tensors+1:2 kinematics
tensors+1:2 kinematics
Continuum
Mechanics
of Solids
Lallit Anand
Department of Mechanical Engineering, Massachusetts Institute of Technology,
Cambridge, MA 02139, USA
Sanjay Govindjee
Department of Civil and Environmental Engineering, University of California,
Berkeley, Berkeley, CA 94720, USA
1
3
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Lallit Anand and Sanjay Govindjee 2020
The moral rights of the authors have been asserted
First Edition published in 2020
Reprinted with corrections in 2023
Preface
Continuum Mechanics is a useful model of materials which views a material as being contin-
uously divisible, making no reference to its discrete structure at microscopic length scales
well below those of the application or phenomenon of interest. Solid Mechanics and Fluid
Mechanics are subsets of Continuum Mechanics. The focus of this book is Solid Mechanics,
which is concerned with the deformation, flow, and fracture of solid materials. A study of
Solid Mechanics is of use to anyone who seeks to understand these natural phenomena, and
to anyone who seeks to apply such knowledge to improve the living conditions of society. This
is the central aim of people working in many fields of engineering, e.g., mechanical, materials,
civil, aerospace, nuclear, and bio-engineering. This book, which is aimed at first-year graduate
students in engineering, offers a unified presentation of the major concepts in Solid Mechanics.
In addressing any problem in Solid Mechanics, we need to bring together the following major
considerations:
• Kinematics or the geometry of deformation, in particular the expressions of
stretch/strain and rotation in terms of the gradient of the deformation field.
• Balance of mass.
• The system of forces acting on the body, and the concept of stress.
• Balance of forces and moments in situations where inertial forces can be neglected,
and correspondingly balance of linear momentum and angular momentum when they
cannot.
• Balance of energy and an entropy imbalance, which represent the first two laws of
thermodynamics.
• Constitutive equations which relate the stress to the strain, strain-rate, and temperature.
• The governing partial differential equations, together with suitable boundary conditions
and initial conditions.
While the first five considerations are the same for all continuous bodies, no matter what mate-
rial they are made from, the constitutive relations are characteristic of the material in question,
the stress level, the temperature, and the timescale of the problem under consideration.
The major topics regarding the mechanical response of solids that are covered in this
book include: Elasticity, Viscoelasticity, Plasticity, Fracture, and Fatigue. Because the consti-
tutive theories of Solid Mechanics are complex when presented within the context of finite
deformations, in most of the book we will restrict our attention to situations in which the
deformations may be adequately presumed to be small—a situation commonly encountered
in most structural applications. However, in the last section of the book on Finite Elasticity of
elastomeric materials, we will consider the complete large deformation theory, because for such
materials the deformations in typical applications are usually quite large.
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
vi PREFACE
In addition to these standard topics in Solid Mechanics, because of the growing need for
engineering students to have a knowledge of the coupled multi-physics response of materials
in modern technologies related to the environment and energy, the book also includes
chapters on Thermoelasticity, Chemoelasticity, Poroelasticity, and Piezoelectricity—all within
the (restricted) framework of small deformations.
For pedagogical reasons, we have also prepared a companion book with many fully
solved example problems, Example Problems for Continuum Mechanics of Solids (Anand and
Govindjee, 2020).
The present book is an outgrowth of lecture notes by the first author for first-year graduate
students in engineering at MIT taking 2.071 Mechanics of Solid Materials, and the material
taught at UC Berkeley in CE 231/MSE 211 Introduction to Solid Mechanics by the second
author. Although the book has grown out of lecture notes for a one-semester course, it contains
enough material for a two-semester sequence of subjects, especially if the material on coupled
theories is to be taught. The different topics covered in the book are presented in essentially
self-contained “modules”, and instructors may pick-and-choose the topics that they wish to
focus on in their own classes.
We wish to emphasize that this book is intended as a textbook for classroom teaching or self-
study for first-year graduate students in engineering. It is not intended to be a comprehensive
treatise on all topics of importance in Solid Mechanics. It is a “pragmatic” book designed to
provide a general preparation for first-year graduate students who will pursue further study or
conduct research in specialized sub-fields of Solid Mechanics, or related disciplines, during the
course of their graduate studies.
Lallit Anand
Cambridge, MA
Sanjay Govindjee
Berkeley, CA
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
Acknowledgments
Lallit Anand is indebted to his teacher Morton Gurtin of CMU, who first taught him continuum
mechanics in 1975–76, and once again during the period 2004–2010, when he worked with
Gurtin on writing the book, The Mechanics and Thermodynamics of Continua (Gurtin et al.,
2010). He is also grateful to his colleague David Parks at MIT—who has used a prelimi-
nary version of this book in teaching his classes—for his comments, helpful criticisms, and
his contributions to the chapters on Fracture and Fatigue. Discussions with Mary Boyce,
now at Columbia University, regarding the subject matter of the book are also gratefully
acknowledged.
Sanjay Govindjee is indebted to his many instructors and in particular to his professors at
the Massachusetts Institute of Technology and at Stanford University, who never shied away
from advanced concepts and philosophical discussions. Special gratitude is extended to the late
Juan Carlos Simo, Jerome Sackman, and Egor Popov. Additional thanks are given to current
colleague Robert L. Taylor and countless students whose penetrating questions have formed
his thinking about solid mechanics.
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
Contents
x CONTENTS
II Kinematics 33
3 Kinematics 35
3.1 Motion: Displacement, velocity, and acceleration 35
3.1.1 Example motions 37
Simple shear 37
Elementary elongation 37
Rigid motion 37
Bending deformation 38
3.2 Deformation and displacement gradients 39
3.2.1 Transformation of material line elements 40
3.2.2 Transformation of material volume elements 41
3.2.3 Transformation of material area elements 42
3.3 Stretch and rotation 43
3.3.1 Polar decomposition theorem 44
3.3.2 Properties of the tensors U and C 45
Fiber stretch and strain 46
Angle change and engineering shear strain 47
3.4 Strain 48
3.4.1 Biot strain 48
3.4.2 Green finite strain tensor 48
3.4.3 Hencky’s logarithmic strain tensors 49
3.5 Infinitesimal deformation 49
3.5.1 Infinitesimal strain tensor 49
3.5.2 Infinitesimal rotation tensor 51
3.6 Example infinitesimal homogeneous strain states 52
3.6.1 Uniaxial compression 52
3.6.2 Simple shear 53
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
CONTENTS xi
xii CONTENTS
IV Linear elasticity 95
CONTENTS xiii
xiv CONTENTS
CONTENTS xv
xvi CONTENTS
CONTENTS xvii
xviii CONTENTS
CONTENTS xix
xx CONTENTS
CONTENTS xxi
xxii CONTENTS
28 Fatigue 529
28.1 Introduction 529
28.2 Defect-free approach 532
28.2.1 S-N curves 532
28.2.2 Strain-life approach to design against fatigue failure 534
High-cycle fatigue. Basquin’s relation 535
Low-cycle fatigue. Coffin–Manson relation 536
Strain-life equation for both high-cycle and low-cycle fatigue 537
Mean stress effects on fatigue 538
Cumulative fatigue damage. Miner’s rule 538
28.3 Defect-tolerant approach 539
28.3.1 Fatigue crack growth 539
28.3.2 Engineering approximation of a fatigue crack growth curve 542
28.3.3 Integration of crack growth equation 543
CONTENTS xxiii
xxiv CONTENTS
CONTENTS xxv
Bibliography 679
Index 689
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
PA R T I
The theories of Solid Mechanics are mathematical in nature. Thus in this chapter and the next,
we begin with a discussion of the essential mathematical definitions and operations regarding
vectors and tensors needed to discuss the major topics covered in this book.1
The approach taken is a pragmatic one, without an overly formal emphasis upon mathemati-
cal proofs. The end-goal is to develop, in a reasonably systematic way, the essential mathematics
of vectors and tensors needed to discuss the major aspects of mechanics of solids. The physical
quantities that we will encounter in this text will be of three basic types: scalars, vectors, and
tensors. Scalars are quantities that can be described simply by numbers, for example things like
temperature and density. Vectors are quantities like velocity and displacement—quantities that
have both magnitude and direction. Tensors will be used to describe more complex quantities
like stress and strain, as well as material properties. In this chapter we deal with the algebra of
vectors and tensors. In the next chapter we deal with their calculus.
1
The mathematical preliminaries in this chapter and the next are largely based on the clear expositions by Gurtin
in Gurtin (1981) and Gurtin et al. (2010).
Continuum Mechanics of Solids. Lallit Anand and Sanjay Govindjee, Oxford University Press (2020).
© Lallit Anand and Sanjay Govindjee, 2020.
DOI: 10.1093/oso/9780198864721.001.0001
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
and hence has the value +1, −1, or 0 when {i, j, k} is an even permutation, an odd permuta-
tion, or not a permutation of {1, 2, 3}, respectively. Where convenient, we will use the compact
notation
def
{ei } = {e1 , e2 , e3 },
to denote our positively oriented orthonormal basis.
2
The alternating symbol is sometimes called the permutation symbol.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
= ui v i = u1 v 1 + u2 v 2 + u3 v 3 . (1.1.7)
It should also be recalled that this relation is equivalent to u · v = |u| |v| cos θuv , where | · |
denotes the length of a vector (its magnitude or norm) and θuv is the angle between vectors
u and v. Equation (1.1.4) further implies that the cross product of two vectors
u × v = (uj ej ) × (vk ek )
= u j v k (ej × ek )
= eijk uj vk ei . (1.1.8)
In particular, (1.1.8) implies that the vector u × v has the component form
(u × v)i = eijk uj vk . (1.1.9)
This definition of the cross product is fully compatible with the relation |u × v| = |u| |v| sin θuv ,
which indicates that the magnitude of the cross product of two vectors is equal to the area of
the parallelogram whose edges are defined by u and v.
It can be shown that the permutation symbol is related to the Kronecker delta by
⎡ ⎤ ⎡ ⎤
δi1 δi2 δi3 δi1 δj1 δk1
eijk = det ⎣ δj1 δj2 δj3 ⎦ = det ⎣δi2 δj2 δk2 ⎦ , (1.1.10)
δk1 δk2 δk3 δi3 δj3 δk3
where det[·] represents the determinant of a 3 × 3 matrix. A consequence of (1.1.10) is that
⎡ ⎤
δip δiq δir
eijk epqr = det ⎣ δjp δjq δjr ⎦ ,
δkp δkq δkr (1.1.11)
= δip (δjq δkr − δjr δkq ) − δiq (δjp δkr − δjr δkp ) + δir (δjp δkq − δjq δkp ).
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
Some useful identities which follow from successive contractions of (1.1.11), the substitution
property of the Kronecker delta and the identity δii = 3 are
u · (u × v) = 0,
(1.1.15)
u × v = −v × u,
u × ( v × w) = ( u · w) v − ( u · v ) w.
The operation
def
[u, v, w] = u · (v × w), (1.1.16)
is known as the scalar triple product and is physically equivalent to the (signed) volume of the
parallelepiped whose edges are defined by any three linearly independent vectors u, v, and w.
1.2 Tensors
1.2.1 What is a tensor?
In mathematics the term tensor is a synonym for the phrase “a linear transformation which
maps a vector into a vector.” A tensor S is therefore a linear mapping that assigns to each vector
u a vector
v = Su. (1.2.1)
One might think of a tensor S as a machine with an input and an output: if a vector u is the
input, then the vector v = Su is the output. Linearity of a tensor S is the requirement that
S(u + v) = Su + Sv for all vectors u and v,
(1.2.2)
S(αu) = αSu for all vectors u and scalars α.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
TENSORS 7
3
Other common terminologies for a tensor product are tensor outer product or simply outer product, as well as,
dyadic product.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
With this definition of the components of a tensor, the Cartesian component representation of
the relation v = Su is
vi = ei · v = ei · Su
= ei · S(uj ej )
= (ei · Sej )uj
= Sij uj . (1.2.6)
It should be observed that (1.2.6) indicates that each component of v is a linear combination
of the components of u, where the weights are given by the components of S.
Further, a tensor S has the representation
S = Sij ei ⊗ ej , (1.2.7)
in terms of its components Sij and the basis tensors ei ⊗ ej .
TENSORS 9
The product of the identity tensor 1 with any arbitrary tensor T, is T itself:
(1T)v = 1(Tv) = Tv,
and since this holds for all vectors v we have
1T = T.
Alternatively, in components
(1T)ij = ei · 1Tej
= 1 ei · Tej
= ei · Tej
= Tij .
The components of 1T are equal to the components of T. The two are equal to each other.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
TENSORS 11
= Sij tr(ei ⊗ ej )
= Sij (ei · ej )
= Sii , (1.2.22)
and the trace is well-defined for all tensors. Some useful properties of the trace are
tr(Su ⊗ v) = v · Su,
tr(S ) = tr S,
(1.2.23)
tr(ST) = tr(TS),
tr 1 = 3.
As a consequence of (1.2.23)2 ,
tr S = 0 whenever S is skew. (1.2.24)
A tensor S is deviatoric (or traceless) if
tr S = 0, (1.2.25)
and we refer to
S ≡ dev S
def
= S − 13 (tr S)1 (1.2.26)
as the deviatoric part of S,4 and to
1
3 (tr S)1 (1.2.27)
as the spherical part of S. Trivially,
S = S − 13 (tr S)1 + 13 (tr S)1,
S s1
4
The notation “dev” is useful for denoting the deviatoric part of the product of many tensors; e.g., dev(ST · · · M).
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
where s = 13 tr S, and it is noted that tr dev S = 0. Thus every tensor S admits the decomposition
S = S + s1 (1.2.28)
into a deviatoric tensor and a spherical tensor.
TENSORS 13
Consider a symmetric tensor S (Sij = Sji ) and a skew tensor Ω (Ωij = −Ωji ). Their
double contraction S : Ω = Sij Ωij will always be zero:
Sij Ωij = −Sij Ωji = −Sji Ωji = −Sij Ωij .
Since the only scalar with the property that it is equal to its negative is zero, we conclude
that Sij Ωij = 0 whenever S is symmetric and Ω is skew.
• All the operations and operators we have introduced for vectors and tensors are in one-
to-one correspondence to the same operations and operators for matrices.
For example,
⎡ ⎤⎡ ⎤
S11 S12 S13 u1
[S][u] = ⎣ S21 S22 S23 ⎦ ⎣ u2 ⎦
S31 S32 S33 u3
⎡ ⎤
S11 u1 + S12 u2 + S13 u3
= ⎣ S21 u1 + S22 u2 + S23 u3 ⎦
S31 u1 + S32 u2 + S33 u3
⎡ ⎤
S u
i 1i i
=⎣ S u ⎦
i 2i i
i S3i ui
= [Su],
so that the action of a tensor on a vector is consistent with that of a 3 × 3 matrix on a 3 × 1
matrix. Further, the matrix [S ] of the transpose S of S is identical to the transposition of the
matrix [S]:
⎡ ⎤
S11 S21 S31
[S ] ≡ [S] = ⎣ S12 S22 S32 ⎦ .
S13 S23 S33
14 VECTORS AND TENSORS: ALGEBRA
Similarly, the trace of a tensor S is equivalent to the conventional definition of this quantity
from matrix algebra
tr S ≡ tr [S] = S11 + S22 + S33 = Skk .
= S11 (S22 S33 − S23 S32 ) − S12 (S21 S33 − S23 S31 ) + S13 (S21 S32 − S22 S31 )
= eijk Si1 Sj2 Sk3 = 16 eijk epqr Sip Sjq Skr . (1.2.31)
It is useful also to observe that
det(ST) = det S det T, (1.2.32)
and that
det S = det S. (1.2.33)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
TENSORS 15
e2
e∗2
e1
o
e∗3
e∗1
e3
We now consider how the two representations (vi , Sij ) and (vi∗ , Sij
∗
) in the two coordinate
systems are related to each other.
Let Q be the rotation tensor defined by
TENSORS 17
Thus, the transformation relation for the Cartesian components of a tensor S under a change
of basis are:
∗
Sij = Qik Qjl Skl . (1.2.43)
µ Sµ
For (1.2.45) to possess a non-trivial (μ = 0) solution, the tensor (S −ω1) can not be invertible,
so that, by (1.2.35),
det(S − ω1) = 0.
Thus, the determinant of the 3 × 3 matrix [S] − ω[1] must vanish,
det [S] − ω[1] = 0,
which is simply the classic requirement for a system of homogeneous equations to have a non-
trivial solution. Each eigenvalue ω of a tensor S is, therefore, a solution of a polynomial equation
of the form ω 3 − a1 ω 2 + a2 ω − a3 = 0, where the coefficients are functions of S. A tedious
computation shows that this characteristic equation5 has the explicit form
ω 3 − I1 (S)ω 2 + I2 (S)ω − I3 (S) = 0, (1.2.47)
where I1 (S), I2 (S), and I3 (S), called the principal invariants6 of S, are given by
⎫
I1 (S) = tr S, ⎪
⎪
⎪
⎬
2
I2 (S) = 2 tr(S) − tr(S ) ,
1 2 (1.2.48)
⎪
⎪
⎪
⎭
I3 (S) = det S.
Ik (S) are called invariants because of the way they transform under the group of orthogonal
tensors:
Ik (QSQ ) = Ik (S) for any orthogonal tensor Q. (1.2.49)
The solutions of the characteristic equation (1.2.47), cubic in ω , are the eigenvalues ωi ,
i = 1, 2, 3. Since the principal invariants Ik (S) are always real, the theory of polynomials tells
us that
• The characteristic equation has (i) either three real roots (not necessarily distinct), or (ii)
one real and two complex conjugate roots.
Once the eigenvalues have been determined, the eigenvectors corresponding to each eigenvalue
can be found by substituting back into (1.2.45).
5
The Cayley–Hamilton theorem also tells us that a tensor satisfies its characteristic equation; that is
S3 − I1 (S)S2 + I2 (S)S − I3 (S)1 = 0. (1.2.46)
6
The principal invariants can also be defined in terms of any collection of three vectors a, b, c, with [a, b, c] = 0,
via the relations
I1 (S) = [Sa, b, c] + [a, Sb, c] + [a, b, Sc] /[a, b, c],
I2 (S) = [Sa, Sb, c] + [a, Sb, Sc] + [Sa, b, Sc] /[a, b, c],
I3 (S) = [Sa, Sb, Sc] /[a, b, c].
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
TENSORS 19
Spectral Theorem Let S be symmetric with distinct eigenvalues ωi . Then there is a corre-
sponding orthonormal set {μi } of eigenvectors of S and, what is most important, one uniquely
has that
3
S= ωi μ i ⊗ μ i . (1.2.50)
i=1
The relation (1.2.50), which is called a spectral decomposition of S, gives S as a linear combi-
nation of projections, with each μi ⊗ μi (no sum) a projection tensor onto the eigenvector μi .
Since the eigenvectors {μi } are orthonormal, they can also function as a basis, and in this
basis the matrix representation S is diagonal:
⎡ ⎤
ω1 0 0
[S] = ⎣ 0 ω2 0 ⎦ .
0 0 ω3
When the eigenvalues of S are repeated, the spectral theorem needs to be modified. The
remaining relevant cases are when two of the eigenvalues are the same and when all three
eigenvalues are the same. Respectively, in these cases one has
(a) If ω1 = ω2 = ω3 , then S admits the representation
S = ω1 μ1 ⊗ μ1 + ω2 (1 − μ1 ⊗ μ1 ), (1.2.51)
which indicates that μ1 , as well as any vector orthogonal to μ1 , is an eigenvector of S.
(b) If ω ≡ ω1 = ω2 = ω3 , then S admits the representation
S = ω1, (1.2.52)
which indicates that any vector will qualify as an eigenvector of S and that S is spherical.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
The decomposition (1.2.55) leads to the possibility of using the invariant (tr S) and the
invariants of S as an alternative set of invariants. Since S has only two independent non-zero
invariants 2
tr S , tr S3 , (1.2.56)
it is sometimes convenient to adopt the list
tr S, tr S2 , tr S3 (1.2.57)
as a set of alternative invariants of a tensor S.
B = CA. (1.2.58)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
TENSORS 21
In continuum mechanics, the scalars, vectors, and tensors whose properties that we discussed
in Chapter 1 are used to represent physical quantities such as temperature, displacement, stress,
and the like. These quantities are typically functions of position within a body. For example,
the temperature at the left end of a rod ϑ(xleft ) could differ from the temperature at the right
end of the rod ϑ(xright ), in which case physical experience teaches us that there will be a
flow of energy from the hotter end to the cooler end, and the flow rate is dependent on the
gradient (rate of change with respect to position) of the temperature. In the general case, the
temperature is what we term a scalar field, a function of position, ϑ(x), defined over the domain
of the material body, Ω, of interest. We will also encounter vector fields and tensor fields in what
follows. An example of a vector field would be the displacement field in a deformed body, u(x).
An example of a tensor field would be the stresses in a body under load, σ(x). In each of these
cases we will be interested in computing rates of change (from position to position) and in
integral manipulations of the fields.
The notation Dϕ(x)[h] stands for an operator Dϕ(x) operating linearly on h. The linear
operator Dϕ(x) is the gradient of ϕ at x. Other common notations for the gradient are
grad ϕ(x) and ∇ϕ(x). It should be noted that the computed rate of change will be per unit h.
Thus it is common to assume |h| = 1, so that the directional derivative provides rates of change
of ϕ per unit x. The expression after the last equality in (2.1.1) is often the most convenient way
of computing the directional derivative.
Continuum Mechanics of Solids. Lallit Anand and Sanjay Govindjee, Oxford University Press (2020).
© Lallit Anand and Sanjay Govindjee, 2020.
DOI: 10.1093/oso/9780198864721.001.0001
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
Slope
h
x
Ω
The definition given also applies to vector and tensor fields. Thus for a vector field v(x), the
directional derivative at a point x in the direction h is given by
v(x + αh) − v(x) d
Dv(x)[h] = lim = v(x + αh) , (2.1.2)
α→0 α dα α=0
and similarly for tensor fields. It should be noted that in the case of a vector field, the
directional derivative is also a vector each of whose components gives the rate of change of
the corresponding component of v in the direction of h. The gradient in this case will be a
tensor field (that when applied to h gives the directional derivative of v in the direction of h).
are, respectively, vector and tensor fields, whose components are given by
∂ϕ(x)
ei · grad ϕ(x) = grad ϕ(x) i = ,
∂xi
(2.2.1)
∂vi (x)
ei · grad v(x)ej = grad v(x) ij
= .
∂xj
The latter equalities for each case show that the components of the gradient expressions are
simply the derivatives of the components in the coordinate directions.
To show the correctness of the component expression (2.2.1)1 expand Definition (2.1.1):
d ∂ϕ ∂ϕ ∂ϕ
ϕ(x1 + αh1 , x2 + αh2 , x3 + αh3 ) = h1 + h2 + h3
dα α=0 ∂x 1 ∂x 2 ∂x 3
∂ϕ
= hi .
∂xi
(grad ϕ)i
1
It also common to see the alternate notations ∇· v = div v, ∇× v = curl v, ∇· T = div T, and ∇× T = curl T
for the divergence and curl of vectors and tensors. Also, the definition of the curl of a tensor field is not consistent in the
literature, and one needs to be aware of the definition being used by a particular author. We will employ the definition
given in (2.2.2), but it is common to also see the alternate definition (curl T)ij = eipq Tqj,p .
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
The Laplacian or Laplace operator for scalar fields ϕ and vector fields v are defined as
follows:
∂2ϕ
ϕ = div grad ϕ, ϕ = ,
∂xi ∂xi
(2.2.3)
∂ 2 vi
v = div grad v, vi = ;
∂xj ∂xj
thus, ϕ is a scalar field and v is a vector field. Further, for a tensor field T, T is the tensor
field defined such that
(T)a = (Ta) for every constant vector a, (2.2.4)
or equivalently, using components,
∂Tij
Tij = . (2.2.5)
∂xk ∂xk
Notation:
In indicial notation it is customary to employ a comma to denote partial differentiation with
respect to a spatial coordinate, as follows:
∂ϕ ∂vi ∂Tij
= ϕ,i , = vi,j , = Tij,k , etc.
∂xi ∂xj ∂xk
Thus, the Laplacians of ϕ and v may be written as
Δϕ = ϕ,ii and Δvi = vi,jj ,
and so on. We will use this shorthand notation whenever convenient.
GENERALIZED DERIVATIVES 27
INTEGRAL THEOREMS 29
The derivative of a tensor with itself can be found by considering the function
G(T) = T,
so that
∂T
DG(T)[H] = H.
∂T
Expanding gives
d
(T + αH) = H = IH.
dα α=0
Thus
∂T ∂Tij
= I, = δik δjl .
∂T ∂Tkl
In the case that T is symmetric this result does not possess the correct (minor) symmetries
upon exchange of i ↔ j or k ↔ l. For the symmetric case, with some effort, it can be
shown that the appropriate expression should be ∂ T/∂ T = Isym , ∂Tij /∂Tkl = 12 (δik δjl +
δil δjk ).
Note that these identities follow a general rule: the ni in the surface integral results in the partial
derivative ∂/∂xi in the volume integral; thus, for a tensor of any order,
∂Tij···k
Tij···k nr da = dv. (2.4.3)
∂xr
∂R R
The central utility of the divergence theorem is that it allows one to convert surface expres-
sions/information to volume expressions/information. Later we will see that the divergence
theorem together with the localization theorem will allow us to derive important governing
equations for many physical phenomena.
Considering a domain Ω with volume vol(Ω) and a constant tensor ε, find the value of
the surface integral
1
εij xj ni da.
vol(Ω)
∂Ω
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
INTEGRAL THEOREMS 31
The Stokes theorem plays a very important role in fluid mechanics and theories of elec-
tromagnetic phenomena, to name a few. In the mechanics of solid materials it plays a less
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
important role, but we will see one important application of the theorem in the context of
small deformation mechanics.
PA R T II
Kinematics
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
3 Kinematics
Continuum mechanics, is for the most part, a study of deforming bodies, and kinematics is
the mathematical description of the possible deformations that a body may undergo—it is
the first step in understanding the mechanics of solids. In this chapter we will develop the
necessary tools for describing general deformations and for analyzing them. We will start with
descriptions of deformations of arbitrary magnitude, which leads to non-linear measures of
strain. Following this, we will specialize our results to the important case of small deformations,
which results in linear measures of strain.
Continuum Mechanics of Solids. Lallit Anand and Sanjay Govindjee, Oxford University Press (2020).
© Lallit Anand and Sanjay Govindjee, 2020.
DOI: 10.1093/oso/9780198864721.001.0001
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
36 KINEMATICS
Reference body
Deformed body
X
Bt
u=x−X
e3
(X−o) x
o (x−o)
e2
e1
is the unique image of some X ∈ B , that is the mapping is one-to-one and has a unique
inverse
X = χ−1 (x, t), Xi = χ−1
i (x1 , x2 , x3 , t), (3.1.2)
for every point in the body.
The vector
u(X, t) = χ(X, t) − X, ui (X1 , X2 , X3 , t) = χi (X1 , X2 , X3 , t) − Xi , (3.1.3)
represents the displacement of X at time t.
The partial derivatives
∂
u̇(X, t) = χ̇(X, t) = χ(X, t),
∂t
(3.1.4)
∂2
ü(X, t) = χ̈(X, t) = 2 χ(X, t),
∂t
represent the velocity and acceleration (vectors) of material points, respectively. Using the
inverse map (3.1.2) we can alternatively describe the velocity and acceleration as functions
of (x, t):
v(x, t) ≡ χ̇ χ−1 (x, t), t , (3.1.5)
−1
v̇(x, t) ≡ χ̈ χ (x, t), t . (3.1.6)
These then provide the velocity and acceleration of the material particle currently occupying
the position x in space.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
Simple shear
Simple shear is a deformation that is similar to the deformation that occurs when one slides a
stack of papers or a deck of cards. The motion involves only displacement in the e1 -direction
and that motion is proportional to a parameter γ that controls the magnitude of the motion;
cf. Fig. 3.2. Expressed mathematically, one has
χ(X1 , X2 , X3 , t) = (X1 + γ(t)X2 )e1 + X2 e2 + X3 e3 ,
x1 = X1 + γ(t)X2 , x 2 = X2 , x 3 = X3 . (3.1.7)
Elementary elongation
Elementary elongation is a motion which involves a displacement only in the e1 -direction and
that motion is proportional to a parameter α that controls the magnitude of the motion; cf. Fig.
3.3. Mathematically, one has
χ(X1 , X2 , X3 , t) = α(t)X1 e1 + X2 e2 + X3 e3 , x1 = α(t)X1 , x 2 = X2 , x 3 = X3 .
(3.1.8)
Rigid motion
A rigid motion is a motion in which the distance between any two points in the body remains
constant at all times. Such a motion can be described by a translation vector c(t), a rotation
tensor Q(t), and a fixed point, say, o, as
χ(X, t) = c(t) + Q(t)(X − o). (3.1.9)
e2
1
γ
e1
Fig. 3.2 Simple shear deformation. Dashed outline indicates the deformed body.
e2
αL
L
e1
Fig. 3.3 Elementary elongation deformation. Dashed outline indicates the deformed body.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
38 KINEMATICS
Verify that the distance between any two points in a rigid body remains constant during
a rigid motion. Consider any two material points Xa and Xb . At any time t the distance
between these points is given by
|xa − xb | = |c(t) + Q(t)(Xa − o) − c(t) − Q(t)(Xb − o)|
= |Q(t)(Xa − Xb )|
= Q(t)(Xa − Xb ) · Q(t)(Xa − Xb )
= (Xa − Xb ) · Q(t) Q(t)(Xa − Xb )
= (Xa − Xb ) · (Xa − Xb )
= |Xa − Xb |,
and is hence a constant for all time.
Bending deformation
The first three example deformations are relatively simple and good for testing one’s intuition.
This last example motion is a bit more complex, but at the same time also represents an intuitive
motion. Consider a beam whose axis displaces according to a given function v(X1 , t) and
whose cross-sections rotate (independently of v(X1 , t)) according to a second given function
θ(X1 , t); cf. Fig. 3.4. The resulting motion can be expressed as
χ(X1 , X2 , X3 , t) = [X1 − X2 sin θ(X1 , t)] e1 + [X2 + v(X1 , t) − X2 (1 − cos θ(X1 , t))] e2
+ X3 e3
x1 = X1 − X2 sin θ(X1 , t),
x2 = X2 + v(X1 , t) − X2 (1 − cos θ(X1 , t)) , (3.1.10)
x 3 = X3 .
θ(X1)
e2
v(X1)
e1
Fig. 3.4 Kinematics of a shear deformable beam. Dashed outline indicates the deformed body.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
∂ ∂
F(X, t) = ∇χ = χ(X, t), Fij = χi (X1 , X2 , X3 , t) , (3.2.1)
∂X ∂Xj
∂ ∂
H(X, t) = ∇u = u(X, t), Hij (X1 , X2 , X3 , t) = ui (X1 , X2 , X3 , t).
∂X ∂Xj
(3.2.2)
Thus, since
u(X, t) = x − X = χ(X, t) − X,
the displacement gradient and the deformation gradient tensors are related by
• If at a given time the deformation gradient F, and hence the displacement gradient H, is
independent of X, then the deformation is said to be homogeneous at that time.
To find deformation gradient that corresponds to simple shear, start with (3.1.7) and apply
(3.2.1):
⎡ ⎤
∂χ1 /∂X1 ∂χ1 /∂X2 ∂χ1 /∂X3 ⎡ ⎤
⎢ ⎥ 1 γ 0
⎢ ∂χ /∂X ∂χ /∂X ∂χ /∂X ⎥
[F] = ⎢
⎢
2 1 2 2 2 3 ⎥ = ⎣ 0 1 0 ⎦.
⎥
⎣ ⎦ 0 0 1
∂χ /∂X ∂χ /∂X ∂χ /∂X
3 1 3 2 3 3
With these basic definitions, it is now possible to discuss how lines of material particles,
volumes of material, and areas of material in a body transform during deformation. These
relations will then allow us to deduce sensible definitions for stretch and strain in bodies
undergoing arbitrary deformations.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
40 KINEMATICS
≈ F(X(1) , t)dX,
where in the last step we have ignored the higher order terms in the Taylor series expansion.
Thus in summary, the tensor F maps short material vectors (line elements) from the reference
configuration to the deformed configuration.
Further, since at each time t the mapping x = χ(X, t) is one-to-one, the determinant of F
∂χ
J ≡ det = det F = 0. (3.2.5)
∂X
Reference body
dX (2)
X Deformed body
X(1)
Bt
dx
x(2)
x(1)
Fig. 3.5 Schematic showing how F maps dX in B to dx in Bt , where dX = X(2) − X(1) and dX = X(2) − X(1) ,
as |dX| → 0.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
Since F is not singular, the inverse F−1 exists, and hence the following also holds true:
dX = F−1 dx. (3.2.6)
42 KINEMATICS
Reference body
nR
B Deformed body
daR
Bt n
dvR
da
dv
The quantity J , which was introduced in (3.2.5) as a shorthand notation for the determinant
of F, is called the volumetric Jacobian, or simply the Jacobian of the deformation. It is
the appropriate mathematical quantity for describing how infinitesimal volumes of material
transform from the reference to the deformed configuration. One also observes that J is the
appropriate measure of volumetric stretch (ratio of current to original volume) at a point, and
e = J − 1 is an appropriate measure of volumetric strain.
The ratio of the volume of an infinitesimal volume element of material during rigid motion
to its volume in the reference configuration is unity. To see this compute the Jacobian of
the motion given in (3.1.9).
dv
J= = det F = det Q = 1,
dvR
where the last equality follows from the property that the determinant of a rotation is +1;
see Sec. 1.2.16.
Exploiting expression (1.2.37) for the cofactor of an invertible tensor, gives Nanson’s For-
mula:
nda = J F− nR daR . (3.2.9)
Nanson’s formula tells us that the appropriate mathematical quantity for describing how area
elements transform from the reference to the deformed configuration is the cofactor of the
deformation gradient. From this one can also define area stretch at a point, λarea = da/daR =
|J F− nR |, and area strain earea = λarea −1. Note that these quantities depend on the orientation
of the area element at the point of interest.
To find the area stretch for an area element with normal e1 in three-dimensional simple
shear we need to compute the inverse transpose of the deformation gradient and its
Jacobian:
⎡ ⎤ ⎡ ⎤
1 γ 0 −1 1 −γ 0
[ F] = ⎣ 0 1 0 ⎦ , F =⎣ 0 1 0 ⎦, and
0 0 1 0 0 1
⎡ ⎤
− 1 0 0
F = ⎣ −γ 1 0 ⎦ .
0 0 1
Since J = 1 in simple shear, it follows that
λarea = |F− e1 | = 1 + γ2.
44 KINEMATICS
whereby
V = RUR , (3.3.3)
and
R = FU−1 = V−1 F. (3.3.4)
• U and V are known as the right and left stretch tensors, respectively.
• The left and right stretch tensors are useful in theoretical discussions but are often
problematic to apply because of the need to compute the square root of a tensor. For
that reason, we introduce the right and left Cauchy–Green (deformation) tensors C
and B defined by:
∂χk ∂χk
C = U2 = F F, Cij = Fki Fkj = ,
∂Xi ∂Xj
(3.3.5)
2
∂χi ∂χj
B = V = FF , Bij = Fik Fjk = .
∂Xk ∂Xk
• Note that the stretch tensors, U and V, as well as the Cauchy–Green tensors, C and B,
are symmetric and positive definite.
• Being symmetric and positive definite, U and V admit spectral representations of the
form
3
U= λ i ri ⊗ ri ,
i=1
(3.3.6)
3
V= λ i li ⊗ li ,
i=1
1
The square root of a symmetric tensor is defined via its spectral decomposition (see Sec. 1.2.18). A symmetric
√ def √
tensor S admits the spectral representation S = i si vi ⊗ vi , and its square root is defined as S = i si vi ⊗ vi .
The square root is always taken as the positive square root in this context.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
where
• The tensors C and B have the following forms when expressed in terms of principal
stretches and directions:
3
C= λ2i ri ⊗ ri ,
i=1
(3.3.10)
3
B= λ2i li ⊗ li .
i=1
= dX(1) · U2 dX(2) ,
46 KINEMATICS
Reference body
Deformed body
(2)
dX(1) dX
X
Bt
dx(1)
dx(2)
x
Fig. 3.7 Schematic showing how F maps two non-colinear, infinitesimal line elements at X.
• The stretch λ at X relative to any given material direction e is determined by the right stretch
tensor U(X) through the relation λ(e) = |U(X)e|, or equivalently by λ2 = e · Ce.
• The stretch λ at X takes its extremal values when e is one of the right principal directions;
thus the principal stretches are the extremal values. This result follows because the
necessary equations for maximizing (3.3.16) subject to |e| = 1 are given by the relations
Ue = λe, the eigenvalue equations for U.