ _
w_ ðF Þ ¼ FIK
@FIK
@wðF Þ
RKi ¼ : ð6:23Þ
@FiK
The relations equations (6.20) and (6.21) can be obtained from equation
(6.23), and this is left as an exercise for the reader.
The pairs of tensors fS, E ðor C=2Þg, fR, F g are said to be conjugate
since the stress power relations (6.12) or (6.15) and (6.22) are the trace of S
and the time rate of E ðor C=2Þ and the trace of R and the time rate of F. The
stress tensor s and the rate of deformation tensor D, which appear in the stress
power per unit reference volume, relation, P ¼ JtrðsDÞ, do not give rise to
160 Finite Deformation of an Elastic Solid
conjugate variables since D is not the time derivative of a strain except for
infinitesimal strains where D ’ e,
_ and e is the infinitesimal strain tensor.
^ ðI1 , I2 Þ:
w¼w ð6:24Þ
I3 1 ¼ det C 1 ¼ 0: ð6:25Þ
6.4 Incompressible Hyperelastic Solid 161
@wðC Þ @ det C
SLK ¼ 2 p : ð6:26Þ
@CKL @CKL
c
CIJ CKJ ¼ dIK det C,
CC cT ¼ IdetC,
c
where CKJ is the cofactor of CKJ , and it may be deduced that
@ det C c 1
¼ CKL ¼ CKL det C: ð6:27Þ
@CKL
Then from equations (6.25), (6.27), and the method of Lagrangian multi-
pliers we obtain
@ det C 1
¼ CKL : ð6:28Þ
@CKL
^ ðC Þ
@w 1
SLK ¼ 2 pCKL , ð6:29Þ
@CKL
@wðCÞ 2 @w
^ @Ii
¼+ ,
@CKL i¼1 @Ii @CKL
where
@I1 @I2
¼ dKL , ¼ I1 dKL CKL ,
@CKL @CKL
162 Finite Deformation of an Elastic Solid
@wðF Þ 1
RKi ¼ pFiK , ð6:33Þ
@FiK
^ ^ ^
w ¼ wðk1 , k2 , k3 Þ ¼ wðk2 , k1 , k3 Þ ¼ wðk2 , k3 , k1 Þ, ð6:34Þ
6.5 Alternative Formulation 163
^
wð1, 1, 1Þ ¼ 0:
where the unit vectors l ðiÞ , i 2 f1, 2, 3g are the eigenvectors of C and U,
pffiffiffiffi
corresponding to the ki . If only two eigenvalues of C, k1 and k2 , are dis-
tinct and k2 ¼ k3 , the spectral decomposition of C is
ð1 Þ ð1Þ ð Þ ð Þ
C ¼ k21 l l þ k22 I l 1 l 1 ,
C ¼ k2 I:
I1 ¼ k21 þ k22 þ k23 , I2 ¼ k21 k22 þ k22 k23 þ k23 k21 , I3 ¼ k21 k22 k23 ,
_ ,
so that SB and U are conjugate. Then, since w_ ðU Þ ¼ tr ð@w=@U ÞUÞ
^
B @wðUÞ
S ¼ : ð6:35Þ
@U
164 Finite Deformation of an Elastic Solid
The eigenvalues of U are ðk1 , k2 , k3 Þ and it follows from equation (6.35) that
the principal components of the Biot stress are then given by
@wðk1 , k2 , k3 Þ
SB
i ¼ , i 2 f1, 2, 3g, ð6:36Þ
@ki
@wð1, 1, 1Þ
¼ 0, i 2 f1, 2, 3g:
@ki
SB ¼ US ¼ SU,
that
JRT sR ¼ S B U: ð6:37Þ
The tensor JRT sR and the weighted Cauchy stress Js have the same
eigenvalues values. Consequently, it follows from equation (6.36) and the
principal values of each side of equation (6.37) that the principal compo-
nents of Cauchy stress are given by
^
@w
ri ¼ J 1 ki , i 2 f1, 2, 3g: No summation: ð6:38Þ
@ki
^
L1 L2 L3 dwðk1 , k2 , k3 Þ ¼ r1 l2 l3 dl1 þ r2 l1 l3 dl2 þ r3 l1 l2 dl3 ,
3 ^
dki , J ¼ Ll11Ll22lL3 3 , and the dki are arbitrary the result
^ @w
and, since dw ¼ +i @k i
(6.38) follows.
6.5 Alternative Formulation 165
It should be noted that, for an isotropic elastic solid, the right stretch
tensor, U, S, and S B are coaxial and the left stretch tensor, V, and s are coaxial.
The use of the strain energy function in the form (6.34) along with equa-
tion (6.36) or (6.38) could be the preferred method of solution for problems
with the principal directions of s and/or SB known beforehand. Problems that
involve cylindrical or spherical symmetry are examples. For problems such as
simple shear and the torsion of a cylindrical bar the use of form (6.34) is less
desirable since the principal directions of stress have to be found initially.
For incompressibility, k1 k2 k3 ¼ 1. An assumption, known as the
Valanis-Landel hypothesis [5], for an incompressible isotropic elastic
solid is
^
wðk1 , k2 , k3 Þ ¼ w
~ ðk1 Þ þ w
~ ðk2 Þ þ w
~ ðk3 Þ,
~ ð1Þ ¼ 0 and w9
where w ~ ð1Þ þ w0ð1Þ ¼ 2l: Most strain energy functions that
have been proposed for incompressible isotropic elastic solids satisfy this
hypothesis.
For incompressibility equations (6.36) and (6.38) are replaced by
^
@w p
SB
i ¼ , i 2 f1, 2, 3g: No summation,
@ki ki
and
^
@w
ri ¼ ki p, i 2 f1, 2, 3g: No summation, ð6:39Þ
@ki
respectively.
^
Since k1 k2 k3 ¼ 1, it is possible to eliminate k3 from w to obtain
ðk1 , k2 Þ ¼ w k1 , k2 , k1 1
^
w 1 k2 : ð6:40Þ
^ ^ ^ ^
@w @w @w @k3 @w 2 1 @w
¼ þ ¼ k1 k2 : ð6:41Þ
@k1 @k1 @k3 @k1 @k1 @k3
166 Finite Deformation of an Elastic Solid
@w
r 1 r3 ¼ k 1 : ð6:42Þ
@k1
Similarly
@w
r2 r3 ¼ k 2 : ð6:43Þ
@k2
Equations (6.42) and (6.43) are useful when plane stress problems with
r3 ¼ 0 are considered.
Further aspects of strain energy functions are given in chapters 7
and 8.
EXERCISES
6.1. Show that the basic invariants of the left Cauchy-Green tensor B are of
the same form as those of the right Cauchy-Green tensor C, that is,
1n o
I1 ¼ trB, I2 ¼ ðtrBÞ2 trB2 , I3 ¼ det B:
2
6.2. Show that for consistency with classical linear elastic theory the strain
energy function, wðk1 , k2 , k3 Þ, must satisfy the condition
@2w
ð1, 1, 1Þ ¼ k þ 2ldij ,
@ki @kj
^
~ ðk1 Þ þ w
w k1 , k2, k3 ¼ w ~ ðk2 Þ þ w
~ ðk3 Þ,
6.4 Show that, for an incompressible neoHookean elastic solid, the con-
stitutive equation r ¼ pI þ lB is materially objective.
REFERENCES
7.1 Introduction
168
7.2 Strain Energy Functions and Stress-Strain Relations 169
Solutions are given in this chapter for an incompressible solid based on the
Mooney-Rivlin strain energy function,
l
wðI1 , I2 Þ ¼ fcðI1 3Þ þ ð1 cÞðI2 3Þg, ð7:1Þ
2
where l is the shear modulus for infinitesimal deformation from the natural
reference state and 0 < c < 1 is a constant, or its special case with c ¼ 1, the
neo-Hookean strain energy function
l
wðI1 Þ ¼ ðI1 3Þ: ð7:2Þ
2
It has been found experimentally that, for simple tension, equations (7.1),
with c = 0.6, and (7.2) give results in good results for rubber-like materials
for axial stretches up to about 3.0 and 1.8, respectively. For larger stretches
a strain energy function of the form w ¼ w ~ ðk1 , k2 , k3 Þ, which is not readily
put in the form (6.24), gives results for simple tension that are in good
agreement with experiment. This strain energy function is considered in
the final section of this chapter.
Other strain energy functions for isotropic incompressible solids are dis-
cussed in [5].
The Hadamard strain energy function
l
wðI1 , I2 , I3 Þ ¼ fcðI1 3Þ þ ð1 cÞðI2 3Þ þ H ðI3 Þg, ð7:3Þ
2
has been proposed by Blatz and Ko [6]. In equation (7.4), m is Poisson’s ratio
for infinitesimal deformation from the natural reference state, and as
170 Some Problems of Finite Elastic Deformation
@w @w
s ¼ pI þ 2B 2B1 ð7:5Þ
@I1 @I2
2 @w^ @w^ @w^ @w^ 1
s ¼ pffiffiffiffi I2 þ I3 Iþ B I3 B ð7:6Þ
I3 @I2 @I3 @I1 @I2
x1 ¼ X1 þ kX2 , x2 ¼ X2 , x3 ¼ X3 , ð7:7Þ
2 3
1 k 0
6 7
½F ¼ ½Gradx ¼ 4 0 1 0 5,
0 0 1
2 3 ð7:8Þ
h i 1 þ k2 k 0
6 7
½B ¼ FF T ¼ 4 k 1 0 5,
0 0 1
7.3 Simple Shear of a Rectangular Block 171
X2
θ
X1
X3
Figure 7.1. Simple shear of a rectangular block.
2 3
h i 1 k 0
1
B ¼ 4 k k2 þ 1 0 5,
0 0 1
respectively.
Simple shear is a plane deformation and it is easily verified that
I1 ¼ I2 ¼ 3 þ k2 .
First we consider an incompressible solid and use equations (7.8) and con-
stitutive equation (7.5) to obtain
@w @w
r11 ¼ p þ 2 1 þ k2 2 ,
@I1 @I2
@w @w
r22 ¼ p þ 2 2 1 þ k2 ,
@I1 @I2
ð7:9Þ
@w @w
r33 ¼ p þ 2 2 ,
@I1 @I2
@w @w
r12 ¼2 þ k.
@I1 @I2
It is clear that the normal stresses cannot be obtained from the deformation
only, since the Lagrangian multiplier p depends on a stress boundary con-
dition. If r33 ¼ 0, equations (7.9) give
2@w @w
2 @w @w
r11 ¼ 2k , r22 ¼ 2k , r12 ¼2 þ k, ð7:10Þ
@I1 @I2 @I1 @I2
172 Some Problems of Finite Elastic Deformation
This analysis proceeds in the same way if r11 or r22 is taken as zero. It is
evident from equations (7.10) and (7.11) that, in order to maintain the finite
simple shear given by equation (7.7), normal stress components r11 and r22
are required for the Mooney-Rivlin and r11 for the neo-hookean Hookean,
in addition to r12 . This is a second-order effect that is a form of the Poynting
effect [8]. The normal components r11 and r22 are O k2 for the Mooney-
Rivlin solid, whereas there is a linear relation between r12 and k: The
relation between r12 and k is nonlinear for materials for which
@w=@I1 and or @w=@I2 are functions of k.
We now consider the compressible solid and use the constitutive equa-
tion (7.6). The stress components are now completely determined by the
deformation equation (7.7) and for strain energy function (7.4) are
obtained from equations (7.6) and (7.8) as
x1 ¼ k1 X1 , x2 ¼ k2 X2 , x3 ¼ k3 X3 , k2 ¼ k3 , ð7:13Þ
7.4 Simple Tension 173
where k1 > 1 for simple tension and 0 < k1 < 1 for simple compression. It
follows from equation (7.13) that the matrix of components of the left
Cauchy-Green strain tensor is given by
½B ¼ diag k21 , k22 , k22 , ð7:14Þ
2 2 2 2 4 2 4
I1 ¼ k1 þ 2k2 , I2 ¼ 2k1 k2 þ k2 , I3 ¼ k1 k2 : ð7:15Þ
2 @w 2 @w @w
r22 ¼ r33 ¼ þ k1 þ k22 þ 2k1 k22 ¼ 0: ð7:17Þ
k1 @I1 @I2 @I3
@w l @w @w l 2 4 ð1mÞ=ð12mÞ
¼ , ¼ 0, ¼ k k ,
@I1 2 @I2 @I3 2 1 2
r11
k2 ¼ 0:89643, ¼ 0:75552:
l
174 Some Problems of Finite Elastic Deformation
The nominal stress for simple tension is given by R11 ¼ r11 k22 .
The problem for an incompressible solid is simpler since the incom-
pressibility condition I3 ¼ 1 gives
1=2
k2 ¼ k3 ¼ k1 :
@w 1 @w
2 1
r11 ¼ 2 k1 k1 þ ,
@I1 k1 @I2
ð1 cÞ
r11 ¼ l k21 k1
1 c þ
k1
and
N ð1 cÞ
¼ k1 k2
1 cþ ,
pA2 l k1
where N is the axial force and A is the radius in the undeformed reference
configuration.
For simple tension of the neo-Hookean solid with k1 ¼ 1:25 it follows
from the results of this problem that
r11
k2 ¼ 0:89443, ¼ 0:7625,
l
which differ negligibly from the corresponding values for the compressible
case already considered. This indicates that, for static problems of rubber-
like materials, the incompressible model is justified unless a hydrostatic
compressive stress of the order of magnitude of the bulk modulus is applied.
The Mooney-Rivlin strain energy function and its special case, the
neo-Hookean, give results, for simple tension, with stretch less than about
two that are in good agreement with experiment. For stretches greater than
about three experimentally obtained nominal stress-stretch curves for most
7.5 Extension and Torsion of an Incompressible Cylindrical Bar 175
where
3
+ li ai ¼ 2l,
i¼1
and the ai and li are chosen to fit experimental data. For simple.tension of
pffiffiffi
an incompressible solid the axial stretch k1 ¼ k and k2 ¼ k3 ¼ 1 k and it
follows from equation (7.18) that the axial tensile nominal stress is given by
dwðkÞ 3
R¼ ¼ + li kai 1 kai =21 : ð7:19Þ
dk i¼1
Ogden [5] has shown that equation (7.19) gives a close agreement with
experimental data for several rubbers for stretches up to seven when
Stress-stretch relations, obtained from equations (7.19) and (7.20), and the
Mooney Mooney-Rivlin strain energy function with c ¼ 1 (neo-Hookean)
and c ¼ 0:6, are shown graphically in Figure 7.2.
The z axis of the cylindrical polar system is chosen to coincide with the
axis of the cylinder and the origin is taken at one end of the cylinder, and the
cylindrical polar coordinates of a particle in the spatial and undeformed
reference configurations are denoted by ðr, h, zÞ and ðR, H, ZÞ, respec-
tively, and the corresponding radii of the cylinder are r ¼ a and R ¼ A: De-
formation of the cylinder is a pure torsion with plane cross sections remaining
plane followed by a homogeneous extension or vice versa, and is given by,
1=2
r ¼ Rk , h ¼ H þ DZ, z ¼ kZ, ð7:21Þ
where D is the angle of twist per unit length of the reference configuration.
It is reasonable to expect that a stress system equivalent to a torque T and
axial force N is required to maintain the deformation (7.21).
The physical components of the deformation gradient tensor, left
Cauchy-Green strain tensor B, and B1 are
2 3
k1=2 0 0
6 7
½F ¼ 4 0 k1=2 rD 5,
0 0 k
2 1
3 ð7:22Þ
h i k 0 0
T 6 1 7
½B ¼ FF ¼ 4 0 k þ r 2 D2 krD 5,
0 krD k2
7.5 Extension and Torsion of an Incompressible Cylindrical Bar 177
2 3
h i k 0 0
1 4
B ¼ 0 k rD 5,
2 1 2 2
0 rD k þk r D
respectively.
A solution is now obtained and particularized for the Mooney-Rivlin
material. The procedure given can be adapted to consider any incompress-
ible hyperelastic material, although more complication results when
@w=@I1 and=or @w=@I2 are not constant.
Expressions that involve the Lagrangian multiplier p are obtained
from equations (7.1), (7.5), and (7.22) for the nonzero components of the
Cauchy stress,
1 @w @w
rr ¼ p þ 2 k k , ð7:23Þ
@I1 @I2
1 2 2 @w @w
rh ¼ p þ 2 k þ r D k , ð7:24Þ
@I1 @I2
2 @w 2 1 2 2 @w
rz ¼ p þ 2 k k þk r D , ð7:25Þ
@I1 @I2
@w @w
rhz ¼ rD k þ : ð7:26Þ
@I1 @I2
drr rr rh
þ ¼ 0: ð7:27Þ
dr r
drr 2 @w
¼ 2rD ,
dr @I1
178 Some Problems of Finite Elastic Deformation
Za
@w
rr ¼ 2 rD2 dr. ð7:28Þ
@I1
r
then p is obtained from equations (7.23) and (7.28) and this enables it to be
eliminated from equations (7.24) and (7.25).
It is clear from the analysis so far that equation (7.21) is a controllable
deformation. If the Mooney-Rivlin strain energy function is considered, the
solution is relatively simple since
p cD 2
2
¼ a r2 þ ck1 ð1 cÞk,
l 2
rr cD2 2
¼ a r2 , ð7:29Þ
l 2
rh cD2 2 2
¼ a 3r , ð7:30Þ
l 2
rz nc o cD2 a2
¼ D2 r2 ð1 cÞk1
l 2 2 ð7:31Þ
2
þ ð1 cÞ k k þ c k k1 ,
2
rhz
¼ fcðk 1Þ þ 1grD: ð7:32Þ
l
7.5 Extension and Torsion of an Incompressible Cylindrical Bar 179
Za
T ¼ 2p r2 rhz dr,
0
T
D
3 ¼ 4 ½cðk 1Þ þ 1, ð7:33Þ
plA 2k
¼ DA.
where D
Similarly, the axial force required is given by
Za
N ¼ 2p rrz dr,
0
and with the use of equation (7.28) this gives the nondimensional form
( ! )
N 2
D k2
3 3 3 2
¼ c k k þ ð1 cÞ 1 k þc kk :
plA2 2 2
ð7:34Þ
Equation (7.33) shows that, for the Mooney-Rivlin material, the relation
between torque and angle of twist per unit length is linear for a fixed value
of k. Equation (7.34) shows that if the length of the bar is constrained with
k ¼ 1 an axial force given by
N 2c
D
2 ¼
plA 4
is required, and if the axial force is zero, the length increases. This is a
second-order effect, known as the Poynting effect [8], which does not exist
in linear elasticity theory.
180 Some Problems of Finite Elastic Deformation
r ¼ f ðRÞ, h ¼ H, / ¼ U ð7:35Þ
h i h i h i
½B ¼ diag Q4 , Q2 , Q2 , B1 ¼ diag Q4 , Q2 , Q2 , ð7:36Þ
4 @w 4 @w
rr ¼ p þ 2 Q Q , ð7:37Þ
@I1 @I2
2 @w 2 @w
rh ¼ r/ ¼ p þ 2 Q Q , ð7:38Þ
@I1 @I2
7.6 Spherically Symmetric Expansion of a Thick-Walled Shell 181
are obtained from equations (7.5) and (7.36). The nontrivial equilibrium
equation is
drr 2ðrr rh Þ
þ ¼ 0, ð7:39Þ
dr r
dQ 1 2
¼ Q Q :
dr r
drr drr dQ
Substituting equations (7.37), (7.38), and ¼ in equation (7.39)
dr dQ dr
gives
@w @w
drr
¼ 4 Q4 Q2 Q4 Q2 Q2 Q ð7:40Þ
dQ @I1 @I2
drr n o.
4 2 4 2 2
¼ 2l Q Q c Q Q ð1 cÞ Q Q : ð7:41Þ
dQ
4 !
rr Q Q2 1
¼ þ 2Q c þ 2ð1 cÞ Q þ K, ð7:42Þ
l 2 2
where Qi ¼ A=a and Qo ¼ B=b and Pi and P0 are the internal and
external pressures, respectively. Since the shell is assumed to be
182 Some Problems of Finite Elastic Deformation
!
Qi 4 Q2i 1 P^i
þ 2Qi c þ 2ð1 cÞ Qi þK ¼ , ð7:45Þ
2 2 l
It is desirable to have a relation between Qi ¼ A=a and P^i for a given value
of g ¼ A=B. It follows from equations ð7:44Þ2 and (7.46) that
8 9
P^i < Q4o Q4i =
¼c þ 2ðQ0 Qi Þ
l : 2 ;
8 9 ð7:48Þ
< Q2o Q2i =
1 1
þ 2ð1 cÞ Qo Qi :
: 2 ;
7.6 Spherically Symmetric Expansion of a Thick-Walled Shell 183
3
The incompressibility condition B3 A ¼ b3 a3 gives, after some
manipulation, the relation
Qi
Q0 ¼ n o1=3 : ð7:49Þ
3 3 3
g þ 1 g Qi
d P^i l
¼ 0:
dða=AÞ
P^i l ¼ 0:8164 and a=A ¼ 1:828:
the wall thickness and rr ðrÞ are neglected, and the elementary approximate
relation
2rh t
P^i ’ ð7:50Þ
a
a
Q1 ’ ,
A
l
w ¼ ð2k2 þ k4 Þ,
2
pffiffiffiffiffiffiffiffiffi
where, for the thin-walled shell problem, k ’ a=A. Then, it may be
deduced that
1 dwðkÞ
rh ¼ ;
2 dk
which gives
rh a 2 a 4
’ : ð7:51Þ
l A A
t t T a 3 1
¼ ¼ g 1 : ð7:52Þ
a T a A
.
In Figure 7.4, the relation between P^i l and a=A is shown graphically for
the neo-Hookean model and g ¼ 0:975. The dashed curve is obtained
from equation (7.53) and the solid curve is obtained from equations
(7.48) and (7.49). It is evident that, for thin-walled spherical shell, the
approximate theory is in close agreement with the theory for the thick-
walled shell.
Figure 7.4. Comparison of results for g ¼ 0:975 obtained from thick-walled theory and
approximate thin-walled theory.
186 Some Problems of Finite Elastic Deformation
dr R
¼ : ð7:56Þ
dR kr
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r ¼ aR2 þ b, ð7:57Þ
0 0′
0′ 0
a B
b A
r R
l L
Figure 7.5. Eversion of a cylindrical tube.
7.7 Eversion of a Cylindrical Tube 187
ðr, RÞ b ðrr , rh , rz Þ
ðr, RÞ ¼ , b ¼ 2 , ðrr , rh , rz Þ ¼ ,
A A l
¼ 1, B
A B ¼ ¼ g:
A
" #
2
2R r2 2
½B ¼ diag a , ,a ð7:58Þ
r2 R2
2
aR
rr ¼ p þ , ð7:59Þ
r
r 2
rh ¼ p þ : ð7:60Þ
R
rz ¼ p þ a2 : ð7:61Þ
@rr rr rh
þ ¼0
@r r
gives
Z !
a2 R2 r2 dr
rr ¼ 2 2 : ð7:62Þ
r R r
188 Some Problems of Finite Elastic Deformation
Since rr ðaÞ ¼ rr ðbÞ ¼ 0 the integral (7.62) with limits of integration a and b
is zero, that is,
Zb !
a 2 R2 r2 dr
2 2 ¼ 0: ð7:63Þ
r R r
a
2 3
Zg 2 2
6 aR 17
4 2 2 5RdR ¼ 0: ð7:64Þ
2
aR þ b R
1
The everted tube is assumed to have no resultant axial force applied so that
Zb
2p rrz dr ¼ 0: ð7:65Þ
a
Using equations (7.57) and (7.59) to (7.61), it may be deduced from equa-
tion (7.65) that
Zg " #
a 2 R2 2 aR2 þ b
þ RdR ¼ 0: ð7:66Þ
aR2 þ b a2 R2
1
a ¼ 0:995729, b ¼ 2:32722:
7.8 Pure Bending of an Hyperelastic Plate 189
Since k ¼ a1 > 1 the tube lengthens when it is inverted. The non
dimensional radii in the deformed configuration
are then obtained from equation (7.57). It is easily verified that the
isochoric condition
g2 1 ¼ a1 a2 b2
is satisfied.
The Mooney-Rivlin solid can be considered with some additional
complication.
X1 ¼ S1 , X1 ¼ S2 , X2 ¼ 6L, X3 ¼ 6W=2
and
S2 S1 ¼ H:
X2
r L
a b S1 S2
0 θ
α X1
H
Figure 7.6. Finite deformation pure bending of hyperelastic plate.
r ¼ a, r ¼ b, h ¼ a, h ¼ a, z ¼ w, z ¼ w:
For the pure bending, planes normal to the X2 axis in the undeformed
reference configuration are plane in the deformed configuration and are
in meridian planes containing the X3 axis. Also the stretch, k, in the
X3 direction is uniform, so that z ¼ kX3 , k ¼ w=W; which implies that
anticlastic bending is constrained.
A material plane X3 ¼ K, where S1 < K < S2 is a constant, becomes
a sector of a cylinder of radius r subtending the angle 2a at the center. It
then follows that the circumferential stretch in the spatial configuration is
given by
ra
kh ¼ : ð7:67Þ
L
L
kr ¼ : ð7:68Þ
rak
7.8 Pure Bending of an Hyperelastic Plate 191
And since the r, h, and z directions are principal directions, the physical
components of the left Cauchy strain tensor are given by
" #
L2 r2 a2 2
½B ¼ diag 2 2 2 , 2 ,k : ð7:69Þ
kr a L
A further relation
2 2
L k b a
¼ ð7:70Þ
a 2H
rr p L2
¼ þ 2 2 2, ð7:71Þ
l l kr a
2 2
rh p r a
¼ þ 2 , ð7:72Þ
l l L
rz p
¼ þ k2 , ð7:73Þ
l l
drr ðrr rh Þ
þ ¼0
dr r
192 Some Problems of Finite Elastic Deformation
gives
" #,
1 drr L2 r2 a2
¼ 2 2 2 2 r,
l dr kr a L
" #
rr 1 L 2 r2 a2
¼ þ 2 þ c1 , ð7:74Þ
l 2 k2 r2 a2 L
rr ðaÞ ¼ 0, rr ðbÞ ¼ 0,
that
! !
2 2 2 2 2 2
1 L b a 1 L aa
c1 ¼ 2 2 2þ ¼ 2 2 2þ : ð7:75Þ
2 kb a L2 2 kaa L2
2
L
¼ kab: ð7:76Þ
a2
If k and either a or one of the radii of the cylindrical surfaces are given,
equations (7.70) and (7.76) determine the deformation.
It follows from equations (7.71) and (7.74) that
!
p 1 r2 a2 L2
¼ þ c1 , ð7:77Þ
l 2 L2 k2 r2 a2
7.8 Pure Bending of an Hyperelastic Plate 193
!
rh 1 r2 a2 L2
¼ 3 2 2 2 2 þ c1 , ð7:78Þ
l 2 L kr a
!
2 2 2
rz 1 r a L
¼ k2 þ 2 2 2 2 þ c1 , ð7:79Þ
l 2 L kr a
Zb
Fz ¼ 2a rrz dr, ð7:80Þ
a
Zb
Mz ¼ 2w rrh dr: ð7:81Þ
a
The length of a material line parallel to the X2 axis in the reference con-
figuration is 2L, and in the spatial configuration, this line is bent into an arc
of a circle of radius r and length ra. If k ¼ 1 and the length of the line are
unchanged due to the deformation, it follows from equation (7.76) that
r ¼ r0 where r0 ¼ ðabÞ1=2 , and then from equation (7.68) that the principal
stretches on surface r ¼ r0 are ð1, 1, 1Þ. Consequently, on the surface r ¼ r0
the stress is isotropic and from equations (7.74) to (7.76), (7.78) and (7.79) is
given by
The
respectively, which are solved simultaneously to obtain a and b.
nondimensional resultant force,
Zb
1n 2 2
o
Fz ¼ 2a r 1 þ rÞ ða
ða rÞ þ c1 d
r, ð7:84Þ
2
a
Zb
1 2 2
Mz =2W ¼ r ð
3 aÞ ð
r ar Þ þ c1 d
r, ð7:85Þ
2
a
obtained from equations (7.78) and (7.81). The stress distribution rz ðrÞ at
z ¼ 6w is statically equivalent to the force Fz and a couple that prevents
anticlastic bending.
Numerical results are now given in nondimensional form, for two sets
of data,
¼ 0:1, a ¼ 0:8
ðaÞ H ¼ 0:2, a ¼ 1:2:
and ðbÞ H
For (a),
For (b)
EH 3 1
Mz =2W ¼ , ð7:86Þ
2 r0
12 1 m
H3
Mz =2W ¼ ;
3
r0
which gives
4
Mz =2W ¼ 2:667 3 10
for (b). It can be shown that the difference between the values of Mz =2W
for linear and nonlinear theories increase as H and a increase.
wð AÞ ¼ wi , að AÞ ¼ ai , ð7:88Þ
where A and B are the fixed inner and outer radii, respectively, of the tube.
The problem is to determine the nonzero components of Cauchy stress so
that the axial force and torque required per unit length can be obtained.
Instead of equation (7.78) a stress boundary condition
rrz ð AÞ ¼ s, rrh ¼ q
Fixed
B A
2 3
1 0 0
6 7
½F ¼ 4 ra9 1 0 5,
w9 0 1
2 3
1 ra9 w9
6 2 2 7
½B ¼ 4 ra9 1 þ r a9 ra9w9 5, ð7:90Þ
w9 ra9w9 1 þ w92
2 3
h i 1 þ w92 þ r2 a92 ra9 w9
6 7
B1 ¼ 4 ra9 1 0 5,
w9 0 1
rr p
¼ þ c ð1 cÞ 1 þ w92 þ r2 a92 , ð7:91Þ
l l
rh p
¼ þ c 1 þ r2 a92 ð1 cÞ, ð7:92Þ
l l
rz p
2
¼ þ c 1 þ w9 ð1 cÞ, ð7:93Þ
l l
198 Some Problems of Finite Elastic Deformation
rrh
¼ ra9, ð7:94Þ
l
rrz
¼ w9, ð7:95Þ
l
rhz
¼ cra9w9: ð7:96Þ
l
@rr rr rh
þ ¼ 0, ð7:97Þ
@r r
@rrh 2rrh
þ ¼ 0, ð7:98Þ
@r r
@rrz rrz
þ ¼ 0, ð7:99Þ
@r r
d2 a 3 da
2 þ ¼ 0,
dr r dr
with solution
a ¼ c1 þ c2 r2 ,
7.9 Combined Telescopic and Torsional Shear 199
where c1 and c2 are constants of integration that are evaluated from the
boundary conditions ð7:88Þ2 and ð7:89Þ2 to give
r2 B2
a ¼ ai : ð7:100Þ
A2 B2
2
dw 1 dw
2 þ ¼ 0,
dr r dr
with solution
w ¼ c3 ln r þ c4 ,
where c3 and c4 are constants of integration that are evaluated from the
boundary conditions ð7:84Þ1 and ð7:85Þ1 to give
r A
w ¼ wi ln ln : ð7:101Þ
B B
EXERCISES
7.1. Solve the thick-walled spherical shell problem for the neo-Hookean
solid using the strain energy function in the form
^
w ¼ wðk1 , k2 , k3 Þ:
7.2. Solve the thick-walled axially symmetric cylindrical tube problem for
the neo-Hookean solid using the strain energy function in the form
^
w ¼ wðk1 , k2 , k3 Þ:
REFERENCES
The theory in this chapter is based on the Principle of Local State given in
chapter 4. Kestin [1] has noted that this principle implies that ‘‘the local and
instantaneous gradients of the thermodynamic properties as well as their
local and instantaneous rates of change do not enter into the description of
the state and do not modify the equations of state.’’ It follows that the
thermodynamic potentials are assumed to be of the same form as for an
equilibrium state.
The important potentials in finite deformation thermoelasticity are the
Helmholtz free energy and the internal energy, since, as for linear thermo-
elasticity, the Helmholtz free energy at constant temperature is the iso-
thermal strain energy and the internal energy at constant entropy is the
isentropic strain energy. This indicates the relation between hyperelasticity
and thermoelasticity.
Application of thermoelasticity to finite deformation of idealized
rubber-like materials involves the concept of entropic response and ener-
getic response and this is discussed in this chapter.
202
8.2 Basic Relations for Finite Deformation Thermoelasticity 203
SKL CKL
h ¼ hðS, sÞ ¼ u , ð8:3Þ
2q0
SKL CKL
g ¼ gðS,HÞ ¼ f : ð8:4Þ
2q0
_
u ¼ u^ðF, sÞ, f ¼ f^ðF,HÞ,
_
u ¼ uðE, sÞ and f ¼ f ðE,HÞ,
strain tensor (or E the Lagrangian strain tensor). A Gibbs relation, based
on u, that can be deduced from the first and second laws of thermodynamics is
SKL dEKL
du ¼ þ Hds
q0
ð8:5Þ
SKL dCKL
¼ þ Hds:
2q0
@u @u
du ¼ dCKL þ ds: ð8:6Þ
dCKL ds
@uðC, sÞ @uðC, sÞ
SKL ðC, sÞ ¼ 2q0 , HðC, sÞ ¼ : ð8:7Þ
@CKL @s
S~KL ðC,HÞdCKL
df ¼ sdH, ð8:8Þ
2q0
follows from equations (8.2) and (8.5). It then follows from equation (8.2)
or (8.8) that
@f @f
df ¼ dCKL þ dH: ð8:9Þ
@CKL @h
@f ðC,HÞ @f ðC, HÞ
S~KL ðC,HÞ ¼ 2q0 , sðC, HÞ ¼ : ð8:10Þ
@CKL @H
@f ðC,HÞ
u~ðC,HÞ ¼ f ðC,HÞ H .
@H
@ u~ @sðC, HÞ
cc ¼ ¼H : ð8:11Þ
@H @H
The following relation is obtained from equations (8.2), (8.10), and (8.11)
@ u~ @s
S~KL ðC,HÞ ¼ 2q0 2q0 H :
@CKL @CKL
This shows that the stress consists of two parts, one due to change of in-
ternal energy produced by deformation and the other due to the change of
entropy. The first part is described as energetic and the second as entropic.
Equations ð8:7Þ1 and ð8:10Þ1 indicate that the strain energy w is given
by
@ 2 f ðC,HÞ
cc ðC, HÞ ¼ H ,
@H2
206 Finite Deformation Thermoelasticity
ZH
@f @f cc ðC, H9Þ
¼ dH9: ð8:12Þ
@H @H H¼H0 H9
H0
@f
f ¼ u~ þ H , ð8:13Þ
@H
( )
@f 1
¼ f C; H0 u~ C, H0 ,
@H H¼H0 H0
( ) ZH
@f 1 cc ðC, H9Þ
¼ f C, H0 u~ C, H0 dH9: ð8:14Þ
@H H0 H9
H0
ZH
u~ðC, HÞ ¼ cc ðC, H9ÞdH9 þ u~ðC, H0 Þ: ð8:15Þ
H0
The relation
H H
f ðC, HÞ ¼ f ðC, H0 Þ 1 u~ðC, H0 Þ
H0 H0
ZH ð8:16Þ
H
1 cc ðC, H9ÞdH9
H9
H0
8.2 Basic Relations for Finite Deformation Thermoelasticity 207
QK @H
_ þ S~KL C_ KL 2
q0 ðHs_ uÞ > 0, ð8:17Þ
H @XK
where Q is the referential heat flux that is related to the spatial heat flux, q,
by
q ¼ J 1 FQ:
Qk @H
q0 sh_ þ f_ þ S~KL C_ KL 2 > 0, ð8:18Þ
H @XK
@u _ @u
u_ ¼ CKL þ _
s,
@CKL @s
which follows from equation (8.6), in equation (8.17) gives the entropy
inequality for a thermoelastic solid,
1 @u _ @u QK @H
SKL q0 CKL þ q0 H s_ > 0: ð8:19Þ
2 @CKL @s h @XK
Substituting
@f _ @f _
f_ ¼ CKL þ H,
@CKL @h
208 Finite Deformation Thermoelasticity
that follows from equation (8.9), into equation (8.18) gives an alternative
form
1~ @f @f _ QK @H
SKL q0 C_ KL q0 s þ H > 0: ð8:20Þ
2 @CKL @H H @XK
Inequality (8.20) must hold for all values of C_ and H,_ and this gives an
alternative derivation of equations (8.10) since the factors of C_ and H
_ in
equation (8.20) must be zero. The same reasoning applied to equations
(8.19) gives an alternative derivation of equations (8.7).
It follows from equations (8.7), and (8.19) or (8.10) and (8.20) that
QK @H @H
c_ ¼ 2 > 0, or QK > 0, ð8:21Þ
q0 H @X K @X K
where c_ is the rate of entropy production per unit mass. As an exercise the
reader should show that the spatial forms of equations (8.21) are
qk @H @H
c_ ¼ 2 > 0 or qK > 0:
qH @xk @xk
pffiffiffiffiffiffiffiffiffiffiffi
where J ¼ det½F ¼ detC . Henceforth, equation (8.22) is assumed in this
section.
Equations ð8:11Þ1 and (8.22) imply that the specific heat at constant
deformation is a function of temperature only or constant.
The bulk modulus K for infinitesimal deformation of most rubbers is
up to three orders of magnitude greater than the corresponding shear
modulus l; however, the volume coefficient of thermal expansion is about
210 Finite Deformation Thermoelasticity
three times a typical value for a metal. This has motivated a model that is
mechanically incompressible but has a nonzero coefficient of volume ther-
mal expansion [4]. We avoid the use of this model since it is thermodynam-
ically inconsistent. It may be reasonable in a thermoelastic problem
involving a rubber-like material to neglect the dilatation, due to mechanical
compressibility, compared with that due to volume thermal expansion and
this is essentially the same as assuming the mechanically incompressible
model with a nonzero coefficient of volume thermal expansion. However, it
is possible that the dilatation due to the isotropic part of the stress tensor
could be of the same order of magnitude as that due to thermal expansion
depending on the particular problem.
In this section we consider both incompressible and compressible
isotropic models for rubber-like solids and express the deformation in
terms of the principal stretches ki , i 2 f1, 2, 3g. In the present discussion
an incompressible solid is defined as a solid capable of isochoric deforma-
tion only so that the volume coefficient of thermal expansion is zero. This
model is thermodynamically consistent but not strictly physical. The tem-
perature in the natural reference configuration is denoted by H0 , and it is
assumed that jH H0 j=H0 < < 1 and jJ 1 < < 1j, for the problems
considered.
A physically realistic form for the isothermal strain energy function
per unit volume of the natural reference configuration, for an isotropic
compressible rubber rubber-like solid at temperature H0 is [2],
where l and K are the isothermal shear and bulk moduli, respectively, for
infinitesimal deformation from the natural reference configuration, / is
a nondimensional symmetric function of the principal stretches,
ki , i 2 f1, 2, 3g, which vanishes when k1 ¼ k2 ¼ k3 , and g is a nondimen-
sional function of J ¼ k1 k2 k3 that satisfies the conditions
gð1Þ ¼ 0, g9ð1Þ ¼ 0 and g0 ¼ 1, where a prime denotes differentiation with
respect to J. It may be shown that the compressible models given by equa-
tions (7.3) and (7.4) can be put in the form (8.22). For most rubbers the ratio
8.3 Ideal Rubber-Like Materials 211
l=K is of order of magnitude 103 , and it follows that for finite distortion
the dilatation may be considered infinitesimal unless the isotropic part of
the Cauchy stress is very large compared with l, Ogden [5] has proposed an
empirical relation
K 1
p¼ J J 10 , ð8:24Þ
9
( )
1 J 9 1
g¼ þ ln J : ð8:25Þ
9 9 9
p ¼ K ðJ 1Þ, ð8:26Þ
and
1
g ¼ ðJ 1Þ2 , ð8:27Þ
2
respectively, with negligible difference from equations (8.24) and (8.25) for
0:98 < J < 1:02. Consequently, equations (8.26) and (8.27) are assumed in
what follows.
Since, for the problems considered, it is assumed that
jH H0 j=H0 < < 1 the approximation cc ¼ constant is adopted, and from
equations ð8:11Þ1 and ð8:22Þ,
u~2 ¼ cc ðH H0 Þ,
212 Finite Deformation Thermoelasticity
H H H
f ðki , HÞ ¼ f ðki , H0 Þ 1 u~1 ð J Þ cc H ln þ cc ðH H0 Þ:
H0 H0 H0
ð8:28Þ
It has been shown, from the statistical theory of rubber elasticity [6], that
an ideal rubber is incompressible, is strictly entropic, that is, the internal
energy can be expressed as a function u ¼ uðHÞ of temperature, only, as for
a perfect gas and has the Helmholtz free energy, at temperature H0 , given, by
l 2
w=q0 ¼ f ðki , H0 Þ ¼ k1 þ k22 þ k23 3 , k1 k2 k3 ¼ 1. ð8:29Þ
2q0
l 2
w ¼ q0 f ðk,H0 Þ ¼ k þ 2=k 3 , ð8:30Þ
2
and
dw dw
2 2 1
R¼ ¼l kk , r¼k ¼l k k : ð8:31Þ
dk dk
8.3 Ideal Rubber-Like Materials 213
H H
f ðk, HÞ ¼ f ðk, H0 Þ cH ln þ cc ðH H0 Þ: ð8:32Þ
H0 H0
The nominal stress obtained from equations ð8:31Þ1 and (8.32) is then
given by
H df ðk,H0 Þ H dw
Rðk,HÞ ¼ q0 ¼ , ð8:33Þ
H0 dk H0 dk
@f ðk,HÞ f ðk,H0 Þ H
s¼ ¼ cc ln , ð8:34Þ
@H H0 H0
which shows that s is of the form s ¼ s1 ðkÞ þ s2 ðHÞ. It follows from equa-
tions (8.33) and (8.34) that, for isothermal simple tension,
@s
R ¼ q0 H ,
@k
and this indicates that the stress is obtained from the change of entropy
with stretch for isothermal deformation.
214 Finite Deformation Thermoelasticity
We now consider isentropic simple tension from the reference state. It may
be deduced from equation (8.34) that
f ðk,H0 Þ
H ¼ H0 exp ; ð8:35Þ
cc H0
wðkÞ
H H0 ¼ H0 exp 1
q0 cH0
3
q = 906:5 kg:m , l = 420kPa and
1 1
cc = 1662 J:kg K , H0 = 293:15K,
h i
w
R ¼ exp l k k2 ,
q0 cc H0
l 2 2 2 2=3
1
2
w ¼ q0 f ðki ,H0 Þ ¼ k1 þ k2 þ k3 3J þ K ðJ 1Þ : ð8:36Þ
2 2
H w H H
f ðki ,HÞ ¼ 1 u1 ð J Þ cc H ln þ cc ðH H0 Þ, ð8:37Þ
H0 q0 H0 H0
@f w u1 ð J Þ H
s¼ ¼ cc ln : ð8:38Þ
@H q0 H0 H0 H0
@f H @w H uð J Þ @J
d~
SB
i ¼ q0 ¼ q0 1 : i 2 f1, 2, 3g: ð8:39Þ
@ki H0 @ki H0 dJ @ki
aKH0
u1 ¼ ðJ 1Þ, ð8:40Þ
q0
/ðkr , k, HÞ ¼ 0 ð8:41Þ
is obtained from SB
2 ¼ 0, where, for isothermal simple tension H is replaced
by H0 . For isentropic simple tension from the natural reference state, with
J ¼ 1 and s ¼ 0,
w u1 ð J Þ H
cc ln ¼ 0, ð8:42Þ
q0 H 0 H0 H0
w u1 ð J Þ
H ¼ H0 exp cc ð8:43Þ
q0 c c H 0 H0
EXERCISES
8.1. Determine the Gibbs relations for the enthalpy, h ¼ hðS, sÞ, and the
Gibbs free energy, g ¼ gðS,HÞ.
References 219
@h @h @g @g
H¼ , CKL ¼ 2qo , s¼ , CKL ¼ 2q0 :
@s @SKL @H @SKL
REFERENCES
Ideal elastic solids and inviscid fluids do not exhibit dissipative effects. Con-
sequently, the only irreversible processes possible in these media are heat
conduction and shock wave propagation. Deformation and flow of dissipa-
tive media involve dissipation, due to hysteresis effects, of mechanical energy
as heat and this is an additional contribution to irreversibility. Entropy pro-
duction is a measure of irreversibility, and for a dissipative medium, it has
contributions from dissipation of mechanical energy, heat conduction, and
shock wave propagation. We will not be concerned with shock wave propa-
gation in this chapter. This topic is covered in the classic text by Whitham [1].
The dissipative media discussed in this chapter are Newtonian viscous
fluids, Stokesian fluids, and linear viscoelastic solids. Other dissipative media
are discussed in texts on rheology, for example, Tanner [2].
2
r þ pI ¼ 2gD þ n g trD I,
3
ð9:1Þ
2
rij þ pdij ¼ 2gdij þ n g dkk dij ,
3
220
9.1 Newtonian Viscous Fluids 221
where p is the thermodynamic pressure and g and n are the shear and
bulk viscosity coefficients, respectively, which are temperature depen-
dent in real fluids. The shear viscosity g of a gas increases with temper-
ature increase and the viscosity of a liquid decreases with temperature
increase. Temperature dependences of g and n are neglected in what
follows, and this restricts the theory presented to relatively small tem-
perature changes. Constitutive equation (9.1) can be decomposed into
dilatational and deviatoric parts
rkk
þ p ¼ ndkk ð9:2Þ
3
and
rkk
rij ¼ sij þ dij ,
3
ð9:4Þ
dkk
dij ¼ d9ij þ dij ,
3
nd2kk pdkk 2ldij9 dij9
and , ð9:5Þ
q q
rij dij 1 2
¼ pdkk þ 2gdij dij þ n g d2kk
q q 3
222 Dissipative Media
can also be obtained from equation (9.1) and substitution of equation (9.4)
gives, after some manipulation,
rij dij 1 2
¼ pdkk þ 2gdij9 dij9 þ ndkk : ð9:6Þ
q q
In equation (9.6) the stress power has been decomposed into three parts,
a nondissipative dilatational part pdkk =q, and dissipative parts, 2gd9ij d9ij q
due to shear deformation and nd2kk q due to dilatation. This is consistent
with equation (9.5). The dissipative parts are rates of dissipation per unit
mass of mechanical energy as heat. It follows that the rates of entropy
production per unit mass due to shear deformation and dilatation are
2
2gdij9 dij9 nd
c_ s ¼ and c_ d ¼ kk ,
qH qH
qi @H
c_ c ¼ :
qH2 @xi
Equation (9.8) indicates that a dissipation potential, / ¼ gdij dij q,
dkk ¼ 0, exists, that is , half the rate of dissipation of mechanical energy or
stress power 2gdij dij q per unit mass, so that
@/
sij ¼ q : ð9:9Þ
@dij
2
r ¼ pI þ a1 D þ a2 D ; trD ¼ 0, ð9:10Þ
1 1
I 1 ¼ 0, I2 ¼ trD92 , I3 ¼ Det½D9 ¼ trD93 :
2 3
P ¼ a1 trD2 þ a2 trD3
224 Dissipative Media
e
σ Figure 9.1. Elastic element.
The uniaxial stress and strain are denoted by r and e, respectively, and are
related by Hooke’s law,
r ¼ Ee, ð9:11Þ
r ¼ ge,
_ ð9:12Þ
r
e ¼ e 1 þ e2 ¼ þ e2 ,
E
e
σ Figure 9.2. Viscous element.
226 Dissipative Media
σ
e1 e2
σ1
σ2
and the time derivative and r ¼ ge_2 give the constitutive equation for
a Maxwell material,
E
r_ þ r ¼ Ee:
_ ð9:13Þ
g
It then follows from equation (9.13) and the initial condition rð0þ Þ ¼ Ee0
that
Et
rðtÞ ¼ Ee0 exp H ðtÞ: ð9:14Þ
g
Equation (9.14) indicates relaxation, that is, the stress decays with time and
Et
GðtÞ ¼ Eexp ð9:15Þ
g
rðtÞ ¼ GðtÞe0 :
Er
_ t > 0þ :
¼ Ee, ð9:16Þ
g
Then, it follows from equation (9.16) and initial condition eð0þ Þ that
1 t
e ¼ r0 þ : ð9:17Þ
E g
Equation (9.17) indicates creep and that the Maxwell model represents
a fluid since e increases monotonically due to a constant stress and
1 t
J ðtÞ ¼ þ : ð9:18Þ
E g
is known as the creep function. In general the creep function is the strain
eðtÞ that results from the application of stress rðtÞ ¼ r0 H ðtÞ, that is,
eðtÞ ¼ J ðtÞr0 :
The Maxwell model is now used to illustrate the superposition principle for
linear viscoelasticity. Suppose a stress rðtÞH ðtÞ is applied to the Maxwell
element, with rð0Þ ¼ r , as indicated in Figure 9.4.
The response deðtÞ due to stress increment dr applied at t ¼ s,
1 ðt sÞ
de ¼ dr þ , t > s,
E g
228 Dissipative Media
σ(t)
dσ(τ)
σ*
0 τ τ + dτ
t
Figure 9.4. Application of stress increment.
follows from equation (9.17). Since the governing equation (9.13) is linear
a superposition principle holds. Consequently,
Zt
1 t drðsÞ 1 t s
eðtÞ ¼ r þ þ þ ds
E g ds E g
þ
0
or
Zt
drðsÞ 1 t s
eðtÞ ¼ þ ds:
ds E g
0
r ¼ r1 þ r2 ,
so that, with the use of equations (9.11) and (9.12), the constitutive equa-
tion is
E r
e_ þ e¼ : ð9:19Þ
g g
Suppose the Kelvin-Voigt element is subjected to a sudden application of
axial stress rðtÞ ¼ H ðtÞr0 ; then from equation (9.19),
E r0
e_ þ e¼ : ð9:20Þ
g g
9.3 Linear Viscoelastic Medium 229
r0 Et
e¼ 1 exp : ð9:21Þ
E g
1 Et
J ðtÞ ¼ 1 exp : ð9:22Þ
E g
Et9
r ¼ Eeðt9Þ ¼ r0 1 exp ,
g
that is, when the strain is held fixed, the stress immediately relaxes.
When the strain rate approaches infinity the viscous element
approaches rigidity, consequently the impact modulus is infinite. Substitut-
ing eðtÞ ¼ e0 H ðtÞ in equation (9.19) gives
where dðtÞ is the Dirac delta function. The Dirac delta function dðtÞ, and the
unit step function H ðtÞ belong to a class known as generalized functions.
The unit function has already been defined and the Dirac delta function is
defined by
Z‘ Z‘
dðtÞ ¼ 0; t 6¼ 0, and dðtÞdt ¼ 1or uðtÞdðtÞdt ¼ uð0Þ,
‘ ‘
230 Dissipative Media
E2 σ2 η
e2 e3
σ1
E1 e1= e
and can be regarded, in a nonrigorous way, as the time derivative of H ðtÞ. The
theory of generalized functions is given in a text by Lighthill [7]. It is evident
from equation (9.23) that the Kelvin-Voigt model has an infinite impact
modulus, and this indicates that it may not be suitable for certain applications.
There are many more complicated spring-dashpot models than the
Maxwell and Kelvin-Voigt models. The so-called standard model consists
of a spring and Maxwell element in parallel as indicated in Figure 9.5 and
involves three parameters, two stiffness parameters, E1 and E2 , and a vis-
cosity g. It has been proposed as a more realistic model than the Kelvin-
Voigt model for a linear viscoelastic solid. The standard model is used to
illustrate the theory in the remainder of this chapter; however, the expres-
sions for the hereditary integrals given are valid in general for isotropic
linear viscoelasticity.
Referring to Figure 9.5,
e2 þ e3 ¼ e, ð9:24Þ
and
r1 þ r2 ¼ r: ð9:25Þ
It follows from equations (9.11) and (9.12) and the time derivative of equa-
tion (9.24) that
r_ 2 r2
þ ¼ e;
_ ð9:26Þ
E2 g
and from equations (9.25) and (9.11) and the time derivative of equation
(9.25) that
9.3 Linear Viscoelastic Medium 231
r2 ¼ r E1 e, ð9:27Þ
r_ 2 ¼ r_ E1 e:
_ ð9:28Þ
E2 E1 E2
r_ þ r ¼ ðE1 þ E2 Þe_ þ e: ð9:29Þ
g g
r ¼ E1 e þ E2 ðe qÞ ð9:30Þ
and
E2
q_ ðtÞ ¼ ðe qÞ, ð9:31Þ
g
E2 E1 E2
r_ þ r¼ e0 ,
g g
with solution
E2 t
r ¼ bexp þ E1 e0 , ð9:32Þ
g
232 Dissipative Media
E2 t
r ðt Þ ¼ E2 exp þ E1 e0 : ð9:33Þ
g
It then follows from equation (9.33) that the relaxation function for the
standard model is
E2 t
GðtÞ ¼ E2 exp þ E1 , ð9:34Þ
g
or
Zt
deðsÞ
r ðt Þ ¼ Gðt sÞ ds: ð9:35bÞ
ds
0
G(0)
G(t)
t
9.3 Linear Viscoelastic Medium 233
E2 E1 E2
r0 ¼ ðE1 þ E2 Þe_ þ e,
g g
with solution
E1 E2 t r0
e ¼ cexp þ , ð9:36Þ
gðE1 þ E2 Þ E1
1 E2 E1 E2 t
e¼ exp r0 :
E1 E1 ðE1 þ E2 Þ gðE1 þ E2 Þ
1 E2 E1 E2 t
J ðtÞ ¼ exp : ð9:37Þ
E1 E1 ðE1 þ E2 Þ gðE1 þ E2 Þ
e(t)
de(τ)
e*
τ τ + dτ
t
Figure 9.7. Application of strain increment.
234 Dissipative Media
J(∞)
J(t)
Figure 9.8. Typical creep function.
J(0)
Zt
drðsÞ
eðtÞ ¼ J ðtÞr þ J ðt sÞ ds ð9:38Þ
ds
0þ
or
Zt
drðsÞ
eðtÞ ¼ J ð t sÞ ds: ð9:38bÞ
ds
0
The forms (9.35) and (9.38), for the hereditary integrals, are valid for all
discrete linear viscoelastic models and are valid for continuous systems.
A continuum is now considered and distortion and dilation relations
are separated by expressing Cauchy stress, r, and infinitesimal strain, e, in
terms of the deviatoric and isotropic parts,
e~
rij ¼ sij þ rdij , eij ¼ eij9 þ dij ,
3
where r ¼ rkk =3 and e~ ¼ ekk .
The relaxation functions for distortion and dilatation described as
the shear and bulk relaxation functions are denoted by Gl ðtÞ and GK ðtÞ,
respectively, and the corresponding creep functions are denoted by Jl ðtÞ
and JK ðtÞ, respectively, and are obtained by replacing E1 and E2 in equa-
tions (9.34) and (9.37) by 2l1 and 2l2 for the distortional deformation and
K1 and K2 for the dilatational deformation. Axial stress and strain in the
relations already obtained for the one-dimensional viscoelastic model are
9.3 Linear Viscoelastic Medium 235
now replaced by sij and eij9 for distortional deformation and r and e~ for
dilatation.
The hereditary integrals are
Zt
deij9
sij ¼ Gl ðt sÞ ds, ð9:39Þ
ds
0
Zt
d~
e
r ¼ GK ðt sÞ ds, ð9:40Þ
ds
0
Zt
dsij
eij9 ¼ Jl ðt sÞ ds, ð9:41Þ
ds
0
Zt
d
r
e~ ¼ JK ðt sÞ ds: ð9:42Þ
ds
0
1 K2 K1 K2 t
JK ¼ exp : ð9:46Þ
K1 K1 ðK1 þ K2 Þ g^ðK1 þ K2 Þ
and
It should be noted that equations (9.47) and (9.48) do not imply that
Gl ðtÞJl ðtÞ ¼ 1 and Gl ðtÞJl ðtÞ ¼ 1, for all t. The correct result [5] is obtained
from
where the superposed bar denotes the Laplace transform and a is replaced
by l or K for shear or dilatational deformation, respectively. The Laplace
transform fðsÞ of f ðtÞ is defined by
Z‘
fðsÞ ¼ f ðtÞest dt,
0
since Gl ð‘Þ GK ð‘Þ ¼ 2l1 =K1 and, for linear elasticity,
3K ¼ 2lð1 þ mÞ=ð1 2mÞ, where m is Poisson’s ratio.
Similarly, it may be shown that
Jl ðtÞ 1þm
¼ :
JK ðtÞ 3ð1 2mÞ
It may also be shown, for equations (9.43) and (9.44), that the relaxation
times are equal, that is,
g g^
¼
2l2 K2
if Gl ðtÞ GK ðtÞ is constant for all t.
where k ¼ K 2l=3, K and l are the relaxation bulk and shear moduli,
respectively. It then follows that the stress power per unit volume is
The stress power is composed of two parts, a dissipative part, 2ge_ij9 e_ij9, and
a part, 2leij e_ij þ kekk e_ll , which is the rate of change of stored elastic energy. It
is reasonable to assume that, in general, the stress power is composed of
a dissipative part and a rate of stored energy part. This is now shown for
a more difficult case, a standard material. Continuum forms of the differential
constitutive equation (9.29), for the shear and dilatational deformation, are
2l2 4l l
s_ij þ sij ¼ 2ðl1 þ l2 Þe_ij9 þ 1 2 eij9 ð9:49Þ
g g
238 Dissipative Media
and
K2 K1 K2
r_ þ r ¼ ðK1 þ K2 Þe~_ þ e~, ð9:50Þ
g^ g^
respectively. In equations (9.49) and (9.50) it is not assumed that the vis-
cosity coefficient g is the same for shear and dilatation.
The stress power can be decomposed into a deviatoric or shear part,
Pl ¼ sij e_ij9, and a dilatational part PK ¼ re~_ so that rij e_ij ¼ Pl þ PK . A fur-
ther decomposition of each of these parts into a dissipative part and a rate
of stored energy part is possible. The part for the deviatoric deformation
governed by equation (9.49) is now considered. Equation (9.49) is equiva-
lent to [8],
and
2l2
and
K2
q~_ ¼ ðe~ q~Þ, ð9:54Þ
g^
where the tensor q with components qij ¼ qij9 þ ðq~=3Þdij , q~ ¼ qkk , is a so-called
internal variable and is analogous to the strain of the dashpot element in
Figure 9.5. More complicated viscoelastic materials whose spring and dash-
pot representations have more than one dashpot have a corresponding
number of internal variables.
9.4 Viscoelastic Dissipation 239
The stress power for the deviatoric part of the deformation obtained
from equation (9.51) is
and with the use of equation (9.52) can be put in the form
sij e_ij9 ¼ 2l1 eij9 e_ij9 þ 2l2 eij9 qij9 e_ij9 q_ ij9 þ gq_ ij9 q_ ij9: ð9:55Þ
If the parameters l1 , l2 , and g, and sij and eij9 are known, qij9 can be obtained
from equation (9.51). The first two terms on the right-hand side of equation
(9.55) represent the rate of storage of elastic strain energy and the third
term represents the rate of dissipation of mechanical energy. This is evident
if we consider the spring-dashpot model shown in Figure 9.5, since elastic
energy is stored in the springs and mechanical energy is dissipated in the
dashpot. Similarly, the part of the stress power for the dilatational defor-
mation can be put in the form
A model for linear viscoelasticity has been proposed [10], which assumes
that the dilatational deformation is purely elastic, so that equations (9.40)
and (9.42) are replaced by
r ¼ K~
e:
All the relations already obtained for the deviatoric deformation of the
standard model are valid, but for purely dilatational deformation linear
elasticity theory holds. This model is somewhat similar to that for a com-
pressible Newtonian viscous fluid with zero bulk viscosity and would be
completely analogous for the Kelvin-Voigt viscoelastic model. A disad-
vantage of the model is that Poisson’s ratio is not constant. It may be
shown that for the standard model for deviatoric deformation along with
240 Dissipative Media
purely elastic dilatation that the equilibrium and impact Poisson’s ratios
are given by
3K 2l1
me ¼
8K þ 2l1
and
3K 2ðl1 þ l2 Þ
m¼ ,
8K þ 2ðl1 þ l2 Þ
respectively.
The present treatment of linear viscoelasticity has so far neglected ther-
mal effects, although the material properties may be temperature dependent.
It is assumed that neglect of thermal effects may be justified if
jH H0 j=H0 << 1, where H0 is the temperature in the undeformed reference
configuration.
@qi
q f_ þ Hs
_ þ sH
_ qr rij e_ij þ ¼ 0, ð9:56Þ
@xi
where qi are the components of the heat flux vector, and the Clausius-
Duhem in equality is
It should be noted that equations (9.56) and (9.57) are valid only for in-
finitesimal deformation and ðH H0 Þ=H0 << 1. Consequently, it is reason-
able to assume that the thermal conductivity k is constant and Fourier’s law
of heat conduction is given by
@H
qi ¼ k :
@xi
r ¼ K e~ aðH H0 Þ , ð9:58Þ
242 Dissipative Media
and
@f @f @f _ @f @f
f_ ¼ _ þ e~_ þ
e9 Hþ g_ i þ q_ ij ð9:61Þ
@eij9 e
@~ @H @gi @qij
1 @f 1 @f _ @f _
sij e_ij9 þ r e~ s þ H ð9:62Þ
q @eij9 q e
@~ @H
@f @f 1
g_ i q_ ij qi gi > 0:
@gi @qij qh
@f @f @f
rij9 ¼ q , r~ ¼ q , s ¼ , ð9:63Þ
@eij9 e
@~ @h
9.5 Some Thermodynamic Considerations in Linear Thermoelasticity 243
and f does not depend on gi . Then, equations (9.62) and (9.58) become
@f 1
q_ ij qi gi > 0, ð9:64Þ
@qij qh
respectively.
The elimination of gi from equation (9.60) is also justified by invoking the
Principle of Local State. In equation (9.64), the first term is independent of
gi ; consequently,
@f
q_ ij > 0, ð9:66Þ
@qij
and this is the rate of dissipation of mechanical energy. It follows from the
mechanical dissipation term that appears in equation (9.55) and the expres-
sion (9.66) for mechanical dissipation that
@f
q q_ ij ¼ gq_ ij q_ ij : ð9:67Þ
@qij
@f
qdf ¼ 2ðl1 þ l2 Þeij9 2l2 qij deij9 þ K fe~ aðH H0 Þgd~
e
ð9:69Þ
þ qsdH þ 2l2 qij eij9 dqij :
H H0
seij9¼~e¼qij¼0 ¼ c0 ; ð9:70Þ
H0
2
e~
qf ¼ ðl1 þ l2 Þeij9 eij9 2l2 qij eij9 þ K
2
ð9:71Þ
c0 ðH H0 Þ2
KaðH H0 Þ~
e þ l2 qij qij ;
2 H0
Ka~e H H0
s¼ þ c0 ,
q H0
@s
c0 ¼ H :
@H
1 1
then
e12 t
q_ 12 ¼ exp : ð9:75Þ
T T
The rate of mechanical energy dissipation per unit volume, obtained from
equation (9.55), is given by
D ¼ 2gq212 ,
he ti2
12
D ¼ 2g exp : ð9:76Þ
T T
246 Dissipative Media
The mechanical energy dissipated, per unit volume when equilibrium has
been reached is given by
e 2 Z‘ h ti2 e2
12 2
Wd ¼ 2g exp dt ¼ g 12 ¼ 2l2 e12 , ð9:77Þ
T T T
0
EXERCISES
9.1. Show that equations (9.1) and (9.7) are frame indifferent.
9.2. Show that relation (9.10) is frame indifferent and trr=3 ¼ p if and
only is the fluid is in equilibrium, that is, D[0.
9.3. Determine the nonzero components of the extra stress r þ pI, given by
equations (9.7) and (9.10) for simple shear flow with velocity field
v1 ¼ kx2 , v2 ¼ 0, v3 ¼ 0:
9.4. (a) Verify that J ð0Þ ¼ ðGð0ÞÞ1 for linear viscoelastic solids and fluids,
and J ð‘Þ ¼ ðGð‘ÞÞ1 for linear viscoelastic solids.
σ3 E2
e2 E1 σ
e Exercise 5. Chapter 9.
e1
σ2
η
where e12 is a constant strain. Show that for the standard material
s12 ¼ 0, t < 0,
Tn to
s12 ¼ 2l1 e12 ðtÞ þ 2l2 e12 1 exp , 0 < t < t9,
t9 T
t
T t t9
s12 ¼ 2l1 e12 þ 2l2 e12 exp exp , t > t9,
t9 T T
E2 e
rðeÞ ¼ E1 e þ ge_ 1 exp :
ge_
E0
E1 η1
E2 η2
σ Exercise 8. Chapter 9.
EN ηN
Determine the set of relaxation times Tk ; k 2 f1; 2; :::N gs, the relaxa-
tion function and the creep function. The set of relaxation functions is
known as the relaxation spectrum.
REFERENCES
A1.1 Introduction
Rectangular Cartesian coordinate systems are well suited for presenting the
basic concepts of continuum mechanics since many mathematical com-
plications that arise with other coordinate systems, for example,
convected systems, are avoided. However, for the solution of specific prob-
lems, the use of rectangular Cartesian coordinate systems may result in con-
siderable difficulties, and the use of other systems may be desirable in order to
take advantage of symmetry aspects of the problem. For example, in the
analysis of the expansion of a thick-walled cylindrical tube, the use of cylin-
drical polar coordinates has an obvious advantage. We introduce a simple
theory of curvilinear coordinates in this appendix and specialize it for orthog-
onal curvilinear systems, in particular cylindrical and spherical. We avoid the
use of the general theory of tensor components referred to curvilinear coor-
dinates by considering what are known as the physical components of tensors
that are derived for orthogonal coordinate systems. Superscripts are used to
denote curvilinear coordinates.
x1 ¼ y1 , x2 ¼ y2 , x3 ¼ y3
249
250 Orthogonal Curvilinear Coordinate Systems
xi ¼ xi h1 , h2 , h3 : ðA1:2Þ
A superscript, rather than a suffix, is used for hj , since hj are not Cartesian
coordinates. However, the summation convention applies as for suffixes.
A point with rectangular coordinate xi may also be specified by the values
of hj at the point since equation (A1.2) represent three families of surfaces
and the intersection of three surfaces hj ¼ Pj , where each Pj is a constant,
locates a point. The Pj , j 2 f1, 2, 3g are the curvilinear coordinates of the
point. It is assumed that the functions (A1.2) are at least of class C1 , that is,
continuous with continuous first derivatives, and since equation (A1.2)
must be invertible,
@xi
det 6¼ 0,
@hj
@x
gi ¼ : ðA1:3Þ
@hi
A1.2 Curvilinear Coordinates 251
θ 3curve
θ 2surface θ 1surface
g3
g1 g2
Figure 1.1A. Curvilinear coordinates.
θ 3surface
θ 2curve
θ 1curve
The base vectors for rectangular Cartesian coordinate systems are given by
ej ¼ @x=@xj . Consequently,
@xs @hi
gi ¼ es , es ¼ gi : ðA1:4Þ
@hi @xs
The vectors gi are not in general unit vectors nor are they in general con-
stants, except for the special case of oblique Cartesian coordinates, that is,
when equations (A1.1) represent three nonorthogonal families of planes.
Also the vectors gi are nondimensional if and only if the corresponding hi
have the dimension of length.
Referring to Figure 1.2A, the point P has position vector x ¼ xi ei and
curvilinear coordinates hj , and the neighboring point has position vector,
x þ dx ¼ ðxi þ dxi Þei , and curvilinear coordinates hj þ dhj .
If PP# ¼ ds,
2
ds ¼ dx dx ¼ ei dxi ej dxj ,
where
@xk @xk
gij ¼ gi gj ¼ , ðA1:6Þ
@hi @hj
252 Orthogonal Curvilinear Coordinate Systems
x3
dx P′
e3 P
x
x + dx
0 e2 x2
e1
x1
A1.3 Orthogonality
gij ¼ 0 if i 6¼ j,
2 2 2
ds2 ¼ g11 dh1 þg22 dh2 þg33 dh3 , ðA1:7Þ
where
2 2 2
@x1 @x2 @x3
gii ¼ þ þ , ðs:c:sÞ: ðA1:8Þ
@hi @h i
@hi
g g
bi ¼ i ¼ i , ðs:c:sÞ: ðA1:10Þ
gi hi
The orthogonal triad of unit base vectors, bi , at P forms a basis for a local-
ized Cartesian coordinate system; however, in general as P moves, the
orientation of the triad changes. It follows that the vectors bi , are nondi-
mensional and of constant unit magnitude but in general are not constant
vectors.
Consider the vector field uðxÞ where x is the position vector of any
point P as indicated in Figure 1.2A. The vector uðxÞ can be expressed in
terms of its components with respect to the triad of unit base vectors bi , that
is, with respect to the local Cartesian coordinate system at P as
u ¼ ui bi :
h1 ¼ r ¼ 0P#, h2 ¼ / ¼ x1 ^
0P#, h3 ¼ z ¼ PP#:
254 Orthogonal Curvilinear Coordinate Systems
Consequently,
1=2 x2
r ¼ x21 þ x22 , / ¼ arctan , z ¼ x3 :
x1
h1 ¼ 1, h2 ¼ r, h3 ¼ 1, ðA1:12Þ
The unit base vectors, b1 , b2 , and b3 are in the directions r, /, and z in-
creasing, respectively, as indicated in Figure 1.3A.
The spherical polar coordinates of P are indicated in Figure 1.4A.
x3 b3
z
0
x2
φ Figure 1.3A. Cylindrical polar coordinates.
b2
r
P′
b1
x1
A1.4 Cylindrical and Spherical Polar Coordinate Systems 255
x3
b1
P
b2
θ r
0
x2
φ b3 Figure 1.4A. Spherical polar coordinates.
P′
x1
1 2 3
h ¼ r ¼ 0P, h ¼ h ¼ x3 ^0P, h ¼ / ¼ x1 0P#:
Consequently,
where 0 < h < p, 0 < / < 2p. The angles h and / are the colatitude and
azimuth angles, respectively. Equation (A1.13) can be inverted to give
2 2 2
1=2 x21 þ x22 x2
r ¼ x1 þ x2 þ x3 , h ¼ arctan , / ¼ arctan ;
x3 x1
h1 ¼ 1, h2 ¼ r, h3 ¼ r sin h, ðA1:14Þ
2 2 2 2 2 2 2
ds ¼ dr þ r dh þ r sin hd/ :
256 Orthogonal Curvilinear Coordinate Systems
@ðÞ @ðÞ
= [ ei , =[ ei , ðA1:15Þ
@xi @xi
S ¼ S~ij bi bj ,
where S~ij denotes the physical components of S . Since the unit base vectors,
bi ; are not, in general, constants, this must be considered when S is operated
on by a differential operator such as = .
If C is a scalar field, then at a point in the field
dC
=C bi ¼ ,
dsi
where bi are the unit base vectors given by equation (A1.10) and si is the
distance along the coordinate curve hi increasing.
Henceforth, the summation convention is suspended in this appendix.
It follows from equations (A1.7) and (A1.9) that
3 1 @ðÞ
=[ + i bi : ðA1:16Þ
i¼1 hi @h
3
1 @C
=C[ + i bi
i¼1 hi @h
@bi
ðA1:19Þ
@hj
@ x @ hj bj @ ðhi bi Þ
i j ¼ i ¼ , i 6¼ j, ðs:c:sÞ
@h @h @h @hj
258 Orthogonal Curvilinear Coordinate Systems
@bi 1 @hj
j ¼ bj , ðA1:21Þ
@h hi @hi
@bi @ bj 3 bk
¼
@hi @hi
2 3
0 b/ 0
@bi 4
¼ 0 br 05 ðA1:23Þ
@hj 0 0 0
2 3
0 bh b/ sin h
@bi
¼ 40 br b/ cos h 5 ðA1:24Þ
@hj 0 0 br sin h bh cos h
3 1 @u
divu ¼ + i bi ,
i¼1 hi @h
1 @ ðu1 h2 h3 Þ @ ðu2 h1 h3 Þ @ ðu3 h1 h2 Þ
divu ¼ þ þ : ðA1:25Þ
h1 h2 h3 @h1 @h2 @h3
is obtained from equations (A1.12) and (A1.25), and for spherical polar
coordinates
2
1 @ u r r 1 @ ðuh sin hÞ 1 @u/
divu ¼ 2 þ þ
r @r r sin h @h r sin h @/
curlu ¼ = 3 u,
3 1 @u
curlu ¼ + i 3 bi ,
h
i¼1 i @h
260 Orthogonal Curvilinear Coordinate Systems
2 3
h1 b1 h2 b2 h3 b3
1 6 @ @ @ 7
curlu ¼ det6
4 1
7:
5 ðA1:26Þ
h 1 h2 h3 @h @h2 @h3
h1 u1 h2 u2 h3 u3
1 @uz @u/ @ur @uz 1 @ ru/ 1 @uz
curlu ¼ br þ b/ þ bz
r @/ @z @z @z r @r r @/
where a second summation arises when @u @hi is evaluated using equations
(A1.21) and (A1.22). It follows from equations (A1.23) and (A1.27) that,
for cylindrical polar coordinates, the matrix of components of = u is
2 3
@ur 1 @ur @ur
a/
6 @r r
@/ @z 7
6 7
½= u ¼ 6 @u/ 1 @u/
a/
@u/
7,
4 @r r @/ @z 5
@uz 1 @uz @uz
@r r @/ @z
A1.6 The =2 Operator 261
and from equations (A1.24) and (A1.27) that, for spherical polar
coordinates,
2
3
@uR 1 @ur 1 @ur
@r r @h uh u/ sin h
r sin h @/
6
7
6 h 7
½= u ¼ 6 @u 1 @uh
r @h þ uR
1 @uh
r sin h @/ u/ cos h 7:
4 @r 5
@u/ 1 @u/ 1 @u/
@r r @h r sin h @/ þ uh cos h þ ur sin h
2
3 @2
= [ +
i¼1 @xi @xi
=2 [ div grad ¼ = =,
2 1 @ h2 h3 @ðÞ @ h3 h1 @ðÞ @ h1 h2 @ðÞ
= [ þ 2 þ 3 ,
h1 h2 h3 @h1 h1 @h1 @h h2 @h2 @h h3 @h3
ðA1:28Þ
@2 1@ 1 @2 @2
=2 [ 2þ þ 2 2þ 2: ðA1:29Þ
@r r @r r @/ @z
262 Orthogonal Curvilinear Coordinate Systems
1 @ 2 @ðÞ 1 @ @ðÞ
=2 [ 2 r þ 2 sin h
r @r @r r sin h @h @h
ðA1:30Þ
1 @ 2 ðÞ
þ 2 2 2 :
r sin h @/
2 2 ur
2 @u/ 2 u/ 2 @ur 2
= u ¼ br = ur 2 2 þ b/ = u/ 2 þ 2 þ bz = u z :
r r @/ r r @/
The terms that result from the nonconstant base vectors, br and b/ , are
clearly evident.
Appendix 2
F ¼ =X x,
@xi
F¼ ei EK ,
@XK
263
264 Physical Components of the Deformation Gradient Tensor
x ¼ rbr : ðA:2:4Þ
@ 1 @ @
F¼ ðrbr þ zbz Þ BR þ ðrbr þ zbz Þ BF þ ðrbr þ zbz Þ BZ :
@R R @F @Z
ðA2:5Þ
2 3
0 b/ 0
@bi 4
j ¼ 0 br 0 5, i, j 2 fr, /, zg,
@h 0 0 0
2 3 2 3 2 3 2 3 2 3
b b b b b
@ 4 r5 @ 4 r 5 @r @ 4 r 5 @/ @ 4 r 5 @z 4 / 5 @/
b/ ¼ b/ þ b/ þ b/ ¼ br :
@R @r @R @/ @R @z @R @R
bz bz bz bz 0
ðA2:6Þ
Physical Components of the Deformation Gradient Tensor 265
Similarly
2 3 2 3
b b
@ 4 r 5 4 / 5 @/
b/ ¼ br ðA2:7Þ
@F @F
bz 0
and
2 3 2 3
br b/
@ 4 5 4 @/
b/ ¼ br 5 : ðA2:8Þ
@Z @Z
bz 0
@r 1 @r @r
F¼ br B R þ br B U þ br B Z þ
@R R @F @Z
@/ r @/ @/
r b/ BR þ b/ B U þ r b/ BZ þ
@R R @F @Z
@z 1 @z @z
bz B R þ bz BU þ bz BZ :
@R R @F @Z
Consequently,
2 3 2 @r 1 @r @r
3
FrR FrF FrZ @R R @F @Z
4 F/R 6 @/ @/ 7
F/F F/Z 5 ¼ 4 r @R r @/
R @F r @Z 5: ðA2:7Þ
FzR FzF FzZ @z 1 @z @z
@R R @F @Z
2 3
2 @r 1 @r 1
3
@r
FrR FrH FrU @R R @Q R sin h @F
4 FhR 6 @h 7 ðA2:8Þ
FhH FhU 5 ¼ 4 @h
r @R r @h
R @Q
r
R sin h @F 5
F/R F/H F/U @/ r @/ r @/
r sin h @R R sin h @Q R @F
Legendre Transformation
@/ @/
d/ ¼ dx þ dy: ðA3:1Þ
@x @y
Let
@/ @/
p¼ and q ¼ ; ðA3:2Þ
@x @y
w ¼ / px qy, ðA3:4Þ
with differential,
266
Legendre Transformation 267
dw ¼ xdp ydq,
where
@w @w
x¼ , y¼ : ðA3:6Þ
@p @q
w ¼ / px, ðA3:7Þ
with differential
dw ¼ qdy xdp,
where
@w @w
x¼ , q¼ : ðA3:9Þ
@p @y
f ¼ u Hs:
268 Legendre Transformation
q ¼ S=ð2q0 Þ:
SKL CKL
gðS, HÞ ¼ u Hs:
2q0
@g @g
sðS, HÞ ¼ , CKL ðS, HÞ ¼ 2q0 :
@H @SKL
Appendix 4
a. x þ y ¼ y þ x (commutative property),
b. x þ ðy þ zÞ ¼ ðx þ yÞ þ z (associative property),
c. There exists a zero vector, 0, such that x þ 0 ¼ x,
d. For every x there is a negative, ðxÞ, such that x þ ðxÞ ¼ 0.
x y ¼ x1 y1 þ x2 y2 þ x3 y3 ¼ xi yi , ðA4:1Þ
h. x y ¼ y x (commutative property),
i. x ðy þ zÞ ¼ x y þ x z (distributive property),
j. x x > 0 (equality holds when x ¼ 0 )
269
270 Linear Vector Spaces
a1 x1 þ a2 x2 þ a3 x3 þ an xn ¼ 0
holds only when all the numbers a1 ; a2 ; a3 a4 are zero, otherwise the vec-
tors are linearly dependent. The basis of a linear vector space is any linearly
independent system maximum order. For example, for the three-dimensional
space of vectors of elementary vector analysis any three non coplanar vectors
x, y, and z is a basis. This means any vector u of the space can be expressed as
a linear combination
u ¼ a1 x þ a2 y þ a3 z
u ¼ u1 e1 þ u2 e2 þ u3 e3
as shown chapter 1.
Linear Vector Spaces 271
n
qða, bÞ ¼ + ai bi :
i¼1
X ¼ Xij ei ej ,
h T i
tr ei ej ðek et Þ ¼ dik djt :
A scalar product
qðX, Y Þ ¼ X Y ¼ tr XY T
272
Index 273