Vous êtes sur la page 1sur 15

Chapter 1

Introduction and Mathematical


Preliminaries

Abstract The theory of elasticity comprises a consistent set of equations which


uniquely describe the state of stress, strain, and displacement at each point within an
elastic deformable body. Engineering approaches are often based on a strength-of-
materials formulation with its various specialized derivatives such as the theories of
rods, beams, plates, and shells. The distinguishing feature between the various
alternative approaches and the theory of elasticity is the pointwise description, as
opposed to sectional description embodied in elasticity. The basic elements of the
theory are equilibrium equations relating the stresses, kinematic equations relating
the strains and displacements, constitutive equations relating the stresses and strains,
boundary conditions relating to the physical domain, and uniqueness constraints
relating to the applicability of the solution. In this chapter, the mathematical prelim-
inaries for the presentation, such as vector algebra, integral theorems, indicial
notation, and Cartesian tensors, are introduced.

1.1 Scope

The theory of elasticity comprises a consistent set of equations which uniquely


describe the state of stress, strain, and displacement at each point within an elastic
deformable body. Solutions of these equations fall into the realm of applied math-
ematics, while applications of such solutions are of engineering interest. When
elasticity is selected as the basis for an engineering solution, a rigor is accepted
that distinguishes this approach from the alternatives, which are mainly based on the
strength-of-materials with its various specialized derivatives such as the theories of
rods, beams, plates, and shells. The distinguishing feature between the various
alternative approaches and the theory of elasticity is the pointwise description
embodied in elasticity, without resort to expedients such as Navier’s hypothesis of
plane sections remaining plane.
The theory of elasticity contains equilibrium equations relating the stresses,
kinematic equations relating the strains and displacements, constitutive equations
relating the stresses and strains, boundary conditions relating to the physical domain,
and uniqueness constraints relating to the applicability of the solution. Origination of

© Springer International Publishing AG, part of Springer Nature 2018 1


P. L. Gould, Y. Feng, Introduction to Linear Elasticity,
https://doi.org/10.1007/978-3-319-73885-7_1
2 1 Introduction and Mathematical Preliminaries

the theory of elasticity is attributed to Claude Louis Marie Henri Navier, Siméon
Denis Poisson, and George Green in the first half of the nineteenth century [1].
In subsequent chapters, each component of the theory will be developed in full
from fundamental principles of physics and mathematics. Some limited applications
will then be presented to illustrate the potency of the theory as well as its limitations.

1.2 Vector Algebra

A vector is a directed line segment in the physical sense. Referred to the unit basis
vectors (ex, ey, ez) in the Cartesian coordinate system (x, y, z), an arbitrary vector
A may be written in component form as
A ¼ Ax ex þ Ay ey þ Az ez : ð1:1Þ

Alternately, the Cartesian system could be numerically designated as (x1, x2, x3),
whereupon

A ¼ A 1 e1 þ A 2 e2 þ A 3 e3 : ð1:2Þ

The latter form is common in elasticity. An example is vector r in Fig. 1.1, where
the unit vectors e1, e2, and e3 are identified [2]. Beyond the physical representation, it
is often sufficient to deal with the components alone as ordered triples,

A ¼ ðA1 ; A2 ; A3 Þ: ð1:3Þ

The length or magnitude of A is given by


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
jAj ¼ A21 þ A22 þ A23 : ð1:4Þ

Vector equality, addition, and subtraction are trivial, while vector multiplication
has two forms. The inner, dot, or scalar product is

C¼ AB
¼ A1 B1 þ A2 B2 þ A3 B3 ð1:5Þ
¼ jAjjBj cos θAB :

Also, there is the outer, cross, or vector product

C¼ AB
ð1:6aÞ
¼ ðA2 B3  A3 B2 Þe1 þ ðA3 B1  A1 B3 Þe2 þ ðA1 B2  A2 B1 Þe3 ,
1.3 Scalar and Vector Fields 3

x3

x3

x3 x2
x3
e3
e3 e2
x2
e1 0 e2
e1 x1
x1

x2

x1 x2

x1

Fig. 1.1 Cartesian coordinate systems (After Tauchert [2]). Reproduced by permission

which is conveniently evaluated as a determinant


 
 e1 e2 e3 

C ¼ A  B ¼  A1 A2 A3 ; ð1:6bÞ
 B1 B2 B3 

C is perpendicular to the plane containing A and B.

1.3 Scalar and Vector Fields

1.3.1 Definitions

A scalar quantity expressed as a function of the Cartesian coordinates such as

f ðx1 ; x2 ; x3 Þ ¼ constant ð1:7Þ

is known as a scalar field. An example is the temperature at a point. A vector


quantity similarly expressed, such as A(x1, x2, x3), is called a vector field. An
example is the velocity of a particle. We are especially concerned with changes or
derivatives of these fields.
4 1 Introduction and Mathematical Preliminaries

1.3.2 Gradient

The gradient of a scalar field f is defined as


 
∂f ∂f ∂f ∂f ∂f ∂f
grad f ¼ —f ¼ e1 þ e2 þ e3 ¼ ; ; , ð1:8Þ
∂x1 ∂x2 ∂x3 ∂x1 ∂x2 ∂x3

where grad f is a vector point function which is orthogonal to the surface f ¼ constant,
everywhere. Conversely, the components of grad f may be found by the appropriate
dot product, for example,

∂f
¼ e1  —f : ð1:9Þ
∂x1

1.3.3 Operators

The del operator — may be treated as a vector

∂ð Þ ∂ð Þ ∂ð Þ
—ð Þ ¼ e1 þ e2 þ e3 ð1:10aÞ
∂x1 ∂x2 ∂x3
while the higher-order operators are written as
2 2 2
∂ ∂ ∂
∇2 ð Þ ¼ —  —ð Þ ¼ ð Þþ 2ð Þþ 2ð Þ ð1:10bÞ
∂x21 ∂x2 ∂x3

and
 
∇4 ð Þ ¼ ∇ 2 ∇ 2 ð Þ
4 4 4
∂ ∂ ∂
¼ ð Þ þ 4ð Þ þ 4ð Þ
∂x1
4 ∂x2 ∂x3 ð1:10cÞ
4 4 4
∂ ∂ ∂
þ2 ð Þ þ 2 2 2 ð Þ þ 2 2 2 ð Þ:
∂x1 ∂x2
2 2 ∂x1 ∂x3 ∂x2 ∂x3

Both ∇2 ( ) and ∇4 ( ) are scalars and are used frequently in the following
chapters. These operators are general because the operation performed is indepen-
dent of any particular coordinate system or invariant. However, the forms given in
(1.10a), (1.10b), and (1.10c) are for Cartesian coordinates; for curvilinear coordi-
nates, such as cylindrical coordinates, the operators must be appropriately
transformed. This is developed in Sect. 7.4.
1.3 Scalar and Vector Fields 5

1.3.4 Divergence

The divergence of a vector field A, (1.2), is defined as

div A ¼ —  A
∂A1 ∂A2 ∂A3 ð1:11Þ
¼ þ þ ,
∂x1 ∂x2 ∂x3
which is a scalar conveniently written as ΔA.

1.3.5 Curl

Since there are two forms of vector multiplication, it is natural to expect another
derivative form of A. The curl of A is defined as

curl A ¼ —  A
 
 e1 e2 e3 
 
 ∂ð Þ ∂ð Þ ∂ð Þ  ð1:12Þ
¼ 

 ∂x1 ∂x2 ∂x3 
 A A2 A3 
1

in determinant form.

1.3.6 Integral Theorems

Two integral theorems relating vector fields are particularly useful for transforming
between contour, area, and volume integrals: Green’s theorem and the divergence
theorem.
First, considering two functions P(x, y) and Q(x, y), which are continuous and
have continuous first partial derivatives (C1 continuous) in a two-dimensional
domain D, Green’s theorem states that
þ ð  
∂Q ∂P
ðPdx þ QdyÞ ¼  dxdy, ð1:13Þ
C A ∂x ∂y

where A is a closed region within D bounded by C.


Then, considering a continuously differentiable vector point function G in D, the
divergence theorem states that
6 1 Introduction and Mathematical Preliminaries

ð ð
—  G dV ¼ n  G dA, ð1:14Þ
V A

where V is the volume bounded by the oriented surface A and n is the positive normal
to A.

1.4 Indicial Notation

One of the conveniences of modern treatments of the theory of elasticity is the use of
shorthand notation to facilitate the mathematical manipulation of lengthy equations.
Referring to the ordered triple representation for A in (1.3), the three Cartesian
components can be symbolized as Ai, where the subscript or index i is understood to
take the sequential values 1, 2, 3. If we have nine quantities, we may employ a
double-subscripted notation Dij, where i and j range from 1 to 3 in turn. Later, we
will associate these nine components with a higher form of a vector, called a tensor.
Further, we may have 27 quantities, Cijk, etc. While i and j range as stated, an
exception is made when two subscripts are identical, such as Djj. The Einstein
summation convention states that a subscript appearing twice is summed from 1 to
3. No subscript can appear more than twice in a single term. As an example, we have
the inner product, (1.5), rewritten as

X
3
Ai Bi ¼ Ai Bi
i¼1
ð1:15aÞ
¼ A1 B1 þ A2 B2 þ A3 B3 :

Also,

Djj ¼ D11 þ D22 þ D33 : ð1:15bÞ

It is apparent from the preceding examples that there are two distinct types of
indices. The first type appears only once in each term of the equation and ranges
from 1 to 3. It is called a free index. The second type appears twice in a single term
and is summed from 1 to 3. Since it is immaterial which letter is used in this context,
a repeated subscript is called a dummy index. That is, Dii ¼ Djj ¼ Dkk. From the
preceding discussion, it may be deduced that the number of individual terms
represented by a single product is 3k, where k is the number of free indices. There
are some situations where double subscripts occur but the summation convention is
not intended. This is indicated by enclosing the subscripts in parentheses [3]. For
example, the individual components D11, D22, and D33 could be represented by
D(ii)i ¼ 1, 2, 3. Also in the following sections, the summation convention is applied
only to the indices i, j, k unless otherwise stated.
1.5 Coordinate Rotation 7

The product of the three components of a vector is expressed by the Pi


convention:

Y
3
Ai ¼ A1 A2 A3 : ð1:16Þ
i¼1

Partial differentiation may also be abbreviated using the comma convention:

∂Ai
¼ Ai, j : ð1:17aÞ
∂xj

Since both i and j are free indices, (1.17a) represents 32 ¼ 9 quantities. With
repeated indices, the equation becomes

∂Ai
¼ Ai, i
∂xi ð1:17bÞ
¼ ΔA

which is the divergence as defined in (1.11). Further,

∂Dij
¼ Dij, j
∂xj ð1:17cÞ
¼ Di1, 1 þ Di2, 2 þ Di3, 3

that takes on 31 ¼ 3 values for each i ¼ 1, 2, 3. This example combines the


summation and comma conventions.

1.5 Coordinate Rotation

In Fig. 1.1 (see [2]), we show a position vector to point P, r, resolved into
components with respect to two different Cartesian systems, xi and x0i , having a
common origin. The unit vectors in the x0i system are indicated as e0i in the figure.
First, we consider the
 point P with coordinates
P(x1, x2, x3) ¼ P(xi) in the
unprimed system and P x01 ; x02 ; x03 ¼ P x0i in the primed system. The linear trans-
formation between the coordinates of P is given by

x01 ¼ α11 x1 þ α12 x2 þ α13 x3


x02 ¼ α21 x1 þ α22 x2 þ α23 x3 ð1:18aÞ
x03 ¼ α31 x1 þ α32 x2 þ α33 x3
8 1 Introduction and Mathematical Preliminaries

or

x0i ¼ αij xj ð1:18bÞ

using the summation convention. Each of the nine quantities α ij is the cosine of the
angle between the i th primed and the jth unprimed axis, that is,

 ∂xj 
αij ¼ cos x0i ; xj ¼ 0 ¼ e0i  ej ¼ cos e0i ; ej : ð1:19Þ
∂xi

The αij’s are known as direction cosines and are conveniently arranged in tabular
form for computation:

x1 x2 x3
x1′ α11 α12 α13
ð1:20Þ
x2′ α 21 α 22 α 23
x3′ α 31 α 32 α 33

It is emphasized that, in general, αij 6¼ αji. From a computational standpoint,


(1.19) indicates that expressing the unit vectors in the x0i coordinate system,e0i , in
terms of those in the xi system, ei, is tantamount to evaluating the corresponding αij
terms. A numerical example is given in Sect. 2.2.6 and 2.4.3.
We next consider the position vector r and recognize that the components are
related by (1.18a) and (1.18b). Conversely, any quantity which obeys this transfor-
mation law is a vector. This somewhat indirect definition of a vector proves to be
convenient for defining higher-order quantities, Cartesian tensors.
From a computational standpoint, it is often convenient to carry out the trans-
formations indicated in (1.18a) and (1.18b) in matrix form as

x0 ¼ Rx ð1:21Þ

in which
0 0 0
x0 ¼ fx1 , x2 , x3 g ð1:22aÞ
x ¼ fx1 , x2 , x3 g ð1:22bÞ
2 3
α11 α12 α13
R ¼ 4 α21 α22 α23 5: ð1:22cÞ
α31 α32 α33

The brackets { } indicate a column vector, the transpose { }T or b c is a row matrix,


and R may be called a rotation matrix.
1.7 Algebra of Cartesian Tensors 9

1.6 Cartesian Tensors

A tensor of order n is a set of 3n quantities which transform from one coordinate


system, xi, to another, x0i , by a specified law, as follows:

n Order Transformation law


0 Zero (scalar) A0i ¼ Ai
1 One (vector) A0i ¼ αij Aj
2 Two (dyadic) A0ij ¼ αik αjl Akl
3 Three A0ijk ¼ αil αjm αkn Almn
4 Four A0ijkl ¼ αim αjn αkp αlq Amnpq

Order zero and order one tensors are familiar physical quantities, whereas the
higher-order tensors are useful to describe physical quantities with a corresponding
number of associated directions. Second-order tensors (dyadics) are particularly
prevalent in elasticity, and the transformation may be carried out in a matrix format,
analogous to (1.21), as

A0 ¼ RART , ð1:23Þ

in which
2 3
A011 A012 A013
6 7
A0 ¼ 4 A021 A022 A023 5 ð1:24Þ
A031 A032 A033

and A is similar.
It may be helpful to visualize a tensor of order n as having n unit vectors or
directions associated with each component. Thus, a scalar has no directional asso-
ciation (isotropic), and a vector is directed in one direction. A second-order tensor
has two associated directions, perhaps one direction in which it acts and another
defining the surface on which it is acting.

1.7 Algebra of Cartesian Tensors

Tensor arithmetic and algebra are similar to matrix operations in regard to addition,
subtraction, equality, and scalar multiplication. Multiplication of two tensors of
order n and m produces a new tensor of order n + m. For example,

Ai Bjk ¼ C ijk : ð1:25Þ

For two repeated indices, the summation convention holds, as shown in (1.15b).
10 1 Introduction and Mathematical Preliminaries

1.8 Operational Tensors

Additional tensor operations are facilitated by the use of the Kronecker delta δij
defined such that

δij ¼ 1 if i¼j
ð1:26Þ
δij ¼ 0 if i 6¼ j

and the permutation tensor εijk defined such that

1
εijk ¼ ði  jÞðj  kÞðk  iÞ: ð1:27Þ
2
Thus,

εijk ¼ 0 If any two of i, j, k are equal


εijk ¼ 1 For an even permutation (forward on the number line 1, 2, 3, 1, 2, 3, . . .)
εijk ¼ 1 For an odd permutation (backward on the number line 3, 2, 1, 3, 2, 1, . . .)

Hence, ε112 ¼ 0, ε231 ¼ þ1, and ε321 ¼ 1.


The Kronecker delta δij is used to change the subscripts in a tensor by multipli-
cation, as illustrated in the following:

δij Ai ¼ δ1j A1 þ δ2j A2 þ δ3j A3 ¼ Aj


δij Djk ¼ δi1 D1k þ δi2 D2k þ δi3 D3k ¼ Dik ð1:28Þ
δij Cijk ¼ δ11 C11k þ δ22 C22k þ δ33 C 33k ¼ C iik ¼ Ak :

The last illustration, in which two subscripts in Cijk were made identical by δij,
results in Cijk being changed from a third- to a first-order tensor. This is known as
contraction and generally reduces the order of the original tensor by two.
The Kronecker delta δij is also useful in vector algebra and for coordinate trans-
formations. Starting with the dot product of two unit vectors

ei  ej ¼ δij ð1:29Þ

and seeking the component of a vector A ¼ Aiei in the j-direction, Aj, we do the
following:

ej  A ¼ ej  ei Ai
¼ δji Ai ð1:30Þ
¼ Aj :
1.8 Operational Tensors 11

Next, we consider (1.29) for transformed coordinates:

δij ¼ e0i  e0j



¼ ðαik ek Þ  αjl el
¼ αik αjl ek  el ð1:31Þ
¼ αik αjl δkl
¼ αik αjk :

Equation (1.31) is useful for demonstrating some important properties of direc-


tion cosines.
First, taking i ¼ j and summing only on k, we get

1 ¼ αðiÞk αðiÞk
ð1:32Þ
¼ α2i1 þ α2i2 þ α2i3 :

Expanding (1.32), we have

i ¼ 1 : α211 þ α212 þ α213 ¼ 1


i ¼ 2 : α221 þ α222 þ α223 ¼ 1
i ¼ 3 : α231 þ α232 þ α233 ¼ 1,

which is the normality property of direction cosines.


Then, taking i 6¼ j, we get

0 ¼ αik αjk
ð1:33Þ
¼ αi1 αj1 þ αi2 αj2 þ αi3 αj3 :

Expanding (1.33), we have

i ¼ 1, j ¼ 2 : α11 α21 þ α12 α22 þ α13 α23 ¼ 0


i ¼ 1, j ¼ 3 : α11 α31 þ α12 α32 þ α13 α33 ¼ 0
i ¼ 2, j ¼ 1 : α21 α11 þ α22 α12 þ α23 α13 ¼ 0
i ¼ 2, j ¼ 3 : α21 α31 þ α22 α32 þ α23 α33 ¼ 0
i ¼ 3, j ¼ 1 : α31 α11 þ α32 α12 þ α33 α13 ¼ 0
i ¼ 3, j ¼ 2 : α31 α21 þ α32 α22 þ α33 α23 ¼ 0,

which is the orthogonality property of direction cosines.


12 1 Introduction and Mathematical Preliminaries

The permutation tensor εijk is useful for vector cross-product operations. If we take

εijk Aj Bk ei ¼ ðε123 A2 B3 þ ε132 A3 B2 Þe1 þ ðε213 A1 B3 þ ε231 A3 B2 Þe2


þðε312 A1 B2 þ ε321 A2 B1 Þe3
ð1:34Þ
εijk Aj Bk ei ¼ ðA2 B3  A3 B2 Þe1 þ ðA3 B1  A1 B3 Þe2
þðA1 B2  A2 B1 Þe3 ,

we obtain an expression which is identical to (1.6a) and (1.6b). Thus, εijkAjBk gives
the components of A  B.

1.9 Computational Examples

To illustrate computations and manipulations using Cartesian tensors, we present the


following illustrations:
1. Show that δijδjk ¼ δik:
Expanding δijδjk, we get

δij δjk ¼ δi1 δ1k þ δi2 δ2k þ δi3 δ3k ¼ 1  δik for a selected
:
i ¼ 1, 2, or 3

2. Show that εijkAjAk ¼ 0:


Expanding εijkAjAk, we find

εijk Aj Ak ¼ ε123 A2 A3 þ ε132 A3 A2 þ ε231 A3 A1 þ ε213 A1 A3


þ ε312 A1 A2 þ ε321 A2 A1
¼ þ1  1 þ 1  1 þ 1  1
¼ 0:

3. Prove that the product of two first-order tensors is a second-order tensor:


Let Ai and Bj be two first-order tensors and Cij be their product; then

Ai ¼ αik Ak
Bj ¼ αjl Bl
Ai Bj ¼ αik αjl Ak Bl ¼ αik αjl Ckl ¼ Cij

so that Cij transforms as a second-order tensor.


1.11 Introduction to MuPAD 13

1.10 Fundamentals of MATLAB

The core ideas of MATLAB are matrices or arrays that are the basic computation
units. As a computational tool, MATLAB has been widely used in the field of
scientific computation. In this book, we use MATLAB to demonstrate some com-
putational applications to elasticity. Several examples solved analytically will be
solved using MATLAB. As a useful tool to implement and test algorithms, codes
written in MATLAB can be translated to other codes such as Fortran or C.
2 tensor σ, defined
In the computational framework, the components of the stress 3 in
σ 11 σ 12 σ 13
Section 2.2.2, can be represented in a 3  3 matrix σ ¼ 4 σ 21 σ 22 σ 23 5. The
σ 31 σ 32 σ 33
matrix notation of the stress tensor makes it easy for computational implementation.
Similarly, vectors such as the normal n can be denoted as a 31 matrix n ¼
{n1, n2, n3}.
MATLAB offers a very intuitive way of data representation. The stress tensor can
be input as

sig = [sig_11, sig_12, sig_13;


sig_21, sig_22, sig_23;
sig_31, sig_32, sig_33];

where each element is the corresponding component of the stress tensor. The normal
vector can be input as

n = [n_1; n_2; n_3];

where n_1 to n_3 are the corresponding vector components.


Frequently used operations are the matrix product and dot product. MATLAB
uses “*” for the matrix product. For example, A*B¼C can be input as C ¼ A * B.
The dot product between the two vectors, e.g., v1*v2 ¼ v3, can be calculated by
v3 ¼ v1 * v2. Therefore, in a state of stress with stress tensor sig, the traction force
on the plane with a normal vector n can be calculated by sig*n. For other
command usage, please refer to Appendix I.

1.11 Introduction to MuPAD

Many elasticity problems involve presenting analytical solutions. Although analyt-


ical solutions are not obtainable for all problems, they provide a useful way to verify
computational results.
There are two ways to obtain analytical solutions in MATLAB. One is to use the
“symbolic toolbox” and the other is to use the embedded package “MuPAD.”
14 1 Introduction and Mathematical Preliminaries

Although the commands are different between the two, both tools interface within
MATLAB and can exchange data and variables via the MATLAB workspace.
In the command window, simply type in “mupad” and the MuPAD interface
window will pop out. An important feature of the MuPAD is the visualization of the
symbols and equations. This is one of the major advantages of the “symbolic
toolbox.” For example, create a symbol “σ” and use “sig” for representation in the
scripts.

In this case, we can use sig for all the symbolic calculations, but in the final
results, symbol “σ” will be shown. Therefore, if we use “sig” for computation, the
command prompt output will show symbol “σ” rather than “sig”. This is useful for
visualizing all the equations. For other command usage, please refer to Appendix II.
In the following chapters, we demonstrate the use of MuPAD using typical
examples in elasticity.

1.12 Exercises

1.1 Show that (after [2]):


(a) δijδij ¼ 3
(b) εijkεkji ¼ 6
(c) αijbjk ¼ αinbnk
(d) δijδjkδki ¼ 3
(e) δijCjkl ¼ Cikl
1.2 Prove that if δij ¼ δ’ij, then δij is a second-order tensor (after [2]).
1.3 Prove that the product of a first- and a second-order tensor is a third-order tensor.
1.4 Two first-order tensors are related by

Ai ¼ Cij Bj :

Prove that Cij is a second-order tensor.


1.5 Show that if Bi is a first-order tensor, Bi,j is a second-order tensor [2].
1.6 If a square matrix C has the property

C 1 ¼ C T ,

it is said to be orthogonal. This property is used in the derivation of (1.23). Show


that the matrix of direction cosines R as defined in (1.22c) is orthogonal.
1.7 Write the operator —2( ) defined in (1.10b) in indicial notation.
References 15

1.8 Write the divergence theorem defined in (1.14) in indicial notation.


1.9 Show that curl A ¼ —  A ¼ εijkAj,iek.

References

1. Westergaard HM (1964) Theory of elasticity and plasticity. Dover Publications, Inc., New York
2. Tauchert TR (1974) Energy principles in structural mechanics. McGraw-Hill Book Company
Inc., New York
3. Ma Y, Desai CS (1990) Alternative definition of finite strains. J Eng Mech, ASCE 116
(4):901–919

Vous aimerez peut-être aussi