04.06.2013 Views

C:\book\booktex\start.DVI 12

C:\book\booktex\start.DVI 12

C:\book\booktex\start.DVI 12

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Introduction to<br />

Tensor Calculus<br />

and<br />

Continuum Mechanics<br />

by J.H. Heinbockel<br />

Department of Mathematics and Statistics<br />

Old Dominion University


PREFACE<br />

This is an introductory text which presents fundamental concepts from the subject<br />

areas of tensor calculus, differential geometry and continuum mechanics. The material<br />

presented is suitable for a two semester course in applied mathematics and is flexible<br />

enough to be presented to either upper level undergraduate or beginning graduate students<br />

majoring in applied mathematics, engineering or physics. The presentation assumes the<br />

students have some knowledge from the areas of matrix theory, linear algebra and advanced<br />

calculus. Each section includes many illustrative worked examples. At the end of each<br />

section there is a large collection of exercises which range in difficulty. Many new ideas<br />

are presented in the exercises and so the students should be encouraged to read all the<br />

exercises.<br />

The purpose of preparing these notes is to condense into an introductory text the basic<br />

definitions and techniques arising in tensor calculus, differential geometry and continuum<br />

mechanics. In particular, the material is presented to (i) develop a physical understanding<br />

of the mathematical concepts associated with tensor calculus and (ii) develop the basic<br />

equations of tensor calculus, differential geometry and continuum mechanics which arise<br />

in engineering applications. From these basic equations one can go on to develop more<br />

sophisticated models of applied mathematics. The material is presented in an informal<br />

manner and uses mathematics which minimizes excessive formalism.<br />

The material has been divided into two parts. The first part deals with an introduction<br />

to tensor calculus and differential geometry which covers such things as the indicial<br />

notation, tensor algebra, covariant differentiation, dual tensors, bilinear and multilinear<br />

forms, special tensors, the Riemann Christoffel tensor, space curves, surface curves, curvature<br />

and fundamental quadratic forms. The second part emphasizes the application of<br />

tensor algebra and calculus to a wide variety of applied areas from engineering and physics.<br />

The selected applications are from the areas of dynamics, elasticity, fluids and electromagnetic<br />

theory. The continuum mechanics portion focuses on an introduction of the basic<br />

concepts from linear elasticity and fluids. The Appendix A contains units of measurements<br />

from the Système International d’Unitès along with some selected physical constants. The<br />

Appendix B contains a listing of Christoffel symbols of the second kind associated with<br />

various coordinate systems. The Appendix C is a summary of useful vector identities.<br />

J.H. Heinbockel, 1996


Copyright c○1996 by J.H. Heinbockel. All rights reserved.<br />

Reproduction and distribution of these notes is allowable provided it is for non-profit<br />

purposes only.


INTRODUCTION TO<br />

TENSOR CALCULUS<br />

AND<br />

CONTINUUM MECHANICS<br />

PART 1: INTRODUCTION TO TENSOR CALCULUS<br />

§1.1 INDEX NOTATION . . . . . . . . . . . . . . . . . . 1<br />

Exercise 1.1 . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

§1.2 TENSOR CONCEPTS AND TRANSFORMATIONS . . . . 35<br />

Exercise 1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

§1.3 SPECIAL TENSORS . . . . . . . . . . . . . . . . . . 65<br />

Exercise 1.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

§1.4 DERIVATIVE OF A TENSOR . . . . . . . . . . . . . . 108<br />

Exercise 1.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>12</strong>3<br />

§1.5 DIFFERENTIAL GEOMETRY AND RELATIVITY . . . . <strong>12</strong>9<br />

Exercise 1.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . 162<br />

PART 2: INTRODUCTION TO CONTINUUM MECHANICS<br />

§2.1 TENSOR NOTATION FOR VECTOR QUANTITIES . . . . 171<br />

Exercise 2.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 182<br />

§2.2 DYNAMICS . . . . . . . . . . . . . . . . . . . . . . 187<br />

Exercise 2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 206<br />

§2.3 BASIC EQUATIONS OF CONTINUUM MECHANICS . . . 211<br />

Exercise 2.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 238<br />

§2.4 CONTINUUM MECHANICS (SOLIDS) . . . . . . . . . 243<br />

Exercise 2.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 272<br />

§2.5 CONTINUUM MECHANICS (FLUIDS) . . . . . . . . . 282<br />

Exercise 2.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . 317<br />

§2.6 ELECTRIC AND MAGNETIC FIELDS . . . . . . . . . . 325<br />

Exercise 2.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . 347<br />

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . 352<br />

APPENDIX A UNITS OF MEASUREMENT . . . . . . . 353<br />

APPENDIX B CHRISTOFFEL SYMBOLS OF SECOND KIND 355<br />

APPENDIX C VECTOR IDENTITIES . . . . . . . . . . 362<br />

INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . 363


PART 1: INTRODUCTION TO TENSOR CALCULUS<br />

A scalar field describes a one-to-one correspondence between a single scalar number and a point. An ndimensional<br />

vector field is described by a one-to-one correspondence between n-numbers and a point. Let us<br />

generalize these concepts by assigning n-squared numbers to a single point or n-cubed numbers to a single<br />

point. When these numbers obey certain transformation laws they become examples of tensor fields. In<br />

general, scalar fields are referred to as tensor fields of rank or order zero whereas vector fields are called<br />

tensor fields of rank or order one.<br />

Closely associated with tensor calculus is the indicial or index notation. In section 1 the indicial<br />

notation is defined and illustrated. We also define and investigate scalar, vector and tensor fields when they<br />

are subjected to various coordinate transformations. It turns out that tensors have certain properties which<br />

are independent of the coordinate system used to describe the tensor. Because of these useful properties,<br />

we can use tensors to represent various fundamental laws occurring in physics, engineering, science and<br />

mathematics. These representations are extremely useful as they are independent of the coordinate systems<br />

considered.<br />

§1.1 INDEX NOTATION<br />

Two vectors A and B can be expressed in the component form<br />

A = A1 e1 + A2 e2 + A3 e3 and B = B1 e1 + B2 e2 + B3 e3,<br />

where e1, e2 and e3 are orthogonal unit basis vectors. Often when no confusion arises, the vectors A and<br />

B are expressed for brevity sake as number triples. For example, we can write<br />

A =(A1, A2, A3) and B =(B1, B2, B3)<br />

where it is understood that only the components of the vectors A and B are given. The unit vectors would<br />

be represented<br />

e1 =(1, 0, 0), e2 =(0, 1, 0), e3 =(0, 0, 1).<br />

A still shorter notation, depicting the vectors A and B is the index or indicial notation. In the index notation,<br />

the quantities<br />

Ai, i =1, 2, 3 and Bp, p =1, 2, 3<br />

represent the components of the vectors A and B. This notation focuses attention only on the components of<br />

the vectors and employs a dummy subscript whose range over the integers is specified. The symbol Ai refers<br />

to all of the components of the vector A simultaneously. The dummy subscript i can have any of the integer<br />

values 1, 2or3. For i = 1 we focus attention on the A1 component of the vector A. Setting i =2focuses<br />

attention on the second component A2 of the vector A and similarly when i = 3 we can focus attention on<br />

the third component of A. The subscript i is a dummy subscript and may be replaced by another letter, say<br />

p, so long as one specifies the integer values that this dummy subscript can have.<br />

1


2<br />

It is also convenient at this time to mention that higher dimensional vectors may be defined as ordered<br />

n−tuples. For example, the vector<br />

X =(X1,X2,...,XN )<br />

with components Xi, i=1, 2,...,N is called a N−dimensional vector. Another notation used to represent<br />

this vector is<br />

X = X1 e1 + X2 e2 + ···+ XN eN<br />

where<br />

e1, e2,..., eN<br />

are linearly independent unit base vectors. Note that many of the operations that occur in the use of the<br />

index notation apply not only for three dimensional vectors, but also for N−dimensional vectors.<br />

In future sections it is necessary to define quantities which can be represented by a letter with subscripts<br />

or superscripts attached. Such quantities are referred to as systems. When these quantities obey certain<br />

transformation laws they are referred to as tensor systems. For example, quantities like<br />

A k ij<br />

e ijk<br />

δij<br />

δ j<br />

i<br />

The subscripts or superscripts are referred to as indices or suffixes. When such quantities arise, the indices<br />

must conform to the following rules:<br />

1. They are lower case Latin or Greek letters.<br />

2. The letters at the end of the alphabet (u, v, w, x, y, z) are never employed as indices.<br />

The number of subscripts and superscripts determines the order of the system. A system with one index<br />

is a first order system. A system with two indices is called a second order system. In general, a system with<br />

N indices is called a Nth order system. A system with no indices is called a scalar or zeroth order system.<br />

The type of system depends upon the number of subscripts or superscripts occurring in an expression.<br />

For example, A i jk and B m st , (all indices range 1 to N), are of the same type because they have the same<br />

number of subscripts and superscripts. In contrast, the systems A i jk and C mn<br />

p are not of the same type<br />

because one system has two superscripts and the other system has only one superscript. For certain systems<br />

the number of subscripts and superscripts is important. In other systems it is not of importance. The<br />

meaning and importance attached to sub- and superscripts will be addressed later in this section.<br />

In the use of superscripts one must not confuse “powers ”of a quantity with the superscripts. For<br />

example, if we replace the independent variables (x, y, z) bythesymbols(x 1 , x 2 , x 3 ), then we are letting<br />

y = x 2 where x 2 is a variable and not x raised to a power. Similarly, the substitution z = x 3 is the<br />

replacement of z by the variable x 3 and this should not be confused with x raised to a power. In order to<br />

write a superscript quantity to a power, use parentheses. For example, (x 2 ) 3 is the variable x 2 cubed. One<br />

of the reasons for introducing the superscript variables is that many equations of mathematics and physics<br />

can be made to take on a concise and compact form.<br />

There is a range convention associated with the indices. This convention states that whenever there<br />

is an expression where the indices occur unrepeated it is to be understood that each of the subscripts or<br />

superscripts can take on any of the integer values 1, 2,...,N where N is a specified integer. For example,<br />

A i<br />

Bj<br />

aij.


the Kronecker delta symbol δij, defined by δij =1ifi = j and δij =0fori= j, withi, j ranging over the<br />

values 1,2,3, represents the 9 quantities<br />

δ11 =1<br />

δ21 =0<br />

δ31 =0<br />

δ<strong>12</strong> =0<br />

δ22 =1<br />

δ32 =0<br />

δ13 =0<br />

δ23 =0<br />

δ33 =1.<br />

The symbol δij refers to all of the components of the system simultaneously. As another example, consider<br />

the equation<br />

em · en = δmn m, n =1, 2, 3 (1.1.1)<br />

the subscripts m, n occur unrepeated on the left side of the equation and hence must also occur on the right<br />

hand side of the equation. These indices are called “free ”indices and can take on any of the values 1, 2or3<br />

as specified by the range. Since there are three choices for the value for m and three choices for a value of<br />

n we find that equation (1.1.1) represents nine equations simultaneously. These nine equations are<br />

e1 · e1 =1<br />

e2 · e1 =0<br />

e3 · e1 =0<br />

Symmetric and Skew-Symmetric Systems<br />

e1 · e2 =0<br />

e2 · e2 =1<br />

e3 · e2 =0<br />

e1 · e3 =0<br />

e2 · e3 =0<br />

e3 · e3 =1.<br />

A system defined by subscripts and superscripts ranging over a set of values is said to be symmetric<br />

in two of its indices if the components are unchanged when the indices are interchanged. For example, the<br />

third order system Tijk is symmetric in the indices i and k if<br />

Tijk = Tkji for all values of i, j and k.<br />

A system defined by subscripts and superscripts is said to be skew-symmetric in two of its indices if the<br />

components change sign when the indices are interchanged. For example, the fourth order system Tijkl is<br />

skew-symmetric in the indices i and l if<br />

Tijkl = −Tljki for all values of ijk and l.<br />

As another example, consider the third order system aprs, p,r,s =1, 2, 3 which is completely skewsymmetric<br />

in all of its indices. We would then have<br />

aprs = −apsr = aspr = −asrp = arsp = −arps.<br />

It is left as an exercise to show this completely skew- symmetric systems has 27 elements, 21 of which are<br />

zero. The 6 nonzero elements are all related to one another thru the above equations when (p, r, s) =(1, 2, 3).<br />

This is expressed as saying that the above system has only one independent component.<br />

3


4<br />

Summation Convention<br />

The summation convention states that whenever there arises an expression where there is an index which<br />

occurs twice on the same side of any equation, or term within an equation, it is understood to represent a<br />

summation on these repeated indices. The summation being over the integer values specified by the range. A<br />

repeated index is called a summation index, while an unrepeated index is called a free index. The summation<br />

convention requires that one must never allow a summation index to appear more than twice in any given<br />

expression. Because of this rule it is sometimes necessary to replace one dummy summation symbol by<br />

some other dummy symbol in order to avoid having three or more indices occurring on the same side of<br />

the equation. The index notation is a very powerful notation and can be used to concisely represent many<br />

complex equations. For the remainder of this section there is presented additional definitions and examples<br />

to illustrated the power of the indicial notation. This notation is then employed to define tensor components<br />

and associated operations with tensors.<br />

EXAMPLE 1.1-1 The two equations<br />

y1 = a11x1 + a<strong>12</strong>x2<br />

y2 = a21x1 + a22x2<br />

can be represented as one equation by introducing a dummy index, say k, and expressing the above equations<br />

as<br />

yk = ak1x1 + ak2x2, k =1, 2.<br />

The range convention states that k is free to have any one of the values 1 or 2, (k is a free index). This<br />

equation can now be written in the form<br />

yk =<br />

2<br />

i=1<br />

akixi = ak1x1 + ak2x2<br />

where i is the dummy summation index. When the summation sign is removed and the summation convention<br />

is adopted we have<br />

yk = akixi<br />

i, k =1, 2.<br />

Since the subscript i repeats itself, the summation convention requires that a summation be performed by<br />

letting the summation subscript take on the values specified by the range and then summing the results.<br />

The index k which appears only once on the left and only once on the right hand side of the equation is<br />

called a free index. It should be noted that both k and i are dummy subscripts and can be replaced by other<br />

letters. For example, we can write<br />

yn = anmxm<br />

n, m =1, 2<br />

where m is the summation index and n is the free index. Summing on m produces<br />

yn = an1x1 + an2x2<br />

and letting the free index n take on the values of 1 and 2 we produce the original two equations.


EXAMPLE 1.1-2. For yi = aijxj, i,j=1, 2, 3andxi = bijzj, i,j=1, 2, 3solvefortheyvariables in<br />

terms of the z variables.<br />

Solution: In matrix form the given equations can be expressed:<br />

⎛<br />

⎝ y1<br />

⎞ ⎛<br />

⎠ = ⎝ a11<br />

⎞ ⎛<br />

a<strong>12</strong> a13<br />

⎠ ⎝ x1<br />

⎞ ⎛<br />

⎠ and ⎝ x1<br />

⎞ ⎛<br />

⎠ = ⎝ b11<br />

⎞ ⎛<br />

b<strong>12</strong> b13<br />

⎠ ⎝ z1<br />

⎞<br />

⎠ .<br />

y2<br />

y3<br />

a21 a22 a23<br />

a31 a32 a33<br />

Now solve for the y variables in terms of the z variables and obtain<br />

⎛<br />

⎝ y1<br />

⎞ ⎛<br />

⎠ = ⎝ a11 a<strong>12</strong><br />

⎞ ⎛<br />

a13<br />

⎠<br />

y2<br />

y3<br />

x2<br />

x3<br />

a21 a22 a23<br />

a31 a32 a33<br />

x2<br />

x3<br />

⎝ b11 b<strong>12</strong> b13<br />

b21 b22 b23<br />

b31 b32 b33<br />

b21 b22 b23<br />

b31 b32 b33<br />

⎞ ⎛<br />

⎠<br />

⎝ z1<br />

z2<br />

The index notation employs indices that are dummy indices and so we can write<br />

z3<br />

⎞<br />

⎠ .<br />

yn = anmxm, n,m =1, 2, 3 and xm = bmjzj, m,j =1, 2, 3.<br />

Here we have purposely changed the indices so that when we substitute for xm, from one equation into the<br />

other, a summation index does not repeat itself more than twice. Substituting we find the indicial form of<br />

the above matrix equation as<br />

yn = anmbmjzj, m,n,j =1, 2, 3<br />

where n is the free index and m, j are the dummy summation indices. It is left as an exercise to expand<br />

both the matrix equation and the indicial equation and verify that they are different ways of representing<br />

the same thing.<br />

EXAMPLE 1.1-3. The dot product of two vectors Aq, q=1, 2, 3andBj, j=1, 2, 3 can be represented<br />

with the index notation by the product AiBi = AB cos θ i =1, 2, 3, A = | A|, B = | B|. Since the<br />

subscript i is repeated it is understood to represent a summation index.<br />

specified, there results<br />

A1B1 + A2B2 + A3B3 = AB cos θ.<br />

Summing on i over the range<br />

Observe that the index notation employs dummy indices. At times these indices are altered in order to<br />

conform to the above summation rules, without attention being brought to the change. As in this example,<br />

the indices q and j are dummy indices and can be changed to other letters if one desires. Also, in the future,<br />

if the range of the indices is not stated it is assumed that the range is over the integer values 1, 2and3.<br />

To systems containing subscripts and superscripts one can apply certain algebraic operations. We<br />

present in an informal way the operations of addition, multiplication and contraction.<br />

z2<br />

z3<br />

5


6<br />

Addition, Multiplication and Contraction<br />

The algebraic operation of addition or subtraction applies to systems of the same type and order. That<br />

is we can add or subtract like components in systems. For example, the sum of A i jk and B i jk is again a<br />

system of the same type and is denoted by C i jk = A i jk + B i jk, where like components are added.<br />

The product of two systems is obtained by multiplying each component of the first system with each<br />

component of the second system. Such a product is called an outer product. The order of the resulting<br />

product system is the sum of the orders of the two systems involved in forming the product. For example,<br />

if A i j is a second order system and B mnl is a third order system, with all indices having the range 1 to N,<br />

then the product system is fifth order and is denoted C imnl<br />

j = A i jB mnl . The product system represents N 5<br />

terms constructed from all possible products of the components from A i j with the components from B mnl .<br />

The operation of contraction occurs when a lower index is set equal to an upper index and the summation<br />

convention is invoked. For example, if we have a fifth order system C imnl<br />

j<br />

we form the system<br />

C mnl = C jmnl<br />

j<br />

= C 1mnl<br />

1<br />

+ C 2mnl<br />

2<br />

+ ···+ C Nmnl<br />

N .<br />

and we set i = j and sum, then<br />

Here the symbol C mnl is used to represent the third order system that results when the contraction is<br />

performed. Whenever a contraction is performed, the resulting system is always of order 2 less than the<br />

original system. Under certain special conditions it is permissible to perform a contraction on two lower case<br />

indices. These special conditions will be considered later in the section.<br />

The above operations will be more formally defined after we have explained what tensors are.<br />

The e-permutation symbol and Kronecker delta<br />

Two symbols that are used quite frequently with the indicial notation are the e-permutation symbol<br />

and the Kronecker delta. The e-permutation symbol is sometimes referred to as the alternating tensor. The<br />

e-permutation symbol, as the name suggests, deals with permutations. A permutation is an arrangement of<br />

things. When the order of the arrangement is changed, a new permutation results. A transposition is an<br />

interchange of two consecutive terms in an arrangement. As an example, let us change the digits 1 2 3 to<br />

3 2 1 by making a sequence of transpositions. Starting with the digits in the order 1 2 3 we interchange 2 and<br />

3 (first transposition) to obtain 1 3 2. Next, interchange the digits 1 and 3 ( second transposition) to obtain<br />

3<strong>12</strong>. Finally, interchange the digits 1 and 2 (third transposition) to achieve 3 2 1. Here the total number<br />

of transpositions of 1 2 3 to 3 2 1 is three, an odd number. Other transpositions of 1 2 3 to 3 2 1 can also be<br />

written. However, these are also an odd number of transpositions.


EXAMPLE 1.1-4. The total number of possible ways of arranging the digits 1 2 3 is six. We have<br />

three choices for the first digit. Having chosen the first digit, there are only two choices left for the second<br />

digit. Hence the remaining number is for the last digit. The product (3)(2)(1) = 3! = 6 is the number of<br />

permutations of the digits 1, 2 and 3. These six permutations are<br />

1 2 3 even permutation<br />

1 3 2 odd permutation<br />

3 1 2 even permutation<br />

3 2 1 odd permutation<br />

2 3 1 even permutation<br />

2 1 3 odd permutation.<br />

Here a permutation of 1 2 3 is called even or odd depending upon whether there is an even or odd number<br />

of transpositions of the digits. A mnemonic device to remember the even and odd permutations of <strong>12</strong>3<br />

is illustrated in the figure 1.1-1. Note that even permutations of <strong>12</strong>3 are obtained by selecting any three<br />

consecutive numbers from the sequence <strong>12</strong>3<strong>12</strong>3 and the odd permutations result by selecting any three<br />

consecutive numbers from the sequence 321321.<br />

Figure 1.1-1. Permutations of <strong>12</strong>3.<br />

In general, the number of permutations of n things taken m at a time is given by the relation<br />

P (n, m) =n(n − 1)(n − 2) ···(n − m +1).<br />

By selecting a subset of m objects from a collection of n objects, m ≤ n, without regard to the ordering is<br />

called a combination of n objects taken m at a time. For example, combinations of 3 numbers taken from<br />

the set {1, 2, 3, 4} are (<strong>12</strong>3), (<strong>12</strong>4), (134), (234). Note that ordering of a combination is not considered. That<br />

is, the permutations (<strong>12</strong>3), (132), (231), (213), (3<strong>12</strong>), (321) are considered equal. In general, the number of<br />

<br />

n<br />

<br />

n!<br />

combinations of n objects taken m at a time is given by C(n, m) = =<br />

m m!(n − m)! where n<br />

m are the<br />

binomial coefficients which occur in the expansion<br />

(a + b) n =<br />

n<br />

m=0<br />

<br />

n<br />

<br />

a<br />

m<br />

n−m b m .<br />

7


8<br />

The definition of permutations can be used to define the e-permutation symbol.<br />

Definition: (e-Permutation symbol or alternating tensor)<br />

The e-permutation symbol is defined<br />

e ijk...l ⎧<br />

⎪⎨ 1 if ijk...l is an even permutation of the integers <strong>12</strong>3 ...n<br />

= eijk...l = −1<br />

⎪⎩<br />

0<br />

if ijk...l is an odd permutation of the integers <strong>12</strong>3 ...n<br />

in all other cases<br />

EXAMPLE 1.1-5. Find e6<strong>12</strong>453.<br />

Solution: To determine whether 6<strong>12</strong>453 is an even or odd permutation of <strong>12</strong>3456 we write down the given<br />

numbers and below them we write the integers 1 through 6. Like numbers are then connected by a line and<br />

we obtain figure 1.1-2.<br />

Figure 1.1-2. Permutations of <strong>12</strong>3456.<br />

In figure 1.1-2, there are seven intersections of the lines connecting like numbers. The number of<br />

intersections is an odd number and shows that an odd number of transpositions must be performed. These<br />

results imply e6<strong>12</strong>453 = −1.<br />

Another definition used quite frequently in the representation of mathematical and engineering quantities<br />

is the Kronecker delta which we now define in terms of both subscripts and superscripts.<br />

Definition: (Kronecker delta) The Kronecker delta is defined:<br />

δij = δ j<br />

i =<br />

1 if i equals j<br />

0 if i is different from j


EXAMPLE 1.1-6. Some examples of the e−permutation symbol and Kronecker delta are:<br />

e<strong>12</strong>3 = e <strong>12</strong>3 =+1<br />

e213 = e 213 = −1<br />

e1<strong>12</strong> = e 1<strong>12</strong> =0<br />

δ 1 1 =1<br />

δ 1 2 =0<br />

δ 1 3 =0<br />

δ<strong>12</strong> =0<br />

δ22 =1<br />

δ32 =0.<br />

EXAMPLE 1.1-7. When an index of the Kronecker delta δij is involved in the summation convention,<br />

the effect is that of replacing one index with a different index. For example, let aij denote the elements of an<br />

N × N matrix. Here i and j are allowed to range over the integer values 1, 2,...,N. Consider the product<br />

aijδik<br />

where the range of i, j, k is 1, 2,...,N. The index i is repeated and therefore it is understood to represent<br />

a summation over the range. The index i is called a summation index. The other indices j and k are free<br />

indices. They are free to be assigned any values from the range of the indices. They are not involved in any<br />

summations and their values, whatever you choose to assign them, are fixed. Let us assign a value of j and<br />

k to the values of j and k. The underscore is to remind you that these values for j and k are fixed and not<br />

to be summed. When we perform the summation over the summation index i we assign values to i from the<br />

range and then sum over these values. Performing the indicated summation we obtain<br />

aijδik = a1jδ1k + a2jδ2k + ···+ akjδkk + ···+ aNjδNk.<br />

In this summation the Kronecker delta is zero everywhere the subscripts are different and equals one where<br />

the subscripts are the same. There is only one term in this summation which is nonzero. It is that term<br />

where the summation index i was equal to the fixed value k This gives the result<br />

akjδkk = akj<br />

where the underscore is to remind you that the quantities have fixed values and are not to be summed.<br />

Dropping the underscores we write<br />

aijδik = akj<br />

Here we have substituted the index i by k and so when the Kronecker delta is used in a summation process<br />

it is known as a substitution operator. This substitution property of the Kronecker delta can be used to<br />

simplify a variety of expressions involving the index notation. Some examples are:<br />

Bijδjs = Bis<br />

δjkδkm = δjm<br />

eijkδimδjnδkp = emnp.<br />

Some texts adopt the notation that if indices are capital letters, then no summation is to be performed.<br />

For example,<br />

aKJδKK = aKJ<br />

9


10<br />

as δKK represents a single term because of the capital letters. Another notation which is used to denote no<br />

summation of the indices is to put parenthesis about the indices which are not to be summed. For example,<br />

a (k)jδ (k)(k) = akj,<br />

since δ (k)(k) represents a single term and the parentheses indicate that no summation is to be performed.<br />

At any time we may employ either the underscore notation, the capital letter notation or the parenthesis<br />

notation to denote that no summation of the indices is to be performed. To avoid confusion altogether, one<br />

can write out parenthetical expressions such as “(no summation on k)”.<br />

EXAMPLE 1.1-8. In the Kronecker delta symbol δi j we set j equal to i and perform a summation. This<br />

operation is called a contraction. There results δi i , which is to be summed over the range of the index i.<br />

Utilizing the range 1, 2,...,N we have<br />

δ i i = δ 1 1 + δ 2 2 + ···+ δ N N<br />

δ i i =1+1+···+1<br />

δ i i<br />

= N.<br />

In three dimension we have δi j ,i,j=1, 2, 3and<br />

δ k k = δ1 1 + δ2 2 + δ3 3 =3.<br />

In certain circumstances the Kronecker delta can be written with only subscripts. For example,<br />

δij, i,j =1, 2, 3. We shall find that these circumstances allow us to perform a contraction on the lower<br />

indices so that δii =3.<br />

EXAMPLE 1.1-9. The determinant of a matrix A =(aij) can be represented in the indicial notation.<br />

Employing the e-permutation symbol the determinant of an N × N matrix is expressed<br />

|A| = eij...ka1ia2j ···aNk<br />

where eij...k is an Nth order system. In the special case of a 2 × 2 matrix we write<br />

|A| = eija1ia2j<br />

where the summation is over the range 1,2 and the e-permutation symbol is of order 2. In the special case<br />

of a 3 × 3matrixwehave<br />

<br />

<br />

<br />

a11 a<strong>12</strong> a13 <br />

<br />

|A| = <br />

a21 a22 a23 <br />

<br />

<br />

= eijkai1aj2ak3 = eijka1ia2ja3k<br />

a31 a32 a33<br />

where i, j, k are the summation indices and the summation is over the range 1,2,3. Here eijk denotes the<br />

e-permutation symbol of order 3. Note that by interchanging the rows of the 3 × 3 matrix we can obtain


more general results. Consider (p, q, r) as some permutation of the integers (1, 2, 3), and observe that the<br />

determinant can be expressed <br />

<br />

ap1<br />

∆= aq1 <br />

<br />

ap2<br />

aq2<br />

<br />

ap3 <br />

<br />

aq3 <br />

= eijkapiaqjark.<br />

<br />

We can then write<br />

ar1 ar2 ar3<br />

I f (p, q, r) is an even permutation of (1, 2, 3) then ∆ = |A|<br />

I f (p, q, r) is an odd permutation of (1, 2, 3) then ∆ = −|A|<br />

I f (p, q, r) is not a permutation of (1, 2, 3) then ∆ = 0.<br />

eijkapiaqjark = epqr|A|.<br />

Each of the above results can be verified by performing the indicated summations. A more formal proof of<br />

the above result is given in EXAMPLE 1.1-25, later in this section.<br />

EXAMPLE 1.1-10. The expression eijkBijCi is meaningless since the index i repeats itself more than<br />

twice and the summation convention does not allow this.<br />

EXAMPLE 1.1-11.<br />

The cross product of the unit vectors e1, e2, e3 can be represented in the index notation by<br />

⎧<br />

⎪⎨ ek if (i, j, k) is an even permutation of (1, 2, 3)<br />

ei × ej = − ek<br />

⎪⎩<br />

0<br />

if (i, j, k) is an odd permutation of (1, 2, 3)<br />

in all other cases<br />

This result can be written in the form ei × ej = ekij ek. This later result can be verified by summing on the<br />

index k and writing out all 9 possible combinations for i and j.<br />

EXAMPLE 1.1-<strong>12</strong>. Given the vectors Ap, p=1, 2, 3andBp, p=1, 2, 3 the cross product of these two<br />

vectors is a vector Cp, p=1, 2, 3withcomponents<br />

The quantities Ci represent the components of the cross product vector<br />

Ci = eijkAjBk, i,j,k =1, 2, 3. (1.1.2)<br />

C = A × B = C1 e1 + C2 e2 + C3 e3.<br />

The equation (1.1.2), which defines the components of C, is to be summed over each of the indices which<br />

repeats itself. We have summing on the index k<br />

Ci = eij1AjB1 + eij2AjB2 + eij3AjB3. (1.1.3)<br />

11


<strong>12</strong><br />

We next sum on the index j which repeats itself in each term of equation (1.1.3). This gives<br />

Ci = ei11A1B1 + ei21A2B1 + ei31A3B1<br />

+ ei<strong>12</strong>A1B2 + ei22A2B2 + ei32A3B2<br />

+ ei13A1B3 + ei23A2B3 + ei33A3B3.<br />

(1.1.4)<br />

Now we are left with i being a free index which can have any of the values of 1, 2or3. Letting i =1, then<br />

letting i =2, and finally letting i = 3 produces the cross product components<br />

C1 = A2B3 − A3B2<br />

C2 = A3B1 − A1B3<br />

C3 = A1B2 − A2B1.<br />

The cross product can also be expressed in the form A × B = eijkAjBk ei. This result can be verified by<br />

summing over the indices i,j and k.<br />

EXAMPLE 1.1-13. Show<br />

eijk = −eikj = ejki for i, j, k =1, 2, 3<br />

Solution: The array ikjrepresents an odd number of transpositions of the indices ijkand to each<br />

transposition there is a sign change of the e-permutation symbol. Similarly, jkiis an even transposition<br />

of ijkand so there is no sign change of the e-permutation symbol. The above holds regardless of the<br />

numerical values assigned to the indices i, j, k.<br />

The e-δ Identity<br />

An identity relating the e-permutation symbol and the Kronecker delta, which is useful in the simplification<br />

of tensor expressions, is the e-δ identity. This identity can be expressed in different forms. The<br />

subscript form for this identity is<br />

eijkeimn = δjmδkn − δjnδkm, i,j,k,m,n=1, 2, 3<br />

where i is the summation index and j, k, m, n are free indices. A device used to remember the positions of<br />

the subscripts is given in the figure 1.1-3.<br />

The subscripts on the four Kronecker delta’s on the right-hand side of the e-δ identity then are read<br />

(first)(second)-(outer)(inner).<br />

This refers to the positions following the summation index. Thus, j, m are the first indices after the summation<br />

index and k, n are the second indices after the summation index. The indices j, n are outer indices<br />

when compared to the inner indices k, m as the indices are viewed as written on the left-hand side of the<br />

identity.


Figure 1.1-3. Mnemonic device for position of subscripts.<br />

Another form of this identity employs both subscripts and superscripts and has the form<br />

e ijk eimn = δ j mδk n − δj nδk m . (1.1.5)<br />

One way of proving this identity is to observe the equation (1.1.5) has the free indices j, k, m, n. Each<br />

of these indices can have any of the values of 1, 2or3. There are 3 choices we can assign to each of j, k, m<br />

or n and this gives a total of 34 = 81 possible equations represented by the identity from equation (1.1.5).<br />

By writing out all 81 of these equations we can verify that the identity is true for all possible combinations<br />

that can be assigned to the free indices.<br />

An alternate proof of the e − δ identity is to consider the determinant<br />

<br />

<br />

δ<br />

<br />

<br />

<br />

1 1 δ1 2 δ1 3<br />

δ2 1 δ2 2 δ2 <br />

<br />

<br />

<br />

3 <br />

=<br />

<br />

<br />

1 0 0<br />

<br />

<br />

0 1 0<br />

<br />

0 0 1<br />

=1.<br />

δ 3 1 δ 3 2 δ 3 3<br />

By performing a permutation of the rows of this matrix we can use the permutation symbol and write<br />

<br />

<br />

δ<br />

<br />

<br />

<br />

i 1 δi 2 δi δ<br />

3<br />

j<br />

1 δ j<br />

2 δ j<br />

<br />

<br />

<br />

<br />

3 <br />

= eijk .<br />

δ k 1 δ k 2 δ k 3<br />

By performing a permutation of the columns, we can write<br />

<br />

<br />

δ<br />

<br />

<br />

<br />

i r δi s δi δ<br />

t<br />

j r δj s δ j<br />

<br />

<br />

<br />

<br />

t <br />

= eijkerst. δ k r δ k s δ k t<br />

Now perform a contraction on the indices i and r to obtain<br />

<br />

<br />

<br />

δ<br />

<br />

<br />

<br />

i i δi s δi δ<br />

t<br />

j<br />

i δj s δ j<br />

<br />

<br />

<br />

<br />

t <br />

= eijkeist. δ k i δ k s δ k t<br />

Summing on i we have δ i i = δ1 1 + δ 2 2 + δ 3 3 = 3 and expand the determinant to obtain the desired result<br />

δ j sδ k t − δ j<br />

t δ k s = e ijk eist.<br />

13


14<br />

Generalized Kronecker delta<br />

The generalized Kronecker delta is defined by the (n × n) determinant<br />

δ ij...k<br />

mn...p =<br />

For example, in three dimensions we can write<br />

δ ijk<br />

mnp =<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

.<br />

.<br />

<br />

δi m δi n ··· δi p<br />

δj m δj n ··· δj p<br />

.<br />

. ..<br />

.<br />

δ k m δk n ··· δ k p<br />

δi m δi n δi p<br />

δj m δj n δj p<br />

δk m δk n δk p<br />

<br />

<br />

<br />

<br />

<br />

.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

= eijkemnp. Performing a contraction on the indices k and p we obtain the fourth order system<br />

δ rs<br />

mn = δrsp mnp = erspemnp = e prs epmn = δ r mδs n − δr nδs m .<br />

As an exercise one can verify that the definition of the e-permutation symbol can also be defined in terms<br />

of the generalized Kronecker delta as<br />

1 2 3··· N<br />

ej1j2j3···jN = δj1j2j3···jN .<br />

Additional definitions and results employing the generalized Kronecker delta are found in the exercises.<br />

In section 1.3 we shall show that the Kronecker delta and epsilon permutation symbol are numerical tensors<br />

whichhavefixedcomponentsineverycoordinatesystem.<br />

Additional Applications of the Indicial Notation<br />

The indicial notation, together with the e − δ identity, can be used to prove various vector identities.<br />

EXAMPLE 1.1-14. Show, using the index notation, that A × B = − B × A<br />

Solution: Let<br />

C = A × B = C1 e1 + C2 e2 + C3 e3 = Ci ei<br />

D =<br />

and let<br />

B × A = D1 e1 + D2 e2 + D3 e3 = Di ei.<br />

We have shown that the components of the cross products can be represented in the index notation by<br />

Ci = eijkAjBk and Di = eijkBjAk.<br />

We desire to show that Di = −Ci for all values of i. Consider the following manipulations: Let Bj = Bsδsj<br />

and Ak = Amδmk and write<br />

Di = eijkBjAk = eijkBsδsjAmδmk<br />

(1.1.6)<br />

where all indices have the range 1, 2, 3. In the expression (1.1.6) note that no summation index appears<br />

more than twice because if an index appeared more than twice the summation convention would become<br />

meaningless. By rearranging terms in equation (1.1.6) we have<br />

Di = eijkδsjδmkBsAm = eismBsAm.


In this expression the indices s and m are dummy summation indices and can be replaced by any other<br />

letters. We replace s by k and m by j to obtain<br />

Di = eikjAjBk = −eijkAjBk = −Ci.<br />

Consequently, we find that D = − C or B × A = − A × B. That is, D = Di ei = −Ci ei = − C.<br />

Note 1. The expressions<br />

Ci = eijkAjBk and Cm = emnpAnBp<br />

with all indices having the range 1, 2, 3, appear to be different because different letters are used as subscripts.<br />

It must be remembered that certain indices are summed according to the summation convention<br />

and the other indices are free indices and can take on any values from the assigned range. Thus, after<br />

summation, when numerical values are substituted for the indices involved, none of the dummy letters<br />

used to represent the components appear in the answer.<br />

Note 2. A second important point is that when one is working with expressions involving the index notation,<br />

the indices can be changed directly. For example, in the above expression for Di we could have replaced<br />

j by k and k by j simultaneously (so that no index repeats itself more than twice) to obtain<br />

Di = eijkBjAk = eikjBkAj = −eijkAjBk = −Ci.<br />

Note 3. Be careful in switching back and forth between the vector notation and index notation. Observe that a<br />

vector A can be represented<br />

A = Ai ei<br />

or its components can be represented<br />

A · ei = Ai, i =1, 2, 3.<br />

Do not set a vector equal to a scalar. That is, do not make the mistake of writing A = Ai as this is a<br />

misuse of the equal sign. It is not possible for a vector to equal a scalar because they are two entirely<br />

different quantities. A vector has both magnitude and direction while a scalar has only magnitude.<br />

EXAMPLE 1.1-15. Verify the vector identity<br />

Solution: Let<br />

A · ( B × C)= B · ( C × A)<br />

B × C = D = Di ei where Di = eijkBjCk and let<br />

C × A = F = Fi ei where Fi = eijkCjAk<br />

where all indices have the range 1, 2, 3. To prove the above identity, we have<br />

A · ( B × C)= A · D = AiDi = AieijkBjCk<br />

= Bj(eijkAiCk)<br />

= Bj(ejkiCkAi)<br />

15


16<br />

since eijk = ejki. We also observe from the expression<br />

Fi = eijkCjAk<br />

that we may obtain, by permuting the symbols, the equivalent expression<br />

This allows us to write<br />

Fj = ejkiCkAi.<br />

A · ( B × C)=BjFj = B · F = B · ( C × A)<br />

which was to be shown.<br />

The quantity A · ( B × C) is called a triple scalar product. The above index representation of the triple<br />

scalar product implies that it can be represented as a determinant (See example 1.1-9). We can write<br />

A · ( B × <br />

<br />

<br />

C)= <br />

<br />

<br />

A1 A2 A3<br />

B1 B2 B3<br />

C1 C2 C3<br />

<br />

<br />

<br />

<br />

= eijkAiBjCk<br />

<br />

A physical interpretation that can be assigned to this triple scalar product is that its absolute value represents<br />

the volume of the parallelepiped formed by the three noncoplaner vectors A, B, C. The absolute value is<br />

needed because sometimes the triple scalar product is negative. This physical interpretation can be obtained<br />

from an analysis of the figure 1.1-4.<br />

Figure 1.1-4. Triple scalar product and volume


In figure 1.1-4 observe that: (i) | B × C| is the area of the parallelogram PQRS. (ii) the unit vector<br />

en = B × C<br />

| B × C|<br />

is normal to the plane containing the vectors B and C. (iii) The dot product<br />

<br />

<br />

A · en = <br />

A · B × C<br />

| B × <br />

<br />

<br />

= h<br />

C|<br />

equals the projection of A on en which represents the height of the parallelepiped. These results demonstrate<br />

that<br />

<br />

<br />

A · ( B × <br />

<br />

C) = | B × C| h = (area of base)(height) = volume.<br />

EXAMPLE 1.1-16. Verify the vector identity<br />

( A × B) × ( C × D)= C( D · A × B) − D( C · A × B)<br />

Solution: Let F = A × B = Fi ei and E = C × D = Ei ei. These vectors have the components<br />

Fi = eijkAjBk and Em = emnpCnDp<br />

where all indices have the range 1, 2, 3. The vector G = F × E = Gi ei has the components<br />

From the identity eqim = emqi this can be expressed<br />

Gq = eqimFiEm = eqimeijkemnpAjBkCnDp.<br />

Gq =(emqiemnp)eijkAjBkCnDp<br />

which is now in a form where we can use the e − δ identity applied to the term in parentheses to produce<br />

Simplifying this expression we have:<br />

Gq =(δqnδip − δqpδin)eijkAjBkCnDp.<br />

Gq = eijk [(Dpδip)(Cnδqn)AjBk − (Dpδqp)(Cnδin)AjBk]<br />

= eijk [DiCqAjBk − DqCiAjBk]<br />

which are the vector components of the vector<br />

= Cq [DieijkAjBk] − Dq [CieijkAjBk]<br />

C( D · A × B) − D( C · A × B).<br />

17


18<br />

Transformation Equations<br />

Consider two sets of N independent variables which are denoted by the barred and unbarred symbols<br />

x i and x i with i = 1,...,N. The independent variables x i ,i =1,...,N can be thought of as defining<br />

the coordinates of a point in a N−dimensional space. Similarly, the independent barred variables define a<br />

point in some other N−dimensional space. These coordinates are assumed to be real quantities and are not<br />

complex quantities. Further, we assume that these variables are related by a set of transformation equations.<br />

x i = x i (x 1 , x 2 ,...,x N ) i =1,...,N. (1.1.7)<br />

It is assumed that these transformation equations are independent. A necessary and sufficient condition that<br />

these transformation equations be independent is that the Jacobian determinant be different from zero, that<br />

is<br />

J( x<br />

x )=<br />

<br />

<br />

<br />

∂x<br />

<br />

i<br />

∂¯x j<br />

<br />

<br />

<br />

=<br />

<br />

∂x<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

1<br />

∂x1 ∂x 1<br />

∂x2 ∂x ··· 1<br />

∂xN ∂x 2<br />

∂x1 ∂x 2<br />

∂x2 ∂x ··· 2<br />

∂xN <br />

<br />

<br />

<br />

<br />

<br />

.<br />

. .<br />

.<br />

.. .<br />

= 0.<br />

. <br />

<br />

<br />

∂x N<br />

∂x 1<br />

This assumption allows us to obtain a set of inverse relations<br />

∂x N<br />

∂x2 ∂x ··· N<br />

∂xN x i = x i (x 1 ,x 2 ,...,x N ) i =1,...,N, (1.1.8)<br />

where the x ′ s are determined in terms of the x ′ s. Throughout our discussions it is to be understood that the<br />

given transformation equations are real and continuous. Further all derivatives that appear in our discussions<br />

are assumed to exist and be continuous in the domain of the variables considered.<br />

EXAMPLE 1.1-17. The following is an example of a set of transformation equations of the form<br />

defined by equations (1.1.7) and (1.1.8) in the case N =3. Consider the transformation from cylindrical<br />

coordinates (r, α, z) to spherical coordinates (ρ, β, α). From the geometry of the figure 1.1-5 we can find the<br />

transformation equations<br />

r = ρ sin β<br />

with inverse transformation<br />

Now make the substitutions<br />

α = α 0


Figure 1.1-5. Cylindrical and Spherical Coordinates<br />

The resulting transformations then have the forms of the equations (1.1.7) and (1.1.8).<br />

Calculation of Derivatives<br />

We now consider the chain rule applied to the differentiation of a function of the bar variables. We<br />

represent this differentiation in the indicial notation. Let Φ = Φ(x1 , x2 ,...,xn ) be a scalar function of the<br />

variables xi , i =1,...,N and let these variables be related to the set of variables xi , with i =1,...,N by<br />

the transformation equations (1.1.7) and (1.1.8). The partial derivatives of Φ with respect to the variables<br />

xi can be expressed in the indicial notation as<br />

∂Φ ∂Φ<br />

=<br />

∂xi ∂xj ∂xj ∂Φ<br />

=<br />

∂xi ∂x1 ∂x1 ∂Φ<br />

+<br />

∂xi ∂x2 ∂x2 ∂Φ<br />

+ ···+<br />

∂xi ∂xN ∂xN ∂xi (1.1.9)<br />

for any fixed value of i satisfying 1 ≤ i ≤ N.<br />

The second partial derivatives of Φ can also be expressed in the index notation. Differentiation of<br />

equation (1.1.9) partially with respect to xm produces<br />

∂2Φ ∂xi ∂Φ<br />

=<br />

∂xm ∂xj ∂2xj ∂xi ∂<br />

+<br />

∂xm ∂xm <br />

∂Φ<br />

∂xj j ∂x<br />

. (1.1.10)<br />

∂xi This result is nothing more than an application of the general rule for differentiating a product of two<br />

quantities. To evaluate the derivative of the bracketed term in equation (1.1.10) it must be remembered that<br />

the quantity inside the brackets is a function of the bar variables. Let<br />

G = ∂Φ<br />

∂x j = G(x1 , x 2 ,...,x N )<br />

to emphasize this dependence upon the bar variables, then the derivative of G is<br />

∂G ∂G<br />

=<br />

∂xm ∂xk ∂xj∂x k<br />

This is just an application of the basic rule from equation (1.1.9) with Φ replaced by G. Hence the derivative<br />

from equation (1.1.10) can be expressed<br />

∂x k<br />

∂x m = ∂2 Φ<br />

∂2Φ ∂xi ∂Φ<br />

=<br />

∂xm ∂xj ∂2xj ∂xi∂xm + ∂2Φ ∂xj∂x k<br />

where i, m arefreeindicesandj, k are dummy summation indices.<br />

∂xk . (1.1.11)<br />

∂xm ∂x j<br />

∂x i<br />

∂x k<br />

∂x m<br />

(1.1.<strong>12</strong>)<br />

19


20<br />

EXAMPLE 1.1-18. Let Φ = Φ(r, θ) wherer, θ are polar coordinates related to the Cartesian coordinates<br />

(x, y) by the transformation equations x = r cos θ y = r sin θ. Find the partial derivatives ∂Φ ∂<br />

and<br />

∂x<br />

2Φ ∂x2 Solution: The partial derivative of Φ with respect to x is found from the relation (1.1.9) and can be written<br />

∂Φ<br />

∂x<br />

∂Φ ∂r<br />

=<br />

∂r ∂x<br />

+ ∂Φ<br />

∂θ<br />

∂θ<br />

. (1.1.13)<br />

∂x<br />

The second partial derivative is obtained by differentiating the first partial derivative. From the product<br />

rule for differentiation we can write<br />

∂2Φ ∂Φ ∂<br />

=<br />

∂x2 ∂r<br />

2 <br />

r ∂r ∂ ∂Φ<br />

+ +<br />

∂x2 ∂x ∂x ∂r<br />

∂Φ ∂<br />

∂θ<br />

2 <br />

θ ∂θ ∂ ∂Φ<br />

+ . (1.1.14)<br />

∂x2 ∂x ∂x ∂θ<br />

To further simplify (1.1.14) it must be remembered that the terms inside the brackets are to be treated as<br />

functions of the variables r and θ and that the derivative of these terms can be evaluated by reapplying the<br />

basic rule from equation (1.1.13) with Φ replaced by ∂Φ<br />

∂r<br />

∂Φ<br />

and then Φ replaced by ∂θ . This gives<br />

∂2Φ ∂Φ ∂<br />

=<br />

∂x2 ∂r<br />

2 2 r ∂r ∂ Φ<br />

+<br />

∂x2 ∂x ∂r2 ∂r<br />

∂x + ∂2 <br />

Φ ∂θ<br />

∂r∂θ ∂x<br />

+ ∂Φ ∂<br />

∂θ<br />

2 2 θ ∂θ ∂ Φ ∂r<br />

+<br />

∂x2 ∂x ∂θ∂r ∂x + ∂2Φ ∂θ2 <br />

∂θ<br />

.<br />

∂x<br />

(1.1.15)<br />

From the transformation equations we obtain the relations r 2 = x 2 +y 2<br />

and tan θ = y<br />

and from<br />

x<br />

these relations we can calculate all the necessary derivatives needed for the simplification of the equations<br />

(1.1.13) and (1.1.15). These derivatives are:<br />

2r ∂r<br />

=2x or<br />

∂x<br />

sec 2 θ ∂θ y<br />

= −<br />

∂x x2 ∂2r ∂θ<br />

= − sin θ<br />

∂x2 ∂x = sin2 θ<br />

r<br />

or<br />

∂r x<br />

=<br />

∂x r =cosθ<br />

∂θ y sin θ<br />

= − = −<br />

∂x r2 r<br />

∂2 ∂θ<br />

θ −r cos θ ∂x =<br />

∂x2 r2 +sinθ ∂r<br />

∂x<br />

= 2sinθ cos θ<br />

.<br />

r<br />

Therefore, the derivatives from equations (1.1.13) and (1.1.15) can be expressed in the form<br />

∂Φ<br />

∂x<br />

= ∂Φ<br />

∂r<br />

∂2Φ ∂Φ<br />

=<br />

∂x2 ∂r<br />

∂Φ sin θ<br />

cos θ −<br />

∂θ r<br />

sin 2 θ<br />

r +2∂Φ<br />

sin θ cos θ<br />

∂θ r2 + ∂2Φ ∂r2 cos2 θ − 2 ∂2Φ cos θ sin θ<br />

∂r∂θ r<br />

+ ∂2Φ ∂θ2 sin 2 θ<br />

r2 .<br />

By letting x 1 = r, x 2 = θ, x 1 = x, x 2 = y and performing the indicated summations in the equations (1.1.9)<br />

and (1.1.<strong>12</strong>) there is produced the same results as above.<br />

Vector Identities in Cartesian Coordinates<br />

Employing the substitutions x 1 = x, x 2 = y, x 3 = z, where superscript variables are employed and<br />

denoting the unit vectors in Cartesian coordinates by e1, e2, e3, we illustrated how various vector operations<br />

are written by using the index notation.


Gradient. In Cartesian coordinates the gradient of a scalar field is<br />

grad φ = ∂φ<br />

∂x e1 + ∂φ<br />

∂y e2 + ∂φ<br />

∂z e3.<br />

The index notation focuses attention only on the components of the gradient. In Cartesian coordinates these<br />

components are represented using a comma subscript to denote the derivative<br />

ej · grad φ = φ,j = ∂φ<br />

, j =1, 2, 3.<br />

∂xj The comma notation will be discussed in section 4. For now we use it to denote derivatives. For example<br />

φ ,j = ∂φ<br />

∂xj , φ,jk = ∂2φ ∂xj , etc.<br />

∂xk Divergence. In Cartesian coordinates the divergence of a vector field A is a scalar field and can be<br />

represented<br />

∇· A = div A = ∂A1<br />

∂x<br />

+ ∂A2<br />

∂y<br />

+ ∂A3<br />

∂z .<br />

Employing the summation convention and index notation, the divergence in Cartesian coordinates can be<br />

represented<br />

∇· A = div A = Ai,i = ∂Ai ∂A1 ∂A2 ∂A3<br />

= + +<br />

∂xi ∂x1 ∂x2 ∂x3 where i is the dummy summation index.<br />

Curl. To represent the vector B =curl A = ∇× A in Cartesian coordinates, we note that the index<br />

notation focuses attention only on the components of this vector. The components Bi, i =1, 2, 3of B can<br />

be represented<br />

Bi = ei · curl A = eijkAk,j, for i, j, k =1, 2, 3<br />

where eijk is the permutation symbol introduced earlier and Ak,j = ∂Ak<br />

∂x j . To verify this representation of the<br />

curl A we need only perform the summations indicated by the repeated indices. We have summing on j that<br />

Bi = ei1kAk,1 + ei2kAk,2 + ei3kAk,3.<br />

Now summing each term on the repeated index k gives us<br />

Bi = ei<strong>12</strong>A2,1 + ei13A3,1 + ei21A1,2 + ei23A3,2 + ei31A1,3 + ei32A2,3<br />

Here i is a free index which can take on any of the values 1, 2or3. Consequently, we have<br />

For i =1, B1 = A3,2 − A2,3 = ∂A3 ∂A2<br />

−<br />

∂x2 ∂x3 For i =2, B2 = A1,3 − A3,1 = ∂A1 ∂A3<br />

−<br />

∂x3 ∂x1 For i =3, B3 = A2,1 − A1,2 = ∂A2 ∂A1<br />

−<br />

∂x1 ∂x2 which verifies the index notation representation of curl A in Cartesian coordinates.<br />

21


22<br />

Other Operations. The following examples illustrate how the index notation can be used to represent<br />

additional vector operators in Cartesian coordinates.<br />

1. In index notation the components of the vector ( B ·∇) A are<br />

{( B ·∇) A}· ep = Ap,qBq<br />

p, q =1, 2, 3<br />

This can be verified by performing the indicated summations. We have by summing on the repeated<br />

index q<br />

Ap,qBq = Ap,1B1 + Ap,2B2 + Ap,3B3.<br />

The index p is now a free index which can have any of the values 1, 2or3. We have:<br />

for p =1, A1,qBq = A1,1B1 + A1,2B2 + A1,3B3<br />

= ∂A1<br />

∂x1 B1 + ∂A1<br />

∂x2 B2 + ∂A1<br />

for p =2,<br />

∂x<br />

A2,qBq = A2,1B1 + A2,2B2 + A2,3B3<br />

3 B3<br />

= ∂A2<br />

∂x1 B1 + ∂A2<br />

∂x2 B2 + ∂A2<br />

for p =3,<br />

∂x<br />

A3,qBq = A3,1B1 + A3,2B2 + A3,3B3<br />

= ∂A3<br />

∂x1 B1 + ∂A3<br />

∂x2 B2 + ∂A3<br />

∂x<br />

2. The scalar ( B ·∇)φ has the following form when expressed in the index notation:<br />

( B ·∇)φ = Biφ,i = B1φ,1 + B2φ,2 + B3φ,3<br />

∂φ ∂φ ∂φ<br />

= B1 + B2 + B3 .<br />

∂x1 ∂x2 ∂x3 3 B3<br />

3 B3<br />

3. The components of the vector ( B ×∇)φ is expressed in the index notation by<br />

<br />

ei · ( <br />

B ×∇)φ = eijkBjφ,k.<br />

This can be verified by performing the indicated summations and is left as an exercise.<br />

4. The scalar ( B ×∇) · A may be expressed in the index notation. It has the form<br />

( B ×∇) · A = eijkBjAi,k.<br />

This can also be verified by performing the indicated summations and is left as an exercise.<br />

5. The vector components of ∇2A in the index notation are represented<br />

The proof of this is left as an exercise.<br />

ep ·∇ 2 A = Ap,qq.


EXAMPLE 1.1-19. In Cartesian coordinates prove the vector identity<br />

curl (f A)=∇×(f A)=(∇f) × A + f(∇× A).<br />

Solution: Let B =curl(f A) and write the components as<br />

Bi = eijk(fAk),j<br />

= eijk [fAk,j + f,jAk]<br />

= feijkAk,j + eijkf,jAk.<br />

This index form can now be expressed in the vector form<br />

B =curl(f A)=f(∇× A)+(∇f) × A<br />

EXAMPLE 1.1-20. Prove the vector identity ∇·( A + B)=∇· A + ∇· B<br />

Solution: Let A + B = C and write this vector equation in the index notation as Ai + Bi = Ci. We then<br />

have<br />

∇· C = Ci,i =(Ai + Bi),i = Ai,i + Bi,i = ∇· A + ∇· B.<br />

EXAMPLE 1.1-21. In Cartesian coordinates prove the vector identity ( A ·∇)f = A ·∇f<br />

Solution: In the index notation we write<br />

( A ·∇)f = Aif,i = A1f,1 + A2f,2 + A3f,3<br />

∂f ∂f ∂f<br />

= A1 + A2 + A3<br />

∂x1 ∂x2 ∂x3 = A ·∇f.<br />

EXAMPLE 1.1-22. In Cartesian coordinates prove the vector identity<br />

∇×( A × B)= A(∇· B) − B(∇· A)+( B ·∇) A − ( A ·∇) B<br />

Solution: The pth component of the vector ∇×( A × B)is<br />

ep · [∇×( A × B)] = epqk[ekjiAjBi],q<br />

= epqkekjiAjBi,q + epqkekjiAj,qBi<br />

By applying the e − δ identity, the above expression simplifies to the desired result. That is,<br />

In vector form this is expressed<br />

ep · [∇×( A × B)] = (δpjδqi − δpiδqj)AjBi,q +(δpjδqi − δpiδqj)Aj,qBi<br />

= ApBi,i − AqBp,q + Ap,qBq − Aq,qBp<br />

∇×( A × B)= A(∇· B) − ( A ·∇) B +( B ·∇) A − B(∇· A)<br />

23


24<br />

EXAMPLE 1.1-23. In Cartesian coordinates prove the vector identity ∇×(∇× A)=∇(∇· A) −∇ 2 A<br />

Solution: We have for the ith component of ∇× A is given by ei · [∇× A]=eijkAk,j and consequently the<br />

pth component of ∇×(∇× A)is<br />

The e − δ identity produces<br />

ep · [∇×(∇× A)] = epqr[erjkAk,j],q<br />

= epqrerjkAk,jq.<br />

ep · [∇×(∇× A)] = (δpjδqk − δpkδqj)Ak,jq<br />

= Ak,pk − Ap,qq.<br />

Expressing this result in vector form we have ∇×(∇× A)=∇(∇· A) −∇ 2 A.<br />

Indicial Form of Integral Theorems<br />

The divergence theorem, in both vector and indicial notation, can be written<br />

<br />

<br />

<br />

div · F dτ= F · n dσ<br />

Fi,i dτ = Fini dσ i =1, 2, 3 (1.1.16)<br />

V<br />

S<br />

V<br />

where ni are the direction cosines of the unit exterior normal to the surface, dτ is a volume element and dσ<br />

is an element of surface area. Note that in using the indicial notation the volume and surface integrals are<br />

to be extended over the range specified by the indices. This suggests that the divergence theorem can be<br />

applied to vectors in n−dimensional spaces.<br />

The vector form and indicial notation for the Stokes theorem are<br />

<br />

(∇×<br />

S<br />

<br />

<br />

<br />

F ) · n dσ = F · dr<br />

eijkFk,jni dσ = Fi dx<br />

C<br />

S<br />

C<br />

i<br />

i, j, k =1, 2, 3 (1.1.17)<br />

and the Green’s theorem in the plane, which is a special case of the Stoke’s theorem, can be expressed<br />

<br />

∂F2 ∂F1<br />

− dxdy = F1 dx + F2 dy<br />

∂x ∂y<br />

C<br />

<br />

<br />

e3jkFk,j dS = Fi dx<br />

S<br />

C<br />

i<br />

i, j, k =1, 2 (1.1.18)<br />

Other forms of the above integral theorems are<br />

<br />

∇φdτ =<br />

V<br />

S<br />

S<br />

φ n dσ<br />

obtained from the divergence theorem by letting F = φ C where C is a constant vector. By replacing F by<br />

F × C in the divergence theorem one can derive<br />

<br />

∇×<br />

V<br />

<br />

F dτ = − F × ndσ.<br />

S<br />

In the divergence theorem make the substitution F = φ∇ψ to obtain<br />

<br />

<br />

2<br />

(φ∇ ψ +(∇φ) · (∇ψ) dτ = (φ∇ψ) · n dσ.<br />

V<br />

S


The Green’s identity <br />

V<br />

<br />

2 2<br />

φ∇ ψ − ψ∇ φ dτ = (φ∇ψ − ψ∇φ) · n dσ<br />

is obtained by first letting F = φ∇ψ in the divergence theorem and then letting F = ψ∇φ in the divergence<br />

theorem and then subtracting the results.<br />

Determinants, Cofactors<br />

For A =(aij), i,j=1,...,n an n × n matrix, the determinant of A can be written as<br />

det A = |A| = ei1i2i3...ina1i1a2i2a3i3 ...anin.<br />

This gives a summation of the n! permutations of products formed from the elements of the matrix A. The<br />

result is a single number called the determinant of A.<br />

EXAMPLE 1.1-24. Inthecasen =2wehave<br />

<br />

<br />

|A| = <br />

a11 a<strong>12</strong><br />

a21 a22<br />

S<br />

<br />

<br />

<br />

= enma1na2m<br />

= e1ma11a2m + e2ma<strong>12</strong>a2m<br />

= e<strong>12</strong>a11a22 + e21a<strong>12</strong>a21<br />

= a11a22 − a<strong>12</strong>a21<br />

EXAMPLE 1.1-25. Inthecasen = 3 we can use either of the notations<br />

⎛<br />

A = ⎝ a11 a<strong>12</strong><br />

⎞<br />

a13<br />

⎠ or<br />

⎛<br />

A =<br />

a21 a22 a23<br />

a31 a32 a33<br />

and represent the determinant of A in any of the forms<br />

det A = eijka1ia2ja3k<br />

det A = eijkai1aj2ak3<br />

det A = eijka i 1a j<br />

2 ak 3<br />

det A = eijka 1 i a 2 ja 3 k.<br />

⎝ a11 a1 2 a1 3<br />

a2 1 a22 a23 a3 1 a32 a33 These represent row and column expansions of the determinant.<br />

An important identity results if we examine the quantity Brst = eijka i ra j sa k t . Itisaneasyexerciseto<br />

change the dummy summation indices and rearrange terms in this expression. For example,<br />

Brst = eijka i r aj s ak t = ekjia k r aj s ai t = ekjia i t aj s ak r = −eijka i t aj s ak r<br />

⎞<br />

⎠<br />

= −Btsr,<br />

and by considering other permutations of the indices, one can establish that Brst is completely skewsymmetric.<br />

In the exercises it is shown that any third order completely skew-symmetric system satisfies<br />

Brst = B<strong>12</strong>3erst. But B<strong>12</strong>3 =detA and so we arrive at the identity<br />

Brst = eijka i r aj s ak t<br />

= |A|erst.<br />

25


26<br />

Other forms of this identity are<br />

e ijk a r i as j at k<br />

= |A|erst<br />

Consider the representation of the determinant<br />

<br />

<br />

<br />

|A| = <br />

<br />

<br />

and eijkairajsakt = |A|erst. (1.1.19)<br />

a1 1 a<strong>12</strong> a13 a2 1 a2 2 a2 3<br />

a3 1 a32 a33 by use of the indicial notation. By column expansions, this determinant can be represented<br />

|A| = ersta r 1 as 2 at 3<br />

and if one uses row expansions the determinant can be expressed as<br />

<br />

<br />

<br />

<br />

<br />

<br />

(1.1.20)<br />

|A| = e ijk a 1 i a2j a3k . (1.1.21)<br />

Define Ai m as the cofactor of the element am i in the determinant |A|. From the equation (1.1.20) the cofactor<br />

of ar 1 is obtained by deleting this element and we find<br />

The result (1.1.20) can then be expressed in the form<br />

A 1 r = ersta s 2a t 3. (1.1.22)<br />

|A| = a r 1A1r = a11 A11 + a21 A<strong>12</strong> + a31 A13 . (1.1.23)<br />

That is, the determinant |A| is obtained by multiplying each element in the first column by its corresponding<br />

cofactor and summing the result. Observe also that from the equation (1.1.20) we find the additional<br />

cofactors<br />

A 2 s = ersta r 1at3 and A 3 t = ersta r 1as2 . (1.1.24)<br />

Hence, the equation (1.1.20) can also be expressed in one of the forms<br />

|A| = a s 2 A2 s = a1 2 A2 1 + a2 2 A2 2 + a3 2 A2 3<br />

|A| = a t 3A 3 t = a 1 3A 3 1 + a 2 3A 3 2 + a 3 3A 3 3<br />

The results from equations (1.1.22) and (1.1.24) can be written in a slightly different form with the indicial<br />

notation. From the notation for a generalized Kronecker delta defined by<br />

the above cofactors can be written in the form<br />

e ijk elmn = δ ijk<br />

lmn ,<br />

A 1 r = e <strong>12</strong>3 ersta s 2a t 3 = 1<br />

2! e1jk ersta s ja t k = 1<br />

2! δ1jk<br />

rst a s ja t k<br />

A 2 r = e<strong>12</strong>3 esrta s 1 at 3<br />

A 3 r = e<strong>12</strong>3 etsra t 1 as 2<br />

1<br />

=<br />

2! e2jkersta s jat 1<br />

k =<br />

2! δ2jk rst asj atk 1<br />

=<br />

2! e3jkersta s jat 1<br />

k =<br />

2! δ3jk rst asj atk .


These cofactors are then combined into the single equation<br />

A i r<br />

= 1<br />

2! δijk<br />

rst as j at k<br />

(1.1.25)<br />

which represents the cofactor of ar i . When the elements from any row (or column) are multiplied by their<br />

corresponding cofactors, and the results summed, we obtain the value of the determinant. Whenever the<br />

elements from any row (or column) are multiplied by the cofactor elements from a different row (or column),<br />

and the results summed, we get zero. This can be illustrated by considering the summation<br />

a m r Ai m<br />

Here we have used the e − δ identity to obtain<br />

1<br />

=<br />

2! δijk mstasj atk am 1<br />

r =<br />

2! eijkemsta m r asj atk = 1<br />

2! eijkerjk|A| = 1<br />

2! δijk<br />

rjk |A| = δi r |A|<br />

δ ijk<br />

rjk = eijk erjk = e jik ejrk = δ i r δk k − δi k δk r =3δi r − δi r =2δi r<br />

which was used to simplify the above result.<br />

As an exercise one can show that an alternate form of the above summation of elements by its cofactors<br />

is<br />

a r m Am i = |A|δr i .<br />

27


28<br />

EXERCISE 1.1<br />

◮ 1. Simplify each of the following by employing the summation property of the Kronecker delta. Perform<br />

sums on the summation indices only if your are unsure of the result.<br />

(a) eijkδkn<br />

(b) eijkδisδjm<br />

(c) eijkδisδjmδkn<br />

(d) aijδin<br />

◮ 2. Simplify and perform the indicated summations over the range 1, 2, 3<br />

(a) δii<br />

(b) δijδij<br />

(c) eijkAiAjAk<br />

(d) eijkeijk<br />

(e) eijkδjk<br />

(e) δijδjn<br />

(f) δijδjnδni<br />

(f) AiBjδji − BmAnδmn<br />

◮ 3. Express each of the following in index notation. Be careful of the notation you use. Note that A = Ai<br />

is an incorrect notation because a vector can not equal a scalar. The notation A · ei = Ai should be used to<br />

express the ith component of a vector.<br />

(a) A · ( B × C)<br />

(b) A × ( B × C)<br />

(c) B( A · C)<br />

(d) B( A · C) − C( A · B)<br />

◮ 4. Show the e permutation symbol satisfies: (a) eijk = ejki = ekij (b) eijk = −ejik = −eikj = −ekji<br />

◮ 5. Use index notation to verify the vector identity A × ( B × C)= B( A · C) − C( A · B)<br />

◮ 6. Let yi = aijxj and xm = aimzi where the range of the indices is 1, 2<br />

(a) Solve for yi in terms of zi using the indicial notation and check your result<br />

to be sure that no index repeats itself more than twice.<br />

(b) Perform the indicated summations and write out expressions<br />

for y1,y2 in terms of z1,z2<br />

(c) Express the above equations in matrix form. Expand the matrix<br />

equations and check the solution obtained in part (b).<br />

◮ 7. Use the e − δ identity to simplify (a) eijkejik (b) eijkejki<br />

◮ 8. Prove the following vector identities:<br />

(a) A · ( B × C)= B · ( C × A)= C · ( A × B) triple scalar product<br />

(b) ( A × B) × C = B( A · C) − A( B · C)<br />

◮ 9. Prove the following vector identities:<br />

(a) ( A × B) · ( C × D)=( A · C)( B · D) − ( A · D)( B · C)<br />

(b) A × ( B × C)+ B × ( C × A)+ C × ( A × B)=0<br />

(c) ( A × B) × ( C × D)= B( A · C × D) − A( B · C × D)


◮ 10. For A =(1, −1, 0) and B =(4, −3, 2) find using the index notation,<br />

(a) Ci = eijkAjBk, i =1, 2, 3<br />

(b) AiBi<br />

◮ 11. Represent the differential equations<br />

using the index notation.<br />

(c) What do the results in (a) and (b) represent?<br />

dy1<br />

dt = a11y1 + a<strong>12</strong>y2<br />

and<br />

dy2<br />

dt = a21y1 + a22y2<br />

◮ <strong>12</strong>.<br />

Let Φ = Φ(r, θ) wherer, θ are polar coordinates related to Cartesian coordinates (x, y) by the transformation<br />

equations x = r cos θ and y = r sin θ.<br />

(a) Find the partial derivatives<br />

∂Φ<br />

,<br />

∂y<br />

and<br />

∂2Φ ∂y2 (b) Combine the result in part (a) with the result from EXAMPLE 1.1-18 to calculate the Laplacian<br />

in polar coordinates.<br />

∇ 2 Φ= ∂2Φ ∂x2 + ∂2Φ ∂y2 ◮ 13. (Index notation) Let a11 =3, a<strong>12</strong> =4, a21 =5, a22 =6.<br />

Calculate the quantity C = aijaij, i,j=1, 2.<br />

◮ 14. Show the moments of inertia Iij defined by<br />

<br />

I11 = (y 2 + z 2 )ρ(x, y, z) dτ<br />

R<br />

<br />

I22 =<br />

R<br />

<br />

I33 =<br />

R<br />

(x 2 + z 2 )ρ(x, y, z) dτ<br />

(x 2 + y 2 )ρ(x, y, z) dτ<br />

can be represented in the index notation as Iij =<br />

<br />

R<br />

x1 = x, x2 = y, x3 = z and dτ = dxdydz is an element of volume.<br />

<br />

I23 = I32 = −<br />

R<br />

R<br />

yzρ(x, y, z) dτ<br />

<br />

I<strong>12</strong> = I21 = − xyρ(x, y, z) dτ<br />

<br />

I13 = I31 = − xzρ(x, y, z) dτ,<br />

◮ 15. Determine if the following relation is true or false. Justify your answer.<br />

Hint: Let em =(δ1m,δ2m,δ3m).<br />

ei · ( ej × ek) =(ei × ej) · ek = eijk, i,j,k =1, 2, 3.<br />

R<br />

x m x m δij − x i x j ρdτ, where ρ is the density,<br />

◮ 16. Without substituting values for i, l =1, 2, 3 calculate all nine terms of the given quantities<br />

(a) B il =(δ i jAk + δ i kAj)e jkl<br />

(b) Ail =(δ m i B k + δ k i B m )emlk<br />

◮ 17. Let Amnx m y n = 0 for arbitrary x i and y i , i =1, 2, 3, and show that Aij = 0 for all values of i, j.<br />

29


30<br />

◮ 18.<br />

(a) For amn,m,n=1, 2, 3 skew-symmetric, show that amnx m x n =0.<br />

(b) Let amnxmxn =0, m,n =1, 2, 3 for all values of xi ,i =1, 2, 3andshowthatamn must be skew-<br />

symmetric.<br />

◮ 19. Let A and B denote 3 × 3 matrices with elements aij and bij respectively. Show that if C = AB is a<br />

matrix product, then det(C) =det(A) · det(B).<br />

Hint: Use the result from example 1.1-9.<br />

◮ 20.<br />

(a) Let u 1 ,u 2 ,u 3 be functions of the variables s 1 ,s 2 ,s 3 . Further, assume that s 1 ,s 2 ,s 3 areinturneach<br />

functions of the variables x 1 , x 2 , x 3 . Let<br />

respect to the x ′ s. Show that<br />

(b) Note that ∂xi<br />

∂¯x j<br />

∂¯x j ∂xi<br />

=<br />

∂xm of the transformation (1.1.7).<br />

<br />

<br />

<br />

∂u<br />

<br />

m<br />

∂xn <br />

<br />

<br />

= ∂(u1 ,u2 , u3 )<br />

<br />

<br />

<br />

∂u<br />

<br />

i<br />

∂xm <br />

<br />

<br />

=<br />

<br />

<br />

<br />

∂u<br />

<br />

i<br />

∂sj ∂sj ∂xm <br />

<br />

<br />

=<br />

<br />

<br />

<br />

∂u<br />

<br />

i<br />

∂sj <br />

<br />

<br />

·<br />

<br />

<br />

<br />

∂s<br />

<br />

j<br />

∂xm <br />

<br />

<br />

.<br />

∂(x 1 ,x 2 , x 3 ) denote the Jacobian of the u′ s with<br />

∂xm = δi m and show that J( x ¯x<br />

x<br />

¯x )·J( x )=1,whereJ( ¯x ) is the Jacobian determinant<br />

◮ 21. A third order system aℓmn with ℓ, m, n =1, 2, 3 is said to be symmetric in two of its subscripts if the<br />

components are unaltered when these subscripts are interchanged. When aℓmn is completely symmetric then<br />

aℓmn = amℓn = aℓnm = amnℓ = anmℓ = anℓm. Whenever this third order system is completely symmetric,<br />

then: (i) How many components are there? (ii) How many of these components are distinct?<br />

Hint: Consider the three cases (i) ℓ = m = n (ii) ℓ = m = n (iii) ℓ = m = n.<br />

◮ 22. A third order system bℓmn with ℓ, m, n =1, 2, 3 is said to be skew-symmetric in two of its subscripts<br />

if the components change sign when the subscripts are interchanged. A completely skew-symmetric third<br />

order system satisfies bℓmn = −bmℓn = bmnℓ = −bnmℓ = bnℓm = −bℓnm. (i) How many components does<br />

a completely skew-symmetric system have? (ii) How many of these components are zero? (iii) How many<br />

components can be different from zero? (iv) Show that there is one distinct component b<strong>12</strong>3 and that<br />

bℓmn = eℓmnb<strong>12</strong>3.<br />

Hint: Consider the three cases (i) ℓ = m = n (ii) ℓ = m = n (iii) ℓ = m = n.<br />

◮ 23. Let i, j, k =1, 2, 3 and assume that eijkσjk = 0 for all values of i. What does this equation tell you<br />

about the values σij, i, j =1, 2, 3?<br />

◮ 24. Assume that Amn and Bmn are symmetric for m, n =1, 2, 3. Let Amnx m x n = Bmnx m x n for arbitrary<br />

values of xi ,i=1, 2, 3, and show that Aij = Bij for all values of i and j.<br />

◮ 25. Assume Bmn is symmetric and Bmnx m x n = 0 for arbitrary values of x i ,i=1, 2, 3, show that Bij =0.


◮ 26. (Generalized Kronecker delta) Define the generalized Kronecker delta as the n×n determinant<br />

δ ij...k<br />

mn...p =<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

.<br />

<br />

(a) Show eijk = δ <strong>12</strong>3<br />

ijk<br />

(b) Show e ijk = δ ijk<br />

<strong>12</strong>3<br />

(c) Show δ ij mn = e ij emn<br />

δi m δi n ··· δi p<br />

δj m δj n ··· δj p<br />

.<br />

. ..<br />

δ k m δk n ··· δ k p<br />

(d) Define δ rs<br />

mn = δrsp mnp (summation on p)<br />

and show δ rs<br />

mn = δr mδs n − δr nδs m<br />

.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

where δ r s<br />

is the Kronecker delta.<br />

Note that by combining the above result with the result from part (c)<br />

we obtain the two dimensional form of the e − δ identity e rs emn = δ r mδ s n − δ r nδ s m.<br />

(e) Define δ r m<br />

(f) Show δ rst<br />

rst =3!<br />

◮ 27. Let A i r denote the cofactor of ar i<br />

= 1<br />

2 δrn<br />

mn (summation on n) and show δ rst<br />

pst =2δr p<br />

<br />

<br />

<br />

in the determinant <br />

<br />

<br />

a1 1 a<strong>12</strong> a13 a2 1 a22 a23 a3 1 a3 2 a3 3<br />

<br />

<br />

<br />

<br />

as given by equation (1.1.25).<br />

<br />

(a) Show e rst A i r = eijk a s j at k (b) Show erstA r i = eijka j s ak t<br />

◮ 28. (a) Show that if Aijk = Ajik, i, j, k =1, 2, 3 there is a total of 27 elements, but only 18 are distinct.<br />

(b) Show that for i, j, k =1, 2,...,N there are N 3 elements, but only N 2 (N +1)/2 aredistinct.<br />

◮ 29. Let aij = BiBj for i, j =1, 2, 3whereB1,B2,B3 are arbitrary constants. Calculate det(aij) =|A|.<br />

◮ 30.<br />

(a) For A =(aij), i,j=1, 2, 3, show |A| = eijkai1aj2ak3.<br />

(b) For A =(a i j ),i,j=1, 2, 3, show |A| = eijka i 1 aj<br />

2 ak 3 .<br />

(c) For A =(a i j), i,j=1, 2, 3, show |A| = e ijk a 1 i a 2 ja 3 k.<br />

(d) For I =(δ i j), i,j=1, 2, 3, show |I| =1.<br />

◮ 31. Let |A| = eijkai1aj2ak3 and define Aim as the cofactor of aim. Show the determinant can be<br />

expressed in any of the forms:<br />

(a) |A| = Ai1ai1 where Ai1 = eijkaj2ak3<br />

(b) |A| = Aj2aj2 where Ai2 = ejikaj1ak3<br />

(c) |A| = Ak3ak3 where Ai3 = ejkiaj1ak2<br />

31


32<br />

◮ 32. Show the results in problem 31 can be written in the forms:<br />

Ai1 = 1<br />

2! e1steijkajsakt, Ai2 = 1<br />

2! e2steijkajsakt, Ai3 = 1<br />

2! e3steijkajsakt, or Aim = 1<br />

2! emsteijkajsakt<br />

◮ 33. Use the results in problems 31 and 32 to prove that apmAim = |A|δip.<br />

◮ 34.<br />

⎛<br />

1<br />

Let (aij) = ⎝ 1<br />

2<br />

0<br />

⎞<br />

1<br />

3⎠and<br />

calculate C = aijaij, i,j=1, 2, 3.<br />

2 3 2<br />

◮ 35. Let<br />

a111 = −1, a1<strong>12</strong> =3, a<strong>12</strong>1 =4, a<strong>12</strong>2 =2<br />

a211 =1, a2<strong>12</strong> =5, a221 =2, a222 = −2<br />

and calculate the quantity C = aijkaijk, i,j,k =1, 2.<br />

◮ 36. Let<br />

a1111 =2, a11<strong>12</strong> =1, a1<strong>12</strong>1 =3, a1<strong>12</strong>2 =1<br />

a<strong>12</strong>11 =5, a<strong>12</strong><strong>12</strong> = −2, a<strong>12</strong>21 =4, a<strong>12</strong>22 = −2<br />

a2111 =1, a21<strong>12</strong> =0, a2<strong>12</strong>1 = −2, a2<strong>12</strong>2 = −1<br />

a2211 = −2, a22<strong>12</strong> =1, a2221 =2, a2222 =2<br />

and calculate the quantity C = aijklaijkl, i,j,k,l=1, 2.<br />

◮ 37. Simplify the expressions:<br />

(a) (Aijkl + Ajkli + Aklij + Alijk)xixjxkxl<br />

(b) (Pijk + Pjki + Pkij)x i x j x k<br />

(c)<br />

∂x i<br />

∂x j<br />

(d) aij<br />

∂2xi ∂xt∂xs ∂x j<br />

∂x<br />

∂<br />

r − ami<br />

2xm ∂xs∂xt ∂xi ∂xr ◮ 38. Let g denote the determinant of the matrix having the components gij, i,j=1, 2, 3. Show that<br />

(a)<br />

<br />

<br />

g1r<br />

gerst = <br />

g2r<br />

<br />

g1s<br />

g2s<br />

<br />

g1t <br />

<br />

g2t <br />

<br />

<br />

(b)<br />

<br />

<br />

gir<br />

gersteijk = <br />

gjr<br />

<br />

gis<br />

gjs<br />

<br />

git <br />

<br />

gjt <br />

<br />

<br />

◮ 39. Show that e ijk emnp = δ ijk<br />

mnp =<br />

g3r g3s g3t<br />

<br />

<br />

<br />

<br />

<br />

<br />

δi m δi n δi δ<br />

p<br />

j m δj n δj p<br />

δk m δk n δk p<br />

◮ 40. Show that e ijk emnpA mnp = A ijk − A ikj + A kij − A jik + A jki − A kji<br />

Hint: Use the results from problem 39.<br />

◮ 41. Show that<br />

(a) e ij eij =2!<br />

(b) e ijk eijk =3!<br />

<br />

<br />

<br />

<br />

<br />

<br />

(c) e ijkl eijkl =4!<br />

gkr gks gkt<br />

(d) Guess at the result e i1i2...in ei1i2...in


◮ 42. Determine if the following statement is true or false. Justify your answer. eijkAiBjCk = eijkAjBkCi.<br />

◮ 43. Let aij, i,j=1, 2 denote the components of a 2 × 2matrixA, which are functions of time t.<br />

<br />

(a) Expand both |A| = eijai1aj2 and |A| = <br />

a11<br />

<br />

a<strong>12</strong> <br />

<br />

a21 a22 to verify that these representations are the same.<br />

(b) Verify the equivalence of the derivative relations<br />

d|A|<br />

dt<br />

dai1<br />

= eij<br />

dt aj2<br />

daj2<br />

+ eijai1<br />

dt<br />

and d|A|<br />

dt =<br />

<br />

<br />

<br />

<br />

da11 da<strong>12</strong><br />

dt dt<br />

a21 a22<br />

<br />

<br />

<br />

+<br />

<br />

<br />

<br />

a11<br />

da21<br />

dt<br />

(c) Let aij, i,j=1, 2, 3 denote the components of a 3 × 3matrixA, which are functions of time t. Develop<br />

appropriate relations, expand them and verify, similar to parts (a) and (b) above, the representation of<br />

a determinant and its derivative.<br />

◮ 44. For f = f(x1 ,x2 ,x3 )andφ = φ(f) differentiable scalar functions, use the indicial notation to find a<br />

formula to calculate grad φ.<br />

◮ 45. Use the indicial notation to prove (a) ∇×∇φ = 0 (b) ∇·∇× A =0<br />

◮ 46. If Aij is symmetric and Bij is skew-symmetric, i, j =1, 2, 3, then calculate C = AijBij.<br />

◮ 47. Assume Aij = Aij(x 1 , x 2 , x 3 )andAij = Aij(x 1 ,x 2 ,x 3 )fori, j =1, 2, 3 are related by the expression<br />

∂x<br />

Amn = Aij<br />

i<br />

∂xm ∂xj ∂Amn<br />

n . Calculate the derivative .<br />

∂x k ∂x<br />

◮ 48. Prove that if any two rows (or two columns) of a matrix are interchanged, then the value of the<br />

determinant of the matrix is multiplied by minus one. Construct your proof using 3 × 3 matrices.<br />

◮ 49. Prove that if two rows (or columns) of a matrix are proportional, then the value of the determinant<br />

of the matrix is zero. Construct your proof using 3 × 3 matrices.<br />

◮ 50. Prove that if a row (or column) of a matrix is altered by adding some constant multiple of some other<br />

row (or column), then the value of the determinant of the matrix remains unchanged. Construct your proof<br />

using 3 × 3 matrices.<br />

◮ 51. Simplify the expression φ = eijkeℓmnAiℓAjmAkn.<br />

◮ 52. Let Aijk denote a third order system where i, j, k =1, 2. (a) How many components does this system<br />

have? (b) Let Aijk be skew-symmetric in the last pair of indices, how many independent components does<br />

the system have?<br />

◮ 53. Let Aijk denote a third order system where i, j, k =1, 2, 3. (a) How many components does this<br />

system have? (b) In addition let Aijk = Ajik and Aikj = −Aijk and determine the number of distinct<br />

nonzero components for Aijk.<br />

a<strong>12</strong><br />

da22<br />

dt<br />

<br />

<br />

<br />

<br />

33


34<br />

◮ 54. Show that every second order system Tij can be expressed as the sum of a symmetric system Aij and<br />

skew-symmetric system Bij. Find Aij and Bij in terms of the components of Tij.<br />

◮ 55. Consider the system Aijk, i,j,k =1, 2, 3, 4.<br />

(a) How many components does this system have?<br />

(b) Assume Aijk is skew-symmetric in the last pair of indices, how many independent components does this<br />

system have?<br />

(c) Assume that in addition to being skew-symmetric in the last pair of indices, Aijk + Ajki + Akij =0is<br />

satisfied for all values of i, j, and k, then how many independent components does the system have?<br />

◮ 56. (a) Write the equation of a line r = r0 + t A in indicial form. (b) Write the equation of the plane<br />

n · (r − r0) = 0 in indicial form. (c) Write the equation of a general line in scalar form. (d) Write the<br />

equation of a plane in scalar form. (e) Find the equation of the line defined by the intersection of the<br />

planes 2x +3y +6z =<strong>12</strong>and6x +3y + z =6. (f) Find the equation of the plane through the points<br />

(5, 3, 2), (3, 1, 5), (1, 3, 3). Find also the normal to this plane.<br />

◮ 57. The angle 0 ≤ θ ≤ π between two skew lines in space is defined as the angle between their direction<br />

vectors when these vectors are placed at the origin. Show that for two lines with direction numbers ai and<br />

bi i =1, 2, 3, the cosine of the angle between these lines satisfies<br />

cos θ =<br />

aibi<br />

√ √<br />

aiai bibi<br />

◮ 58. Let aij = −aji for i, j =1, 2,...,N and prove that for N odd det(aij) =0.<br />

◮ 59. Let λ = Aijxixj where Aij = Aji and calculate (a)<br />

∂λ<br />

∂xm<br />

(b)<br />

∂ 2 λ<br />

∂xm∂xk<br />

◮ 60. Given an arbitrary nonzero vector Uk, k =1, 2, 3, define the matrix elements aij = eijkUk, whereeijk<br />

is the e-permutation symbol. Determine if aij is symmetric or skew-symmetric. Suppose Uk is defined by<br />

the above equation for arbitrary nonzero aij, thensolveforUkin terms of the aij.<br />

◮ 61. If Aij = AiBj = 0foralli, j values and Aij = Aji for i, j =1, 2,...,N, show that Aij = λBiBj<br />

where λ is a constant. State what λ is.<br />

◮ 62. Assume that Aijkm, withi, j, k, m =1, 2, 3, is completely skew-symmetric. How many independent<br />

components does this quantity have?<br />

◮ 63. Consider Rijkm, i, j, k, m =1, 2, 3, 4. (a) How many components does this quantity have? (b) If<br />

Rijkm = −Rijmk = −Rjikm then how many independent components does Rijkm have? (c) If in addition<br />

Rijkm = Rkmij determine the number of independent components.<br />

◮ 64. Let xi = aij ¯xj, i, j =1, 2, 3 denote a change of variables from a barred system of coordinates to an<br />

unbarred system of coordinates and assume that Āi = aijAj where aij are constants, Āi is a function of the<br />

∂<br />

¯xj variables and Aj is a function of the xj variables. Calculate<br />

Āi<br />

.<br />

∂¯xm


§1.2 TENSOR CONCEPTS AND TRANSFORMATIONS<br />

For e1, e2, e3 independent orthogonal unit vectors (base vectors), we may write any vector A as<br />

A = A1 e1 + A2 e2 + A3 e3<br />

where (A1,A2,A3) are the coordinates of A relative to the base vectors chosen. These components are the<br />

projection of A onto the base vectors and<br />

A =( A · e1) e1 +( A · e2) e2 +( A · e3) e3.<br />

Select any three independent orthogonal vectors, ( E1, E2, E3), not necessarily of unit length, we can then<br />

write<br />

e1 = E1<br />

| E1| , e2 = E2<br />

| E2| , e3 = E3<br />

| E3| ,<br />

and consequently, the vector A can be expressed as<br />

<br />

A<br />

A<br />

· E1 <br />

=<br />

E1 · <br />

A · E2 <br />

E1 +<br />

E1 E2 · <br />

A · E3 <br />

E2 +<br />

E2 E3 · <br />

E3.<br />

E3<br />

Here we say that<br />

A · E (i)<br />

E (i) · , i =1, 2, 3<br />

E (i)<br />

are the components of A relative to the chosen base vectors E1, E2, E3. Recall that the parenthesis about<br />

the subscript i denotes that there is no summation on this subscript. It is then treated as a free subscript<br />

which can have any of the values 1, 2or3.<br />

Reciprocal Basis<br />

Consider a set of any three independent vectors ( E1, E2, E3) which are not necessarily orthogonal, nor of<br />

unit length. In order to represent the vector A in terms of these vectors we must find components (A1 ,A2 ,A3 )<br />

such that<br />

A = A 1 E1 + A 2 E2 + A 3 E3. <br />

This can be done by taking appropriate projections and obtaining three equations and three unknowns from<br />

which the components are determined. A much easier way to find the components (A1 ,A2 ,A3 )istoconstruct<br />

a reciprocal basis ( E1 , E2 , E3 ). Recall that two bases ( E1, E2, E3) and( E1 , E2 , E3 ) are said to be reciprocal<br />

if they satisfy the condition<br />

Ei · E j = δ j<br />

i =<br />

<br />

1 if i = j<br />

0 if i = j .<br />

Note that E2 · E 1 = δ 1 2 =0 andE3 · E 1 = δ 1 3 = 0 so that the vector E1 is perpendicular to both the<br />

vectors E2 and E3. (i.e. A vector from one basis is orthogonal to two of the vectors from the other basis.)<br />

We can therefore write E1 = V −1E2 × E3 where V is a constant to be determined. By taking the dot<br />

product of both sides of this equation with the vector E1 we find that V = E1 · ( E2 × E3) isthevolume<br />

of the parallelepiped formed by the three vectors E1, E2, E3 when their origins are made to coincide. In a<br />

35


36<br />

similar manner it can be demonstrated that for ( E1, E2, E3) a given set of basis vectors, then the reciprocal<br />

basis vectors are determined from the relations<br />

E 1 = 1<br />

V E2 × E3, E 2 1<br />

=<br />

V E3 × E1, E 3 1<br />

=<br />

V E1 × E2,<br />

where V = E1 · ( E2 × E3) = 0 is a triple scalar product and represents the volume of the parallelepiped<br />

having the basis vectors for its sides.<br />

Let ( E1, E2, E3) and( E1 , E2 , E3 ) denote a system of reciprocal bases. We can represent any vector A<br />

with respect to either of these bases. If we select the basis ( E1, E2, E3) and represent A in the form<br />

A = A 1 E1 + A 2 E2 + A 3 E3, (1.2.1)<br />

then the components (A1 ,A2 ,A3 )of A relative to the basis vectors ( E1, E2, E3) are called the contravariant<br />

components of A. These components can be determined from the equations<br />

A · E 1 = A 1 , A · E 2 2<br />

= A , A · E 3 3<br />

= A .<br />

Similarly, if we choose the reciprocal basis ( E1 , E2 , E3 ) and represent A in the form<br />

A = A1 E 1 + A2 E 2 + A3 E 3 , (1.2.2)<br />

then the components (A1,A2,A3) relative to the basis ( E 1 , E 2 , E 3 ) are called the covariant components of<br />

A. These components can be determined from the relations<br />

A · E1 = A1, A · E2 = A2, A · E3 = A3.<br />

The contravariant and covariant components are different ways of representing the same vector with respect<br />

to a set of reciprocal basis vectors. There is a simple relationship between these components which we now<br />

develop. We introduce the notation<br />

Ei · Ej = gij = gji, and E i · E j = g ij = g ji<br />

(1.2.3)<br />

where gij are called the metric components of the space and gij are called the conjugate metric components<br />

of the space. We can then write<br />

A · E1 = A1( E 1 · E1)+A2( E 2 · E1)+A3( E 3 · E1) =A1<br />

or<br />

A · E1 = A 1 ( E1 · E1)+A 2 ( E2 · E1)+A 3 ( E3 · E1) =A1<br />

A1 = A 1 g11 + A 2 g<strong>12</strong> + A 3 g13. (1.2.4)<br />

In a similar manner, by considering the dot products A · E2 and A · E3 one can establish the results<br />

A2 = A 1 g21 + A 2 g22 + A 3 g23<br />

These results can be expressed with the index notation as<br />

A3 = A 1 g31 + A 2 g32 + A 3 g33.<br />

Ai = gikA k . (1.2.6)<br />

Forming the dot products A · E 1 , A · E 2 , A · E 3 it can be verified that<br />

A i = g ik Ak. (1.2.7)<br />

The equations (1.2.6) and (1.2.7) are relations which exist between the contravariant and covariant components<br />

of the vector A. Similarly, if for some value j we have Ej = α E1 + β E2 + γ E3, then one can show<br />

that E j = g ij Ei. This is left as an exercise.


Coordinate Transformations<br />

Consider a coordinate transformation from a set of coordinates (x, y, z) to(u, v, w) defined by a set of<br />

transformation equations<br />

x = x(u, v, w)<br />

y = y(u, v, w)<br />

z = z(u, v, w)<br />

(1.2.8)<br />

It is assumed that these transformations are single valued, continuous and possess the inverse transformation<br />

u = u(x, y, z)<br />

v = v(x, y, z)<br />

w = w(x, y, z).<br />

(1.2.9)<br />

These transformation equations define a set of coordinate surfaces and coordinate curves. The coordinate<br />

surfaces are defined by the equations<br />

u(x, y, z) =c1<br />

v(x, y, z) =c2<br />

w(x, y, z) =c3<br />

where c1,c2,c3 are constants. These surfaces intersect in the coordinate curves<br />

where<br />

(1.2.10)<br />

r(u, c2,c3), r(c1,v,c3), r(c1,c2,w), (1.2.11)<br />

r(u, v, w) =x(u, v, w) e1 + y(u, v, w) e2 + z(u, v, w) e3.<br />

The general situation is illustrated in the figure 1.2-1.<br />

Consider the vectors<br />

E 1 =gradu = ∇u, E 2 =gradv = ∇v, E 3 =gradw = ∇w (1.2.<strong>12</strong>)<br />

evaluated at the common point of intersection (c1,c2,c3) of the coordinate surfaces. The system of vectors<br />

( E1 , E2 , E3 ) can be selected as a system of basis vectors which are normal to the coordinate surfaces.<br />

Similarly, the vectors<br />

E1 = ∂r<br />

∂u , E2 = ∂r<br />

∂v , E3 = ∂r<br />

(1.2.13)<br />

∂w<br />

when evaluated at the common point of intersection (c1,c2,c3) forms a system of vectors ( E1, E2, E3) which<br />

we can select as a basis. This basis is a set of tangent vectors to the coordinate curves. It is now demonstrated<br />

that the normal basis ( E1 , E2 , E3 ) and the tangential basis ( E1, E2, E3) are a set of reciprocal bases.<br />

Recall that r = x e1 + y e2 + z e3 denotes the position vector of a variable point. By substitution for<br />

x, y, z from (1.2.8) there results<br />

r = r(u, v, w) =x(u, v, w) e1 + y(u, v, w) e2 + z(u, v, w) e3. (1.2.14)<br />

37


38<br />

A small change in r is denoted<br />

Figure 1.2-1. Coordinate curves and coordinate surfaces.<br />

dr = dx e1 + dy e2 + dz e3 = ∂r ∂r ∂r<br />

du + dv + dw (1.2.15)<br />

∂u ∂v ∂w<br />

where<br />

∂r ∂x<br />

=<br />

∂u ∂u e1 + ∂y<br />

∂u e2 + ∂z<br />

∂u e3<br />

∂r ∂x<br />

=<br />

∂v ∂v e1 + ∂y<br />

∂v e2 + ∂z<br />

∂v e3<br />

∂r ∂x<br />

=<br />

∂w ∂w e1 + ∂y<br />

∂w e2 + ∂z<br />

∂w e3.<br />

(1.2.16)<br />

In terms of the u, v, w coordinates, this change can be thought of as moving along the diagonal of a parallelepiped<br />

having the vector sides ∂r<br />

∂u du,<br />

∂r<br />

∂r<br />

dv, and<br />

∂v ∂w dw.<br />

Assume u = u(x, y, z) is defined by equation (1.2.9) and differentiate this relation to obtain<br />

du = ∂u ∂u ∂u<br />

dx + dy + dz. (1.2.17)<br />

∂x ∂y ∂z<br />

The equation (1.2.15) enables us to represent this differential in the form:<br />

du =gradu · dr<br />

<br />

∂r ∂r ∂r<br />

du =gradu · du + dv +<br />

∂u ∂v ∂w dw<br />

<br />

<br />

du = grad u · ∂r<br />

<br />

du + grad u ·<br />

∂u<br />

∂r<br />

<br />

dv + grad u ·<br />

∂v<br />

∂r<br />

<br />

dw.<br />

∂w<br />

By comparing like terms in this last equation we find that<br />

(1.2.18)<br />

E 1 · E1 =1, E 1 · E2 =0, E 1 · E3 =0. (1.2.19)<br />

Similarly, from the other equations in equation (1.2.9) which define v = v(x, y, z),<br />

can be demonstrated that<br />

and w = w(x, y, z) it<br />

<br />

dv = grad v · ∂r<br />

<br />

du + grad v ·<br />

∂u<br />

∂r<br />

<br />

dv + grad v ·<br />

∂v<br />

∂r<br />

<br />

dw<br />

∂w<br />

(1.2.20)


and<br />

<br />

dw = grad w · ∂r<br />

<br />

du + grad w ·<br />

∂u<br />

∂r<br />

<br />

dv + grad w ·<br />

∂v<br />

∂r<br />

<br />

dw.<br />

∂w<br />

(1.2.21)<br />

By comparing like terms in equations (1.2.20) and (1.2.21) we find<br />

E 2 · E1 =0, E 2 · E2 =1, E 2 · E3 =0<br />

E 3 · E1 =0, E 3 · E2 =0, E 3 · E3 =1.<br />

(1.2.22)<br />

The equations (1.2.22) and (1.2.19) show us that the basis vectors defined by equations (1.2.<strong>12</strong>) and (1.2.13)<br />

are reciprocal.<br />

Introducing the notation<br />

(x 1 ,x 2 ,x 3 )=(u, v, w) (y 1 ,y 2 ,y 3 )=(x, y, z) (1.2.23)<br />

where the x ′ s denote the generalized coordinates and the y ′ s denote the rectangular Cartesian coordinates,<br />

the above equations can be expressed in a more concise form with the index notation. For example, if<br />

x i = x i (x, y, z) =x i (y 1 ,y 2 ,y 3 ), and y i = y i (u, v, w) =y i (x 1 ,x 2 ,x 3 ), i =1, 2, 3 (1.2.24)<br />

then the reciprocal basis vectors can be represented<br />

E i =gradx i , i =1, 2, 3 (1.2.25)<br />

and<br />

Ei = ∂r<br />

, i =1, 2, 3. (1.2.26)<br />

∂xi We now show that these basis vectors are reciprocal. Observe that r = r(x 1 ,x2 ,x3 )with<br />

and consequently<br />

dr = ∂r<br />

dxm<br />

∂xm (1.2.27)<br />

dx i =gradx i · dr =gradx i · ∂r<br />

∂xm dxm <br />

= E i<br />

· Em dx m = δ i m dx m , i =1, 2, 3 (1.2.28)<br />

Comparing like terms in this last equation establishes the result that<br />

which demonstrates that the basis vectors are reciprocal.<br />

E i · Em = δ i m , i,m =1, 2, 3 (1.2.29)<br />

39


40<br />

Scalars, Vectors and Tensors<br />

Tensors are quantities which obey certain transformation laws. That is, scalars, vectors, matrices<br />

and higher order arrays can be thought of as components of a tensor quantity. We shall be interested in<br />

finding how these components are represented in various coordinate systems. We desire knowledge of these<br />

transformation laws in order that we can represent various physical laws in a form which is independent of<br />

the coordinate system chosen. Before defining different types of tensors let us examine what we mean by a<br />

coordinate transformation.<br />

Coordinate transformations of the type found in equations (1.2.8) and (1.2.9) can be generalized to<br />

higher dimensions. Let xi ,i = 1, 2,...,N denote N variables. These quantities can be thought of as<br />

representing a variable point (x1 ,x2 ,...,xN )inanNdimensional space VN . Another set of N quantities,<br />

call them barred quantities, xi ,i=1, 2,...,N, can be used to represent a variable point (x1 , x2 ,...,xN )in<br />

an N dimensional space V N . When the x ′ s are related to the x ′ s by equations of the form<br />

x i = x i (x 1 , x 2 ,...,x N ), i =1, 2,...,N (1.2.30)<br />

then a transformation is said to exist between the coordinates x i and x i ,i=1, 2,...,N. Whenever the<br />

relations (1.2.30) are functionally independent, single valued and possess partial derivatives such that the<br />

Jacobian of the transformation<br />

<br />

x<br />

1 2 N<br />

x ,x ,...,x<br />

J = J<br />

x x1 , x2 ,...,xN <br />

∂x<br />

<br />

<br />

= <br />

<br />

<br />

1<br />

∂x1 ∂x 1<br />

∂x2 ∂x ... 1<br />

∂xN <br />

<br />

<br />

<br />

. . ... .<br />

<br />

(1.2.31)<br />

<br />

<br />

∂x N<br />

∂x 1<br />

is different from zero, then there exists an inverse transformation<br />

∂x N<br />

∂x2 ∂x ... N<br />

∂xN x i = x i (x 1 ,x 2 ,...,x N ), i =1, 2,...,N. (1.2.32)<br />

For brevity the transformation equations (1.2.30) and (1.2.32) are sometimes expressed by the notation<br />

x i = x i (x), i=1,...,N and x i = x i (x), i=1,...,N. (1.2.33)<br />

Consider a sequence of transformations from x to ¯x andthenfrom¯xto ¯x coordinates. For simplicity<br />

let ¯x = y and ¯x = z. If we denote by T1,T2 and T3 the transformations<br />

T1 : y i = y i (x 1 ,...,x N ) i =1,...,N or T1x = y<br />

T2 : z i = z i (y 1 ,...,y N ) i =1,...,N or T2y = z<br />

Then the transformation T3 obtained by substituting T1 into T2 is called the product of two successive<br />

transformations and is written<br />

T3 : z i = z i (y 1 (x 1 ,...,x N ),...,y N (x 1 ,...,x N )) i =1,...,N or T3x = T2T1x = z.<br />

This product transformation is denoted symbolically by T3 = T2T1.<br />

The Jacobian of the product transformation is equal to the product of Jacobians associated with the<br />

product transformation and J3 = J2J1.


Transformations Form a Group<br />

AgroupGis a nonempty set of elements together with a law, for combining the elements. The combined<br />

elements are denoted by a product. Thus, if a and b are elements in G then no matter how you define the<br />

law for combining elements, the product combination is denoted ab. ThesetGandcombining law forms a<br />

group if the following properties are satisfied:<br />

(i) For all a, b ∈ G, thenab ∈ G. This is called the closure property.<br />

(ii) There exists an identity element I such that for all a ∈ G we have Ia = aI = a.<br />

(iii) There exists an inverse element. That is, for all a ∈ G there exists an inverse element a−1 such that<br />

aa −1 = a −1 a = I.<br />

(iv) The associative law holds under the combining law and a(bc) =(ab)c for all a, b, c ∈ G.<br />

For example, the set of elements G = {1, −1,i,−i}, wherei2 = −1 together with the combining law of<br />

ordinary multiplication, forms a group. This can be seen from the multiplication table.<br />

× 1 -1 i -i<br />

1 1 -1 i -i<br />

-1 -1 1 -i i<br />

-i -i i 1 -1<br />

i i -i -1 1<br />

The set of all coordinate transformations of the form found in equation (1.2.30), with Jacobian different<br />

from zero, forms a group because:<br />

(i) The product transformation, which consists of two successive transformations, belongs to the set of<br />

transformations. (closure)<br />

(ii) The identity transformation exists in the special case that x and x are the same coordinates.<br />

(iii) The inverse transformation exists because the Jacobian of each individual transformation is different<br />

from zero.<br />

(iv) The associative law is satisfied in that the transformations satisfy the property T3(T2T1) =(T3T2)T1.<br />

When the given transformation equations contain a parameter the combining law is often times represented<br />

as a product of symbolic operators. For example, we denote by Tα a transformation of coordinates<br />

having a parameter α. The inverse transformation can be denoted by T −1<br />

α and one can write Tαx = x or<br />

x = T −1<br />

α x. We let Tβ denote the same transformation, but with a parameter β, then the transitive property<br />

is expressed symbolically by TαTβ = Tγ where the product TαTβ represents the result of performing two<br />

successive transformations. The first coordinate transformation uses the given transformation equations and<br />

uses the parameter α in these equations. This transformation is then followed by another coordinate transformation<br />

using the same set of transformation equations, but this time the parameter value is β. The above<br />

symbolic product is used to demonstrate that the result of applying two successive transformations produces<br />

a result which is equivalent to performing a single transformation of coordinates having the parameter value<br />

γ. Usually some relationship can then be established between the parameter values α, β and γ.<br />

41


42<br />

Figure 1.2-2. Cylindrical coordinates.<br />

In this symbolic notation, we let Tθ denote the identity transformation. That is, using the parameter<br />

value of θ in the given set of transformation equations produces the identity transformation. The inverse<br />

transformation can then be expressed in the form of finding the parameter value β such that TαTβ = Tθ.<br />

Cartesian Coordinates<br />

At times it is convenient to introduce an orthogonal Cartesian coordinate system having coordinates<br />

y i , i =1, 2,...,N. This space is denoted EN and represents an N-dimensional Euclidean space. Whenever<br />

the generalized independent coordinates x i ,i=1,...,N are functions of the y ′ s, and these equations are<br />

functionally independent, then there exists independent transformation equations<br />

y i = y i (x 1 ,x 2 ,...,x N ), i =1, 2,...,N, (1.2.34)<br />

with Jacobian different from zero. Similarly, if there is some other set of generalized coordinates, say a barred<br />

system x i ,i=1,...,N where the x ′ s are independent functions of the y ′ s, then there will exist another set<br />

of independent transformation equations<br />

y i = y i (x 1 , x 2 ,...,x N ), i =1, 2,...,N, (1.2.35)<br />

with Jacobian different from zero. The transformations found in the equations (1.2.34) and (1.2.35) imply<br />

that there exists relations between the x ′ s and x ′ s of the form (1.2.30) with inverse transformations of the<br />

form (1.2.32). It should be remembered that the concepts and ideas developed in this section can be applied<br />

to a space VN of any finite dimension. Two dimensional surfaces (N = 2) and three dimensional spaces<br />

(N = 3) will occupy most of our applications. In relativity, one must consider spaces where N =4.<br />

EXAMPLE 1.2-1. (cylindrical coordinates (r, θ, z)) Consider the transformation<br />

x = x(r, θ, z) =r cos θ y = y(r, θ, z) =r sin θ z = z(r, θ, z) =z<br />

from rectangular coordinates (x, y, z) to cylindrical coordinates (r, θ, z), illustrated in the figure 1.2-2. By<br />

letting<br />

y 1 = x, y 2 = y, y 3 = z x 1 = r, x 2 = θ, x 3 = z<br />

the above set of equations are examples of the transformation equations (1.2.8) with u = r, v = θ, w = z as<br />

the generalized coordinates.


EXAMPLE 1.2.2. (Spherical Coordinates) (ρ, θ, φ)<br />

Consider the transformation<br />

x = x(ρ, θ, φ) =ρ sin θ cos φ y = y(ρ, θ, φ) =ρ sin θ sin φ z = z(ρ, θ, φ) =ρ cos θ<br />

from rectangular coordinates (x, y, z) to spherical coordinates (ρ, θ, φ). By letting<br />

y 1 = x, y 2 = y, y 3 = z x 1 = ρ, x 2 = θ,x 3 = φ<br />

the above set of equations has the form found in equation (1.2.8) with u = ρ, v = θ, w = φ the generalized<br />

coordinates. One could place bars over the x ′ s in this example in order to distinguish these coordinates from<br />

the x ′ s of the previous example. The spherical coordinates (ρ, θ, φ) are illustrated in the figure 1.2-3.<br />

Scalar Functions and Invariance<br />

Figure 1.2-3. Spherical coordinates.<br />

We are now at a point where we can begin to define what tensor quantities are. The first definition is<br />

for a scalar invariant or tensor of order zero.<br />

43


44<br />

Definition: ( Absolute scalar field) Assume there exists a coordinate<br />

transformation of the type (1.2.30) with Jacobian J different from zero. Let<br />

the scalar function<br />

f = f(x 1 ,x 2 ,...,x N ) (1.2.36)<br />

be a function of the coordinates xi ,i=1,...,N in a space VN . Whenever<br />

there exists a function<br />

f = f(x 1 , x 2 ,...,x N ) (1.2.37)<br />

which is a function of the coordinates x i ,i=1,...,N such that f = J W f,<br />

then f is called a tensor of rank or order zero of weight W in the space VN .<br />

Whenever W = 0, the scalar f is called the component of an absolute scalar<br />

field and is referred to as an absolute tensor of rank or order zero.<br />

That is, an absolute scalar field is an invariant object in the space VN with respect to the group of<br />

coordinate transformations. It has a single component in each coordinate system. For any scalar function<br />

of the type defined by equation (1.2.36), we can substitute the transformation equations (1.2.30) and obtain<br />

f = f(x 1 ,...,x N )=f(x 1 (x),...,x N (x)) = f(x 1 ,...,x N ). (1.2.38)<br />

Vector Transformation, Contravariant Components<br />

In VN consider a curve C defined by the set of parametric equations<br />

C : x i = x i (t), i =1,...,N<br />

where t is a parameter. The tangent vector to the curve C is the vector<br />

1<br />

T<br />

dx dx2<br />

= ,<br />

dt dt ,...,dxN<br />

<br />

.<br />

dt<br />

In index notation, which focuses attention on the components, this tangent vector is denoted<br />

T i = dxi<br />

, i =1,...,N.<br />

dt<br />

For a coordinate transformation of the type defined by equation (1.2.30) with its inverse transformation<br />

defined by equation (1.2.32), the curve C is represented in the barred space by<br />

x i = x i (x 1 (t),x 2 (t),...,x N (t)) = x i (t), i =1,...,N,<br />

with t unchanged. The tangent to the curve in the barred system of coordinates is represented by<br />

dx i<br />

dt<br />

= ∂xi<br />

∂x j<br />

dxj , i =1,...,N. (1.2.39)<br />

dt


Letting T i ,i=1,...,N denote the components of this tangent vector in the barred system of coordinates,<br />

the equation (1.2.39) can then be expressed in the form<br />

T i = ∂xi<br />

∂x j T j , i,j =1,...,N. (1.2.40)<br />

This equation is said to define the transformation law associated with an absolute contravariant tensor of<br />

rank or order one. In the case N = 3 the matrix form of this transformation is represented<br />

⎛ 1<br />

T<br />

⎝ T 2<br />

T 3<br />

⎞ ⎛<br />

∂x<br />

⎠ ⎜<br />

= ⎝<br />

1<br />

∂x1 ∂x 1<br />

∂x2 ∂x 1<br />

∂x3 ∂x 2<br />

∂x1 ∂x 2<br />

∂x2 ∂x 2<br />

∂x3 ⎞ ⎛<br />

1 T<br />

⎟<br />

⎠ ⎝ T 2<br />

T 3<br />

⎞<br />

⎠ (1.2.41)<br />

A more general definition is<br />

∂x 3<br />

∂x 1<br />

∂x 3<br />

∂x 2<br />

∂x 3<br />

∂x 3<br />

Definition: (Contravariant tensor) Whenever N quantities A i in<br />

acoordinatesystem(x 1 ,...,x N ) are related to N quantities A i in a<br />

coordinate system (x1 ,...,xN ) such that the Jacobian J is different<br />

from zero, then if the transformation law<br />

A i = J<br />

W ∂xi<br />

Aj<br />

∂xj is satisfied, these quantities are called the components of a relative tensor<br />

of rank or order one with weight W . Whenever W = 0 these quantities<br />

are called the components of an absolute tensor of rank or order one.<br />

We see that the above transformation law satisfies the group properties.<br />

EXAMPLE 1.2-3. (Transitive Property of Contravariant Transformation)<br />

Show that successive contravariant transformations is also a contravariant transformation.<br />

Solution: Consider the transformation of a vector from an unbarred to a barred system of coordinates. A<br />

vector or absolute tensor of rank one A i = A i (x), i=1,...,N will transform like the equation (1.2.40) and<br />

A i (x) = ∂xi<br />

∂x j Aj (x). (1.2.42)<br />

Another transformation from x → x coordinates will produce the components<br />

A i<br />

(x) = ∂xi<br />

∂x j Aj (x) (1.2.43)<br />

Here we have used the notation A j (x) to emphasize the dependence of the components A j upon the x<br />

coordinates. Changing indices and substituting equation (1.2.42) into (1.2.43) we find<br />

A i<br />

(x) = ∂xi<br />

∂xj ∂xj ∂xm Am (x). (1.2.44)<br />

45


46<br />

From the fact that<br />

the equation (1.2.44) simplifies to<br />

∂x i<br />

∂x j<br />

∂x j<br />

∂x<br />

∂xi<br />

= ,<br />

m ∂xm A i<br />

(x) = ∂xi<br />

∂x m Am (x) (1.2.45)<br />

and hence this transformation is also contravariant. We express this by saying that the above are transitive<br />

with respect to the group of coordinate transformations.<br />

Note that from the chain rule one can write<br />

∂x m<br />

∂x j<br />

∂xj ∂xm<br />

=<br />

∂xn ∂x1 Do not make the mistake of writing<br />

∂x m<br />

∂x 2<br />

∂x1 ∂xm<br />

+<br />

∂xn ∂x2 ∂x2 ∂xm<br />

=<br />

∂xn ∂xn or<br />

∂x2 ∂xm<br />

+<br />

∂xn ∂x3 ∂x m<br />

∂x 3<br />

∂x3 ∂xm<br />

=<br />

∂xn ∂xn = δm n .<br />

∂x3 ∂xm<br />

=<br />

∂xn ∂xn as these expressions are incorrect. Note that there are no summations in these terms, whereas there is a<br />

summation index in the representation of the chain rule.<br />

Vector Transformation, Covariant Components<br />

Consider a scalar invariant A(x) =A(x) which is a shorthand notation for the equation<br />

A(x 1 ,x 2 ,...,x n )=A(x 1 , x 2 ,...,x n )<br />

involving the coordinate transformation of equation (1.2.30). By the chain rule we differentiate this invariant<br />

and find that the components of the gradient must satisfy<br />

Let<br />

Aj = ∂A<br />

∂xj and Ai = ∂A<br />

, i ∂x<br />

then equation (1.2.46) can be expressed as the transformation law<br />

∂A ∂A<br />

i =<br />

∂x ∂xj ∂xj i . (1.2.46)<br />

∂x<br />

Ai = Aj<br />

∂xj i . (1.2.47)<br />

∂x<br />

This is the transformation law for an absolute covariant tensor of rank or order one. A more general definition<br />

is


Definition: (Covariant tensor) Whenever N quantities Ai in a<br />

coordinate system (x1 ,...,xN ) are related to N quantities Ai in a coordinate<br />

system (x1 ,...,xN ), with Jacobian J different from zero, such<br />

that the transformation law<br />

Ai = J<br />

W ∂xj<br />

i<br />

Aj<br />

(1.2.48)<br />

∂x<br />

is satisfied, then these quantities are called the components of a relative<br />

covariant tensor of rank or order one having a weight of W . Whenever<br />

W = 0, these quantities are called the components of an absolute<br />

covariant tensor of rank or order one.<br />

Again we note that the above transformation satisfies the group properties. Absolute tensors of rank or<br />

order one are referred to as vectors while absolute tensors of rank or order zero are referred to as scalars.<br />

EXAMPLE 1.2-4. (Transitive Property of Covariant Transformation)<br />

Consider a sequence of transformation laws of the type defined by the equation (1.2.47)<br />

x → x<br />

x → x<br />

Ai(x) =Aj(x) ∂xj<br />

∂x i<br />

Ak(x) =Am(x) ∂xm<br />

∂x k<br />

We can therefore express the transformation of the components associated with the coordinate transformation<br />

x → x and<br />

<br />

Ak(x) = Aj(x) ∂xj<br />

∂xm m ∂x ∂xj<br />

= Aj(x) ,<br />

k k<br />

∂x ∂x<br />

which demonstrates the transitive property of a covariant transformation.<br />

Higher Order Tensors<br />

We have shown that first order tensors are quantities which obey certain transformation laws. Higher<br />

order tensors are defined in a similar manner and also satisfy the group properties. We assume that we are<br />

given transformations of the type illustrated in equations (1.2.30) and (1.2.32) which are single valued and<br />

continuous with Jacobian J different from zero. Further, the quantities x i and x i ,i=1,...,n represent the<br />

coordinates in any two coordinate systems. The following transformation laws define second order and third<br />

order tensors.<br />

47


48<br />

Definition: (Second order contravariant tensor) Whenever N-squared quantities A ij<br />

in a coordinate system (x 1 ,...,x N ) are related to N-squared quantities A mn in a coordinate<br />

system (x 1 ,...,x N ) such that the transformation law<br />

A mn (x) =A ij W ∂xm<br />

(x)J<br />

∂xi ∂xn ∂xj (1.2.49)<br />

is satisfied, then these quantities are called components of a relative contravariant tensor of<br />

rank or order two with weight W . Whenever W = 0 these quantities are called the components<br />

of an absolute contravariant tensor of rank or order two.<br />

Definition: (Second order covariant tensor) Whenever N-squared quantities<br />

Aij in a coordinate system (x 1 ,...,x N ) are related to N-squared quantities Amn<br />

in a coordinate system (x 1 ,...,x N ) such that the transformation law<br />

W ∂xi<br />

Amn(x) =Aij(x)J<br />

∂xm ∂xj ∂xn (1.2.50)<br />

is satisfied, then these quantities are called components of a relative covariant tensor<br />

of rank or order two with weight W . Whenever W = 0 these quantities are called<br />

the components of an absolute covariant tensor of rank or order two.<br />

Definition: (Second order mixed tensor) Whenever N-squared quantities<br />

A i j in a coordinate system (x1 ,...,x N ) are related to N-squared quantities A m<br />

n in<br />

acoordinatesystem(x 1 ,...,x N ) such that the transformation law<br />

A m<br />

n (x) =A i W ∂xm<br />

j(x)J<br />

∂xi ∂xj ∂xn (1.2.51)<br />

is satisfied, then these quantities are called components of a relative mixed tensor of<br />

rank or order two with weight W . Whenever W = 0 these quantities are called the<br />

components of an absolute mixed tensor of rank or order two. It is contravariant<br />

of order one and covariant of order one.<br />

Higher order tensors are defined in a similar manner. For example, if we can find N-cubed quantities<br />

Am np such that<br />

A i<br />

jk(x) =A γ W ∂xi<br />

αβ (x)J<br />

∂xγ ∂xα ∂xj ∂xβ ∂xk (1.2.52)<br />

then this is a relative mixed tensor of order three with weight W .<br />

covariant of order two.<br />

It is contravariant of order one and


General Definition<br />

In general a mixed tensor of rank or order (m + n)<br />

T i1i2...im<br />

j1j2...jn<br />

is contravariant of order m and covariant of order n if it obeys the transformation law<br />

T i1i2...im<br />

j1j2...jn =<br />

<br />

J<br />

<br />

x<br />

W x<br />

T a1a2...am<br />

b1b2...bn<br />

∂xi1 ∂xa1 ∂xi2 ∂xim ∂xb1<br />

··· · a2 am ∂x ∂x ∂xj1 ∂xb2 ···∂xbn<br />

j2 ∂x ∂xjn (1.2.53)<br />

(1.2.54)<br />

where<br />

<br />

x<br />

<br />

<br />

J = <br />

∂x<br />

<br />

x ∂x<br />

= ∂(x1 ,x2 ,...,xN )<br />

∂(x1 , x2 ,...,xN )<br />

is the Jacobian of the transformation. When W = 0 the tensor is called an absolute tensor, otherwise it is<br />

called a relative tensor of weight W.<br />

Here superscripts are used to denote contravariant components and subscripts are used to denote covariant<br />

components. Thus, if we are given the tensor components in one coordinate system, then the components<br />

in any other coordinate system are determined by the transformation law of equation (1.2.54). Throughout<br />

the remainder of this text one should treat all tensors as absolute tensors unless specified otherwise.<br />

Dyads and Polyads<br />

Note that vectors can be represented in bold face type with the notation<br />

A = AiE i<br />

This notation can also be generalized to tensor quantities. Higher order tensors can also be denoted by bold<br />

face type. For example the tensor components Tij and Bijk can be represented in terms of the basis vectors<br />

E i ,i=1,...,N by using a notation which is similar to that for the representation of vectors. For example,<br />

T = TijE i E j<br />

B = BijkE i E j E k .<br />

Here T denotes a tensor with components Tij and B denotes a tensor with components Bijk. The quantities<br />

EiEj are called unit dyads and EiEjEk are called unit triads. There is no multiplication sign between the<br />

basis vectors. This notation is called a polyad notation. A further generalization of this notation is the<br />

representation of an arbitrary tensor using the basis and reciprocal basis vectors in bold type. For example,<br />

a mixed tensor would have the polyadic representation<br />

T = T ij...k<br />

lm...n EiEj ...EkE l E m ...E n .<br />

A dyadic is formed by the outer or direct product of two vectors. For example, the outer product of the<br />

vectors<br />

a = a1E 1 + a2E 2 + a3E 3<br />

and b = b1E 1 + b2E 2 + b3E 3<br />

49


50<br />

gives the dyad<br />

In general, a dyad can be represented<br />

ab =a1b1E 1 E 1 + a1b2E 1 E 2 + a1b3E 1 E 3<br />

a2b1E 2 E 1 + a2b2E 2 E 2 + a2b3E 2 E 3<br />

a3b1E 3 E 1 + a3b2E 3 E 2 + a3b3E 3 E 3 .<br />

A = AijE i E j<br />

i, j =1,...,N<br />

where the summation convention is in effect for the repeated indices. The coefficients Aij are called the<br />

coefficients of the dyad. When the coefficients are written as an N × N array it is called a matrix. Every<br />

second order tensor can be written as a linear combination of dyads. The dyads form a basis for the second<br />

order tensors. As the example above illustrates, the nine dyads {E1E1 , E1E2 ,...,E3E3 }, associated with<br />

the outer products of three dimensional base vectors, constitute a basis for the second order tensor A = ab<br />

having the components Aij = aibj with i, j =1, 2, 3. Similarly, a triad has the form<br />

T = TijkE i E j E k Sum on repeated indices<br />

where i, j, k have the range 1, 2,...,N.The set of outer or direct products { EiEjEk },withi, j, k =1,...,N<br />

constitutes a basis for all third order tensors. Tensor components with mixed suffixes like Ci jk are associated<br />

with triad basis of the form<br />

C = C i jk EiE j E k<br />

where i, j, k have the range 1, 2,...N.Dyads are associated with the outer product of two vectors, while triads,<br />

tetrads,... are associated with higher-order outer products. These higher-order outer or direct products are<br />

referred to as polyads.<br />

The polyad notation is a generalization of the vector notation. The subject of how polyad components<br />

transform between coordinate systems is the subject of tensor calculus.<br />

In Cartesian coordinates we have Ei = Ei = ei and a dyadic with components called dyads is written<br />

A = Aij ei ej or<br />

A =A11 e1 e1 + A<strong>12</strong> e1 e2 + A13 e1 e3<br />

A21 e2 e1 + A22 e2 e2 + A23 e2 e3<br />

A31 e3 e1 + A32 e3 e2 + A33 e3 e3<br />

where the terms ei ej are called unit dyads. Note that a dyadic has nine components as compared with a<br />

vector which has only three components. The conjugate dyadic Ac is defined by a transposition of the unit<br />

vectors in A, toobtain<br />

Ac =A11 e1 e1 + A<strong>12</strong> e2 e1 + A13 e3 e1<br />

A21 e1 e2 + A22 e2 e2 + A23 e3 e2<br />

A31 e1 e3 + A32 e2 e3 + A33 e3 e3


If a dyadic equals its conjugate A = Ac, thenAij = Aji and the dyadic is called symmetric. If a dyadic<br />

equals the negative of its conjugate A = −Ac, thenAij = −Aji and the dyadic is called skew-symmetric. A<br />

special dyadic called the identical dyadic or idemfactor is defined by<br />

J = e1 e1 + e2 e2 + e3 e3.<br />

This dyadic has the property that pre or post dot product multiplication of J with a vector V produces the<br />

same vector V.For example,<br />

V · J =(V1e1 + V2 e2 + V3 e3) · J<br />

= V1 e1 · e1 e1 + V2 e2 · e2 e2 + V3 e3 · e3 e3 = V<br />

and J · V = J · (V1 e1 + V2 e2 + V3 e3)<br />

= V1 e1 e1 · e1 + V2 e2 e2 · e2 + V3 e3 e3 · e3 = V<br />

A dyadic operation often used in physics and chemistry is the double dot product A : B where A and<br />

B are both dyadics. Here both dyadics are expanded using the distributive law of multiplication, and then<br />

each unit dyad pair ei ej : em en are combined according to the rule<br />

ei ej : em en =(ei · em)( ej · en).<br />

For example, if A = Aij ei ej and B = Bij ei ej, then the double dot product A : B is calculated as follows.<br />

A : B =(Aij ei ej) :(Bmn em en) =AijBmn( ei ej : em en) =AijBmn( ei · em)( ej · en)<br />

= AijBmnδimδjn = AmjBmj<br />

= A11B11 + A<strong>12</strong>B<strong>12</strong> + A13B13<br />

+ A21B21 + A22B22 + A23B23<br />

+ A31B31 + A32B32 + A33B33<br />

When operating with dyads, triads and polyads, there is a definite order to the way vectors and polyad<br />

components are represented. For example, for A = Ai ei and B = Bi ei vectors with outer product<br />

A B = AmBn em en = φ<br />

there is produced the dyadic φ with components AmBn. In comparison, the outer product<br />

B A = BmAn em en = ψ<br />

produces the dyadic ψ with components BmAn. That is<br />

φ = A B =A1B1 e1 e1 + A1B2 e1 e2 + A1B3 e1 e3<br />

A2B1 e2 e1 + A2B2 e2 e2 + A2B3 e2 e3<br />

A3B1 e3 e1 + A3B2 e3 e2 + A3B3 e3 e3<br />

and ψ = B A =B1A1 e1 e1 + B1A2 e1 e2 + B1A3 e1 e3<br />

B2A1 e2 e1 + B2A2 e2 e2 + B2A3 e2 e3<br />

B3A1 e3 e1 + B3A2 e3 e2 + B3A3 e3 e3<br />

are different dyadics.<br />

The scalar dot product of a dyad with a vector C is defined for both pre and post multiplication as<br />

φ · C = A B · C = A( B · C)<br />

C · φ = C · A B =( C · A) B<br />

These products are, in general, not equal.<br />

51


52<br />

Operations Using Tensors<br />

The following are some important tensor operations which are used to derive special equations and to<br />

prove various identities.<br />

Addition and Subtraction<br />

Tensors of the same type and weight can be added or subtracted. For example, two third order mixed<br />

tensors, when added, produce another third order mixed tensor. Let A i jk and Bi jk<br />

mixed tensors. Their sum is denoted<br />

C i jk = A i jk + B i jk.<br />

denote two third order<br />

That is, like components are added. The sum is also a mixed tensor as we now verify. By hypothesis A i jk<br />

and B i jk<br />

are third order mixed tensors and hence must obey the transformation laws<br />

A i<br />

jk = Am ∂x<br />

np<br />

i<br />

∂xm ∂xn ∂xj ∂xp ∂xk B i<br />

jk = Bm ∂x<br />

np<br />

i<br />

∂xm ∂xn ∂xj ∂xp . k ∂x<br />

We let C i<br />

jk = Aijk<br />

+ Bijk<br />

denote the sum in the transformed coordinates. Then the addition of the above<br />

transformation equations produces<br />

C i<br />

jk =<br />

<br />

A i<br />

<br />

jk + Bijk<br />

= A m np + Bm ∂x<br />

np<br />

i<br />

∂xm ∂xn ∂xj ∂xp ∂xk = Cm ∂x<br />

np<br />

i<br />

∂xm ∂xn ∂xj ∂xp . k ∂x<br />

Consequently, the sum transforms as a mixed third order tensor.<br />

Multiplication (Outer Product)<br />

The product of two tensors is also a tensor. The rank or order of the resulting tensor is the sum of<br />

the ranks of the tensors occurring in the multiplication. As an example, let Ai jk<br />

tensor and let Bl m<br />

order tensor<br />

denote a mixed third order<br />

denote a mixed second order tensor. The outer product of these two tensors is the fifth<br />

C il<br />

jkm = AijkBl m , i,j,k,l,m=1, 2,...,N.<br />

Here all indices are free indices as i, j, k, l, m take on any of the integer values 1, 2,...,N. Let A i<br />

jk and Blm<br />

denote the components of the given tensors in the barred system of coordinates. We define C il<br />

jkm as the<br />

outer product of these components. Observe that Cil jkm is a tensor for by hypothesis Ai jk and Bl m are tensors<br />

and hence obey the transformation laws<br />

A α<br />

βγ = Ai ∂x<br />

jk<br />

α<br />

∂xi ∂xj ∂xβ ∂xk ∂xγ B δ<br />

ɛ = B l m<br />

The outer product of these components produces<br />

which demonstrates that C il<br />

jkm<br />

analyzed in a similar way.<br />

∂xδ ∂xl ∂xm ɛ .<br />

∂x<br />

C αδ<br />

βγɛ = AαβγBδɛ<br />

= AijkBl ∂x<br />

m<br />

α<br />

∂xi ∂xj ∂xβ ∂xk ∂xγ ∂xδ ∂xl ∂xm ∂xɛ (1.2.55)<br />

= C il ∂x<br />

jkm<br />

α<br />

∂xi ∂xj ∂xβ ∂xk ∂xγ ∂xδ ∂xl ∂xm ∂xɛ (1.2.56)<br />

transforms as a mixed fifth order absolute tensor. Other outer products are


Contraction<br />

The operation of contraction on any mixed tensor of rank m is performed when an upper index is<br />

set equal to a lower index and the summation convention is invoked. When the summation is performed<br />

over the repeated indices the resulting quantity is also a tensor of rank or order (m − 2). For example, let<br />

Ai jk , i,j,k =1, 2,...,N denote a mixed tensor and perform a contraction by setting j equal to i. We obtain<br />

A i ik = A11k + A22k + ···+ ANNk where k is a free index. To show that Ak is a tensor, we let A i<br />

ik = Ak denote the contraction on the<br />

transformed components of Ai jk . By hypothesis Aijk is a mixed tensor and hence the components must<br />

= Ak<br />

(1.2.57)<br />

satisfy the transformation law<br />

A i<br />

jk = Am ∂x<br />

np<br />

i<br />

∂xm ∂xn ∂xj ∂xp . k ∂x<br />

Now execute a contraction by setting j equal to i and perform a summation over the repeated index. We<br />

find<br />

A i<br />

ik = Ak = A m ∂x<br />

np<br />

i<br />

∂xm ∂xn ∂xi ∂xp ∂xk = Am ∂x<br />

np<br />

n<br />

∂xm ∂xp ∂xk = A m npδ n ∂x<br />

m<br />

p<br />

∂xk = An ∂x<br />

np<br />

p ∂x<br />

= Ap k ∂x p<br />

(1.2.58)<br />

. k ∂x<br />

Hence, the contraction produces a tensor of rank two less than the original tensor. Contractions on other<br />

mixed tensors can be analyzed in a similar manner.<br />

New tensors can be constructed from old tensors by performing a contraction on an upper and lower<br />

index. This process can be repeated as long as there is an upper and lower index upon which to perform the<br />

contraction. Each time a contraction is performed the rank of the resulting tensor is two less than the rank<br />

of the original tensor.<br />

Multiplication (Inner Product)<br />

The inner product of two tensors is obtained by:<br />

(i) first taking the outer product of the given tensors and<br />

(ii) performing a contraction on two of the indices.<br />

EXAMPLE 1.2-5. (Inner product)<br />

Let Ai and Bj denote the components of two first order tensors (vectors). The outer product of these<br />

tensors is<br />

The inner product of these tensors is the scalar<br />

C i j = A i Bj, i,j=1, 2,...,N.<br />

C = A i Bi = A 1 B1 + A 2 B2 + ···+ A N BN .<br />

Note that in some situations the inner product is performed by employing only subscript indices. For<br />

example, the above inner product is sometimes expressed as<br />

C = AiBi = A1B1 + A2B2 + ···AN BN .<br />

This notation is discussed later when Cartesian tensors are considered.<br />

53


54<br />

Quotient Law<br />

Assume Bqs r and Cs p are arbitrary absolute tensors. Further assume we have a quantity A(ijk) which<br />

we think might be a third order mixed tensor Ai jk . By showing that the equation<br />

A r qp Bqs<br />

r = Cs p<br />

is satisfied, then it follows that Ar qp must be a tensor. This is an example of the quotient law. Obviously,<br />

this result can be generalized to apply to tensors of any order or rank. To prove the above assertion we shall<br />

show from the above equation that A i jk is a tensor. Let xi and x i denote a barred and unbarred system of<br />

coordinates which are related by transformations of the form defined by equation (1.2.30). In the barred<br />

system, we assume that<br />

(1.2.59)<br />

A r<br />

qpBqs r = Csp<br />

where by hypothesis B ij<br />

k and Cl m are arbitrary absolute tensors and therefore must satisfy the transformation<br />

equations<br />

We substitute for B qs<br />

r<br />

B qs<br />

r<br />

C s<br />

∂x<br />

= Bij<br />

k<br />

q<br />

p = Cl m<br />

∂xi ∂xs ∂xj ∂xk ∂xr ∂xs ∂xl ∂xm p .<br />

∂x<br />

and Csp<br />

in the equation (1.2.59) and obtain the equation<br />

A r<br />

<br />

qp B ij ∂x<br />

k<br />

q<br />

∂xi ∂xs ∂xj ∂xk ∂xr <br />

= C l ∂x<br />

m<br />

s<br />

∂xl ∂xm ∂xp <br />

= A r qmBql ∂x<br />

r<br />

s<br />

∂xl ∂xm p .<br />

∂x<br />

Since the summation indices are dummy indices they can be replaced by other symbols. We change l to j,<br />

q to i and r to k and write the above equation as<br />

∂xs ∂xj <br />

A r ∂x<br />

qp<br />

q<br />

∂xi ∂xk ∂xr − Ak ∂x<br />

im<br />

m<br />

∂xp <br />

B ij<br />

k =0.<br />

Use inner multiplication by ∂xn<br />

s ∂x and simplify this equation to the form<br />

Because B in<br />

k<br />

δ n j<br />

<br />

A r<br />

qp<br />

<br />

A r<br />

qp<br />

∂xq ∂xi ∂xq ∂xi ∂xk ∂xr − Akim ∂xk ∂xr − Akim ∂xm ∂xp ∂x m<br />

∂x p<br />

<br />

B ij<br />

k =0 or<br />

<br />

B in<br />

k =0.<br />

is an arbitrary tensor, the quantity inside the brackets is zero and therefore<br />

A r<br />

qp<br />

∂xk ∂xr − Ak ∂x<br />

im<br />

m<br />

p =0.<br />

∂x<br />

This equation is simplified by inner multiplication by ∂xi<br />

∂xj ∂x l<br />

∂xk to obtain<br />

∂xq ∂xi δ q<br />

j δl rArqp − Ak ∂x<br />

im<br />

m<br />

∂xp ∂xi ∂xj A l<br />

jp = A k im<br />

∂x m<br />

∂x p<br />

∂x i<br />

∂x j<br />

∂x l<br />

=0 or<br />

∂xk ∂x l<br />

∂x k<br />

which is the transformation law for a third order mixed tensor.


EXERCISE 1.2<br />

◮ 1. Consider the transformation equations representing a rotation of axes through an angle α.<br />

Tα :<br />

1 x 1 2 = x cos α − x sin α<br />

x2 = x1 sin α + x2 cos α<br />

Treat α as a parameter and show this set of transformations constitutes a group by finding the value of α<br />

which:<br />

(i) gives the identity transformation.<br />

(ii) gives the inverse transformation.<br />

(iii) show the transformation is transitive in that a transformation with α = θ1 followed by a transformation<br />

with α = θ2 is equivalent to the transformation using α = θ1 + θ2.<br />

◮ 2. Show the transformation<br />

1 1 x = αx<br />

Tα :<br />

x2 = 1<br />

αx2 forms a group with α as a parameter. Find the value of α such that:<br />

(i) the identity transformation exists.<br />

(ii) the inverse transformation exists.<br />

(iii) the transitive property is satisfied.<br />

◮ 3. Show the given transformation forms a group with parameter α.<br />

Tα :<br />

x 1 = x 1<br />

1−αx 1<br />

x 2 = x2<br />

1−αx 1<br />

◮ 4. Consider the Lorentz transformation from relativity theory having the velocity parameter V, c is the<br />

speed of light and x4 = t is time.<br />

⎧<br />

TV :<br />

⎪⎨ x2 = x2 ⎪⎩<br />

x1 = x1−Vx 4<br />

<br />

V 1− 2<br />

c2 x 3 = x 3<br />

x4 = x4− Vx1<br />

c2 <br />

V 1− 2<br />

c2 Show this set of transformations constitutes a group, by establishing:<br />

(i) V = 0 gives the identity transformation T0.<br />

(ii) TV2 · TV1 = T0 requires that V2 = −V1.<br />

(iii) TV2 · TV1 = TV3 requires that<br />

V3 = V1 + V2<br />

1+ V1V2<br />

c2 .<br />

◮ 5. For ( E1, E2, E3) an arbitrary independent basis, (a) Verify that<br />

E 1 = 1<br />

V E2 × E3,<br />

E 2 = 1<br />

V E3 × E1,<br />

E 3 = 1<br />

V E1 × E2<br />

is a reciprocal basis, where V = E1 · ( E2 × E3) (b) Show that E j = g ij Ei.<br />

55


56<br />

Figure 1.2-4. Cylindrical coordinates (r, β, z).<br />

◮ 6. For the cylindrical coordinates (r, β, z) illustrated in the figure 1.2-4.<br />

(a) Write out the transformation equations from rectangular (x, y, z) coordinates to cylindrical (r, β, z)<br />

coordinates. Also write out the inverse transformation.<br />

(b) Determine the following basis vectors in cylindrical coordinates and represent your results in terms of<br />

cylindrical coordinates.<br />

(i) The tangential basis E1, E2, E3. (ii)The normal basis E1 , E2 , E3 . (iii) êr, êβ, êz<br />

where êr, êβ, êz are normalized vectors in the directions of the tangential basis.<br />

(c) A vector A = Ax e1 + Ay e2 + Az e3 can be represented in any of the forms:<br />

A = A 1 E1 + A 2 E2 + A 3 E3<br />

A = A1 E 1 + A2 E 2 + A3 E 3<br />

A = Arêr + Aβêβ + Azêz<br />

depending upon the basis vectors selected . In terms of the components Ax,Ay,Az<br />

(i) Solve for the contravariant components A1 ,A2 ,A3 .<br />

(ii) Solve for the covariant components A1,A2,A3.<br />

(iii) Solve for the components Ar,Aβ,Az. Express all results in cylindrical coordinates. (Note the<br />

components Ar,Aβ,Az are referred to as physical components. Physical components are considered in<br />

more detail in a later section.)


Figure 1.2-5. Spherical coordinates (ρ, α, β).<br />

◮ 7. For the spherical coordinates (ρ, α, β) illustrated in the figure 1.2-5.<br />

(a) Write out the transformation equations from rectangular (x, y, z) coordinates to spherical (ρ, α, β) coordinates.<br />

Also write out the equations which describe the inverse transformation.<br />

(b) Determine the following basis vectors in spherical coordinates<br />

(i) The tangential basis E1, E2, E3.<br />

(ii) The normal basis E1 , E2 , E3 .<br />

(iii) êρ, êα, êβ which are normalized vectors in the directions of the tangential basis. Express all results<br />

in terms of spherical coordinates.<br />

(c) A vector A = Ax e1 + Ay e2 + Az e3 can be represented in any of the forms:<br />

A = A 1 E1 + A 2 E2 + A 3 E3<br />

A = A1 E 1 + A2 E 2 + A3 E 3<br />

A = Aρêρ + Aαêα + Aβêβ<br />

depending upon the basis vectors selected . Calculate, in terms of the coordinates (ρ, α, β) andthe<br />

components Ax,Ay,Az<br />

(i) The contravariant components A1 ,A2 ,A3 .<br />

(ii) The covariant components A1,A2,A3.<br />

(iii) The components Aρ,Aα,Aβ which are called physical components.<br />

◮ 8. Work the problems 6,7 and then let (x1 ,x2 ,x3 )=(r, β, z) denote the coordinates in the cylindrical<br />

system and let (x1 , x2 , x3 )=(ρ, α, β) denote the coordinates in the spherical system.<br />

(a) Write the transformation equations x → x from cylindrical to spherical coordinates. Also find the<br />

inverse transformations. ( Hint: See the figures 1.2-4 and 1.2-5.)<br />

(b) Use the results from part (a) and the results from problems 6,7 to verify that<br />

∂x<br />

Ai = Aj<br />

j<br />

∂xi for i =1, 2, 3.<br />

(i.e. Substitute Aj from problem 6 to get Āi given in problem 7.)<br />

57


58<br />

(c) Use the results from part (a) and the results from problems 6,7 to verify that<br />

A i j ∂xi<br />

= A<br />

∂xj for i =1, 2, 3.<br />

(i.e. Substitute Aj from problem 6 to get Āi given by problem 7.)<br />

◮ 9. Pick two arbitrary noncolinear vectors in the x, y plane, say<br />

V1 =5e1 + e2 and V2 = e1 +5e2<br />

and let V3 = e3 be a unit vector perpendicular to both V1 and V2. The vectors V1 and V2 can be thought of<br />

as defining an oblique coordinate system, as illustrated in the figure 1.2-6.<br />

(a) Find the reciprocal basis ( V 1 , V 2 , V 3 ).<br />

(b) Let<br />

r = x e1 + y e2 + z e3 = α V1 + β V2 + γ V3<br />

and show that<br />

(c) Show<br />

α = 5x y<br />

−<br />

24 24<br />

β = − x 5y<br />

+<br />

24 24<br />

γ = z<br />

x =5α + β<br />

y = α +5β<br />

z = γ<br />

(d) For γ = γ0 constant, show the coordinate lines are described by α = constant<br />

and sketch some of these coordinate lines. (See figure 1.2-6.)<br />

and β = constant,<br />

(e) Find the metrics gij and conjugate metrices gij associated with the (α, β, γ) space.<br />

Figure 1.2-6. Oblique coordinates.


◮ 10. Consider the transformation equations<br />

substituted into the position vector<br />

Define the basis vectors<br />

with the reciprocal basis<br />

where<br />

E 1 = 1<br />

V E2 × E3,<br />

x = x(u, v, w)<br />

y = y(u, v, w)<br />

z = z(u, v, w)<br />

r = x e1 + y e2 + z e3.<br />

( E1, E2, E3) =<br />

Let v = E1 · ( E2 × E3 )andshowthatv · V =1.<br />

◮ 11. Given the coordinate transformation<br />

<br />

∂r ∂r ∂r<br />

, ,<br />

∂u ∂v ∂w<br />

E 2 = 1<br />

V E3 × E1,<br />

V = E1 · ( E2 × E3).<br />

x = −u − 2v y = −u − v z = z<br />

(a) Find and illustrate graphically some of the coordinate curves.<br />

(b) For r = r(u, v, z) a position vector, define the basis vectors<br />

E1 = ∂r<br />

∂u ,<br />

E2 = ∂r<br />

∂v ,<br />

E 3 = 1<br />

V E1 × E2.<br />

E3 = ∂r<br />

∂z .<br />

Calculate these vectors and then calculate the reciprocal basis E1 , E2 , E3 .<br />

(c) With respect to the basis vectors in (b) find the contravariant components Ai associated with the vector<br />

A = α1 e1 + α2 e2 + α3 e3<br />

where (α1,α2,α3) areconstants.<br />

(d) Find the covariant components Ai associated with the vector A given in part (c).<br />

(e) Calculate the metric tensor gij and conjugate metric tensor gij .<br />

(f) From the results (e), verify that gijg jk = δ k i<br />

(g) Use the results from (c)(d) and (e) to verify that Ai = gikA k<br />

(h) Use the results from (c)(d) and (e) to verify that A i = g ik Ak<br />

(i) Find the projection of the vector A on unit vectors in the directions E1, E2, E3.<br />

(j) Find the projection of the vector A on unit vectors the directions E1 , E2 , E3 .<br />

59


60<br />

◮ <strong>12</strong>. For r = y i ei where y i = y i (x 1 ,x 2 ,x 3 ), i =1, 2, 3 we have by definition<br />

and consequently<br />

Ej = ∂r ∂yi<br />

=<br />

∂xj gij = Ei · Ej = ∂ym<br />

∂x i<br />

∂x j ei. From this relation show that E m = ∂x m<br />

∂y m<br />

∂x j , and gij = E i · E j = ∂xi<br />

∂y m<br />

◮ 13. Consider the set of all coordinate transformations of the form<br />

y i = a i j xj + b i<br />

where a i j and bi are constants and the determinant of a i j<br />

tions forms a group.<br />

ej<br />

∂yj ∂xj , i,j,m=1,...,3<br />

∂ym is different from zero. Show this set of transforma-<br />

◮ 14. For αi , βi constants and t a parameter, xi = αi + tβi,i =1, 2, 3 is the parametric representation of<br />

a straight line. Find the parametric equation of the line which passes through the two points (1, 2, 3) and<br />

(14, 7, −3). What does the vector dr<br />

dt represent?<br />

◮ 15. A surface can be represented using two parameters u, v by introducing the parametric equations<br />

x i = x i (u, v), i =1, 2, 3, a < u < b and c


◮ 17. The equation of a plane is defined in terms of two parameters u and v and has the form<br />

x i = αi u + βi v + γi<br />

i =1, 2, 3,<br />

where αi βi and γi are constants. Find the equation of the plane which passes through the points (1, 2, 3),<br />

(14, 7, −3) and (5, 5, 5). What does this problem have to do with the position vector r(u, v), the vectors<br />

∂r ∂r<br />

∂u , ∂v<br />

and r(0, 0)? Hint: See problem 15.<br />

◮ 18. Determine the points of intersection of the curve x 1 = t, x 2 =(t) 2 , x 3 =(t) 3 with the plane<br />

8 x 1 − 5 x 2 + x 3 − 4=0.<br />

◮ 19. Verify the relations Veijk E k = Ei × Ej and v −1 e ijk Ek = E i × E j where v = E1 · ( E2 × E3 )and<br />

V = E1 · ( E2 × E3)..<br />

◮ 20. Let ¯x i and x i , i =1, 2, 3 be related by the linear transformation ¯x i = c i j xj ,wherec i j<br />

such that the determinant c = det(c i j ) is different from zero. Let γn m denote the cofactor of cm n<br />

the determinant c.<br />

(a) Show that c i j γj<br />

k = γi j cj<br />

k = δi k .<br />

(b) Show the inverse transformation can be expressed x i = γ i j ¯x j .<br />

(c) Show that if A i is a contravariant vector, then its transformed components are Āp = c p qA q .<br />

(d) Show that if Ai is a covariant vector, then its transformed components are Āi = γ p<br />

i Ap.<br />

are constants<br />

divided by<br />

◮ 21. Show that the outer product of two contravariant vectors Ai and Bi , i =1, 2, 3 results in a second<br />

order contravariant tensor.<br />

◮ 22. Show that for the position vector r = y i (x 1 ,x 2 ,x 3 ) ei the element of arc length squared is<br />

ds 2 = dr · dr = gijdx i dx j where gij = Ei · Ej = ∂ym<br />

∂x i<br />

∂ym .<br />

∂xj ◮ 23. For Ai jk ,Bm n and C p<br />

tq absolute tensors, show that if Ai jkBk n = Ci jn then AijkB<br />

k<br />

n = C i<br />

jn.<br />

◮ 24. Let Aij denote an absolute covariant tensor of order 2. Show that the determinant A = det(Aij) is<br />

an invariant of weight 2 and (A) is an invariant of weight 1.<br />

◮ 25. Let Bij denote an absolute contravariant tensor of order 2. Show that the determinant B = det(Bij )<br />

is an invariant of weight −2 and √ B is an invariant of weight −1.<br />

◮ 26.<br />

(a) Write out the contravariant components of the following vectors<br />

(i) E1 (ii) E2 (iii) E3 where Ei = ∂r<br />

∂x i for i =1, 2, 3.<br />

(b) Write out the covariant components of the following vectors<br />

(i) E 1<br />

(ii) E 2<br />

(ii) E 3 where E i =gradx i , for i =1, 2, 3.<br />

61


62<br />

◮ 27. Let Aij and A ij denote absolute second order tensors. Show that λ = AijA ij is a scalar invariant.<br />

◮ 28. Assume that aij, i, j =1, 2, 3, 4 is a skew-symmetric second order absolute tensor. (a) Show that<br />

bijk = ∂ajk<br />

∂x<br />

∂x<br />

∂x k<br />

∂aki ∂aij<br />

+ + i j<br />

is a third order tensor. (b) Show bijk is skew-symmetric in all pairs of indices and (c) determine the number<br />

of independent components this tensor has.<br />

◮ 29. Show the linear forms A1x + B1y + C1 and A2x + B2y + C2, with respect to the group of rotations<br />

and translations x = x cos θ − y sin θ + h and y = x sin θ + y cos θ + k, have the forms A1x + B1y + C1 and<br />

A2x + B2y + C2. Also show that the quantities A1B2 − A2B1 and A1A2 + B1B2 are invariants.<br />

◮ 30. Show that the curvature of a curve y = f(x) isκ = ± y ′′ (1 + y ′2 ) −3/2 and that this curvature remains<br />

invariant under the group of rotations given in the problem 1. Hint: Calculate dy dy dx<br />

dx = dx<br />

◮ 31. Show that when the equation of a curve is given in the parametric form x = x(t), y= y(t), then<br />

˙x¨y − ˙y¨x<br />

the curvature is κ = ±<br />

(˙x 2 +˙y2 and remains invariant under the change of parameter t = t(t), where<br />

) 3/2<br />

˙x = dx<br />

dt , etc.<br />

◮ 32. Let A ij<br />

k<br />

denote a third order mixed tensor. (a) Show that the contraction Aij<br />

i is a first order<br />

which is not a tensor. This shows<br />

that in general, the process of contraction does not always apply to indices at the same level.<br />

contravariant tensor. (b) Show that contraction of i and j produces A ii<br />

k<br />

◮ 33. Let φ = φ(x 1 ,x 2 ,...,x N ) denote an absolute scalar invariant. (a) Is the quantity ∂φ<br />

∂x i atensor?(b)<br />

Is the quantity ∂2 φ<br />

∂x i ∂x j atensor?<br />

◮ 34. Consider the second order absolute tensor aij, i,j=1, 2wherea11 =1,a<strong>12</strong> =2,a21 =3anda22 =4.<br />

Find the components of aij under the transformation of coordinates x1 = x1 + x2 and x2 = x1 − x2 .<br />

◮ 35. Let Ai, Bi denote the components of two covariant absolute tensors of order one. Show that<br />

Cij = AiBj is an absolute second order covariant tensor.<br />

◮ 36. Let A i denote the components of an absolute contravariant tensor of order one and let Bi denote the<br />

components of an absolute covariant tensor of order one, show that C i j = Ai Bj transforms as an absolute<br />

mixed tensor of order two.<br />

◮ 37. (a) Show the sum and difference of two tensors of the same kind is also a tensor of this kind. (b) Show<br />

that the outer product of two tensors is a tensor. Do parts (a) (b) in the special case where one tensor Ai is a relative tensor of weight 4 and the other tensor B j<br />

k is a relative tensor of weight 3. What is the weight<br />

of the outer product tensor T ij<br />

k = AiB j<br />

k in this special case?<br />

◮ 38. Let A ij<br />

km denote the components of a mixed tensor of weight M. Form the contraction Bj m = Aij im<br />

and determine how Bj m transforms. What is its weight?<br />

◮ 39. Let A i j<br />

contraction S = A i i<br />

denote the components of an absolute mixed tensor of order two. Show that the scalar<br />

is an invariant.<br />

dx .


◮ 40. Let A i = A i (x 1 ,x 2 ,...,x N ) denote the components of an absolute contravariant tensor. Form the<br />

quantity B i j<br />

∂Ai = ∂xj and determine if Bi j transforms like a tensor.<br />

◮ 41. Let Ai denote the components of a covariant vector. (a) Show that aij = ∂Ai<br />

components of a second order tensor. (b) Show that<br />

∂xj ∂Aj<br />

−<br />

∂xi are the<br />

∂aij ∂ajk ∂aki<br />

+ + =0.<br />

∂xk ∂xi ∂xj ◮ 42. Show that xi = KeijkAjBk, withK= 0 and arbitrary, is a general solution of the system of equations<br />

Aix i =0,Bix i =0,i=1, 2, 3. Give a geometric interpretation of this result in terms of vectors.<br />

◮ 43. Given the vector A = y e1 + z e2 + x e3 where e1, e2, e3 denote a set of unit basis vectors which<br />

define a set of orthogonal x, y, z axes. Let E1 =3e1 +4e2, E2 =4e1 +7e2 and E3 = e3 denote a set of<br />

basis vectors which define a set of u, v, w axes. (a) Find the coordinate transformation between these two<br />

sets of axes. (b) Find a set of reciprocal vectors E1 , E3 , E3 . (c) Calculate the covariant components of A.<br />

(d) Calculate the contravariant components of A.<br />

◮ 44. Let A = Aij ei ej denote a dyadic. Show that<br />

A : Ac = A11A11 + A<strong>12</strong>A21 + A13A31 + A21A<strong>12</strong> + A22A22 + A23A32 + A31A13 + A32A23 + A23A33<br />

◮ 45. Let A = Ai ei, B = Bi ei, C = Ci ei, D = Di ei denote vectors and let φ = A B, ψ = C D denote<br />

dyadics which are the outer products involving the above vectors. Show that the double dot product satisfies<br />

φ : ψ = A B : C D =( A · C)( B · D)<br />

◮ 46. Show that if aij is a symmetric tensor in one coordinate system, then it is symmetric in all coordinate<br />

systems.<br />

◮ 47. Write the transformation laws for the given tensors. (a) A k ij (b) A ij<br />

k (c) A ijk<br />

m<br />

∂x<br />

◮ 48. Show that if Ai = Aj<br />

j<br />

andunbarredsystems.<br />

∂x i ,thenAi = Aj ∂xj<br />

◮ 49.<br />

(a) Show that under the linear homogeneous transformation<br />

the quadratic form<br />

∂x i . Note that this is equivalent to interchanging the bar<br />

x1 =a 1 1 x1 + a 2 1 x2<br />

x2 =a 1 2x1 + a 2 2x2<br />

Q(x1,x2) =g11(x1) 2 +2g<strong>12</strong>x1x2 + g22(x2) 2 becomes Q(x1, x2) =g 11 (x1) 2 +2g <strong>12</strong> x1x2 + g 22 (x2) 2<br />

where gij = g11a j<br />

1ai1 + g<strong>12</strong>(a i 1aj2 + aj1<br />

ai2 )+g22a i 2aj2 .<br />

(b) Show F = g11g22 − (g<strong>12</strong>) 2 is a relative invariant of weight 2 of the quadratic form Q(x1,x2) with respect<br />

to the group of linear homogeneous transformations. i.e. Show that F =∆ 2 F where F = g 11g 22 −(g <strong>12</strong>) 2<br />

and ∆ = (a 1 1a 2 2 − a 2 1a 1 2).<br />

63


64<br />

◮ 50. Let ai and bi for i =1,...,n denote arbitrary vectors and form the dyadic<br />

By definition the first scalar invariant of Φ is<br />

Φ=a1b1 + a2b2 + ···+ anbn.<br />

φ1 = a1 · b1 + a2 · b2 + ···+ an · bn<br />

where a dot product operator has been placed between the vectors. The first vector invariant of Φ is defined<br />

φ = a1 × b1 + a2 × b2 + ···+ an × bn<br />

where a vector cross product operator has been placed between the vectors.<br />

(a) Show that the first scalar and vector invariant of<br />

are respectively 1 and e1 + e3.<br />

Φ= e1 e2 + e2 e3 + e3 e3<br />

(b) From the vector f = f1 e1 + f2 e2 + f3 e3 one can form the dyadic ∇f having the matrix components<br />

⎛ ∂f1 ∂f2 ∂f3 ⎞<br />

∂x ∂x ∂x<br />

∂f1 ∂f2 ∂f3<br />

∇f = ⎝<br />

⎠<br />

∂y ∂y ∂y .<br />

∂f1<br />

∂z<br />

∂f2<br />

∂z<br />

∂f3<br />

∂z<br />

Show that this dyadic has the first scalar and vector invariants given by<br />

∇·f = ∂f1 ∂f2 ∂f3<br />

+ +<br />

∂x ∂y ∂z<br />

<br />

∂f3 ∂f2 ∂f1<br />

∇×f = − e1 +<br />

∂y ∂z ∂z<br />

<br />

∂f3 ∂f2 ∂f1<br />

− e2 + − e3<br />

∂x ∂x ∂y<br />

◮ 51. Let Φ denote the dyadic given in problem 50. The dyadic Φ2 defined by<br />

Φ2 = 1 <br />

ai × ajbi × bj<br />

2<br />

i,j<br />

is called the Gibbs second dyadic of Φ, where the summation is taken over all permutations of i and j. When<br />

i = j the dyad vanishes. Note that the permutations i, j and j, i give the same dyad and so occurs twice<br />

in the final sum. The factor 1/2 removes this doubling. Associated with the Gibbs dyad Φ2 are the scalar<br />

invariants<br />

φ2 = 1 <br />

(ai × aj) · (bi × bj)<br />

2<br />

Show that the dyad<br />

has<br />

φ3 = 1<br />

6<br />

i,j<br />

<br />

(ai × aj · ak)(bi × bj · bk)<br />

i,j,k<br />

Φ=as+ tq+ cu<br />

the first scalar invariant φ1 = a · s + b · t + c · u<br />

the first vector invariant φ = a × s + b × t + c × u<br />

Gibbs second dyad Φ2 = b × ct × u + c × au × s + a × bs × t<br />

second scalar of Φ φ2 =(b × c) · (t · u)+(c × a) · (u × s)+(a × b) · (s × t)<br />

third scalar of Φ φ3 =(a × b · c)(s × t · u)


◮ 52. (Spherical Trigonometry) Construct a spherical triangle ABC on the surface of a unit sphere with<br />

sides and angles less than 180 degrees. Denote by a,b cthe unit vectors from the origin of the sphere to the<br />

vertices A,B and C. Make the construction such that a·(b×c) is positive with a, b, c forming a right-handed<br />

system. Let α, β, γ denote the angles between these unit vectors such that<br />

a · b =cosγ c · a =cosβ b · c =cosα. (1)<br />

The great circles through the vertices A,B,C then make up the sides of the spherical triangle where side α<br />

is opposite vertex A, side β is opposite vertex B and side γ is opposite the vertex C. The angles A,B and C<br />

between the various planes formed by the vectors a, b and c are called the interior dihedral angles of the<br />

spherical triangle. Note that the cross products<br />

a × b =sinγ c b× c =sinα a c× a =sinβ b (2)<br />

define unit vectors a, b and c perpendicular to the planes determined by the unit vectors a, b and c. The<br />

dot products<br />

a · b =cosγ b · c =cosα c · a =cosβ (3)<br />

define the angles α,β and γ which are called the exterior dihedral angles at the vertices A,B and C and are<br />

such that<br />

α = π − A β = π − B γ = π − C. (4)<br />

(a) Using appropriate scaling, show that the vectors a, b, c and a, b, c form a reciprocal set.<br />

(b) Show that a · (b × c) =sinαa · a =sinβb · b =sinγc · c<br />

(c) Show that a · (b × c) =sinαa · a =sinβb · b =sinγc · c<br />

(d) Using parts (b) and (c) show that<br />

sin α sin β sin γ<br />

= =<br />

sin α sin β sin γ<br />

(e) Use the results from equation (4) to derive the law of sines for spherical triangles<br />

sin α sin β sin γ<br />

= =<br />

sin A sin B sin C<br />

(f) Using the equations (2) show that<br />

and hence show that<br />

In a similar manner show also that<br />

sin β sin γb · c =(c × a) · (a × b) =(c · a)(a · b) − b · c<br />

cos α =cosβ cos γ − sin β sin γ cos α.<br />

cos α =cosβ cos γ − sin β sin γ cos α.<br />

(g) Using part (f) derive the law of cosines for spherical triangles<br />

cos α =cosβ cos γ +sinβ sin γ cos A<br />

cos A = − cos B cos C +sinB sin C cos α<br />

A cyclic permutation of the symbols produces similar results involving the other angles and sides of the<br />

spherical triangle.<br />

65


§1.3 SPECIAL TENSORS<br />

Knowing how tensors are defined and recognizing a tensor when it pops up in front of you are two<br />

different things. Some quantities, which are tensors, frequently arise in applied problems and you should<br />

learn to recognize these special tensors when they occur. In this section some important tensor quantities<br />

are defined. We also consider how these special tensors can in turn be used to define other tensors.<br />

Metric Tensor<br />

Define yi ,i=1,...,N as independent coordinates in an N dimensional orthogonal Cartesian coordinate<br />

system. The distance squared between two points y i<br />

and y i + dy i expression<br />

, i =1,...,N is defined by the<br />

ds 2 = dy m dy m =(dy 1 ) 2 +(dy 2 ) 2 + ···+(dy N ) 2 . (1.3.1)<br />

Assume that the coordinates y i are related to a set of independent generalized coordinates x i ,i=1,...,N<br />

by a set of transformation equations<br />

y i = y i (x 1 ,x 2 ,...,x N ), i =1,...,N. (1.3.2)<br />

To emphasize that each y i depends upon the x coordinates we sometimes use the notation y i = y i (x), for<br />

i =1,...,N. The differential of each coordinate can be written as<br />

dy m = ∂ym<br />

∂x j dxj , m =1,...,N, (1.3.3)<br />

and consequently in the x-generalized coordinates the distance squared, found from the equation (1.3.1),<br />

becomes a quadratic form. Substituting equation (1.3.3) into equation (1.3.1) we find<br />

where<br />

ds 2 = ∂ym<br />

∂x i<br />

gij = ∂ym<br />

∂x i<br />

∂y m<br />

∂x j dxi dx j = gij dx i dx j<br />

(1.3.4)<br />

∂ym , i,j =1,...,N (1.3.5)<br />

∂xj are called the metrices of the space defined by the coordinates xi ,i=1,...,N. Here the gij are functions of<br />

the x coordinates and is sometimes written as gij = gij(x). Further, the metrices gij are symmetric in the<br />

indices i and j so that gij = gji for all values of i and j over the range of the indices. If we transform to<br />

another coordinate system, say xi ,i=1,...,N, then the element of arc length squared is expressed in terms<br />

of the barred coordinates and ds2 = gij dxidxj , where gij = gij(x) is a function of the barred coordinates.<br />

The following example demonstrates that these metrices are second order covariant tensors.<br />

65


66<br />

EXAMPLE 1.3-1. Show the metric components gij are covariant tensors of the second order.<br />

Solution: In a coordinate system xi ,i=1,...,N the element of arc length squared is<br />

ds 2 = gijdx i dx j<br />

while in a coordinate system x i ,i=1,...,N the element of arc length squared is represented in the form<br />

(1.3.6)<br />

ds 2 = g mn dx m dx n . (1.3.7)<br />

The element of arc length squared is to be an invariant and so we require that<br />

g mndx m dx n = gijdx i dx j<br />

(1.3.8)<br />

Here it is assumed that there exists a coordinate transformation of the form defined by equation (1.2.30)<br />

together with an inverse transformation, as in equation (1.2.32), which relates the barred and unbarred<br />

coordinates. In general, if x i = x i (x), then for i =1,...,N we have<br />

dx i = ∂xi<br />

∂xm dxm and dx j = ∂xj<br />

∂x<br />

Substituting these differentials in equation (1.3.8) gives us the result<br />

gmndx m dx n ∂x<br />

= gij<br />

i<br />

∂xm ∂x j<br />

∂x n dxm dx n<br />

For arbitrary changes in dx m this equation implies that g mn = gij<br />

as a second order absolute covariant tensor.<br />

or<br />

n dxn<br />

<br />

∂x<br />

gmn − gij<br />

i<br />

∂xm ∂xj ∂xn <br />

dx m dx n =0<br />

∂xi ∂xm (1.3.9)<br />

∂x j<br />

∂x n and consequently gij transforms<br />

EXAMPLE 1.3-2. (Curvilinear coordinates) Consider a set of general transformation equations from<br />

rectangular coordinates (x, y, z) to curvilinear coordinates (u, v, w). These transformation equations and the<br />

corresponding inverse transformations are represented<br />

x = x(u, v, w)<br />

y = y(u, v, w)<br />

z = z(u, v, w).<br />

u = u(x, y, z)<br />

v = v(x, y, z)<br />

w = w(x, y, z)<br />

(1.3.10)<br />

Here y 1 = x, y 2 = y, y 3 = z and x 1 = u, x 2 = v, x 3 = w are the Cartesian and generalized coordinates<br />

and N =3. The intersection of the coordinate surfaces u = c1,v = c2 and w = c3 define coordinate curves<br />

of the curvilinear coordinate system. The substitution of the given transformation equations (1.3.10) into<br />

the position vector r = x e1 + y e2 + z e3 produces the position vector which is a function of the generalized<br />

coordinates and<br />

r = r(u, v, w) =x(u, v, w) e1 + y(u, v, w) e2 + z(u, v, w) e3


and consequently dr = ∂r ∂r ∂r<br />

du + dv + dw, where<br />

∂u ∂v ∂w<br />

E1 = ∂r ∂x<br />

=<br />

∂u ∂u e1 + ∂y<br />

∂u e2 + ∂z<br />

∂u e3<br />

E2 = ∂r ∂x<br />

=<br />

∂v ∂v e1 + ∂y<br />

∂v e2 + ∂z<br />

∂v e3<br />

E3 = ∂r ∂x<br />

=<br />

∂w ∂w e1 + ∂y<br />

∂w e2 + ∂z<br />

∂w e3.<br />

are tangent vectors to the coordinate curves. The element of arc length in the curvilinear coordinates is<br />

ds 2 = dr · dr = ∂r ∂r ∂r ∂r ∂r ∂r<br />

· dudu + · dudv + ·<br />

∂u ∂u ∂u ∂v ∂u ∂w dudw<br />

+ ∂r ∂r ∂r ∂r ∂r ∂r<br />

· dvdu + · dvdv + ·<br />

∂v ∂u ∂v ∂v ∂v ∂w dvdw<br />

+ ∂r ∂r ∂r ∂r ∂r ∂r<br />

· dwdu + · dwdv + ·<br />

∂w ∂u ∂w ∂v ∂w ∂w dwdw.<br />

(1.3.11)<br />

(1.3.<strong>12</strong>)<br />

Utilizing the summation convention, the above can be expressed in the index notation.<br />

quantities<br />

Define the<br />

g11 = ∂r ∂r<br />

·<br />

∂u ∂u<br />

g21 = ∂r ∂r<br />

·<br />

∂v ∂u<br />

g31 = ∂r<br />

g<strong>12</strong> =<br />

∂r<br />

·<br />

∂w ∂u<br />

∂r ∂r<br />

·<br />

∂u ∂v<br />

g22 = ∂r ∂r<br />

·<br />

∂v ∂v<br />

g32 = ∂r<br />

g13 =<br />

∂r<br />

·<br />

∂w ∂v<br />

∂r ∂r<br />

·<br />

∂u ∂w<br />

g23 = ∂r ∂r<br />

·<br />

∂v ∂w<br />

g33 = ∂r ∂r<br />

·<br />

∂w ∂w<br />

and let x 1 = u, x 2 = v, x 3 = w. Then the above element of arc length can be expressed as<br />

where<br />

ds 2 = Ei · Ej dx i dx j = gijdx i dx j , i,j =1, 2, 3<br />

gij = Ei · Ej = ∂r ∂r<br />

·<br />

∂xi ∂x<br />

∂ym , i,j free indices (1.3.13)<br />

∂xj ∂ym<br />

= j ∂xi are called the metric components of the curvilinear coordinate system. The metric components may be<br />

thought of as the elements of a symmetric matrix, since gij = gji. In the rectangular coordinate system<br />

x, y, z, the element of arc length squared is ds2 = dx2 + dy2 + dz2 . In this space the metric components are<br />

⎛<br />

1<br />

gij = ⎝ 0<br />

0<br />

1<br />

⎞<br />

0<br />

0⎠<br />

.<br />

0 0 1<br />

67


68<br />

EXAMPLE 1.3-3. (Cylindrical coordinates (r, θ, z))<br />

The transformation equations from rectangular coordinates to cylindrical coordinates can be expressed<br />

as x = r cos θ, y = r sin θ, z = z. Here y 1 = x, y 2 = y, y 3 = z and x 1 = r, x 2 = θ, x 3 = z, and the<br />

position vector can be expressed r = r(r, θ, z) =r cos θ e1 + r sin θ e2 + z e3. The derivatives of this position<br />

vector are calculated and we find<br />

E1 = ∂r<br />

∂r =cosθe1 +sinθe2, E2 = ∂r<br />

∂θ = −r sin θ e1 + r cos θ e2, E3 = ∂r<br />

= e3.<br />

∂z<br />

From the results in equation (1.3.13), the metric components of this space are<br />

⎛<br />

1<br />

gij = ⎝ 0<br />

0<br />

r<br />

0<br />

2 ⎞<br />

0 ⎠ .<br />

0 0 1<br />

Wenotethatsincegij =0wheni = j, the coordinate system is orthogonal.<br />

Given a set of transformations of the form found in equation (1.3.10), one can readily determine the<br />

metric components associated with the generalized coordinates. For future reference we list several different<br />

coordinate systems together with their metric components. Each of the listed coordinate systems are<br />

orthogonal and so gij =0fori= j. The metric components of these orthogonal systems have the form<br />

and the element of arc length squared is<br />

1. Cartesian coordinates (x, y, z)<br />

⎛<br />

gij = ⎝ h21 0 0<br />

0 h2 2 0<br />

0 0 h2 ⎞<br />

⎠<br />

3<br />

ds 2 = h 2 1(dx 1 ) 2 + h 2 2(dx 2 ) 2 + h 2 3(dx 3 ) 2 .<br />

x = x<br />

y = y<br />

z = z<br />

h1 =1<br />

h2 =1<br />

h3 =1<br />

The coordinate curves are formed by the intersection of the coordinate surfaces<br />

x =Constant, y =Constant and z =Constant.


2. Cylindrical coordinates (r, θ, z)<br />

Figure 1.3-1. Cylindrical coordinates.<br />

x = r cos θ<br />

y = r sin θ<br />

z = z<br />

r ≥ 0<br />

0 ≤ θ ≤ 2π<br />

−∞


70<br />

Figure 1.3-2. Spherical coordinates.<br />

The coordinate curves, illustrated in the figure 1.3-3, are formed by the intersection of the coordinate<br />

surfaces<br />

5. Parabolic coordinates (ξ,η,φ)<br />

x 2 = −2ξ 2 (y − ξ2<br />

) Parabolic cylinders<br />

2<br />

x 2 =2η 2 (y + η2<br />

)<br />

2<br />

Parabolic cylinders<br />

z = Constant Planes.<br />

Figure 1.3-3. Parabolic cylindrical coordinates in plane z =0.<br />

x = ξη cos φ<br />

y = ξη sin φ<br />

z = 1<br />

2 (ξ2 − η 2 )<br />

ξ ≥ 0<br />

η ≥ 0<br />

0


The coordinate curves, illustrated in the figure 1.3-4, are formed by the intersection of the coordinate<br />

surfaces<br />

6. Elliptic cylindrical coordinates (ξ,η,z)<br />

x 2 + y 2 = −2ξ 2 (z − ξ2<br />

) Paraboloids<br />

2<br />

x 2 + y 2 =2η 2 (z + η2<br />

)<br />

2<br />

Paraboloids<br />

y = x tan φ Planes.<br />

Figure 1.3-4. Parabolic coordinates, φ = π/4.<br />

x =coshξ cos η<br />

y =sinhξ sin η<br />

z = z<br />

ξ ≥ 0<br />

0 ≤ η ≤ 2π<br />

−∞


72<br />

7. Elliptic coordinates (ξ,η,φ)<br />

Figure 1.3-5. Elliptic cylindrical coordinates in the plane z =0.<br />

x = (1 − η 2 )(ξ 2 − 1) cos φ<br />

y = (1 − η 2 )(ξ 2 − 1) sin φ<br />

z = ξη<br />

1 ≤ ξ


Figure 1.3-6. Elliptic coordinates φ = π/4.<br />

Figure 1.3-7. Bipolar coordinates.<br />

The coordinate curves, illustrated in the figure 1.3-7, are formed by the intersection of the coordinate<br />

surfaces<br />

(x − a coth v) 2 + y 2 = a2<br />

sinh 2 v<br />

x 2 +(y − a cot u) 2 = a2<br />

Cylinders<br />

sin 2 u<br />

Cylinders<br />

z = Constant Planes.<br />

73


74<br />

9. Conical coordinates (u, v, w)<br />

x = uvw<br />

y = u<br />

<br />

(v2 − a2 )(w2 − a2 )<br />

a a2 − b2 z = u<br />

<br />

(v2 − b2 )(w2 − b2 )<br />

b b2 − a2 ab , b2 >v 2 >a 2 >w 2 , u ≥ 0<br />

h 2 1 =1<br />

h 2 2 =<br />

h 2 3 =<br />

u 2 (v 2 − w 2 )<br />

(v 2 − a 2 )(b 2 − v 2 )<br />

u 2 (v 2 − w 2 )<br />

(w 2 − a 2 )(w 2 − b 2 )<br />

The coordinate curves, illustrated in the figure 1.3-8, are formed by the intersection of the coordinate<br />

surfaces<br />

x 2 + y 2 + z 2 = u 2<br />

Spheres<br />

x2 y2<br />

+<br />

v2 v2 z2<br />

+<br />

− a2 v2 =0,<br />

− b2 Cones<br />

x2 +<br />

w2 y2<br />

10. Prolate spheroidal coordinates (u, v, φ)<br />

w2 +<br />

− a2 z2<br />

w2 =0, Cones.<br />

− b2 Figure 1.3-8. Conical coordinates.<br />

x = a sinh u sin v cosφ , u ≥ 0<br />

y = a sinh u sin v sin φ, 0 ≤ v ≤ π<br />

z = a cosh u cos v, 0 ≤ φ


11. Oblate spheroidal coordinates (ξ,η,φ)<br />

Figure 1.3-9. Prolate spheroidal coordinates<br />

x = a cosh ξ cos η cosφ , ξ ≥ 0<br />

y = a cosh ξ cos η sin φ, − π π<br />

≤ η ≤<br />

2 2<br />

z = a sinh ξ sin η, 0 ≤ φ ≤ 2π<br />

h 2 1 = h2 2<br />

h 2 2 = a2 (sinh 2 ξ +sin 2 η)<br />

h 2 3 = a 2 cosh 2 ξ cos 2 η<br />

The coordinate curves, illustrated in the figure 1.3-10, are formed by the intersection of the coordinate<br />

surfaces<br />

x2 +<br />

(a cosh ξ) 2<br />

x2 +<br />

(a cos η) 2<br />

<strong>12</strong>. Toroidal coordinates (u, v, φ)<br />

y2 +<br />

(a cosh ξ) 2<br />

y2 −<br />

(a cos η) 2<br />

a sinh v cos φ<br />

x = ,<br />

cosh v − cos u<br />

0 ≤ u


76<br />

Figure 1.3-10. Oblate spheroidal coordinates<br />

Figure 1.3-11. Toroidal coordinates<br />

is a mixed second order tensor.<br />

EXAMPLE 1.3-4. Show the Kronecker delta δi j<br />

Solution: Assume we have a coordinate transformation xi = xi (x),i =1,...,N of the form (1.2.30) and<br />

possessing an inverse transformation of the form (1.2.32). Let δ i<br />

j and δi j denote the Kronecker delta in the<br />

barred and unbarred system of coordinates. By definition the Kronecker delta is defined<br />

δ i<br />

j = δi j =<br />

<br />

0,<br />

1,<br />

if<br />

if<br />

i = j<br />

i = j .


Employing the chain rule we write<br />

∂xm ∂xm<br />

n =<br />

∂x ∂xi ∂xi ∂xm<br />

n =<br />

∂x ∂xi ∂xk ∂xn δi k<br />

(1.3.14)<br />

By hypothesis, the xi ,i=1,...,N are independent coordinates and therefore we have ∂xm<br />

n ∂x = δmn<br />

and (1.3.14)<br />

simplifies to<br />

δ m<br />

n = δi ∂x<br />

k<br />

m<br />

∂xi ∂xk n .<br />

∂x<br />

Therefore, the Kronecker delta transforms as a mixed second order tensor.<br />

Conjugate Metric Tensor<br />

Let g denote the determinant of the matrix having the metric tensor gij,i,j =1,...,N as its elements.<br />

In our study of cofactor elements of a matrix we have shown that<br />

cof(g1j)g1k + cof(g2j)g2k + ...+ cof(gNj)gNk = gδ j<br />

k . (1.3.15)<br />

We can use this fact to find the elements in the inverse matrix associated with the matrix having the<br />

components gij. Theelementsofthisinversematrixare<br />

g ij = 1<br />

g cof(gij) (1.3.16)<br />

and are called the conjugate metric components. We examine the summation g ij gik and find:<br />

The equation<br />

g ij gik = g 1j g1k + g 2j g2k + ...+ g Nj gNk<br />

= 1<br />

g [cof(g1j)g1k + cof(g2j)g2k + ...+ cof(gNj)gNk]<br />

= 1<br />

<br />

gδ<br />

g<br />

j<br />

<br />

k = δ j<br />

k<br />

g ij gik = δ j<br />

k<br />

(1.3.17)<br />

is an example where we can use the quotient law to show gij is a second order contravariant tensor. Because<br />

of the symmetry of gij and gij the equation (1.3.17) can be represented in other forms.<br />

EXAMPLE 1.3-5. Let Ai and Ai denote respectively the covariant and contravariant components of a<br />

vector A. Show these components are related by the equations<br />

Ai = gijA j<br />

A k = g jk Aj<br />

where gij and g ij are the metric and conjugate metric components of the space.<br />

(1.3.18)<br />

(1.3.19)<br />

77


78<br />

Solution: We multiply the equation (1.3.18) by g im (inner product) and use equation (1.3.17) to simplify<br />

the results. This produces the equation g im Ai = g im gijA j = δ m j A j = A m . Changing indices produces the<br />

result given in equation (1.3.19). Conversely, if we start with equation (1.3.19) and multiply by gkm (inner<br />

product) we obtain gkmA k = gkmg jk Aj = δ j mAj = Am which is another form of the equation (1.3.18) with<br />

the indices changed.<br />

Notice the consequences of what the equations (1.3.18) and (1.3.19) imply when we are in an orthogonal<br />

Cartesian coordinate system where<br />

In this special case, we have<br />

⎛<br />

1<br />

gij = ⎝ 0<br />

0<br />

1<br />

⎞<br />

0<br />

0⎠<br />

and g<br />

0 0 1<br />

ij ⎛<br />

1<br />

= ⎝ 0<br />

0<br />

1<br />

⎞<br />

0<br />

0⎠<br />

.<br />

0 0 1<br />

A1 = g11A 1 + g<strong>12</strong>A 2 + g13A 3 = A 1<br />

A2 = g21A 1 + g22A 2 + g23A 3 = A 2<br />

A3 = g31A 1 + g32A 2 + g33A 3 = A 3 .<br />

These equations tell us that in a Cartesian coordinate system the contravariant and covariant components<br />

are identically the same.<br />

EXAMPLE 1.3-6. We have previously shown that if Ai is a covariant tensor of rank 1 its components in<br />

a barred system of coordinates are<br />

Ai = Aj<br />

Solve for the Aj in terms of the Aj. (i.e. find the inverse transformation).<br />

Solution: Multiply equation (1.3.20) by ∂xi<br />

∂xm (inner product) and obtain<br />

Ai<br />

∂x i<br />

= Aj<br />

∂xm In the above product we have ∂xj<br />

∂xi ∂xi ∂xj<br />

=<br />

∂xm coordinates. This reduces equation (1.3.21) to the form<br />

∂xj i . (1.3.20)<br />

∂x<br />

∂x j<br />

∂x i<br />

∂xi . (1.3.21)<br />

∂xm ∂x m = δj m since xj and x m are assumed to be independent<br />

∂x i<br />

Ai<br />

∂xm = Ajδ j m = Am<br />

(1.3.22)<br />

which is the desired inverse transformation.<br />

This result can be obtained in another way. Examine the transformation equation (1.3.20) and ask the<br />

question, “When we have two coordinate systems, say a barred and an unbarred system, does it matter which<br />

system we call the barred system?” With some thought it should be obvious that it doesn’t matter which<br />

system you label as the barred system. Therefore, we can interchange the barred and unbarred symbols in<br />

equation (1.3.20) and obtain the result Ai = Aj<br />

a different set of indices.<br />

∂xj which is the same form as equation (1.3.22), but with<br />

∂xi


Associated Tensors<br />

Associated tensors can be constructed by taking the inner product of known tensors with either the<br />

metric or conjugate metric tensor.<br />

Definition: (Associated tensor) Any tensor constructed by multiplying (inner<br />

product) a given tensor with the metric or conjugate metric tensor is called an<br />

associated tensor.<br />

Associated tensors are different ways of representing a tensor. The multiplication of a tensor by the<br />

metric or conjugate metric tensor has the effect of lowering or raising indices. For example the covariant<br />

and contravariant components of a vector are different representations of the same vector in different forms.<br />

These forms are associated with one another by way of the metric and conjugate metric tensor and<br />

g ij Ai = A j<br />

gijA j = Ai.<br />

EXAMPLE 1.3-7. The following are some examples of associated tensors.<br />

A j = g ij Ai<br />

A m .jk = g mi Aijk<br />

A .nm<br />

i.. = g mk g nj Aijk<br />

Aj = gijA i<br />

A i.k<br />

m = gmjA ijk<br />

Amjk = gimA i .jk<br />

Sometimes ‘dots’are used as indices in order to represent the location of the index that was raised or lowered.<br />

If a tensor is symmetric, the position of the index is immaterial and so a dot is not needed. For example, if<br />

Amn is a symmetric tensor, then it is easy to show that An .m and A.n m<br />

as An m without confusion.<br />

are equal and therefore can be written<br />

Higher order tensors are similarly related. For example, if we find a fourth order covariant tensor Tijkm<br />

we can then construct the fourth order contravariant tensor T pqrs from the relation<br />

T pqrs = g pi g qj g rk g sm Tijkm.<br />

This fourth order tensor can also be expressed as a mixed tensor. Some mixed tensors associated with<br />

the given fourth order covariant tensor are:<br />

T p<br />

.jkm = gpi Tijkm, T pq<br />

..km = gqj T p<br />

.jkm .<br />

79


80<br />

Riemann Space VN<br />

A Riemannian space VN is said to exist if the element of arc length squared has the form<br />

ds 2 = gijdx i dx j<br />

(1.3.23)<br />

where the metrices gij = gij(x 1 ,x 2 ,...,x N ) are continuous functions of the coordinates and are different<br />

from constants. In the special case gij = δij the Riemannian space VN reduces to a Euclidean space EN .<br />

The element of arc length squared defined by equation (1.3.23) is called the Riemannian metric and any<br />

geometry which results by using this metric is called a Riemannian geometry. A space VN is called flat if<br />

it is possible to find a coordinate transformation where the element of arclength squared is ds 2 = ɛi(dx i ) 2<br />

where each ɛi is either +1 or −1. A space which is not flat is called curved.<br />

Geometry in VN<br />

Given two vectors A = A i Ei and B = B j Ej, then their dot product can be represented<br />

A · B = A i B j Ei · Ej = gijA i B j = AjB j = A i Bi = g ij AjBi = | A|| B| cos θ. (1.3.24)<br />

Consequently, in an N dimensional Riemannian space VN the dot or inner product of two vectors A and B<br />

is defined:<br />

gijA i B j = AjB j = A i Bi = g ij AjBi = AB cos θ. (1.3.25)<br />

In this definition A is the magnitude of the vector Ai , the quantity B is the magnitude of the vector Bi and<br />

θ is the angle between the vectors when their origins are made to coincide. In the special case that θ =90◦ we have gijAiB j = 0 as the condition that must be satisfied in order that the given vectors Ai and Bi are<br />

orthogonal to one another. Consider also the special case of equation (1.3.25) when Ai = Bi and θ =0. In<br />

this case the equations (1.3.25) inform us that<br />

g in AnAi = A i Ai = ginA i A n =(A) 2 . (1.3.26)<br />

From this equation one can determine the magnitude of the vector A i . The magnitudes A and B can be<br />

written A =(ginA i A n ) 1<br />

2 and B =(gpqB p B q ) 1<br />

2 and so we can express equation (1.3.24) in the form<br />

cos θ =<br />

gijA i B j<br />

(gmnA m A n ) 1<br />

2 (gpqB p B q ) 1<br />

2<br />

. (1.3.27)<br />

An import application of the above concepts arises in the dynamics of rigid body motion. Note that if a<br />

vector A i has constant magnitude and the magnitude of dAi<br />

dt is different from zero, then the vectors A i and<br />

dA i<br />

dt<br />

i dAj<br />

must be orthogonal to one another due to the fact that gijA<br />

dt =0. As an example, consider the unit<br />

vectors e1, e2 and e3 on a rotating system of Cartesian axes. We have for constants ci, i =1, 6that<br />

d e1<br />

dt = c1 e2 + c2 e3<br />

d e2<br />

dt = c3 e3 + c4 e1<br />

d e3<br />

dt = c5 e1 + c6 e2<br />

because the derivative of any ei (i fixed) constant vector must lie in a plane containing the vectors ej and<br />

ek, (j = i , k = i and j = k), since any vector in this plane must be perpendicular to ei.


The above definition of a dot product in VN can be used to define unit vectors in VN .<br />

EXAMPLE 1.3-8. (Unit vectors)<br />

Definition: (Unit vector) Whenever the magnitude of a vector<br />

Ai is unity, the vector is called a unit vector. In this case we<br />

have<br />

gijA i A j =1. (1.3.28)<br />

In VN the element of arc length squared is expressed ds 2 = gij dx i dx j which can be expressed in the<br />

dx<br />

form 1 = gij<br />

i dx<br />

ds<br />

j<br />

dxi<br />

. This equation states that the vector ,i=1,...,N is a unit vector. One application<br />

ds ds<br />

of this equation is to consider a particle moving along a curve in VN which is described by the parametric<br />

equations xi = xi (t), for i =1,...,N. The vector V i = dxi<br />

dt ,i=1,...,N represents a velocity vector of the<br />

particle. By chain rule differentiation we have<br />

where V = ds<br />

dt<br />

V i = dxi<br />

dt<br />

= dxi<br />

ds<br />

ds<br />

dt<br />

dxi<br />

= V , (1.3.29)<br />

ds<br />

dxi<br />

is the scalar speed of the particle and ds is a unit tangent vector to the curve. The equation<br />

(1.3.29) shows that the velocity is directed along the tangent to the curve and has a magnitude V. That is<br />

2 ds<br />

=(V )<br />

dt<br />

2 = gijV i V j .<br />

EXAMPLE 1.3-9. (Curvilinear coordinates)<br />

Find an expression for the cosine of the angles between the coordinate curves associated with the<br />

transformation equations<br />

x = x(u, v, w), y = y(u, v, w), z = z(u, v, w).<br />

81


82<br />

Figure 1.3-<strong>12</strong>. Angles between curvilinear coordinates.<br />

Solution: Let y 1 = x, y 2 = y, y 3 = z and x 1 = u, x 2 = v, x 3 = w denote the Cartesian and curvilinear<br />

coordinates respectively. With reference to the figure 1.3-<strong>12</strong> we can interpret the intersection of the surfaces<br />

v = c2 and w = c3 as the curve r = r(u, c2,c3) which is a function of the parameter u. By moving only along<br />

thiscurvewehavedr = ∂r<br />

du and consequently<br />

∂u<br />

or<br />

This equation shows that the vector dx1<br />

ds<br />

be represented by tr 1<br />

(1) = √ δ g11 r 1 .<br />

ds 2 = dr · dr = ∂r ∂r<br />

·<br />

∂u ∂u dudu = g11(dx 1 ) 2 ,<br />

1= dr dr<br />

· = g11<br />

ds ds<br />

dx 1<br />

ds<br />

2<br />

.<br />

= 1<br />

√ g11 is a unit vector along this curve. This tangent vector can<br />

The curve which is defined by the intersection of the surfaces u = c1 and w = c3 has the unit tangent<br />

= 1<br />

√ g22 δ r 2. Similarly, the curve which is defined as the intersection of the surfaces u = c1 and<br />

vector tr (2)<br />

v = c2 has the unit tangent vector tr (3)<br />

unit vectors tr (1) and tr (2) , is obtained from the result of equation (1.3.25). We find<br />

cos θ<strong>12</strong> = gpqt p<br />

(1) tq<br />

(2)<br />

For θ13 the angle between the directions t i (1) and ti (3)<br />

= 1<br />

√ g33 δ r 3. The cosine of the angle θ<strong>12</strong>, which is the angle between the<br />

1<br />

= gpq √ δ<br />

g11<br />

p 1<br />

1 √<br />

g22<br />

we find<br />

g13<br />

cos θ13 = √ √ .<br />

g11 g33<br />

Finally, for θ23 the angle between the directions t i (2) and ti (3)<br />

δ q<br />

2 =<br />

we find<br />

g<strong>12</strong><br />

√ √ .<br />

g11 g22<br />

cos θ23 = √ √ .<br />

g22 g33<br />

When θ13 = θ<strong>12</strong> = θ23 =90◦ , we have g<strong>12</strong> = g13 = g23 = 0 and the coordinate curves which make up the<br />

curvilinear coordinate system are orthogonal to one another.<br />

In an orthogonal coordinate system we adopt the notation<br />

g23<br />

g11 =(h1) 2 , g22 =(h2) 2 , g33 =(h3) 2<br />

and gij =0,i= j.


Epsilon Permutation Symbol<br />

Associated with the e−permutation symbols there are the epsilon permutation symbols defined by the<br />

relations<br />

ɛijk = √ geijk and ɛ ijk = 1<br />

√ g e ijk<br />

(1.3.30)<br />

where g is the determinant of the metrices gij.<br />

It can be demonstrated that the eijk permutation symbol is a relative tensor of weight −1 whereasthe<br />

ɛijk permutation symbol is an absolute tensor. Similarly, the eijk permutation symbol is a relative tensor of<br />

weight +1 and the corresponding ɛijk permutation symbol is an absolute tensor.<br />

EXAMPLE 1.3-10. (ɛ permutation symbol)<br />

Show that eijk is a relative tensor of weight −1 and the corresponding ɛijk permutation symbol is an<br />

absolute tensor.<br />

Solution: Examine the Jacobian<br />

and make the substitution<br />

J<br />

<br />

x<br />

<br />

=<br />

x<br />

From the definition of a determinant we may write<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

∂x 1<br />

∂x1 ∂x 2<br />

∂x1 ∂x 3<br />

∂x1 ∂x 1<br />

∂x2 ∂x 2<br />

∂x2 ∂x 3<br />

∂x2 ∂x 1<br />

∂x3 ∂x 2<br />

∂x3 ∂x 3<br />

∂x3 a i j = ∂xi<br />

j ,i,j=1, 2, 3.<br />

∂x<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

eijka i ma j na k p = J( x<br />

x )emnp. (1.3.31)<br />

By definition, emnp = emnp in all coordinate systems and hence equation (1.3.31) can be expressed in the<br />

form<br />

<br />

J( x<br />

x )<br />

−1 ∂x<br />

eijk<br />

i<br />

∂xm ∂xj ∂xn ∂xk p = emnp<br />

∂x<br />

(1.3.32)<br />

which demonstrates that eijk transforms as a relative tensor of weight −1.<br />

Wehavepreviouslyshownthemetrictensorgij<br />

according to the rule gij = gmn<br />

is a second order covariant tensor and transforms<br />

∂x m<br />

∂x i<br />

∂xn j . Taking the determinant of this result we find<br />

∂x<br />

<br />

<br />

g = |gij| = |gmn| <br />

<br />

∂x m<br />

∂x i<br />

<br />

<br />

<br />

<br />

2<br />

<br />

= g J( x<br />

x )<br />

2 (1.3.33)<br />

where g is the determinant of (gij) andg is the determinant of (g ij ). This result demonstrates that g is a<br />

scalar invariant of weight +2. Taking the square root of this result we find that<br />

√ x<br />

g = gJ( ). (1.3.34)<br />

x<br />

Consequently, we call √ g a scalar invariant of weight +1. Now multiply both sides of equation (1.3.32) by<br />

√<br />

g and use (1.3.34) to verify the relation<br />

√ ∂x<br />

geijk<br />

i<br />

∂xm ∂xj ∂xn ∂xk ∂xp = g emnp. (1.3.35)<br />

This equation demonstrates that the quantity ɛijk = √ geijk transforms like an absolute tensor.<br />

83


84<br />

Figure 1.3-14. Translation followed by rotation of axes<br />

In a similar manner one can show e ijk is a relative tensor of weight +1 and ɛ ijk = 1<br />

√ g e ijk is an absolute<br />

tensor. This is left as an exercise.<br />

Another exercise found at the end of this section is to show that a generalization of the e − δ identity<br />

is the epsilon identity<br />

g ij ɛiptɛjrs = gprgts − gpsgtr. (1.3.36)<br />

Cartesian Tensors<br />

Consider the motion of a rigid rod in two dimensions. No matter how complicated the movement of<br />

the rod is we can describe the motion as a translation followed by a rotation. Consider the rigid rod AB<br />

illustrated in the figure 1.3-13.<br />

Figure 1.3-13. Motion of rigid rod<br />

In this figure there is a before and after picture of the rod’s position. By moving the point B to B ′ we<br />

have a translation. This is then followed by a rotation holding B fixed.


Figure 1.3-15. Rotation of axes<br />

A similar situation exists in three dimensions. Consider two sets of Cartesian axes, say a barred and<br />

unbarred system as illustrated in the figure 1.3-14. Let us translate the origin 0 to 0 and then rotate the<br />

(x, y, z) axes until they coincide with the (x, y, z) axes. We consider first the rotation of axes when the<br />

origins 0 and 0 coincide as the translational distance can be represented by a vector bk ,k=1, 2, 3. When<br />

the origin 0 is translated to 0 we have the situation illustrated in the figure 1.3-15, where the barred axes<br />

can be thought of as a transformation due to rotation.<br />

Let<br />

r = x e1 + y e2 + z e3<br />

(1.3.37)<br />

denote the position vector of a variable point P with coordinates (x, y, z) with respect to the origin 0 and the<br />

unit vectors e1, e2, e3. This same point, when referenced with respect to the origin 0 and the unit vectors<br />

ê1, ê2, ê3, has the representation<br />

r = x ê1 + y ê2 + z ê3. (1.3.38)<br />

By considering the projections of r upon the barred and unbarred axes we can construct the transformation<br />

equations relating the barred and unbarred axes. We calculate the projections of r onto the x, y and z axes<br />

and find:<br />

r · e1 = x = x( ê1 · e1)+y( ê2 · e1)+z( ê3 · e1)<br />

r · e2 = y = x( ê1 · e2)+y( ê2 · e2)+z( ê3 · e2)<br />

r · e3 = z = x( ê1 · e3)+y( ê2 · e3)+z( ê3 · e3).<br />

We also calculate the projection of r onto the x, y, z axes and find:<br />

r · ê1 = x = x( e1 · ê1)+y( e2 · ê1)+z( e3 · ê1)<br />

r · ê2 = y = x( e1 · ê2)+y( e2 · ê2)+z( e3 · ê2)<br />

r · ê3 = z = x( e1 · ê3)+y( e2 · ê3)+z( e3 · ê3).<br />

(1.3.39)<br />

(1.3.40)<br />

By introducing the notation (y1,y2,y3) =(x, y, z) (y1, y2, y3)=(x, y, z) and defining θij as the angle<br />

between the unit vectors ei and êj, we can represent the above transformation equations in a more concise<br />

85


86<br />

form. We observe that the direction cosines can be written as<br />

ℓ11 = e1 · ê1 =cosθ11<br />

ℓ21 = e2 · ê1 =cosθ21<br />

ℓ31 = e3 · ê1 =cosθ31<br />

ℓ<strong>12</strong> = e1 · ê2 =cosθ<strong>12</strong><br />

ℓ22 = e2 · ê2 =cosθ22<br />

ℓ32 = e3 · ê2 =cosθ32<br />

which enables us to write the equations (1.3.39) and (1.3.40) in the form<br />

Using the index notation we represent the unit vectors as:<br />

ℓ13 = e1 · ê3 =cosθ13<br />

ℓ23 = e2 · ê3 =cosθ23<br />

ℓ33 = e3 · ê3 =cosθ33<br />

(1.3.41)<br />

yi = ℓijy j and y i = ℓjiyj. (1.3.42)<br />

êr = ℓpr ep or ep = ℓpr êr (1.3.43)<br />

where ℓpr are the direction cosines. In both the barred and unbarred system the unit vectors are orthogonal<br />

and consequently we must have the dot products<br />

êr · êp = δrp and em · en = δmn (1.3.44)<br />

where δij is the Kronecker delta. Substituting equation (1.3.43) into equation (1.3.44) we find the direction<br />

cosines ℓij must satisfy the relations:<br />

The relations<br />

êr · ês = ℓpr ep · ℓms em = ℓprℓms ep · em = ℓprℓmsδpm = ℓmrℓms = δrs<br />

and er · es = ℓrm êm · ℓsn ên = ℓrmℓsn êm · ên = ℓrmℓsnδmn = ℓrmℓsm = δrs.<br />

ℓmrℓms = δrs and ℓrmℓsm = δrs, (1.3.45)<br />

with summation index m, are important relations which are satisfied by the direction cosines associated with<br />

a rotation of axes.<br />

Combining the rotation and translation equations we find<br />

yi = ℓijy j<br />

<br />

+ bi<br />

<br />

. (1.3.46)<br />

rotation translation<br />

We multiply this equation by ℓik and make use of the relations (1.3.45) to find the inverse transformation<br />

These transformations are called linear or affine transformations.<br />

y k = ℓik(yi − bi). (1.3.47)<br />

Consider the xi axes as fixed, while the xi axes are rotating with respect to the xi axes where both sets<br />

of axes have a common origin. Let A = Ai ei denote a vector fixed in and rotating with the xi axes. We<br />

denote by d <br />

A<br />

and<br />

dt<br />

d <br />

A<br />

the derivatives of<br />

dt<br />

A with respect to the fixed (f) and rotating (r) axes. We can<br />

f<br />

r


write, with respect to the fixed axes, that d <br />

A<br />

<br />

dt =<br />

f<br />

dAi<br />

dt ei<br />

i d ei<br />

d ei<br />

+ A . Note that is the derivative of a<br />

dt dt<br />

vector with constant magnitude. Therefore there exists constants ωi, i =1,...,6 such that<br />

d e1<br />

dt = ω3 e2 − ω2 e3<br />

d e2<br />

dt = ω1 e3 − ω4 e1<br />

d e3<br />

dt = ω5 e1 − ω6 e2<br />

i.e. see page 80. From the dot product e1 · e2 = 0 we obtain by differentiation e1 · de2<br />

dt<br />

de1<br />

+ dt · e2 =0<br />

which implies ω4 = ω3. Similarly, from the dot products e1 · e3 and e2 · e3 we obtain by differentiation the<br />

additional relations ω5 = ω2 and ω6 = ω1. The derivative of A with respect to the fixed axes can now be<br />

represented<br />

d <br />

A <br />

<br />

dt<br />

= dAi<br />

dt ei +(ω2A3 − ω3A2) e1 +(ω3A1 − ω1A3) e2 +(ω1A2 − ω2A1) e3 = d <br />

A<br />

<br />

dt<br />

+ ω × A<br />

f<br />

where ω = ωi ei is called an angular velocity vector of the rotating system. The term ω × A represents the<br />

velocity of the rotating system relative to the fixed system and d <br />

A<br />

<br />

dt =<br />

r<br />

dAi<br />

dt ei represents the derivative with<br />

respect to the rotating system.<br />

Employing the special transformation equations (1.3.46) let us examine how tensor quantities transform<br />

when subjected to a translation and rotation of axes. These are our special transformation laws for Cartesian<br />

tensors. We examine only the transformation laws for first and second order Cartesian tensor as higher order<br />

transformation laws are easily discerned. We have previously shown that in general the first and second order<br />

tensor quantities satisfy the transformation laws:<br />

∂yj<br />

Ai = Aj<br />

∂yi A i = A j ∂y i<br />

∂yj<br />

A mn = A ij ∂y m<br />

∂yi<br />

Amn = Aij<br />

∂y n<br />

∂yj<br />

∂yi ∂yj<br />

∂ym ∂yn A m<br />

n = Ai ∂ym ∂yj<br />

j<br />

∂yi ∂yn r<br />

(1.3.48)<br />

(1.3.49)<br />

(1.3.50)<br />

(1.3.51)<br />

(1.3.52)<br />

For the special case of Cartesian tensors we assume that yi and yi,i=1, 2, 3 are linearly independent. We<br />

differentiate the equations (1.3.46) and (1.3.47) and find<br />

∂yj = ℓij<br />

∂yk ∂yi<br />

∂yk ∂yi<br />

= ℓijδjk = ℓik, and = ℓik = ℓikδim = ℓmk.<br />

∂yk ∂ym ∂ym<br />

Substituting these derivatives into the transformation equations (1.3.48) through (1.3.52) we produce the<br />

transformation equations<br />

Ai = Ajℓji<br />

A i = A j ℓji<br />

A mn = A ij ℓimℓjn<br />

Amn = Aijℓimℓjn<br />

A m<br />

n = Aij ℓimℓjn.<br />

87


88<br />

Figure 1.3-16. Transformation to curvilinear coordinates<br />

These are the transformation laws when moving from one orthogonal system to another. In this case the<br />

direction cosines ℓim are constants and satisfy the relations given in equation (1.3.45). The transformation<br />

laws for higher ordered tensors are similar in nature to those given above.<br />

In the unbarred system (y1,y2,y3) the metric tensor and conjugate metric tensor are:<br />

gij = δij and g ij = δij<br />

where δij is the Kronecker delta. In the barred system of coordinates, which is also orthogonal, we have<br />

From the orthogonality relations (1.3.45) we find<br />

Weexaminetheassociatedtensors<br />

g ij = ∂ym<br />

∂y i<br />

∂ym<br />

.<br />

∂yj g ij = ℓmiℓmj = δij and g ij = δij.<br />

A i = g ij Aj<br />

A ij = g im g jn Amn<br />

A i n = gim Amn<br />

Ai = gijA j<br />

Amn = gmignjA ij<br />

A i ij<br />

n = gnjA<br />

and find that the contravariant and covariant components are identical to one another. This holds also in<br />

the barred system of coordinates. Also note that these special circumstances allow the representation of<br />

contractions using subscript quantities only. This type of a contraction is not allowed for general tensors. It<br />

is left as an exercise to try a contraction on a general tensor using only subscripts to see what happens. Note<br />

that such a contraction does not produce a tensor. These special situations are considered in the exercises.<br />

Physical Components<br />

We have previously shown an arbitrary vector A can be represented in many forms depending upon<br />

the coordinate system and basis vectors selected. For example, consider the figure 1.3-16 which illustrates a<br />

Cartesian coordinate system and a curvilinear coordinate system.


Figure 1.3-17. Physical components<br />

In the Cartesian coordinate system we can represent a vector A as<br />

A = Ax e1 + Ay e2 + Az e3<br />

where ( e1, e2, e3) are the basis vectors. Consider a coordinate transformation to a more general coordinate<br />

system, say (x 1 ,x 2 ,x 3 ). The vector A can be represented with contravariant components as<br />

A = A 1 E1 + A 2 E2 + A 3 E3<br />

(1.3.53)<br />

with respect to the tangential basis vectors ( E1, E2, E3). Alternatively, the same vector A can be represented<br />

in the form<br />

A = A1 E 1 + A2 E 2 + A3 E 3<br />

(1.3.54)<br />

having covariant components with respect to the gradient basis vectors ( E 1 , E 2 , E 3 ). These equations are<br />

just different ways of representing the same vector. In the above representations the basis vectors need not<br />

be orthogonal and they need not be unit vectors. In general, the physical dimensions of the components Ai and Aj are not the same.<br />

The physical components of the vector A in a direction is defined as the projection of A upon a unit<br />

vector in the desired direction. For example, the physical component of A in the direction E1 is<br />

Similarly, the physical component of A in the direction E 1 is<br />

A · E1<br />

| A1<br />

=<br />

E1| | E1| = projection of A on E1. (1.3.58)<br />

A · E 1<br />

| E 1 |<br />

= A1<br />

| E 1 | = projection of A on E 1 . (1.3.59)<br />

EXAMPLE 1.3-11. (Physical components) Let α, β, γ denote nonzero positive constants such that the<br />

product relation αγ = 1 is satisfied. Consider the nonorthogonal basis vectors<br />

illustrated in the figure 1.3-17.<br />

E1 = α e1, E2 = β e1 + γ e2, E3 = e3<br />

89


90<br />

It is readily verified that the reciprocal basis is<br />

E 1 = γ e1 − β e2, E 2 = α e2, E 3 = e3.<br />

Consider the problem of representing the vector A = Ax e1 + Ay e2 in the contravariant vector form<br />

This vector has the contravariant components<br />

A = A 1 E1 + A 2 E2 or tensor form A i ,i=1, 2.<br />

A 1 = A · E 1 = γAx − βAy and A 2 = A · E 2 = αAy.<br />

Alternatively, this same vector can be represented as the covariant vector<br />

A = A1 E 1 + A2 E 2 which has the tensor form Ai, i=1, 2.<br />

The covariant components are found from the relations<br />

A1 = A · E1 = αAx<br />

A2 = A · E2 = βAx + γAy.<br />

The physical components of A in the directions E 1 and E 2 are found to be:<br />

A · E1 | E1 |<br />

E 2<br />

A ·<br />

| E2 |<br />

A1<br />

=<br />

| E1 | = γAx − βAy<br />

<br />

γ2 + β2 = A(1)<br />

A2<br />

=<br />

| E2 αAy<br />

=<br />

| α = Ay = A(2).<br />

Note that these same results are obtained from the dot product relations using either form of the vector A.<br />

For example, we can write<br />

A · E 1<br />

and A · E 2<br />

| E1 | = A1( E1 · E1 )+A2( E2 · E1 )<br />

| E1 |<br />

| E2 | = A1( E1 · E2 )+A2( E2 · E2 )<br />

| E2 |<br />

= A(1)<br />

= A(2).<br />

In general, the physical components of a vector A in a direction of a unit vector λi is the generalized<br />

dot product in VN . This dot product is an invariant and can be expressed<br />

gijA i λ j = A i λi = Aiλ i = projection of A in direction of λ i


Physical Components For Orthogonal Coordinates<br />

where<br />

In orthogonal coordinates observe the element of arc length squared in V3 is<br />

ds 2 = gijdx i dx j =(h1) 2 (dx 1 ) 2 +(h2) 2 (dx 2 ) 2 +(h3) 2 (dx 3 ) 2<br />

In this case the curvilinear coordinates are orthogonal and<br />

⎛<br />

gij = ⎝ (h1) 2 0 0<br />

0 (h2) 2 0<br />

0 0 (h3) 2<br />

⎞<br />

⎠ . (1.3.60)<br />

h 2 (i) = g (i)(i) i not summed and gij =0,i= j.<br />

At an arbitrary point in this coordinate system we take λi ,i =1, 2, 3 as a unit vector in the direction<br />

of the coordinate x1 . We then obtain<br />

This is a unit vector since<br />

λ 1 = dx1<br />

ds , λ2 =0, λ 3 =0.<br />

1=gijλ i λ j = g11λ 1 λ 1 = h 2 1 (λ1 ) 2<br />

or λ1 = 1 . Here the curvilinear coordinate system is orthogonal and in this case the physical component<br />

h1<br />

of a vector Ai , in the direction xi , is the projection of Ai on λi in V3. The projection in the x1 direction is<br />

determined from<br />

A(1) = gijA i λ j = g11A 1 λ 1 = h 2 1 1<br />

1A = h1A<br />

h1<br />

1 .<br />

Similarly, we choose unit vectors µ i and ν i ,i=1, 2, 3inthex 2 and x 3 directions. These unit vectors<br />

can be represented<br />

µ 1 =0,<br />

ν 1 =0,<br />

µ 2 = dx2<br />

ds<br />

ν 2 =0,<br />

1<br />

= ,<br />

h2<br />

µ 3 =0<br />

ν 3 = dx3<br />

ds<br />

= 1<br />

h3<br />

and the physical components of the vector A i in these directions are calculated as<br />

A(2) = h2A 2<br />

and A(3) = h3A 3 .<br />

In summary, we can say that in an orthogonal coordinate system the physical components of a contravariant<br />

tensor of order one can be determined from the equations<br />

A(i) =h (i)A (i) = √ g (i)(i)A (i) , i =1, 2 or 3 no summation on i,<br />

which is a short hand notation for the physical components (h1A 1 ,h2A 2 ,h3A 3 ). In an orthogonal coordinate<br />

system the nonzero conjugate metric components are<br />

g (i)(i) = 1<br />

, i =1, 2, or3 no summation on i.<br />

g (i)(i)<br />

91


92<br />

These components are needed to calculate the physical components associated with a covariant tensor of<br />

order one. For example, in the x1−direction, we have the covariant components<br />

1<br />

λ1 = g11λ 1 = h 2 1<br />

h1<br />

and consequently the projection in V3 can be represented<br />

= h1, λ2 =0, λ3 =0<br />

gijA i λ j = gijA i g jm λm = Ajg jm λm = A1λ1g 11 = A1h1<br />

h2 1<br />

In a similar manner we calculate the relations<br />

A(2) = A2<br />

h2<br />

and A(3) = A3<br />

h3<br />

1<br />

= A1<br />

h1<br />

= A(1).<br />

for the other physical components in the directions x 2 and x 3 . These physical components can be represented<br />

in the short hand notation<br />

A(i) = A (i)<br />

h (i)<br />

= A (i)<br />

√ , i =1, 2 or 3 no summation on i.<br />

g(i)(i)<br />

In an orthogonal coordinate system the physical components associated with both the contravariant and<br />

covariant components are the same. To show this we note that when Aigij = Aj is summed on i we obtain<br />

Since gij =0fori = j this equation reduces to<br />

Another form for this equation is<br />

A 1 g1j + A 2 g2j + A 3 g3j = Aj.<br />

A (i) g (i)(i) = A (i), i not summed.<br />

A(i) =A (i)√ g(i)(i) = A (i)<br />

√ g(i)(i)<br />

i not summed,<br />

which demonstrates that the physical components associated with the contravariant and covariant components<br />

are identical.<br />

NOTATION The physical components are sometimes expressed by symbols with subscripts which represent<br />

the coordinate curve along which the projection is taken. For example, let Hi denote the contravariant<br />

components of a first order tensor. The following are some examples of the representation of the physical<br />

components of Hi in various coordinate systems:<br />

orthogonal coordinate tensor physical<br />

coordinates system components components<br />

general (x 1 ,x 2 ,x 3 ) H i H(1),H(2),H(3)<br />

rectangular (x, y, z) H i Hx,Hy,Hz<br />

cylindrical (r, θ, z) H i Hr,Hθ,Hz<br />

spherical (ρ, θ, φ) H i Hρ,Hθ,Hφ<br />

general (u, v, w) H i Hu,Hv,Hw


Higher Order Tensors<br />

The physical components associated with higher ordered tensors are defined by projections in VN just<br />

like the case with first order tensors. For an nth ordered tensor Tij...k we can select n unit vectors λ i ,µ i ,...,ν i<br />

and form the inner product (projection)<br />

Tij...kλ i µ j ...ν k .<br />

When projecting the tensor components onto the coordinate curves, there are N choices for each of the unit<br />

vectors. This produces N n physical components.<br />

The above inner product represents the physical component of the tensor Tij...k along the directions of<br />

the unit vectors λ i ,µ i ,...,ν i . The selected unit vectors may or may not be orthogonal. In the cases where<br />

the selected unit vectors are all orthogonal to one another, the calculation of the physical components is<br />

greatly simplified. By relabeling the unit vectors λi (m) ,λi (n) ,...,λi (p) where (m), (n), ..., (p) representoneof<br />

the N directions, the physical components of a general nth order tensor is represented<br />

T (mn...p)=Tij...kλ i (m) λj<br />

(n) ...λk (p)<br />

EXAMPLE 1.3-<strong>12</strong>. (Physical components)<br />

In an orthogonal curvilinear coordinate system V3 with metric gij, i,j=1, 2, 3, find the physical components<br />

of<br />

(i) the second order tensor Aij. (ii) the second order tensor A ij . (iii) the second order tensor A i j .<br />

Solution: The physical components of Amn,m,n =1, 2, 3 along the directions of two unit vectors λi and<br />

µ i is defined as the inner product in V3. These physical components can be expressed<br />

A(ij) =Amnλ m (i) µn (j)<br />

i, j =1, 2, 3,<br />

where the subscripts (i) and(j) represent one of the coordinate directions. Dropping the subscripts (i) and<br />

(j), we make the observation that in an orthogonal curvilinear coordinate system there are three choices for<br />

the direction of the unit vector λi and also three choices for the direction of the unit vector µ i . These three<br />

choices represent the directions along the x1 ,x2 or x3 coordinate curves which emanate from a point of the<br />

curvilinear coordinate system. This produces a total of nine possible physical components associated with<br />

the tensor Amn.<br />

For example, we can obtain the components of the unit vector λ i ,i=1, 2, 3inthex 1 direction directly<br />

from an examination of the element of arc length squared<br />

By setting dx 2 = dx 3 = 0, we find<br />

ds 2 =(h1) 2 (dx 1 ) 2 +(h2) 2 (dx 2 ) 2 +(h3) 2 (dx 3 ) 2 .<br />

dx 1<br />

ds<br />

1<br />

= = λ<br />

h1<br />

1 , λ 2 =0, λ 3 =0.<br />

This is the vector λ i (1) ,i =1, 2, 3. Similarly, if we choose to select the unit vector λi ,i =1, 2, 3inthex 2<br />

direction, we set dx 1 = dx 3 = 0 in the element of arc length squared and find the components<br />

λ 1 =0, λ 2 = dx2<br />

ds<br />

1<br />

= , λ<br />

h2<br />

3 =0.<br />

93


94<br />

This is the vector λ i (2) ,i=1, 2, 3. Finally, if we select λi ,i=1, 2, 3inthex 3 direction, we set dx 1 = dx 2 =0<br />

in the element of arc length squared and determine the unit vector<br />

λ 1 =0, λ 2 =0, λ 3 = dx3<br />

ds<br />

1<br />

= .<br />

h3<br />

This is the vector λ i (3) ,i =1, 2, 3. Similarly, the unit vector µi can be selected as one of the above three<br />

directions. Examining all nine possible combinations for selecting the unit vectors, we calculate the physical<br />

components in an orthogonal coordinate system as:<br />

A(11) = A11<br />

h1h1<br />

A(21) = A21<br />

h1h2<br />

A(31) = A31<br />

h3h1<br />

A(<strong>12</strong>) = A<strong>12</strong><br />

h1h2<br />

A(22) = A22<br />

h2h2<br />

A(32) = A32<br />

h3h2<br />

These results can be written in the more compact form<br />

For mixed tensors we have<br />

A(ij) = A (i)(j)<br />

h (i)h (j)<br />

A(13) = A13<br />

h1h3<br />

A(23) = A23<br />

h2h3<br />

A(33) = A33<br />

h3h3<br />

no summation on i or j . (1.3.61)<br />

A i j = gim Amj = g i1 A1j + g i2 A2j + g i3 A3j. (1.3.62)<br />

From the fact gij =0fori= j, together with the physical components from equation (1.3.61), the equation<br />

(1.3.62) reduces to<br />

A (i)<br />

(j) = g(i)(i) A (i)(j) = 1<br />

h2 · h (i)h (j)A(ij) no summation on i and i, j =1, 2or3.<br />

(i)<br />

This can also be written in the form<br />

A(ij) =A (i) h (i)<br />

(j) h (j)<br />

Hence, the physical components associated with the mixed tensor A i j<br />

canbeexpressedas<br />

A(11) = A 1 1<br />

A(21) = A 2 h2<br />

1<br />

h1<br />

A(31) = A 3 h3<br />

1<br />

h1<br />

no summation on i or j. (1.3.63)<br />

A(<strong>12</strong>) = A 1 h1<br />

2<br />

h2<br />

A(22) = A 2 2<br />

A(32) = A 3 h3<br />

2<br />

h2<br />

For second order contravariant tensors we may write<br />

in an orthogonal coordinate system<br />

A(13) = A 1 h1<br />

3<br />

h3<br />

A(23) = A 2 h2<br />

3<br />

h3<br />

A(33) = A 3 3.<br />

A ij gjm = A i m = Ai1 g1m + A i2 g2m + A i3 g3m.


Weusethefactgij =0fori = j together with the physical components from equation (1.3.63) to reduce the<br />

above equation to the form A (i)<br />

(m) = A(i)(m) g (m)(m)<br />

we have<br />

no summation on m . In terms of physical components<br />

h (m)<br />

A(im) =A<br />

h (i)<br />

(i)(m) h 2 (m) or A(im) =A (i)(m) h (i)h (m). no summation i, m =1, 2, 3 (1.3.64)<br />

Examining the results from equation (1.3.64) we find that the physical components associated with the<br />

contravariant tensor Aij , in an orthogonal coordinate system, can be written as:<br />

A(11) = A 11 h1h1<br />

A(21) = A 21 h2h1<br />

A(31) = A 31 h3h1<br />

Physical Components in General<br />

A(<strong>12</strong>) = A <strong>12</strong> h1h2<br />

A(22) = A 22 h2h2<br />

A(32) = A 32 h3h2<br />

A(13) = A 13 h1h3<br />

A(23) = A 23 h2h3<br />

A(33) = A 33 h3h3.<br />

In an orthogonal curvilinear coordinate system, the physical components associated with the nth order<br />

tensor Tij...kl along the curvilinear coordinate directions can be represented:<br />

T (ij...kl)=<br />

T (i)(j)...(k)(l)<br />

h (i)h (j) ...h (k)h (l)<br />

no summations.<br />

These physical components can be related to the various tensors associated with Tij...kl. For example, in<br />

an orthogonal coordinate system, the physical components associated with the mixed tensor T ij...m<br />

n...kl can be<br />

expressed as:<br />

T (ij...mn...kl)=T (i)(j)...(m) h (i)h (j) ...h (m)<br />

(n)...(k)(l)<br />

no summations. (1.3.65)<br />

h (n) ...h (k)h (l)<br />

EXAMPLE 1.3-13. (Physical components) Let xi = xi (t),i =1, 2, 3 denote the position vector of a<br />

particle which moves as a function of time t. Assume there exists a coordinate transformation xi = xi (x), for<br />

i =1, 2, 3, of the form given by equations (1.2.33). The position of the particle when referenced with respect<br />

to the barred system of coordinates can be found by substitution. The generalized velocity of the particle<br />

in the unbarred system is a vector with components<br />

v i = dxi<br />

,i=1, 2, 3.<br />

dt<br />

The generalized velocity components of the same particle in the barred system is obtained from the chain<br />

rule. We find this velocity is represented by<br />

v i = dxi<br />

dt<br />

∂xi<br />

=<br />

∂xj dxj dt<br />

This equation implies that the contravariant quantities<br />

(v 1 ,v 2 ,v 3 )=( dx1<br />

dt<br />

= ∂xi<br />

∂x j vj .<br />

dx2 dx3<br />

, ,<br />

dt dt )<br />

95


96<br />

are tensor quantities. These quantities are called the components of the generalized velocity. The coordinates<br />

x 1 ,x 2 ,x 3 are generalized coordinates. This means we can select any set of three independent variables for<br />

the representation of the motion. The variables selected might not have the same dimensions. For example,<br />

in cylindrical coordinates we let (x 1 = r, x 2 = θ, x 3 = z). Here x 1 and x 3 have dimensions of distance but x 2<br />

has dimensions of angular displacement. The generalized velocities are<br />

v 1 = dx1<br />

dt<br />

dr<br />

=<br />

dt , v2 = dx2<br />

dt<br />

dθ<br />

=<br />

dt , v3 = dx3<br />

dt<br />

= dz<br />

dt .<br />

Here v 1 and v 3 have units of length divided by time while v 2 has the units of angular velocity or angular<br />

change divided by time. Clearly, these dimensions are not all the same. Let us examine the physical<br />

components of the generalized velocities. We find in cylindrical coordinates h1 =1,h2 = r, h3 =1andthe<br />

physical components of the velocity have the forms:<br />

vr = v(1) = v 1 h1 = dr<br />

dt , vθ = v(2) = v 2 h2 = r dθ<br />

dt , vz = v(3) = v 3 h3 = dz<br />

dt .<br />

Now the physical components of the velocity all have the same units of length divided by time.<br />

Additional examples of the use of physical components are considered later. For the time being, just<br />

remember that when tensor equations are derived, the equations are valid in any generalized coordinate<br />

system. In particular, we are interested in the representation of physical laws which are to be invariant and<br />

independent of the coordinate system used to represent these laws. Once a tensor equation is derived, we<br />

can chose any type of generalized coordinates and expand the tensor equations. Before using any expanded<br />

tensor equations we must replace all the tensor components by their corresponding physical components in<br />

order that the equations are dimensionally homogeneous. It is these expanded equations, expressed in terms<br />

of the physical components, which are used to solve applied problems.<br />

Tensors and Multilinear Forms<br />

Tensors can be thought of as being created by multilinear forms defined on some vector space V. Let<br />

us define on a vector space V a linear form, a bilinear form and a general multilinear form. We can then<br />

illustrate how tensors are created from these forms.<br />

Definition: (Linear form) Let V denote a vector space which<br />

contains vectors x, x1,x2,....A linear form in x is a scalar function<br />

ϕ(x) having a single vector argument x which satisfies the linearity<br />

properties:<br />

(i) ϕ(x1 + x2) =ϕ(x1)+ϕ(x2)<br />

(ii) ϕ(µx1) =µϕ(x1)<br />

for all arbitrary vectors x1,x2 in V and all real numbers µ.<br />

(1.3.66)


An example of a linear form is the dot product relation<br />

ϕ(x) = A · x (1.3.67)<br />

where A is a constant vector and x is an arbitrary vector belonging to the vector space V.<br />

Note that a linear form in x can be expressed in terms of the components of the vector x and the base<br />

vectors ( e1, e2, e3) used to represent x. To show this, we write the vector x in the component form<br />

x = x i ei = x 1 e1 + x 2 e2 + x 3 e3,<br />

where xi ,i =1, 2, 3arethecomponentsofxwith respect to the basis vectors ( e1, e2, e3). By the linearity<br />

property of ϕ we can write<br />

ϕ(x) =ϕ(x i ei) =ϕ(x 1 e1 + x 2 e2 + x 3 e3)<br />

= ϕ(x 1 e1)+ϕ(x 2 e2)+ϕ(x 3 e3)<br />

= x 1 ϕ( e1)+x 2 ϕ( e2)+x 3 ϕ( e3) =x i ϕ( ei)<br />

Thus we can write ϕ(x) =xiϕ( ei) and by defining the quantity ϕ( ei) =ai as a tensor we obtain ϕ(x) =xiai. Note that if we change basis from ( e1, e2, e3) to( E1, E2, E3) then the components of x also must change.<br />

Letting xi denote the components of x with respect to the new basis, we would have<br />

x = x i Ei and ϕ(x) =ϕ(x i Ei) =x i ϕ( Ei).<br />

The linear form ϕ defines a new tensor ai = ϕ( Ei) sothatϕ(x) =xiai. Whenever there is a definite relation<br />

between the basis vectors ( e1, e2, e3) and( E1, E2, E3), say,<br />

Ei = ∂xj<br />

ej, i ∂x<br />

then there exists a definite relation between the tensors ai and ai. This relation is<br />

ai = ϕ( Ei) =ϕ( ∂xj<br />

∂xi ej) = ∂xj<br />

∂xi ϕ( ej) = ∂xj<br />

i aj.<br />

∂x<br />

This is the transformation law for an absolute covariant tensor of rank or order one.<br />

The above idea is now extended to higher order tensors.<br />

Definition: ( Bilinear form) A bilinear form in x and y is a<br />

scalar function ϕ(x, y) with two vector arguments, which satisfies<br />

the linearity properties:<br />

(i) ϕ(x1 + x2,y1) =ϕ(x1,y1)+ϕ(x2,y1)<br />

(ii) ϕ(x1,y1 + y2) =ϕ(x1,y1)+ϕ(x1,y2)<br />

(iii) ϕ(µx1,y1) =µϕ(x1,y1)<br />

(iv) ϕ(x1,µy1) =µϕ(x1,y1)<br />

(1.3.68)<br />

for arbitrary vectors x1,x2,y1,y2 in the vector space V and for all<br />

real numbers µ.<br />

97


98<br />

Note in the definition of a bilinear form that the scalar function ϕ is linear in both the arguments x and<br />

y. An example of a bilinear form is the dot product relation<br />

ϕ(x, y) =x · y (1.3.69)<br />

where both x and y belong to the same vector space V.<br />

The definition of a bilinear form suggests how multilinear forms can be defined.<br />

Definition: (Multilinear forms) A multilinear form of degree M or a M degree<br />

linear form in the vector arguments<br />

is a scalar function<br />

x1,x2,...,xM<br />

ϕ(x1,x2,...,xM )<br />

of M vector arguments which satisfies the property that it is a linear form in each of its<br />

arguments. That is, ϕ must satisfy for each j =1, 2,...,M the properties:<br />

(i) ϕ(x1,...,xj1 + xj2,...xM )=ϕ(x1,...,xj1,...,xM )+ϕ(x1,...,xj2,...,xM )<br />

(ii) ϕ(x1,...,µxj,...,xM )=µϕ(x1,...,xj,...,xM )<br />

(1.3.70)<br />

for all arbitrary vectors x1,...,xM in the vector space V and all real numbers µ.<br />

An example of a third degree multilinear form or trilinear form is the triple scalar product<br />

ϕ(x, y,z) =x · (y × z). (1.3.71)<br />

Note that multilinear forms are independent of the coordinate system selected and depend only upon the<br />

vector arguments. In a three dimensional vector space we select the basis vectors ( e1, e2, e3) and represent<br />

all vectors with respect to this basis set. For example, if x, y,z are three vectors we can represent these<br />

vectors in the component forms<br />

x = x i ei, y = y j ej, z = z k ek (1.3.72)<br />

where we have employed the summation convention on the repeated indices i, j and k. Substituting equations<br />

(1.3.72) into equation (1.3.71) we obtain<br />

ϕ(x i ei,y j ej,z k ek) =x i y j z k ϕ( ei, ej, ek), (1.3.73)<br />

since ϕ is linear in all its arguments. By defining the tensor quantity<br />

ϕ( ei, ej, ek) =eijk<br />

(1.3.74)


(See exercise 1.1, problem 15) the trilinear form, given by equation (1.3.71), with vectors from equations<br />

(1.3.72), can be expressed as<br />

ϕ(x, y,z) =eijkx i y j z k , i,j,k =1, 2, 3. (1.3.75)<br />

The coefficients eijk of the trilinear form is called a third order tensor. It is the familiar permutation symbol<br />

considered earlier.<br />

In a multilinear form of degree M, ϕ(x, y,...,z), the M arguments can be represented in a component<br />

form with respect to a set of basis vectors ( e1, e2, e3). Let these vectors have components xi ,yi ,zi ,i=1, 2, 3<br />

with respect to the selected basis vectors. We then can write<br />

x = x i ei, y = y j ej, z = z k ek.<br />

Substituting these vectors into the M degree multilinear form produces<br />

ϕ(x i ei,y j ej,...,z k ek) =x i y j ···z k ϕ( ei, ej,..., ek). (1.3.76)<br />

Consequently, the multilinear form defines a set of coefficients<br />

aij...k = ϕ( ei, ej,..., ek) (1.3.77)<br />

which are referred to as the components of a tensor of order M. The tensor is thus created by the multilinear<br />

form and has M indices if ϕ is of degree M.<br />

Note that if we change to a different set of basis vectors, say, ( E1, E2, E3) the multilinear form defines<br />

anewtensor<br />

aij...k = ϕ( Ei, Ej,..., Ek). (1.3.78)<br />

This new tensor has a bar over it to distinguish it from the previous tensor. A definite relation exists between<br />

the new and old basis vectors and consequently there exists a definite relation between the components of<br />

the barred and unbarred tensors components. Recall that if we are given a set of transformation equations<br />

y i = y i (x 1 ,x 2 ,x 3 ),i=1, 2, 3, (1.3.79)<br />

from rectangular to generalized curvilinear coordinates, we can express the basis vectors in the new system<br />

by the equations<br />

Ei = ∂yj<br />

∂xi ej, i =1, 2, 3. (1.3.80)<br />

For example, see equations (1.3.11) with y1 = x, y2 = y, y3 = z,x1 = u, x2 = v, x3 = w. Substituting<br />

equations (1.3.80) into equations (1.3.78) we obtain<br />

aij...k = ϕ( ∂yα ∂yβ ∂yγ<br />

eα, eβ,...,<br />

∂xi ∂xj ∂x<br />

By the linearity property of ϕ, this equation is expressible in the form<br />

∂yβ ∂yγ<br />

...<br />

∂xj k eγ).<br />

aij...k = ∂yα<br />

∂xi ∂xk ϕ( eα, eβ,..., eγ)<br />

aij...k = ∂yα<br />

∂xi ∂yβ ∂yγ<br />

... aαβ...γ<br />

∂xj ∂xk This is the familiar transformation law for a covariant tensor of degree M. By selecting reciprocal basis<br />

vectors the corresponding transformation laws for contravariant vectors can be determined.<br />

The above examples illustrate that tensors can be considered as quantities derivable from multilinear<br />

forms defined on some vector space.<br />

99


100<br />

Dual Tensors<br />

The e-permutation symbol is often used to generate new tensors from given tensors. For Ti1i2...im<br />

skew-symmetric tensor, we define the tensor<br />

a<br />

ˆT j1j2...jn−m = 1<br />

m! ej1j2...jn−mi1i2...imTi1i2...im m ≤ n (1.3.81)<br />

as the dual tensor associated with Ti1i2...im. Note that the e-permutation symbol or alternating tensor has<br />

a weight of +1 and consequently the dual tensor will have a higher weight than the original tensor.<br />

The e-permutation symbol has the following properties<br />

e i1i2...iN ei1i2...iN = N!<br />

e i1i2...iN ej1j2...jN = δ i1i2...iN<br />

j1j2...jN<br />

ek1k2...kmi1i2...iN −me j1j2...jmi1i2...iN −m =(N − m)!δ j1j2...jm<br />

k1k2...km<br />

δ j1j2...jm<br />

k1k2...km Tj1j2...jm = m!Tk1k2...km.<br />

(1.3.82)<br />

Using the above properties we can solve for the skew-symmetric tensor in terms of the dual tensor. We find<br />

1<br />

Ti1i2...im =<br />

(n − m)! ei1i2...imj1j2...jn−m ˆ T j1j2...jn−m . (1.3.83)<br />

For example, if Aij i, j =1, 2, 3 is a skew-symmetric tensor, we may associate with it the dual tensor<br />

V i = 1<br />

2! eijkAjk, which is a first order tensor or vector. Note that Aij has the components<br />

⎛<br />

⎞<br />

⎝ 0 A<strong>12</strong> A13<br />

⎠ (1.3.84)<br />

−A<strong>12</strong> 0 A23<br />

−A13 −A23 0<br />

and consequently, the components of the vector V are<br />

(V 1 ,V 2 ,V 3 )=(A23,A31,A<strong>12</strong>). (1.3.85)<br />

Note that the vector components have a cyclic order to the indices which comes from the cyclic properties<br />

of the e-permutation symbol.<br />

As another example, consider the fourth order skew-symmetric tensor Aijkl, i,j,k,l=1,...,n.Wecan<br />

associate with this tensor any of the dual tensor quantities<br />

V = 1<br />

4! eijkl Aijkl<br />

V i = 1<br />

4! eijklm Ajklm<br />

V ij = 1<br />

4! eijklmn Aklmn<br />

V ijk = 1<br />

4! eijklmnp Almnp<br />

V ijkl = 1<br />

4! eijklmnpr Amnpr<br />

Applications of dual tensors can be found in section 2.2.<br />

(1.3.86)


EXERCISE 1.3<br />

◮ 1.<br />

∂x<br />

(a) From the transformation law for the second order tensor gij = gab<br />

a<br />

∂xi ∂xb ∂xj solve for the gab in terms of gij. (b) Show that if gij is symmetric in one coordinate system it is symmetric in all coordinate systems.<br />

(c) Let g = det(gij )andg = det(gij) andshowthatg = gJ2 ( x<br />

x ) and consequently g = √ gJ( x<br />

). This<br />

x<br />

shows that g is a scalar invariant of weight 2 and √ g is a scalar invariant of weight 1.<br />

◮ 2. For<br />

gij = ∂ym<br />

∂x i<br />

∂ym ∂xj show that g ij = ∂xi<br />

∂ym ∂xj ∂ym ◮ 3. Show that in a curvilinear coordinate system which is orthogonal we have:<br />

<br />

<br />

◮ 4. Show that g = det(gij) = <br />

∂y<br />

<br />

i<br />

∂xj <br />

<br />

<br />

<br />

(a) g = det(gij) =g11g22g33<br />

(b) gmn = g mn =0 form= n<br />

(c) g NN = 1<br />

for N =1, 2, 3 (no summation on N)<br />

2<br />

gNN<br />

= J 2 , where J is the Jacobian.<br />

◮ 5. Define the quantities h1 = hu = | ∂r<br />

∂u |, h2 = hv = | ∂r<br />

∂v |, h3 = hw = | ∂r<br />

| and construct the unit<br />

∂w<br />

vectors<br />

eu = 1 ∂r<br />

h1 ∂u , ev = 1 ∂r<br />

h2 ∂v , ew = 1 ∂r<br />

h3 ∂w .<br />

(a) Assume the coordinate system is orthogonal and show that<br />

g11 = h 2 1 =<br />

g22 = h 2 2 =<br />

g33 = h 2 3 =<br />

<br />

∂x<br />

∂u<br />

<br />

∂x<br />

∂v<br />

<br />

∂x<br />

∂w<br />

2<br />

2<br />

2<br />

<br />

∂y<br />

+<br />

∂u<br />

<br />

∂y<br />

+<br />

∂v<br />

<br />

∂y<br />

+<br />

∂w<br />

2<br />

2<br />

2 ∂z<br />

+ ,<br />

∂u<br />

2 ∂z<br />

+ ,<br />

∂v<br />

2 ∂z<br />

+ .<br />

∂w<br />

(b) Show that dr can be expressed in the form dr = h1 eu du + h2 ev dv + h3 ew dw.<br />

(c) Show that the volume of the elemental parallelepiped having dr as diagonal can be represented<br />

Hint:<br />

dτ = √ gdudvdw= Jdudvdw=<br />

| A · ( B × <br />

<br />

<br />

C)| = <br />

<br />

<br />

2<br />

∂(x, y, z)<br />

∂(u, v, w) dudvdw.<br />

A1 A2 A3<br />

B1 B2 B3<br />

C1 C2 C3<br />

<br />

<br />

<br />

<br />

<br />

<br />

101


102<br />

Figure 1.3-18 Oblique cylindrical coordinates.<br />

◮ 6. For the change dr given in problem 5, show the elemental parallelepiped with diagonal dr has:<br />

(a) the element of area dS1 =<br />

dvdw in the u =constant surface.<br />

<br />

(b) The element of area dS2 =<br />

<br />

(c) the element of area dS3 =<br />

g22g33 − g 2 23<br />

g33g11 − g 2 13<br />

g11g22 − g 2 <strong>12</strong><br />

dudw in the v =constant surface.<br />

dudv in the w =constant surface.<br />

(d) What do the above elements of area reduce to in the special case the curvilinear coordinates are orthogonal?<br />

Hint:<br />

| A × <br />

B| = ( A × B) · ( A × B)<br />

<br />

= ( A · A)( B · B) − ( A · B)( A · .<br />

B)<br />

◮ 7. In Cartesian coordinates you are given the affine transformation. xi = ℓijxj where<br />

x1 = 1<br />

15 (5x1 − 14x2 +2x3), x2 = − 1<br />

3 (2x1 + x2 +2x3), x3 = 1<br />

15 (10x1 +2x2 − 11x3)<br />

(a) Show the transformation is orthogonal.<br />

(b) A vector A(x1,x2,x3) in the unbarred system has the components<br />

A1 =(x1) 2 , A2 =(x2) 2<br />

A3 =(x3) 2 .<br />

Find the components of this vector in the barred system of coordinates.<br />

◮ 8. Calculate the metric and conjugate metric tensors in cylindrical coordinates (r, θ, z).<br />

◮ 9. Calculate the metric and conjugate metric tensors in spherical coordinates (ρ, θ, φ).<br />

◮ 10. Calculate the metric and conjugate metric tensors in parabolic cylindrical coordinates (ξ,η,z).<br />

◮ 11. Calculate the metric and conjugate metric components in elliptic cylindrical coordinates (ξ,η,z).<br />

◮ <strong>12</strong>. Calculate the metric and conjugate metric components for the oblique cylindrical coordinates (r, φ, η),<br />

illustrated in figure 1.3-18, where x = r cosφ , y = r sin φ + η cos α, z = η sin α and α is a parameter<br />

0


◮ 13. Calculate the metric and conjugate metric tensor associated with the toroidal surface coordinates<br />

(ξ,η) illustrated in the figure 1.3-19, where<br />

x =(a + b cos ξ)cosη<br />

y =(a + b cos ξ)sinη<br />

z = b sin ξ<br />

a>b>0<br />

0


104<br />

Figure 1.3-20. Spherical surface coordinates<br />

◮ 18. Given the fourth order tensor Cikmp = λδikδmp + µ(δimδkp + δipδkm)+ν(δimδkp − δipδkm) whereλ, µ<br />

and ν are scalars and δij is the Kronecker delta. Show that under an orthogonal transformation of rotation of<br />

axes with xi = ℓijxj where ℓrsℓis = ℓmrℓmi = δri the components of the above tensor are unaltered. Any<br />

tensor whose components are unaltered under an orthogonal transformation is called an ‘isotropic’ tensor.<br />

Another way of stating this problem is to say “Show Cikmp is an isotropic tensor.”<br />

◮ 19. Assume Aijl is a third order covariant tensor and B pqmn is a fourth order contravariant tensor. Prove<br />

that AiklB klmn is a mixed tensor of order three, with one covariant and two contravariant indices.<br />

◮ 20. Assume that Tmnrs is an absolute tensor. Show that if Tijkl + Tijlk = 0 in the coordinate system x r<br />

then T ijkl + T ijlk = 0 in any other coordinate system x r .<br />

◮ 21. Show that<br />

Hint: See problem 38, Exercise 1.1<br />

<br />

<br />

<br />

ɛijkɛrst = <br />

<br />

<br />

gir gis git<br />

gjr gjs gjt<br />

gkr gks gkt<br />

◮ 22. Determine if the tensor equation ɛmnpɛmij +ɛmnjɛmpi = ɛmniɛmpj is true or false. Justify your answer.<br />

◮ 23. Prove the epsilon identity g ij ɛiptɛjrs = gprgts − gpsgtr. Hint: See problem 38, Exercise 1.1<br />

◮ 24. Let A rs denote a skew-symmetric contravariant tensor and let cr = 1<br />

2 ɛrmnA mn where<br />

ɛrmn = √ germn. Show that cr are the components of a covariant tensor. Write out all the components.<br />

◮ 25. Let Ars denote a skew-symmetric covariant tensor and let c r = 1<br />

2 ɛrmnAmn where ɛ rmn = 1<br />

√ e<br />

g rmn .<br />

Show that cr are the components of a contravariant tensor. Write out all the components.


◮ 26. Let ApqB qs<br />

r = Cs pr where Bqs r is a relative tensor of weight ω1 and Cs pr is a relative tensor of weight<br />

ω2. Prove that Apq is a relative tensor of weight (ω2 − ω1).<br />

◮ 27. When Ai j is an absolute tensor prove that √ gAi j is a relative tensor of weight +1.<br />

◮ 28. When A i j<br />

1<br />

is an absolute tensor prove that √<br />

g Ai j is a relative tensor of weight −1.<br />

◮ 29.<br />

(a) Show eijk is a relative tensor of weight +1.<br />

(b) Show ɛ ijk = 1<br />

√ g e ijk is an absolute tensor. Hint: See example 1.1-25.<br />

◮ 30. The equation of a surface can be represented by an equation of the form Φ(x 1 ,x 2 ,x 3 )=constant.<br />

Show that a unit normal vector to the surface can be represented by the vector<br />

n i =<br />

ij ∂Φ g ∂xj (g mn ∂Φ<br />

∂x m<br />

∂Φ<br />

∂x n ) 1<br />

2<br />

◮ 31. Assume that g ij = λgij with λ a nonzero constant. Find and calculate g ij in terms of g ij .<br />

◮ 32. Determine if the following tensor equation is true. Justify your answer.<br />

Hint: See problem 21, Exercise 1.1.<br />

ɛrjkA r i + ɛirkA r j + ɛijrA r k = ɛijkA r r .<br />

◮ 33. Show that for Ci and C i associated tensors, and C i = ɛ ijk AjBk, then Ci = ɛijkA j B k<br />

◮ 34. Prove that ɛ ijk and ɛijk are associated tensors. Hint: Consider the determinant of gij.<br />

◮ 35. Show ɛ ijk AiBjCk = ɛijkA i B j C k .<br />

◮ 36. Let T i j ,i,j=1, 2, 3 denote a second order mixed tensor. Show that the given quantities are scalar<br />

invariants.<br />

◮ 37.<br />

(i) I1 = T i<br />

i<br />

(ii) I2 = 1<br />

2<br />

(iii) I3 = det|T i j |<br />

.<br />

(T i<br />

i ) 2 − T i m<br />

(a) Assume Aij and Bij ,i,j=1, 2, 3 are absolute contravariant tensors, and determine if the inner product<br />

C ik = A ij B jk is an absolute tensor?<br />

(b) Assume that the condition ∂xj<br />

∂xn ∂xj ∂xm = δnm is satisfied, and determine whether the inner product in<br />

part (a) is a tensor?<br />

(c) Consider only transformations which are a rotation and translation of axes yi = ℓijyj + bi, where ℓij are<br />

direction cosines for the rotation of axes. Show that ∂yj ∂yj = δnm<br />

∂yn ∂ym<br />

m<br />

T<br />

i<br />

105


106<br />

◮ 38. For Aijk a Cartesian tensor, determine if a contraction on the indices i and j is allowed. That<br />

is, determine if the quantity Ak = Aiik, (summation on i) is a tensor. Hint: See part(c) of the previous<br />

problem.<br />

◮ 39. Prove the e-δ identity e ijk eimn = δ j mδ k n − δ j nδ k m.<br />

◮ 40. Consider the vector Vk, k=1, 2, 3 and define the matrix (aij) having the elements aij = eijkVk,<br />

where eijk is the e−permutation symbol.<br />

(a) Solve for Vi in terms of amn by multiplying both sides of the given equation by eijl and note the e − δ<br />

identity allows us to simplify the result.<br />

(b) Sum the given expression on k and then assign values to the free indices (i,j=1,2,3) and compare your<br />

results with part (a).<br />

(c) Is aij symmetric, skew-symmetric, or neither?<br />

◮ 41. It can be shown that the continuity equation of fluid dynamics can be expressed in the tensor form<br />

1<br />

√ g<br />

∂<br />

∂x r (√ gϱV r )+ ∂ϱ<br />

∂t =0,<br />

where ϱ is the density of the fluid, t is time, V r ,withr =1, 2, 3 are the velocity components and g = |gij|<br />

is the determinant of the metric tensor. Employing the summation convention and replacing the tensor<br />

components of velocity by their physical components, express the continuity equation in<br />

(a) Cartesian coordinates (x, y, z) with physical components Vx,Vy,Vz.<br />

(b) Cylindrical coordinates (r, θ, z) with physical components Vr,Vθ,Vz.<br />

(c) Spherical coordinates (ρ, θ, φ) with physical components Vρ,Vθ,Vφ.<br />

◮ 42. Let x 1 ,x 2 ,x 3 denote a set of skewed coordinates with respect to the Cartesian coordinates y 1 ,y 2 ,y 3 .<br />

Assume that E1, E2, E3 are unit vectors in the directions of the x 1 ,x 2 and x 3 axes respectively. If the unit<br />

vectors satisfy the relations<br />

E1 · E1 =1<br />

E2 · E2 =1<br />

E3 · E3 =1<br />

then calculate the metrices gij and conjugate metrices g ij .<br />

E1 · E2 =cosθ<strong>12</strong><br />

E1 · E3 =cosθ13<br />

E2 · E3 =cosθ23,<br />

◮ 43. Let Aij, i,j=1, 2, 3, 4 denote the skew-symmetric second rank tensor<br />

⎛<br />

0 a b<br />

⎞<br />

c<br />

⎜ −a<br />

Aij = ⎝<br />

−b<br />

0<br />

−d<br />

d<br />

0<br />

e⎟<br />

⎠ ,<br />

f<br />

−c −e −f 0<br />

where a, b, c, d, e, f are complex constants. Calculate the components of the dual tensor<br />

V ij = 1<br />

2 eijkl Akl.


◮ 44. In Cartesian coordinates the vorticity tensor at a point in a fluid medium is defined<br />

ωij = 1<br />

<br />

∂Vj ∂Vi<br />

−<br />

2 ∂xi ∂xj <br />

where Vi are the velocity components of the fluid at the point. The vorticity vector at a point in a fluid<br />

medium in Cartesian coordinates is defined by ω i = 1<br />

2 eijkωjk. Show that these tensors are dual tensors.<br />

◮ 45. Write out the relation between each of the components of the dual tensors<br />

ˆT ij = 1<br />

2 eijklTkl i, j, k, l =1, 2, 3, 4<br />

and show that if ijkl is an even permutation of <strong>12</strong>34, then ˆ T ij = Tkl.<br />

◮ 46. Consider the general affine transformation ¯xi = aijxj where (x1 ,x2 ,x3 )=(x, y, z) withinverse<br />

transformation xi = bij ¯xj. Determine (a) the image of the plane Ax + By + Cz + D = 0 under this<br />

transformation and (b) the image of a second degree conic section<br />

Ax 2 +2Bxy + Cy 2 + Dx + Ey + F =0.<br />

◮ 47. Using a multilinear form of degree M, derive the transformation law for a contravariant vector of<br />

degree M.<br />

◮ 48. Let g denote the determinant of gij and show that ∂g ∂gij<br />

= ggij .<br />

∂xk ∂xk ◮ 49. We have shown that for a rotation of xyz axes with respect to a set of fixed ¯x¯y¯z axes, the derivative<br />

of a vector A with respect to an observer on the barred axes is given by<br />

d <br />

A <br />

<br />

dt =<br />

f<br />

d <br />

A<br />

<br />

dt + ω ×<br />

r<br />

A.<br />

Introduce the operators<br />

Df A = d <br />

A<br />

<br />

dt = derivative in fixed system<br />

f<br />

Dr A = d <br />

A<br />

<br />

dt = derivative in rotating system<br />

r<br />

(a) Show that Df A =(Dr + ω×) A.<br />

(b) Consider the special case that the vector A is the position vector r. Show that Dfr =(Dr + ω×)r<br />

produces <br />

<br />

V <br />

=<br />

f<br />

<br />

<br />

V <br />

+ ω × r where<br />

r<br />

<br />

<br />

V <br />

represents the velocity of a particle relative to the fixed system<br />

<br />

f<br />

<br />

and V <br />

represents the velocity of a particle with respect to the rotating system of coordinates.<br />

r <br />

<br />

<br />

<br />

(c) Show that a <br />

= a <br />

+ ω × (ω × r) wherea<br />

represents the acceleration of a particle relative to the<br />

f r<br />

f<br />

<br />

fixed system and a represents the acceleration of a particle with respect to the rotating system.<br />

r<br />

(d) Show in the special case ω is a constant that<br />

<br />

<br />

a =2ω × V + ω × (ω × r)<br />

f<br />

where V is the velocity of the particle relative to the rotating system. The term 2ω × V is referred to<br />

as the Coriolis acceleration and the term ω × (ω × r) is referred to as the centripetal acceleration.<br />

107


108<br />

§1.4 DERIVATIVE OF A TENSOR<br />

In this section we develop some additional operations associated with tensors. Historically, one of the<br />

basic problems of the tensor calculus was to try and find a tensor quantity which is a function of the metric<br />

tensor gij and some of its derivatives ∂gij<br />

∂ 2 gij<br />

,<br />

∂xm ∂xm , .... A solution of this problem is the fourth order<br />

∂xn Riemann Christoffel tensor Rijkl to be developed shortly. In order to understand how this tensor was arrived<br />

at, we must first develop some preliminary relationships involving Christoffel symbols.<br />

Christoffel Symbols<br />

Let us consider the metric tensor gij which we know satisfies the transformation law<br />

Define the quantity<br />

(α, β, γ) = ∂gαβ ∂gab<br />

γ =<br />

∂x ∂xc ∂xc ∂xγ ∂xa ∂xα ∂x<br />

gαβ = gab<br />

a<br />

∂xα ∂xb . β ∂x<br />

∂x b<br />

∂x<br />

∂<br />

+ gab β 2xa ∂xα∂xγ ∂x b<br />

∂x<br />

β + gab<br />

∂xa ∂xα ∂ 2 x b<br />

∂x β ∂x γ<br />

and form the combination of terms 1<br />

[(α, β, γ)+(β,γ,α) − (γ,α,β)] to obtain the result<br />

2<br />

<br />

1 ∂gαβ<br />

2 ∂xγ + ∂gβγ ∂xα − ∂gγα ∂xβ <br />

= 1<br />

<br />

∂gab<br />

2 ∂x<br />

∂x<br />

∂x b<br />

∂gbc ∂gca<br />

+ − c a<br />

∂x a<br />

∂x α<br />

∂x b<br />

∂x β<br />

∂x c<br />

∂x<br />

γ + gab<br />

∂xb ∂xβ ∂2xa ∂xα γ . (1.4.1)<br />

∂x<br />

In this equation the combination of derivatives occurring inside the brackets is called a Christoffel symbol<br />

of the first kind and is defined by the notation<br />

[ac, b] =[ca, b] = 1<br />

<br />

∂gab ∂gbc ∂gac<br />

+ −<br />

2 ∂xc ∂xa ∂xb <br />

. (1.4.2)<br />

The equation (1.4.1) defines the transformation for a Christoffel symbol of the first kind and can be expressed<br />

as<br />

[α γ,β]=[ac, b] ∂xa<br />

∂xα ∂xb ∂xβ ∂xc ∂<br />

γ + gab<br />

∂x 2xa ∂xα∂xγ ∂xb . (1.4.3)<br />

β ∂x<br />

Observe that the Christoffel symbol of the first kind [ac, b] does not transform like a tensor. However, it is<br />

symmetric in the indices a and c.<br />

At this time it is convenient to use the equation (1.4.3) to develop an expression for the second derivative<br />

term which occurs in that equation as this second derivative term arises in some of our future considerations.<br />

To solve for this second derivative we can multiply equation (1.4.3) by ∂xβ<br />

∂xd gde and simplify the result to the<br />

form<br />

∂2xe ∂xα∂x γ = −gde [ac, d] ∂xa<br />

∂xα ∂xc γ + [αγ,β]∂xβ<br />

∂x ∂xd gde . (1.4.4)<br />

The transformation g de λµ ∂xd<br />

= g<br />

∂xλ ∂xe µ allows us to express the equation (1.4.4) in the form<br />

∂x<br />

∂2xe ∂xα∂xγ = −gde [ac, d] ∂xa<br />

∂xα ∂xc ∂xγ + gβµ [αγ,β] ∂xe<br />

µ . (1.4.5)<br />

∂x


Define the Christoffel symbol of the second kind as<br />

<br />

i i<br />

= = g<br />

jk kj<br />

iα [jk,α]= 1<br />

2 giα<br />

<br />

∂gkα<br />

∂x<br />

∂x<br />

∂x α<br />

∂gjα ∂gjk<br />

+ − j k<br />

<br />

. (1.4.6)<br />

This Christoffel symbol of the second kind is symmetric in the indices j and k and from equation (1.4.5) we<br />

see that it satisfies the transformation law<br />

e<br />

a<br />

µ ∂x e ∂x<br />

µ =<br />

αγ ∂x ac ∂xα ∂xc ∂xγ + ∂2xe ∂xα γ . (1.4.7)<br />

∂x<br />

Observe that the Christoffel symbol of the second kind does not transform like a tensor quantity. We can use<br />

the relation defined by equation (1.4.7) to express the second derivative of the transformation equations in<br />

terms of the Christoffel symbols of the second kind. At times it will be convenient to represent the Christoffel<br />

symbols with a subscript to indicate the metric from which they are calculated. Thus, an alternative notation<br />

<br />

<br />

i<br />

i<br />

for is the notation .<br />

jk<br />

jk<br />

g<br />

EXAMPLE 1.4-1. (Christoffel symbols) Solve for the Christoffel symbol of the first kind in terms of<br />

the Christoffel symbol of the second kind.<br />

Solution: By the definition from equation (1.4.6) we have<br />

<br />

i<br />

= g<br />

jk<br />

iα [jk,α].<br />

We multiply this equation by gβi and find<br />

and so<br />

[jk,α]=gαi<br />

gβi<br />

<br />

i<br />

= δ<br />

jk<br />

α β [jk,α]=[jk,β]<br />

<br />

<br />

i<br />

1<br />

N<br />

= gα1 + ···+ gαN .<br />

jk jk<br />

jk<br />

EXAMPLE 1.4-2. (Christoffel symbols of first kind)<br />

Derive formulas to find the Christoffel symbols of the first kind in a generalized orthogonal coordinate<br />

system with metric coefficients<br />

gij =0 for i = j and g (i)(i) = h 2 (i) , i =1, 2, 3<br />

where i is not summed.<br />

Solution: In an orthogonal coordinate system where gij =0fori= j we observe that<br />

[ab, c] = 1<br />

<br />

∂gac<br />

2 ∂x<br />

∂x<br />

∂gbc ∂gab<br />

+ − b a<br />

Here there are 3 3 = 27 quantities to calculate. We consider the following cases:<br />

∂x c<br />

<br />

. (1.4.8)<br />

109


110<br />

CASE I Let a = b = c = i, then the equation (1.4.8) simplifies to<br />

[ab, c] =[ii, i] = 1 ∂gii<br />

2 ∂xi (no summation on i). (1.4.9)<br />

From this equation we can calculate any of the Christoffel symbols<br />

[11, 1], [22, 2], or [33, 3].<br />

CASE II Let a = b = i = c, then the equation (1.4.8) simplifies to the form<br />

[ab, c] =[ii, c] =− 1 ∂gii<br />

2 ∂xc (no summation on i and i = c). (1.4.10)<br />

since, gic =0fori= c. This equation shows how we may calculate any of the six Christoffel symbols<br />

[11, 2], [11, 3], [22, 1], [22, 3], [33, 1], [33, 2].<br />

CASE III Let a = c = i = b, and noting that gib =0fori= b, it can be verified that the equation (1.4.8)<br />

simplifies to the form<br />

[ab, c] =[ib, i] =[bi, i] = 1 ∂gii<br />

2 ∂xb (no summation on i and i = b). (1.4.11)<br />

From this equation we can calculate any of the twelve Christoffel symbols<br />

[<strong>12</strong>, 1] = [21, 1]<br />

[32, 3] = [23, 3]<br />

[13, 1] = [31, 1]<br />

[31, 3] = [13, 3]<br />

[21, 2] = [<strong>12</strong>, 2]<br />

[23, 2] = [32, 2]<br />

CASE IV Let a = b = c and show that the equation (1.4.8) reduces to<br />

This represents the six Christoffel symbols<br />

[ab, c] =0, (a = b = c.)<br />

[<strong>12</strong>, 3] = [21, 3] = [23, 1] = [32, 1] = [31, 2] = [13, 2] = 0.<br />

From the Cases I,II,III,IV all twenty seven Christoffel symbols of the first kind can be determined. In<br />

practice, only the nonzero Christoffel symbols are listed.<br />

EXAMPLE 1.4-3. (Christoffel symbols of the first kind)Find the nonzero Christoffel symbols of the<br />

first kind in cylindrical coordinates.<br />

Solution: From the results of example 1.4-2 we find that for x 1 = r, x 2 = θ, x 3 = z and<br />

g11 =1, g22 =(x 1 ) 2 = r 2 , g33 =1<br />

the nonzero Christoffel symbols of the first kind in cylindrical coordinates are:<br />

[22, 1] = − 1 ∂g22<br />

2 ∂x1 = −x1 = −r<br />

[21, 2] = [<strong>12</strong>, 2] = 1 ∂g22<br />

2 ∂x1 = x1 = r.


EXAMPLE 1.4-4. (Christoffel symbols of the second kind)<br />

Find formulas for the calculation of the Christoffel symbols of the second kind in a generalized orthogonal<br />

coordinate system with metric coefficients<br />

gij =0 for i = j and g (i)(i) = h 2 (i) , i =1, 2, 3<br />

where i is not summed.<br />

Solution: By definition we have<br />

<br />

i<br />

= g<br />

jk<br />

im [jk,m]=g i1 [jk,1] + g i2 [jk,2] + g i3 [jk,3] (1.4.<strong>12</strong>)<br />

By hypothesis the coordinate system is orthogonal and so<br />

g ij =0 for i = j and g ii = 1<br />

gii<br />

i not summed.<br />

The only nonzero term in the equation (1.4.<strong>12</strong>) occurs when m = i and consequently<br />

<br />

i<br />

= g<br />

jk<br />

ii [jk,i]= [jk,i]<br />

gii<br />

no summation on i. (1.4.13)<br />

We can now consider the four cases considered in the example 1.4-2.<br />

CASE I Let j = k = i and show<br />

<br />

i<br />

=<br />

ii<br />

[ii, i]<br />

=<br />

gii<br />

1 ∂gii 1 ∂<br />

=<br />

2gii ∂xi 2 ∂xi ln gii no summation on i. (1.4.14)<br />

CASE II Let k = j = i and show<br />

<br />

i<br />

=<br />

jj<br />

[jj,i]<br />

=<br />

gii<br />

−1 ∂gjj<br />

2gii ∂xi no summation on i or j. (1.4.15)<br />

CASE III Let i = j = k and verify that<br />

<br />

j j<br />

= =<br />

jk kj<br />

[jk,j]<br />

=<br />

gjj<br />

1 ∂gjj 1 ∂<br />

=<br />

2gjj ∂xk 2 ∂xk ln gjj no summation on i or j. (1.4.16)<br />

CASE IV For the case i = j = k we find<br />

<br />

i<br />

=<br />

jk<br />

[jk,i]<br />

=0,<br />

gii<br />

i = j = k no summation on i.<br />

The above cases represent all 27 terms.<br />

111


1<strong>12</strong><br />

EXAMPLE 1.4-5. (Notation) In the case of cylindrical coordinates we can use the above relations and<br />

find the nonzero Christoffel symbols of the second kind:<br />

<br />

1<br />

= −<br />

22<br />

1 ∂g22<br />

2g11 ∂x1 = −x1 = −r<br />

<br />

2 2<br />

1 1<br />

=<br />

= =<br />

<strong>12</strong> 21<br />

1 x1 r<br />

= 1 ∂g22<br />

2g22 ∂x<br />

Note 1: The notation for the above Christoffel symbols are based upon the assumption that x 1 = r, x 2 = θ<br />

and x 3 = z. However, in tensor calculus the choice of the coordinates can be arbitrary. We could just as well<br />

have defined x 1 = z,x 2 = r and x 3 = θ. In this latter case, the numbering system of the Christoffel symbols<br />

changes. To avoid confusion, an alternate method of writing the Christoffel symbols is to use coordinates in<br />

place of the integers 1,2 and 3. For example, in cylindrical coordinates we can write<br />

<br />

θ θ<br />

= =<br />

rθ θr<br />

1<br />

<br />

r<br />

and = −r.<br />

r θθ<br />

If we define x 1 = r, x 2 = θ, x 3 = z, then the nonzero Christoffel symbols are written as<br />

<br />

2 2<br />

= =<br />

<strong>12</strong> 21<br />

1<br />

r<br />

and<br />

<br />

1<br />

= −r.<br />

22<br />

In contrast, if we define x 1 = z,x 2 = r, x 3 = θ, then the nonzero Christoffel symbols are written<br />

<br />

3 3<br />

= =<br />

23 32<br />

1<br />

r<br />

and<br />

<br />

2<br />

= −r.<br />

33<br />

Note 2: Some textbooks use the notation Γa,bc for Christoffel symbols of the first kind and Γ d bc = gda Γa,bc for<br />

Christoffel symbols of the second kind. This notation is not used in these notes since the notation suggests<br />

that the Christoffel symbols are third order tensors, which is not true. The Christoffel symbols of the first<br />

and second kind are not tensors. This fact is clearly illustrated by the transformation equations (1.4.3) and<br />

(1.4.7).<br />

Covariant Differentiation<br />

Let Ai denote a covariant tensor of rank 1 which obeys the transformation law<br />

∂x<br />

Aα = Ai<br />

i<br />

α . (1.4.17)<br />

∂x<br />

Differentiate this relation with respect to xβ and show<br />

∂Aα<br />

∂x<br />

β = Ai<br />

∂ 2 x i<br />

∂xα ∂Ai<br />

+ β ∂x ∂xj ∂x j<br />

∂x β<br />

∂xi α . (1.4.18)<br />

∂x<br />

Now use the relation from equation (1.4.7) to eliminate the second derivative term from (1.4.18) and express<br />

it in the form<br />

i<br />

j<br />

∂Aα σ ∂x i ∂x<br />

= Ai β σ −<br />

∂x αβ ∂x jk ∂xα ∂xk ∂xβ <br />

+ ∂Ai<br />

∂xj ∂xj ∂xβ ∂xi α .<br />

∂x<br />

(1.4.19)


Employing the equation (1.4.17), with α replaced by σ, the equation (1.4.19) is expressible in the form<br />

<br />

∂Aα σ<br />

− Aσ = β ∂x αβ<br />

∂Aj<br />

∂xk ∂xj ∂xα ∂xk j i ∂x<br />

− Ai β ∂x jk ∂xα ∂xk ∂xβ (1.4.20)<br />

or alternatively <br />

∂Aα<br />

− Aσ β ∂x<br />

<br />

σ<br />

αβ<br />

j<br />

∂Aj i ∂x<br />

= − Ai<br />

∂xk jk ∂xα ∂xk . (1.4.21)<br />

β ∂x<br />

Define the quantity<br />

Aj,k = ∂Aj<br />

<br />

i<br />

− Ai<br />

(1.4.22)<br />

∂xk jk<br />

as the covariant derivative of Aj with respect to xk . The equation (1.4.21) demonstrates that the covariant<br />

derivative of a covariant tensor produces a second order tensor which satisfies the transformation law<br />

Aα,β = Aj,k<br />

∂xj ∂xα Other notations frequently used to denote the covariant derivative are:<br />

∂xk . (1.4.23)<br />

β ∂x<br />

Aj,k = Aj;k = A j/k = ∇kAj = Aj|k. (1.4.24)<br />

In the special case where gij are constants the Christoffel symbols of the second kind are zero, and consequently<br />

the covariant derivative reduces to Aj,k = ∂Aj<br />

. That is, under the special circumstances where the<br />

∂xk Christoffel symbols of the second kind are zero, the covariant derivative reduces to an ordinary derivative.<br />

Covariant Derivative of Contravariant Tensor<br />

form<br />

A contravariant tensor A i obeys the transformation law A i = A<br />

A i α ∂xi<br />

= A<br />

∂xα α ∂xi<br />

which can be expressed in the<br />

∂xα (1.4.24)<br />

by interchanging the barred and unbarred quantities. We write the transformation law in the form of equation<br />

(1.4.24) in order to make use of the second derivative relation from the previously derived equation (1.4.7).<br />

Differentiate equation (1.4.24) with respect to xj to obtain the relation<br />

∂Ai ∂xj = Aα ∂2xi ∂xα∂x β<br />

∂xβ ∂Aα<br />

+<br />

∂xj ∂xβ ∂xβ ∂xj ∂xi α . (1.4.25)<br />

∂x<br />

Changing the indices in equation (1.4.25) and substituting for the second derivative term, using the relation<br />

from equation (1.4.7), produces the equation<br />

∂Ai i<br />

m<br />

σ ∂x i ∂x<br />

= Aα<br />

∂xj σ −<br />

αβ ∂x mk ∂xα ∂xk ∂xβ <br />

∂xβ ∂Aα<br />

+<br />

∂xj ∂xβ ∂xβ ∂xj ∂xi α . (1.4.26)<br />

∂x<br />

Applying the relation found in equation (1.4.24), with i replaced by m, together with the relation<br />

∂x β<br />

∂x j<br />

∂x k<br />

∂x β = δk j ,<br />

113


114<br />

we simplify equation (1.4.26) to the form<br />

i ∂A<br />

+<br />

∂xj <br />

i<br />

A<br />

mj<br />

m<br />

<br />

=<br />

∂A σ<br />

+ β ∂x<br />

<br />

σ<br />

A<br />

αβ<br />

α<br />

<br />

∂xβ ∂xj ∂xi σ . (1.4.27)<br />

∂x<br />

Define the quantity<br />

A i ,j = ∂Ai<br />

<br />

i<br />

+ A<br />

∂xj mj<br />

m<br />

(1.4.28)<br />

as the covariant derivative of the contravariant tensor Ai . The equation (1.4.27) demonstrates that a covariant<br />

derivative of a contravariant tensor will transform like a mixed second order tensor and<br />

A i ,j = A σ ∂x<br />

,β<br />

β<br />

∂xj ∂xi σ . (1.4.29)<br />

∂x<br />

Again it should be observed that for the condition where gij are constants we have A i ∂Ai<br />

,j = and the<br />

∂xj covariant derivative of a contravariant tensor reduces to an ordinary derivative in this special case.<br />

In a similar manner the covariant derivative of second rank tensors can be derived. We find these<br />

derivatives have the forms:<br />

Aij,k = ∂Aij<br />

<br />

σ<br />

σ<br />

− Aσj − Aiσ<br />

∂xk ik jk<br />

A i j,k = ∂Aij ∂xk + Aσ <br />

i<br />

j − A<br />

σk<br />

i <br />

σ<br />

σ<br />

jk<br />

A ij<br />

<br />

∂Aij i<br />

,k = + Aσj + A<br />

∂xk σk<br />

iσ<br />

<br />

j<br />

.<br />

σk<br />

(1.4.30)<br />

In general, the covariant derivative of a mixed tensor<br />

of rank n has the form<br />

A ij...k<br />

lm...p,q<br />

= ∂Aij...k<br />

lm...p<br />

∂x q<br />

+ Aσj...k<br />

lm...p<br />

− A ij...k<br />

σm...p<br />

A ij...k<br />

lm...p<br />

<br />

i<br />

+ A<br />

σq<br />

iσ...k<br />

<br />

j<br />

lm...p + ···+ A<br />

σq<br />

ij...σ<br />

<br />

k<br />

lm...p σq<br />

<br />

σ<br />

− A<br />

lq<br />

ij...k<br />

<br />

σ<br />

lσ...p −···−A<br />

mq<br />

ij...k<br />

<br />

σ<br />

lm...σ pq<br />

(1.4.31)<br />

and this derivative is a tensor of rank n +1. Note the pattern of the + signs for the contravariant indices<br />

and the − signs for the covariant indices.<br />

Observe that the covariant derivative of an nth order tensor produces an n+1st order tensor, the indices<br />

of these higher order tensors can also be raised and lowered by multiplication by the metric or conjugate<br />

metric tensor. For example we can write<br />

g im Ajk|m = Ajk| i<br />

and g im A jk |m = A jk | i


Rules for Covariant Differentiation<br />

The rules for covariant differentiation are the same as for ordinary differentiation. That is:<br />

(i) The covariant derivative of a sum is the sum of the covariant derivatives.<br />

(ii) The covariant derivative of a product of tensors is the first times the covariant derivative of the second<br />

plus the second times the covariant derivative of the first.<br />

(iii) Higher derivatives are defined as derivatives of derivatives. Be careful in calculating higher order derivativesasingeneral<br />

Ai,jk = Ai,kj.<br />

EXAMPLE 1.4-6. (Covariant differentiation) Calculate the second covariant derivative Ai,jk.<br />

Solution: The covariant derivative of Ai is<br />

Ai,j = ∂Ai<br />

<br />

σ<br />

− Aσ .<br />

∂xj ij<br />

By definition, the second covariant derivative is the covariant derivative of a covariant derivative and hence<br />

Ai,jk =(Ai,j) ,k = ∂<br />

∂xk <br />

∂Ai σ<br />

m<br />

m<br />

− Aσ − Am,j − Ai,m .<br />

∂xj ij<br />

ik<br />

jk<br />

Simplifying this expression one obtains<br />

Ai,jk = ∂2 Ai<br />

−<br />

∂xj ∂Aσ<br />

−<br />

∂xk ∂xk <br />

σ ∂<br />

− Aσ<br />

ij ∂xk <br />

σ<br />

ij<br />

<br />

∂Am σ m ∂Ai σ m<br />

− Aσ<br />

− − Aσ<br />

.<br />

∂xj mj ik ∂xm im jk<br />

Rearranging terms, the second covariant derivative can be expressed in the form<br />

∂xj ∂Aσ<br />

−<br />

∂xk ∂xk <br />

σ<br />

−<br />

ij<br />

∂Am<br />

∂xj <br />

m<br />

−<br />

ik<br />

∂Ai<br />

∂xm <br />

m<br />

jk<br />

<br />

∂<br />

− Aσ<br />

∂xk <br />

σ σ m m σ<br />

−<br />

−<br />

.<br />

ij im jk ik mj<br />

Ai,jk = ∂2 Ai<br />

(1.4.32)<br />

115


116<br />

Riemann Christoffel Tensor<br />

where<br />

Utilizing the equation (1.4.32), it is left as an exercise to show that<br />

R σ ijk<br />

Ai,jk − Ai,kj = AσR σ ijk<br />

∂<br />

=<br />

∂xj <br />

σ<br />

−<br />

ik<br />

∂<br />

∂xk <br />

σ<br />

+<br />

ij<br />

<br />

m σ<br />

−<br />

ik mj<br />

is called the Riemann Christoffel tensor. The covariant form of this tensor is<br />

<br />

m σ<br />

<br />

ij mk<br />

(1.4.33)<br />

Rhjkl = gihR i jkl. (1.4.34)<br />

It is an easy exercise to show that this covariant form can be expressed in either of the forms<br />

Rinjk = ∂<br />

<br />

∂<br />

s<br />

s<br />

[nk, i] − [nj, i]+[ik, s] − [ij, s]<br />

∂xj ∂xk nj<br />

nk<br />

or Rijkl = 1<br />

2 ∂ gil<br />

2 ∂xj∂xk − ∂2gjl ∂xi∂xk − ∂2gik ∂xj∂xl + ∂2gjk ∂xi∂xl <br />

+ g αβ ([jk,β][il, α] − [jl,β][ik, α]) .<br />

From these forms we find that the Riemann Christoffel tensor is skew symmetric in the first two indices<br />

and the last two indices as well as being symmetric in the interchange of the first pair and last pairs of<br />

indices and consequently<br />

Rjikl = −Rijkl Rijlk = −Rijkl Rklij = Rijkl.<br />

In a two dimensional space there are only four components of the Riemann Christoffel tensor to consider.<br />

These four components are either +R<strong>12</strong><strong>12</strong> or −R<strong>12</strong><strong>12</strong> since they are all related by<br />

R<strong>12</strong><strong>12</strong> = −R21<strong>12</strong> = R2<strong>12</strong>1 = −R<strong>12</strong>21.<br />

In a Cartesian coordinate system Rhijk = 0. The Riemann Christoffel tensor is important because it occurs<br />

in differential geometry and relativity which are two areas of interest to be considered later. Additional<br />

properties of this tensor are found in the exercises of section 1.5.


Physical Interpretation of Covariant Differentiation<br />

In a system of generalized coordinates (x 1 ,x 2 ,x 3 ) we can construct the basis vectors ( E1, E2, E3). These<br />

basis vectors change with position. That is, each basis vector is a function of the coordinates at which they<br />

are evaluated. We can emphasize this dependence by writing<br />

Ei = Ei(x 1 ,x 2 ,x 3 )= ∂r<br />

∂x i<br />

i =1, 2, 3.<br />

Associated with these basis vectors we have the reciprocal basis vectors<br />

E i = E i (x 1 ,x 2 ,x 3 ), i =1, 2, 3<br />

which are also functions of position. A vector A can be represented in terms of contravariant components as<br />

A = A 1 E1 + A 2 E2 + A 3 E3 = A j Ej<br />

or it can be represented in terms of covariant components as<br />

A change in the vector A is represented as<br />

where from equation (1.4.35) we find<br />

(1.4.35)<br />

A = A1 E 1 + A2 E 2 + A3 E 3 = Aj E j . (1.4.36)<br />

d A = ∂ A<br />

dxk<br />

∂xk ∂ A<br />

∂xk = Aj ∂ Ej ∂Aj<br />

+<br />

∂xk or alternatively from equation (1.4.36) we may write<br />

∂ A<br />

= Aj<br />

∂xk ∂x k Ej<br />

We define the covariant derivative of the covariant components as<br />

Ai,k = ∂ A<br />

∂xk · Ei = ∂Ai<br />

∂x<br />

(1.4.37)<br />

∂ Ej ∂Aj<br />

+<br />

∂xk ∂xk E j . (1.4.38)<br />

k + Aj<br />

The covariant derivative of the contravariant components are defined by the relation<br />

Introduce the notation<br />

∂ <br />

Ej m<br />

= Em<br />

∂xk jk<br />

We then have<br />

∂ E j<br />

∂x k · Ei. (1.4.39)<br />

A i ,k = ∂ A<br />

∂x k · E i = ∂Ai<br />

∂x k + Aj ∂ Ej<br />

∂x k · E i . (1.4.40)<br />

and<br />

E i · ∂ <br />

Ej m<br />

= Em ·<br />

∂xk jk<br />

E i <br />

m<br />

= δ<br />

jk<br />

i m =<br />

<br />

i<br />

jk<br />

∂ Ej <br />

j<br />

= − E<br />

∂xk mk<br />

m . (1.4.41)<br />

(1.4.42)<br />

117


118<br />

and<br />

Ei · ∂ Ej <br />

j<br />

= − E<br />

∂xk mk<br />

m · <br />

j<br />

Ei = − δ<br />

mk<br />

m <br />

j<br />

i = − .<br />

ik<br />

(1.4.43)<br />

Then equations (1.4.39) and (1.4.40) become<br />

Ai,k = ∂Ai<br />

<br />

j<br />

−<br />

∂xk ik<br />

A i ,k<br />

∂Ai<br />

= +<br />

∂xk i<br />

jk<br />

<br />

Aj<br />

<br />

A j ,<br />

which is consistent with our earlier definitions from equations (1.4.22) and (1.4.28). Here the first term of<br />

the covariant derivative represents the rate of change of the tensor field as we move along a coordinate curve.<br />

The second term in the covariant derivative represents the change in the local basis vectors as we move<br />

along the coordinate curves. This is the physical interpretation associated with the Christoffel symbols of<br />

the second kind.<br />

We make the observation that the derivatives of the basis vectors in equations (1.4.39) and (1.4.40) are<br />

related since<br />

and consequently<br />

Hence we can express equation (1.4.39) in the form<br />

Ei · E j = δ j<br />

i<br />

∂<br />

∂xk ( Ei · E j )= Ei · ∂ Ej ∂xk + ∂ Ei<br />

∂xk · E j =0<br />

or<br />

Ai,k = ∂Ai<br />

Ei · ∂ Ej ∂xk = − E j · ∂ Ei<br />

∂xk ∂xk − Aj E j · ∂ Ei<br />

∂x<br />

We write the first equation in (1.4.41) in the form<br />

∂ <br />

Ej m<br />

= gim<br />

∂xk jk<br />

E i =[jk,i] E i<br />

and consequently<br />

and<br />

∂ Ej<br />

∂xk · E m <br />

i<br />

= Ei ·<br />

jk<br />

E m <br />

i<br />

= δ<br />

jk<br />

m <br />

m<br />

i =<br />

jk<br />

∂ Ej<br />

∂x k · Em =[jk,i] E i · Em =[jk,i]δ i m =[jk,m].<br />

k . (1.4.44)<br />

(1.4.45)<br />

(1.4.46)<br />

These results also reduce the equations (1.4.40) and (1.4.44) to our previous forms for the covariant derivatives.<br />

The equations (1.4.41) are representations of the vectors ∂ Ei<br />

∂x k and ∂ E j<br />

∂x k in terms of the basis vectors and<br />

reciprocal basis vectors of the space. The covariant derivative relations then take into account how these<br />

vectors change with position and affect changes in the tensor field.<br />

The Christoffel symbols in equations (1.4.46) are symmetric in the indices j and k since<br />

∂ Ej ∂<br />

=<br />

∂xk ∂xk <br />

∂r<br />

∂xj <br />

= ∂<br />

∂xj <br />

∂r<br />

∂xk <br />

= ∂ Ek<br />

. (1.4.47)<br />

∂xj


The equations (1.4.46) and (1.4.47) enable us to write<br />

[jk,m]= Em · ∂ <br />

Ej 1<br />

= Em ·<br />

∂xk 2<br />

∂ Ej<br />

∂xk + Em · ∂ Ek<br />

∂xj <br />

= 1<br />

<br />

∂<br />

2 ∂xk <br />

Em · <br />

Ej + ∂<br />

∂xj <br />

Em · <br />

Ek − Ej · ∂ Em<br />

∂xk − Ek · ∂ Em<br />

∂xj <br />

= 1<br />

<br />

∂<br />

2 ∂xk <br />

Em · <br />

Ej + ∂<br />

∂xj <br />

Em · <br />

Ek − Ej · ∂ Ek<br />

∂xm − Ek · ∂ Ej<br />

∂xm <br />

= 1<br />

<br />

∂<br />

2 ∂xk <br />

Em · <br />

Ej + ∂<br />

∂xj <br />

Em · <br />

Ek − ∂<br />

∂xm <br />

Ej · <br />

Ek<br />

<br />

= 1<br />

<br />

∂gmj ∂gmk ∂gjk<br />

+ −<br />

2 ∂xk ∂xj ∂xm <br />

=[kj, m]<br />

which again agrees with our previous result.<br />

For future reference we make the observation that if the vector A is represented in the form A = AjEj, <br />

involving contravariant components, then we may write<br />

d A = ∂ A<br />

∂xk dxk <br />

∂A<br />

=<br />

j<br />

∂xk Ej + A j ∂ Ej<br />

∂xk <br />

dx k<br />

j ∂A<br />

=<br />

∂xk Ej + A j<br />

<br />

i<br />

Ei dx<br />

jk<br />

k<br />

<br />

j ∂A j<br />

= + A<br />

∂xk mk<br />

m<br />

<br />

Ej dx k = A j<br />

,k dxk (1.4.48)<br />

Ej. <br />

Similarly, if the vector A is represented in the form A = Aj Ej involving covariant components it is left as<br />

an exercise to show that<br />

Ricci’s Theorem<br />

d A = Aj,k dx k E j<br />

Ricci’s theorem states that the covariant derivative of the metric tensor vanishes and gik,l =0.<br />

Proof: We have<br />

gik,l = ∂gik<br />

<br />

m m<br />

− gim − gmk<br />

∂xl kl il<br />

(1.4.49)<br />

gik,l = ∂gik<br />

− [kl, i] − [il, k]<br />

∂xl gik,l = ∂gik<br />

<br />

1 ∂gik ∂gil ∂gkl<br />

− + −<br />

∂xl 2 ∂xl ∂xk ∂xi <br />

− 1<br />

<br />

∂gik ∂gkl ∂gil<br />

+ −<br />

2 ∂xl ∂xi ∂xk <br />

=0.<br />

Because of Ricci’s theorem the components of the metric tensor can be regarded as constants during covariant<br />

differentiation.<br />

EXAMPLE 1.4-7. (Covariant differentiation) Show that δ i j,k =0.<br />

Solution<br />

δ i j,k = ∂δi j<br />

∂xk + δσ <br />

i<br />

j − δ<br />

σk<br />

i <br />

σ i i<br />

σ = − =0.<br />

jk jk jk<br />

119


<strong>12</strong>0<br />

EXAMPLE 1.4-8. (Covariant differentiation) Show that g ij<br />

,k =0.<br />

Solution: Since gijg jk = δk i we take the covariant derivative of this expression and find<br />

(gijg jk ),l = δ k i,l =0<br />

gijg jk<br />

,l + gij,lg jk =0.<br />

But gij,l = 0 by Ricci’s theorem and hence gijg jk<br />

,l =0. We multiply this expression by gim and obtain<br />

g im gijg jk<br />

,l = δm j g jk<br />

,l = gmk ,l =0<br />

which demonstrates that the covariant derivative of the conjugate metric tensor is also zero.<br />

EXAMPLE 1.4-9. (Covariant differentiation) Some additional examples of covariant differentiation<br />

are:<br />

Intrinsic or Absolute Differentiation<br />

(i) (gilA l ),k = gilA l ,k = Ai,k<br />

(ii) (gimgjnA ij ) ,k = gimgjnA ij<br />

,k = Amn,k<br />

The intrinsic or absolute derivative of a covariant vector Ai taken along a curve x i = x i (t),i =1,...,N<br />

is defined as the inner product of the covariant derivative with the tangent vector to the curve. The intrinsic<br />

derivative is represented<br />

δAi dx<br />

= Ai,j<br />

δt j<br />

dt<br />

δAi<br />

δt =<br />

j<br />

∂Ai α dx<br />

− Aα<br />

∂xj ij dt<br />

j<br />

δAi dAi α dx<br />

= − Aα<br />

δt dt ij dt .<br />

(1.4.50)<br />

Similarly, the absolute or intrinsic derivative of a contravariant tensor A i is represented<br />

δAi δt = Ai dx<br />

,j<br />

j<br />

dt<br />

dAi<br />

=<br />

dt +<br />

<br />

i k dxj<br />

A<br />

jk dt .<br />

The intrinsic or absolute derivative is used to differentiate sums and products in the same manner as used<br />

in ordinary differentiation. Also if the coordinate system is Cartesian the intrinsic derivative becomes an<br />

ordinary derivative.<br />

The intrinsic derivative of higher order tensors is similarly defined as an inner product of the covariant<br />

derivative with the tangent vector to the given curve. For example,<br />

δA ij<br />

klm<br />

δt<br />

dx<br />

= Aij<br />

klm,p<br />

p<br />

dt<br />

is the intrinsic derivative of the fifth order mixed tensor A ij<br />

klm .


EXAMPLE 1.4-10. (Generalized velocity and acceleration) Let t denote time and let xi = xi (t)<br />

for i =1,...,N, denote the position vector of a particle in the generalized coordinates (x1 ,...,xN ). From<br />

the transformation equations (1.2.30), the position vector of the same particle in the barred system of<br />

coordinates, (x 1 , x 2 ,...,x N ), is<br />

x i = x i (x 1 (t),x 2 (t),...,x N (t)) = x i (t), i =1,...,N.<br />

The generalized velocity is v i = dxi<br />

dt ,i=1,...,N. The quantity vi transforms as a tensor since by definition<br />

v i = dxi<br />

dt<br />

∂xi<br />

=<br />

∂xj dxj dt<br />

= ∂xi<br />

∂x j vj . (1.4.51)<br />

Let us now find an expression for the generalized acceleration. Write equation (1.4.51) in the form<br />

and differentiate with respect to time to obtain<br />

The equation (1.4.53) demonstrates that dvi<br />

dt<br />

v j i ∂xj<br />

= v<br />

∂xi dvj dt = vi ∂2xj ∂xi∂x k<br />

dxk dvi ∂x<br />

+<br />

dt dt<br />

j<br />

∂xi (1.4.52)<br />

(1.4.53)<br />

does not transform like a tensor. From the equation (1.4.7)<br />

previously derived, we change indices and write equation (1.4.53) in the form<br />

Rearranging terms we find<br />

dv j<br />

dt<br />

= vi dxk<br />

dt<br />

j<br />

a<br />

σ ∂x j ∂x<br />

σ −<br />

ik ∂x ac ∂xi ∂xc ∂xk <br />

+ ∂xj<br />

∂xi dvi dt .<br />

∂vj ∂xk dxk dt +<br />

a<br />

c<br />

j ∂x ∂x<br />

vi<br />

ac i ∂x ∂xk dxk <br />

=<br />

dt<br />

∂xj<br />

∂xi ∂vi ∂xk dxk dt +<br />

<br />

σ i ∂xj<br />

v<br />

ik ∂xσ dxk dt<br />

<br />

j ∂v j<br />

+ v<br />

∂xk ak<br />

a<br />

k dx<br />

dt =<br />

<br />

∂vσ <br />

σ<br />

+ v k ∂x ik<br />

i<br />

<br />

dxk ∂x<br />

dt<br />

j<br />

∂xσ δv j<br />

δt<br />

= δvσ<br />

δt<br />

∂xj σ .<br />

∂x<br />

The above equation illustrates that the intrinsic derivative of the velocity is a tensor quantity. This derivative<br />

is called the generalized acceleration and is denoted<br />

f i = δvi<br />

δt = vi dx<br />

,j<br />

j<br />

dt<br />

dvi<br />

=<br />

dt +<br />

<br />

i<br />

mn<br />

To summarize, we have shown that if<br />

v m v n = d2xi +<br />

dt2 <br />

i<br />

m dx<br />

mn dt<br />

x i = x i (t), i =1,...,N is the generalized position vector, then<br />

v i = dxi<br />

, i =1,...,N is the generalized velocity, and<br />

dt<br />

f i = δvi<br />

δt = vi ,j<br />

dxj , i =1,...,N is the generalized acceleration.<br />

dt<br />

or<br />

dxn , i =1,...,N (1.4.54)<br />

dt<br />

<strong>12</strong>1


<strong>12</strong>2<br />

Parallel Vector Fields<br />

Let y i = y i (t), i =1, 2, 3 denote a space curve Cin a Cartesian coordinate system and let Y i define a<br />

constant vector in this system. Construct at each point of the curve C the vector Y i . This produces a field<br />

of parallel vectors along the curve C. What happens to the curve and the field of parallel vectors when we<br />

transform to an arbitrary coordinate system using the transformation equations<br />

with inverse transformation<br />

y i = y i (x 1 ,x 2 ,x 3 ), i =1, 2, 3<br />

x i = x i (y 1 ,y 2 ,y 3 ), i =1, 2, 3?<br />

The space curve Cin the new coordinates is obtained directly from the transformation equations and can<br />

be written<br />

x i = x i (y 1 (t),y 2 (t),y 3 (t)) = x i (t), i =1, 2, 3.<br />

The field of parallel vectors Y i become X i in the new coordinates where<br />

Y i = X<br />

j ∂yi<br />

. (1.4.55)<br />

∂xj Since the components of Y i are constants, their derivatives will be zero and consequently we obtain by<br />

differentiating the equation (1.4.55), with respect to the parameter t, that the field of parallel vectors Xi must satisfy the differential equation<br />

dX j<br />

dt<br />

∂yi ∂xj + Xj ∂2y i<br />

∂xj∂xm dxm dt<br />

= dY i<br />

dt<br />

=0. (1.4.56)<br />

Changing symbols in the equation (1.4.7) and setting the Christoffel symbol to zero in the Cartesian system<br />

of coordinates, we represent equation (1.4.7) in the form<br />

∂2y i<br />

∂xj i α ∂y<br />

=<br />

∂xm jm ∂xα and consequently, the equation (1.4.56) can be reduced to the form<br />

δX j<br />

δt<br />

j dX<br />

=<br />

dt +<br />

<br />

j k dxm<br />

X<br />

km dt<br />

=0. (1.4.57)<br />

The equation (1.4.57) is the differential equation which must be satisfied by a parallel field of vectors X i<br />

along an arbitrary curve x i (t).


EXERCISE 1.4<br />

◮ 1. Find the nonzero Christoffel symbols of the first and second kind in cylindrical coordinates<br />

(x 1 ,x 2 ,x 3 )=(r, θ, z), where x = r cos θ, y = r sin θ, z = z.<br />

◮ 2. Find the nonzero Christoffel symbols of the first and second kind in spherical coordinates<br />

(x 1 ,x 2 ,x 3 )=(ρ, θ, φ), where x = ρ sin θ cos φ, y = ρ sin θ sin φ, z = ρ cos θ.<br />

◮ 3. Find the nonzero Christoffel symbols of the first and second kind in parabolic cylindrical coordinates<br />

(x 1 ,x 2 ,x 3 )=(ξ,η,z), where x = ξη, y = 1<br />

2 (ξ2 − η 2 ), z = z.<br />

◮ 4. Find the nonzero Christoffel symbols of the first and second kind in parabolic coordinates<br />

(x 1 ,x 2 ,x 3 )=(ξ,η,φ), where x = ξη cos φ, y = ξη sin φ, z = 1<br />

2 (ξ2 − η 2 ).<br />

◮ 5. Find the nonzero Christoffel symbols of the first and second kind in elliptic cylindrical coordinates<br />

(x 1 ,x 2 ,x 3 )=(ξ,η,z), where x =coshξ cos η, y =sinhξ sin η, z = z.<br />

◮ 6. Find the nonzero Christoffel symbols of the first and second kind for the oblique cylindrical coordinates<br />

(x 1 ,x 2 ,x 3 )=(r, φ, η), where x = r cos φ, y = r sin φ+η cos α, z = η sin α with 0


<strong>12</strong>4<br />

◮ 13.<br />

(a) Show √ g is a relative scalar of weight +1.<br />

(b) Use the results from problem 9(c) and problem 43 of the exercises to show that<br />

( √ g),k = ∂√ <br />

g m √g<br />

−<br />

=0.<br />

∂xk <br />

km<br />

m<br />

(c) Show that =<br />

km<br />

∂<br />

∂xk ln(√g)= 1 ∂g<br />

.<br />

2g ∂xk <br />

m<br />

◮ 14. Use the result from problem 9(b) to show =<br />

km<br />

∂<br />

∂xk ln(√g)= 1 ∂g<br />

.<br />

2g ∂xk Hint: Expand the covariant derivative ɛrst,p and then substitute ɛrst = √ gerst. Simplify by inner<br />

multiplication with erst<br />

√<br />

g and note the Exercise 1.1, problem 26.<br />

◮ 15. Calculate the covariant derivative A i ,m and then contract on m and i to show that<br />

A i ,i = 1<br />

√ g<br />

∂<br />

∂xi √ i<br />

gA .<br />

◮ 16. Show 1 ∂<br />

√<br />

g ∂xj √ ij<br />

gg <br />

i<br />

+ g<br />

pq<br />

pq =0. Hint: See problem 14.<br />

◮ 17. Prove that the covariant derivative of a sum equals the sum of the covariant derivatives.<br />

Hint: Assume Ci = Ai + Bi and write out the covariant derivative for Ci,j.<br />

◮ 18. Let Ci j = AiBj and prove that the covariant derivative of a product equals the first term times the<br />

covariant derivative of the second term plus the second term times the covariant derivative of the first term.<br />

◮ 19. Start with the transformation law Āij<br />

∂x<br />

= Aαβ<br />

α<br />

∂¯x i<br />

∂xβ and take an ordinary derivative of both sides<br />

∂¯x j<br />

with respect to ¯x k and hence derive the relation for Aij,k given in (1.4.30).<br />

◮ 20. Start with the transformation law A ij =<br />

∂xi<br />

Āαβ<br />

∂¯x α<br />

with respect to xk and hence derive the relation for A ij<br />

,k<br />

◮ 21. Find the covariant derivatives of<br />

(a) A ijk<br />

∂xj and take an ordinary derivative of both sides<br />

∂¯x β<br />

given in (1.4.30).<br />

(b) A ij<br />

k (c) A i jk (d) Aijk<br />

◮ 22. Find the intrinsic derivative along the curve x i = x i (t), i =1,...,N for<br />

◮ 23.<br />

(a) A ijk<br />

(a) Assume A = AiEi and show that d A = Ai ,k dxk Ei. <br />

(b) Assume A = Ai Ei and show that d A = Ai,k dxk E i .<br />

(b) A ij<br />

k (c) A i jk (d) Aijk


◮ 24. (parallel vector field) Imagine a vector field A i = A i (x 1 ,x 2 ,x 3 ) which is a function of position.<br />

Assume that at all points along a curve x i = x i (t),i =1, 2, 3 the vector field points in the same direction,<br />

we would then have a parallel vector field or homogeneous vector field. Assume A is a constant, then<br />

d A = ∂ A<br />

∂xk dxk =0. Show that for a parallel vector field the condition Ai,k = 0 must be satisfied.<br />

◮ 25. Show that<br />

∂[ik, n]<br />

∂x j<br />

∂<br />

= gnσ<br />

∂xj ◮ 26. Show Ar,s − As,r = ∂Ar ∂As<br />

− .<br />

∂xs ∂xr <br />

<br />

σ<br />

σ<br />

+([nj, σ]+[σj, n]) .<br />

ik<br />

ik<br />

◮ 27. In cylindrical coordinates you are given the contravariant vector components<br />

A 1 = r A 2 =cosθ A 3 = z sin θ<br />

(a) Find the physical components Ar, Aθ, and Az.<br />

(b) Denote the physical components of A i ,j<br />

Find these physical components.<br />

, i, j =1, 2, 3, by<br />

◮ 28. Find the covariant form of the contravariant tensor C i = ɛ ijk Ak,j.<br />

Express your answer in terms of A k ,j .<br />

Arr Arθ Arz<br />

Aθr Aθθ Aθz<br />

Azr Azθ Azz.<br />

◮ 29. In Cartesian coordinates let x denote the magnitude of the position vector xi. Show that (a) x ,j = 1<br />

x xj<br />

(b) x ,ij = 1<br />

x δij − 1<br />

x3 xixj (c) x ,ii = 2<br />

1<br />

. (d) LetU =<br />

x x , x = 0,andshowthatU ,ij = −δij 3xixj<br />

+<br />

x3 x5 and<br />

U ,ii =0.<br />

◮ 30. Consider a two dimensional space with element of arc length squared<br />

ds 2 = g11(du 1 ) 2 + g22(du 2 ) 2<br />

and metric<br />

<br />

g11<br />

gij =<br />

0<br />

<br />

0<br />

g22<br />

where u 1 ,u 2 are surface coordinates.<br />

(a) Find formulas to calculate the Christoffel symbols of the first kind.<br />

(b) Find formulas to calculate the Christoffel symbols of the second kind.<br />

◮ 31. Find the metric tensor and Christoffel symbols of the first and second kind associated with the<br />

two dimensional space describing points on a cylinder of radius a. Let u 1 = θ and u 2 = z denote surface<br />

coordinates where<br />

x = a cos θ = a cos u 1<br />

y = a sin θ = a sin u 1<br />

z = z = u 2<br />

<strong>12</strong>5


<strong>12</strong>6<br />

◮ 32. Find the metric tensor and Christoffel symbols of the first and second kind associated with the<br />

two dimensional space describing points on a sphere of radius a. Let u 1 = θ and u 2 = φ denote surface<br />

coordinates where<br />

x = a sin θ cos φ = a sin u 1 cos u 2<br />

y = a sin θ sin φ = a sin u 1 sin u 2<br />

z = a cos θ = a cos u 1<br />

◮ 33. Find the metric tensor and Christoffel symbols of the first and second kind associated with the<br />

two dimensional space describing points on a torus having the parameters a and b and surface coordinates<br />

u 1 = ξ, u 2 = η. illustrated in the figure 1.3-19. The points on the surface of the torus are given in terms<br />

of the surface coordinates by the equations<br />

x =(a + b cos ξ)cosη<br />

y =(a + b cos ξ)sinη<br />

z = b sin ξ<br />

◮ 34. Prove that eijka m b j c k u i ,m + eijka i b m c k u j ,m + eijka i b j c m u k ,m = ur ,reijka i b j c k . Hint: See Exercise 1.3,<br />

problem 32 and Exercise 1.1, problem 21.<br />

◮ 35. Calculate the second covariant derivative A i ,jk.<br />

◮ 36. Show that σ ij 1 ∂<br />

,j = √<br />

g ∂xj √ ij<br />

gσ + σ mn<br />

<br />

i<br />

mn<br />

◮ 37. Find the contravariant, covariant and physical components of velocity and acceleration in (a) Cartesian<br />

coordinates and (b) cylindrical coordinates.<br />

◮ 38. Find the contravariant, covariant and physical components of velocity and acceleration in spherical<br />

coordinates.<br />

◮ 39. In spherical coordinates (ρ, θ, φ) show that the acceleration components can be represented in terms<br />

of the velocity components as<br />

Hint: Calculate ˙vρ, ˙vθ, ˙vφ.<br />

fρ =˙vρ − v2 θ + v2 φ<br />

, fθ =˙vθ +<br />

ρ<br />

vρvθ<br />

ρ − v2 φ<br />

ρ tan θ , fφ =˙vφ + vρvφ vθvφ<br />

+<br />

ρ ρ tan θ<br />

◮ 40. The divergence of a vector Ai is Ai ,i . That is, perform a contraction on the covariant derivative<br />

Ai ,j to obtain Ai ,i . Calculate the divergence in (a) Cartesian coordinates (b) cylindrical coordinates and (c)<br />

spherical coordinates.<br />

◮ 41. If S is a scalar invariant of weight one and Ai jk is a third order relative tensor of weight W ,show<br />

is an absolute tensor.<br />

that S −W A i jk


◮ 42. Let ¯ Y i ,i =1, 2, 3 denote the components of a field of parallel vectors along the curve C defined by<br />

the equations y i =¯y i (t), i =1, 2, 3inaspacewithmetrictensor¯gij, i, j =1, 2, 3. Assume that ¯ Y i and d¯yi<br />

dt<br />

are unit vectors such that at each point of the curve ¯ C we have<br />

¯gij ¯ i d¯yj<br />

Y<br />

dt<br />

=cosθ = Constant.<br />

(i.e. The field of parallel vectors makes a constant angle θ with the tangent to each point of the curve ¯ C.)<br />

Show that if ¯ Y i and ¯y i (t) undergo a transformation x i = x i (¯y 1 , ¯y 2 , ¯y 3 ), i =1, 2, 3 then the transformed<br />

vector Xm = ¯ i ∂xm Y ∂ ¯y j makes a constant angle with the tangent vector to the transformed curve C given by<br />

xi = xi (¯y 1 (t), ¯y 2 (t), ¯y 3 (t)).<br />

◮ 43. Let J denote the Jacobian determinant | ∂xi<br />

∂x j |. Differentiate J with respect to xm and show that<br />

Hint: See Exercise 1.1, problem 27 and (1.4.7).<br />

p <br />

∂J α ∂x r<br />

= J<br />

− J .<br />

∂xm αp ∂xm rm<br />

◮ 44. Assume that φ is a relative scalar of weight W so that φ = J W φ. Differentiate this relation with<br />

respect to xk . Use the result from problem 42 to obtain the transformation law:<br />

<br />

∂φ α<br />

− W φ = J k ∂x αk<br />

W<br />

m<br />

∂φ r ∂x<br />

− W φ .<br />

∂xm mr<br />

k ∂x<br />

The quantity inside the brackets is called the covariant derivative of a relative scalar of weight W. The<br />

covariant derivative of a relative scalar of weight W is defined as<br />

φ ,k = ∂φ<br />

<br />

r<br />

− W φ<br />

∂xk kr<br />

and this definition has an extra term involving the weight.<br />

It can be shown that similar results hold for relative tensors of weight W. For example, the covariant<br />

derivative of first and second order relative tensors of weight W have the forms<br />

T i ,k<br />

i ∂T i<br />

= +<br />

∂xk T i j,k = ∂T i j<br />

+<br />

∂xk km<br />

i<br />

kσ<br />

<br />

T m <br />

r<br />

− W<br />

<br />

T σ j −<br />

σ<br />

jk<br />

kr<br />

<br />

T i<br />

<br />

T i σ − W<br />

<br />

r<br />

T<br />

kr<br />

i j<br />

When the weight term is zero these covariant derivatives reduce to the results given in our previous definitions.<br />

◮ 45. Let dxi<br />

dt = vi denote a generalized velocity and define the scalar function of kinetic energy T of a<br />

particle with mass m as<br />

T = 1<br />

2 mgij v i v j = 1<br />

2 mgij ˙x i ˙x j .<br />

Show that the intrinsic derivative of T is the same as an ordinary derivative of T. (i.e. Show that δT<br />

δT<br />

= dT<br />

dt .)<br />

<strong>12</strong>7


<strong>12</strong>8<br />

◮ 46. Verify the relations<br />

∂gij<br />

∂xk = −gmj<br />

∂g<br />

gni<br />

nm<br />

∂xk ∂gin ∂xk = −gmn ij ∂gjm<br />

g<br />

∂xk ◮ 47. Assume that Bijk is an absolute tensor. Is the quantity T jk = 1 ∂<br />

√<br />

g ∂xi √ ijk<br />

gB a tensor? Justify<br />

your answer. If your answer is “no”, explain your answer and determine if there any conditions you can<br />

impose upon Bijk such that the above quantity will be a tensor?<br />

◮ 48. The e-permutation symbol can be used to define various vector products. Let Ai,Bi,Ci,Di<br />

i =1,...,N denote vectors, then expand and verify the following products:<br />

(a) In two dimensions<br />

R =eijAiBj a scalar determinant.<br />

(b) In three dimensions<br />

(c) In four dimensions<br />

with similar products in higher dimensions.<br />

◮ 49. Expand the curl operator for:<br />

(a) Two dimensions B = eijAj,i<br />

(b) Three dimensions Bi = eijkAk,j<br />

(c) Four dimensions Bij = eijkmAm,k<br />

Ri =eijAj a vector (rotation).<br />

S =eijkAiBjCk a scalar determinant.<br />

Si =eijkBjCk a vector cross product.<br />

Sij =eijkCk a skew-symmetric matrix<br />

T =eijkmAiBjCkDm a scalar determinant.<br />

Ti =eijkmBjCkDm 4-dimensional cross product.<br />

Tij =eijkmCkDm skew-symmetric matrix.<br />

Tijk =eikmDm skew-symmetric tensor.


§1.5 DIFFERENTIAL GEOMETRY AND RELATIVITY<br />

In this section we will examine some fundamental properties of curves and surfaces. In particular, at<br />

each point of a space curve we can construct a moving coordinate system consisting of a tangent vector, a<br />

normal vector and a binormal vector which is perpendicular to both the tangent and normal vectors. How<br />

these vectors change as we move along the space curve brings up the subjects of curvature and torsion<br />

associated with a space curve. The curvature is a measure of how the tangent vector to the curve is changing<br />

and the torsion is a measure of the twisting of the curve out of a plane. We will find that straight lines have<br />

zero curvature and plane curves have zero torsion.<br />

In a similar fashion, associated with every smooth surface there are two coordinate surface curves and<br />

a normal surface vector through each point on the surface. The coordinate surface curves have tangent<br />

vectors which together with the normal surface vectors create a set of basis vectors. These vectors can be<br />

used to define such things as a two dimensional surface metric and a second order curvature tensor. The<br />

coordinate curves have tangent vectors which together with the surface normal form a coordinate system at<br />

each point of the surface. How these surface vectors change brings into consideration two different curvatures.<br />

A normal curvature and a tangential curvature (geodesic curvature). How these curvatures are related to<br />

the curvature tensor and to the Riemann Christoffel tensor, introduced in the last section, as well as other<br />

interesting relationships between the various surface vectors and curvatures, is the subject area of differential<br />

geometry.<br />

Also presented in this section is a brief introduction to relativity where again the Riemann Christoffel<br />

tensor will occur. Properties of this important tensor are developed in the exercises of this section.<br />

Space Curves and Curvature<br />

For xi = xi (s),i =1, 2, 3, a 3-dimensional space curve in a Riemannian space Vn with metric tensor gij,<br />

and arc length parameter s, the vector T i = dxi<br />

ds represents a tangent vector to the curve at a point P on<br />

the curve. The vector T i is a unit vector because<br />

gijT i T j dx<br />

= gij<br />

i dx<br />

ds<br />

j<br />

ds<br />

=1. (1.5.1)<br />

Differentiate intrinsically, with respect to arc length, the relation (1.5.1) and verify that<br />

which implies that<br />

j<br />

i δT δT<br />

gijT + gij<br />

δs i<br />

δs T j =0, (1.5.2)<br />

i<br />

j δT<br />

gijT =0.<br />

δs<br />

(1.5.3)<br />

i<br />

δT<br />

δs is perpendicular to the tangent vector T i . Define the unit normal vector N i to the<br />

i<br />

δT<br />

δs and write<br />

N i = 1 δT<br />

κ<br />

i<br />

δs<br />

(1.5.4)<br />

Hence, the vector<br />

space curve to be in the same direction as the vector<br />

where κ is a scale factor, called the curvature, and is selected such that<br />

gijN i N j δT<br />

= 1 which implies gij<br />

i δT<br />

δs<br />

j<br />

δs = κ2 . (1.5.5)<br />

<strong>12</strong>9


130<br />

The reciprocal of curvature is called the radius of curvature. The curvature measures the rate of change of<br />

the tangent vector to the curve as the arc length varies. By differentiating intrinsically, with respect to arc<br />

length s, the relation gijT i N j = 0 we find that<br />

Consequently, the curvature κ can be determined from the relation<br />

i δNj<br />

gijT<br />

δs<br />

i δNj δT<br />

gijT + gij<br />

δs i<br />

δs N j =0. (1.5.6)<br />

δT<br />

= −gij<br />

i<br />

δs N j = −gijκN i N j = −κ (1.5.7)<br />

which defines the sign of the curvature. In a similar fashion we differentiate the relation (1.5.5) and find that<br />

i δNj<br />

gijN =0.<br />

δs<br />

(1.5.8)<br />

This later equation indicates that the vector δNj<br />

δs is perpendicular to the unit normal N i . The equation<br />

(1.5.3) indicates that T i is also perpendicular to N i and hence any linear combination of these vectors will<br />

also be perpendicular to N i . The unit binormal vector is defined by selecting the linear combination<br />

δNj δs<br />

and then scaling it into a unit vector by defining<br />

B j = 1<br />

j δN<br />

τ δs<br />

+ κT j<br />

<br />

j<br />

+ κT<br />

(1.5.9)<br />

(1.5.10)<br />

where τ is a scalar called the torsion. The sign of τ is selected such that the vectors T i ,Ni and Bi form a<br />

right handed system with ɛijkT iN jB k = 1 and the magnitude of τ is selected such that Bi is a unit vector<br />

satisfying<br />

gijB i B j =1. (1.5.11)<br />

The triad of vectors T i ,Ni ,Bi at a point on the curve form three planes. The plane containing T i and Bi is<br />

called the rectifying plane. The plane containing N i and Bi is called the normal plane. The plane containing<br />

T i and N i is called the osculating plane. The reciprocal of the torsion is called the radius of torsion. The<br />

torsion measures the rate of change of the osculating plane. The vectors T i ,Ni and Bi form a right-handed<br />

orthogonal system at a point on the space curve and satisfy the relation<br />

B i = ɛ ijk TjNk. (1.5.<strong>12</strong>)<br />

By using the equation (1.5.10) it can be shown that B i is perpendicular to both the vectors T i and N i since<br />

gijB i T j =0 and gijB i N j =0.<br />

It is left as an exercise to show that the binormal vector B i satisfies the relation δBi<br />

δs = −τNi . The three<br />

relations<br />

δT i<br />

i<br />

= κN<br />

δs<br />

δN i<br />

δs = τBi − κT i<br />

δB i<br />

δs<br />

= −τNi<br />

(1.5.13)


are known as the Frenet-Serret formulas of differential geometry.<br />

Surfaces and Curvature<br />

Let us examine surfaces in a Cartesian frame of reference and then later we can generalize our results<br />

to other coordinate systems. A surface in Euclidean 3-dimensional space can be defined in several different<br />

ways. Explicitly, z = f(x, y), implicitly, F (x, y, z) = 0 or parametrically by defining a set of parametric<br />

equations of the form<br />

x = x(u, v), y = y(u, v), z = z(u, v)<br />

which contain two independent parameters u, v called surface coordinates. For example, the equations<br />

x = a sin θ cos φ, y = a sin θ sin φ, z = a cos θ<br />

are the parametric equations which define a spherical surface of radius a with parameters u = θ and v = φ.<br />

See for example figure 1.3-20 in section 1.3. By eliminating the parameters u, v one can derive the implicit<br />

form of the surface and by solving for z one obtains the explicit form of the surface. Using the parametric<br />

form of a surface we can define the position vector to a point on the surface which is then represented in<br />

terms of the parameters u, v as<br />

r = r(u, v) =x(u, v) e1 + y(u, v) e2 + z(u, v) e3. (1.5.14)<br />

The coordinates (u, v) are called the curvilinear coordinates of a point on the surface. The functions<br />

x(u, v),y(u, v),z(u, v) are assumed to be real and differentiable such that ∂r ∂r<br />

∂u × ∂v =0. The curves<br />

r(u, c2) and r(c1,v) (1.5.15)<br />

with c1,c2 constants, then define two surface curves called coordinate curves, which intersect at the surface<br />

coordinates (c1,c2). The family of curves defined by equations (1.5.15) with equally spaced constant values<br />

evaluated at the<br />

surface coordinates (c1,c2) on the surface, are tangent vectors to the coordinate curves through the point<br />

ci,ci +∆ci,ci +2∆ci,... define a surface coordinate grid system. The vectors ∂r ∂r<br />

∂u and ∂v<br />

and are basis vectors for any vector lying in the surface. Letting (x, y, z) =(y 1 ,y 2 ,y 3 )and(u, v) =(u 1 ,u 2 )<br />

and utilizing the summation convention, we can write the position vector in the form<br />

r = r(u 1 ,u 2 )=y i (u 1 ,u 2 ) ei. (1.5.16)<br />

The tangent vectors to the coordinate curves at a point P can then be represented as the basis vectors<br />

Eα = ∂r ∂yi<br />

=<br />

∂uα ∂uα ei, α =1, 2 (1.5.17)<br />

where the partial derivatives are to be evaluated at the point P where the coordinate curves on the surface<br />

intersect. From these basis vectors we construct a unit normal vector to the surface at the point P by<br />

calculating the cross product of the tangent vector ru = ∂r<br />

. A unit normal is then<br />

∂u and rv = ∂r<br />

∂v<br />

n = n(u, v) = E1 × E2<br />

| E1 × E2| = ru × rv<br />

|ru × rv|<br />

(1.5.18)<br />

131


132<br />

and is such that the vectors E1, E2 and n form a right-handed system of coordinates.<br />

If we transform from one set of curvilinear coordinates (u, v) toanotherset(ū, ¯v), which are determined<br />

by a set of transformation laws<br />

u = u(ū, ¯v), v = v(ū, ¯v),<br />

the equation of the surface becomes<br />

r = r(ū, ¯v) =x(u(ū, ¯v),v(ū, ¯v)) e1 + y(u(ū, ¯v),v(ū, ¯v)) e2 + z(u(ū, ¯v),v(ū, ¯v)) e3<br />

and the tangent vectors to the new coordinate curves are<br />

∂r ∂r ∂u ∂r ∂v<br />

= +<br />

∂ū ∂u ∂ū ∂v ∂ū and<br />

∂r ∂r ∂u ∂r ∂v<br />

= +<br />

∂¯v ∂u ∂¯v ∂v ∂¯v .<br />

Using the indicial notation this result can be represented as<br />

∂yi ∂yi<br />

=<br />

∂ūα ∂uβ ∂uβ .<br />

∂ūα This is the transformation law connecting the two systems of basis vectors on the surface.<br />

A curve on the surface is defined by a relation f(u, v) = 0 between the curvilinear coordinates. Another<br />

way to represent a curve on the surface is to represent it in a parametric form where u = u(t) andv = v(t),<br />

where t is a parameter. The vector<br />

dr ∂r du ∂r dv<br />

= +<br />

dt ∂u dt ∂v dt<br />

is tangent to the curve on the surface.<br />

An element of arc length with respect to the surface coordinates is represented by<br />

ds 2 = dr · dr = ∂r ∂r<br />

·<br />

∂uα ∂uβ duα du β = aαβdu α du β<br />

(1.5.19)<br />

where aαβ = ∂r<br />

∂uα · ∂r<br />

∂uβ with α, β =1, 2 defines a surface metric. This element of arc length on the surface is<br />

often written as the quadratic form<br />

A = ds 2 = E(du) 2 +2Fdudv+ G(dv) 2 = 1<br />

E (Edu+ Fdv)2 2 EG − F<br />

+ dv<br />

E<br />

2<br />

(1.5.20)<br />

and called the first fundamental form of the surface. Observe that for ds2 to be positive definite the quantities<br />

E and EG − F 2 must be positive.<br />

The surface metric associated with the two dimensional surface is defined by<br />

aαβ = Eα · Eβ = ∂r ∂r ∂yi<br />

· =<br />

∂uα ∂uβ ∂uα ∂yi , α,β =1, 2 (1.5.21)<br />

∂uβ with conjugate metric tensor a αβ defined such that a αβ aβγ = δ α γ . Here the surface is embedded in a three<br />

dimensional space with metric gij and aαβ is the two dimensional surface metric. In the equation (1.5.20)<br />

the quantities E,F,G are functions of the surface coordinates u, v and are determined from the relations<br />

E =a11 = ∂r ∂r ∂yi<br />

· =<br />

∂u ∂u ∂u1 ∂yi ∂u1 F =a<strong>12</strong> = ∂r ∂r ∂yi<br />

· =<br />

∂u ∂v ∂u1 G =a22 = ∂r ∂r ∂yi<br />

· =<br />

∂v ∂v ∂u2 ∂yi ∂u2 ∂y i<br />

∂u 2<br />

(1.5.22)


Here and throughout the remainder of this section, we adopt the convention that Greek letters have the<br />

range 1,2, while Latin letters have the range 1,2,3.<br />

Construct at a general point P on the surface the unit normal vector n at this point. Also construct a<br />

plane which contains this unit surface normal vector n. Observe that there are an infinite number of planes<br />

which contain this unit surface normal. For now, select one of these planes, then later on we will consider<br />

all such planes. Let r = r(s) denote the position vector defining a curve C which is the intersection of the<br />

selected plane with the surface, where s is the arc length along the curve, which is measured from some fixed<br />

point on the curve. Let us find the curvature of this curve of intersection. The vector T = dr<br />

ds , evaluated<br />

at the point P, is a unit tangent vector to the curve C and lies in the tangent plane to the surface at the<br />

point P. Here we are using ordinary differentiation rather than intrinsic differentiation because we are in<br />

a Cartesian system of coordinates. Differentiating the relation T · T =1, with respect to arc length s we<br />

find that T · dT<br />

dT<br />

ds = 0 which implies that the vector ds is perpendicular to the tangent vector T. Since the<br />

coordinate system is Cartesian we can treat the curve of intersection C as a space curve, then the vector<br />

K = dT<br />

ds , evaluated at point P, is defined as the curvature vector with curvature | K| = κ and radius of<br />

curvature R =1/κ. A unit normal N to the space curve is taken in the same direction as dT<br />

ds so that the<br />

curvature will always be positive. We can then write K = κ N = d T<br />

. Consider the geometry of figure 1.5-1<br />

ds<br />

and define on the surface a unit vector u = n × T which is perpendicular to both the surface tangent vector<br />

T and the surface normal vector n, such that the vectors T i ,ui and ni forms a right-handed system.<br />

Figure 1.5-1 Surface curve with tangent plane and a normal plane.<br />

133


134<br />

The direction of u in relation to T is in the same sense as the surface tangents E1 and E2. Note that<br />

the vector dT<br />

ds is perpendicular to the tangent vector T and lies in the plane which contains the vectors n<br />

and u. We can therefore write the curvature vector K in the component form<br />

K = d T<br />

ds = κ (n) n + κ (g) u = Kn + Kg<br />

(1.5.23)<br />

where κ (n) is called the normal curvature and κ (g) is called the geodesic curvature. The subscripts are not<br />

indices. These curvatures can be calculated as follows. From the orthogonality condition n · T =0weobtain<br />

by differentiation with respect to arc length s the result n · d T<br />

ds + T · dn<br />

=0. Consequently, the normal<br />

ds<br />

curvature is determined from the dot product relation<br />

n · K = κ (n) = − T · dn<br />

ds<br />

= − dr<br />

ds<br />

dn<br />

· . (1.5.24)<br />

ds<br />

By taking the dot product of u with equation (1.5.23) we find that the geodesic curvature is determined<br />

from the triple scalar product relation<br />

Normal Curvature<br />

κ (g) = u · d T<br />

ds =(n × T ) · d T<br />

. (1.5.25)<br />

ds<br />

The equation (1.5.24) can be expressed in terms of a quadratic form by writing<br />

κ (n) ds 2 = −dr · dn. (1.5.26)<br />

The unit normal to the surface n and position vector r are functions of the surface coordinates u, v with<br />

dr = ∂r<br />

∂u<br />

du + ∂r<br />

∂v<br />

dv and dn = ∂n<br />

∂u<br />

We define the quadratic form<br />

<br />

∂r ∂r<br />

B = −dr · dn = − du +<br />

∂u ∂v dv<br />

<br />

∂n ∂n<br />

· du +<br />

∂u ∂v dv<br />

<br />

where<br />

e = − ∂r<br />

∂u<br />

B = e(du) 2 +2fdudv + g(dv) 2 = bαβ du α du β<br />

<br />

∂n<br />

∂r ∂n ∂n ∂r<br />

· , 2f = − · + · , g = −<br />

∂u ∂u ∂v ∂u ∂v<br />

∂r ∂n<br />

·<br />

∂v ∂v<br />

and bαβ α, β =1, 2 is called the curvature tensor and a αγ bαβ = b γ<br />

β<br />

∂n<br />

du + dv. (1.5.27)<br />

∂v<br />

(1.5.28)<br />

(1.5.29)<br />

is an associated curvature tensor.<br />

The quadratic form of equation (1.5.28) is called the second fundamental form of the surface. Alternative<br />

methods for calculating the coefficients of this quadratic form result from the following considerations. The<br />

unit surface normal is perpendicular to the tangent vectors to the coordinate curves at the point P and<br />

therefore we have the orthogonality relationships<br />

∂r<br />

· n =0 and<br />

∂u<br />

∂r<br />

· n =0. (1.5.30)<br />

∂v


Observe that by differentiating the relations in equation (1.5.30), with respect to both u and v, one can<br />

derive the results<br />

e = ∂2r ∂r ∂n<br />

· n = − · = b11<br />

∂u2 ∂u ∂u<br />

f = ∂2r ∂r ∂n ∂n ∂r<br />

· n = − · = − ·<br />

∂u∂v ∂u ∂v ∂u ∂v = b21 = b<strong>12</strong><br />

g = ∂2 (1.5.31)<br />

r ∂r ∂n<br />

· n = − · = b22<br />

∂v2 ∂v ∂v<br />

and consequently the curvature tensor can be expressed as<br />

bαβ = − ∂r ∂n<br />

· . (1.5.32)<br />

∂uα ∂uβ The quadratic forms from equations (1.5.20) and (1.5.28) enable us to represent the normal curvature<br />

in the form of a ratio of quadratic forms. We find from equation (1.5.26) that the normal curvature in the<br />

direction du<br />

dv is<br />

κ (n) = B<br />

A = e(du)2 +2fdudv+ g(dv) 2<br />

E(du) 2 .<br />

+2Fdudv+ G(dv) 2 (1.5.33)<br />

If we write the unit tangent vector to the curve in the form T = dr ∂r<br />

ds = ∂uα du α<br />

ds and express the derivative<br />

of the unit surface normal with respect to arc length as dn ∂n<br />

ds = ∂uβ du β<br />

ds , then the normal curvature can be<br />

expressed in the form<br />

κ (n) = − T · dn<br />

<br />

∂r ∂n<br />

= − ·<br />

ds ∂uα ∂uβ α du du<br />

ds<br />

β<br />

ds<br />

(1.5.34)<br />

= bαβdu α du β<br />

ds 2<br />

= bαβduαduβ aαβduα .<br />

duβ Observe that the curvature tensor is a second order symmetric tensor.<br />

In the previous discussions, the plane containing the unit normal vector was arbitrary. Let us now<br />

consider all such planes that pass through this unit surface normal. As we vary the plane containing the unit<br />

surface normal n at P we get different curves of intersection with the surface. Each curve has a curvature<br />

associated with it. By examining all such planes we can find the maximum and minimum normal curvatures<br />

associated with the surface. We write equation (1.5.33) in the form<br />

κ (n) =<br />

e +2fλ+ gλ2<br />

E +2Fλ+ Gλ 2<br />

where λ = dv<br />

du . From the theory of proportions we can also write this equation in the form<br />

κ (n) =<br />

(1.5.35)<br />

(e + fλ)+λ(f + gλ) f + gλ e + fλ<br />

= = . (1.5.36)<br />

(E + Fλ)+λ(F + Gλ) F + Gλ E + Fλ<br />

Consequently, the curvature κ will satisfy the differential equations<br />

(e − κE)du +(f − κF )dv =0 and (f − κF )du +(g − κG)dv =0. (1.5.37)<br />

The maximum and minimum curvatures occur in those directions λ where dκ (n)<br />

dλ =0. Calculating the derivative<br />

of κ (n) with respect to λ and setting the derivative to zero we obtain a quadratic equation in λ<br />

(Fg− Gf)λ 2 +(Eg − Ge)λ +(Ef − Fe)=0, (Fg − Gf) = 0.<br />

135


136<br />

This equation has two roots λ1 and λ2 which satisfy<br />

Eg − Ge<br />

λ1 + λ2 = −<br />

Fg− Gf<br />

and λ1λ2 =<br />

Ef − Fe<br />

, (1.5.38)<br />

Fg− Gf<br />

where Fg − Gf = 0. The curvatures κ (1),κ (2) corresponding to the roots λ1 and λ2 are called the principal<br />

curvatures at the point P. Several quantities of interest that are related to κ (1) and κ (2) are: (1) the principal<br />

radii of curvature Ri =1/κi,i =1, 2; (2) H = 1<br />

2 (κ (1) + κ (2)) called the mean curvature and K = κ (1)κ (2)<br />

called the total curvature or Gaussian curvature of the surface. Observe that the roots λ1 and λ2 determine<br />

two directions on the surface<br />

dr1<br />

du<br />

∂r ∂r<br />

= +<br />

∂u ∂v λ1<br />

If these directions are orthogonal we will have<br />

This requires that<br />

dr1<br />

du<br />

· dr2<br />

du =(∂r<br />

∂u<br />

and<br />

dr2<br />

du<br />

∂r ∂r<br />

= +<br />

∂u ∂v λ2.<br />

∂r ∂r ∂r<br />

+ λ1)( +<br />

∂v ∂u ∂v λ2) =0.<br />

Gλ1λ2 + F (λ1 + λ2)+E =0. (1.5.39)<br />

It is left as an exercise to verify that this is indeed the case and so the directions determined by the principal<br />

curvatures must be orthogonal. In the case where Fg− Gf =0wehavethatF =0andf = 0 because the<br />

coordinate curves are orthogonal and G must be positive. In this special case there are still two directions<br />

determined by the differential equations (1.5.37) with dv =0,du arbitrary, and du =0,dv arbitrary. From<br />

the differential equations (1.5.37) we find these directions correspond to<br />

κ (1) = e<br />

E<br />

and κ (2) = g<br />

G .<br />

Weletλ α = duα<br />

ds denote a unit vector on the surface satisfying aαβλ α λ β =1. Then the equation (1.5.34)<br />

can be written as κ (n) = bαβλ α λ β or we can write (bαβ − κ (n) aαβ)λ α λ β =0. The maximum and minimum<br />

normal curvature occurs in those directions λ α where<br />

(bαβ − κ (n) aαβ)λ α =0<br />

and so κ (n) must be a root of the determinant equation |bαβ − κ (n) aαβ| =0or<br />

|a αγ bαβ − κ (n)δ γ<br />

β | =<br />

<br />

<br />

<br />

b11 − κ (n)<br />

b 2 1<br />

b1 2<br />

b2 2 − κ (n)<br />

<br />

<br />

<br />

= κ2 (n) − bαβa αβ κ (n) + b<br />

=0. (1.5.40)<br />

a<br />

This is a quadratic equation in κ (n) of the form κ2 (n) − (κ (1) + κ (2))κ (n) + κ (1)κ (2) =0. In other words the<br />

principal curvatures κ (1) and κ (2) are the eigenvalues of the matrix with elements b γ<br />

β = aαγbαβ. Observe that<br />

from the determinant equation in κ (n) we can directly find the total curvature or Gaussian curvature which<br />

is an invariant given by K = κ (1)κ (2) = |bα β | = |aαγbγβ| = b/a. The mean curvature is also an invariant<br />

obtained from H = 1<br />

2 (κ (1) + κ (2)) = 1<br />

2aαβbαβ, where a = a11a22 − a<strong>12</strong>a21 and b = b11b22 − b<strong>12</strong>b21 are the<br />

determinants formed from the surface metric tensor and curvature tensor components.


The equations of Gauss, Weingarten and Codazzi<br />

At each point on a space curve we can construct a unit tangent T , a unit normal N and unit binormal<br />

B. The derivatives of these vectors, with respect to arc length, can also be represented as linear combinations<br />

of the base vectors T, N, B. See for example the Frenet-Serret formulas from equations (1.5.13). In a similar<br />

fashion the surface vectors ru,rv, n form a basis and the derivatives of these basis vectors with respect to<br />

the surface coordinates u, v can also be expressed as linear combinations of the basis vectors ru,rv, n. For<br />

example, the derivatives ruu,ruv,rvv can be expressed as linear combinations of ru,rv, n. Wecanwrite<br />

ruu = c1ru + c2rv + c3n<br />

ruv = c4ru + c5rv + c6n<br />

rvv = c7ru + c8rv + c9n<br />

(1.5.41)<br />

where c1,...,c9 are constants to be determined. It is an easy exercise (see exercise 1.5, problem 8) to show<br />

that these equations can be written in the indicial notation as<br />

∂2r ∂uα <br />

γ ∂r<br />

=<br />

∂uβ αβ ∂uγ + bαβn.<br />

These equations are known as the Gauss equations.<br />

(1.5.42)<br />

In a similar fashion the derivatives of the normal vector can be represented as linear combinations of<br />

the surface basis vectors. If we write<br />

∂n<br />

∂u = c1ru + c2rv<br />

∂n<br />

∂v = c3ru + c4rv<br />

or<br />

∂r<br />

∂u = c∗ ∂n<br />

1<br />

∂u + c∗ ∂n<br />

2<br />

∂v<br />

∂r<br />

∂v = c∗ ∂n<br />

3<br />

∂u + c∗ ∂n<br />

4<br />

∂v<br />

(1.5.43)<br />

where c1,...,c4 and c∗ 1 ,...,c∗4 are constants. These equations are known as the Weingarten equations. It<br />

is easily demonstrated (see exercise 1.5, problem 9) that the Weingarten equations can be written in the<br />

indicial form<br />

∂n<br />

∂uα = −bβ ∂r<br />

α<br />

∂uβ (1.5.44)<br />

where b β α = a βγ bγα is the mixed second order form of the curvature tensor.<br />

The equations of Gauss produce a system of partial differential equations defining the surface coordinates<br />

xi as a function of the curvilinear coordinates u and v. The equations are not independent as certain<br />

compatibility conditions must be satisfied. In particular, it is required that the mixed partial derivatives<br />

must satisfy<br />

Wecalculate<br />

∂3r ∂uα∂uβ =<br />

∂uδ ∂3r ∂uα∂uβ 2 γ ∂ r<br />

=<br />

∂uδ αβ ∂uγ ∂<br />

+<br />

∂uδ ∂3r ∂uα∂uδ .<br />

∂uβ <br />

γ<br />

αβ<br />

∂uδ ∂r ∂n ∂bαβ<br />

+ bαβ + n<br />

∂uγ ∂uδ ∂uδ and use the equations of Gauss and Weingarten to express this derivative in the form<br />

∂3r ∂uα∂uβ ⎡ <br />

ω<br />

⎢<br />

∂<br />

αβ<br />

= ⎢<br />

∂uδ ⎣ ∂uδ +<br />

<br />

γ ω<br />

− bαβb<br />

αβ γδ<br />

ω ⎤<br />

<br />

⎥<br />

∂r γ<br />

δ ⎦ + bγδ +<br />

∂uω αβ<br />

∂bαβ<br />

∂uδ <br />

n.<br />

137


138<br />

Forming the difference<br />

∂3r ∂uα∂uβ ∂<br />

−<br />

∂uδ 3r ∂uα∂uδ =0<br />

∂uβ we find that the coefficients of the independent vectors n and ∂r<br />

∂uω equal to zero produces the Codazzi equations<br />

must be zero. Setting the coefficient of n<br />

<br />

γ<br />

γ<br />

bγδ − bγβ +<br />

αβ αδ<br />

∂bαβ<br />

∂uδ ∂bαδ<br />

− =0.<br />

∂uβ (1.5.45)<br />

These equations are sometimes referred to as the Mainardi-Codazzi equations. Equating to zero the coefficient<br />

of ∂r<br />

∂uω we find that R δ αγβ = bαβb δ γ − bαγb δ β or changing indices we have the covariant form<br />

aωδR δ αβγ = Rωαβγ = bωβbαγ − bωγbαβ, (1.5.46)<br />

where<br />

R δ ∂<br />

αγβ =<br />

∂uγ <br />

δ<br />

−<br />

αβ<br />

∂<br />

∂uβ <br />

δ ω δ ω δ<br />

+<br />

−<br />

αγ αβ ωγ αγ ωβ<br />

is the mixed Riemann curvature tensor.<br />

EXAMPLE 1.5-1<br />

(1.5.47)<br />

Show that the Gaussian or total curvature K = κ (1)κ (2) depends only upon the metric aαβ and is<br />

K = R<strong>12</strong><strong>12</strong><br />

where a = det(aαβ).<br />

a<br />

Solution:<br />

Utilizing the two-dimensional alternating tensor eαβ and the property of determinants we can write<br />

e γδ K = e αβ b γ α bδ β<br />

ζ and δ to obtain<br />

where from page 137, K = |bγ<br />

β | = |aαγ bαβ|. Now multiply by eγζ and then contract on<br />

eγδe γδ K = eγδe αβ b γ α bδ β =2K<br />

2K = eγδe αβ (a γµ bαu) a δν <br />

bβν<br />

But eγδa γµ a δν = ae µν so that 2K = e αβ ae µν bαµbβν. Using √ ae µν = ɛ µν we have 2K = ɛ µν ɛ αβ bαµbβν.<br />

Interchanging indices we can write<br />

2K = ɛ βγ ɛ ωα bωβbαγ and 2K = ɛ γβ ɛ ωα bωγbαβ.<br />

Adding these last two results we find that 4K = ɛ βγ ɛ ωγ (bωβbαγ − bωγbαβ) =ɛ βγ ɛ ωγ Rωαβγ. Now multiply<br />

both sides by ɛστɛλν to obtain 4Kɛστɛλν = δ βγ<br />

στ δ ωα<br />

λν Rωαβγ. From exercise 1.5, problem 16, the Riemann<br />

curvature tensor Rijkl is skew symmetric in the (i, j), (k, l) as well as being symmetric in the (ij), (kl) pair<br />

of indices. Consequently, δ βγ<br />

στ δ ωα<br />

λν Rωαβγ =4Rλνστ and hence Rλνστ = Kɛστɛλν and we have the special case<br />

where K √ √<br />

ae<strong>12</strong> ae<strong>12</strong> = R<strong>12</strong><strong>12</strong> or K = R<strong>12</strong><strong>12</strong><br />

b<br />

. A much simpler way to obtain this result is to observe K =<br />

a a<br />

(bottom of page 137) and note from equation (1.5.46) that R<strong>12</strong><strong>12</strong> = b11b22 − b<strong>12</strong>b21 = b.<br />

Note that on a surface ds2 = aαβduαduβ where aαβ are the metrices for the surface. This metric is a<br />

∂u<br />

tensor and satisfies āγδ = aαβ<br />

α<br />

∂ūγ ∂uβ and by taking determinants we find<br />

∂ūδ ā =<br />

<br />

<br />

āγδ<br />

<br />

<br />

<br />

∂uα<br />

∂ūγ <br />

<br />

<br />

∂uβ<br />

∂ūδ <br />

<br />

= aJ 2


where J is the Jacobian of the surface coordinate transformation. Here the curvature tensor for the surface<br />

Rαβγδ has only one independent component since R<strong>12</strong><strong>12</strong> = R2<strong>12</strong>1 = −R<strong>12</strong>21 = −R21<strong>12</strong> (See exercises 20,21).<br />

From the transformation law<br />

∂u<br />

¯Rɛηλµ = Rαβγδ<br />

α<br />

∂ūɛ ∂uβ ∂ūη ∂uγ ∂ūλ ∂uδ ∂ū µ<br />

one can sum over the repeated indices and show that ¯ R<strong>12</strong><strong>12</strong> = R<strong>12</strong><strong>12</strong>J 2 and consequently<br />

¯R<strong>12</strong><strong>12</strong><br />

ā<br />

= R<strong>12</strong><strong>12</strong><br />

a<br />

= K<br />

which shows that the Gaussian curvature is a scalar invariant in V2.<br />

Geodesic Curvature<br />

For C an arbitrary curve on a given surface the curvature vector K, associated with this curve, is<br />

the vector sum of the normal curvature κ (n) n and geodesic curvature κ (g) u and lies in a plane which<br />

is perpendicular to the tangent vector to the given curve on the surface. The geodesic curvature κ (g) is<br />

obtained from the equation (1.5.25) and can be represented<br />

Substituting into this expression the vectors<br />

κ (g) = u · K = u · d T<br />

ds =(n × T ) · d T<br />

ds =<br />

<br />

T × d <br />

T<br />

· n.<br />

ds<br />

T = dr<br />

ds<br />

du<br />

= ru<br />

ds<br />

dv<br />

+ rv<br />

ds<br />

d T<br />

ds = K = ruu(u ′ ) 2 +2ruvu ′ v ′ + rvv(v ′ ) 2 + ruu ′′ + rvv ′′ ,<br />

where ′ = d<br />

ds , and by utilizing the results from problem 10 of the exercises following this section, we find<br />

that the geodesic curvature can be represented as<br />

<br />

2<br />

κ (g) = (u<br />

11<br />

′ ) 3 <br />

2 1<br />

+ 2 −<br />

<strong>12</strong> 11<br />

<br />

<br />

2<br />

1<br />

− 2<br />

22<br />

22<br />

<br />

1<br />

u<br />

<strong>12</strong><br />

′ (v ′ ) 2 −<br />

(u ′ ) 2 v ′ +<br />

<br />

(v ′ ) 3 +(u ′ v ′′ − u ′′ v ′ <br />

EG<br />

) − F 2 .<br />

(1.5.48)<br />

This equation indicates that the geodesic curvature is only a function of the surface metrices E,F,G and<br />

the derivatives u ′ ,v ′ ,u ′′ ,v ′′ . When the geodesic curvature is zero the curve is called a geodesic curve. Such<br />

curves are often times, but not always, the lines of shortest distance between two points on a surface. For<br />

example, the great circle on a sphere which passes through two given points on the sphere is a geodesic curve.<br />

If you erase that part of the circle which represents the shortest distance between two points on the circle<br />

you are left with a geodesic curve connecting the two points, however, the path is not the shortest distance<br />

between the two points.<br />

For plane curves we let u = x and v = y so that the geodesic curvature reduces to<br />

kg = u ′ v ′′ − u ′′ v ′ = dφ<br />

ds<br />

139


140<br />

where φ is the angle between the tangent T to the curve and the unit vector e1.<br />

Geodesics are curves on the surface where the geodesic curvature is zero. Since kg = 0 along a geodesic<br />

surface curve, then at every point on this surface curve the normal N to the curve will be in the same<br />

direction as the normal n to the surface. In this case, we have ru · n =0andrv · n = 0 which reduces to<br />

since the vectors n and d T<br />

ds<br />

d T<br />

ds · ru =0 and d T<br />

ds · rv =0, (1.5.49)<br />

have the same direction. In particular, we may write<br />

∂r du ∂r dv<br />

= +<br />

∂u ds ∂v ds = ru u ′ + rv v ′<br />

d T<br />

ds = ruu (u ′ ) 2 +2ruv u ′ v ′ + rvv (v ′ ) 2 + ru u ′′ + rv v ′′<br />

T = dr<br />

ds<br />

Consequently, the equations (1.5.49) become<br />

d T<br />

ds · ru =(ruu · ru)(u ′ ) 2 +2(ruv · ru) u ′ v ′ +(rvv · ru)(v ′ ) 2 + Eu ′′ + Fv ′′ =0<br />

d T<br />

ds · rv =(ruu · rv)(u ′ ) 2 +2(ruv · rv) u ′ v ′ +(rvv · rv)(v ′ ) 2 + Fu ′′ + Gv ′′ . (1.5.50)<br />

=0.<br />

Utilizing the results from exercise 1.5,(See problems 4,5 and 6), we can eliminate v ′′ from the equations<br />

(1.5.50) to obtain<br />

d2 2 <br />

u 1 du 1 du dv<br />

+<br />

+2<br />

ds2 11 ds <strong>12</strong> ds ds +<br />

2 1 dv<br />

=0<br />

22 ds<br />

and eliminating u ′′ from the equations (1.5.50) produces the equation<br />

d2 2 <br />

v 2 du 2 du dv<br />

+<br />

+2<br />

ds2 11 ds <strong>12</strong> ds ds +<br />

2 2 dv<br />

=0.<br />

22 ds<br />

In tensor form, these last two equations are written<br />

d2uα <br />

α du<br />

+<br />

ds2 βγ a<br />

β<br />

ds<br />

du γ<br />

ds<br />

=0, α,β,γ =1, 2 (1.5.51)<br />

where u = u 1 and v = u 2 . The equations (1.5.51) are the differential equations defining a geodesic curve on<br />

a surface. We will find that these same type of equations arise in considering the shortest distance between<br />

two points in a generalized coordinate system. See for example problem 18 in exercise 2.2.


Tensor Derivatives<br />

Let uα = uα (t) denote the parametric equations of a curve on the surface defined by the parametric<br />

equations xi = xi (u1 ,u2 ). We can then represent the surface curve in the spatial geometry since the surface<br />

curve can be represented in the spatial coordinates through the representation x i = x i (u 1 (t),u 2 (t)) = x i (t).<br />

Recall that for xi = xi (t) a given curve C , the intrinsic derivative of a vector field Ai along C is defined as<br />

the inner product of the covariant derivative of the vector field with the tangent vector to the curve. This<br />

intrinsic derivative is written<br />

δAi δt = Ai dx<br />

,j<br />

j<br />

dt =<br />

<br />

∂Ai <br />

i<br />

+ A<br />

∂xj jk g<br />

k<br />

<br />

dxj dt<br />

or<br />

δA i<br />

δt<br />

dAi<br />

=<br />

dt +<br />

<br />

i k dxj<br />

A<br />

jk g dt<br />

where the subscript g indicates that the Christoffel symbol is formed from the spatial metric gij. If Aα is a<br />

surface vector defined along the curve C, the intrinsic derivative is represented<br />

δAα δt = Aα du<br />

,β<br />

β<br />

dt =<br />

<br />

α ∂A α<br />

+ A<br />

∂uβ βγ a<br />

γ<br />

β du<br />

dt<br />

or<br />

δAα dAα<br />

=<br />

δt dt +<br />

<br />

α γ duβ<br />

A<br />

βγ a dt<br />

where the subscript a denotes that the Christoffel is formed from the surface metric aαβ.<br />

Similarly, the formulas for the intrinsic derivative of a covariant spatial vector Ai or covariant surface<br />

vector Aα are given by<br />

δAi dAi<br />

=<br />

δt dt −<br />

<br />

k dx<br />

Ak<br />

ij g<br />

j<br />

dt<br />

and<br />

δAα dAα<br />

=<br />

δt dt −<br />

<br />

γ du<br />

Aα<br />

αβ a<br />

β<br />

dt .<br />

Consider a mixed tensor T i α which is contravariant with respect to a transformation of space coordinates<br />

xi and covariant with respect to a transformation of surface coordinates uα . For T i α defined over the surface<br />

curve C, which can also be viewed as a space curve C, define the scalar invariant Ψ = Ψ(t) =T i αAiB α where<br />

Ai is a parallel vector field along the curve C when it is viewed as a space curve and Bα is also a parallel<br />

vector field along the curve C when it is viewed as a surface curve. Recall that these parallel vector fields<br />

must satisfy the differential equations<br />

δAi<br />

δt<br />

dAi<br />

=<br />

dt −<br />

<br />

k dx<br />

Ak<br />

ij g<br />

j<br />

dt<br />

=0 and<br />

δB α<br />

δt<br />

dBα<br />

=<br />

dt +<br />

<br />

α γ duβ<br />

B<br />

βγ a dt<br />

=0. (1.5.52)<br />

The scalar invariant Ψ is a function of the parameter t of the space curve since both the tensor and the<br />

parallel vector fields are to be evaluated along the curve C. By differentiating the function Ψ with respect<br />

to the parameter t there results<br />

dΨ<br />

dt = dT i α<br />

dt AiB α + T i dAi<br />

α<br />

dt Bα + T i αAi dBα . (1.5.53)<br />

dt<br />

141


142<br />

But the vectors Ai and Bα are parallel vector fields and must satisfy the relations given by equations (1.5.52).<br />

This implies that equation (1.5.53) can be written in the form<br />

dΨ<br />

dt =<br />

<br />

dT i α<br />

dt +<br />

<br />

i<br />

T<br />

kj g<br />

k dx<br />

α<br />

j<br />

dt −<br />

<br />

γ<br />

T<br />

βα a<br />

i du<br />

γ<br />

β<br />

<br />

AiB<br />

dt<br />

α . (1.5.54)<br />

The quantity inside the brackets of equation (1.5.54) is defined as the intrinsic tensor derivative with respect<br />

to the parameter t along the curve C. This intrinsic tensor derivative is written<br />

δT i α<br />

dt = dT i α<br />

dt +<br />

<br />

i<br />

T<br />

kj g<br />

k dx<br />

α<br />

j<br />

dt −<br />

<br />

γ<br />

T<br />

βα a<br />

i du<br />

γ<br />

β<br />

. (1.5.55)<br />

dt<br />

The spatial representation of the curve C is related to the surface representation of the curve C through the<br />

defining equations. Therefore, we can express the equation (1.5.55) in the form<br />

δT i α<br />

dt =<br />

<br />

∂T i <br />

α i<br />

+ T<br />

∂uβ kj g<br />

k ∂x<br />

α<br />

j <br />

γ<br />

− T<br />

∂uβ βα a<br />

i <br />

du<br />

γ<br />

β<br />

(1.5.56)<br />

dt<br />

The quantity inside the brackets is a mixed tensor which is defined as the tensor derivative of T i α with<br />

respect to the surface coordinates uβ . The tensor derivative of the mixed tensor T i α with respect to the<br />

surface coordinates u β is written<br />

T i α,β = ∂T i α<br />

+<br />

∂uβ In general, given a mixed tensor T i...j<br />

α...β<br />

<br />

i<br />

<br />

kj g<br />

T k α<br />

∂x j<br />

−<br />

∂uβ <br />

γ<br />

T<br />

βα a<br />

i γ.<br />

which is contravariant with respect to transformations of the<br />

space coordinates and covariant with respect to transformations of the surface coordinates, then we can<br />

define the scalar field along the surface curve C as<br />

Ψ(t) =T i...j<br />

α...β Ai ···AjB α ···B β<br />

(1.5.57)<br />

where Ai,...,Aj and B α ,...,B β are parallel vector fields along the curve C. The intrinsic tensor derivative<br />

is then derived by differentiating the equation (1.5.57) with respect to the parameter t.<br />

Tensor derivatives of the metric tensors gij,aαβ and the alternating tensors ɛijk,ɛαβ and their associated<br />

tensors are all zero. Hence, they can be treated as constants during the tensor differentiation process.<br />

Generalizations<br />

In a Riemannian space Vn with metric gij and curvilinear coordinates x i ,i =1, 2, 3, the equations of a<br />

surface can be written in the parametric form x i = x i (u 1 ,u 2 )whereu α ,α =1, 2 are called the curvilinear<br />

coordinates of the surface. Since<br />

dx i = ∂xi<br />

duα<br />

(1.5.58)<br />

∂uα then a small change duα on the surface results in change dxi in the space coordinates. Hence an element of<br />

arc length on the surface can be represented in terms of the curvilinear coordinates of the surface. This same<br />

element of arc length can also be represented in terms of the curvilinear coordinates of the space. Thus, an<br />

element of arc length squared in terms of the surface coordinates is represented<br />

ds 2 = aαβdu α du β<br />

(1.5.59)


where aαβ is the metric of the surface. This same element when viewed as a spatial element is represented<br />

By equating the equations (1.5.59) and (1.5.60) we find that<br />

gijdx i dx j = gij<br />

ds 2 = gijdx i dx j . (1.5.60)<br />

∂xi ∂uα ∂x j<br />

∂u β duα du β = aαβdu α du β . (1.5.61)<br />

The equation (1.5.61) shows that the surface metric is related to the spatial metric and can be calculated<br />

∂x<br />

from the relation aαβ = gij<br />

i<br />

∂uα ∂xj . This equation reduces to the equation (1.5.21) in the special case of<br />

∂uβ Cartesian coordinates. In the surface coordinates we define the quadratic form A = aαβdu α du β as the first<br />

fundamental form of the surface. The tangent vector to the coordinate curves defining the surface are given<br />

by ∂xi<br />

∂u α and can be viewed as either a covariant surface vector or a contravariant spatial vector. We define<br />

this vector as<br />

x i α = ∂xi<br />

, i =1, 2, 3, α =1, 2. (1.5.62)<br />

∂uα Any vector which is a linear combination of the tangent vectors to the coordinate curves is called a surface<br />

vector. A surface vector A α can also be viewed as a spatial vector A i . The relation between the spatial<br />

representation and surface representation is Ai = Aαxi α. The surface representation Aα ,α =1, 2andthe<br />

spatial representation Ai ,i=1, 2, 3 define the same direction and magnitude since<br />

gijA i A j = gijA α x i αA β x j<br />

β = gijx i αx j<br />

β Aα A β = aαβA α A β .<br />

Consider any two surface vectors A α and B α and their spatial representations A i and B i where<br />

A i = A α x i α and B i = B α x i α. (1.5.63)<br />

These vectors are tangent to the surface and so a unit normal vector to the surface can be defined from the<br />

cross product relation<br />

niAB sin θ = ɛijkA j B k<br />

(1.5.64)<br />

where A, B are the magnitudes of A i ,B i and θ is the angle between the vectors when their origins are made<br />

to coincide. Substituting equations (1.5.63) into the equation (1.5.64) we find<br />

niAB sin θ = ɛijkA α x j αBβ x k β . (1.5.65)<br />

In terms of the surface metric we have AB sin θ = ɛαβA α B β so that equation (1.5.65) can be written in the<br />

form<br />

which for arbitrary surface vectors implies<br />

(niɛαβ − ɛijkx j αx k β)A α B β =0 (1.5.66)<br />

niɛαβ = ɛijkx j αxkβ or ni = 1<br />

2 ɛαβɛijkx j αxkβ . (1.5.67)<br />

The equation (1.5.67) defines a unit normal vector to the surface in terms of the tangent vectors to the<br />

coordinate curves. This unit normal vector is related to the covariant derivative of the surface tangents as<br />

143


144<br />

is now demonstrated. By using the results from equation (1.5.50), the tensor derivative of equation (1.5.59),<br />

with respect to the surface coordinates, produces<br />

x i α,β = ∂2xi ∂uα <br />

i<br />

+ x<br />

∂uβ pq g<br />

p αx q<br />

β −<br />

<br />

σ<br />

x<br />

αβ a<br />

i σ<br />

(1.5.68)<br />

where the subscripts on the Christoffel symbols refer to the metric from which they are calculated. Also the<br />

tensor derivative of the equation (1.5.57) produces the result<br />

gijx i α,γx j<br />

β + gijx i αx j<br />

β,γ = aαβ,γ =0. (1.5.69)<br />

Interchanging the indices α, β, γ cyclically in the equation (1.5.69) one can verify that<br />

gijx i α,βx j γ =0. (1.5.70)<br />

The equation (1.5.70) indicates that in terms of the space coordinates, the vector xi α,β is perpendicular to<br />

the surface tangent vector xi γ and so must have the same direction as the unit surface normal ni . Therefore,<br />

there must exist a second order tensor bαβ such that<br />

bαβn i = x i α,β . (1.5.71)<br />

By using the relation gijn i n j = 1 we can transform equation (1.5.71) to the form<br />

bαβ = gijn j x i α,β<br />

1<br />

=<br />

2 ɛγδɛijkx i α,βxjγ xkδ . (1.5.72)<br />

The second order symmetric tensor bαβ is called the curvature tensor and the quadratic form<br />

B = bαβdu α du β<br />

(1.5.73)<br />

is called the second fundamental form of the surface.<br />

Consider also the tensor derivative with respect to the surface coordinates of the unit normal vector to<br />

the surface. This derivative is<br />

n i <br />

∂ni i<br />

,α = + n<br />

∂uα jk g<br />

j x k α . (1.5.74)<br />

Taking the tensor derivative of gijninj = 1 with respect to the surface coordinates produces the result<br />

gijninj ,α = 0 which shows that the vector nj ,α is perpendicular to ni and must lie in the tangent plane to the<br />

surface. It can therefore be expressed as a linear combination of the surface tangent vectors xi α and written<br />

in the form<br />

n i ,α = η β αx i β<br />

(1.5.75)<br />

where the coefficients η β α can be written in terms of the surface metric components aαβ and the curvature<br />

components bαβ as follows. The unit vector n i is normal to the surface so that<br />

gijn i x j α<br />

=0. (1.5.76)


The tensor derivative of this equation with respect to the surface coordinates gives<br />

gijn i βxjα + gijn i x j<br />

α,β =0. (1.5.77)<br />

Substitute into equation (1.5.77) the relations from equations (1.5.57), (1.5.71) and (1.5.75) and show that<br />

Solving the equation (1.5.78) for the coefficients η γ<br />

β we find<br />

bαβ = −aαγη γ<br />

β . (1.5.78)<br />

η γ<br />

β = −aαγ bαβ. (1.5.79)<br />

Now substituting equation (1.5.79) into the equation (1.5.75) produces the Weingarten formula<br />

n i ,α = −aγβbγαx i β . (1.5.80)<br />

This is a relation for the derivative of the unit normal in terms of the surface metric, curvature tensor and<br />

surface tangents.<br />

A third fundamental form of the surface is given by the quadratic form<br />

where cαβ is defined as the symmetric surface tensor<br />

C = cαβdu α du β<br />

(1.5.81)<br />

cαβ = gijn i ,αn j<br />

,β . (1.5.82)<br />

By using the Weingarten formula in the equation (1.5.81) one can verify that<br />

Geodesic Coordinates<br />

cαβ = a γδ bαγbβδ. (1.5.83)<br />

In a Cartesian coordinate system the metric tensor gij is a constant and consequently the Christoffel<br />

symbols are zero at all points of the space. This is because the Christoffel symbols are dependent upon<br />

the derivatives of the metric tensor which is constant. If the space VN is not Cartesian then the Christoffel<br />

symbols do not vanish at all points of the space. However, it is possible to find a coordinate system where<br />

the Christoffel symbols will all vanish at a given point P of the space. Such coordinates are called geodesic<br />

coordinates of the point P.<br />

Consider a two dimensional surface with surface coordinates uα and surface metric aαβ. If we transform<br />

to some other two dimensional coordinate system, say ūα with metric āαβ, where the two coordinates are<br />

related by transformation equations of the form<br />

u α = u α (ū 1 , ū 2 ), α =1, 2, (1.5.84)<br />

145


146<br />

then from the transformation equation (1.4.7) we can write, after changing symbols,<br />

<br />

δ<br />

βγ<br />

ā<br />

∂u α<br />

=<br />

∂ū δ<br />

<br />

α<br />

δɛ<br />

a<br />

∂uδ ∂ū β<br />

∂uɛ ∂ū γ + ∂2uα ∂ū β . (1.5.85)<br />

∂ū γ<br />

This is a relationship between the Christoffel symbols in the two coordinate systems. If<br />

apointP , then for that particular point the equation (1.5.85) reduces to<br />

∂2uα ∂ū β <br />

α<br />

= −<br />

∂ū γ δɛ<br />

a<br />

∂uδ ∂ū β<br />

∂u ɛ<br />

∂ū γ<br />

<br />

δ<br />

vanishes at<br />

βγ<br />

ā<br />

(1.5.86)<br />

where all terms are evaluated at the<br />

point P. Conversely, if the equation (1.5.86) is satisfied at the point P,<br />

δ<br />

then the Christoffel symbol must be zero at this point. Consider the special coordinate transforma-<br />

βγ<br />

ā<br />

tion<br />

u α = u α 0 +ū α − 1<br />

<br />

α<br />

ū<br />

2 βγ a<br />

β ū α<br />

(1.5.87)<br />

where uα 0 are the surface coordinates of the point P. The point P in the new coordinates is given by<br />

ū α =0. We now differentiate the relation (1.5.87) to see if it satisfies the equation (1.5.86). We calculate<br />

the derivatives<br />

∂uα ∂ū τ = δα τ − 1<br />

<br />

α<br />

ū<br />

2 βτ a<br />

β − 1<br />

<br />

α<br />

ū<br />

2 τγ a<br />

γ<br />

<br />

<br />

(1.5.88)<br />

uα =0<br />

and<br />

∂2uα ∂ū τ <br />

α<br />

<br />

<br />

= −<br />

∂ū σ (1.5.89)<br />

τσ a<br />

uα =0<br />

where these derivative are evaluated at ū α =0. We find the derivative equations (1.5.88) and (1.5.89) do<br />

satisfy the equation (1.5.86) locally at the point P. Hence, the Christoffel symbols will all be zero at this<br />

particular point. The new coordinates can then be called geodesic coordinates.<br />

Riemann Christoffel Tensor<br />

Consider the Riemann Christoffel tensor defined by the equation (1.4.33). Various properties of this<br />

tensor are derived in the exercises at the end of this section. We will be particularly interested in the<br />

Riemann Christoffel tensor in a two dimensional space with metric aαβ and coordinates uα . We find the<br />

Riemann Christoffel tensor has the form<br />

R δ .αβγ = ∂<br />

∂uβ <br />

δ<br />

−<br />

αγ<br />

∂<br />

∂uγ <br />

δ τ δ τ δ<br />

+<br />

−<br />

(1.5.90)<br />

αβ αγ βτ αβ γτ<br />

where the Christoffel symbols are evaluated with respect to the surface metric. The above tensor has the<br />

associated tensor<br />

Rσαβγ = aσδR δ .αβγ<br />

which is skew-symmetric in the indices (σ, α) and(β,γ) such that<br />

(1.5.91)<br />

Rσαβγ = −Rασβγ and Rσαβγ = −Rσαγβ. (1.5.92)<br />

The two dimensional alternating tensor is used to define the constant<br />

K = 1<br />

4 ɛαβɛ γδ Rαβγδ<br />

(1.5.93)


(see example 1.5-1) which is an invariant of the surface and called the Gaussian curvature or total curvature.<br />

In the exercises following this section it is shown that the Riemann Christoffel tensor of the surface can be<br />

expressed in terms of the total curvature and the alternating tensors as<br />

Consider the second tensor derivative of x r α<br />

which can be shown to satisfy the relation<br />

Rαβγδ = Kɛαβɛγδ. (1.5.94)<br />

which is given by<br />

x r α,βγ = ∂xr <br />

α,β r<br />

+ x<br />

∂uγ mn g<br />

r α,βxnγ −<br />

<br />

δ<br />

x<br />

αγ a<br />

r δ,β −<br />

<br />

δ<br />

x<br />

βγ a<br />

r α,γ<br />

(1.5.95)<br />

x r α,βγ − x r α,γβ = R δ .αβγx r δ. (1.5.96)<br />

Using the relation (1.5.96) we can now derive some interesting properties relating to the tensors aαβ,bαβ,<br />

cαβ, Rαβγδ, the mean curvature H and the total curvature K.<br />

Consider the tensor derivative of the equation (1.5.71) which can be written<br />

x i α,βγ = bαβ,γn i + bαβn i ,γ<br />

(1.5.97)<br />

where<br />

bαβ,γ = ∂bαβ<br />

<br />

σ<br />

σ<br />

− bσβ − bασ. (1.5.98)<br />

∂uα αγ a βγ a<br />

By using the Weingarten formula, given in equation (1.5.80), the equation (1.5.97) can be expressed in the<br />

form<br />

x i α,βγ = bαβ,γn i − bαβa τσ bτγx i σ<br />

and by using the equations (1.5.98) and (1.5.99) it can be established that<br />

(1.5.99)<br />

x r α,βγ − xr α,γβ =(bαβ,γ − bαγ,β)n r − a τδ (bαβbτγ − bαγbτβ)x r δ . (1.5.100)<br />

Now by equating the results from the equations (1.5.96) and (1.5.100) we arrive at the relation<br />

R δ .αβγx r δ =(bαβ,γ − bαγ,β)n r − a τδ (bαβbτγ − bαγbτβ)x r δ. (1.5.101)<br />

Multiplying the equation (1.5.101) by nr and using the results from the equation (1.5.76) there results the<br />

Codazzi equations<br />

bαβ,γ − bαγ,β =0. (1.5.102)<br />

Multiplying the equation (1.5.101) by grmx m σ and simplifying one can derive the Gauss equations of the<br />

surface<br />

Rσαβγ = bαγbσβ − bαβbσγ. (1.5.103)<br />

By using the Gauss equations (1.5.103) the equation (1.5.94) can be written as<br />

Kɛσαɛβγ = bαγbσβ − bαβbσγ. (1.5.104)<br />

147


148<br />

Another form of equation (1.5.104) is obtained by using the equation (1.5.83) together with the relation<br />

aαβ = −a σγ ɛσαɛβγ. It is left as an exercise to verify the resulting form<br />

−Kaαβ = cαβ − a σγ bσγbαβ. (1.5.106)<br />

Define the quantity<br />

H = 1<br />

2 aσγbσγ as the mean curvature of the surface, then the equation (1.5.106) can be written in the form<br />

By multiplying the equation (1.5.108) by du α du β and summing, we find<br />

is a relation connecting the first, second and third fundamental forms.<br />

(1.5.107)<br />

cαβ − 2Hbαβ + Kaαβ =0. (1.5.108)<br />

C − 2HB+ KA=0 (1.5.109)<br />

EXAMPLE 1.5-2<br />

In a two dimensional space the Riemann Christoffel tensor has only one nonzero independent component<br />

R<strong>12</strong><strong>12</strong>. ( See Exercise 1.5, problem number 21.) Consequently, the equation (1.5.104) can be written in the<br />

form K √ √<br />

ae<strong>12</strong> ae<strong>12</strong> = b22b11 − b21b<strong>12</strong> and solving for the Gaussian curvature K we find<br />

Surface Curvature<br />

K = b22b11 − b<strong>12</strong>b21<br />

=<br />

a11a22 − a<strong>12</strong>a21<br />

b<br />

a<br />

R<strong>12</strong><strong>12</strong><br />

= . (1.5.110)<br />

a<br />

For a surface curve u α = u α (s),α =1, 2 lying upon a surface x i = x i (u 1 ,u 2 ),i =1, 2, 3, we have a two<br />

dimensional space embedded in a three dimensional space. Thus, if t α = duα<br />

is a unit tangent vector to<br />

ds<br />

du<br />

the surface curve then aαβ<br />

α du<br />

ds<br />

β<br />

vector to the space curve x i = x i (u 1 (s),u 2 (s)) with T i = dxi<br />

ds<br />

The surface vector tα and the space vector T i are related by<br />

ds = aαβt α t β =1. This same vector can be represented as the unit tangent<br />

. That is we will have gij<br />

dx i<br />

ds<br />

dx j<br />

ds = gijT i T j =1.<br />

T i = ∂xi<br />

∂uα duα ds = xiα tα . (1.5.111)<br />

The surface vector t α is a unit vector so that aαβt α t β =1. If we differentiate this equation intrinsically with<br />

α δtβ<br />

respect to the parameter s, we find that aαβt<br />

δs<br />

=0. This shows that the surface vector δtα<br />

δs<br />

is perpendicular<br />

to the surface vector tα . Let uα denote a unit normal vector in the surface plane which is orthogonal to the<br />

tangent vector tα . The direction of uα is selected such that ɛαβtαuβ =1. Therefore, there exists a scalar κ (g)<br />

such that<br />

δt α<br />

δs = κ (g)u α<br />

(1.5.1<strong>12</strong>)


where κ (g) is called the geodesic curvature of the curve. In a similar manner it can be shown that δuα<br />

δs<br />

is a surface vector orthogonal to t α . Let δuα<br />

δs = αtα where α is a scalar constant to be determined. By<br />

differentiating the relation aαβt α u β = 0 intrinsically and simplifying we find that α = −κ (g) and therefore<br />

δu α<br />

δs = −κ (g)t α . (1.5.113)<br />

The equations (1.5.1<strong>12</strong>) and (1.5.113) are sometimes referred to as the Frenet-Serret formula for a curve<br />

relative to a surface.<br />

Taking the intrinsic derivative of equation (1.5.111), with respect to the parameter s, we find that<br />

δT i<br />

δs = xi δt<br />

α<br />

α<br />

δs + xi du<br />

α,β<br />

β<br />

ds tα . (1.5.114)<br />

Treating the curve as a space curve we use the Frenet formulas (1.5.13). If we treat the curve as a surface<br />

curve, then we use the Frenet formulas (1.5.1<strong>12</strong>) and (1.5.113). In this way the equation (1.5.114) can be<br />

writtenintheform<br />

κN i = x i α κ (g)u α + x i α,β tβ t α . (1.5.115)<br />

By using the results from equation (1.5.71) in equation (1.5.115) we obtain<br />

κN i = κ (g)u i + bαβn i t α t β<br />

(1.5.116)<br />

where ui is the space vector counterpart of the surface vector uα . Let θ denote the angle between the surface<br />

normal ni and the principal normal N i ,thenwehavethatcosθ = niN i . Hence, by multiplying the equation<br />

(1.5.116) by ni we obtain<br />

κ cos θ = bαβt α t β . (1.5.117)<br />

Consequently, for all curves on the surface with the same tangent vector tα ,thequantityκcos θ will remain<br />

constant. This result is known as Meusnier’s theorem. Note also that κ cos θ = κ (n) is the normal component<br />

of the curvature and κ sin θ = κ (g) is the geodesic component of the curvature. Therefore, we write the<br />

equation (1.5.117) as<br />

κ (n) = bαβt α t β<br />

(1.5.118)<br />

which represents the normal curvature of the surface in the direction tα . The equation (1.5.118) can also be<br />

writtenintheform<br />

du<br />

κ (n) = bαβ<br />

α du<br />

ds<br />

β<br />

which is a ratio of quadratic forms.<br />

B<br />

=<br />

ds A<br />

(1.5.119)<br />

The surface directions for which κ (n) has a maximum or minimum value is determined from the equation<br />

(1.5.119) which is written as<br />

(bαβ − κ (n)aαβ)λ α λ β =0. (1.5.<strong>12</strong>0)<br />

The direction giving a maximum or minimum value to κ (n) must then satisfy<br />

(bαβ − κ (n)aαβ)λ β =0 (1.5.<strong>12</strong>1)<br />

149


150<br />

so that κ (n) must be a root of the determinant equation<br />

The expanded form of equation (1.5.<strong>12</strong>2) can be written as<br />

det(bαβ − κ (n)aαβ) =0. (1.5.<strong>12</strong>2)<br />

κ 2 (n) − aαβbαβκ (n) + b<br />

=0 (1.5.<strong>12</strong>3)<br />

a<br />

where a = a11a22 − a<strong>12</strong>a21 and b = b11b22 − b<strong>12</strong>b21. Using the definition given in equation (1.5.107) and using<br />

the result from equation (1.5.110), the equation (1.5.<strong>12</strong>3) can be expressed in the form<br />

The roots κ (1) and κ (2) of the equation (1.5.<strong>12</strong>4) then satisfy the relations<br />

and<br />

κ 2 (n) − 2Hκ (n) + K =0. (1.5.<strong>12</strong>4)<br />

H = 1<br />

2 (κ (1) + κ (2)) (1.5.<strong>12</strong>5)<br />

K = κ (1)κ (2). (1.5.<strong>12</strong>6)<br />

Here H is the mean value of the principal curvatures and K is the Gaussian or total curvature which is the<br />

product of the principal curvatures. It is readily verified that<br />

H =<br />

Eg − 2fF + eG<br />

2(EG − F 2 )<br />

and K =<br />

are invariants obtained from the surface metric and curvature tensor.<br />

eg − f 2<br />

EG − F 2<br />

Relativity<br />

Sir Isaac Newton and Albert Einstein viewed the world differently when it came to describing gravity and<br />

the motion of the planets. In this brief introduction to relativity we will compare the Newtonian equations<br />

with the relativistic equations in describing planetary motion. We begin with an examination of Newtonian<br />

systems.<br />

Newton’s viewpoint of planetary motion is a multiple bodied problem, but for simplicity we consider<br />

only a two body problem, say the sun and some planet where the motion takes place in a plane. Newton’s<br />

law of gravitation states that two masses m and M are attracted toward each other with a force of magnitude<br />

GmM<br />

ρ2 ,whereGisaconstant, ρ is the distance between the masses, m is the mass of the planet and M is the<br />

mass of the sun. One can construct an x, y plane containing the two masses with the origin located at the<br />

center of mass of the sun. Let eρ =cosφe1 +sinφe2 denote a unit vector at the origin of this coordinate<br />

systemandpointinginthedirectionofthemassm. The vector force of attraction of mass M on mass m is<br />

given by the relation<br />

F = −GmM<br />

ρ 2<br />

eρ. (1.5.<strong>12</strong>7)


Figure 1.5-2. Parabolic and elliptic conic sections<br />

The equation of motion of mass m with respect to mass M is obtained from Newton’s second law. Let<br />

ρ = ρ eρ denote the position vector of mass m with respect to the origin. Newton’s second law can then be<br />

written in any of the forms<br />

F = −GmM<br />

ρ 2<br />

eρ = m d2 ρ<br />

dt 2 = md V<br />

dt<br />

= −GmM<br />

ρ 3 ρ (1.5.<strong>12</strong>8)<br />

and from this equation we can show that the motion of the mass m can be described as a conic section.<br />

Recall that a conic section is defined as a locus of points p(x, y) such that the distance of p from a fixed<br />

point (or points), called a focus (foci), is proportional to the distance of the point p from a fixed line, called<br />

a directrix, that does not contain the fixed point. The constant of proportionality is called the eccentricity<br />

and is denoted by the symbol ɛ. For ɛ = 1 a parabola results; for 0 ≤ ɛ ≤ 1 an ellipse results; for ɛ>1a<br />

hyperbola results; and if ɛ = 0 the conic section is a circle.<br />

With reference to figure 1.5-2, a conic section is defined in terms of the ratio FP = ɛ where FP = ρ and<br />

PD<br />

PD =2q− ρ cos φ. Fromtheɛratio we solve for ρ and obtain the polar representation for the conic section<br />

ρ =<br />

p<br />

1+ɛ cos φ<br />

(1.5.<strong>12</strong>9)<br />

151


152<br />

where p =2qɛ and the angle φ is known as the true anomaly associated with the orbit. The quantity p is<br />

called the semi-parameter of the conic section. (Note that when φ = π<br />

2 ,thenρ = p.) A more general form<br />

of the above equation is<br />

ρ =<br />

p<br />

1+ɛ cos(φ − φ0)<br />

or u = 1<br />

ρ = A[1 + ɛ cos(φ − φ0)], (1.5.130)<br />

where φ0 is an arbitrary starting anomaly. An additional symbol a, knownasthesemi-majoraxesofan<br />

elliptical orbit can be introduced where q, p, ɛ, a are related by<br />

p<br />

1+ɛ = q = a(1 − ɛ) or p = a(1 − ɛ2 ). (1.5.131)<br />

To show that the equation (1.5.<strong>12</strong>8) produces a conic section for the motion of mass m with respect to<br />

mass M we will show that one form of the solution of equation (1.5.<strong>12</strong>8) is given by the equation (1.5.<strong>12</strong>9).<br />

To verify this we use the following vector identities:<br />

From the equation (1.5.<strong>12</strong>8) we find that<br />

<br />

d<br />

ρ ×<br />

dt<br />

dρ<br />

<br />

dt<br />

so that an integration of equation (1.5.133) produces<br />

ρ × eρ =0<br />

<br />

d<br />

ρ ×<br />

dt<br />

dρ<br />

<br />

=ρ ×<br />

dt<br />

d2ρ dt2 d eρ<br />

eρ ·<br />

dt =0<br />

<br />

d eρ d eρ<br />

eρ × eρ × = −<br />

dt dt .<br />

(1.5.132)<br />

= ρ × d2ρ GM<br />

= −<br />

dt2 ρ2 ρ × eρ = 0 (1.5.133)<br />

ρ × dρ<br />

dt = h = constant. (1.5.134)<br />

The quantity H = ρ × m V = ρ × m dρ<br />

dt is the angular momentum of the mass m so that the quantity h<br />

represents the angular momentum per unit mass. The equation (1.5.134) tells us that h is a constant for our<br />

two body system. Note that because h is constant we have<br />

d<br />

<br />

V × h =<br />

dt<br />

d V<br />

dt × h = − GM<br />

ρ2 eρ<br />

<br />

× ρ × dρ<br />

<br />

dt<br />

= − GM<br />

ρ2 eρ<br />

d eρ dρ<br />

× [ρ eρ × (ρ +<br />

dt dt eρ)]<br />

and consequently an integration produces<br />

= − GM<br />

ρ 2 eρ × ( eρ ×<br />

V × h = GM eρ + C<br />

d eρ<br />

dt )ρ2 = GM<br />

d eρ<br />

dt


where C is a vector constant of integration. The triple scalar product formula gives us<br />

ρ · ( V × h)= h · (ρ × dρ<br />

dt )=h2 = GM ρ · eρ + ρ · C<br />

or<br />

h 2 = GMρ + Cρcos φ (1.5.135)<br />

where φ is the angle between the vectors C and ρ. From the equation (1.5.135) we find that<br />

ρ =<br />

p<br />

1+ɛ cos φ<br />

(1.5.136)<br />

where p = h2 /GM and ɛ = C/GM. This result is known as Kepler’s first law and implies that when ɛ


154<br />

The substitution ρ = 1<br />

u<br />

Figure 1.5-3. Relative motion of two inertial systems.<br />

can be used to represent the equation (1.5.142) in the form<br />

2 du<br />

dφ<br />

+ u 2 − 2GM<br />

h<br />

E<br />

u + =0 (1.5.143)<br />

2 h2 which is a form we will return to later in this section. Note that we can separate the variables in equations<br />

(1.5.142) or (1.5.143). The results can then be integrate to produce the equation (1.5.130).<br />

Newton also considered the relative motion of two inertial systems, say S and S. Consider two such<br />

systems as depicted in the figure 1.5-3 where the S system is moving in the x−direction with speed v relative<br />

to the system S.<br />

For a Newtonian system, if at time t = 0 we have clocks in both systems which coincide, than at time t<br />

apointP (x, y, z) intheSsystem can be described by the transformation equations<br />

x =x + vt<br />

y =y<br />

z =z<br />

t =t<br />

or<br />

x =x − vt<br />

y =y<br />

z =z<br />

t =t.<br />

(1.5.144)<br />

These are the transformation equation of Newton’s relativity sometimes referred to as a Galilean transformation.<br />

Before Einstein the principle of relativity required that velocities be additive and obey Galileo’s velocity<br />

addition rule<br />

V P/R = V P/Q + V Q/R. (1.5.145)


That is, the velocity of P with respect to R equals the velocity of P with respect to Q plus the velocity of Q<br />

with respect to R. For example, a person (P ) running north at 3 km/hr on a train (Q) moving north at 60<br />

km/hr with respect to the ground (R) has a velocity of 63 km/hr with respect to the ground. What happens<br />

when (P ) is a light wave moving on a train (Q) which is moving with velocity V relative to the ground? Are<br />

the velocities still additive? This type of question led to the famous Michelson-Morley experiment which<br />

has been labeled as the starting point for relativity. Einstein’s answer to the above question was ”NO” and<br />

required that V P/R = V P/Q = c =speed of light be a universal constant.<br />

In contrast to the Newtonian equations, Einstein considered the motion of light from the origins 0 and<br />

0 of the systems S and S. IftheSsystem moves with velocity v relative to the S system and at time t =0<br />

a light signal is sent from the S system to the S system, then this light signal will move out in a spherical<br />

wave front and lie on the sphere<br />

x 2 + y 2 + z 2 = c 2 t 2<br />

(1.5.146)<br />

where c is the speed of light. Conversely, if a light signal is sent out from the S system at time t = 0, it will<br />

lie on the spherical wave front<br />

x 2 + y 2 + z 2 = c 2 t 2 . (1.5.147)<br />

Observe that the Newtonian equations (1.5.144) do not satisfy the equations (1.5.146) and (1.5.147) identically.<br />

If y = y and z = z then the space variables (x, x) andtimevariables(t, t) must somehow be related.<br />

Einstein suggested the following transformation equations between these variables<br />

x = γ(x − vt) and x = γ(x + vt) (1.5.148)<br />

where γ is a constant to be determined. The differentials of equations (1.5.148) produce<br />

from which we obtain the ratios<br />

dx γ(dx − vdt)<br />

=<br />

γ(dx + vdt) dx<br />

When dx<br />

dt<br />

= dx<br />

dt<br />

dx = γ(dx − vdt) and dx = γ(dx + vdt) (1.5.149)<br />

or<br />

1<br />

γ(1 + v dx<br />

dt<br />

= c, the speed of light, the equation (1.5.150) requires that<br />

γ 2 =(1− v2<br />

)−1<br />

c2 From the equations (1.5.148) we eliminate x and find<br />

) = γ(1 − v<br />

dx<br />

dt<br />

). (1.5.150)<br />

or γ =(1− v2<br />

c 2 )−1/2 . (1.5.151)<br />

t = γ(t − v<br />

x). (1.5.152)<br />

c2 We can now replace the Newtonian equations (1.5.144) by the relativistic transformation equations<br />

x =γ(x + vt)<br />

y =y<br />

z =z<br />

t =γ(t + v<br />

x)<br />

c2 or<br />

x =γ(x − vt)<br />

y =y<br />

z =z<br />

t =γ(t − v<br />

x)<br />

c2 (1.5.153)<br />

155


156<br />

where γ is given by equation (1.5.151). These equations are also known as the Lorentz transformation.<br />

Note that for v


where g11 = −1, g22 = −ρ2 , g33 = −ρ2 sin 2 θ, g44 = c2 and gij =0fori= j. The negative signs are<br />

2 2 2 = c − v is positive when v


158<br />

Subtracting the first equation from the third equation gives<br />

The second equation in (1.5.164) then becomes<br />

du dv<br />

+<br />

dρ dρ =0 or u + v = c1 = constant. (1.5.165)<br />

ρ du<br />

=1− eu<br />

dρ<br />

Separate the variables in equation (1.5.166) and integrate to obtain the result<br />

e u =<br />

where c2 is a constant of integration and consequently<br />

e v = e c1−u = e c1<br />

1<br />

1 − c2<br />

ρ<br />

(1.5.166)<br />

(1.5.167)<br />

<br />

1 − c2<br />

<br />

. (1.5.168)<br />

ρ<br />

The constant c1 is selected such that g44 approaches c2 as ρ increases without bound. This produces the<br />

metrices<br />

g11 = −1<br />

1 − c2 , g22 = −ρ<br />

ρ<br />

2 , g33 = −ρ 2 sin 2 θ, g44 = c 2 (1 − c2<br />

) (1.5.169)<br />

ρ<br />

where c2 is a constant still to be determined. The metrices given by equation (1.5.169) are now used to<br />

expand the equations (1.5.157) representing the geodesics in this four dimensional space. The differential<br />

equations representing the geodesics are found to be<br />

d2 2 ρ 1 du dρ<br />

+ − ρe<br />

ds2 2 dρ ds<br />

−u<br />

2 dθ<br />

− ρe<br />

ds<br />

−u sin 2 2 dφ<br />

θ +<br />

ds<br />

1<br />

2 dv dt<br />

ev−u =0<br />

2 dρ ds<br />

d<br />

(1.5.170)<br />

2 2 θ 2 dθ dρ<br />

dφ<br />

+ − sin θ cos θ =0<br />

ds2 ρ ds ds ds<br />

(1.5.171)<br />

d2φ 2 dφ dρ θ dφ dθ<br />

+ +2cos =0<br />

ds2 ρ ds ds sin θ ds ds<br />

(1.5.172)<br />

d2t dv dt dρ<br />

+ =0.<br />

ds2 dρ ds ds<br />

(1.5.173)<br />

is a constant. This<br />

value of θ also simplifies the equations (1.5.170) and (1.5.172). The equation (1.5.172) becomes an exact<br />

differential equation<br />

<br />

d 2 dφ<br />

ρ =0<br />

ds ds<br />

or<br />

2 dφ<br />

ρ<br />

ds = c4, (1.5.174)<br />

and the equation (1.5.173) also becomes an exact differential<br />

<br />

d dt<br />

ds ds ev<br />

<br />

=0 or<br />

dt<br />

ds ev = c5, (1.5.175)<br />

The equation (1.5.171) is identically satisfied if we examine planar orbits where θ = π<br />

2<br />

where c4 and c5 are constants of integration. This leaves the equation (1.5.170) which determines ρ. Substituting<br />

the results from equations (1.5.174) and (1.5.175), together with the relation (1.5.161), the equation<br />

(1.5.170) reduces to<br />

d2ρ c2<br />

+<br />

ds2 2ρ2 + c2c2 4 c2<br />

− (1 −<br />

2ρ4 ρ ) c24 =0. (1.5.176)<br />

ρ3


By the chain rule we have<br />

d2ρ ds2 = d2ρ dφ2 2 dφ<br />

+<br />

ds<br />

dρ d<br />

dφ<br />

2φ ds2 = d2ρ dφ2 and so equation (1.5.176) can be written in the form<br />

The substitution ρ = 1<br />

u<br />

d2ρ 2<br />

−<br />

dφ2 ρ<br />

c2 4<br />

+<br />

ρ4 dρ<br />

dφ<br />

2 2 −2c4 ρ5 <br />

2 dρ<br />

+<br />

dφ<br />

c2 ρ<br />

2<br />

2<br />

c2 +<br />

4<br />

c2<br />

2 −<br />

<br />

1 − c2<br />

<br />

ρ =0. (1.5.177)<br />

ρ<br />

reduces the equation (1.5.177) to the form<br />

d2u c2<br />

+ u −<br />

dφ2 2c2 =<br />

4<br />

3<br />

2 c2u 2 . (1.5.178)<br />

Multiply the equation (1.5.178) by 2 du<br />

dφ and integrate with respect to φ to obtain<br />

2 du<br />

dφ<br />

+ u 2 − c2<br />

c2 u = c2u<br />

4<br />

3 + c6. (1.5.179)<br />

where c6 is a constant of integration. To determine the constant c6 we write the equation (1.5.161) in the<br />

special case θ = π<br />

2 and use the substitutions from the equations (1.5.174) and (1.5.175) to obtain<br />

e u<br />

2 dρ<br />

= e<br />

ds<br />

u<br />

<br />

dρ<br />

dφ<br />

2 dφ<br />

=1− ρ<br />

ds<br />

2<br />

2 dφ<br />

+ e<br />

ds<br />

v<br />

2 dt<br />

ds<br />

or 2 <br />

dρ<br />

+ 1 −<br />

dφ<br />

c2<br />

<br />

ρ<br />

ρ<br />

2 <br />

+ 1 − c2<br />

ρ − c25 c2 4 ρ<br />

c2 =0. (1.5.180)<br />

4<br />

reduces the equation (1.5.180) to the form<br />

2 du<br />

+ u<br />

dφ<br />

2 − c2u 3 + 1<br />

c2 4<br />

Now comparing the equations (1.5.181) and (1.5.179) we select<br />

2 c5 1<br />

c6 = − 1<br />

c2 c2 4<br />

so that the equation (1.5.179) takes on the form<br />

2 du<br />

+ u<br />

dφ<br />

2 − c2<br />

c2 <br />

u + 1 −<br />

4<br />

c25 c2 <br />

1<br />

c2 4<br />

= c2u 3<br />

(1.5.182)<br />

Now we can compare our relativistic equation (1.5.182) with our Newtonian equation (1.5.143). In order<br />

that the two equations almost agree we select the constants c2,c4,c5 so that<br />

The substitution ρ = 1<br />

u<br />

c2<br />

c2 4<br />

= 2GM<br />

h 2<br />

and<br />

− c2<br />

c2 u −<br />

4<br />

c25 c2c2 =0. (1.5.181)<br />

4<br />

1 − c2 5<br />

c 2<br />

c 2 4<br />

= E<br />

. (1.5.183)<br />

h2 The equations (1.5.183) are only two equations in three unknowns and so we use the additional equation<br />

dφ<br />

lim ρ2 = lim<br />

ρ→∞ dt ρ→∞<br />

dφ ds<br />

ρ2 = h (1.5.184)<br />

ds dt<br />

159


160<br />

which is obtained from equation (1.5.141). Substituting equations (1.5.174) and (1.5.175) into equation<br />

(1.5.184), rearranging terms and taking the limit we find that<br />

From equations (1.5.183) and (1.5.185) we obtain the results that<br />

c 2 5<br />

c4c2 = h. (1.5.185)<br />

c5<br />

c2<br />

=<br />

1+ E<br />

c2 , c2 = 2GM<br />

c2 <br />

1<br />

1+E/c 2<br />

<br />

, c4 =<br />

h<br />

c 1+E/c 2<br />

These values substituted into equation (1.5.181) produce the differential equation<br />

Let α = c2<br />

c2 =<br />

4<br />

2GM<br />

h2 2 du<br />

dφ<br />

+ u 2 − 2GM<br />

h<br />

(1.5.186)<br />

E 2GM<br />

u + = 2 h2 c2 <br />

1<br />

1+E/c2 <br />

u 3 . (1.5.187)<br />

and β = c2 = 2GM<br />

c 2 ( 1<br />

1+E/c 2 ) then the differential equation (1.5.178) can be written as<br />

We know the solution to equation (1.5.143) is given by<br />

d2u α 3<br />

+ u − =<br />

dφ2 2 2 βu2 . (1.5.188)<br />

u = 1<br />

ρ = A(1 + ɛ cos(φ − φ0)) (1.5.189)<br />

and so we assume a solution to equation (1.5.188) of this same general form. We know that A is small and so<br />

we make the assumption that the solution of equation (1.5.188) given by equation (1.5.189) is such that φ0 is<br />

approximately constant and varies slowly as a function of Aφ. Observethatifφ0 = φ0(Aφ), then dφ0<br />

dφ = φ′ 0A<br />

and d2φ0 dφ2 = φ ′′<br />

0A2 , where primes denote differentiation with respect to the argument of the function. (i.e.<br />

Aφ for this problem.) The derivatives of equation (1.5.189) produce<br />

du<br />

dφ = − ɛA sin(φ − φ0)(1 − φ ′ 0A)<br />

d2u dφ2 =ɛA3 sin(φ − φ0)φ ′′<br />

0 − ɛA cos(φ − φ0)(1 − 2Aφ ′ 0 + A 2 (φ ′ 0) 2 )<br />

= − ɛA cos(φ − φ0)+2ɛA 2 φ ′ 0 cos(φ − φ0)+O(A 3 ).<br />

Substituting these derivatives into the differential equation (1.5.188) produces the equations<br />

2ɛA 2 φ ′ 0 cos(φ − φ0)+A − α<br />

2<br />

= 3β<br />

2<br />

A 2 +2ɛA 2 cos(φ − φ0)+ɛ 2 A 2 cos 2 (φ − φ0) + O(A 3 ).<br />

Now A is small so that terms O(A3 ) can be neglected. Equating the constant terms and the coefficient of<br />

the cos(φ − φ0) terms we obtain the equations<br />

A − α<br />

2<br />

= 3β<br />

2 A2<br />

2ɛA 2 φ ′ 0 =3βɛA 2 + 3β<br />

2 ɛ2 A 2 cos(φ − φ0).<br />

Treating φ0 as essentially constant, the above system has the approximate solutions<br />

A ≈ α<br />

2<br />

φ0 ≈ 3β 3β<br />

Aφ +<br />

2 4 Aɛ sin(φ − φ0) (1.5.190)


The solutions given by equations (1.5.190) tells us that φ0 varies slowly with time. For ɛ less than 1, the<br />

elliptical motion is affected by this change in φ0. It causes the semi-major axis of the ellipse to slowly rotate<br />

at a rate given by dφ0<br />

dt . Using the following values for the planet Mercury<br />

G =6.67(10 −8 )dynecm 2 /g 2<br />

M =1.99(10 33 )g<br />

a =5.78(10 <strong>12</strong> )cm<br />

ɛ =0.206<br />

c =3(10 10 )cm/sec<br />

β ≈ 2GM<br />

c2 =2.95(105 )cm<br />

h ≈ GMa(1 − ɛ2 )=2.71(10 19 )cm 2 /sec<br />

dφ<br />

dt ≈<br />

<br />

GM<br />

a3 1/2 sec −1 Kepler’s third law<br />

we calculate the slow rate of rotation of the semi-major axis to be approximately<br />

dφ0<br />

dt<br />

= dφ0<br />

dφ<br />

dφ<br />

dt<br />

3<br />

≈<br />

2 βAdφ<br />

<br />

GM<br />

≈ 3<br />

dt ch<br />

2 GM<br />

a 3<br />

1/2<br />

=6.628(10 −14 )rad/sec<br />

=43.01 seconds of arc per century.<br />

(1.5.191)<br />

(1.5.192)<br />

This slow variation in Mercury’s semi-major axis has been observed and measured and is in agreement with<br />

the above value. Newtonian mechanics could not account for the changes in Mercury’s semi-major axis, but<br />

Einstein’s theory of relativity does give this prediction. The resulting solution of equation (1.5.188) can be<br />

viewed as being caused by the curvature of the space-time continuum.<br />

The contracted curvature tensor Gij set equal to zero is just one of many conditions that can be assumed<br />

in order to arrive at a metric for the space-time continuum. Any assumption on the value of Gij relates to<br />

imposing some kind of curvature on the space. Within the large expanse of our universe only our imaginations<br />

limit us as to how space, time and matter interact. You can also imagine the existence of other tensor metrics<br />

in higher dimensional spaces where the geodesics within the space-time continuum give rise to the motion<br />

of other physical quantities.<br />

This short introduction to relativity is concluded with a quote from the NASA News@hg.nasa.gov news<br />

release, spring 1998, Release:98-51. “An international team of NASA and university researchers has found<br />

the first direct evidence of a phenomenon predicted 80 years ago using Einstein’s theory of general relativity–<br />

that the Earth is dragging space and time around itself as it rotates.”The news release explains that the<br />

effect is known as frame dragging and goes on to say “Frame dragging is like what happens if a bowling<br />

ball spins in a thick fluid such as molasses. As the ball spins, it pulls the molasses around itself. Anything<br />

stuck in the molasses will also move around the ball. Similarly, as the Earth rotates it pulls space-time in<br />

its vicinity around itself. This will shift the orbits of satellites near the Earth.”This research is reported in<br />

the journal Science.<br />

161


162<br />

EXERCISE 1.5<br />

◮ 1. Let κ = δ T<br />

δs · N and τ = δ N<br />

δs · B. Assume in turn that each of the intrinsic derivatives of T, N, B are<br />

some linear combination of T, N, B and hence derive the Frenet-Serret formulas of differential geometry.<br />

◮ 2. Determine the given surfaces. Describe and sketch the curvilinear coordinates upon each surface.<br />

(a) r(u, v) =u e1 + v e2 (b) r(u, v) =u cos v e1 + u sin v e2 (c) r(u, v) = 2uv2<br />

u2 + v2 e1 + 2u2v u2 e2.<br />

+ v2 ◮ 3. Determine the given surfaces and describe the curvilinear coordinates upon the surface. Use some<br />

graphics package to plot the surface and illustrate the coordinate curves on the surface. Find element of<br />

area dS in terms of u and v.<br />

(a) r(u, v) =a sin u cos v e1 + b sin u sin v e2 + c cos u e3 a, b, c constants 0 ≤ u, v ≤ 2π<br />

(b) r(u, v) =(4+vsin u<br />

2 )cosue1 +(4+v sin u<br />

2 )sinue2 + v cos u<br />

2 e3 − 1 ≤ v ≤ 1, 0 ≤ u ≤ 2π<br />

(c) r(u, v) =au cos v e1 + bu sin v e2 + cu e3<br />

(d) r(u, v) =u cos v e1 + u sin v e2 + αv e3 α constant<br />

(e) r(u, v) =a cos v e1 + b sin v e2 + u e3 a, b constant<br />

(f) r(u, v) =u cos v e1 + u sin v e2 + u 2 e3<br />

◮ 4.<br />

<br />

E<br />

Consider a two dimensional space with metric tensor (aαβ) =<br />

F<br />

<br />

F<br />

. Assume that the surface is<br />

G<br />

described by equations of the form y i = y i (u, v) and that any point on the surface is given by the position<br />

vector r = r(u, v) =yi ei. Show that the metrices E,F,G are functions of the parameters u, v and are given<br />

by<br />

E = ru · ru, F = ru · rv, G = rv · rv where ru = ∂r<br />

∂u and rv = ∂r<br />

∂v .<br />

◮ 5. For the metric given in problem 4 show that the Christoffel symbols of the first kind are given by<br />

[1 1, 1] = ru · ruu [1 2, 1] = [2 1, 1] = ru · ruv [2 2, 1] = ru · rvv<br />

[1 1, 2] = rv · ruu<br />

[1 2, 2] = [2 1, 2] = rv · ruv<br />

which can be represented [αβ,γ]= ∂2r ∂uα ∂r<br />

· ,<br />

∂uβ ∂uγ α,β,γ =1, 2.<br />

◮ 6. Show that the results in problem 5 can also be written in the form<br />

[1 1, 1] = 1<br />

2 Eu [1 2, 1] = [2 1, 1] = 1<br />

2 Ev<br />

[1 1, 2] = Fu − 1<br />

2 Ev<br />

where the subscripts indicate partial differentiation.<br />

[1 2, 2] = [2 1, 2] = 1<br />

2 Gu<br />

[2 2, 2] = rv · rvv<br />

[2 2, 1] = Fv − 1<br />

2 Gu<br />

[2 2, 2] = 1<br />

2 Gv<br />

◮ 7. For the metric<br />

<br />

given<br />

<br />

in problem 4, show that the Christoffel symbols of the second kind can be<br />

γ<br />

expressed in the form = a<br />

αβ<br />

γδ [αβ,δ], α,β,γ =1, 2 and produce the results<br />

<br />

1<br />

=<br />

11<br />

GEu − 2FFu + FEv<br />

2(EG − F 2 )<br />

<br />

1<br />

=<br />

22<br />

2GFv − GGu − FGv<br />

2(EG − F 2 <br />

1 1<br />

= =<br />

<strong>12</strong> 21<br />

)<br />

GEv − FGu<br />

2(EG − F 2 )<br />

<br />

2 2<br />

= =<br />

<strong>12</strong> 21<br />

EGu − FEv<br />

2(EG − F 2 <br />

2<br />

=<br />

11<br />

)<br />

2EFu − EEv − FEu<br />

2(EG − F 2 )<br />

<br />

2<br />

=<br />

22<br />

EGv − 2FFv + FGu<br />

2(EG − F 2 )<br />

where the subscripts indicate partial differentiation.


◮ 8. Derive the Gauss equations by assuming that<br />

ruu = c1ru + c2rv + c3 n , ruv = c4ru + c5rv + c6 n , rvv = c7ru + c8rv + c9n<br />

where c1,...,c9 are constants<br />

<br />

determined<br />

<br />

by<br />

<br />

taking dot products<br />

<br />

of<br />

<br />

the above<br />

<br />

vectors<br />

<br />

with the vectors ru,rv,<br />

1<br />

2<br />

1<br />

2<br />

and n. Show that c1 = , c2 = , c3 = e, c4 = , c5 = , c6 = f,<br />

<br />

11 11<br />

<strong>12</strong> <strong>12</strong><br />

1<br />

2<br />

∂<br />

c7 = , c8 = , c9 = g Show the Gauss equations can be written<br />

22 22<br />

2r ∂uα <br />

γ ∂r<br />

=<br />

+bαβn.<br />

∂uβ αβ ∂uγ ◮ 9. Derive the Weingarten equations<br />

and show<br />

fF − eG<br />

c1 =<br />

EG − F 2<br />

eF − fE<br />

c2 =<br />

EG − F 2<br />

nu = c1ru + c2rv<br />

nv = c3ru + c4rv<br />

gF − fG<br />

c3 =<br />

EG − F 2<br />

fF − gE<br />

c4 =<br />

EG − F 2<br />

and<br />

c ∗ 1 =<br />

c ∗ 2<br />

ru = c ∗ 1 nu + c ∗ 2 nv<br />

rv = c ∗ 3 nu + c ∗ 4 nv<br />

fF − gE<br />

eg − f 2<br />

= fE − eF<br />

eg − f 2<br />

c ∗ 3 =<br />

c ∗ 4<br />

fG− gF<br />

eg − f 2<br />

= fF − eG<br />

eg − f 2<br />

The constants in the above equations are determined in a manner similar to that suggested in problem 8.<br />

Show that the Weingarten equations can be written in the form<br />

∂n<br />

∂u α = −bβ α<br />

∂r<br />

.<br />

∂uβ ◮ 10. Using n = ru × rv<br />

√ , the results from exercise 1.1, problem 9(a), and the results from problem 5,<br />

EG − F 2<br />

verify that<br />

<br />

2 EG<br />

(ru × ruu) · n =<br />

− F 2<br />

11<br />

<br />

2 EG<br />

(ru × ruv) · n =<br />

− F 2<br />

<strong>12</strong><br />

<br />

1 EG<br />

(rv × ruu) · n = −<br />

− F 2<br />

11<br />

<br />

2 EG<br />

(ru × rvv) · n =<br />

− F 2<br />

22<br />

<br />

1 EG<br />

(rv × ruv) · n = −<br />

− F 2<br />

21<br />

<br />

1 EG<br />

(rv × rvv) · n = −<br />

− F 2<br />

22<br />

(ru × rv) · n = EG − F 2<br />

and then derive the formula for the geodesic curvature given by equation (1.5.48).<br />

Hint:(n × T ) · d T<br />

ds =( T × d T<br />

ds ) · n and aαδ ]βγ,δ]=<br />

α<br />

βγ<br />

<br />

.<br />

163


164<br />

◮ 11. Verify the equation (1.5.39) which shows that the normal curvature directions are orthogonal. i.e.<br />

verify that Gλ1λ2 + F (λ1 + λ2)+E =0.<br />

◮ <strong>12</strong>. Verify that δ βγ<br />

στ δ ωα<br />

λν Rωαβγ =4Rλνστ .<br />

◮ 13. Find the first fundamental form and unit normal to the surface defined by z = f(x, y).<br />

◮ 14. Verify<br />

Ai,jk − Ai,kj = AσR σ .ijk<br />

where<br />

R σ .ijk = ∂<br />

∂xj <br />

σ<br />

−<br />

ik<br />

∂<br />

∂xk <br />

σ n σ n σ<br />

+<br />

−<br />

.<br />

ij ik nj ij nk<br />

which is sometimes written<br />

<br />

<br />

∂<br />

Rinjk = ∂x<br />

<br />

<br />

j<br />

∂<br />

∂xk [nj, k]<br />

<br />

<br />

<br />

<br />

<br />

[nk, i] +<br />

<br />

<br />

s<br />

nj<br />

<br />

[ij, s]<br />

<br />

s <br />

<br />

nk <br />

<br />

[ik, s] <br />

◮ 15. For Rijkl = giσR σ .jkl show<br />

which is sometimes written<br />

◮ 16. Show<br />

Rijkl = 1<br />

2<br />

Rinjk = ∂<br />

<br />

∂<br />

s<br />

s<br />

[nk, i] − [nj, i]+[ik, s] − [ij, s]<br />

∂xj ∂xk nj<br />

nk<br />

R σ .ijk =<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

∂<br />

∂xj <br />

σ<br />

ij<br />

∂<br />

∂xk σ<br />

ik<br />

<br />

<br />

<br />

<br />

<br />

+ <br />

<br />

<br />

<br />

<br />

n<br />

ik<br />

<br />

σ<br />

nk<br />

<br />

n <br />

<br />

ij <br />

<br />

σ <br />

nj <br />

2 ∂ gil<br />

∂xj∂xk − ∂2gjl ∂xi∂xk − ∂2gik ∂xj∂xl + ∂2gjk ∂xi∂xl <br />

+ g αβ ([jk,β][il, α] − [jl,β][ik, α]) .<br />

◮ 17. Use the results from problem 15 to show<br />

(i) Rjikl = −Rijkl, (ii) Rijlk = −Rijkl, (iii) Rklij = Rijkl<br />

Hence, the tensor Rijkl is skew-symmetric in the indices i, j and k, l. Also the tensor Rijkl is symmetric with<br />

respect to the (ij) and(kl) pair of indices.<br />

◮ 18. Verify the following cyclic properties of the Riemann Christoffel symbol:<br />

(i) Rnijk + Rnjki + Rnkij = 0 first index fixed<br />

(ii) Rinjk + Rjnki + Rknij = 0 second index fixed<br />

(iii) Rijnk + Rjkni + Rkinj = 0 third index fixed<br />

(iv) Rikjn + Rkjin + Rjikn = 0 fourth index fixed<br />

◮ 19. By employing the results from the previous problems, show all components of the form:<br />

Riijk, Rinjj, Riijj, Riiii, (no summation on i or j) must be zero.


◮ 20. Find the number of independent components associated with the Riemann Christoffel tensor<br />

Rijkm, i,j,k,m=1, 2,...,N.There are N 4 components to examine in an N−dimensional space. Many of<br />

these components are zero and many of the nonzero components are related to one another by symmetries<br />

or the cyclic properties. Verify the following cases:<br />

CASE I We examine components of the form Rinin, i = n with no summation of i or n. The first index<br />

can be chosen in N ways and therefore with i = n the second index can be chosen in N − 1 ways. Observe<br />

that Rinin = Rnini, (no summation on i or n) and so one half of the total combinations are repeated. This<br />

leaves M1 = 1<br />

2 N(N − 1) components of the form Rinin. The quantity M1 can also be thought of as the<br />

number of distinct pairs of indices (i, n).<br />

CASE II We next examine components of the form Rinji, i = n = j where there is no summation on<br />

the index i. We have previously shown that the first pair of indices can be chosen in M1 ways. Therefore,<br />

the third index can be selected in N − 2 ways and consequently there are M2 = 1<br />

2N(N − 1)(N − 2) distinct<br />

components of the form Rinji with i = n = j.<br />

CASE III Next examine components of the form Rinjk where i = n = j = k. From CASE I the first pairs<br />

of indices (i, n) can be chosen in M1 ways. Taking into account symmetries, it can be shown that the second<br />

pair of indices can be chosen in 1<br />

1<br />

2 (N − 2)(N − 3) ways. This implies that there are 4N(N − 1)(N − 2)(N − 3)<br />

ways of choosing the indices i, n, j and k with i = n = j = k. By symmetry the pairs (i, n) and(j, k) canbe<br />

interchanged and therefore only one half of these combinations are distinct. This leaves<br />

1<br />

N(N − 1)(N − 2)(N − 3)<br />

8<br />

distinct pairs of indices. Also from the cyclic relations we find that only two thirds of the above components<br />

are distinct. This produces<br />

N(N − 1)(N − 2)(N − 3)<br />

M3 =<br />

<strong>12</strong><br />

distinct components of the form Rinjk with i = n = j = k.<br />

Adding the above components from each case we find there are<br />

distinct and independent components.<br />

Verify the entries in the following table:<br />

M4 = M1 + M2 + M3 = N 2 (N 2 − 1)<br />

<strong>12</strong><br />

Dimension of space N 1 2 3 4 5<br />

Number of components N 4 1 16 81 256 625<br />

M4 = Independent components of Rijkm 0 1 6 20 50<br />

Note 1: A one dimensional space can not be curved and all one dimensional spaces are Euclidean. (i.e. if we have<br />

an element of arc length squared given by ds2 = f(x)(dx) 2 , we can make the coordinate transformation<br />

<br />

2 2 f(x)dx = du and reduce the arc length squared to the form ds = du .)<br />

Note 2: In a two dimensional space, the indices can only take on the values 1 and 2. In this special case there<br />

are 16 possible components. It can be shown that the only nonvanishing components are:<br />

R<strong>12</strong><strong>12</strong> = −R<strong>12</strong>21 = −R21<strong>12</strong> = R2<strong>12</strong>1.<br />

165


166<br />

For these nonvanishing components only one independent component exists. By convention, the component<br />

R<strong>12</strong><strong>12</strong> is selected as the single independent component and all other nonzero components are<br />

expressed in terms of this component.<br />

Find the nonvanishing independent components Rijkl for i, j, k, l =1, 2, 3, 4andshowthat<br />

R<strong>12</strong><strong>12</strong><br />

R1313<br />

R2323<br />

R1414<br />

R2424<br />

R3434<br />

R<strong>12</strong>31<br />

R1421<br />

R1341<br />

R2132<br />

can be selected as the twenty independent components.<br />

R2142<br />

R2342<br />

R3213<br />

R3243<br />

R3143<br />

R4<strong>12</strong>4<br />

R4314<br />

R4234<br />

R1324<br />

R1432<br />

◮ 21.<br />

(a) For N =2showR<strong>12</strong><strong>12</strong> is the only nonzero independent component and<br />

R<strong>12</strong><strong>12</strong> = R2<strong>12</strong>1 = −R<strong>12</strong>21 = −R21<strong>12</strong>.<br />

(b) Show that on the surface of a sphere of radius r0 we have R<strong>12</strong><strong>12</strong> = r2 0 sin 2 θ.<br />

◮ 22. Show for N =2that<br />

R<strong>12</strong><strong>12</strong> = R<strong>12</strong><strong>12</strong>J 2 = R<strong>12</strong><strong>12</strong><br />

<br />

<br />

<br />

∂x<br />

<br />

∂x<br />

<br />

◮ 23. Define Rij = Rs .ijs as the Ricci tensor and Gij = Ri 1<br />

j − 2δi jR as the Einstein tensor, where Ri j = gikRkj and R = Ri i . Show that<br />

(a) Rjk = g ab Rjabk<br />

(b) Rij = ∂2 log √ g<br />

∂xi −<br />

∂xj (c) R i ijk =0<br />

√<br />

b ∂ log g ∂<br />

−<br />

ij ∂xb ∂xa <br />

a b a<br />

+<br />

ij ia jb<br />

◮ 24. By employing the results from the previous problem show that in the case N =2wehave<br />

R11<br />

g11<br />

= R22<br />

g22<br />

= R<strong>12</strong><br />

g<strong>12</strong><br />

= − R<strong>12</strong><strong>12</strong><br />

g<br />

where g is the determinant of gij.<br />

◮ 25. Consider the case N =2wherewehaveg<strong>12</strong> = g21 = 0 and show that<br />

(a) R<strong>12</strong> = R21 =0<br />

(b) R11g22 = R22g11 = R<strong>12</strong>21<br />

2<br />

(c) R = 2R<strong>12</strong>21<br />

g11g22<br />

(d) Rij = 1<br />

2 Rgij, where R = g ij Rij<br />

The scalar invariant R is known as the Einstein curvature of the surface and the tensor G i j = Ri j<br />

1 − 2δi jR is<br />

known as the Einstein tensor.<br />

◮ 26. For N =3showthatR<strong>12</strong><strong>12</strong>,R1313,R2323,R<strong>12</strong>13,R2<strong>12</strong>3,R3132 are independent components of the<br />

Riemann Christoffel tensor.


a11 0<br />

◮ 27. For N =2andaαβ =<br />

show that<br />

0 a22<br />

◮ 28. For N =2andaαβ =<br />

K = 1<br />

2 √ <br />

∂<br />

a ∂u1 K = R<strong>12</strong><strong>12</strong><br />

a<br />

a<strong>12</strong><br />

a11<br />

a11 a<strong>12</strong><br />

√ a<br />

1<br />

= −<br />

2 √ <br />

∂<br />

a ∂u1 <br />

1<br />

√a<br />

a21 a22<br />

∂a11 1<br />

− √<br />

∂u2 a<br />

<br />

show that<br />

∂a22<br />

∂u1 <br />

+ ∂<br />

∂u2 <br />

1<br />

√a<br />

∂a22<br />

∂u1 <br />

+ ∂<br />

∂u2 <br />

2 ∂a<strong>12</strong> 1<br />

√a − √<br />

∂u1 a<br />

∂a11<br />

∂u2 <br />

.<br />

∂a11 a<strong>12</strong><br />

−<br />

∂u2 a11<br />

√ a<br />

∂a11<br />

∂u1 <br />

.<br />

Check your results by setting a<strong>12</strong> = a21 = 0 and comparing this answer with that given in the problem 27.<br />

◮ 29. Write out the Frenet-Serret formulas (1.5.1<strong>12</strong>)(1.5.113) for surface curves in terms of Christoffel<br />

symbols of the second kind.<br />

◮ 30.<br />

(a) Use the fact that for n =2wehaveR<strong>12</strong><strong>12</strong> = R2<strong>12</strong>1 = −R21<strong>12</strong> = −R<strong>12</strong>21 together with eαβ, eαβ the two<br />

dimensional alternating tensors to show that the equation (1.5.110) can be written as<br />

Rαβγδ = Kɛαβɛγδ where ɛαβ = √ aeαβ and ɛ αβ = 1<br />

√ a e αβ<br />

are the corresponding epsilon tensors.<br />

(b) Show that from the result in part (a) we obtain 1<br />

4 Rαβγδɛ αβ ɛ γδ = K.<br />

Hint: See equations (1.3.82),(1.5.93) and (1.5.94).<br />

◮ 31. Verify the result given by the equation (1.5.100).<br />

◮ 32. Show that a αβ cαβ =4H 2 − 2K.<br />

◮ 33. Find equations for the principal curvatures associated with the surface<br />

x = u, y = v, z = f(u, v).<br />

◮ 34. Geodesics on a sphere Let (θ, φ) denote the surface coordinates of the sphere of radius ρ defined<br />

by the parametric equations<br />

x = ρ sin θ cos φ, y = ρ sin θ sin φ, z = ρ cos θ. (1)<br />

Consider also a plane which passes through the origin with normal having the direction numbers (n1,n2,n3).<br />

This plane is represented by n1x+n2y +n3z = 0 and intersects the sphere in a great circle which is described<br />

by the relation<br />

n1 sin θ cos φ + n2 sin θ sin φ + n3 cos θ =0. (2)<br />

This is an implicit relation between the surface coordinates θ, φ which describes the great circle lying on the<br />

sphere. We can write this later equation in the form<br />

n1 cos φ + n2 sin φ = −n3<br />

tan θ<br />

(3)<br />

167


168<br />

and in the special case where n1 =cosβ, n2 =sinβ,n3 = − tan α is expressible in the form<br />

cos(φ − β) =<br />

tan α<br />

tan θ<br />

or φ − β =cos −1<br />

<br />

tan α<br />

. (4)<br />

tan θ<br />

The above equation defines an explicit relationship between the surface coordinates which defines a great<br />

circle on the sphere. The arc length squared relation satisfied by the surface coordinates together with the<br />

equation obtained by differentiating equation (4) with respect to arc length s gives the relations<br />

sin 2 θ dφ<br />

ds =<br />

tan α<br />

<br />

1 − tan2 α<br />

tan 2 θ<br />

dθ<br />

ds<br />

ds 2 = ρ 2 dθ 2 + ρ 2 sin 2 θdφ 2<br />

The above equations (1)-(6) are needed to consider the following problem.<br />

(a) Show that the differential equations defining the geodesics on the surface of a sphere (equations (1.5.51))<br />

are<br />

d2θ − sin θ cos θ<br />

ds2 (b) Multiply equation (8) by sin 2 θ and integrate to obtain<br />

where c1 is a constant of integration.<br />

(c) Multiply equation (7) by dθ<br />

ds<br />

where c 2 2<br />

is a constant of integration.<br />

(5)<br />

(6)<br />

2 dφ<br />

=0 (7)<br />

ds<br />

d2φ dθ dφ<br />

+2cotθ =0 (8)<br />

ds2 ds ds<br />

sin 2 θ dφ<br />

ds<br />

(d) Use the equations (5)(6) to show that c2 =1/ρ and c1 =<br />

(e) Show that equations (9) and (10) imply that<br />

= c1<br />

(9)<br />

and use the result of equation (9) to show that an integration produces<br />

2 dθ<br />

=<br />

ds<br />

−c21 sin 2 θ + c22 (10)<br />

tan α<br />

tan θ<br />

dφ<br />

dθ<br />

= tan α<br />

tan 2 θ<br />

sin α<br />

ρ .<br />

sec 2 θ<br />

<br />

1 − tan2 α<br />

tan 2 θ<br />

and making the substitution u = this equation can be integrated to obtain the equation (4). We<br />

can now expand the equation (4) and express the results in terms of x, y, z to obtain the equation (3).<br />

This produces a plane which intersects the sphere in a great circle. Consequently, the geodesics on a<br />

sphere are great circles.


◮ 35. Find the differential equations defining the geodesics on the surface of a cylinder.<br />

◮ 36. Find the differential equations defining the geodesics on the surface of a torus. (See problem 13,<br />

Exercise 1.3)<br />

◮ 37. Find the differential equations defining the geodesics on the surface of revolution<br />

x = r cos φ, y = r sin φ, z = f(r).<br />

Note the curve z = f(x) gives a profile of the surface. The curves r = Constant are the parallels, while the<br />

curves φ = Constant are the meridians of the surface and<br />

ds 2 =(1+f ′2 ) dr 2 + r 2 dφ 2 .<br />

◮ 38. Find the unit normal and tangent plane to an arbitrary point on the right circular cone<br />

x = u sin α cos φ, y = u sin α sin φ, z = u cos α.<br />

This is a surface of revolution with r = u sin α and f(r) =r cot α with α constant.<br />

◮ 39. Let s denote arc length and assume the position vector r(s) is analytic about a point s0. Show that<br />

the Taylor series r(s) =r(s0)+hr ′ (s0)+ h2<br />

2! r ′′ (s0)+ h3<br />

3! r ′′′ (s0)+··· about the point s0, withh = s − s0 is<br />

given by r(s) =r(s0)+h T + 1<br />

2 κh2 N + 1<br />

6 h3 (−κ 2 T + κ ′ N + κτ B)+··· which is obtained by differentiating<br />

the Frenet formulas.<br />

◮ 40.<br />

(a) Show that the circular helix defined by x = a cos t, y = a sin t, z = bt with a, b constants, has the<br />

property that any tangent to the curve makes a constant angle with the line defining the z-axis.<br />

(i.e. T · e3 =cosα = constant.)<br />

(b) Show also that N · e3 = 0 and consequently e3 is parallel to the rectifying plane, which implies that<br />

e3 = T cos α + B sin α.<br />

(c) Differentiate the result in part (b) and show that κ/τ =tanαis a constant.<br />

◮ 41. Consider a space curve xi = xi(s) in Cartesian coordinates.<br />

<br />

<br />

(a) Show that κ = d T<br />

<br />

<br />

=<br />

ds<br />

x ′ ix′ i<br />

(b) Show that τ = 1<br />

κ 2 eijkx ′ i x′′<br />

j x′′′<br />

k . Hint: Consider r ′ · r ′′ × r ′′′<br />

◮ 42.<br />

(a) Find the direction cosines of a normal to a surface z = f(x, y).<br />

(b) Find the direction cosines of a normal to a surface F (x, y, z) =0.<br />

(c) Find the direction cosines of a normal to a surface x = x(u, v),y = y(u, v),z = z(u, v).<br />

◮ 43. Show that for a smooth surface z = f(x, y) the Gaussian curvature at a point on the surface is given<br />

by<br />

K = fxxfyy − f 2 xy<br />

(f 2 x + f 2 .<br />

y +1)2<br />

169


170<br />

◮ 44. Show that for a smooth surface z = f(x, y) themeancurvatureatapointonthesurfaceisgivenby<br />

H = (1 + f 2 y )fxx − 2fxfyfxy +(1+f2 x )fyy<br />

2(f 2 x + f 2 y +1) 3/2<br />

.<br />

◮ 45. Express the Frenet-Serret formulas (1.5.13) in terms of Christoffel symbols of the second kind.<br />

◮ 46. Verify the relation (1.5.106).<br />

◮ 47. In Vn assume that Rij = ρgij and show that ρ = R<br />

n where R = gijRij. This result is known as<br />

Einstein’s gravitational equation at points where matter is present. It is analogous to the Poisson equation<br />

∇2V = ρ from the Newtonian theory of gravitation.<br />

◮ 48. In Vn assume that Rijkl = K(gikgjl − gilgjk) andshowthatR = Kn(1 − n). (Hint: See problem 23.)<br />

◮ 49. Assume gij =0fori= j and verify the following.<br />

(a) Rhijk =0forh= i = j = k<br />

(b) Rhiik = √ 2 ∂<br />

gii<br />

√ gii<br />

∂xh∂xk − ∂√gii ∂xh ∂ log √ ghh<br />

∂xk − ∂√gii ∂xk ∂ log √ gkk<br />

∂xh <br />

for h, i, k unequal.<br />

(c) Rhiih = √ ⎡<br />

√ ⎢<br />

gii ghh ⎣ ∂<br />

∂xh <br />

1 ∂ √ gii<br />

∂xh <br />

+ ∂<br />

∂xi <br />

1 ∂ √ ghh<br />

∂xi n ∂<br />

+<br />

√ gii<br />

∂xm ∂ √ ghh<br />

∂xm ⎤<br />

⎥<br />

⎦ where h = i.<br />

√ghh<br />

√gii<br />

m=1<br />

m=h m=i<br />

◮ 50. Consider a surface of revolution where x = r cos θ, y = r sin θ and z = f(r) is a given function of r.<br />

(a) Show in this V2 we have ds 2 =(1+(f ′ ) 2 )dr 2 + r 2 dθ 2 where ′ = d<br />

ds .<br />

(b) Show the geodesic equations in this V2 are<br />

d2r ds2 + f ′ f ′′<br />

1+(f ′ ) 2<br />

d2θ 2 dθ dr<br />

+<br />

ds2 r ds ds =0<br />

2 dr<br />

−<br />

ds<br />

(c) Solve the second equation in part (b) to obtain dθ<br />

ds<br />

r<br />

1+(f ′ ) 2<br />

dθ = ± a1+(f ′ ) 2<br />

r √ r2 dr which theoretically can be integrated.<br />

− a2 2 dθ<br />

=0<br />

ds<br />

a<br />

= . Substitute this result for ds in part (a) to show<br />

r2


PART 2: INTRODUCTION TO CONTINUUM MECHANICS<br />

In the following sections we develop some applications of tensor calculus in the areas of dynamics,<br />

elasticity, fluids and electricity and magnetism. We begin by first developing generalized expressions for the<br />

vector operations of gradient, divergence, and curl. Also generalized expressions for other vector operators<br />

are considered in order that tensor equations can be converted to vector equations. We construct a table to<br />

aid in the translating of generalized tensor equations to vector form and vice versa.<br />

The basic equations of continuum mechanics are developed in the later sections. These equations are<br />

developed in both Cartesian and generalized tensor form and then converted to vector form.<br />

§2.1 TENSOR NOTATION FOR SCALAR AND VECTOR QUANTITIES<br />

We consider the tensor representation of some vector expressions. Our goal is to develop the ability to<br />

convert vector equations to tensor form as well as being able to represent tensor equations in vector form.<br />

In this section the basic equations of continuum mechanics are represented using both a vector notation and<br />

the indicial notation which focuses attention on the tensor components. In order to move back and forth<br />

between these notations, the representation of vector quantities in tensor form is now considered.<br />

Gradient<br />

For Φ = Φ(x 1 ,x 2 ,...,x N ) a scalar function of the coordinates x i ,i =1,...,N , the gradient of Φ is<br />

defined as the covariant vector<br />

The contravariant form of the gradient is<br />

Φ,i = ∂Φ<br />

, i =1,...,N. (2.1.1)<br />

∂xi g im Φ,m. (2.1.2)<br />

Note, if C i = g im Φ,m, i =1, 2, 3 are the tensor components of the gradient then in an orthogonal coordinate<br />

system we will have<br />

C 1 = g 11 Φ,1, C 2 = g 22 Φ,2, C 3 = g 33 Φ,3.<br />

We note that in an orthogonal coordinate system that gii =1/h2 i , (no sum on i), i =1, 2, 3 and hence<br />

replacing the tensor components by their equivalent physical components there results the equations<br />

C(1)<br />

h1<br />

= 1<br />

h 2 1<br />

∂Φ<br />

,<br />

∂x1 C(2)<br />

h2<br />

= 1<br />

h 2 2<br />

∂Φ<br />

,<br />

∂x2 Simplifying, we find the physical components of the gradient are<br />

C(1) = 1<br />

h1<br />

∂Φ<br />

1<br />

, C(2) =<br />

∂x1 h2<br />

C(3)<br />

h3<br />

= 1<br />

h 2 3<br />

∂Φ<br />

1<br />

, C(3) =<br />

∂x2 h3<br />

∂Φ<br />

.<br />

∂x3 ∂Φ<br />

.<br />

∂x3 These results are only valid when the coordinate system is orthogonal and gij =0fori = j and gii = h 2 i ,<br />

with i =1, 2, 3, and where i is not summed.<br />

171


172<br />

Divergence<br />

The divergence of a contravariant tensor Ar is obtained by taking the covariant derivative with respect<br />

to xk and then performing a contraction. This produces<br />

div A r = A r ,r. (2.1.3)<br />

Still another form for the divergence is obtained by simplifying the expression (2.1.3). The covariant derivative<br />

can be represented<br />

A r <br />

∂Ar r<br />

,k = + A<br />

∂xk mk<br />

m .<br />

Upon contracting the indices r and k and using the result from Exercise 1.4, problem 13, we obtain<br />

A r ,r<br />

∂Ar 1 ∂(<br />

= +<br />

∂xr √<br />

g<br />

√ g)<br />

Am<br />

∂xm <br />

√g ∂Ar ∂xr + Ar ∂√g ∂xr <br />

A r ,r = 1<br />

√<br />

g<br />

A r ,r = 1 ∂<br />

√<br />

g ∂xr (√gA r ) .<br />

(2.1.4)<br />

EXAMPLE 2.1-1. (Divergence) Find the representation of the divergence of a vector Ar in spherical<br />

coordinates (ρ, θ, φ). Solution: In spherical coordinates we have<br />

x 1 = ρ, x 2 = θ, x 3 = φ with gij =0 for i = j and<br />

g11 = h 2 1 =1, g22 = h 2 2 = ρ2 , g33 = h 2 3 = ρ2 sin 2 θ.<br />

The determinant of gij is g = |gij| = ρ 4 sin 2 θ and √ g = ρ 2 sin θ. Employing the relation (2.1.4) we find<br />

div A r = 1<br />

<br />

∂<br />

√<br />

g ∂x1 (√gA 1 )+ ∂<br />

∂x2 (√gA 2 )+ ∂<br />

∂x3 (√gA 3 <br />

) .<br />

In terms of the physical components this equation becomes<br />

div A r = 1<br />

<br />

∂<br />

√<br />

g ∂ρ (√g A(1)<br />

)+<br />

h1<br />

∂<br />

∂θ (√g A(2)<br />

)+<br />

h2<br />

∂<br />

∂φ (√g A(3)<br />

<br />

) .<br />

h3<br />

By using the notation<br />

A(1) = Aρ, A(2) = Aθ, A(3) = Aφ<br />

for the physical components, the divergence can be expressed in either of the forms:<br />

∂<br />

∂ρ (ρ2 sin θAρ)+ ∂<br />

div A r 1<br />

=<br />

ρ2 sin θ<br />

∂θ (ρ2 sin θ Aθ<br />

ρ<br />

div A r = 1<br />

ρ2 ∂<br />

∂ρ (ρ2Aρ)+ 1 ∂<br />

1 ∂Aφ<br />

(sin θAθ)+<br />

ρ sin θ ∂θ ρ sin θ ∂φ .<br />

∂<br />

)+<br />

∂φ (ρ2 sin θ Aφ<br />

ρ sin θ )<br />

<br />

or


Curl<br />

The contravariant components of the vector C =curl A are represented<br />

In expanded form this representation becomes:<br />

C 1 = 1<br />

<br />

∂A3 ∂A2<br />

√ −<br />

g ∂x2 ∂x3 <br />

C 2 = 1<br />

√ g<br />

C 3 = 1<br />

√ g<br />

C i = ɛ ijk Ak,j. (2.1.5)<br />

<br />

∂A1 ∂A3<br />

−<br />

∂x3 ∂x1 <br />

∂A2 ∂A1<br />

−<br />

∂x1 ∂x2 <br />

<br />

.<br />

(2.1.6)<br />

EXAMPLE 2.1-2. (Curl) Find the representation for the components of curl A in spherical coordinates<br />

(ρ, θ, φ).<br />

Solution: In spherical coordinates we have :x 1 = ρ, x 2 = θ, x 3 = φ with gij =0fori = j and<br />

g11 = h 2 1 =1, g22 = h 2 2 = ρ2 , g33 = h 2 3 = ρ2 sin 2 θ.<br />

The determinant of gij is g = |gij| = ρ 4 sin 2 θ with √ g = ρ 2 sin θ. The relations (2.1.6) are tensor equations<br />

representing the components of the vector curl A. To find the components of curl A in spherical components<br />

we write the equations (2.1.6) in terms of their physical components. These equations take on the form:<br />

We employ the notations<br />

C(1)<br />

h1<br />

C(2)<br />

h2<br />

C(3)<br />

h3<br />

= 1<br />

<br />

∂<br />

√<br />

g ∂θ (h3A(3)) − ∂<br />

∂φ (h2A(2))<br />

<br />

= 1<br />

<br />

∂<br />

√<br />

g ∂φ (h1A(1)) − ∂<br />

∂ρ (h3A(3))<br />

<br />

= 1<br />

<br />

∂<br />

√<br />

g ∂ρ (h2A(2)) − ∂<br />

∂θ (h1A(1))<br />

<br />

.<br />

C(1) = Cρ, C(2) = Cθ, C(3) = Cφ, A(1) = Aρ, A(2) = Aθ, A(3) = Aφ<br />

(2.1.7)<br />

to denote the physical components, and find the components of the vector curl A, in spherical coordinates,<br />

are expressible in the form:<br />

1<br />

Cρ =<br />

ρ2 <br />

∂<br />

sin θ ∂θ (ρ sin θAφ) − ∂<br />

∂φ (ρAθ)<br />

<br />

Cθ = 1<br />

<br />

∂<br />

ρ sin θ ∂φ (Aρ) − ∂<br />

<br />

(ρ sin θAφ)<br />

∂ρ<br />

Cφ = 1<br />

<br />

∂<br />

ρ ∂ρ (ρAθ) − ∂<br />

∂θ (Aρ)<br />

<br />

.<br />

(2.1.8)<br />

173


174<br />

Laplacian<br />

The Laplacian ∇ 2 U has the contravariant form<br />

∇ 2 U = g ij U,ij =(g ij U,i),j =<br />

Expanding this expression produces the equations:<br />

∇ 2 U = ∂<br />

∂xj <br />

ij ∂U<br />

g<br />

∂xi <br />

im ∂U<br />

+ g<br />

∂xi <br />

j<br />

mj<br />

∇ 2 U = ∂<br />

∂xj <br />

ij ∂U<br />

g<br />

∂xi <br />

+ 1 ∂<br />

√<br />

g<br />

√ g ∂U<br />

gij<br />

∂xj ∂xi ∇ 2 U = 1<br />

<br />

√g ∂<br />

√<br />

g ∂xj <br />

ij ∂U<br />

g<br />

∂xi <br />

ij ∂U<br />

+ g<br />

∂xi ∂ √ g<br />

∂xj <br />

∇ 2 U = 1 ∂<br />

√<br />

g ∂xj <br />

√ggij ∂U<br />

∂xi <br />

.<br />

In orthogonal coordinates we have g ij =0fori = j and<br />

and so (2.1.10) when expanded reduces to the form<br />

∇ 2 U =<br />

1<br />

h1h2h3<br />

g11 = h 2 1 , g22 = h 2 2 , g33 = h 2 3<br />

<br />

ij ∂U<br />

g<br />

∂xi <br />

. (2.1.9)<br />

,j<br />

(2.1.10)<br />

<br />

∂<br />

∂x1 <br />

h2h3 ∂U<br />

h1 ∂x1 <br />

+ ∂<br />

∂x2 <br />

h1h3 ∂U<br />

h2 ∂x2 <br />

+ ∂<br />

∂x3 <br />

h1h2 ∂U<br />

h3 ∂x3 <br />

. (2.1.11)<br />

This representation is only valid in an orthogonal system of coordinates.<br />

EXAMPLE 2.1-3. (Laplacian) Find the Laplacian in spherical coordinates.<br />

Solution: Utilizing the results given in the previous example we find the Laplacian in spherical coordinates<br />

has the form<br />

∇ 2 1<br />

U =<br />

ρ2 <br />

∂<br />

ρ<br />

sin θ ∂ρ<br />

2 sin θ ∂U<br />

<br />

+<br />

∂ρ<br />

∂<br />

<br />

sin θ<br />

∂θ<br />

∂U<br />

<br />

+<br />

∂θ<br />

∂<br />

This simplifies to<br />

<br />

1 ∂U<br />

.<br />

∂φ sin θ ∂φ<br />

(2.1.<strong>12</strong>)<br />

∇ 2 U = ∂2U 2 ∂U 1<br />

+ +<br />

∂ρ2 ρ ∂ρ ρ2 ∂2U cot θ<br />

+<br />

∂θ2 ρ2 ∂U<br />

∂θ +<br />

1<br />

ρ 2 sin 2 θ<br />

The table 1 gives the vector and tensor representation for various quantities of interest.<br />

∂2U . (2.1.13)<br />

∂φ2


VECTOR GENERAL TENSOR CARTESIAN TENSOR<br />

A A i or Ai Ai<br />

A · B A i Bi = gijA i B j = AiB i<br />

A i Bi = g ij AiBj<br />

C = A × B C i = 1<br />

√ g e ijk AjBk<br />

AiBi<br />

Ci = eijkAjBk<br />

∇ Φ = grad Φ g im Φ,m Φ,i = ∂Φ<br />

∂x i<br />

∇· A =div A g mn Am,n = A r ,r = 1 ∂<br />

√<br />

g ∂xr (√gA r ) Ai,i = ∂Ai<br />

∂xi ∇× A = C =curl A C i = ɛ ijk ∂Ak<br />

Ak,j Ci = eijk<br />

∂xj ∇ 2 U g mn U ,mn = 1<br />

√ g<br />

∂<br />

∂xj <br />

√ggij ∂U<br />

∂xi <br />

∂<br />

∂xi <br />

∂U<br />

∂xi <br />

C =(A ·∇) B C i = A m B i ,m Ci = Am<br />

C = A(∇· B) C i = A i B j<br />

,j<br />

∂Bi<br />

∂xm ∂Bm<br />

Ci = Ai<br />

∂xm C = ∇ 2 A i jm i<br />

C = g A ,mj or Ci = g jm Ai,mj Ci = ∂<br />

∂xm <br />

∂Ai<br />

∂xm <br />

<br />

A ·∇ φ g im A i φ ,m Aiφ,i<br />

<br />

∇ ∇· <br />

A<br />

<br />

∇× ∇× <br />

A<br />

g im A r <br />

,r ,m<br />

ɛijkg jm ɛ kst <br />

At,s ,m<br />

Table 1 Vector and tensor representations.<br />

∂ 2 Aj<br />

∂xj∂xi<br />

∂ 2 Ar<br />

∂xi∂xr<br />

− ∂2 Ai<br />

∂xj∂xj<br />

175


176<br />

EXAMPLE 2.1-4. (Maxwell’s equations) In the study of electrodynamics there arises the following<br />

vectors and scalars:<br />

E =Electric force vector, [ E]=Newton/coulomb<br />

B =Magnetic force vector, [ B]=Weber/m 2<br />

D =Displacement vector, [ D]=coulomb/m 2<br />

H =Auxilary magnetic force vector, [ H]=ampere/m<br />

J =Free current density, [ J]=ampere/m 2<br />

ϱ =free charge density, [ϱ] =coulomb/m 3<br />

The above quantities arise in the representation of the following laws:<br />

Faraday’s Law This law states the line integral of the electromagnetic force around a loop is proportional<br />

to the rate of flux of magnetic induction through the loop. This gives rise to the first electromagnetic field<br />

equation:<br />

∇× E = − ∂ B<br />

or ɛ<br />

∂t<br />

ijk Ek,j = − ∂Bi<br />

. (2.1.15)<br />

∂t<br />

Ampere’s Law This law states the line integral of the magnetic force vector around a closed loop is<br />

proportional to the sum of the current through the loop and the rate of flux of the displacement vector<br />

through the loop. This produces the second electromagnetic field equation:<br />

∇× H = J + ∂ D<br />

∂t<br />

or ɛ ijk Hk,j = J i + ∂Di<br />

. (2.1.16)<br />

∂t<br />

Gauss’s Law for Electricity This law states that the flux of the electric force vector through a closed<br />

surface is proportional to the total charge enclosed by the surface. This results in the third electromagnetic<br />

field equation:<br />

∇· D = ϱ or<br />

1 ∂<br />

√<br />

g ∂xi √ i<br />

gD = ϱ. (2.1.17)<br />

Gauss’s Law for Magnetism This law states the magnetic flux through any closed volume is zero. This<br />

produces the fourth electromagnetic field equation:<br />

∇· B =0 or<br />

1 ∂<br />

√<br />

g ∂xi √ i<br />

gB =0. (2.1.18)<br />

The four electromagnetic field equations are referred to as Maxwell’s equations. These equations arise<br />

in the study of electrodynamics and can be represented in other forms. These other forms will depend upon<br />

such things as the material assumptions and units of measurements used. Note that the tensor equations<br />

(2.1.15) through (2.1.18) are representations of Maxwell’s equations in a form which is independent of the<br />

coordinate system chosen.<br />

In applications, the tensor quantities must be expressed in terms of their physical components. In a<br />

general orthogonal curvilinear coordinate system we will have<br />

g11 = h 2 1 , g22 = h 2 2 , g33 = h 2 3 , and gij =0 for i = j.<br />

This produces the result √ g = h1h2h3. Further, if we represent the physical components of<br />

Di,Bi,Ei,Hi by D(i),B(i),E(i), and H(i)


the Maxwell equations can be represented by the equations in table 2. The tables 3, 4 and 5 are the<br />

representation of Maxwell’s equations in rectangular, cylindrical, and spherical coordinates. These latter<br />

tables are special cases associated with the more general table 2.<br />

1<br />

h1h2h3<br />

1<br />

h1h2h3<br />

1<br />

h1h2h3<br />

1<br />

h1h2h3<br />

1<br />

h1h2h3<br />

1<br />

h1h2h3<br />

1<br />

h1h2h3<br />

1<br />

h1h2h3<br />

<br />

∂<br />

∂x2 (h3E(3)) − ∂<br />

<br />

(h2E(2)) = −<br />

∂x3 1<br />

<br />

∂<br />

∂x3 (h1E(1)) − ∂<br />

<br />

(h3E(3)) = −<br />

∂x1 1<br />

<br />

∂<br />

= − 1<br />

∂x1 (h2E(2)) − ∂<br />

(h1E(1))<br />

∂x2 h1<br />

h2<br />

h3<br />

∂B(1)<br />

∂t<br />

∂B(2)<br />

∂t<br />

∂B(3)<br />

∂t<br />

<br />

∂<br />

∂x2 (h3H(3)) − ∂<br />

<br />

(h2H(2)) =<br />

∂x3 J(1)<br />

+<br />

h1<br />

1 ∂D(1)<br />

h1 ∂t<br />

<br />

∂<br />

∂x3 (h1H(1)) − ∂<br />

<br />

(h3H(3)) =<br />

∂x1 J(2)<br />

+<br />

h2<br />

1 ∂D(2)<br />

h2 ∂t<br />

<br />

∂<br />

∂x1 (h2H(2)) − ∂<br />

<br />

(h1H(1)) =<br />

∂x2 J(3)<br />

+<br />

h3<br />

1 ∂D(3)<br />

h3 ∂t<br />

<br />

∂<br />

∂x1 <br />

<br />

D(1)<br />

h1h2h3 + ∂<br />

∂x2 <br />

<br />

D(2)<br />

h1h2h3 + ∂<br />

∂x3 <br />

<br />

D(3)<br />

h1h2h3 = ϱ<br />

h1<br />

<br />

∂<br />

∂x1 <br />

<br />

B(1)<br />

h1h2h3 + ∂<br />

∂x2 <br />

<br />

B(2)<br />

h1h2h3 + ∂<br />

∂x3 <br />

<br />

B(3)<br />

h1h2h3 =0<br />

h1<br />

Table 2 Maxwell’s equations in generalized orthogonal coordinates.<br />

Note that all the tensor components have been replaced by their physical components.<br />

h2<br />

h2<br />

h3<br />

h3<br />

177


178<br />

∂Ez<br />

∂y<br />

∂Ex<br />

∂z<br />

∂Ey<br />

∂x<br />

− ∂Ey<br />

∂z<br />

− ∂Ez<br />

∂x<br />

− ∂Ex<br />

∂y<br />

= − ∂Bx<br />

∂t<br />

= − ∂By<br />

∂t<br />

= − ∂Bz<br />

∂t<br />

∂Hz<br />

∂y<br />

∂Hx<br />

∂z<br />

∂Hy<br />

∂x<br />

Here we have introduced the notations:<br />

Dx = D(1)<br />

Dy = D(2)<br />

Dz = D(3)<br />

Bx = B(1)<br />

By = B(2)<br />

Bz = B(3)<br />

∂Hy<br />

−<br />

∂z = Jx + ∂Dx<br />

∂t<br />

∂Hz<br />

−<br />

∂x = Jy + ∂Dy<br />

∂t<br />

∂Hx<br />

−<br />

∂y = Jz + ∂Dz<br />

∂t<br />

Hx = H(1)<br />

Hy = H(2)<br />

Hz = H(3)<br />

with x 1 = x, x 2 = y, x 3 = z, h1 = h2 = h3 =1<br />

Jx = J(1)<br />

Jy = J(2)<br />

Jz = J(3)<br />

∂Dx<br />

∂x<br />

∂Bx<br />

∂x<br />

Table 3 Maxwell’s equations Cartesian coordinates<br />

1 ∂Ez<br />

r ∂θ<br />

− ∂Eθ<br />

∂z<br />

= − ∂Br<br />

∂t<br />

∂Er ∂Ez ∂Bθ<br />

− = −<br />

∂z ∂r ∂t<br />

1 ∂<br />

r ∂r (rEθ) − 1 ∂Er ∂Bz<br />

= −<br />

r ∂θ ∂t<br />

1 ∂ 1 ∂Dθ ∂Dz<br />

(rDr)+ + = ϱ<br />

r ∂r r ∂θ ∂z<br />

Here we have introduced the notations:<br />

Dr = D(1)<br />

Dθ = D(2)<br />

Dz = D(3)<br />

Br = B(1)<br />

Bθ = B(2)<br />

Bz = B(3)<br />

1<br />

r<br />

Hr = H(1)<br />

Hθ = H(2)<br />

Hz = H(3)<br />

1 ∂Hz<br />

r ∂θ<br />

∂Hr<br />

∂z<br />

∂<br />

∂r (rHθ) − 1<br />

with x 1 = r, x 2 = θ, x 3 = z, h1 =1, h2 = r, h3 =1.<br />

+ ∂Dy<br />

∂y<br />

+ ∂By<br />

∂y<br />

Ex = E(1)<br />

Ey = E(2)<br />

Ez = E(3)<br />

+ ∂Dz<br />

∂z<br />

∂Hθ<br />

−<br />

∂z = Jr + ∂Dr<br />

∂t<br />

∂Hz<br />

−<br />

∂r = Jθ + ∂Dθ<br />

∂t<br />

∂Hr<br />

r ∂θ = Jz + ∂Dz<br />

∂t<br />

1 ∂ 1 ∂Bθ ∂Bz<br />

(rBr)+ +<br />

r ∂r r ∂θ ∂z =0<br />

Jr = J(1)<br />

Jθ = J(2)<br />

Jz = J(3)<br />

Er = E(1)<br />

Eθ = E(2)<br />

Ez = E(3)<br />

Table 4 Maxwell’s equations in cylindrical coordinates.<br />

= ϱ<br />

+ ∂Bz<br />

∂z =0


1 ∂<br />

ρ sin θ ∂θ (sin θEφ) − ∂Eθ<br />

<br />

= −<br />

∂φ<br />

∂Bρ<br />

∂t<br />

1 ∂Eρ 1 ∂<br />

−<br />

ρ sin θ ∂φ ρ ∂ρ (ρEφ) =− ∂Bθ<br />

∂t<br />

1 ∂<br />

ρ ∂ρ (ρEθ) − 1 ∂Eρ ∂Bφ<br />

= −<br />

ρ ∂θ ∂t<br />

<br />

1 ∂<br />

ρ sin θ ∂θ (sin θHφ) − ∂Hθ<br />

<br />

= Jρ +<br />

∂φ<br />

∂Dρ<br />

∂t<br />

1 ∂Hρ 1 ∂<br />

−<br />

ρ sin θ ∂φ ρ ∂ρ (ρHφ) =Jθ + ∂Dθ<br />

∂t<br />

1 ∂<br />

ρ ∂ρ (ρHθ) − 1 ∂Hρ<br />

ρ ∂θ = Jφ + ∂Dφ<br />

∂t<br />

1<br />

ρ2 ∂<br />

∂ρ (ρ2Dρ)+ 1 ∂<br />

1 ∂Dφ<br />

(sin θDθ)+<br />

ρ sin θ ∂θ ρ sin θ ∂φ =ϱ<br />

1<br />

ρ2 ∂<br />

∂ρ (ρ2Bρ)+ 1 ∂<br />

1 ∂Bφ<br />

(sin θBθ)+<br />

ρ sin θ ∂θ ρ sin θ ∂φ =0<br />

Here we have introduced the notations:<br />

Dρ = D(1)<br />

Dθ = D(2)<br />

Dφ = D(3)<br />

Bρ = B(1)<br />

Bθ = B(2)<br />

Bφ = B(3)<br />

Hρ = H(1)<br />

Hθ = H(2)<br />

Hφ = H(3)<br />

Jρ = J(1)<br />

Jθ = J(2)<br />

Jφ = J(3)<br />

with x 1 = ρ, x 2 = θ, x 3 = φ, h1 =1, h2 = ρ, h3 = ρ sin θ<br />

Table 5 Maxwell’s equations spherical coordinates.<br />

Eigenvalues and Eigenvectors of Symmetric Tensors<br />

Consider the equation<br />

Eρ = E(1)<br />

Eθ = E(2)<br />

Eφ = E(3)<br />

TijAj = λAi, i,j =1, 2, 3, (2.1.19)<br />

where Tij = Tji is symmetric, Ai are the components of a vector and λ is a scalar. Any nonzero solution<br />

Ai of equation (2.1.19) is called an eigenvector of the tensor Tij and the associated scalar λ is called an<br />

eigenvalue. When expanded these equations have the form<br />

(T11 − λ)A1 + T<strong>12</strong>A2 + T13A3 =0<br />

T21A1 +(T22 − λ)A2 + T23A3 =0<br />

T31A1 + T32A2 +(T33 − λ)A3 =0.<br />

The condition for equation (2.1.19) to have a nonzero solution Ai is that the characteristic equation<br />

should be zero. This equation is found from the determinant equation<br />

<br />

<br />

<br />

T11 − λ T<strong>12</strong> T13 <br />

<br />

f(λ) = <br />

T21 T22 − λ T23 <br />

=0, (2.1.20)<br />

T31 T32 T33 − λ <br />

179


180<br />

which when expanded is a cubic equation of the form<br />

where I1,I2 and I3 are invariants defined by the relations<br />

f(λ) =−λ 3 + I1λ 2 − I2λ + I3 =0, (2.1.21)<br />

I1 = Tii<br />

I2 = 1<br />

2 TiiTjj − 1<br />

2 TijTij<br />

I3 = eijkTi1Tj2Tk3.<br />

When Tij is subjected to an orthogonal transformation, where ¯ Tmn = Tijℓimℓjn, then<br />

ℓimℓjn (Tmn − λδmn) = ¯ Tij − λδij and det (Tmn − λδmn) =det <br />

Tij ¯ − λδij .<br />

Hence, the eigenvalues of a second order tensor remain invariant under an orthogonal transformation.<br />

If Tij is real and symmetric then<br />

• the eigenvalues of Tij will be real, and<br />

• the eigenvectors corresponding to distinct eigenvalues will be orthogonal.<br />

(2.1.22)<br />

Proof: To show a quantity is real we show that the conjugate of the quantity equals the given quantity. If<br />

(2.1.19) is satisfied, we multiply by the conjugate Ai and obtain<br />

AiTijAj = λAiAi. (2.1.25)<br />

The right hand side of this equation has the inner product AiAi which is real. It remains to show the left<br />

hand side of equation (2.1.25) is also real. Consider the conjugate of this left hand side and write<br />

AiTijAj = AiT ijAj = AiTjiAj = AiTijAj.<br />

Consequently, the left hand side of equation (2.1.25) is real and the eigenvalue λ can be represented as the<br />

ratio of two real quantities.<br />

Assume that λ (1) and λ (2) are two distinct eigenvalues which produce the unit eigenvectors ˆ L1 and ˆ L2<br />

with components ℓi1 and ℓi2,i=1, 2, 3 respectively. We then have<br />

Consider the products<br />

Tijℓj1 = λ (1)ℓi1 and Tijℓj2 = λ (2)ℓi2. (2.1.26)<br />

λ (1)ℓi1ℓi2 = Tijℓj1ℓi2,<br />

λ (2)ℓi1ℓi2 = ℓi1Tijℓj2 = ℓj1Tjiℓi2.<br />

and subtract these equations. We find that<br />

(2.1.27)<br />

[λ (1) − λ (2)]ℓi1ℓi2 =0. (2.1.28)<br />

By hypothesis, λ (1) is different from λ (2) and consequently the inner product ℓi1ℓi2 must be zero. Therefore,<br />

the eigenvectors corresponding to distinct eigenvalues are orthogonal.


Therefore, associated with distinct eigenvalues λ (i),i=1, 2, 3 there are unit eigenvectors<br />

ˆL (i) = ℓi1 ê1 + ℓi2 ê2 + ℓi3 ê3<br />

with components ℓim,m=1, 2, 3 which are direction cosines and satisfy<br />

The unit eigenvectors satisfy the relations<br />

and can be written as the single equation<br />

Consider the transformation<br />

ℓinℓim = δmn and ℓijℓjm = δim. (2.1.23)<br />

Tijℓj1 = λ (1)ℓi1 Tijℓj2 = λ (2)ℓi2 Tijℓj3 = λ (3)ℓi3<br />

Tijℓjm = λ (m)ℓim, m =1, 2, or3 m not summed.<br />

xi = ℓijxj or xm = ℓmjxj<br />

which represents a rotation of axes, where ℓij are the direction cosines from the eigenvectors of Tij. This is a<br />

linear transformation where the ℓij satisfy equation (2.1.23). Such a transformation is called an orthogonal<br />

transformation. In the new x coordinate system, called principal axes, we have<br />

∂x<br />

T mn = Tij<br />

i<br />

∂xm ∂x j<br />

∂x n = Tijℓimℓjn = λ (n)ℓinℓim = λ (n)δmn (no sum on n). (2.1.24)<br />

This equation shows that in the barred coordinate system there are the components<br />

<br />

T mn =<br />

⎡<br />

⎣ λ (1) 0 0<br />

0 λ (2) 0<br />

0 0 λ (3)<br />

That is, along the principal axes the tensor components Tij are transformed to the components T ij where<br />

T ij =0fori = j. The elements T (i)(i) , i not summed, represent the eigenvalues of the transformation<br />

(2.1.19).<br />

⎤<br />

⎦ .<br />

181


182<br />

EXERCISE 2.1<br />

◮ 1. In cylindrical coordinates (r, θ, z) withf = f(r, θ, z) find the gradient of f.<br />

◮ 2. In cylindrical coordinates (r, θ, z) with A = A(r, θ, z) find div A.<br />

◮ 3. In cylindrical coordinates (r, θ, z) for A = A(r, θ, z) find curl A.<br />

◮ 4. In cylindrical coordinates (r, θ, z) forf = f(r, θ, z) find ∇ 2 f.<br />

◮ 5. In spherical coordinates (ρ, θ, φ) withf = f(ρ, θ, φ) find the gradient of f.<br />

◮ 6. In spherical coordinates (ρ, θ, φ) with A = A(ρ, θ, φ) find div A.<br />

◮ 7. In spherical coordinates (ρ, θ, φ) for A = A(ρ, θ, φ) find curl A.<br />

◮ 8. In spherical coordinates (ρ, θ, φ) forf = f(ρ, θ, φ) find ∇ 2 f.<br />

◮ 9. Let r = x ê1 +y ê2 +z ê3 denote the position vector of a variable point (x, y, z) in Cartesian coordinates.<br />

Let r = |r| denote the distance of this point from the origin. Find in terms of r and r:<br />

(a) grad(r) (b) grad(r m ) (c) grad( 1<br />

) (d) grad(lnr) (e) grad(φ)<br />

r<br />

where φ = φ(r) is an arbitrary function of r.<br />

◮ 10. Let r = x ê1+y ê2+z ê3 denote the position vector of a variable point (x, y, z) in Cartesian coordinates.<br />

Let r = |r| denote the distance of this point from the origin. Find:<br />

where φ = φ(r) is an arbitrary function or r.<br />

(a) div (r) (b) div (r m r) (c) div (r −3 r) (d) div (φr)<br />

◮ 11. Let r = x ê1 + y ê2 + z ê3 denote the position vector of a variable point (x, y, z) in Cartesian<br />

coordinates. Let r = |r| denote the distance of this point from the origin. Find: (a)<br />

where φ = φ(r) is an arbitrary function of r.<br />

curl r (b) curl (φr)<br />

◮ <strong>12</strong>. Expand and simplify the representation for curl (curl A).<br />

◮ 13. Show that the curl of the gradient is zero in generalized coordinates.<br />

◮ 14. Write out the physical components associated with the gradient of φ = φ(x 1 ,x 2 ,x 3 ).<br />

◮ 15. Show that<br />

g im Ai,m = 1<br />

√ g<br />

∂<br />

∂xi √ im i<br />

gg Am = A ,i = 1<br />

√<br />

g<br />

∂<br />

∂xi √ i<br />

gA .


◮ 16. Let r =(r · r) 1/2 = x 2 + y 2 + z 2 ) and calculate (a) ∇ 2 (r) (b) ∇ 2 (1/r) (c) ∇ 2 (r 2 ) (d) ∇ 2 (1/r 2 )<br />

◮ 17. Given the tensor equations Dij = 1<br />

2 (vi,j + vj,i), i,j =1, 2, 3. Let v(1),v(2),v(3) denote the<br />

physical components of v1,v2,v3 and let D(ij) denote the physical components associated with Dij. Assume<br />

the coordinate system (x1 ,x2 ,x3 ) is orthogonal with metric coefficients g (i)(i) = h2 i ,i=1, 2, 3andgij =0<br />

for i = j.<br />

(a) Find expressions for the physical components D(11),D(22) and D(33) in terms of the physical components<br />

v(i),i=1, 2, 3. Answer: D(ii) = 1 ∂V (i) V (j) ∂hi<br />

+ no sum on i.<br />

hi ∂xi hihj ∂xj j=i<br />

(b) Find expressions for the physical components D(<strong>12</strong>),D(13) and D(23) in terms of the physical components<br />

v(i),i=1, 2, 3. Answer: D(ij) = 1<br />

<br />

hi ∂<br />

2 hj ∂xj <br />

V (i)<br />

+<br />

hi<br />

hj ∂<br />

hi ∂xi <br />

V (j)<br />

hj<br />

◮ 18. Write out the tensor equations in problem 17 in Cartesian coordinates.<br />

◮ 19. Write out the tensor equations in problem 17 in cylindrical coordinates.<br />

◮ 20. Write out the tensor equations in problem 17 in spherical coordinates.<br />

◮ 21. Express the vector equation (λ +2µ)∇Φ − 2µ∇×ω + F = 0 intensorform.<br />

◮ 22. Write out the equations in problem 21 for a generalized orthogonal coordinate system in terms of<br />

physical components.<br />

◮ 23. Write out the equations in problem 22 for cylindrical coordinates.<br />

◮ 24. Write out the equations in problem 22 for spherical coordinates.<br />

◮ 25. Use equation (2.1.4) to represent the divergence in parabolic cylindrical coordinates (ξ,η,z).<br />

◮ 26. Use equation (2.1.4) to represent the divergence in parabolic coordinates (ξ,η,φ).<br />

◮ 27. Use equation (2.1.4) to represent the divergence in elliptic cylindrical coordinates (ξ,η,z).<br />

Change the given equations from a vector notation to a tensor notation.<br />

◮ 28. B = v ∇· A +(∇·v) A<br />

◮ 29.<br />

◮ 30.<br />

◮ 31.<br />

◮ 32.<br />

d<br />

dt [ A · ( B × C)] = d A<br />

dt · ( B × C)+ A · ( d B<br />

dt × C)+ A · ( B × d C<br />

dt )<br />

dv ∂v<br />

= +(v ·∇)v<br />

dt ∂t<br />

1 ∂<br />

c<br />

H<br />

∂t = −curl E<br />

d B<br />

dt − ( B ·∇)v + B(∇·v) =0<br />

183


184<br />

Change the given equations from a tensor notation to a vector notation.<br />

◮ 33. ɛ ijk Bk,j + F i =0<br />

◮ 34. gijɛ jkl Bl,k + Fi =0<br />

◮ 35.<br />

∂ϱ<br />

∂t +(ϱvi),i=0<br />

◮ 36. ϱ( ∂vi ∂vi ∂P<br />

+ vm )=−<br />

∂t ∂xm ∂xi + µ ∂2vi ∂xm∂x <br />

1<br />

<strong>12</strong>bh3 − 1<br />

24b2h 2<br />

− 1<br />

24b2 2 1 h <strong>12</strong>b3h m + Fi<br />

<br />

◮ 37. The moment of inertia of an area or second moment of area is defined by Iij = (ymymδij −yiyj) dA<br />

A<br />

where dA is an element of area. Calculate the moment of inertia<br />

<br />

Iij, i,j=1, 2 for the triangle illustrated in<br />

the figure 2.1-1 and show that Iij =<br />

.<br />

Figure 2.1-1 Moments of inertia for a triangle<br />

◮ 38. Use the results from problem 37 and rotate the axes in figure 2.1-1 through an angle θ to a barred<br />

system of coordinates.<br />

(a) Show that in the barred system of coordinates<br />

<br />

I11 + I22 I11 − I22<br />

I11 =<br />

+<br />

cos 2θ + I<strong>12</strong> sin 2θ<br />

2<br />

2<br />

<br />

I11 − I22<br />

I<strong>12</strong> = I21 = −<br />

sin 2θ + I<strong>12</strong> cos 2θ<br />

2<br />

<br />

I11 + I22 I11 − I22<br />

I22 =<br />

−<br />

cos 2θ − I<strong>12</strong> sin 2θ<br />

2<br />

2<br />

(b) For what value of θ will I11 have a maximum value?<br />

(c) Show that when I11 is a maximum, we will have I22 a minimum and I<strong>12</strong> = I21 =0.


Figure 2.1-2 Mohr’s circle<br />

◮ 39. Otto Mohr1 gave the following physical interpretation to the results obtained in problem 38:<br />

• Plot the points A(I11,I<strong>12</strong>) andB(I22, −I<strong>12</strong>) as illustrated in the figure 2.1-2<br />

• Draw the line AB and calculate the point C where this line intersects the I axes. Show the point C<br />

has the coordinates<br />

( I11 + I22<br />

, 0)<br />

2<br />

• Calculate the radius of the circle with center at the point C and with diagonal AB and show this<br />

radius is<br />

<br />

I11<br />

2 − I22<br />

r =<br />

+ I<br />

2<br />

2 <strong>12</strong><br />

• Show the maximum and minimum values of I occur where the constructed circle intersects the I axes.<br />

Show that Imax = I11 = I11 + I22<br />

2<br />

+ r Imin = I22 = I11 + I22<br />

2<br />

◮ 40. Show directly that the eigenvalues of the symmetric matrix Iij =<br />

λ2 = Imin where Imax and Imin are given in problem 39.<br />

− r.<br />

I11 I<strong>12</strong><br />

I21 I22<br />

<br />

are λ1 = Imax and<br />

◮ 41. Find the principal axes and moments of inertia for the triangle given in problem 37 and summarize<br />

your results from problems 37,38,39, and 40.<br />

◮ 42. Verify for orthogonal coordinates the relations<br />

or<br />

<br />

∇× <br />

A · ê (i) =<br />

∇× A =<br />

1<br />

h1h2h3<br />

◮ 43. Verify for orthogonal coordinates the relation<br />

<br />

∇×(∇× <br />

A) · ê (i) =<br />

3<br />

e (i)jk<br />

h1h2h3<br />

k=1<br />

∂(h (k)A(k))<br />

h (i)<br />

∂xj<br />

<br />

<br />

<br />

h1 ê1 h2 ê2 h3 ê3 <br />

<br />

∂ ∂<br />

∂ <br />

∂x1 ∂x2 ∂x3 <br />

h1A(1) h2A(2) h3A(3) .<br />

3<br />

e (i)jrersm<br />

m=1<br />

h (i)<br />

h1h2h3<br />

1 Christian Otto Mohr (1835-1918) German civil engineer.<br />

∂<br />

∂xj<br />

h 2 (r)<br />

h1h2h3<br />

<br />

∂(h (m)A(m))<br />

∂xs<br />

185


186<br />

◮ 44. Verify for orthogonal coordinates the relation<br />

<br />

∇ ∇· <br />

A · ê (i) = 1<br />

<br />

∂ 1 ∂(h2h3A(1))<br />

h (i) ∂x (i) h1h2h3 ∂x1<br />

◮ 45. Verify the relation<br />

<br />

( A ·∇) <br />

B · ê (i) =<br />

3<br />

k=1<br />

A(k) ∂B(i)<br />

+<br />

h (k) ∂xk<br />

<br />

k=i<br />

◮ 46. The Gauss divergence theorem is written<br />

<br />

1<br />

<br />

∂F ∂F2 ∂F3<br />

+ + dτ =<br />

∂x ∂y ∂z<br />

B(k)<br />

hkh (i)<br />

+ ∂(h1h3A(2))<br />

∂x2<br />

+ ∂(h1h2A(3))<br />

∂x3<br />

<br />

A(i) ∂h (i)<br />

− A(k)<br />

∂xk<br />

∂hk<br />

<br />

∂x (i)<br />

n1F 1 + n2F 2 + n3F 3 dσ<br />

V<br />

S<br />

where V is the volume within a simple closed surface S. Here it is assumed that F i = F i (x, y, z) are<br />

continuous functions with continuous first order derivatives throughout V and ni are the direction cosines<br />

of the outward normal to S, dτ is an element of volume and dσ is an element of surface area.<br />

(a) Show that in a Cartesian coordinate system<br />

F i ∂F1<br />

,i =<br />

∂x<br />

<br />

and that the tensor form of this theorem is<br />

(b) Write the vector form of this theorem.<br />

(c) Show that if we define<br />

V<br />

+ ∂F2<br />

∂y<br />

F i<br />

,i dτ =<br />

∂F3<br />

+<br />

∂z<br />

<br />

F<br />

S<br />

i ni dσ.<br />

ur = ∂u<br />

∂xr , vr = ∂v<br />

∂xr and Fr = grmF m = uvr<br />

then F i<br />

,i = gimFi,m = g im (uvi,m + umvi)<br />

(d) Show that another form of the Gauss divergence theorem is<br />

<br />

g im <br />

umvi dτ = uvmn m <br />

dσ −<br />

V<br />

S<br />

ug<br />

V<br />

im vi,m dτ<br />

Write out the above equation in Cartesian coordinates.<br />

⎛<br />

1<br />

◮ 47. Find the eigenvalues and eigenvectors associated with the matrix A = ⎝ 1<br />

1<br />

2<br />

⎞<br />

2<br />

1⎠<br />

.<br />

2 1 1<br />

Show that the eigenvectors are orthogonal.<br />

⎛<br />

1<br />

◮ 48. Find the eigenvalues and eigenvectors associated with the matrix A = ⎝ 2<br />

2<br />

1<br />

⎞<br />

1<br />

0⎠<br />

.<br />

1 0 1<br />

Show that the eigenvectors are orthogonal.<br />

⎛<br />

1<br />

◮ 49. Find the eigenvalues and eigenvectors associated with the matrix A = ⎝ 1<br />

1<br />

1<br />

⎞<br />

0<br />

1⎠<br />

.<br />

0<br />

Show that the eigenvectors are orthogonal.<br />

1 1<br />

◮ 50. The harmonic and biharmonic functions or potential functions occur in the mathematical modeling<br />

of many physical problems. Any solution of Laplace’s equation ∇2Φ = 0 is called a harmonic function and<br />

any solution of the biharmonic equation ∇4Φ = 0 is called a biharmonic function.<br />

(a) Expand the Laplace equation in Cartesian, cylindrical and spherical coordinates.<br />

(b) Expand the biharmonic equation in two dimensional Cartesian and polar coordinates.<br />

Hint: Consider ∇ 4 Φ=∇ 2 (∇ 2 Φ). In Cartesian coordinates ∇ 2 Φ=Φ,ii and ∇ 4 Φ=Φ,iijj.


§2.2 DYNAMICS<br />

Dynamics is concerned with studying the motion of particles and rigid bodies. By studying the motion<br />

of a single hypothetical particle, one can discern the motion of a system of particles. This in turn leads to<br />

the study of the motion of individual points in a continuous deformable medium.<br />

Particle Movement<br />

The trajectory of a particle in a generalized coordinate system is described by the parametric equations<br />

x i = x i (t), i =1,...,N (2.2.1)<br />

where t is a time parameter. If the coordinates are changed to a barred system by introducing a coordinate<br />

transformation<br />

x i = x i (x 1 ,x 2 ,...,x N ), i =1,...,N<br />

then the trajectory of the particle in the barred system of coordinates is<br />

x i = x i (x 1 (t),x 2 (t),...,x N (t)), i =1,...,N. (2.2.2)<br />

The generalized velocity of the particle in the unbarred system is defined by<br />

v i = dxi<br />

, i =1,...,N. (2.2.3)<br />

dt<br />

By the chain rule differentiation of the transformation equations (2.2.2) one can verify that the velocity in<br />

the barred system is<br />

v r = dxr ∂xr<br />

=<br />

dt ∂xj dxj ∂xr<br />

=<br />

dt ∂xj vj , r =1,...,N. (2.2.4)<br />

Consequently, the generalized velocity vi is a first order contravariant tensor. The speed of the particle is<br />

obtained from the magnitude of the velocity and is<br />

v 2 = gijv i v j .<br />

The generalized acceleration f i of the particle is defined as the intrinsic derivative of the generalized velocity.<br />

The generalized acceleration has the form<br />

f i = δvi<br />

δt = vi dx<br />

,n<br />

n<br />

dt<br />

and the magnitude of the acceleration is<br />

dvi<br />

=<br />

dt +<br />

<br />

i<br />

mn<br />

f 2 = gijf i f j .<br />

v m v n = d2xi +<br />

dt2 <br />

i<br />

m dx dx<br />

mn dt<br />

n<br />

dt<br />

(2.2.5)<br />

187


188<br />

Frenet-Serret Formulas<br />

Figure 2.2-1 Tangent, normal and binormal to point P on curve.<br />

The parametric equations (2.2.1) describe a curve in our generalized space. With reference to the figure<br />

2.2-1 we wish to define at each point P of the curve the following orthogonal unit vectors:<br />

T i = unit tangent vector at each point P.<br />

N i = unit normal vector at each point P.<br />

B i = unit binormal vector at each point P.<br />

These vectors define the osculating, normal and rectifying planes illustrated in the figure 2.2-1.<br />

In the generalized coordinates the arc length squared is<br />

Define T i = dxi<br />

ds<br />

ds 2 = gijdx i dx j .<br />

as the tangent vector to the parametric curve defined by equation (2.2.1). This vector is a<br />

unit tangent vector because if we write the element of arc length squared in the form<br />

dx<br />

1=gij<br />

i dx<br />

ds<br />

j<br />

ds = gijT i T j , (2.2.6)<br />

we obtain the generalized dot product for T i . This generalized dot product implies that the tangent vector<br />

is a unit vector. Differentiating the equation (2.2.6) intrinsically with respect to arc length s along the curve<br />

produces<br />

which simplifies to<br />

δT<br />

gmn<br />

m<br />

δs T n n<br />

m δT<br />

+ gmnT<br />

δs =0,<br />

m<br />

n δT<br />

gmnT =0. (2.2.7)<br />

δs


The equation (2.2.7) is a statement that the vector<br />

vector is defined as<br />

N i = 1 δT<br />

κ<br />

i<br />

δs<br />

δT m<br />

δs is orthogonal to the vector T m . The unit normal<br />

or Ni = 1<br />

κ<br />

δTi<br />

, (2.2.8)<br />

δs<br />

where κ is a scalar called the curvature and is chosen such that the magnitude of N i is unity. The reciprocal<br />

of the curvature is R = 1<br />

κ , which is called the radius of curvature. The curvature of a straight line is zero<br />

while the curvature of a circle is a constant. The curvature measures the rate of change of the tangent vector<br />

as the arc length varies.<br />

The equation (2.2.7) can be expressed in the form<br />

gijT i N j =0. (2.2.9)<br />

Taking the intrinsic derivative of equation (2.2.9) with respect to the arc length s produces<br />

or<br />

i δNj<br />

gijT<br />

δs<br />

The generalized dot product can be written<br />

i δNj δT<br />

gijT + gij<br />

δs i<br />

δs N j =0<br />

δT<br />

= −gij<br />

i<br />

δs N j = −κgijN i N j = −κ. (2.2.10)<br />

gijT i T j =1,<br />

and consequently we can express equation (2.2.10) in the form<br />

Consequently, the vector<br />

i δNj<br />

gijT<br />

δs = −κgijT i T j<br />

δN j<br />

δs<br />

or gijT i<br />

<br />

j δN j<br />

+ κT =0. (2.2.11)<br />

δs<br />

+ κT j<br />

(2.2.<strong>12</strong>)<br />

is orthogonal to T i . In a similar manner, we can use the relation gijN i N j = 1 and differentiate intrinsically<br />

with respect to the arc length s to show that<br />

i δNj<br />

gijN<br />

δs =0.<br />

This in turn can be expressed in the form<br />

gijN i<br />

<br />

j δN j<br />

+ κT =0.<br />

δs<br />

This form of the equation implies that the vector represented in equation (2.2.<strong>12</strong>) is also orthogonal to the<br />

unit normal N i . We define the unit binormal vector as<br />

B i = 1<br />

<br />

i δN i<br />

+ κT or Bi =<br />

τ δs 1<br />

<br />

δNi<br />

+ κTi<br />

(2.2.13)<br />

τ δs<br />

where τ is a scalar called the torsion. The torsion is chosen such that the binormal vector is a unit vector.<br />

The torsion measures the rate of change of the osculating plane and consequently, the torsion τ is a measure<br />

189


190<br />

of the twisting of the curve out of a plane. The value τ = 0 corresponds to a plane curve. The vectors<br />

T i ,N i ,B i , i =1, 2, 3 satisfy the cross product relation<br />

B i = ɛ ijk TjNk.<br />

If we differentiate this relation intrinsically with respect to arc length s we find<br />

δB i<br />

δs<br />

= ɛijk<br />

<br />

Tj<br />

δNk<br />

δs<br />

δTj<br />

+<br />

δs Nk<br />

<br />

= ɛ ijk [Tj(τBk − κTk)+κNjNk]<br />

= τɛ ijk TjBk = −τɛ ikj BkTj = −τN i .<br />

The relations (2.2.8),(2.2.13) and (2.2.14) are now summarized and written<br />

δT i<br />

i<br />

= κN<br />

δs<br />

δN i<br />

δs = τBi − κT i<br />

δB i<br />

δs = −τNi .<br />

These equations are known as the Frenet-Serret formulas of differential geometry.<br />

Velocity and Acceleration<br />

Chain rule differentiation of the generalized velocity is expressible in the form<br />

v i = dxi<br />

dt<br />

= dxi<br />

ds<br />

(2.2.14)<br />

(2.2.15)<br />

ds<br />

dt = T i v, (2.2.16)<br />

where v = ds<br />

dt is the speed of the particle and is the magnitude of vi . The vector T i is the unit tangent vector<br />

to the trajectory curve at the time t. The equation (2.2.16) is a statement of the fact that the velocity of a<br />

particle is always in the direction of the tangent vector to the curve and has the speed v.<br />

By chain rule differentiation, the generalized acceleration is expressible in the form<br />

f r = δvr<br />

δt<br />

dv<br />

=<br />

dt T r r δT<br />

+ v<br />

δt<br />

= dv<br />

dt T r + v<br />

δT r<br />

δs<br />

ds<br />

dt<br />

= dv<br />

dt T r + κv 2 N r .<br />

(2.2.17)<br />

The equation (2.2.17) states that the acceleration lies in the osculating plane. Further, the equation (2.2.17)<br />

indicates that the tangential component of the acceleration is dv<br />

dt , while the normal component of the acceleration<br />

is κv 2 .


Work and Potential Energy<br />

Define M as the constant mass of the particle as it moves along the curve defined by equation (2.2.1).<br />

Also let Qr denote the components of a force vector (in appropriate units of measurements) which acts upon<br />

the particle. Newton’s second law of motion can then be expressed in the form<br />

Q r = Mf r<br />

or Qr = Mfr. (2.2.18)<br />

The work done W in moving a particle from a point P0 to a point P1 along a curve x r = x r (t),r =1, 2, 3,<br />

with parameter t, is represented by a summation of the tangential components of the forces acting along the<br />

path and is defined as the line integral<br />

P1<br />

dx<br />

W = Qr<br />

P0<br />

r P1<br />

ds = Qr dx<br />

ds P0<br />

r t1<br />

dx<br />

= Qr<br />

t0<br />

r t1<br />

dt = Qrv<br />

dt t0<br />

r dt (2.2.19)<br />

where Qr = grsQs is the covariant form of the force vector, t is the time parameter and s is arc length along<br />

the curve.<br />

Conservative Systems<br />

If the force vector is conservative it means that the force is derivable from a scalar potential function<br />

V = V (x 1 ,x 2 ,...,x N ) such that Qr = −V ,r = − ∂V<br />

, r =1,...,N. (2.2.20)<br />

∂xr In this case the equation (2.2.19) can be integrated and we find that to within an additive constant we will<br />

have V = −W. The potential function V is called the potential energy of the particle and the work done<br />

becomes the change in potential energy between the starting and end points and is independent of the path<br />

connecting the points.<br />

Lagrange’s Equations of Motion<br />

The kinetic energy T of the particle is defined as one half the mass times the velocity squared and can<br />

be expressed in any of the forms<br />

T = 1<br />

2 M<br />

2 ds<br />

=<br />

dt<br />

1<br />

2 Mv2 = 1<br />

2 Mgmnv m v n = 1<br />

2 Mgmn ˙x m ˙x n , (2.2.21)<br />

where the dot notation denotes differentiation with respect to time. It is an easy exercise to calculate the<br />

derivatives<br />

∂T<br />

and thereby verify the relation<br />

d<br />

dt<br />

∂ ˙x r = Mgrm ˙x m<br />

∂T<br />

∂ ˙x r<br />

∂T<br />

∂x<br />

<br />

= M<br />

1<br />

= r 2<br />

<br />

grm¨x m + ∂grm<br />

∂xn ˙xn ˙x m<br />

<br />

M ∂gmn<br />

∂x r ˙xm ˙x n ,<br />

(2.2.22)<br />

<br />

d ∂T<br />

dt ∂ ˙x r<br />

<br />

− ∂T<br />

∂xr = Mfr = Qr, r =1,...,N. (2.2.23)<br />

191


192<br />

This equation is called the Lagrange’s form of the equations of motion.<br />

EXAMPLE 2.2-1. (Equations of motion in spherical coordinates) Find the Lagrange’s form of<br />

the equations of motion in spherical coordinates.<br />

Solution: Let x 1 = ρ, x 2 = θ, x 3 = φ then the element of arc length squared in spherical coordinates has<br />

the form<br />

ds 2 =(dρ) 2 + ρ 2 (dθ) 2 + ρ 2 sin 2 θ(dφ) 2 .<br />

The element of arc length squared can be used to construct the kinetic energy. For example,<br />

T = 1<br />

2 M<br />

2 ds<br />

=<br />

dt<br />

1<br />

2 M<br />

<br />

(˙ρ) 2 + ρ 2 ( ˙ θ) 2 + ρ 2 sin 2 θ( ˙ φ) 2<br />

.<br />

The Lagrange form of the equations of motion of a particle are found from the relations (2.2.23) and are<br />

calculated to be:<br />

Mf1 = Q1 = d<br />

<br />

∂T<br />

−<br />

dt ∂ ˙ρ<br />

∂T<br />

<br />

= M ¨ρ − ρ(<br />

∂ρ ˙ θ) 2 − ρ sin 2 θ( ˙ φ) 2<br />

Mf2 = Q2 = d<br />

<br />

∂T<br />

dt ∂ ˙ <br />

−<br />

θ<br />

∂T<br />

<br />

d<br />

<br />

= M ρ<br />

∂θ dt<br />

2 <br />

θ˙<br />

− ρ 2 sin θ cos θ( ˙ φ) 2<br />

<br />

Mf3 = Q3 = d<br />

<br />

∂T<br />

dt ∂ ˙ <br />

−<br />

φ<br />

∂T<br />

<br />

d<br />

<br />

= M ρ<br />

∂φ dt<br />

2 sin 2 θ ˙ <br />

φ<br />

<br />

.<br />

In terms of physical components we have<br />

<br />

Qρ = M ¨ρ − ρ( ˙ θ) 2 − ρ sin 2 θ( ˙ φ) 2<br />

Qθ = M<br />

<br />

d<br />

<br />

ρ<br />

ρ dt<br />

2 <br />

θ˙<br />

− ρ 2 sin θ cos θ( ˙ φ) 2<br />

<br />

Qφ = M<br />

<br />

d<br />

<br />

ρ<br />

ρ sin θ dt<br />

2 sin 2 θ ˙ <br />

φ<br />

<br />

.<br />

Euler-Lagrange Equations of Motion<br />

Starting with the Lagrange’s form of the equations of motion from equation (2.2.23), we assume that<br />

the external force Qr is derivable from a potential function V as specified by the equation (2.2.20). That is,<br />

we assume the system is conservative and express the equations of motion in the form<br />

<br />

d ∂T<br />

dt ∂ ˙x r<br />

<br />

− ∂T ∂V<br />

= −<br />

∂xr ∂xr = Qr, r =1,...,N (2.2.24)<br />

The Lagrangian is defined by the equation<br />

L = T − V = T (x 1 ,...,x N , ˙x 1 ,..., ˙x N ) − V (x 1 ,...,x N )=L(x i , ˙x i ). (2.2.25)<br />

Employing the defining equation (2.2.25), it is readily verified that the equations of motion are expressible<br />

in the form<br />

<br />

d ∂L<br />

dt ∂ ˙x r<br />

<br />

− ∂L<br />

=0, r =1,...,N, (2.2.26)<br />

∂xr which are called the Euler-Lagrange form for the equations of motion.


Figure 2.2-2 Simply pulley system<br />

EXAMPLE 2.2-2. (Simple pulley system)<br />

illustrated in the figure 2.2-2.<br />

Find the equation of motion for the simply pulley system<br />

Solution: The given system has only one degree of freedom, say y1. It is assumed that<br />

The kinetic energy of the system is<br />

y1 + y2 = ℓ =aconstant.<br />

T = 1<br />

2 (m1 + m2)˙y 2 1 .<br />

Let y1 increase by an amount dy1 and show the work done by gravity can be expressed as<br />

dW = m1gdy1 + m2gdy2<br />

dW = m1gdy1 − m2gdy1<br />

dW =(m1 − m2)gdy1 = Q1 dy1.<br />

Here Q1 =(m1− m2)g is the external force acting on the system where g is the acceleration of gravity. The<br />

Lagrange equation of motion is<br />

<br />

d ∂T<br />

−<br />

dt ∂ ˙y1<br />

∂T<br />

∂y1<br />

= Q1<br />

or<br />

(m1 + m2)¨y1 =(m1 − m2)g.<br />

Initial conditions must be applied to y1 and ˙y1 before this equation can be solved.<br />

193


194<br />

EXAMPLE 2.2-3. (Simple pendulum) Find the equation of motion for the pendulum system illustrated<br />

in the figure 2.2-3.<br />

Solution: Choose the angle θ illustrated in the figure 2.2-3 as the generalized coordinate. If the pendulum<br />

is moved from a vertical position through an angle θ, we observe that the mass m moves up a distance<br />

h = ℓ − ℓ cos θ. The work done in moving this mass a vertical distance h is<br />

W = −mgh = −mgℓ(1 − cos θ),<br />

since the force is −mg in this coordinate system. In moving the pendulum through an angle θ, the arc length<br />

s swept out by the mass m is s = ℓθ. This implies that the kinetic energy can be expressed<br />

T = 1<br />

2 m<br />

2 ds<br />

=<br />

dt<br />

1<br />

2 m<br />

<br />

ℓ ˙ 2 θ = 1<br />

2 mℓ2 ( ˙ θ) 2 .<br />

The Lagrangian of the system is<br />

Figure 2.2-3 Simple pendulum system<br />

L = T − V = 1<br />

2 mℓ2 ( ˙ θ) 2 − mgℓ(1 − cos θ)<br />

and from this we find the equation of motion<br />

<br />

d ∂L<br />

dt ∂ ˙ <br />

−<br />

θ<br />

∂L<br />

d<br />

<br />

=0 or mℓ<br />

∂θ dt<br />

2 <br />

θ˙<br />

− mgℓ(− sin θ) =0.<br />

This in turn simplifies to the equation<br />

¨θ + g<br />

sin θ =0.<br />

ℓ<br />

This equation together with a set of initial conditions for θ and ˙ θ represents the nonlinear differential equation<br />

which describes the motion of a pendulum without damping.


EXAMPLE 2.2-4. (Compound pendulum) Find the equations of motion for the compound pendulum<br />

illustrated in the figure 2.2-4.<br />

Solution: Choose for the generalized coordinates the angles x 1 = θ1 and x 2 = θ2 illustrated in the figure<br />

2.2-4. To find the potential function V for this system we consider the work done as the masses m1 and<br />

m2 are moved. Consider independent motions of the angles θ1 and θ2. Imagine the compound pendulum<br />

initially in the vertical position as illustrated in the figure 2.2-4(a). Now let m1 be displaced due to a change<br />

in θ1 and obtain the figure 2.2-4(b). The work done to achieve this position is<br />

W1 = −(m1 + m2)gh1 = −(m1 + m2)gL1(1 − cos θ1).<br />

Starting from the position in figure 2.2-4(b) we now let θ2 undergo a displacement and achieve the configuration<br />

in the figure 2.2-4(c).<br />

Figure 2.2-4 Compound pendulum<br />

The work done due to the displacement θ2 can be represented<br />

W2 = −m2gh2 = −m2gL2(1 − cos θ2).<br />

Since the potential energy V satisfies V = −W to within an additive constant, we can write<br />

V = −W = −W1 − W2 = −(m1 + m2)gL1 cos θ1 − m2gL2 cos θ2 + constant,<br />

where the constant term in the potential energy has been neglected since it does not contribute anything to<br />

the equations of motion. (i.e. the derivative of a constant is zero.)<br />

The kinetic energy term for this system can be represented<br />

T = 1<br />

2 m1<br />

2 ds1<br />

+<br />

dt<br />

1<br />

2 m2<br />

2 ds2<br />

dt<br />

T = 1<br />

2 m1(˙x 2 1 +˙y2 1 )+1<br />

2 m2(˙x 2 2 +˙y2 2 ),<br />

(2.2.27)<br />

195


196<br />

where<br />

(x1,y1) =(L1 sin θ1 , −L1 cos θ1)<br />

(x2,y2) =(L1 sin θ1 + L2 sin θ2, −L1 cos θ1 − L2 cos θ2)<br />

(2.2.28)<br />

are the coordinates of the masses m1 and m2 respectively. Substituting the equations (2.2.28) into equation<br />

(2.2.27) and simplifying produces the kinetic energy expression<br />

T = 1<br />

2 (m1 + m2)L 2 1 ˙ θ 2 1 + m2L1L2 ˙ θ1 ˙ θ2 cos(θ1 − θ2)+ 1<br />

2 m2L 2 2 ˙ θ 2 2. (2.2.29)<br />

Writing the Lagrangian as L = T − V , the equations describing the motion of the compound pendulum<br />

are obtained from the Lagrangian equations<br />

<br />

d ∂L<br />

dt ∂ ˙ <br />

−<br />

θ1<br />

∂L<br />

<br />

d ∂L<br />

=0 and<br />

∂θ1<br />

dt ∂ ˙ <br />

−<br />

θ2<br />

∂L<br />

=0.<br />

∂θ2<br />

Calculating the necessary derivatives, substituting them into the Lagrangian equations of motion and then<br />

simplifying we derive the equations of motion<br />

L1 ¨ θ1 +<br />

m2<br />

m1 + m2<br />

L2 ¨ θ2 cos(θ1 − θ2)+<br />

m2<br />

m1 + m2<br />

L2( ˙ θ2) 2 sin(θ1 − θ2)+g sin θ1 =0<br />

L1 ¨ θ1 cos(θ1 − θ2)+L2 ¨ θ2 − L1( ˙ θ1) 2 sin(θ1 − θ2)+g sin θ2 =0.<br />

These equations are a set of coupled, second order nonlinear ordinary differential equations. These equations<br />

are subject to initial conditions being imposed upon the angular displacements (θ1,θ2) and the angular<br />

velocities ( ˙ θ1, ˙ θ2).<br />

Alternative Derivation of Lagrange’s Equations of Motion<br />

Let c denote a given curve represented in the parametric form<br />

x i = x i (t), i =1,...,N, t0 ≤ t ≤ t1<br />

and let P0,P1 denote two points on this curve corresponding to the parameter values t0 and t1 respectively.<br />

Let c denote another curve which also passes through the two points P0 and P1 as illustrated in the figure<br />

2.2-5.<br />

The curve c is represented in the parametric form<br />

x i = x i (t) =x i (t)+ɛη i (t), i =1,...,N, t0 ≤ t ≤ t1<br />

in terms of a parameter ɛ. In this representation the function η i (t) must satisfy the end conditions<br />

η i (t0) =0 and η i (t1) =0 i =1,...,N<br />

since the curve c is assumed to pass through the end points P0 and P1.<br />

Consider the line integral<br />

I(ɛ) =<br />

t1<br />

L(t, x i + ɛη i , ˙x i + ɛ ˙η i ) dt, (2.2.30)<br />

t0


where<br />

Figure 2.2-5. Motion along curves c and c<br />

L = T − V = L(t, x i , ˙x i )<br />

is the Lagrangian evaluated along the curve c. We ask the question, “What conditions must be satisfied by<br />

the curve c in order that the integral I(ɛ) haveanextremumvaluewhenɛiszero?”If the integral I(ɛ) has<br />

a minimum value when ɛ is zero it follows that its derivative with respect to ɛ will be zero at this value and<br />

we will have<br />

<br />

dI(ɛ) <br />

<br />

dɛ <br />

ɛ=0<br />

=0.<br />

Employing the definition<br />

<br />

dI <br />

<br />

I(ɛ) − I(0)<br />

dɛ = lim<br />

= I<br />

ɛ→0<br />

ɛ=0<br />

ɛ<br />

′ (0) = 0<br />

we expand the Lagrangian in equation (2.2.30) in a series about the point ɛ =0. Substituting the expansion<br />

L(t, x i + ɛη i , ˙x i + ɛ ˙η i )=L(t, x i , ˙x i <br />

∂L<br />

)+ɛ<br />

<br />

+ ɛ 2 [ ]+···<br />

into equation (2.2.30) we calculate the derivative<br />

I ′ I(ɛ) − I(0)<br />

(0) = lim<br />

ɛ→0 ɛ<br />

∂xi ηi + ∂L<br />

˙ηi<br />

∂ ˙x i<br />

t1<br />

∂L<br />

= lim<br />

ɛ→0<br />

t0 ∂xi ηi (t)+ ∂L<br />

∂ ˙x i ˙ηi <br />

(t) dt + ɛ [ ]+···=0,<br />

where we have neglected higher order powers of ɛ since ɛ is approaching zero. Analysis of this equation<br />

informs us that the integral I has a minimum value at ɛ = 0 provided that the integral<br />

t1<br />

∂L<br />

δI =<br />

∂xi ηi (t)+ ∂L<br />

∂ ˙x i ˙ηi <br />

(t) dt =0 (2.2.31)<br />

t0<br />

197


198<br />

is satisfied. Integrating the second term of this integral by parts we find<br />

δI =<br />

t1<br />

t0<br />

∂L<br />

∂xi ηi <br />

∂L<br />

dt +<br />

∂ ˙x i ηi t1 (t)<br />

t0<br />

t1<br />

−<br />

t0<br />

d<br />

dt<br />

<br />

∂L<br />

∂ ˙x i<br />

<br />

η i (t) dt =0. (2.2.32)<br />

The end condition on ηi (t) makes the middle term in equation (2.2.32) vanish and we are left with the<br />

integral<br />

t1<br />

δI = η i <br />

∂L d ∂L<br />

(t) −<br />

∂xi dt ∂ ˙x i<br />

<br />

dt =0, (2.2.33)<br />

t0<br />

which must equal zero for all ηi (t). Since ηi (t) is arbitrary, the only way the integral in equation (2.2.33) can<br />

be zero for all ηi (t) is for the term inside the brackets to vanish. This produces the result that the integral<br />

of the Lagrangian is an extremum when the Euler-Lagrange equations<br />

d<br />

dt<br />

<br />

∂L<br />

∂ ˙x i<br />

<br />

− ∂L<br />

=0, i =1,...,N (2.2.34)<br />

∂xi are satisfied. This is a necessary condition for the integral I(ɛ) to have a minimum value.<br />

In general, any line integral of the form<br />

I =<br />

t1<br />

φ(t, x i , ˙x i ) dt (2.2.35)<br />

t0<br />

has an extremum value if the curve c defined by x i = x i (t), i = 1,...,N satisfies the Euler-Lagrange<br />

equations<br />

<br />

d ∂φ<br />

dt ∂ ˙x i<br />

<br />

− ∂φ<br />

=0, i =1,...,N. (2.2.36)<br />

∂xi The above derivation is a special case of (2.2.36) when φ = L. Note that the equations of motion equations<br />

(2.2.34) are just another form of the equations (2.2.24). Note also that<br />

δT<br />

δt<br />

<br />

δ 1<br />

=<br />

δt 2 mgijv i v j<br />

<br />

= mgijv i f j = mfiv i = mfi ˙x i<br />

and if we assume that the force Qi is derivable from a potential function V ,thenmfi = Qi = − ∂V<br />

,so<br />

∂xi that δT<br />

δt = mfi ˙x i = Qi ˙x i = − ∂V<br />

∂xi ˙x i = − δV δ<br />

or (T + V )=0orT + V = h = constant called the energy<br />

δt δt<br />

constant of the system.<br />

Action Integral<br />

The equations of motion (2.2.34) or (2.2.24) are interpreted as describing geodesics in a space whose<br />

line-element is<br />

ds 2 =2m(h − V )gjkdx j dx k<br />

where V is the potential function for the force system and T + V = h is the energy constant of the motion.<br />

The integral of ds along a curve C between two points P1 and P2 is called an action integral and is<br />

A = √ P2<br />

2m<br />

P1<br />

<br />

dx<br />

(h − V )gjk<br />

j<br />

dτ<br />

dxk 1/2<br />

dτ<br />


where τ is a parameter used to describe the curve C. The principle of stationary action states that of all<br />

curves through the points P1 and P2 the one which makes the action an extremum is the curve specified by<br />

Newton’s second law. The extremum is usually a minimum. To show this let<br />

φ = √ <br />

dx<br />

2m (h − V )gjk<br />

j<br />

dτ<br />

in equation (2.2.36). Using the notation ˙x k = dxk<br />

dτ<br />

∂φ<br />

=2m<br />

∂ ˙x i<br />

∂φ<br />

=2m (h − V )∂gjk<br />

∂xi 2φ<br />

φ (h − V )gik ˙x k<br />

we find that<br />

dxk 1/2<br />

dτ<br />

∂V<br />

∂xi ˙x j ˙x k − 2m<br />

2φ ∂xi gjk ˙x j ˙x k .<br />

The equation (2.2.36) which describe the extremum trajectories are found to be<br />

<br />

d 2m<br />

dt φ (h − V )gik ˙x k<br />

<br />

− 2m<br />

(h − V )∂gjk<br />

2φ<br />

∂V<br />

∂xi ˙x j ˙x k + 2m<br />

φ ∂xi gjk ˙x j ˙x k =0.<br />

By changing variables from τ to t where dt<br />

dτ = √ mφ<br />

√ 2(h−V ) we find that the trajectory for an extremum must<br />

satisfy the equation<br />

m d<br />

<br />

gik<br />

dt<br />

dxk <br />

dt<br />

− m ∂gjk<br />

2 ∂xi dxj dx<br />

dt<br />

k<br />

dt<br />

∂V<br />

+ =0<br />

∂xi which are the same equations as (2.2.24). (i.e. See also the equations (2.2.22).)<br />

Dynamics of Rigid Body Motion<br />

Let us derive the equations of motion of a rigid body which is rotating due to external forces acting<br />

upon it. We neglect any translational motion of the body since this type of motion can be discerned using<br />

our knowledge of particle dynamics. The derivation of the equations of motion is restricted to Cartesian<br />

tensors and rotational motion.<br />

Consider a system of N particles rotating with angular velocity ωi, i=1, 2, 3, about a line L through<br />

the center of mass of the system. Let V (α) denote the velocity of the αth particle which has mass m (α) and<br />

position x (α)<br />

i , i =1, 2, 3 with respect to an origin on the line L. Without loss of generality we can assume<br />

that the origin of the coordinate system is also at the center of mass of the system of particles, as this choice<br />

of an origin simplifies the derivation. The velocity components for each particle is obtained by taking cross<br />

products and we can write<br />

V (α) = ω × r (α)<br />

or V (α)<br />

i<br />

(α)<br />

= eijkωjx . (2.2.37)<br />

The kinetic energy of the system of particles is written as the sum of the kinetic energies of each<br />

individual particle and is<br />

T = 1<br />

2<br />

N<br />

m (α)V<br />

α=1<br />

(α)<br />

i i<br />

(α) 1<br />

V =<br />

2<br />

N<br />

α=1<br />

m (α)eijkωjx (α)<br />

k<br />

k<br />

(α)<br />

eimnωmx n<br />

. (2.2.38)<br />

199


200<br />

Employing the e − δ identity the equation (2.2.38) can be simplified to the form<br />

T = 1<br />

2<br />

N<br />

m (α)<br />

α=1<br />

<br />

ωmωmx (α)<br />

k x(α)<br />

(α)<br />

k − ωnωkx k x(α)<br />

<br />

n .<br />

Define the second moments and products of inertia by the equation<br />

Iij =<br />

N<br />

m (α)<br />

α=1<br />

<br />

x (α)<br />

k x(α)<br />

k δij − x (α)<br />

i x(α)<br />

<br />

j<br />

(2.2.39)<br />

and write the kinetic energy in the form<br />

T = 1<br />

2 Iijωiωj. (2.2.40)<br />

Similarly, the angular momentum of the system of particles can also be represented in terms of the<br />

second moments and products of inertia. The angular momentum of a system of particles is defined as a<br />

summation of the moments of the linear momentum of each individual particle and is<br />

Hi =<br />

N<br />

α=1<br />

m (α)eijkx (α)<br />

j v(α)<br />

k =<br />

N<br />

α=1<br />

The e − δ identity simplifies the equation (2.2.41) to the form<br />

Hi = ωj<br />

α=1<br />

N<br />

m (α)<br />

m (α)eijkx (α)<br />

j<br />

(α)<br />

ekmnωmx n . (2.2.41)<br />

<br />

x (α)<br />

n x(α) n δij − x (α)<br />

j x(α)<br />

<br />

i = ωjIji. (2.2.42)<br />

The equations of motion of a rigid body is obtained by applying Newton’s second law of motion to the<br />

system of N particles. The equation of motion of the αth particle is written<br />

m (α)¨x (α)<br />

i<br />

Summing equation (2.2.43) over all particles gives the result<br />

N<br />

α=1<br />

m (α)¨x (α)<br />

i<br />

(α)<br />

= F i . (2.2.43)<br />

=<br />

N<br />

α=1<br />

F (α)<br />

i . (2.2.44)<br />

This represents the translational equations of motion of the rigid body. The equation (2.2.44) represents the<br />

rate of change of linear momentum being equal to the total external force acting upon the system. Taking<br />

produces<br />

the cross product of equation (2.2.43) with the position vector x (α)<br />

j<br />

and summing over all particles we find the equation<br />

N<br />

α=1<br />

m (α)¨x (α)<br />

t erstx (α)<br />

s = erstx (α)<br />

s F (α)<br />

t<br />

m (α)erstx (α)<br />

s ¨x(α) t =<br />

N<br />

α=1<br />

erstx (α) (α)<br />

s F t . (2.2.45)


The equations (2.2.44) and (2.2.45) represent the conservation of linear and angular momentum and can be<br />

writtenintheforms<br />

<br />

N<br />

d<br />

m (α) ˙x<br />

dt<br />

α=1<br />

(α)<br />

<br />

N<br />

r = F<br />

α=1<br />

(α)<br />

r<br />

(2.2.46)<br />

and<br />

<br />

N<br />

d<br />

m (α)erstx<br />

dt<br />

α=1<br />

(α)<br />

s ˙x (α)<br />

<br />

N<br />

t = erstx<br />

α=1<br />

(α) (α)<br />

s F t . (2.2.47)<br />

By definition we have Gr = m (α) ˙x (α)<br />

r<br />

representing the linear momentum, Fr = F (α)<br />

r<br />

the total force<br />

acting on the system of particles, Hr = m (α)erstx (α)<br />

s ˙x (α)<br />

t is the angular momentum of the system relative<br />

to the origin, and Mr = erstx (α)<br />

s F (α)<br />

t is the total moment of the system relative to the origin. We can<br />

therefore express the equations (2.2.46) and (2.2.47) in the form<br />

dGr<br />

dt = Fr (2.2.48)<br />

and<br />

dHr<br />

dt = Mr. (2.2.49)<br />

The equation (2.2.49) expresses the fact that the rate of change of angular momentum is equal to the<br />

moment of the external forces about the origin. These equations show that the motion of a system of<br />

particles can be studied by considering the motion of the center of mass of the system (translational motion)<br />

and simultaneously considering the motion of points about the center of mass (rotational motion).<br />

We now develop some relations in order to express the equations (2.2.49) in an alternate form. Toward<br />

this purpose we consider first the concepts of relative motion and angular velocity.<br />

Relative Motion and Angular Velocity<br />

Consider two different reference frames denoted by S and S. Both reference frames are Cartesian<br />

coordinates with axes xi and xi , i =1, 2, 3, respectively. The reference frame S is fixed in space and is<br />

called an inertial reference frame or space-fixed reference system of axes. The reference frame S is fixed<br />

to and rotates with the rigid body and is called a body-fixed system of axes. Again, for convenience, it<br />

is assumed that the origins of both reference systems are fixed at the center of mass of the rigid body.<br />

Further, we let the system S have the basis vectors ei,i=1, 2, 3, while the reference system S has the basis<br />

vectors êi ,i =1, 2, 3. The transformation equations between the two sets of reference axes are the affine<br />

transformations<br />

xi = ℓjixj and xi = ℓijxj (2.2.50)<br />

where ℓij = ℓij(t) are direction cosines which are functions of time t (i.e. the ℓij are the cosines of the<br />

angles between the barred and unbarred axes where the barred axes are rotating relative to the space-fixed<br />

unbarred axes.) The direction cosines satisfy the relations<br />

ℓijℓik = δjk and ℓijℓkj = δik. (2.2.51)<br />

201


202<br />

EXAMPLE 2.2-5. (Euler angles φ, θ, ψ) Consider the following sequence of transformations which<br />

are used in celestial mechanics. First a rotation about the x3 axis taking the xi axes to the yi axes<br />

⎛<br />

⎝ y1<br />

⎞ ⎛<br />

cos φ<br />

y2 ⎠ = ⎝ − sin φ<br />

sin φ<br />

cos φ<br />

⎞ ⎛<br />

0<br />

0 ⎠ ⎝<br />

y3 0 0 1<br />

x1<br />

⎞<br />

x2 ⎠<br />

x3<br />

where the rotation angle φ is called the longitude of the ascending node. Second, a rotation about the y1<br />

axis taking the yi axes to the y ′ i axes<br />

⎛<br />

⎝ y′ 1<br />

y ′ 2<br />

y ′ ⎞ ⎛<br />

1<br />

⎠ = ⎝ 0<br />

3 0<br />

0<br />

cosθ<br />

− sin θ<br />

⎞ ⎛<br />

0<br />

sin θ ⎠ ⎝<br />

cos θ<br />

y1<br />

⎞<br />

y2 ⎠<br />

y3<br />

where the rotation angle θ is called the angle of inclination of the orbital plane. Finally, a rotation about<br />

the y ′ 3 axis taking the y′ i axes to the ¯xi axes<br />

⎛<br />

⎝ ¯x1<br />

⎞ ⎛<br />

⎞ ⎛<br />

cos ψ sin ψ 0<br />

¯x2 ⎠ = ⎝ − sin ψ cos ψ 0 ⎠ ⎝<br />

¯x3 0 0 1<br />

y′ 1<br />

y ′ 2<br />

y ′ ⎞<br />

⎠<br />

3<br />

where the rotation angle ψ is called the argument of perigee. The Euler angle θ is the angle ¯x30x3, the angle<br />

φ is the angle x10y1 and ψ is the angle y10¯x1. These angles are illustrated in the figure 2.2-6. Note also that<br />

the rotation vectors associated with these transformations are vectors of magnitude ˙ φ, ˙ θ, ˙ ψ in the directions<br />

indicated in the figure 2.2-6.<br />

Figure 2.2-6. Euler angles.<br />

By combining the above transformations there results the transformation equations (2.2.50)<br />

⎛<br />

⎝ ¯x1<br />

⎞ ⎛<br />

⎞ ⎛<br />

cos ψ cos φ − cos θ sin φ sin ψ cos ψ sin φ +cosθcos φ sin ψ sin ψ sin θ<br />

¯x2 ⎠ = ⎝ − sin ψ cos φ − cos θ sin φ cos ψ − sin ψ sin φ +cosθcos φ cos ψ cos ψ sin θ ⎠ ⎝<br />

¯x3<br />

sin θ sin φ − sin θ cos φ cos θ<br />

x1<br />

⎞<br />

x2 ⎠ .<br />

x3<br />

It is left as an exercise to verify that the transformation matrix is orthogonal and the components ℓji<br />

satisfy the relations (2.2.51).


Consider the velocity of a point which is rotating with the rigid body. Denote by vi = vi(S), for<br />

i =1, 2, 3, the velocity components relative to the S reference frame and by vi = vi(S), i =1, 2, 3the<br />

velocity components of the same point relative to the body-fixed axes. In terms of the basis vectors we can<br />

write<br />

V = v1(S) ê1 + v2(S) ê2 + v3(S) ê3 = dxi<br />

dt êi<br />

as the velocity in the S reference frame. Similarly, we write<br />

(2.2.52)<br />

<br />

V = v1(S)e1 + v2(S)e2 + v3(S)e3 = dxi<br />

dt ei<br />

(2.2.53)<br />

as the velocity components relative to the body-fixed reference frame. There are occasions when it is desirable<br />

to represent V in the S frame of reference and V in the S frame of reference. In these instances we can write<br />

and<br />

V = v1(S)e1 + v2(S)e2 + v3(S)e3<br />

(2.2.54)<br />

<br />

V = v1(S) ê1 + v2(S) ê2 + v3(S) ê3. (2.2.55)<br />

Here we have adopted the notation that vi(S) are the velocity components relative to the S reference frame<br />

and vi(S) are the same velocity components relative to the S reference frame. Similarly, vi(S) denotes the<br />

velocity components relative to the S reference frame, while vi(S) denotes the same velocity components<br />

relative to the S reference frame.<br />

Here both V and V are vectors and so their components are first order tensors and satisfy the transformation<br />

laws<br />

vi(S) =ℓjivj(S) =ℓji ˙xj and vi(S) =ℓijvj(S) =ℓij ˙xj. (2.2.56)<br />

The equations (2.2.56) define the relative velocity components as functions of time t. By differentiating the<br />

equations (2.2.50) we obtain<br />

dxi<br />

dt = vi(S) =ℓji ˙xj + ˙ ℓjixj<br />

(2.2.57)<br />

and<br />

dxi<br />

dt = vi(S) =ℓij ˙xj + ˙ ℓijxj. (2.2.58)<br />

Multiply the equation (2.2.57) by ℓmi and multiply the equation (2.2.58) by ℓim and derive the relations<br />

and<br />

vm(S) =vm(S)+ℓmi ˙ ℓjixj<br />

(2.2.59)<br />

vm(S) =vm(S)+ℓim ˙ ℓijxj. (2.2.60)<br />

The equations (2.2.59) and (2.2.60) describe the transformation laws of the velocity components upon changing<br />

from the S to the S reference frame. These equations can be expressed in terms of the angular velocity<br />

by making certain substitutions which are now defined.<br />

The first order angular velocity vector ωi is related to the second order skew-symmetric angular velocity<br />

tensor ωij by the defining equation<br />

ωmn = eimnωi. (2.2.61)<br />

203


204<br />

The equation (2.2.61) implies that ωi and ωij are dual tensors and<br />

ωi = 1<br />

2 eijkωjk.<br />

Also the velocity of a point which is rotating about the origin relative to the S frame of reference is vi(S) =<br />

eijkωjxk which can also be written in the form vm(S) =−ωmkxk. Since the barred axes rotate with the rigid<br />

body, then a particle in the barred reference frame will have vm(S) = 0, since the coordinates of a point<br />

in the rigid body will be constants with respect to this reference frame. Consequently, we write equation<br />

(2.2.59) in the form 0 = vm(S)+ℓmi ˙ ℓjixj which implies that<br />

vm(S) =−ℓmi ˙ ℓjixj = −ωmkxk or ωmj = ωmj(S,S)=ℓmi ˙ ℓji.<br />

This equation is interpreted as describing the angular velocity tensor of S relative to S. Sinceωij is a tensor,<br />

it can be represented in the barred system by<br />

ωmn(S,S)=ℓimℓjnωij(S,S)<br />

= ℓimℓjnℓis ˙ ℓjs<br />

= δmsℓjn ˙ ℓjs<br />

= ℓjn ˙ ℓjm<br />

(2.2.62)<br />

By differentiating the equations (2.2.51) it is an easy exercise to show that ωij is skew-symmetric. The<br />

second order angular velocity tensor can be used to write the equations (2.2.59) and (2.2.60) in the forms<br />

vm(S) =vm(S)+ωmj(S,S)xj<br />

vm(S) =vm(S)+ωjm(S,S)xj<br />

(2.2.63)<br />

The above relations are now employed to derive the celebrated Euler’s equations of motion of a rigid body.<br />

Euler’s Equations of Motion<br />

We desire to find the equations of motion of a rigid body which is subjected to external forces. These<br />

equations are the formulas (2.2.49), and we now proceed to write these equations in a slightly different form.<br />

Similar to the introduction of the angular velocity tensor, given in equation (2.2.61), we now introduce the<br />

following tensors<br />

1. The fourth order moment of inertia tensor Imnst which is related to the second order moment of<br />

inertia tensor Iij by the equations<br />

Imnst = 1<br />

2 ejmneistIij or Iij = 1<br />

2 Ipqrseipqejrs<br />

(2.2.64)<br />

2. The second order angular momentum tensor Hjk which is related to the angular momentum vector<br />

Hi by the equation<br />

Hi = 1<br />

2 eijkHjk or Hjk = eijkHi (2.2.65)<br />

3. The second order moment tensor Mjk which is related to the moment Mi by the relation<br />

Mi = 1<br />

2 eijkMjk or Mjk = eijkMi. (2.2.66)


Now if we multiply equation (2.2.49) by erjk, thenitcanbewrittenintheform<br />

dHij<br />

dt = Mij. (2.2.67)<br />

Similarly, if we multiply the equation (2.2.42) by eimn, then it can be expressed in the alternate form<br />

Hmn = eimnωjIji = Imnstωst<br />

and because of this relation the equation (2.2.67) can be expressed as<br />

d<br />

dt (Iijstωst) =Mij. (2.2.68)<br />

We write this equation in the barred system of coordinates where Ipqrs will be a constant and consequently<br />

its derivative will be zero. We employ the transformation equations<br />

Iijst = ℓipℓjqℓsrℓtkIpqrk<br />

ωij = ℓsiℓtjωst<br />

M pq = ℓipℓjqMij<br />

and then multiply the equation (2.2.68) by ℓipℓjq and simplify to obtain<br />

d <br />

ℓipℓjq ℓiαℓjβIαβrkωrk = M pq.<br />

dt<br />

Expand all terms in this equation and take note that the derivative of the Iαβrk is zero. The expanded<br />

equation then simplifies to<br />

dωrk<br />

Ipqrk<br />

dt +(δαuδpvδβq + δpαδβuδqv) Iαβrkωrkωuv = M pq. (2.2.69)<br />

Substitute into equation (2.2.69) the relations from equations (2.2.61),(2.2.64) and (2.2.66), and then multiply<br />

by empq and simplify to obtain the Euler’s equations of motion<br />

dωi<br />

Iim<br />

dt − etmjIijωiωt = M m. (2.2.70)<br />

Dropping the bar notation and performing the indicated summations over the range 1,2,3 we find the<br />

Euler equations have the form<br />

dω1<br />

I11<br />

dt<br />

dω1<br />

I<strong>12</strong><br />

dt<br />

dω2<br />

+ I21<br />

dt<br />

dω2<br />

+ I22<br />

dt<br />

dω3<br />

+ I31<br />

dω3<br />

+ I32<br />

dt +(I13ω1 + I23ω2 + I33ω3) ω2 − (I<strong>12</strong>ω1 + I22ω2 + I32ω3) ω3 = M1<br />

dt +(I11ω1 + I21ω2 + I31ω3) ω3 − (I13ω1 + I23ω2 + I33ω3) ω1 = M2<br />

(2.2.71)<br />

dω1 dω2 dω3<br />

I13 + I23 + I33<br />

dt dt dt +(I<strong>12</strong>ω1 + I22ω2 + I32ω3) ω1 − (I11ω1 + I21ω2 + I31ω3) ω2 = M3.<br />

In the special case where the barred axes are principal axes, then Iij =0fori= j and the Euler’s<br />

equations reduces to the system of nonlinear differential equations<br />

dω1<br />

I11<br />

dt +(I33 − I22)ω2ω3 = M1<br />

dω2<br />

I22<br />

dt +(I11 − I33)ω3ω1 = M2<br />

dω3<br />

I33<br />

(2.2.72)<br />

dt +(I22 − I11)ω1ω2 = M3.<br />

In the case of constant coefficients and constant moments the solutions of the above differential equations<br />

can be expressed in terms of Jacobi elliptic functions.<br />

205


206<br />

EXERCISE 2.2<br />

◮ 1. Find a set of parametric equations for the straight line which passes through the points P1(1, 1, 1) and<br />

P2(2, 3, 4). Find the unit tangent vector to any point on this line.<br />

◮ 2. Consider the space curve x = 1<br />

2 sin2 t, y = 1 1<br />

2t − 4 sin 2t, z =sintwhere t is a parameter. Find the unit<br />

vectors T i ,Bi ,Ni ,i=1, 2, 3 at the point where t = π.<br />

◮ 3. A claim has been made that the space curve x = t, y = t2 ,z= t3 intersects the plane 11x-6y+z=6 in<br />

three distinct points. Determine if this claim is true or false. Justify your answer and find the three points<br />

of intersection if they exist.<br />

◮ 4. Find a set of parametric equations xi = xi(s1,s2),i =1, 2, 3 for the plane which passes through the<br />

points P1(3, 0, 0), P2(0, 4, 0) and P3(0, 0, 5). Find a unit normal to this plane.<br />

◮ 5. For the helix x =sint y =cost z = 2<br />

t find the equation of the tangent plane to the curve at the<br />

π<br />

point where t = π/4. Find the equation of the tangent line to the curve at the point where t = π/4.<br />

◮ 6. Express the generalized velocity and acceleration in cylindrical coordinates. Find the physical components<br />

of velocity and acceleration in cylindrical coordinates.<br />

◮ 7. Express the generalized velocity and acceleration in spherical coordinates. Find the physical components<br />

of velocity and acceleration in spherical coordinates.<br />

◮ 8. Verify the derivative ∂T<br />

∂ ˙x r = Mgrm ˙x m .<br />

◮ 9. Verify the derivative d<br />

<br />

∂T<br />

dt ∂ ˙x r<br />

<br />

◮ 10. Verify the derivative ∂T<br />

∂x<br />

1<br />

= r 2<br />

= M<br />

<br />

grm¨x m + ∂grm<br />

∂xn ˙xn ˙x m<br />

<br />

.<br />

M ∂gmn<br />

∂x r ˙xm ˙x n .<br />

◮ 11. Use the results from problems 8,9 and 10 to derive the Lagrange’s form for the equations of motion<br />

defined by equation (2.2.23).<br />

◮ <strong>12</strong>. Expand equation (2.2.39) and write out all the components of the moment of inertia tensor Iij.<br />

◮ 13. For ρ the density of a continuous material and dτ an element of volume inside a region R where the<br />

material is situated, we write ρdτ as an element of mass inside R. Find an equation which describes the<br />

center of mass of the region R.<br />

◮ 14. Use the equation (2.2.68) to derive the equation (2.2.69).<br />

◮ 15. Drop the bar notation and expand the equation (2.2.70) and derive the equations (2.2.71).<br />

◮ 16. Verify the Euler transformation, given in example 2.2-5, is orthogonal.


Figure 2.2-7. Pulley and mass system<br />

◮ 17. For the pulley and mass system illustrated in the figure 2.2-7 let<br />

a = the radius of each pulley.<br />

ℓ1 = the length of the upper chord.<br />

ℓ2 = the length of the lower chord.<br />

Neglect the weight of the pulley and find the equations of motion for the pulley mass system.<br />

◮ 18. Let φ = ds<br />

dt , where s is the arc length between two points on a curve in generalized coordinates.<br />

(a) Write the arc length in general coordinates as ds = gmn ˙x m ˙x ndt and show the integral I, defined by<br />

equation (2.2.35), represents the distance between two points on a curve.<br />

(b) Using the Euler-Lagrange equations (2.2.36) show that the shortest distance between two points in a<br />

i<br />

jk<br />

<br />

˙x j ˙x k =˙x i<br />

generalized space is the curve defined by the equations: ¨x i +<br />

dt<br />

(c) Show in the special case t = s the equations in part (b) reduce to d2xi j i dx dx<br />

+<br />

ds2 jk ds<br />

k<br />

ds =0,for<br />

i =1,...,N. An examination of equation (1.5.51) shows that the above curves are geodesic curves.<br />

(d) Show that the shortest distance between two points in a plane is a straight line.<br />

(e) Consider two points on the surface of a cylinder of radius a. Let u 1 = θ and u 2 = z denote surface<br />

coordinates in the two dimensional space defined by the surface of the cylinder. Show that the shortest<br />

distance between the points where θ =0,z=0andθ = π, z = H is L = a2π2 + H2 .<br />

◮ 19. For T = 1<br />

2mgijviv j the kinetic energy of a particle and V the potential energy of the particle show<br />

that T + V = constant.<br />

Hint: mfi = Qi = − ∂V<br />

∂x i , i =1, 2, 3 and dx i<br />

dt =˙xi = v i ,i=1, 2, 3.<br />

d 2 s<br />

dt 2<br />

ds<br />

207


208<br />

◮ 20. Define H = T + V as the sum of the kinetic energy and potential energy of a particle. The quantity<br />

H = H(xr ,pr) is called the Hamiltonian of the particle and it is expressed in terms of:<br />

• the particle position xi and<br />

• the particle momentum pi = mvi = mgij ˙x j . Here x r and pr are treated as independent variables.<br />

(a) Show that the particle momentum is a covariant tensor of rank 1.<br />

(b) Express the kinetic energy T in terms of the particle momentum.<br />

(c) Show that pi = ∂T<br />

.<br />

∂ ˙x i<br />

(d) Show that dxi ∂H<br />

dpi ∂H<br />

= and = − . These are a set of differential equations describing the<br />

dt ∂pi<br />

dt ∂xi position change and momentum change of the particle and are known as Hamilton’s equations of motion<br />

for a particle.<br />

◮ 21. Let<br />

δT i<br />

δs = κN i and δNi<br />

δs = τBi − κT i and calculate the intrinsic derivative of the cross product<br />

B i = ɛ ijk TjNk and find δBi<br />

δs in terms of the unit normal vector.<br />

◮ 22. For T the kinetic energy of a particle and V the potential energy of a particle, define the Lagrangian<br />

L = L(x i , ˙x i )=T − V = 1<br />

2 Mgij ˙x i ˙x j − V as a function of the independent variables x i , ˙x i . Define the<br />

Hamiltonian H = H(x i ,pi) =T + V = 1<br />

2M gij pipj + V, as a function of the independent variables x i ,pi,<br />

where pi is the momentum vector of the particle and M is the mass of the particle.<br />

(a) Show that pi = ∂T<br />

.<br />

∂ ˙x i<br />

(b) Show that ∂H ∂L<br />

= −<br />

∂xi ∂xi ◮ 23. When the Euler angles, figure 2.2-6, are applied to the motion of rotating objects, θ is the angle<br />

of nutation, φ is the angle of precession and ψ is the angle of spin. Take projections and show that the<br />

time derivative of the Euler angles are related to the angular velocity vector components ωx,ωy,ωz by the<br />

relations<br />

ωx = ˙ θ cos ψ + ˙ φ sin θ sin ψ<br />

ωy = − ˙ θ sin ψ + ˙ φ sin θ cos ψ<br />

ωz = ˙ ψ + ˙ φ cos θ<br />

where ωx,ωy,ωz are the angular velocity components along the x1, x2, x3 axes.<br />

◮ 24. Find the equations of motion for the compound pendulum illustrated in the figure 2.2-8.<br />

◮ 25. Let F = − GMm<br />

r3 r denote the inverse square law force of attraction between the earth and sun, with<br />

G a universal constant, M the mass of the sun, m the mass of the earth and r<br />

r a unit vector from origin<br />

at the center of the sun pointing toward the earth. (a) Write down Newton’s second law, in both vector<br />

and tensor form, which describes the motion of the earth about the sun. (b) Show that d<br />

dt (r × v) =0 and<br />

consequently r × v = r × dr<br />

dt = h =aconstant.


Figure 2.2-8. Compound pendulum<br />

◮ 26. Construct a set of axes fixed and attached to an airplane. Let the x axis be a longitudinal axis running<br />

from the rear to the front of the plane along its center line. Let the y axis run between the wing tips and<br />

let the z axis form a right-handed system of coordinates. The y axis is called a lateral axis and the z axis is<br />

called a normal axis. Define pitch as any angular motion about the lateral axis. Define roll as any angular<br />

motion about the longitudinal axis. Define yaw as any angular motion about the normal axis. Consider two<br />

sets of axes. One set is the x, y, z axes attached to and moving with the aircraft. The other set of axes is<br />

denoted X, Y, Z and is fixed in space ( an inertial set of axes). Describe the pitch, roll and yaw of an aircraft<br />

with respect to the inertial set of axes. Show the transformation is orthogonal. Hint: Consider pitch with<br />

respect to the fixed axes, then consider roll with respect to the pitch axes and finally consider yaw with<br />

respect to the roll axes. This produces three separate transformation matrices which can then be combined<br />

to describe the motions of pitch, roll and yaw of an aircraft.<br />

◮ 27. In Cartesian coordinates let Fi = Fi(x1 ,x2 ,x3 ) denote a force field and let xi = xi (t) denote a curve<br />

C. (a) Show Newton’s second law implies that along the curve C d<br />

<br />

1<br />

dt 2 m<br />

2 i dx<br />

dt<br />

<br />

= Fi(x 1 ,x 2 ,x 3 ) dxi<br />

dt<br />

(no summation on i) and hence<br />

<br />

d 1<br />

dt 2 m<br />

dx 2 2 2 1<br />

2<br />

3<br />

dx dx<br />

+ +<br />

dt dt dt<br />

<br />

= d<br />

<br />

1<br />

dt 2 mv2<br />

<br />

dx<br />

= F1<br />

1 dx<br />

+ F2<br />

dt 2 dx<br />

+ F3<br />

dt 3<br />

dt<br />

(b) Consider two points on the curve C, saypointA, x i (tA) andpointB, x i (tB) and show that the work<br />

done in moving from A to B in the force field Fi is<br />

1<br />

2 mv2<br />

tB B<br />

= Fidx<br />

tA A<br />

1 + F2dx 2 + F3dx 3<br />

209


210<br />

where the right hand side is a line integral along the path C from A to B. (c) Show that if the force field is<br />

derivable from a potential function U(x 1 ,x 2 ,x 3 ) by taking the gradient, then the work done is independent<br />

of the path C and depends only upon the end points A and B.<br />

◮ 28. Find the Lagrangian equations of motion of a spherical pendulum which consists of a bob of mass m<br />

suspended at the end of a wire of length ℓ, which is free to swing in any direction subject to the constraint<br />

that the wire length is constant. Neglect the weight of the wire and show that for the wire attached to the<br />

origin of a right handed x, y, z coordinate system, with the z axis downward, φ the angle between the wire<br />

and the z axis and θ the angle of rotation of the bob from the y axis, that there results the equations of<br />

<br />

d<br />

motion sin<br />

dt<br />

2 φ dθ<br />

<br />

d<br />

=0 and<br />

dt<br />

2 2 φ dθ<br />

− sin φ cos φ +<br />

dt2 dt<br />

g<br />

sin φ =0<br />

ℓ<br />

◮ 29. In Cartesian coordinates show the Frenet formulas can be written<br />

d T<br />

ds = δ × T,<br />

d N<br />

ds = δ × N,<br />

where δ is the Darboux vector and is defined δ = τ T + κ B.<br />

d B<br />

ds = δ × B<br />

◮ 30. Consider the following two cases for rigid body rotation.<br />

Case 1: Rigid body rotation about a fixed line which is called the fixed axis of rotation. Select a point 0<br />

on this fixed axis and denote by e a unit vector from 0 in the direction of the fixed line and denote by êR<br />

a unit vector which is perpendicular to the fixed axis of rotation. The position vector of a general point<br />

in the rigid body can then be represented by a position vector from the point 0 given by r = h e + r0 êR<br />

where h, r0 and e are all constants and the vector êR is fixed in and rotating with the rigid body.<br />

Denote by ω = dθ<br />

dt the scalar angular change with respect to time of the vector êR as it rotates about<br />

the fixed line and define the vector angular velocity as ω = d<br />

(θ e) =dθ e where θ e is defined as the<br />

dt dt<br />

vector angle of rotation.<br />

d êR<br />

(a) Show that = e × êR.<br />

dθ<br />

(b) Show that V = dr d êR d êR dθ<br />

= r0 = r0<br />

dt dt dθ dt = ω × (r0 êR) =ω × (h e + r0 êR) =ω × r.<br />

Case 2: Rigid body rotation about a fixed point 0. Construct at point 0 the unit vector ê1 which is<br />

d ê1<br />

fixed in and rotating with the rigid body. From pages 80,87 we know that must be perpendicular<br />

dt<br />

d ê1<br />

to ê1 and so we can define the vector ê2 as a unit vector which is in the direction of such that<br />

dt<br />

d ê1<br />

dt = α ê2 for some constant α. We can then define the unit vector ê3 from ê3 = ê1 × ê2.<br />

d ê3<br />

(a) Show that<br />

dt , which must be perpendicular to ê3, is also perpendicular to ê1.<br />

d ê3<br />

d ê3<br />

(b) Show that can be written as<br />

dt dt = β ê2 for some constant β.<br />

d ê2<br />

(c) From ê2 = ê3 × ê1 show that<br />

dt =(αê3 − β ê1) × ê2<br />

d ê1<br />

(d) Define ω = α ê3 − β ê1 and show that = ω × ê1,<br />

dt<br />

d ê2<br />

dt<br />

= ω × ê2,<br />

d ê3<br />

dt<br />

= ω × ê3


(e) Let r = x ê1 + y ê2 + z ê3 denote an arbitrary point within the rigid body with respect to the point 0.<br />

Show that dr<br />

= ω × r.<br />

dt<br />

Note that in Case 2 the direction of ω is not fixed as the unit vectors ê3 and ê1 are constantly changing.<br />

In this case the direction ω is called an instantaneous axis of rotation and ω, which also can change in<br />

magnitude and direction, is called the instantaneous angular velocity.<br />

211


§2.3 BASIC EQUATIONS OF CONTINUUM MECHANICS<br />

Continuum mechanics is the study of how materials behave when subjected to external influences.<br />

External influences which affect the properties of a substance are such things as forces, temperature, chemical<br />

reactions, and electric phenomena. Examples of forces are gravitational forces, electromagnetic forces, and<br />

mechanical forces. Solids deform under external forces and so deformations are studied. Fluids move under<br />

external forces and so the velocity of the fluid is studied.<br />

A material is considered to be a continuous media which is a collection of material points interconnected<br />

by internal forces (forces between the atoms making up the material). We concentrate upon the macroscopic<br />

properties rather than the microscopic properties of the material. We treat the material as a body which is<br />

homogeneous and continuous in its makeup.<br />

In this introduction we will only consider solid media and liquid media. In general, most of the ideas<br />

and concepts developed in this section can be applied to any type of material which is assumed to be a<br />

collection of material points held together by some kind of internal forces.<br />

An elastic material is one which deforms under applied forces in such a way that it will return to its<br />

original unloaded state when the applied forces are removed. When a linear relation exists between the<br />

applied forces and material displacements, then the material is called a linear elastic material. In contrast, a<br />

plastic material is one which deforms under applied forces in such a way that it does not return to its original<br />

state after removal of the applied forces. Plastic materials will always exhibit some permanent deformation<br />

after removal of the applied forces. An elastic material is called homogeneous if it has the same properties<br />

throughout. An isotropic material has the same properties, at a point, in all directions about the point.<br />

In this introduction we develop the basic mathematical equations which describe how a continuum<br />

behaves when subjected to external forces. We shall discover that there exists a set of basic equations<br />

associated with all continuous material media. These basic equations are developed for linear elastic materials<br />

and applied to solids and fluids in later sections.<br />

Introduction to Elasticity<br />

Take a rubber band, which has a rectangular cross section, and mark on it a parallelepiped having a<br />

length ℓ, awidthwand a height h, as illustrated in the figure 2.3-1.<br />

Now apply a force F to both ends of the parallelepiped cross section on the rubber band and examine<br />

what happens to the parallelepiped. You will see that:<br />

1. ℓ increases by an amount ∆ℓ.<br />

2. w decreases by an amount ∆w.<br />

3. h decreases by an amount ∆h.<br />

There are many materials which behave in a manner very similar to the rubber band. Most materials,<br />

when subjected to tension forces will break if the change ∆ℓ is only one or two percent of the original length.<br />

The above example introduces us to several concepts which arise in the study of materials when they are<br />

subjected to external forces. The first concept is that of strain which is defined as<br />

strain =<br />

change in length<br />

, (dimensionless).<br />

original length<br />

211


2<strong>12</strong><br />

Figure 2.3-1. Section of a rubber band<br />

When the force F is applied to our rubber band example there arises the strains<br />

∆ℓ<br />

ℓ ,<br />

∆w<br />

w ,<br />

The second concept introduced by our simple example is stress. Stress is defined as a force per unit area. In<br />

particular,<br />

Force<br />

force<br />

stress =<br />

, with dimension of<br />

Area over which force acts unit area .<br />

We will be interested in studying stress and strain in homogeneous, isotropic materials which are in equilibrium<br />

with respect to the force system acting on the material.<br />

Hooke’s Law<br />

For linear elastic materials, where the forces are all one dimensional, the stress and strains are related<br />

by Hooke’s law which has two parts. The Hooke’s law, part one, states that stress is proportional to strain<br />

in the stretch direction, where the Young’s modulus E is the proportionality constant. This is written<br />

<br />

F ∆ℓ<br />

Hooke’s law part 1<br />

= E . (2.3.1)<br />

A ℓ<br />

A graph of stress vs strain is a straight line with slope E in the linear elastic range of the material.<br />

The Hooke’s law, part two, involves the fact that there is a strain contraction perpendicular to the<br />

stretch direction. The strain contraction is the same for both the width and height and is proportional to<br />

the strain in the stretch direction. The proportionality constant being the Poisson’s ratio ν.<br />

Hooke’s law part 2<br />

∆w<br />

w<br />

= ∆h<br />

h<br />

∆h<br />

h .<br />

= −ν ∆ℓ<br />

ℓ<br />

, 0


Figure 2.3-2. Typical Stress-strain curve.<br />

Consider a typical stress-strain curve, such as the one illustrated in the figure 2.3-2, which is obtained<br />

by placing a material in the shape of a rod or wire in a machine capable of performing tensile straining at a<br />

low rate. The engineering stress is the tensile force F divided by the original cross sectional area A0. Note<br />

that during a tensile straining the cross sectional area A of the sample is continually changing and getting<br />

smaller so that the actual stress will be larger than the engineering stress. Observe in the figure 2.3-2 that<br />

the stress-strain relation remains linear up to a point labeled the proportional limit. For stress-strain points<br />

in this linear region the Hooke’s law holds and the material will return to its original shape when the loading<br />

is removed. For points beyond the proportional limit, but less than the yield point, the material no longer<br />

obeys Hooke’s law. In this nonlinear region the material still returns to its original shape when the loading<br />

is removed. The region beyond the yield point is called the plastic region. At the yield point and beyond,<br />

there is a great deal of material deformation while the loading undergoes only small changes. For points<br />

in this plastic region, the material undergoes a permanent deformation and does not return to its original<br />

shape when the loading is removed. In the plastic region there usually occurs deformation due to slipping of<br />

atomic planes within the material. In this introductory section we will restrict our discussions of material<br />

stress-strain properties to the linear region.<br />

EXAMPLE 2.3-1. (One dimensional elasticity) Consider a circular rod with cross sectional area A<br />

which is subjected to an external force F applied to both ends. The figure 2.3-3 illustrates what happens to<br />

the rod after the tension force F is applied. Consider two neighboring points P and Q on the rod, where P<br />

is at the point x and Q is at the point x +∆x. When the force F is applied to the rod it is stretched and<br />

P moves to P ′ and Q moves to Q ′ . We assume that when F is applied to the rod there is a displacement<br />

function u = u(x, t) which describes how each point in the rod moves as a function of time t. If we know the<br />

displacement function u = u(x, t) we would then be able to calculate the following distances in terms of the<br />

displacement function<br />

PP ′ = u(x, t), 0P ′ = x + u(x, t), QQ ′ = u(x +∆x, t) 0Q ′ = x +∆x + u(x +∆x, t).<br />

213


214<br />

Figure 2.3-3. One dimensional rod subjected to tension force<br />

The strain associated with the distance ℓ =∆x = PQ is<br />

e = ∆ℓ<br />

ℓ = P ′ Q ′ − PQ<br />

=<br />

PQ<br />

(0Q′ − 0P ′ ) − (0Q − 0P )<br />

PQ<br />

[x +∆x + u(x +∆x, t) − (x + u(x, t))] − [(x +∆x) − x]<br />

e =<br />

∆x<br />

u(x +∆x, t) − u(x, t)<br />

e = .<br />

∆x<br />

Use the Hooke’s law part(i) and write<br />

F<br />

A<br />

Taking the limit as ∆x approaches zero we find that<br />

u(x +∆x, t) − u(x, t)<br />

= E .<br />

∆x<br />

F<br />

A<br />

∂u(x, t)<br />

= E .<br />

∂x<br />

Hence, the stress is proportional to the spatial derivative of the displacement function.<br />

Normal and Shearing Stresses<br />

Let us consider a more general situation in which we have some material which can be described as<br />

having a surface area S which encloses a volume V. Assume that the density of the material is ϱ and the<br />

material is homogeneous and isotropic. Further assume that the material is subjected to the forces b and t (n)<br />

where b is a body force per unit mass [force/mass], and t (n) is a surface traction per unit area [force/area].<br />

The superscript (n) on the vector is to remind you that we will only be interested in the normal component<br />

of the surface forces. We will neglect body couples, surface couples, and concentrated forces or couples that<br />

act at a single point. If the forces described above are everywhere continuous we can calculate the resultant<br />

force F and resultant moment M acting on the material by constructing various surface and volume integrals<br />

which sum the forces acting upon the material. In particular, the resultant force F acting on our material<br />

can be described by the surface and volume integrals:<br />

<br />

F = t (n) <br />

dS + ϱbdτ (2.3.3)<br />

S<br />

V


Figure 2.3-4. Stress vectors acting upon an element of volume<br />

which is a summation of all the body forces and surface tractions acting upon our material. Here ϱ is the<br />

density of the material, dS is an element of surface area, and dτ is an element of volume.<br />

The resultant moment M about the origin is similarly expressed as<br />

<br />

M =<br />

S<br />

r × t (n) <br />

dS + ϱ(r ×<br />

V<br />

b) dτ. (2.3.4)<br />

The global motion of the material is governed by the Euler equations of motion.<br />

• The time rate of change of linear momentum equals the resultant force or<br />

<br />

d<br />

ϱvdτ =<br />

dt V<br />

<br />

F = t<br />

S<br />

(n) <br />

dS + ϱ<br />

V<br />

bdτ. (2.3.5)<br />

This is a statement concerning the conservation of linear momentum.<br />

• The time rate of change of angular momentum equals the resultant moment or<br />

<br />

d<br />

ϱr × vdτ =<br />

dt V<br />

<br />

M = r × t<br />

S<br />

(n) <br />

dS + ϱ(r ×<br />

V<br />

b) dτ. (2.3.6)<br />

This is a statement concerning conservation of angular momentum.<br />

The Stress Tensor<br />

Define the stress vectors<br />

t 1 = σ 11 ê1 + σ <strong>12</strong> ê2 + σ 13 ê3<br />

t 2 = σ 21 ê1 + σ 22 ê2 + σ 23 ê3<br />

t 3 = σ 31 ê1 + σ 32 ê2 + σ 33 ê3,<br />

(2.3.7)<br />

where σ ij ,i,j=1, 2, 3 is the stress tensor acting at each point of the material. The index i indicates the<br />

coordinate surface x i = a constant, upon which t i acts. The second index j denotes the direction associated<br />

with the components of t i .<br />

215


216<br />

Figure 2.3-5. Stress distribution at a point<br />

For i =1, 2, 3 we adopt the convention of sketching the components of t i in the positive directions if<br />

the exterior normal to the surface xi = constant also points in the positive direction. This gives rise to the<br />

figure 2.3-4 which illustrates the stress vectors acting upon an element of volume in rectangular Cartesian<br />

coordinates. The components σ 11 ,σ 22 ,σ 33 are called normal stresses while the components σ ij ,i= j are<br />

called shearing stresses. The equations (2.3.7) can be written in the more compact form using the indicial<br />

notation as<br />

t i = σ ij êj, i,j =1, 2, 3. (2.3.8)<br />

If we know the stress distribution at three orthogonal interfaces at a point P in a solid body, we can then<br />

determine the stress at the point P with respect to any plane passing through the point P. With reference to<br />

the figure 2.3-5, consider an arbitrary plane passing through the point P which lies within the material body<br />

being considered. Construct the elemental tetrahedron with orthogonal axes parallel to the x 1 = x, x 2 = y<br />

and x 3 = z axes. In this figure we have the following surface tractions:<br />

−t 1<br />

−t 2<br />

−t 3<br />

t (n)<br />

on the surface 0BC<br />

on the surface 0AC<br />

on the surface 0AB<br />

on the surface ABC<br />

The superscript parenthesis n is to remind you that this surface traction depends upon the orientation of<br />

the plane ABC which is determined by a unit normal vector having the direction cosines n1,n2 and n3.


Let<br />

These surface areas are related by the relations<br />

∆S1 = the surface area 0BC<br />

∆S2 = the surface area 0AC<br />

∆S3 = the surface area 0AB<br />

∆S = the surface area ABC .<br />

∆S1 = n1∆S, ∆S2 = n2∆S, ∆S3 = n3∆S (2.3.9)<br />

which can be thought of as projections of ∆S upon the planes xi =constant for i =1, 2, 3.<br />

Cauchy Stress Law<br />

Let t j (n) denote the components of the surface traction on the surface ABC. Thatis,welet<br />

t (n) = t 1(n) ê1 + t 2(n) ê2 + t 3(n) ê3 = t j (n) êj. (2.3.10)<br />

It will be demonstrated that the components t j (n) of the surface traction forces t (n) associated with a plane<br />

through P and having the unit normal with direction cosines n1,n2 and n3, must satisfy the relations<br />

t j (n) = ni σ ij , i,j =1, 2, 3. (2.3.11)<br />

This relation is known as the Cauchy stress law.<br />

Proof: Sum the forces acting on the elemental tetrahedron in the figure 2.3-5. If the body is in equilibrium,<br />

then the sum of these forces must equal zero or<br />

(−t 1 ∆S1)+(−t 2 ∆S2)+(−t 3 ∆S3)+t (n) ∆S =0. (2.3.<strong>12</strong>)<br />

The relations in the equations (2.3.9) are used to simplify the sum of forces in the equation (2.3.<strong>12</strong>). It is<br />

readily verified that the sum of forces simplifies to<br />

Substituting in the relations from equation (2.3.8) we find<br />

or in component form<br />

which is the Cauchy stress law.<br />

t (n) = n1 t 1 + n2 t 2 + n3 t 3 = ni t i . (2.3.13)<br />

t (n) = t j (n) êj = niσ ij êj, i,j =1, 2, 3 (2.3.14)<br />

t j (n) = niσ ij<br />

(2.3.15)<br />

217


218<br />

Conservation of Linear Momentum<br />

Let R denote a region in space where there exists a material volume with density ϱ having surface<br />

tractions and body forces acting upon it. Let vi denote the velocity of the material volume and use Newton’s<br />

second law to set the time rate of change of linear momentum equal to the forces acting upon the volume as<br />

in (2.3.5). We find<br />

<br />

δ<br />

ϱv<br />

δt R<br />

j <br />

dτ = σ<br />

S<br />

ij <br />

ni dS + ϱb<br />

R<br />

j dτ.<br />

Here dτ is an element of volume, dS is an element of surface area, bj are body forces per unit mass, and σij are the stresses. Employing the Gauss divergence theorem, the surface integral term is replaced by a volume<br />

integral and Newton’s second law is expressed in the form<br />

<br />

R<br />

j j ij<br />

ϱf − ϱb − σ ,i dτ =0, (2.3.16)<br />

where f j is the acceleration from equation (1.4.54). Since R is an arbitrary region, the equation (2.3.16)<br />

implies that<br />

σ ij ,i + ϱb j = ϱf j . (2.3.17)<br />

This equation arises from a balance of linear momentum and represents the equations of motion for material<br />

in a continuum. If there is no velocity term, then equation (2.3.17) reduces to an equilibrium equation which<br />

can be written<br />

This equation can also be written in the covariant form<br />

σ ij ,i + ϱb j =0. (2.3.18)<br />

g si σms,i + ϱbm =0,<br />

which reduces to σij,j + ϱbi = 0 in Cartesian coordinates. The equation (2.3.18) is an equilibrium equation<br />

and is one of our fundamental equations describing a continuum.<br />

Conservation of Angular Momentum<br />

The conservation of angular momentum equation (2.3.6) has the Cartesian tensors representation<br />

<br />

<br />

<br />

d<br />

ϱeijkxjvk dτ = eijkxjσpknp dS + ϱeijkxjbk dτ. (2.3.19)<br />

dt R<br />

S<br />

R<br />

Employing the Gauss divergence theorem, the surface integral term is replaced by a volume integral to obtain<br />

<br />

eijkϱ d<br />

dt (xjvk)<br />

<br />

− eijk ϱxjbk + ∂<br />

<br />

(xjσpk) dτ =0.<br />

∂xp (2.3.20)<br />

R<br />

Since equation (2.3.20) must hold for all arbitrary volumes R we conclude that<br />

eijkϱ d<br />

dt (xjvk)<br />

<br />

<br />

∂σpk<br />

=eijk ϱxjbk + xj + σjk<br />

∂xp


Figure 2.3-6. Shearing parallel to the y axis<br />

which can be rewritten in the form<br />

<br />

eijk σjk + xj( ∂σpk<br />

∂xp + ϱbk − ϱ dvk<br />

<br />

) − ϱvjvk =0.<br />

dt<br />

(2.3.21)<br />

In the equation (2.3.21) the middle term is zero because of the equation (2.3.17). Also the last term in<br />

(2.3.21) is zero because eijkvjvk represents the cross product of a vector with itself. The equation (2.3.21)<br />

therefore reduces to<br />

eijkσjk =0, (2.3.22)<br />

which implies (see exercise 1.1, problem 22) that σij = σji for all i and j. Thus, the conservation of angular<br />

momentum requires that the stress tensor be symmetric. Consequently, there are only 6 independent stress<br />

components to be determined. This is another fundamental law for a continuum.<br />

Strain in Two Dimensions<br />

Consider the matrix equation <br />

x<br />

=<br />

y<br />

1 0<br />

β 1<br />

<br />

x<br />

y<br />

(2.3.23)<br />

which can be used to transform points (x, y) topoints(x, y). When this transformation is applied to the<br />

unit square illustrated in the figure 2.3-6(a) we obtain the geometry illustrated in the figure 2.3-6(b) which<br />

represents a shearing parallel to the y axis. If β is very small, we can use the approximation tan β ≈ β and<br />

then this transformation can be thought of as a rotation of the element P1P2 through an angle β to the<br />

position P ′ 1 P ′ 2<br />

when the barred axes are placed atop the unbarred axes.<br />

Similarly, the matrix equation <br />

x<br />

=<br />

y<br />

1 α<br />

0 1<br />

<br />

x<br />

y<br />

(2.3.24)<br />

can be used to represent a shearing of the unit square parallel to the x axis as illustrated in the figure<br />

2.3-7(b).<br />

219


220<br />

Figure 2.3-7. Shearing parallel to the x axis<br />

Figure 2.3-8. Shearing parallel to x and y axes<br />

Again, if α is very small, we may use the approximation tan α ≈ α and interpret α as an angular rotation<br />

of the element P1P4 to the position P ′ 1P ′ 4 . Now let us multiply the matrices given in equations (2.3.23) and<br />

(2.3.24). Note that the order of multiplication is important as can be seen by an examination of the products<br />

<br />

x 1 0 1 α x 1 α x<br />

=<br />

=<br />

y β 1 0 1 y β 1+αβ y<br />

<br />

x 1 α 1 0 x 1+αβ α x<br />

=<br />

=<br />

.<br />

y 0 1 β 1 y β 1 y<br />

(2.3.25)<br />

In equation (2.3.25) we will assume that the product αβ is very, very small and can be neglected. Then the<br />

order of matrix multiplication will be immaterial and the transformation equation (2.3.25) will reduce to<br />

<br />

x 1 α x<br />

=<br />

. (2.3.26)<br />

y β 1 y<br />

Applying this transformation to our unit square we obtain the simultaneous shearing parallel to both the x<br />

and y axes as illustrated in the figure 2.3-8.<br />

This transformation can then be interpreted as the superposition of the two shearing elements depicted<br />

in the figure 2.3-9.<br />

For comparison, we consider also the transformation equation<br />

<br />

x 1 0 x<br />

=<br />

(2.3.27)<br />

y −α 1 y


Figure 2.3-9. Superposition of shearing elements<br />

Figure 2.3-10. Rotation of element P1P2<br />

where α is very small. Applying this transformation to the unit square previously considered we obtain the<br />

results illustrated in the figure 2.3-10.<br />

Note the difference in the direction of shearing associated with the transformation equations (2.3.27)<br />

and (2.3.23) illustrated in the figures 2.3-6 and 2.3-10. If the matrices appearing in the equations (2.3.24)<br />

and (2.3.27) are multiplied and we neglect product terms because α is assumed to be very small, we obtain<br />

the matrix equation<br />

<br />

x 1 α x 1 0 x 0 α x<br />

=<br />

=<br />

+<br />

. (2.3.28)<br />

y −α 1 y 0 1 y −α 0 y<br />

<br />

identity<br />

rotation<br />

This can be interpreted as a superposition of the transformation equations (2.3.24) and (2.3.27) which<br />

represents a rotation of the unit square as illustrated in the figure 2.3-11.<br />

The matrix on the right-hand side of equation (2.3.28) is referred to as a rotation matrix. The ideas<br />

illustrated by the above simple transformations will appear again when we consider the transformation of an<br />

arbitrary small element in a continuum when it under goes a strain. In particular, we will be interested in<br />

extracting the rigid body rotation from a deformed element and treating this rotation separately from the<br />

strain displacement.<br />

221


222<br />

Transformation of an Arbitrary Element<br />

Figure 2.3-11. Rotation of unit square<br />

In two dimensions, we consider a rectangular element ABCD as illustrated in the figure 2.3-<strong>12</strong>.<br />

Let the points ABCD have the coordinates<br />

and denote by<br />

A(x, y), B(x +∆x, y), C(x, y +∆y), D(x +∆x, y +∆y) (2.3.29)<br />

u = u(x, y), v = v(x, y)<br />

the displacement field associated with each of the points in the material continuum when it undergoes a<br />

deformation. Assume that the deformation of the element ABCD in figure 2.3-<strong>12</strong> can be represented by the<br />

matrix equation <br />

x b11 b<strong>12</strong> x<br />

=<br />

(2.3.30)<br />

y b21 b22 y<br />

where the coefficients bij,i,j =1, 2, 3 are to be determined. Let us define u = u(x, y) as the horizontal<br />

displacement of the point (x, y) andv = v(x, y) as the vertical displacement of the same point. We can now<br />

express the displacement of each of the points A, B, C and D in terms of the displacement field u = u(x, y)<br />

and v = v(x, y). Consider first the displacement of the point A to A ′ . Here the coordinates (x, y) deformto<br />

the new coordinates<br />

x = x + u, y = y + v.<br />

That is, the coefficients bij must be chosen such that the equation<br />

<br />

x + u<br />

=<br />

y + v<br />

b11 b<strong>12</strong><br />

b21 b22<br />

<br />

x<br />

y<br />

(2.3.31)<br />

is satisfied. We next examine the displacement of the point B to B ′ . This displacement is described by the<br />

coordinates (x +∆x, y) transforming to (x, y), where<br />

x = x +∆x + u(x +∆x, y), y = y + v(x +∆x, y). (2.3.32)


Figure 2.3-<strong>12</strong>. Displacement of element ABCD to A ′ B ′ C ′ D ′<br />

Expanding u and v in (2.3.32) in Taylor series about the point (x, y) we find<br />

x = x +∆x + u + ∂u<br />

∆x + h.o.t.<br />

∂x<br />

y = y + v + ∂v<br />

∆x + h.o.t.,<br />

∂x<br />

∂u<br />

x + u +∆x + ∂x∆x y + v + ∂v<br />

∂x∆x <br />

=<br />

(2.3.33)<br />

where h.o.t. denotes higher order terms which have been neglected. The equations (2.3.33) require that the<br />

coefficients bij satisfy the matrix equation<br />

<br />

b11 b<strong>12</strong> x +∆x<br />

. (2.3.34)<br />

y<br />

b21 b22<br />

223


224<br />

The displacement of the point C to C ′ is described by the coordinates (x, y +∆y) transforming to (x, y)<br />

where<br />

x = x + u(x, y +∆y), y = y +∆y + v(x, y +∆y). (2.3.35)<br />

Again we expand the displacement field components u and v in a Taylor series about the point (x, y) and<br />

find<br />

x = x + u + ∂u<br />

∆y + h.o.t.<br />

∂y<br />

y = y +∆y + v + ∂v<br />

∆y + h.o.t.<br />

∂y<br />

(2.3.36)<br />

This equation implies that the coefficients bij must be chosen such that<br />

∂u x + u + ∂y ∆y<br />

y + v +∆y + ∂v<br />

∂y∆y <br />

=<br />

b11 b<strong>12</strong><br />

b21 b22<br />

<br />

x<br />

. (2.3.37)<br />

y +∆y<br />

Finally, it can be verified that the point D with coordinates (x +∆x, y +∆y) moves to the point D ′ with<br />

coordinates<br />

x = x +∆x + u(x +∆x, y +∆y), y = y +∆y + v(x +∆x, y +∆y). (2.3.38)<br />

Expanding u and v in a Taylor series about the point (x, y) we find the coefficients bij must be chosen to<br />

satisfy the matrix equation<br />

∂u ∂u<br />

x +∆x + u + ∂x∆x +<br />

y +∆y + v + ∂v<br />

∂x<br />

∂y∆y <br />

=<br />

∆x + ∂v<br />

∂y ∆y<br />

b11 b<strong>12</strong><br />

b21 b22<br />

<br />

x +∆x<br />

. (2.3.39)<br />

y +∆y<br />

The equations (2.3.31),(2.3.34),(2.3.37) and (2.3.39) give rise to the simultaneous equations<br />

b11x + b<strong>12</strong>y = x + u<br />

b21x + b22y = y + v<br />

b11(x +∆x)+b<strong>12</strong>y = x + u +∆x + ∂u<br />

∂x ∆x<br />

b21(x +∆x)+b22y = y + v + ∂v<br />

∂x ∆x<br />

b11x + b<strong>12</strong>(y +∆y) =x + u + ∂u<br />

∂y ∆y<br />

b21x + b22(y +∆y) =y + v +∆y + ∂v<br />

∂y ∆y<br />

b11(x +∆x)+b<strong>12</strong>(y +∆y) =x +∆x + u + ∂u ∂u<br />

∆x +<br />

∂x ∂y ∆y<br />

b21(x +∆x)+b22(y +∆y) =y +∆y + v + ∂v ∂v<br />

∆x +<br />

∂x ∂y ∆y.<br />

It is readily verified that the system of equations (2.3.40) has the solution<br />

b11 =1+ ∂u<br />

∂x<br />

b21 = ∂v<br />

∂x<br />

b<strong>12</strong> = ∂u<br />

∂y<br />

b22 =1+ ∂v<br />

∂y .<br />

(2.3.40)<br />

(2.3.41)


Figure 2.3-13. Change in 45 ◦ line<br />

Hence the transformation equation (2.3.30) can be written as<br />

∂u ∂u<br />

x 1+ ∂x ∂y<br />

=<br />

y<br />

∂v<br />

∂x<br />

1+ ∂v<br />

∂y<br />

<br />

x<br />

. (2.3.42)<br />

y<br />

A physical interpretation associated with this transformation is obtained by writing it in the form:<br />

<br />

x 1 0 x e11 e<strong>12</strong> x ω11 ω<strong>12</strong> x<br />

=<br />

+<br />

+<br />

, (2.3.43)<br />

y 0 1 y e21 e22 y ω21 ω22 y<br />

<br />

identity<br />

strain matrix<br />

rotation matrix<br />

where<br />

e11 = ∂u<br />

∂x<br />

e<strong>12</strong> = 1<br />

<br />

∂v ∂u<br />

+<br />

2 ∂x ∂y<br />

e21 = 1<br />

<br />

∂u ∂v<br />

+<br />

2 ∂y ∂x<br />

e22 = ∂v<br />

∂y<br />

are the elements of a symmetric matrix called the strain matrix and<br />

ω11 =0<br />

ω21 = 1<br />

ω<strong>12</strong> =<br />

∂v ∂u<br />

−<br />

2 ∂x ∂y<br />

1<br />

<br />

∂u ∂v<br />

−<br />

2 ∂y ∂x<br />

ω22 =0<br />

are the elements of a skew symmetric matrix called the rotation matrix.<br />

The strain per unit length in the x-direction associated with the point A in the figure 2.3-<strong>12</strong> is<br />

e11 =<br />

∂u ∆x + ∂x∆x − ∆x<br />

=<br />

∆x<br />

∂u<br />

∂x<br />

and the strain per unit length of the point A in the y direction is<br />

e22 =<br />

∂v ∆y + ∂y∆y − ∆y<br />

∆y<br />

(2.3.44)<br />

(2.3.45)<br />

(2.3.46)<br />

= ∂v<br />

. (2.3.47)<br />

∂y<br />

These are the terms along the main diagonal in the strain matrix. The geometry of the figure 2.3-<strong>12</strong> implies<br />

that<br />

∂v<br />

∂x tan β =<br />

∆x<br />

∆x + ∂u<br />

∂u<br />

∂y<br />

, and tan α =<br />

∂x∆x ∆y<br />

∆y + ∂v . (2.3.48)<br />

∂y∆y For small derivatives associated with the displacements u and v it is assumed that the angles α and β are<br />

small and the equations (2.3.48) therefore reduce to the approximate equations<br />

tan β ≈ β = ∂v<br />

tan α ≈ α =<br />

∂x<br />

∂u<br />

. (2.3.49)<br />

∂y<br />

For a physical interpretation of these terms we consider the deformation of a small rectangular element which<br />

undergoes a shearing as illustrated in the figure 2.3-13.<br />

225


226<br />

The quantity<br />

Figure 2.3-14. Displacement field due to state of strain<br />

α + β =<br />

<br />

∂u ∂v<br />

+ =2e<strong>12</strong> =2e21<br />

∂y ∂x<br />

(2.3.50)<br />

is the change from a ninety degree angle due to the deformation and hence we can write 1<br />

2 (α+β) =e<strong>12</strong> = e21<br />

as representing a change from a 45 ◦ angle due to the deformation. The quantities e21,e<strong>12</strong> are called the<br />

shear strains and the quantity<br />

is called the shear angle.<br />

γ<strong>12</strong> =2e<strong>12</strong><br />

(2.3.51)<br />

In the equation (2.3.45), the quantities ω21 = −ω<strong>12</strong> are the elements of the rigid body rotation matrix<br />

and are interpreted as angles associated with a rotation. The situation is analogous to the transformations<br />

and figures for the deformation of the unit square which was considered earlier.<br />

Strain in Three Dimensions<br />

The development of strain in three dimensions is approached from two different viewpoints. The first<br />

approach considers the derivation using Cartesian tensors and the second approach considers the derivation<br />

of strain using generalized tensors.<br />

Cartesian Tensor Derivation of Strain.<br />

Consider a material which is subjected to external forces such that all the points in the material undergo<br />

a deformation. Let (y1,y2,y3) denote a set of orthogonal Cartesian coordinates, fixed in space, which is<br />

used to describe the deformations within the material. Further, let ui = ui(y1,y2,y3),i =1, 2, 3denotea<br />

displacement field which describes the displacement of each point within the material. With reference to the<br />

figure 2.3-14 let P and Q denote two neighboring points within the material while it is in an unstrained state.<br />

These points move to the points P ′ and Q ′ when the material is in a state of strain. We let yi,i =1, 2, 3<br />

represent the position vector to the general point P in the material, which is in an unstrained state, and<br />

denote by yi + ui,i=1, 2, 3 the position vector of the point P ′ when the material is in a state of strain.


For Q a neighboring point of P which moves to Q ′ when the material is in a state of strain, we have<br />

from the figure 2.3-14 the following vectors:<br />

position of P : yi, i =1, 2, 3<br />

position of P ′ : yi + ui(y1,y2,y3), i =1, 2, 3<br />

position of Q : yi +∆yi, i =1, 2, 3<br />

position of Q ′ : yi +∆yi + ui(y1 +∆y1,y2 +∆y2,y3 +∆y3), i =1, 2, 3<br />

(2.3.52)<br />

Employing our earlier one dimensional definition of strain, we define the strain associated with the point P<br />

L − L0<br />

in the direction PQ as e = , where L0 = PQ and L = P ′ Q ′ . To calculate the strain we need to first<br />

L0<br />

calculate the distances L0 and L. The quantities L2 0 and L2 are easily calculated by considering dot products<br />

of vectors. For example, we have L 2 0 =∆yi∆yi, and the distance L = P ′ Q ′ is the magnitude of the vector<br />

yi +∆yi + ui(y1 +∆y1,y2 +∆y2,y3 +∆y3) − (yi + ui(y1,y2,y3)), i =1, 2, 3.<br />

Expanding the quantity ui(y1 +∆y1,y2 +∆y2,y3 +∆y3) in a Taylor series about the point P and neglecting<br />

higher order terms of the expansion we find that<br />

L 2 =(∆yi + ∂ui<br />

∆ym)(∆yi +<br />

∂ym<br />

∂ui<br />

∆yn).<br />

∂yn<br />

Expanding the terms in this expression produces the equation<br />

L 2 =∆yi∆yi + ∂ui<br />

∆yi∆yn +<br />

∂yn<br />

∂ui<br />

∆ym∆yi +<br />

∂ym<br />

∂ui<br />

∂ym<br />

∂ui<br />

∆ym∆yn.<br />

∂yn<br />

Note that L and L0 are very small and so we express the difference L 2 − L 2 0 in terms of the strain e. We can<br />

write<br />

L 2 − L 2 0 =(L + L0)(L − L0) =(L − L0 +2L0)(L − L0) =(e +2)eL 2 0 .<br />

Now for e very small, and e2 negligible, the above equation produces the approximation<br />

eL 2 0 ≈ L2 − L2 0<br />

=<br />

2<br />

1<br />

<br />

∂um<br />

+<br />

2 ∂yn<br />

∂un<br />

+<br />

∂ym<br />

∂ur<br />

<br />

∂ur<br />

∆ym∆yn.<br />

∂ym ∂yn<br />

The quantities<br />

emn = 1<br />

<br />

∂um<br />

+<br />

2 ∂yn<br />

∂un<br />

+<br />

∂ym<br />

∂ur<br />

<br />

∂ur<br />

(2.3.53)<br />

∂ym ∂yn<br />

is called the Green strain tensor or Lagrangian strain tensor. To show that eij is indeed a tensor, we consider<br />

the transformation yi = ℓijyj +bi, where ℓjiℓki = δjk = ℓijℓik. Note that from the derivative relation ∂yi<br />

∂y = ℓij<br />

j<br />

and the transformation equations ui = ℓijuj,i =1, 2, 3 we can express the strain in the barred system of<br />

coordinates. Performing the necessary calculations produces<br />

eij = 1<br />

<br />

∂ui<br />

+<br />

2 ∂yj ∂uj<br />

+<br />

∂yi ∂ur<br />

<br />

∂ur<br />

∂yi ∂yj = 1<br />

<br />

∂<br />

(ℓikuk)<br />

2 ∂yn<br />

∂yn<br />

+<br />

∂yj ∂<br />

(ℓjkuk)<br />

∂ym<br />

∂ym<br />

+<br />

∂yi ∂<br />

(ℓrsus)<br />

∂yk<br />

∂yk ∂<br />

(ℓrmum)<br />

∂yi ∂yt<br />

∂yt<br />

<br />

∂yj = 1<br />

<br />

<br />

∂um ∂uk<br />

∂us ∂up<br />

ℓimℓnj + ℓjkℓmi + ℓrsℓrpℓkiℓtj<br />

2 ∂yn ∂ym<br />

∂yk ∂yt<br />

= 1<br />

<br />

∂um<br />

+<br />

2 ∂yn<br />

∂un<br />

+<br />

∂ym<br />

∂us<br />

<br />

∂us<br />

ℓimℓnj<br />

∂ym ∂yn<br />

or eij = emnℓimℓnj.Consequently, the strain eij transforms like a second order Cartesian tensor.<br />

227


228<br />

Lagrangian and Eulerian Systems<br />

Let xi denote the initial position of a material particle in a continuum. Assume that at a later time the<br />

particle has moved to another point whose coordinates are xi . Both sets of coordinates are referred to the<br />

same coordinate system. When the final position can be expressed as a function of the initial position and<br />

time we can write x i = x i (x 1 , x 2 , x 3 ,t). Whenever the changes of any physical quantity is represented in terms<br />

of its initial position and time, the representation is referred to as a Lagrangian or material representation of<br />

the quantity. This can be thought of as a transformation of the coordinates. When the Jacobian J( x<br />

x )ofthis<br />

transformation is different from zero, the above set of equations have a unique inverse x i = x i (x 1 ,x 2 ,x 3 ,t),<br />

where the position of the particle is now expressed in terms of its instantaneous position and time. Such a<br />

representation is referred to as an Eulerian or spatial description of the motion.<br />

Let (x1, x2, x3) denote the initial position of a particle whose motion is described by xi = xi(x1, x2, x3,t),<br />

then ui = xi − xi denotes the displacement vector which can by represented in a Lagrangian or Eulerian<br />

form. For example, if<br />

x1 =2(x1 − x2)(e t − 1) + (x2 − x1)(e −t − 1) + x1<br />

x2 =(x1 − x2)(e t − 1) + (x2 − x1)(e −t − 1) + x2<br />

x3 = x3<br />

then the displacement vector can be represented in the Lagrangian form<br />

or the Eulerian form<br />

u1 =2(x1 − x2)(e t − 1) + (x2 − x1)(e −t − 1)<br />

u2 =(x1 − x2)(e t − 1) + (x2 − x1)(e −t − 1)<br />

u3 =0<br />

u1 = x1 − (2x2 − x1)(1 − e −t ) − (x1 − x2)(e −2t − e −t ) − x1e −t<br />

u2 = x2 − (2x2 − x1)(1 − e −t ) − (x2 − x1)(e −2t − e −t ) − x2e −t<br />

u3 =0.<br />

Note that in the Lagrangian system the displacements are expressed in terms of the initial position and<br />

time, while in the Eulerian system the independent variables are the position coordinates and time. Euler<br />

equations describe, as a function of time, how such things as density, pressure, and fluid velocity change at<br />

a fixed point in the medium. In contrast, the Lagrangian viewpoint follows the time history of a moving<br />

individual fluid particle as it moves through the medium.


General Tensor Derivation of Strain.<br />

With reference to the figure 2.3-15 consider the deformation of a point P within a continuum. Let<br />

(y1 ,y2 ,y3 ) denote a Cartesian coordinate system which is fixed in space. We can introduce a coordinate<br />

transformation yi = yi (x1 ,x2 ,x3 ), i =1, 2, 3 and represent all points within the continuum with respect<br />

to a set of generalized coordinates (x1 ,x2 ,x3 ). Let P denote a general point in the continuum while it is<br />

in an unstrained state and assume that this point gets transformed to a point P ′ when the continuum<br />

experiences external forces. If P moves to P ′ , then all points Q which are near P will move to points Q ′<br />

near P ′ . We can imagine that in the unstrained state all the points of the continuum are referenced with<br />

respect to the set of generalized coordinates (x 1 ,x 2 ,x 3 ). After the strain occurs, we can imagine that it will<br />

be convenient to represent all points of the continuum with respect to a new barred system of coordinates<br />

(x 1 , x 2 , x 3 ). We call the original set of coordinates the Lagrangian system of coordinates and the new set<br />

of barred coordinates the Eulerian coordinates. The Eulerian coordinates are assumed to be described by<br />

a set of coordinate transformation equations xi = xi (x1 ,x2 ,x3 ), i =1, 2, 3 with inverse transformations<br />

xi = xi (x1 , x2 , x3 ), i =1, 2, 3, which are assumed to exist. The barred and unbarred coordinates can<br />

be related to a fixed set of Cartesian coordinates yi ,i = 1, 2, 3, and we may assume that there exists<br />

transformation equations<br />

y i = y i (x 1 ,x 2 ,x 3 ), i =1, 2, 3 and y i = y i (x 1 , x 2 , x 3 ), i =1, 2, 3<br />

which relate the barred and unbarred coordinates to the Cartesian axes. In the discussion that follows<br />

be sure to note whether there is a bar over a symbol, as we will be jumping back and forth between the<br />

Lagrangian and Eulerian reference frames.<br />

Figure 2.3-15. Strain in generalized coordinates<br />

In the Lagrangian system of unbarred coordinates we have the basis vectors Ei = ∂r<br />

which produce<br />

∂xi the metrices gij = Ei · Ej. Similarly, in the Eulerian system of barred coordinates we have the basis vectors<br />

<br />

Ei = ∂r<br />

∂xi which produces the metrices Gij = Ei · Ej. These basis vectors are illustrated in the figure 2.3-15.<br />

229


230<br />

We assume that an element of arc length squared ds2 in the unstrained state is deformed to the element<br />

of arc length squared ds2 in the strained state. An element of arc length squared can be expressed in terms<br />

of the barred or unbarred coordinates. For example, in the Lagrangian system, let dr = PQ so that<br />

L 2 0 = dr · dr = ds2 = gijdx i dx j , (2.3.54)<br />

where gij are the metrices in the Lagrangian coordinate system. This same element of arc length squared<br />

canbeexpressedinthebarredsystemby<br />

L 2 0 = ds 2 = gijdx i dx j ∂x<br />

, where gij = gmn<br />

m<br />

∂xi Similarly, in the Eulerian system of coordinates the deformed arc length squared is<br />

∂xn j . (2.3.55)<br />

∂x<br />

L 2 = dr · dr = ds 2 = Gijdx i dx j , (2.3.56)<br />

where Gij are the metrices in the Eulerian system of coordinates. This same element of arc length squared<br />

can be expressed in the Lagrangian system by the relation<br />

where<br />

In the Lagrangian system we have<br />

L 2 = ds 2 = Gijdx i dx j ∂x<br />

, where Gij = Gmn<br />

m<br />

∂xi ∂xn . (2.3.57)<br />

∂xj ds 2 − ds 2 =(Gij − gij)dx i dx j =2eijdx i dx j<br />

eij = 1<br />

2 (Gij − gij) (2.3.58)<br />

is called the Green strain tensor or Lagrangian strain tensor. Alternatively, in the Eulerian system of<br />

coordinates we may write<br />

ds 2 − ds 2 = i j<br />

Gij − gij dx dx =2eijdx i dx j<br />

where<br />

eij = 1 <br />

Gij − gij 2<br />

is called the Almansi strain tensor or Eulerian strain tensor.<br />

(2.3.59)


Note also in the figure 2.3-15 there is the displacement vector u. This vector can be represented in any<br />

of the following forms:<br />

u = u i Ei contravariant, Lagrangian basis<br />

u = ui E i<br />

covariant, Lagrangian reciprocal basis<br />

u = u i Ei contravariant, Eulerian basis<br />

<br />

u = uiE<br />

i<br />

covariant, Eulerian reciprocal basis.<br />

By vector addition we have r + u = r and consequently dr + du = dr. In the Lagrangian frame of reference<br />

at the point P we represent u in the contravariant form u = uiEi and write dr in the form dr = dxiEi. By<br />

use of the equation (1.4.48) we can express du in the form du = ui ,kdxkEi. These substitutions produce the<br />

representation dr =(dx i + u i ,kdx k ) Ei in the Lagrangian coordinate system. We can then express ds2 in the<br />

Lagrangian system. We find<br />

dr · dr = ds 2 =(dx i + u i ,k dxk ) Ei · (dx j + u j ,m dxm ) Ej<br />

=(dx i dx j + u j ,mdx m dx i + u i ,kdx k dx j + u i ,ku j ,mdx k dx m )gij<br />

and consequently from the relation (2.3.58) we derive the representation<br />

eij = 1 <br />

ui,j + uj,i + um,iu<br />

2<br />

m <br />

,j . (2.3.60)<br />

This is the representation of the Lagrangian strain tensor in any system of coordinates. The strain tensor<br />

eij is symmetric. We will restrict our study to small deformations and neglect the product terms in equation<br />

(2.3.60). Under these conditions the equation (2.3.60) reduces to eij = 1<br />

2 (ui,j + uj,i).<br />

If instead, we chose to represent the displacement u with respect to the Eulerian basis, then we can<br />

write<br />

These relations imply that<br />

u = u i Ei with du = u i ,k dxk Ei.<br />

dr = dr − du =(dx i − u i ,k dxk ) Ei.<br />

This representation of dr in the Eulerian frame of reference can be used to calculate the strain eij from the<br />

relation ds 2 − ds 2 . It is left as an exercise to show that there results<br />

eij = 1 <br />

ui,j + uj,i − um,iu<br />

2<br />

m <br />

,j . (2.3.61)<br />

The equation (2.3.61) is the representation of the Eulerian strain tensor in any system of coordinates. Under<br />

conditions of small deformations both the equations (2.3.60) and (2.3.61) reduce to the linearized Lagrangian<br />

and Eulerian strain tensor eij = 1<br />

2 (ui,j + uj,i). In the case of large deformations the equations (2.3.60) and<br />

(2.3.61) describe the strains. In the case of linear elasticity, where the deformations are very small, the<br />

product terms in equations (2.3.60) and (2.3.61) are neglected and the Lagrangian and Eulerian strains<br />

reduce to their linearized forms<br />

eij = 1<br />

2 [ui,j + uj,i] eij = 1<br />

2 [ui,j + uj,i] . (2.3.62)<br />

231


232<br />

Figure 2.3-16. Displacement due to strain<br />

Compressible and Incompressible Material With reference to figure 2.3-16, let xi , i =1, 2, 3denote<br />

the position vector of an arbitrary point P in a continuum before there is a state of strain. Let Q be<br />

a neighboring point of P with position vector xi + dxi , i =1, 2, 3. Also in the figure 2.3-16 there is the<br />

displacement vector u. Here it is assumed that u = u(x 1 ,x2 ,x3 ) denotes the displacement field when the<br />

continuum is in a state of strain. The figure 2.3-16 illustrates that in a state of strain P moves to P ′ and Q<br />

moves to Q ′ . Let us find a relationship between the distance PQbefore the strain and the distance P ′ Q ′ when<br />

the continuum is in a state of strain. For E1, E2, E3 basis functions constructed at P we have previously<br />

shown that if<br />

u(x 1 ,x 2 ,x 3 )=u i Ei then du = u i ,jdx j Ei. <br />

Now for u + du the displacement of the point Q we may use vector addition and write<br />

PQ+ u + du = u + P ′ Q ′ . (2.3.63)<br />

Let PQ = dxiEi = aiEi denote an arbitrary small change in the continuum. This arbitrary displacement<br />

gets deformed to P ′ Q ′ = A i Ei due to the state of strain in the continuum. Employing the equation (2.3.63)<br />

we write<br />

which can be written in the form<br />

dx i + u i ,jdx j = a i + u i ,ja j = A i<br />

δa i = A i − a i = u i ,j aj where dx i = a i ,i=1, 2, 3 (2.3.64)<br />

denotes an arbitrary small change. The tensor ui ,j and the associated tensor ui,j = gitut ,j are in general<br />

not symmetric tensors. However, we know we can express ui,j as the sum of a symmetric (eij) andskew-<br />

symmetric(ωij) tensor. We therefore write<br />

where<br />

ui,j = eij + ωij or u i ,j = ei j + ωi j ,<br />

eij = 1<br />

2 (ui,j + uj,i) = 1<br />

2 (gimu m ,j + gjmu m ,i) and ωij = 1<br />

2 (ui,j − uj,i) = 1<br />

2 (gimu m ,j − gjmu m ,i) .<br />

The deformation of a small quantity ai can therefore be represented by a pure strain Ai − ai = ei sas followed<br />

by a rotation Ai − ai = ωi sas .


Consider now a small element of volume inside a material medium. With reference to the figure 2.3-<br />

17(a) we let a, b,c denote three small arbitrary independent vectors constructed at a general point P within<br />

the material before any external forces are applied. We imagine a, b,c as representing the sides of a small<br />

parallelepiped before any deformation has occurred. When the material is placed in a state of strain the<br />

point P will move to P ′ and the vectors a, b,c will become deformed to the vectors A, B, C as illustrated in<br />

the figure 2.3-17(b). The vectors A, B, C represent the sides of the parallelepiped after the deformation.<br />

Figure 2.3-17. Deformation of a parallelepiped<br />

Let ∆V denote the volume of the parallelepiped with sides a, b,c at P before the strain and let ∆V ′<br />

denote the volume of the deformed parallelepiped after the strain, when it then has sides A, B, C at the<br />

point P ′ . We define the ratio of the change in volume due to the strain divided by the original volume as<br />

the dilatation at the point P. The dilatation is thus expressed as<br />

Θ= ∆V ′ − ∆V<br />

∆V<br />

= dilatation. (2.3.65)<br />

Since ui ,i =1, 2, 3 represents the displacement field due to the strain, we use the result from equation<br />

(2.3.64) and represent the displaced vectors A, B, C in the form<br />

A i = a i + u i ,j aj<br />

B i = b i + u i ,j bj<br />

C i = c i + u i ,jc j<br />

(2.3.66)<br />

where a, b,c are arbitrary small vectors emanating from the point P in the unstrained state. The element of<br />

volume ∆V, before the strain, is calculated from the triple scalar product relation<br />

∆V = a · ( b × c) =eijka i b j c k .<br />

The element of volume ∆V ′ , which occurs due to the strain, is calculated from the triple scalar product<br />

∆V ′ = A · ( B × C)=eijkA i B j C k .<br />

233


234<br />

Substituting the relations from the equations (2.3.66) into the triple scalar product gives<br />

∆V ′ = eijk(a i + u i ,m am )(b j + u j ,n bn )(c k + u k ,p cp ).<br />

Expanding the triple scalar product and employing the result from Exercise 1.4, problem 34, we find the<br />

simplified result gives us the dilatation<br />

Θ= ∆V ′ − ∆V<br />

∆V<br />

= u r ,r<br />

=div(u). (2.3.67)<br />

That is, the dilatation is the divergence of the displacement field. If the divergence of the displacement field<br />

is zero, there is no volume change and the material is said to be incompressible. If the divergence of the<br />

displacement field is different from zero, the material is said to be compressible.<br />

Note that the strain eij is expressible in terms of the displacement field by the relation<br />

eij = 1<br />

2 (ui,j + uj,i), and consequently g mn emn = u r ,r . (2.3.68)<br />

Hence, for an orthogonal system of coordinates the dilatation can be expressed in terms of the strain elements<br />

along the main diagonal.<br />

Conservation of Mass<br />

Consider the material in an arbitrary region R of a continuum. Let ϱ = ϱ(x, y, z, t) denote the density<br />

of the material within the region. Assume that the dimension of the density ϱ is gm/cm3 in the cgs system<br />

of units. We shall assume that the region R is bounded by a closed surface S with exterior unit normal n<br />

defined everywhere on the surface. Further, we let v = v(x, y, z, t) denote a velocity field associated with all<br />

points within the continuum. The velocity field has units of cm/sec in the cgs system of units. Neglecting<br />

sources and sinks, the law of conservation of mass examines all the material entering and leaving a region R.<br />

Enclosed within R is the material mass m where m = ϱdτ with dimensions of gm in the cgs system of<br />

R<br />

units. Here dτ denotes an element of volume inside the region R. The change of mass with time is obtained<br />

by differentiating the above relation. Differentiating the mass produces the equation<br />

∂m<br />

∂t =<br />

<br />

∂ϱ<br />

dτ (2.3.69)<br />

∂t<br />

and has the dimensions of gm/sec.<br />

Consider also the surface integral<br />

<br />

I =<br />

S<br />

R<br />

ϱv · ˆndσ (2.3.70)<br />

where dσ is an element of surface area on the surface S which encloses R and ˆn is the exterior unit normal<br />

vector to the surface S. The dimensions of the integral I is determined by examining the dimensions of each<br />

term in the integrand of I. We find that<br />

[I] = gm cm<br />

·<br />

cm3 sec · cm2 = gm<br />

sec<br />

and so the dimension of I is the same as the dimensions for the change of mass within the region R. The<br />

surface integral I is the flux rate of material crossing the surface of R and represents the change of mass


entering the region if v · ˆn is negative and the change of mass leaving the region if v · ˆn is positive, as ˆn is<br />

always an exterior unit normal vector. Equating the relations from equations (2.3.69) and (2.3.70) we obtain<br />

a mathematical statement for mass conservation<br />

∂m<br />

∂t =<br />

<br />

<br />

∂ϱ<br />

dτ = − ϱv · ndσ. (2.3.71)<br />

R ∂t S<br />

The equation (2.3.71) implies that the rate at which the mass contained in R increases must equal the rate<br />

at which the mass flows into R through the surface S. The negative sign changes the direction of the exterior<br />

normal so that we consider flow of material into the region. Employing the Gauss divergence theorem, the<br />

surface integral in equation (2.3.71) can be replaced by a volume integral and the law of conservation of<br />

mass is then expressible in the form<br />

<br />

∂ϱ<br />

∂t +div(ϱv)<br />

<br />

dτ =0. (2.3.72)<br />

R<br />

Since the region R is an arbitrary volume we conclude that the term inside the brackets must equal zero.<br />

This gives us the continuity equation<br />

∂ϱ<br />

+div(ϱv) =0 (2.3.73)<br />

∂t<br />

which represents the mass conservation law in terms of velocity components. This is the Eulerian representation<br />

of continuity of mass flow.<br />

Equivalent forms of the continuity equation are:<br />

∂ϱ<br />

+ v · grad ϱ + ϱ div v =0<br />

∂t<br />

∂ϱ<br />

∂t<br />

∂ϱ ∂vi<br />

+ vi + ϱ =0<br />

∂xi ∂xi Dϱ ∂vi<br />

+ ϱ =0<br />

Dt ∂xi where Dϱ ∂ϱ ∂ϱ<br />

= +<br />

Dt ∂t ∂xi dxi ∂ϱ ∂ϱ<br />

= +<br />

dt ∂t ∂xi vi is called the material derivative of the density ϱ. Note that the<br />

material derivative contains the expression ∂ϱ<br />

∂xi vi which is known as the convective or advection term. If the<br />

density ϱ = ϱ(x, y, z, t) is a constant we have<br />

Dϱ<br />

Dt<br />

∂ϱ ∂ϱ dx<br />

= +<br />

∂t ∂x dt<br />

∂ϱ dy<br />

+<br />

∂y dt<br />

∂ϱ dz ∂ϱ ∂ϱ<br />

+ = +<br />

∂z dt ∂t ∂xi dxi dt<br />

=0 (2.3.74)<br />

and hence the continuity equation reduces to div (v) =0. Thus, if div (v) is zero, then the material is<br />

incompressible.<br />

EXAMPLE 2.3-2. (Continuity Equation) Find the Lagrangian representation of mass conservation.<br />

Solution: Let (X, Y, Z) denote the initial position of a fluid particle and denote the density of the fluid by<br />

ϱ(X, Y, Z, t) sothatϱ(X, Y, Z, 0) denotes the density at the time t =0. Consider a simple closed region in<br />

our continuum and denote this region by R(0) at time t =0andbyR(t) atsomelatertimet. That is, all<br />

the points in R(0) move in a one-to-one fashion to points in R(t). Initially the mass of material in R(0) is<br />

m(0) = ϱ(X, Y, Z, 0) dτ(0) where dτ(0) = dXdY dZ is an element of volume in R(0). We have after a<br />

R(0)<br />

235


236<br />

<br />

time t has elapsed the mass of material in the region R(t) givenbym(t) = ϱ(X, Y, Z, t) dτ(t) where<br />

R(t) <br />

x,y,z<br />

dτ(t) =dxdydz is a deformed element of volume related to the dτ(0) by dτ(t) =J X,Y,Z dτ(0) where J is<br />

the Jacobian of the Eulerian (x, y, z) variables with respect to the Lagrangian (X, Y, Z) representation.For<br />

mass conservation we require that m(t) =m(0) for all t. This implies that<br />

ϱ(X, Y, Z, t)J = ϱ(X, Y, Z, 0) (2.3.75)<br />

for all time, since the initial region R(0) is arbitrary. The right hand side of equation (2.3.75) is independent<br />

of time and so<br />

d<br />

(ϱ(X, Y, Z, t)J) =0. (2.3.76)<br />

dt<br />

This is the Lagrangian form of the continuity equation which expresses mass conservation. Using the result<br />

that dJ<br />

dt = Jdiv V,(see problem 28, Exercise 2.3), the equation (2.3.76) can be expanded and written in the<br />

form<br />

Dϱ<br />

Dt + ϱ div V =0 (2.3.77)<br />

where Dϱ<br />

Dt is from equation (2.3.74). The form of the continuity equation (2.3.77) is one of the Eulerian forms<br />

previously developed.<br />

In the Eulerian coordinates the continuity equation is written ∂ϱ<br />

∂t +div(ϱv) = 0, while in the Lagrangian<br />

system the continuity equation is written d(ϱJ)<br />

dt =0. Note that the velocity carries the Lagrangian axes and<br />

the density change grad ϱ. This is reflective of the advection term v · grad ϱ. Thus, in order for mass to<br />

be conserved it need not remain stationary. The mass can flow and the density can change. The material<br />

derivative is a transport rule depicting the relation between the Eulerian and Lagrangian viewpoints.<br />

In general, from a Lagrangian viewpoint, any quantity Q(x, y, z, t) which is a function of both position<br />

and time is seen as being transported by the fluid velocity (v1,v2,v3) toQ(x + v1dt, y + v2dt, z + v3dt, t + dt).<br />

Then the time derivative of Q contains both ∂Q<br />

∂t and the advection term v ·∇Q. In terms of mass flow, the<br />

Eulerian viewpoint sees flow into and out of a fixed volume in space, as depicted by the equation (2.3.71),<br />

In contrast, the Lagrangian viewpoint sees the same volume moving with the fluid and consequently<br />

<br />

D<br />

ρdτ =0,<br />

Dt R(t)<br />

where R(t) represents the volume moving with the fluid. Both viewpoints produce the same continuity<br />

equation reflecting the conservation of mass.<br />

Summary of Basic Equations<br />

Let us summarize the basic equations which are valid for all types of a continuum. We have derived:<br />

• Conservation of mass (continuity equation)<br />

∂ϱ<br />

∂t +(ϱvi ),i =0


• Conservation of linear momentum sometimes called the Cauchy equation of motion.<br />

• Conservation of angular momentum<br />

• Strain tensor for linear elasticity<br />

σ ij ,i + ϱb j = ϱf j , j =1, 2, 3.<br />

σij = σji<br />

eij = 1<br />

2 (ui,j + uj,i).<br />

If we assume that the continuum is in equilibrium, and there is no motion, then the velocity and<br />

acceleration terms above will be zero. The continuity equation then implies that the density is a constant.<br />

The conservation of angular momentum equation requires that the stress tensor be symmetric and we need<br />

find only six stresses. The remaining equations reduce to a set of nine equations in the fifteen unknowns:<br />

3 displacements u1,u2,u3<br />

6strains e11,e<strong>12</strong>,e13,e22,e23,e33<br />

6 stresses σ11,σ<strong>12</strong>,σ13,σ22,σ23,σ33<br />

Consequently, we still need additional information if we desire to determine these unknowns.<br />

Note that the above equations do not involve any equations describing the material properties of the<br />

continuum. We would expect solid materials to act differently from liquid material when subjected to external<br />

forces. An equation or equations which describe the material properties are called constitutive equations.<br />

In the following sections we will investigate constitutive equations for solids and liquids. We will restrict<br />

our study to linear elastic materials over a range where there is a linear relationship between the stress and<br />

strain. We will not consider plastic or viscoelastic materials. Viscoelastic materials have the property that<br />

the stress is not only a function of strain but also a function of the rates of change of the stresses and strains<br />

and consequently properties of these materials are time dependent.<br />

237


238<br />

EXERCISE 2.3<br />

◮ 1. Assume an orthogonal coordinate system with metric tensor gij =0fori= j and g (i)(i) = h2 i (no<br />

summation on i). Use the definition of strain<br />

and show that in terms of the physical components<br />

there results the equations:<br />

<br />

t ∂u t<br />

eii = git + u<br />

∂xi mi<br />

m<br />

<br />

∂u<br />

2eij = git<br />

t ∂u<br />

+ gjt<br />

∂xj t<br />

, i = j<br />

∂xi e(ii) = ∂<br />

∂x i<br />

2e(ij) = hi<br />

hj<br />

u(i)<br />

∂<br />

∂x j<br />

hi<br />

ers = 1<br />

2 (ur,s + us,r) = 1 <br />

grtu<br />

2<br />

t ,s + gstu t <br />

,r<br />

e(ij) = eij<br />

hihj<br />

no summation on i or j<br />

u(i) =hiu i no summation on i<br />

u(i)<br />

<br />

+ 1<br />

hi<br />

no summation on i<br />

2h2 3 u(m) ∂<br />

i<br />

hm ∂x<br />

m=1<br />

m<br />

2<br />

hi no summation on i<br />

<br />

+ hj ∂<br />

hi ∂xi <br />

u(j)<br />

, no summation on i or j, i = j.<br />

hj<br />

◮ 2. Use the results from problem 1 to write out all components of the strain tensor in Cartesian coordinates.<br />

Use the notation u(1) = u,u(2) = v,u(3) = w and<br />

to verify the relations:<br />

e(11) = exx, e(22) = eyy, e(33) = ezz, e(<strong>12</strong>) = exy, e(13) = exz, e(23) = eyz<br />

exx = ∂u<br />

∂x<br />

eyy = ∂v<br />

∂y<br />

ezz = ∂w<br />

∂z<br />

exy = 1<br />

<br />

∂v ∂u<br />

+<br />

2 ∂x ∂y<br />

<br />

∂u ∂w<br />

+<br />

∂z ∂x<br />

ezy = 1<br />

<br />

∂w ∂v<br />

+<br />

2 ∂y ∂z<br />

exz = 1<br />

2<br />

◮ 3. Use the results from problem 1 to write out all components of the strain tensor in cylindrical coordinates.<br />

Use the notation u(1) = ur , u(2) = uθ, u(3) = uz and<br />

to verify the relations:<br />

e(11) = err, e(22) = eθθ, e(33) = ezz, e(<strong>12</strong>) = erθ, e(13) = erz, e(23) = eθz<br />

err = ∂ur<br />

∂r<br />

eθθ = 1 ∂uθ<br />

r ∂θ<br />

ezz = ∂uz<br />

∂z<br />

+ ur<br />

r<br />

erθ = 1<br />

<br />

1 ∂ur ∂uθ uθ<br />

+ −<br />

2 r ∂θ ∂r r<br />

erz = 1<br />

<br />

∂uz ∂ur<br />

+<br />

2 ∂r ∂z<br />

eθz = 1<br />

<br />

∂uθ 1 ∂uz<br />

+<br />

2 ∂z r ∂θ


◮ 4. Use the results from problem 1 to write out all components of the strain tensor in spherical coordinates.<br />

Use the notation u(1) = uρ,u(2) = uθ,u(3) = uφ and<br />

to verify the relations<br />

e(11) = eρρ, e(22) = eθθ, e(33) = eφφ, e(<strong>12</strong>) = eρθ, e(13) = eρφ, e(23) = eθφ<br />

eρρ = ∂uρ<br />

∂ρ<br />

eθθ = 1 ∂uθ uρ<br />

+<br />

ρ ∂θ ρ<br />

eφφ = 1 ∂uφ uρ uθ<br />

+ + cot θ<br />

ρ sin θ ∂φ ρ ρ<br />

eρθ = 1<br />

<br />

1 ∂uρ uθ ∂uθ<br />

− +<br />

2 ρ ∂θ ρ ∂ρ<br />

eρφ = 1<br />

<br />

1 ∂uρ uφ ∂uφ<br />

− +<br />

2 ρ sin θ ∂φ ρ ∂ρ<br />

eθφ = 1<br />

<br />

1 ∂uφ uφ 1 ∂uθ<br />

− cot θ +<br />

2 ρ ∂θ ρ ρ sin θ ∂φ<br />

◮ 5. Expand equation (2.3.67) and find the dilatation in terms of the physical components of an orthogonal<br />

system and verify that<br />

Θ=<br />

1<br />

<br />

∂(h2h3u(1))<br />

∂x1 + ∂(h1h3u(2))<br />

∂x2 + ∂(h1h2u(3))<br />

∂x3 <br />

h1h2h3<br />

◮ 6. Verify that the dilatation in Cartesian coordinates is<br />

Θ=exx + eyy + ezz = ∂u ∂v ∂w<br />

+ +<br />

∂x ∂y ∂z .<br />

◮ 7. Verify that the dilatation in cylindrical coordinates is<br />

Θ=err + eθθ + ezz = ∂ur<br />

∂r<br />

◮ 8. Verify that the dilatation in spherical coordinates is<br />

Θ=eρρ + eθθ + eφφ = ∂uρ<br />

∂ρ<br />

1 ∂uθ 1<br />

+ +<br />

r ∂θ r ur + ∂uz<br />

∂z .<br />

1 ∂uθ 2<br />

+ +<br />

ρ ∂θ ρ uρ + 1 ∂uφ<br />

ρ sin θ ∂φ + uθ cot θ<br />

ρ<br />

◮ 9. Show that in an orthogonal set of coordinates the rotation tensor ωij canbewrittenintermsofphysical<br />

components in the form<br />

Hint: See problem 1.<br />

ω(ij) = 1<br />

<br />

∂(hiu(i))<br />

2hihj ∂xj − ∂(hju(j))<br />

∂xi <br />

, no summations<br />

◮ 10. Use the result from problem 9 to verify that in Cartesian coordinates<br />

ωyx = 1<br />

<br />

∂v ∂u<br />

−<br />

2 ∂x ∂y<br />

ωxz = 1<br />

<br />

∂u ∂w<br />

−<br />

2 ∂z ∂x<br />

ωzy = 1<br />

<br />

∂w ∂v<br />

−<br />

2 ∂y ∂z<br />

.<br />

239


240<br />

◮ 11. Use the results from problem 9 to verify that in cylindrical coordinates<br />

ωθr = 1<br />

<br />

∂(ruθ) ∂ur<br />

−<br />

2r ∂r ∂θ<br />

ωrz = 1<br />

<br />

∂ur ∂uz<br />

−<br />

2 ∂z ∂r<br />

ωzθ = 1<br />

<br />

1 ∂uz ∂uθ<br />

−<br />

2 r ∂θ ∂z<br />

◮ <strong>12</strong>. Use the results from problem 9 to verify that in spherical coordinates<br />

ωθρ = 1<br />

<br />

∂(ρuθ) ∂uρ<br />

−<br />

2ρ ∂ρ ∂θ<br />

ωρφ = 1<br />

<br />

1 ∂uρ ∂(ρuφ)<br />

−<br />

2ρ sin θ ∂φ ∂ρ<br />

<br />

1 ∂(uφ sin θ)<br />

ωφθ =<br />

−<br />

2ρ sin θ ∂θ<br />

∂uθ<br />

<br />

∂φ<br />

◮ 13. The conditions for static equilibrium in a linear elastic material are determined from the conservation<br />

law<br />

σ j<br />

i,j + ϱbi =0, i,j =1, 2, 3,<br />

where σ i j are the stress tensor components, bi are the external body forces per unit mass and ϱ is the density<br />

of the material. Assume an orthogonal coordinate system and verify the following results.<br />

(a) Show that<br />

∂<br />

) − [ij, m]σmj<br />

(b) Use the substitutions<br />

σ j 1<br />

i,j = √<br />

g<br />

σ(ij) =σ j hj<br />

i<br />

hi<br />

b(i) = bi<br />

hi<br />

∂xj (√gσ j<br />

i<br />

no summation on i or j<br />

no summation on i<br />

σ(ij) =σ ij hihj no summation on i or j<br />

and express the equilibrium equations in terms of physical components and verify the relations<br />

3<br />

j=1<br />

1 ∂<br />

√<br />

g ∂xj where there is no summation on i.<br />

√ ghiσ(ij)<br />

hj<br />

<br />

− 1<br />

2<br />

3<br />

j=1<br />

σ(jj)<br />

h2 ∂(h<br />

j<br />

2 j )<br />

∂xi + hiϱb(i) =0,<br />

◮ 14. Use the results from problem 13 and verify that the equilibrium equations in Cartesian coordinates<br />

can be expressed<br />

∂σxx<br />

∂x<br />

∂σyx<br />

∂x<br />

∂σzx<br />

∂x<br />

+ ∂σxy<br />

∂y<br />

+ ∂σyy<br />

∂y<br />

+ ∂σzy<br />

∂y<br />

+ ∂σxz<br />

∂z + ϱbx =0<br />

+ ∂σyz<br />

∂z + ϱby =0<br />

+ ∂σzz<br />

∂z + ϱbz =0


◮ 15. Use the results from problem 13 and verify that the equilibrium equations in cylindrical coordinates<br />

can be expressed<br />

∂σrr 1 ∂σrθ ∂σrz 1<br />

+ + +<br />

∂r r ∂θ ∂z r (σrr − σθθ)+ϱbr =0<br />

∂σθr 1 ∂σθθ ∂σθz 2<br />

+ + +<br />

∂r r ∂θ ∂z r σθr + ϱbθ =0<br />

∂σzr 1 ∂σzθ ∂σzz 1<br />

+ + +<br />

∂r r ∂θ ∂z r σzr + ϱbz =0<br />

◮ 16. Use the results from problem 13 and verify that the equilibrium equations in spherical coordinates<br />

can be expressed<br />

∂σρρ<br />

∂ρ<br />

1 ∂σρθ 1 ∂σρφ 1<br />

+ + +<br />

ρ ∂θ ρ sin θ ∂φ ρ (2σρρ − σθθ − σφφ + σρθ cot θ)+ϱbρ =0<br />

∂σθρ<br />

∂ρ<br />

1 ∂σθθ 1 ∂σθφ 1<br />

+ + +<br />

ρ ∂θ ρ sin θ ∂φ ρ (3σρθ +[σθθ − σφφ]cotθ)+ϱbθ =0<br />

∂σφρ<br />

∂ρ<br />

1 ∂σφθ 1 ∂σφφ 1<br />

+ + +<br />

ρ ∂θ ρ sin θ ∂φ ρ (3σρφ +2σθφ cot θ)+ϱbφ =0<br />

◮ 17. Derive the result for the Lagrangian strain defined by the equation (2.3.60).<br />

◮ 18. Derive the result for the Eulerian strain defined by equation (2.3.61).<br />

◮ 19. The equation δa i = u i ,j aj , describes the deformation in an elastic solid subjected to forces. The<br />

quantity δai denotes the difference vector Ai − ai between the undeformed and deformed states.<br />

(a) Let |a| denote the magnitude of the vector ai and show that the strain e in the direction ai can be<br />

represented<br />

e = δ|a|<br />

|a|<br />

= eij<br />

<br />

i j a a<br />

|a|<br />

|a|<br />

= eijλ i λ j ,<br />

where λi is a unit vector in the direction ai .<br />

(b) Show that for λ1 =1,λ2 =0,λ3 = 0 there results e = e11, with similar results applying to vectors λi in<br />

the y and z directions.<br />

Hint: Consider the magnitude squared |a| 2 = gija i a j .<br />

◮ 20. At the point (1, 2, 3) of an elastic solid construct the small vector a = ɛ( 2<br />

3 ê1 + 2<br />

3 ê2 + 1<br />

3 ê3), where<br />

ɛ>0 is a small positive quantity. The solid is subjected to forces such that the following displacement field<br />

results.<br />

u =(xy ê1 + yz ê2 + xz ê3) × 10 −2<br />

Calculate the deformed vector A after the displacement field has been imposed.<br />

◮ 21. For the displacement field<br />

u =(x 2 + yz) ê1 +(xy + z 2 ) ê2 + xyz ê3<br />

(a) Calculate the strain matrix at the point (1, 2, 3).<br />

(b) Calculate the rotation matrix at the point (1, 2, 3).<br />

241


242<br />

◮ 22. Show that for an orthogonal coordinate system the ith component of the convective operator can be<br />

written<br />

[( V ·∇) 3<br />

3<br />

<br />

V (m) ∂A(i) A(m)<br />

A]i =<br />

+<br />

V (i)<br />

∂xm ∂hi<br />

− V (m)∂hm<br />

∂xm ∂xi <br />

m=1<br />

hm<br />

m=1<br />

m=i<br />

◮ 23. Consider a parallelepiped with dimensions ℓ, w, h which has a uniform pressure P applied to each<br />

face. Show that the volume strain can be expressed as<br />

∆V<br />

V<br />

= ∆ℓ<br />

ℓ<br />

+ ∆w<br />

w<br />

+ ∆h<br />

h<br />

hmhi<br />

−3P (1 − 2ν)<br />

= .<br />

E<br />

The quantity k = E/3(1 − 2ν) is called the bulk modulus of elasticity.<br />

◮ 24. Show in Cartesian coordinates the continuity equation is<br />

where (u, v, w) arethevelocitycomponents.<br />

∂ϱ ∂(ϱu) ∂(ϱv) ∂(ϱw)<br />

+ + +<br />

∂t ∂x ∂y ∂z =0,<br />

◮ 25. Show in cylindrical coordinates the continuity equation is<br />

∂ϱ<br />

∂t<br />

1 ∂(rϱVr)<br />

+<br />

r ∂r<br />

where Vr,Vθ,Vz are the velocity components.<br />

1<br />

+<br />

r<br />

∂(ϱVθ)<br />

∂θ<br />

◮ 26. Show in spherical coordinates the continuity equation is<br />

where Vρ,Vθ,Vφ are the velocity components.<br />

+ ∂(ϱVz)<br />

∂z =0<br />

∂ϱ 1<br />

+<br />

∂t ρ2 ∂(ρ2ϱVρ) +<br />

∂ρ<br />

1 ∂(ϱVθ sin θ)<br />

+<br />

ρ sin θ ∂θ<br />

1 ∂(ϱVφ)<br />

ρ sin θ ∂φ =0<br />

◮ 27. (a) Apply a stress σyy to both ends of a square element in a x, y continuum. Illustrate and label<br />

all changes that occur due to this stress. (b) Apply a stress σxx to both ends of a square element in a<br />

x, y continuum. Illustrate and label all changes that occur due to this stress. (c) Use superposition of your<br />

results in parts (a) and (b) and explain each term in the relations<br />

exx = σxx<br />

E<br />

− ν σyy<br />

E<br />

◮ 28. Show that the time derivative of the Jacobian J = J<br />

div V = ∂V1<br />

∂x<br />

+ ∂V2<br />

∂y<br />

+ ∂V3<br />

∂z<br />

and eyy = σyy<br />

E<br />

− ν σxx<br />

E .<br />

<br />

x, y, z<br />

satisfies<br />

X, Y, Z<br />

dJ<br />

dt = J div V where<br />

and V1 = dx<br />

dt , V2 = dy<br />

dt , V3 = dz<br />

dt .<br />

Hint: Let (x, y, z) =(x1,x2,x3) and(X, Y, Z) =(X1,X2,X3), then note that<br />

eijk<br />

∂V1 ∂x2 ∂x3 ∂V1 ∂xm ∂x2 ∂x3 ∂x1 ∂x2 ∂x3 ∂V1<br />

= eijk<br />

= eijk<br />

, etc.<br />

∂Xi ∂Xj ∂Xk ∂xm ∂Xi ∂Xj ∂Xk ∂Xi ∂Xj ∂Xk ∂x1


§2.4 CONTINUUM MECHANICS (SOLIDS)<br />

In this introduction to continuum mechanics we consider the basic equations describing the physical<br />

effects created by external forces acting upon solids and fluids. In addition to the basic equations that<br />

are applicable to all continua, there are equations which are constructed to take into account material<br />

characteristics. These equations are called constitutive equations. For example, in the study of solids the<br />

constitutive equations for a linear elastic material is a set of relations between stress and strain. In the study<br />

of fluids, the constitutive equations consists of a set of relations between stress and rate of strain. Constitutive<br />

equations are usually constructed from some basic axioms. The resulting equations have unknown material<br />

parameters which can be determined from experimental investigations.<br />

One of the basic axioms, used in the study of elastic solids, is that of material invariance. This axiom<br />

requires that certain symmetry conditions of solids are to remain invariant under a set of orthogonal<br />

transformations and translations. This axiom is employed in the next section to simplify the constitutive<br />

equations for elasticity. We begin our study of continuum mechanics by investigating the development of<br />

constitutive equations for linear elastic solids.<br />

Generalized Hooke’s Law<br />

If the continuum material is a linear elastic material, we introduce the generalized Hooke’s law in<br />

Cartesian coordinates<br />

σij = cijklekl, i,j,k,l=1, 2, 3. (2.4.1)<br />

The Hooke’s law is a statement that the stress is proportional to the gradient of the deformation occurring<br />

in the material. These equations assume a linear relationship exists between the components of the stress<br />

tensor and strain tensor and we say stress is a linear function of strain. Such relations are referred to as a<br />

set of constitutive equations. Constitutive equations serve to describe the material properties of the medium<br />

when it is subjected to external forces.<br />

Constitutive Equations<br />

The equations (2.4.1) are constitutive equations which are applicable for materials exhibiting small<br />

deformations when subjected to external forces. The 81 constants cijkl are called the elastic stiffness of the<br />

material. The above relations can also be expressed in the form<br />

eij = sijklσkl, i,j,k,l=1, 2, 3 (2.4.2)<br />

where sijkl are constants called the elastic compliance of the material. Since the stress σij and strain eij<br />

have been shown to be tensors we can conclude that both the elastic stiffness cijkl and elastic compliance<br />

sijkl are fourth order tensors. Due to the symmetry of the stress and strain tensors we find that the elastic<br />

stiffness and elastic compliance tensor must satisfy the relations<br />

cijkl = cjikl = cijlk = cjilk<br />

sijkl = sjikl = sijlk = sjilk<br />

(2.4.3)<br />

and consequently only 36 of the 81 constants are actually independent. If all 36 of the material (crystal)<br />

constants are independent the material is called triclinic and there are no material symmetries.<br />

243


244<br />

Restrictions on Elastic Constants due to Symmetry<br />

The equations (2.4.1) and (2.4.2) can be replaced by an equivalent set of equations which are easier to<br />

analyze. This is accomplished by defining the quantities<br />

where ⎛<br />

and ⎛<br />

e1, e2, e3, e4, e5, e6<br />

σ1, σ2, σ3, σ4, σ5, σ6<br />

⎝ e1 e4 e5<br />

e4 e2 e6<br />

e5 e6 e3<br />

⎝ σ1 σ4 σ5<br />

σ4 σ2 σ6<br />

σ5 σ6 σ3<br />

⎞ ⎛<br />

⎠ =<br />

⎞ ⎛<br />

⎠ =<br />

⎝ e11 e<strong>12</strong> e13<br />

e21 e22 e23<br />

e31 e32 e33<br />

⎝ σ11 σ<strong>12</strong> σ13<br />

σ21 σ22 σ23<br />

σ31 σ32 σ33<br />

Then the generalized Hooke’s law from the equations (2.4.1) and (2.4.2) can be represented in either of<br />

the forms<br />

σi = cijej or ei = sijσj where i, j =1,...,6 (2.4.4)<br />

where cij are constants related to the elastic stiffness and sij are constants related to the elastic compliance.<br />

These constants satisfy the relation<br />

Here<br />

and similarly<br />

relations<br />

where<br />

⎞<br />

⎠<br />

⎞<br />

⎠ .<br />

smicij = δmj where i, m, j =1,...,6 (2.4.5)<br />

<br />

ei, i = j =1, 2, 3<br />

eij =<br />

e1+i+j, i = j, and i =1, or, 2<br />

<br />

σi, i = j =1, 2, 3<br />

σij =<br />

σ1+i+j, i = j, and i =1, or, 2.<br />

These relations show that the constants cij are related to the elastic stiffness coefficients cpqrs by the<br />

cm1 = cij11<br />

cm2 = cij22<br />

cm3 = cij33<br />

cm4 =2cij<strong>12</strong><br />

cm5 =2cij13<br />

cm6 =2cij23<br />

<br />

i, if i = j =1, 2, or 3<br />

m =<br />

1+i + j, if i = j and i =1or2.<br />

A similar type relation holds for the constants sij and spqrs. The above relations can be verified by expanding<br />

the equations (2.4.1) and (2.4.2) and comparing like terms with the expanded form of the equation (2.4.4).


The generalized Hooke’s law can now be expressed in a form where the 36 independent constants can<br />

be examined in more detail under special material symmetries. We will examine the form<br />

⎛ ⎞<br />

e1<br />

⎛<br />

s11 s<strong>12</strong> s13 s14 s15<br />

⎞ ⎛ ⎞<br />

s16 σ1<br />

⎜ e2 ⎟ ⎜ s21<br />

⎜ ⎟ ⎜<br />

⎜ e3 ⎟ ⎜ s31<br />

⎜ ⎟ = ⎜<br />

⎜ e4 ⎟ ⎜ s41<br />

⎝ ⎠ ⎝<br />

s22<br />

s32<br />

s42<br />

s23<br />

s33<br />

s43<br />

s24<br />

s34<br />

s44<br />

s25<br />

s35<br />

s45<br />

s26 ⎟ ⎜ σ2 ⎟<br />

⎟ ⎜ ⎟<br />

s36 ⎟ ⎜ σ3 ⎟<br />

⎟ ⎜ ⎟ .<br />

s46 ⎟ ⎜ σ4 ⎟<br />

⎠ ⎝ ⎠<br />

(2.4.6)<br />

e5<br />

e6<br />

s51 s52 s53 s54 s55 s56<br />

s61 s62 s63 s64 s65 s66<br />

Alternatively, in the arguments that follow, one can examine the equivalent form<br />

⎛ ⎞<br />

σ1<br />

⎛<br />

c11 c<strong>12</strong> c13 c14 c15<br />

⎞ ⎛ ⎞<br />

c16 e1<br />

⎜ σ2 ⎟ ⎜ c21<br />

⎜ ⎟ ⎜<br />

⎜ σ3 ⎟ ⎜ c31<br />

⎜ ⎟ = ⎜<br />

⎜ σ4 ⎟ ⎜ c41<br />

⎝ ⎠ ⎝<br />

c22<br />

c32<br />

c42<br />

c23<br />

c33<br />

c43<br />

c24<br />

c34<br />

c44<br />

c25<br />

c35<br />

c45<br />

c26 ⎟ ⎜ e2 ⎟<br />

⎟ ⎜ ⎟<br />

c36 ⎟ ⎜ e3 ⎟<br />

⎟ ⎜ ⎟ .<br />

c46 ⎟ ⎜ e4 ⎟<br />

⎠ ⎝ ⎠<br />

Material Symmetries<br />

σ5<br />

σ6<br />

c51 c52 c53 c54 c55 c56<br />

c61 c62 c63 c64 c65 c66<br />

A material (crystal) with one plane of symmetry is called an aelotropic material. If we let the x1x2<br />

plane be a plane of symmetry then the equations (2.4.6) must remain invariant under the coordinate<br />

transformation ⎛<br />

⎝ x1<br />

⎞ ⎛<br />

1<br />

x2 ⎠ = ⎝ 0<br />

0<br />

1<br />

0<br />

0<br />

⎞ ⎛<br />

⎠ ⎝<br />

x3 0 0 −1<br />

x1<br />

⎞<br />

x2 ⎠ (2.4.7)<br />

x3<br />

which represents an inversion of the x3 axis. That is, if the x1-x2 plane is a plane of symmetry we should be<br />

able to replace x3 by −x3 and the equations (2.4.6) should remain unchanged. This is equivalent to saying<br />

that a transformation of the type from equation (2.4.7) changes the Hooke’s law to the form ei = sijσj where<br />

the sij remain unaltered because it is the same material. Employing the transformation equations<br />

we examine the stress and strain transformation equations<br />

x1 = x1, x2 = x2, x3 = −x3 (2.4.8)<br />

∂xp ∂xq<br />

σij = σpq<br />

∂xi ∂xj<br />

and<br />

∂xp ∂xq<br />

eij = epq .<br />

∂xi ∂xj<br />

(2.4.9)<br />

If we expand both of the equations (2.4.9) and substitute in the nonzero derivatives<br />

we obtain the relations<br />

∂x1<br />

=1,<br />

∂x1<br />

σ11 = σ11<br />

σ22 = σ22<br />

σ33 = σ33<br />

σ21 = σ21<br />

σ31 = −σ31<br />

σ23 = −σ23<br />

∂x2<br />

=1,<br />

∂x2<br />

e11 = e11<br />

e22 = e22<br />

e33 = e33<br />

e21 = e21<br />

e31 = −e31<br />

e23 = −e23.<br />

σ5<br />

σ6<br />

e5<br />

e6<br />

∂x3<br />

= −1, (2.4.10)<br />

∂x3<br />

(2.4.11)<br />

245


246<br />

We conclude that if the material undergoes a strain, with the x1-x2 plane as a plane of symmetry then<br />

e5 and e6 change sign upon reversal of the x3 axis and e1,e2,e3,e4 remain unchanged. Similarly, we find σ5<br />

and σ6 change sign while σ1,σ2,σ3,σ4 remain unchanged. The equation (2.4.6) then becomes<br />

⎛<br />

e1<br />

⎞ ⎛<br />

s11 s<strong>12</strong> s13 s14 s15<br />

⎞ ⎛<br />

s16 σ1<br />

⎞<br />

⎜ e2 ⎟ ⎜ s21<br />

⎜ ⎟ ⎜<br />

⎜ e3 ⎟ ⎜ s31<br />

⎜ ⎟ = ⎜<br />

⎜ e4 ⎟ ⎜ s41<br />

⎝ ⎠ ⎝<br />

−e5 s51<br />

s22<br />

s32<br />

s42<br />

s52<br />

s23<br />

s33<br />

s43<br />

s53<br />

s24<br />

s34<br />

s44<br />

s54<br />

s25<br />

s35<br />

s45<br />

s55<br />

s26 ⎟ ⎜ σ2 ⎟<br />

⎟ ⎜ ⎟<br />

s36 ⎟ ⎜ σ3 ⎟<br />

⎟ ⎜ ⎟ .<br />

s46 ⎟ ⎜ σ4 ⎟<br />

⎠ ⎝ ⎠<br />

s56 −σ5<br />

(2.4.<strong>12</strong>)<br />

−e6 s61 s62 s63 s64 s65 s66 −σ6<br />

If the stress strain relation for the new orientation of the x3 axis is to have the same form as the<br />

old orientation, then the equations (2.4.6) and (2.4.<strong>12</strong>) must give the same results. Comparison of these<br />

equations we find that<br />

s15 = s16 =0<br />

s25 = s26 =0<br />

s35 = s36 =0<br />

s45 = s46 =0<br />

s51 = s52 = s53 = s54 =0<br />

s61 = s62 = s63 = s64 =0.<br />

(2.4.13)<br />

In summary, from an examination of the equations (2.4.6) and (2.4.<strong>12</strong>) we find that for an aelotropic<br />

material (crystal), with one plane of symmetry, the 36 constants sij reduce to 20 constants and the generalized<br />

Hooke’s law (constitutive equation) has the form<br />

⎛ ⎞<br />

e1<br />

⎛<br />

s11 s<strong>12</strong> s13 s14 0 0<br />

⎞ ⎛ ⎞<br />

σ1<br />

⎜ e2 ⎟ ⎜ s21<br />

⎜ ⎟ ⎜<br />

⎜ e3 ⎟ ⎜ s31<br />

⎜ ⎟ = ⎜<br />

⎜ e4 ⎟ ⎜ s41<br />

⎝ ⎠ ⎝<br />

s22<br />

s32<br />

s42<br />

s23<br />

s33<br />

s43<br />

s24<br />

s34<br />

s44<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

⎟ ⎜ σ2 ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ σ3 ⎟<br />

⎟ ⎜ ⎟ .<br />

⎟ ⎜ σ4 ⎟<br />

⎠ ⎝ ⎠<br />

(2.4.14)<br />

e5<br />

e6<br />

0 0 0 0 s55 s56<br />

0 0 0 0 s65 s66<br />

Alternatively, the Hooke’s law can be represented in the form<br />

⎛ ⎞ ⎛<br />

σ1<br />

⎜ σ2 ⎟ ⎜<br />

⎜ ⎟ ⎜<br />

⎜ σ3 ⎟ ⎜<br />

⎜ ⎟ = ⎜<br />

⎜ σ4 ⎟ ⎜<br />

⎝ ⎠ ⎝<br />

σ5<br />

σ6<br />

c11 c<strong>12</strong> c13 c14 0 0<br />

c21 c22 c23 c24 0 0<br />

c31 c32 c33 c34 0 0<br />

c41 c42 c43 c44 0 0<br />

0 0 0 0 c55 c56<br />

0 0 0 0 c65 c66<br />

⎞ ⎛<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

σ5<br />

σ6<br />

e1<br />

e2<br />

e3<br />

e4<br />

e5<br />

e6<br />

⎞<br />

⎟ .<br />

⎟<br />


Additional Symmetries<br />

If the material (crystal) is such that there is an additional plane of symmetry, say the x2-x3 plane, then<br />

reversal of the x1 axis should leave the equations (2.4.14) unaltered. If there are two planes of symmetry<br />

then there will automatically be a third plane of symmetry. Such a material (crystal) is called orthotropic.<br />

Introducing the additional transformation<br />

x1 = −x1, x2 = x2, x3 = x3<br />

which represents the reversal of the x1 axes, the expanded form of equations (2.4.9) are used to calculate the<br />

effect of such a transformation upon the stress and strain tensor. We find σ1,σ2,σ3,σ6,e1,e2,e3,e6 remain<br />

unchanged while σ4,σ5,e4,e5 change sign. The equation (2.4.14) then becomes<br />

⎛<br />

e1<br />

⎞ ⎛<br />

⎜ e2 ⎟ ⎜<br />

⎜ ⎟ ⎜<br />

⎜ e3 ⎟ ⎜<br />

⎜ ⎟ = ⎜<br />

⎜ −e4 ⎟ ⎜<br />

⎝ ⎠ ⎝<br />

−e5<br />

e6<br />

s11 s<strong>12</strong> s13 s14 0 0<br />

s21 s22 s23 s24 0 0<br />

s31 s32 s33 s34 0 0<br />

s41 s42 s43 s44 0 0<br />

0 0 0 0 s55 s56<br />

0 0 0 0 s65 s66<br />

⎞ ⎛<br />

σ1<br />

⎞<br />

⎟ ⎜ σ2 ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ σ3 ⎟<br />

⎟ ⎜ ⎟ .<br />

⎟ ⎜ −σ4 ⎟<br />

⎠ ⎝ ⎠<br />

−σ5<br />

(2.4.15)<br />

Note that if the constitutive equations (2.4.14) and (2.4.15) are to produce the same results upon reversal<br />

of the x1 axes, then we require that the following coefficients be equated to zero:<br />

This then produces the constitutive equation<br />

or its equivalent form<br />

⎛<br />

⎜<br />

⎝<br />

⎛<br />

⎜<br />

⎝<br />

e1<br />

e2<br />

e3<br />

e4<br />

e5<br />

e6<br />

σ1<br />

σ2<br />

σ3<br />

σ4<br />

σ5<br />

σ6<br />

⎞ ⎛<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ = ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

⎞ ⎛<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ = ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

s14 = s24 = s34 =0<br />

s41 = s42 = s43 =0<br />

s56 = s65 =0.<br />

s11 s<strong>12</strong> s13 0 0 0<br />

s21 s22 s23 0 0 0<br />

s31 s32 s33 0 0 0<br />

0 0 0 s44 0 0<br />

0 0 0 0 s55 0<br />

0 0 0 0 0 s66<br />

c11 c<strong>12</strong> c13 0 0 0<br />

c21 c22 c23 0 0 0<br />

c31 c32 c33 0 0 0<br />

0 0 0 c44 0 0<br />

0 0 0 0 c55 0<br />

0 0 0 0 0 c66<br />

⎞ ⎛<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

⎞ ⎛<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

σ6<br />

σ1<br />

σ2<br />

σ3<br />

σ4<br />

σ5<br />

σ6<br />

e1<br />

e2<br />

e3<br />

e4<br />

e5<br />

e6<br />

⎞<br />

⎟<br />

⎠<br />

⎞<br />

⎟<br />

⎠<br />

(2.4.16)<br />

and the original 36 constants have been reduced to <strong>12</strong> constants. This is the constitutive equation for<br />

orthotropic material (crystals).<br />

247


248<br />

Axis of Symmetry<br />

If in addition to three planes of symmetry there is an axis of symmetry then the material (crystal) is<br />

termed hexagonal. Assume that the x1 axis is an axis of symmetry and consider the effect of the transformation<br />

x 1 = x 1 , x 2 = x 3<br />

x 3 = −x 2<br />

upon the constitutive equations. It is left as an exercise to verify that the constitutive equations reduce to<br />

the form where there are 7 independent constants having either of the forms<br />

⎛ ⎞<br />

e1<br />

⎛<br />

s11 s<strong>12</strong> s<strong>12</strong> 0 0 0<br />

⎞ ⎛ ⎞<br />

σ1<br />

⎜ e2 ⎟ ⎜ s21<br />

⎜ ⎟ ⎜<br />

⎜ e3 ⎟ ⎜<br />

⎜ ⎟ = ⎜<br />

⎜ e4 ⎟ ⎜<br />

⎝ ⎠ ⎝<br />

s22 s23 0 0 0 ⎟ ⎜ σ2 ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ σ3 ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ σ4 ⎟<br />

⎠ ⎝ ⎠<br />

or ⎛<br />

⎜<br />

⎝<br />

e5<br />

e6<br />

σ1<br />

σ2<br />

σ3<br />

σ4<br />

σ5<br />

σ6<br />

⎞ ⎛<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ = ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

s21 s23 s22 0 0 0<br />

0 0 0 s44 0 0<br />

0 0 0 0 s44 0<br />

0 0 0 0 0 s66<br />

c11 c<strong>12</strong> c<strong>12</strong> 0 0 0<br />

c21 c22 c23 0 0 0<br />

c21 c23 c22 0 0 0<br />

0 0 0 c44 0 0<br />

0 0 0 0 c44 0<br />

0 0 0 0 0 c66<br />

Finally, if the material is completely symmetric, the x2 axis is also an axis of symmetry and we can<br />

consider the effect of the transformation<br />

x 1 = −x 3 , x 2 = x 2 , x 3 = x 1<br />

upon the constitutive equations.<br />

constitutive equation to the form<br />

It can be verified that these transformations reduce the Hooke’s law<br />

⎛ ⎞<br />

e1<br />

⎛<br />

s11 s<strong>12</strong> s<strong>12</strong> 0 0 0<br />

⎞ ⎛ ⎞<br />

σ1<br />

⎜ e2 ⎟ ⎜ s<strong>12</strong><br />

⎜ ⎟ ⎜<br />

⎜ e3 ⎟ ⎜<br />

⎜ ⎟ = ⎜<br />

⎜ e4 ⎟ ⎜<br />

⎝ ⎠ ⎝<br />

s11 s<strong>12</strong> 0 0 0 ⎟ ⎜ σ2 ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ σ3 ⎟<br />

⎟ ⎜ ⎟ .<br />

⎟ ⎜ σ4 ⎟<br />

⎠ ⎝ ⎠<br />

(2.4.17)<br />

e5<br />

e6<br />

s<strong>12</strong> s<strong>12</strong> s11 0 0 0<br />

0 0 0 s44 0 0<br />

0 0 0 0 s44 0<br />

0 0 0 0 0 s44<br />

Materials (crystals) with atomic arrangements that exhibit the above symmetries are called isotropic<br />

materials. An equivalent form of (2.4.17) is the relation<br />

⎛ ⎞<br />

σ1<br />

⎛<br />

c11 c<strong>12</strong> c<strong>12</strong> 0 0 0<br />

⎞ ⎛ ⎞<br />

e1<br />

⎜ σ2 ⎟ ⎜ c<strong>12</strong><br />

⎜ ⎟ ⎜<br />

⎜ σ3 ⎟ ⎜<br />

⎜ ⎟ = ⎜<br />

⎜ σ4 ⎟ ⎜<br />

⎝ ⎠ ⎝<br />

c11 c<strong>12</strong> 0 0 0 ⎟ ⎜ e2 ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ e3 ⎟<br />

⎟ ⎜ ⎟ .<br />

⎟ ⎜ e4 ⎟<br />

⎠ ⎝ ⎠<br />

σ5<br />

σ6<br />

c<strong>12</strong> c<strong>12</strong> c11 0 0 0<br />

0 0 0 c44 0 0<br />

0 0 0 0 c44 0<br />

0 0 0 0 0 c44<br />

⎞ ⎛<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

The figure 2.4-1 lists values for the elastic stiffness associated with some metals which are isotropic 1<br />

1Additional constants are given in “International Tables of Selected Constants”, Metals: Thermal and<br />

Mechanical Data, Vol. 16, Edited by S. Allard, Pergamon Press, 1969.<br />

σ5<br />

σ6<br />

e1<br />

e2<br />

e3<br />

e4<br />

e5<br />

e6<br />

σ5<br />

σ6<br />

e5<br />

e6<br />

⎞<br />

⎟ .<br />

⎟<br />


Metal c11 c<strong>12</strong> c44<br />

Na 0.074 0.062 0.042<br />

Pb 0.495 0.423 0.149<br />

Cu 1.684 1.214 0.754<br />

Ni 2.508 1.500 1.235<br />

Cr 3.500 0.678 1.008<br />

Mo 4.630 1.610 1.090<br />

W 5.233 2.045 1.607<br />

Figure 2.4-1. Elastic stiffness coefficients for some metals which are cubic.<br />

Constants are given in units of 10 <strong>12</strong> dynes/cm 2<br />

Under these conditions the stress strain constitutive relations can be written as<br />

Isotropic Material<br />

σ1 = σ11 =(c11 − c<strong>12</strong>)e11 + c<strong>12</strong>(e11 + e22 + e33)<br />

σ2 = σ22 =(c11 − c<strong>12</strong>)e22 + c<strong>12</strong>(e11 + e22 + e33)<br />

σ3 = σ33 =(c11 − c<strong>12</strong>)e33 + c<strong>12</strong>(e11 + e22 + e33)<br />

σ4 = σ<strong>12</strong> = c44e<strong>12</strong><br />

σ5 = σ13 = c44e13<br />

σ6 = σ23 = c44e23.<br />

(2.4.18)<br />

Materials (crystals) which are elastically the same in all directions are called isotropic. We have shown<br />

that for a cubic material which exhibits symmetry with respect to all axes and planes, the constitutive<br />

stress-strain relation reduces to the form found in equation (2.4.17). Define the quantities<br />

s11 = 1<br />

E , s<strong>12</strong> = − ν<br />

E , s44 = 1<br />

2µ<br />

where E is the Young’s Modulus of elasticity, ν is the Poisson’s ratio, and µ is the shear or rigidity modulus.<br />

For isotropic materials the three constants E,ν,µ are not independent as the following example demonstrates.<br />

EXAMPLE 2.4-1. (Elastic constants) For an isotropic material, consider a cross section of material in<br />

the x 1 -x 2 plane which is subjected to pure shearing so that σ4 = σ<strong>12</strong> is the only nonzero stress as illustrated<br />

in the figure 2.4-2.<br />

For the above conditions, the equation (2.4.17) reduces to the single equation<br />

e4 = e<strong>12</strong> = s44σ4 = s44σ<strong>12</strong> or µ = σ<strong>12</strong><br />

and so the shear modulus is the ratio of the shear stress to the shear angle. Now rotate the axes through a<br />

45 degree angle to a barred system of coordinates where<br />

γ<strong>12</strong><br />

x 1 = x 1 cos α − x 2 sin α x 2 = x 1 sin α + x 2 cos α<br />

249


250<br />

Figure 2.4-2. Element subjected to pure shearing<br />

where α = π<br />

4 . Expanding the transformation equations (2.4.9) we find that<br />

and similarly<br />

In the barred system, the Hooke’s law becomes<br />

σ1 = σ11 =cosα sin ασ<strong>12</strong> +sinα cos ασ21 = σ<strong>12</strong> = σ4<br />

σ2 = σ22 = − sin α cos ασ<strong>12</strong> − sin α cos ασ21 = −σ<strong>12</strong> = −σ4,<br />

e1 = e11 = e4, e2 = e22 = −e4.<br />

e1 = s11σ1 + s<strong>12</strong>σ2<br />

or<br />

e4 = s11σ4 − s<strong>12</strong>σ4 = s44σ4.<br />

Hence, the constants s11,s<strong>12</strong>,s44 are related by the relation<br />

s11 − s<strong>12</strong> = s44<br />

or<br />

1 ν<br />

+<br />

E E<br />

1<br />

= . (2.4.19)<br />

2µ<br />

This is an important relation connecting the elastic constants associated with isotropic materials. The<br />

above transformation can also be applied to triclinic, aelotropic, orthotropic, and hexagonal materials to<br />

find relationships between the elastic constants.<br />

Observe also that some texts postulate the existence of a strain energy function U ∗ which has the<br />

property that σij = ∂U∗ . In this case the strain energy function, in the single index notation, is written<br />

∂eij<br />

U ∗ = cijeiej where cij and consequently sij are symmetric. In this case the previous discussed symmetries<br />

give the following results for the nonzero elastic compliances sij : 13 nonzero constants instead of 20 for<br />

aelotropic material, 9 nonzero constants instead of <strong>12</strong> for orthotropic material, and 6 nonzero constants<br />

instead of 7 for hexagonal material. This is because of the additional property that sij = sji be symmetric.


The previous discussion has shown that for an isotropic material the generalized Hooke’s law (constitutive<br />

equations) have the form<br />

e11 = 1<br />

E [σ11 − ν(σ22 + σ33)]<br />

e22 = 1<br />

E [σ22 − ν(σ33 + σ11)]<br />

e33 = 1<br />

E [σ33 − ν(σ11 + σ22)]<br />

e21 = e<strong>12</strong> = 1+ν<br />

E σ<strong>12</strong><br />

e32 = e23 = 1+ν<br />

E σ23<br />

e31 = e13 = 1+ν<br />

E σ13<br />

, (2.4.20)<br />

where equation (2.4.19) holds. These equations can be expressed in the indicial notation and have the form<br />

eij = 1+ν<br />

E σij − ν<br />

E σkkδij, (2.4.21)<br />

where σkk = σ11 + σ22 + σ33 is a stress invariant and δij is the Kronecker delta. We can solve for the stress<br />

in terms of the strain by performing a contraction on i and j in equation (2.4.21). This gives the dilatation<br />

eii = 1+ν<br />

E σii − 3ν<br />

E σkk<br />

1 − 2ν<br />

=<br />

E σkk.<br />

Note that from the result in equation (2.4.21) we are now able to solve for the stress in terms of the strain.<br />

We find<br />

eij = 1+ν<br />

E σij − ν<br />

1 − 2ν ekkδij<br />

E<br />

1+ν eij<br />

νE<br />

= σij −<br />

(1 + ν)(1 − 2ν) ekkδij<br />

or σij = E<br />

1+ν eij<br />

νE<br />

+<br />

(1 + ν)(1 − 2ν) ekkδij.<br />

The tensor equation (2.4.22) represents the six scalar equations<br />

σ11 =<br />

σ22 =<br />

σ33 =<br />

E<br />

(1 + ν)(1 − 2ν) [(1 − ν)e11 + ν(e22 + e33)]<br />

E<br />

(1 + ν)(1 − 2ν) [(1 − ν)e22 + ν(e33 + e11)]<br />

E<br />

(1 + ν)(1 − 2ν) [(1 − ν)e33 + ν(e22 + e11)]<br />

σ<strong>12</strong> = E<br />

1+ν e<strong>12</strong><br />

σ13 = E<br />

1+ν e13<br />

σ23 = E<br />

1+ν e23.<br />

(2.4.22)<br />

251


252<br />

Alternative Approach to Constitutive Equations<br />

The constitutive equation defined by Hooke’s generalized law for isotropic materials can be approached<br />

from another point of view. Consider the generalized Hooke’s law<br />

σij = cijklekl, i,j,k,l=1, 2, 3.<br />

If we transform to a barred system of coordinates, we will have the new Hooke’s law<br />

For an isotropic material we require that<br />

σij = cijklekl, i,j,k,l=1, 2, 3.<br />

cijkl = cijkl.<br />

Tensors whose components are the same in all coordinate systems are called isotropic tensors. We have<br />

previously shown in Exercise 1.3, problem 18, that<br />

cpqrs = λδpqδrs + µ(δprδqs + δpsδqr)+κ(δprδqs − δpsδqr)<br />

is an isotropic tensor when we consider affine type transformations. If we further require the symmetry<br />

conditions found in equations (2.4.3) be satisfied, we find that κ = 0 and consequently the generalized<br />

Hooke’s law must have the form<br />

σpq = cpqrsers =[λδpqδrs + µ(δprδqs + δpsδqr)] ers<br />

σpq = λδpqerr + µ(epq + eqp)<br />

or σpq =2µepq + λerrδpq,<br />

(2.4.23)<br />

where err = e11 + e22 + e33 = Θ is the dilatation. The constants λ and µ are called Lame’s constants.<br />

Comparing the equation (2.4.22) with equation (2.4.23) we find that the constants λ and µ satisfy the<br />

relations<br />

E<br />

νE<br />

µ =<br />

λ =<br />

. (2.4.24)<br />

2(1 + ν)<br />

(1 + ν)(1 − 2ν)<br />

In addition to the constants E,ν,µ,λ, it is sometimes convenient to introduce the constant k, called the bulk<br />

modulus of elasticity, (Exercise 2.3, problem 23), defined by<br />

k =<br />

E<br />

. (2.4.25)<br />

3(1 − 2ν)<br />

The stress-strain constitutive equation (2.4.23) was derived using Cartesian tensors. To generalize the<br />

equation (2.4.23) we consider a transformation from a Cartesian coordinate system yi ,i=1, 2, 3toageneral<br />

coordinate system xi ,i=1, 2, 3. We employ the relations<br />

and<br />

σmn = σij<br />

g ij = ∂ym<br />

∂x i<br />

∂ym ∂xj , gij = ∂xi<br />

∂ym ∂xj ∂ym ∂yi ∂xm ∂yj ∂xn , emn<br />

∂y<br />

= eij<br />

i<br />

∂xm ∂yj ∂xn , or erq<br />

∂x<br />

= eij<br />

i<br />

∂yr ∂x j<br />

∂y q


and convert equation (2.4.23) to a more generalized form. Multiply equation (2.4.23) by ∂yp<br />

∂xm the result<br />

which can be simplified to the form<br />

Dropping the bar notation, we have<br />

The contravariant form of this equation is<br />

σmn = λ ∂yq<br />

∂xm ∂yq ∂xn err + µ (emn + enm) ,<br />

σmn = λg mneijg ij + µ (emn + enm) .<br />

σmn = λgmng ij eij + µ (emn + enm) .<br />

σ sr = λg sr g ij eij + µ (g ms g nr + g ns g mr ) emn.<br />

Employing the equations (2.4.24) the above result can also be expressed in the form<br />

σ rs =<br />

∂yq n and verify<br />

∂x<br />

<br />

E<br />

g<br />

2(1 + ν)<br />

ms g nr + g ns g mr + 2ν<br />

1 − 2ν gsrg mn<br />

<br />

emn. (2.4.26)<br />

This is a more general form for the stress-strain constitutive equations which is valid in all coordinate systems.<br />

Multiplying by gsk and employing the use of associative tensors, one can verify<br />

σ i j<br />

<br />

E<br />

= e<br />

1+ν<br />

i ν<br />

j +<br />

1 − 2ν emm δi <br />

j<br />

or σ i j =2µe i j + λe m mδ i j,<br />

are alternate forms for the equation (2.4.26). As an exercise, solve for the strains in terms of the stresses<br />

and show that<br />

Ee i j =(1+ν)σ i j − νσ m mδ i j.<br />

EXAMPLE 2.4-2. (Hooke’s law) Let us construct a simple example to test the results we have<br />

developed so far. Consider the tension in a cylindrical bar illustrated in the figure 2.4-3.<br />

Figure 2.4-3. Stress in a cylindrical bar<br />

253


254<br />

Assume that<br />

⎛<br />

F<br />

A<br />

σij = ⎝ 0<br />

0<br />

0<br />

⎞<br />

0<br />

0⎠<br />

0 0 0<br />

where F is the constant applied force and A is the cross sectional area of the cylinder. Consequently, the<br />

generalized Hooke’s law (2.4.21) produces the nonzero strains<br />

From these equations we obtain:<br />

The first part of Hooke’s law<br />

The second part of Hooke’s law<br />

e11 = 1+ν<br />

E σ11 − ν<br />

E (σ11 + σ22 + σ33) = σ11<br />

E<br />

e22 = −ν<br />

E σ11<br />

e33 = −ν<br />

E σ11<br />

σ11 = Ee11 or F<br />

= Ee11.<br />

A<br />

lateral contraction −e22<br />

= =<br />

longitudinal extension e11<br />

−e33<br />

= ν = Poisson’s ratio.<br />

e11<br />

This example demonstrates that the generalized Hooke’s law for homogeneous and isotropic materials<br />

reduces to our previous one dimensional result given in (2.3.1) and (2.3.2).<br />

Basic Equations of Elasticity<br />

Assuming the density ϱ is constant, the basic equations of elasticity reduce to the equations representing<br />

conservation of linear momentum and angular momentum together with the strain-displacement relations<br />

and constitutive equations. In these equations the body forces are assumed known. These basic equations<br />

produce 15 equations in 15 unknowns and are a formidable set of equations to solve. Methods for solving<br />

these simultaneous equations are: 1) Express the linear momentum equations in terms of the displacements<br />

ui and obtain a system of partial differential equations. Solve the system of partial differential equations<br />

for the displacements ui and then calculate the corresponding strains. The strains can be used to calculate<br />

the stresses from the constitutive equations. 2) Solve for the stresses and from the stresses calculate the<br />

strains and from the strains calculate the displacements. This converse problem requires some additional<br />

considerations which will be addressed shortly.


Navier’s Equations<br />

Basic Equations of Linear Elasticity<br />

• Conservation of linear momentum.<br />

σ ij<br />

,i + ϱbj = ϱf j<br />

j =1, 2, 3. (2.4.27(a))<br />

where σ ij is the stress tensor, b j is the body force per unit mass and f j is<br />

the acceleration. If there is no motion, then f j = 0 and these equations<br />

reduce to the equilibrium equations<br />

• Conservation of angular momentum. σij = σji<br />

• Strain tensor.<br />

σ ij<br />

,i + ϱbj =0 j =1, 2, 3. (2.4.27(b))<br />

eij = 1<br />

2 (ui,j + uj,i) (2.4.28)<br />

where ui denotes the displacement field.<br />

• Constitutive equation. For a linear elastic isotropic material we have<br />

σ i j<br />

= E<br />

1+ν ei j +<br />

or its equivalent form<br />

E<br />

(1 + ν)(1 − 2ν) ek k δi j i, j =1, 2, 3 (2.4.29(a))<br />

σ i j =2µei j + λer r δi j i, j =1, 2, 3, (2.4.29(b))<br />

where e r r is the dilatation. This produces 15 equations for the 15 unknowns<br />

u1,u2,u3,σ11,σ<strong>12</strong>,σ13,σ22,σ23,σ33,e11,e<strong>12</strong>,e13,e22,e23,e33,<br />

which represents 3 displacements, 6 strains and 6 stresses. In the above<br />

equations it is assumed that the body forces are known.<br />

The equations (2.4.27) through (2.4.29) can be combined and written as one set of equations. The<br />

resulting equations are known as Navier’s equations for the displacements ui over the range i =1, 2, 3. To<br />

derive the Navier’s equations in Cartesian coordinates, we write the equations (2.4.27),(2.4.28) and (2.4.29)<br />

in Cartesian coordinates. We then calculate σij,j in terms of the displacements ui and substitute the results<br />

into the momentum equation (2.4.27(a)). Differentiation of the constitutive equations (2.4.29(b)) produces<br />

σij,j =2µeij,j + λekk,jδij. (2.4.30)<br />

255


256<br />

A contraction of the strain produces the dilatation<br />

err = 1<br />

2 (ur,r + ur,r) =ur,r<br />

From the dilatation we calculate the covariant derivative<br />

(2.4.31)<br />

ekk,j = uk,kj. (2.4.32)<br />

Employing the strain relation from equation (2.4.28), we calculate the covariant derivative<br />

eij,j = 1<br />

2 (ui,jj + uj,ij). (2.4.33)<br />

These results allow us to express the covariant derivative of the stress in terms of the displacement field. We<br />

find<br />

σij,j = µ [ui,jj + uj,ij]+λδijuk,kj<br />

or σij,j =(λ + µ)uk,ki + µui,jj.<br />

Substituting equation (2.4.34) into the linear momentum equation produces the Navier equations:<br />

In vector form these equations can be expressed<br />

(2.4.34)<br />

(λ + µ)uk,ki + µui,jj + ϱbi = ϱfi, i =1, 2, 3. (2.4.35)<br />

(λ + µ)∇ (∇·u)+µ∇ 2 u + ϱ b = ϱ f, (2.4.36)<br />

where u is the displacement vector, b is the body force per unit mass and f is the acceleration. In Cartesian<br />

coordinates these equations have the form:<br />

2 ∂ u1<br />

(λ + µ) +<br />

∂x1∂xi<br />

∂2u2 +<br />

∂x2∂xi<br />

∂2 <br />

u3<br />

+ µ∇<br />

∂x3∂xi<br />

2 ui + ϱbi = ϱ ∂2ui ,<br />

∂t2 for i =1, 2, 3, where<br />

∇ 2 ui = ∂2ui ∂x1 2 + ∂2ui ∂x2 2 + ∂2ui .<br />

∂x3<br />

2<br />

The Navier equations must be satisfied by a set of functions ui = ui(x1,x2,x3) which represent the<br />

displacement at each point inside some prescribed region R. Knowing the displacement field we can calculate<br />

the strain field directly using the equation (2.4.28). Knowledge of the strain field enables us to construct the<br />

corresponding stress field from the constitutive equations.<br />

In the absence of body forces, such as gravity, the solution to equation (2.4.36) can be represented<br />

in the form u = u (1) + u (2) , where u (1) satisfies div u (1) = ∇·u (1) = 0 and the vector u (2) satisfies<br />

curl u (2) = ∇×u (2) =0. The vector field u (1) is called a solenoidal field, while the vector field u (2) is<br />

called an irrotational field. Substituting u into the equation (2.4.36) and setting b =0, we find in Cartesian<br />

coordinates that<br />

2 (1) ∂ u<br />

ϱ<br />

∂t2 + ∂2u (2)<br />

∂t2 <br />

<br />

=(λ + µ)∇ ∇·u (2)<br />

+ µ∇ 2 u (1) + µ∇ 2 u (2) . (2.4.37)


The vector field u (1) can be eliminated from equation (2.4.37) by taking the divergence of both sides of the<br />

equation. This produces<br />

ϱ ∂2 ∇·u (2)<br />

∂t 2<br />

=(λ + µ)∇ 2 (∇·u (2) )+µ∇·∇ 2 u (2) .<br />

The displacement field is assumed to be continuous and so we can interchange the order of the operators ∇ 2<br />

and ∇ and write<br />

This last equation implies that<br />

<br />

∇· ϱ ∂2u (2)<br />

∂t2 − (λ +2µ)∇2u (2)<br />

<br />

=0.<br />

ϱ ∂2 u (2)<br />

∂t 2<br />

=(λ +2µ)∇2 u (2)<br />

and consequently, u (2) is a vector wave which moves with the speed (λ +2µ)/ϱ. Similarly, when the vector<br />

field u (2) is eliminated from the equation (2.4.37), by taking the curl of both sides, we find the vector u (1)<br />

also satisfies a wave equation having the form<br />

ϱ ∂2 u (1)<br />

∂t 2<br />

= µ∇2 u (1) .<br />

This later wave moves with the speed µ/ϱ. The vector u (2) is a compressive wave, while the wave u (1) is<br />

ashearingwave.<br />

The exercises 30 through 38 enable us to write the Navier’s equations in Cartesian, cylindrical or<br />

spherical coordinates. In particular, we have for cartesian coordinates<br />

(λ + µ)( ∂2u ∂x2 + ∂2v ∂x∂y + ∂2w ∂x∂z )+µ(∂2 u<br />

∂x2 + ∂2u ∂y2 + ∂2u ∂z2 )+ϱbx =ϱ ∂2u ∂t2 (λ + µ)( ∂2u ∂x∂y + ∂2v ∂y2 + ∂2w ∂y∂z )+µ( ∂2v ∂x2 + ∂2v ∂y2 + ∂2v ∂z2 )+ϱby =ϱ ∂2v ∂t2 (λ + µ)( ∂2u ∂x∂z + ∂2v ∂y∂z + ∂2w ∂z2 )+µ(∂2 w<br />

∂x2 + ∂2w ∂y2 + ∂2w ∂z2 )+ϱbz =ϱ ∂2w ∂t2 and in cylindrical coordinates<br />

(λ + µ) ∂<br />

<br />

1<br />

∂r r<br />

µ( ∂2ur 1 ∂ur 1<br />

+ +<br />

∂r2 r ∂r r2 ∂2ur µ( ∂2uθ 1 ∂uθ 1<br />

+ +<br />

∂r2 r ∂r r2 ∂2uθ ∂<br />

∂r<br />

<br />

1 ∂uθ ∂uz<br />

(rur)+ + +<br />

r ∂θ ∂z<br />

∂θ2 + ∂2ur ur 2<br />

− −<br />

∂z2 r2 r2 ∂uθ<br />

∂θ )+ϱbr =ϱ ∂2ur ∂t2 (λ + µ) 1<br />

<br />

∂ 1 ∂ 1 ∂uθ ∂uz<br />

(rur)+ + +<br />

r ∂θ r ∂r r ∂θ ∂z<br />

∂θ2 + ∂2uθ 2<br />

+<br />

∂z2 r2 ∂ur uθ<br />

−<br />

∂θ r2 )+ϱbθ =ϱ ∂2uθ ∂t2 (λ + µ) ∂<br />

<br />

1 ∂ 1 ∂uθ ∂uz<br />

(rur)+ + +<br />

∂z r ∂r r ∂θ ∂z<br />

µ( ∂2uz 1 ∂uz 1<br />

+ +<br />

∂r2 r ∂r r2 ∂2uz ∂θ2 + ∂2uz ∂z2 )+ϱbz =ϱ ∂2uz ∂t2 257


258<br />

and in spherical coordinates<br />

(λ + µ) ∂<br />

<br />

1<br />

∂ρ ρ2 ∂<br />

∂ρ (ρ2uρ)+ 1 ∂<br />

ρ sin θ ∂θ (uθ sin θ)+ 1<br />

<br />

∂uφ<br />

+<br />

ρ sin θ ∂φ<br />

µ(∇ 2 uρ − 2<br />

ρ2 uρ − 2<br />

ρ2 ∂uθ<br />

∂θ − 2uθ cot θ<br />

ρ2 2<br />

−<br />

ρ2 ∂uφ<br />

sin θ ∂φ )+ϱbρ =ϱ ∂2uρ ∂t2 (λ + µ) 1<br />

<br />

∂ 1<br />

ρ ∂θ ρ2 ∂<br />

∂ρ (ρ2uρ)+ 1 ∂<br />

ρ sin θ ∂θ (uθ sin θ)+ 1<br />

<br />

∂uφ<br />

+<br />

ρ sin θ ∂φ<br />

µ(∇ 2 uθ + 2<br />

ρ2 ∂uρ<br />

∂θ −<br />

uθ<br />

ρ2 sin 2 2<br />

−<br />

θ ρ2 cos θ<br />

sin 2 ∂uφ<br />

θ ∂φ )+ϱbθ =ϱ ∂2uθ ∂t2 <br />

1 ∂ 1<br />

(λ + µ)<br />

ρ sin θ ∂φ ρ2 ∂<br />

∂ρ (ρ2uρ)+ 1 ∂<br />

ρ sin θ ∂θ (uθ sin θ)+ 1<br />

<br />

∂uφ<br />

+<br />

ρ sin θ ∂φ<br />

µ(∇ 2 uφ −<br />

1<br />

ρ2 sin 2 θ uφ<br />

2<br />

+<br />

ρ2 ∂uρ 2cosθ<br />

+<br />

sin θ ∂φ ρ2 sin 2 ∂uθ<br />

θ ∂φ )+ϱbφ =ϱ ∂2uφ ∂t2 where ∇ 2 is determined from either equation (2.1.<strong>12</strong>) or (2.1.13).<br />

Boundary Conditions<br />

In elasticity the body forces per unit mass (bi,i =1, 2, 3) are assumed known. In addition one of the<br />

following type of boundary conditions is usually prescribed:<br />

• The displacements ui, i =1, 2, 3 are prescribed on the boundary of the region R over which a solution<br />

is desired.<br />

• The stresses (surface tractions) are prescribed on the boundary of the region R over which a solution is<br />

desired.<br />

• The displacements ui,i = 1, 2, 3 are given over one portion of the boundary and stresses (surface<br />

tractions) are specified over the remaining portion of the boundary. This type of boundary condition is<br />

known as a mixed boundary condition.<br />

General Solution of Navier’s Equations<br />

There has been derived a general solution to the Navier’s equations. It is known as the Papkovich-Neuber<br />

solution. In the case of a solid in equilibrium one must solve the equilibrium equations<br />

(λ + µ)∇ (∇·u)+µ∇ 2 u + ϱb =0 or<br />

∇ 2 u + 1<br />

ϱ<br />

∇(∇·u)+<br />

1 − 2ν µ b =0 (ν = 1<br />

2 )<br />

(2.4.38)


THEOREM A general elastostatic solution of the equation (2.4.38) in terms of harmonic potentials φ, is<br />

ψ<br />

u =grad(φ + r · ψ) − 4(1 − ν) ψ (2.4.39)<br />

where φ and ψ are continuous solutions of the equations<br />

∇ 2 φ = −ϱr · b<br />

4µ(1 − ν)<br />

and ∇ 2 ψ =<br />

ϱ b<br />

4µ(1 − ν)<br />

with r = x ê1 + y ê2 + z ê3 a position vector to a general point (x, y, z) within the continuum.<br />

Proof: First we write equation (2.4.38) in the tensor form<br />

Now our problem is to show that equation (2.4.39), in tensor form,<br />

(2.4.40)<br />

ui,kk + 1<br />

1 − 2ν (uj,j) ,i + ϱ<br />

µ bi =0 (2.4.41)<br />

ui = φ,i +(xjψj),i − 4(1 − ν)ψi<br />

is a solution of equation (2.4.41). Toward this purpose, we differentiate equation (2.4.42)<br />

and then contract on i and k giving<br />

ui,k = φ,ik +(xjψj),ik − 4(1 − ν)ψi,k<br />

(2.4.42)<br />

(2.4.43)<br />

ui,i = φ,ii +(xjψj),ii − 4(1 − ν)ψi,i. (2.4.44)<br />

Employing the identity (xjψj),ii =2ψi,i + xiψi,kk the equation (2.4.44) becomes<br />

By differentiating equation (2.4.43) we establish that<br />

We use the hypothesis<br />

φ,kk = −ϱxjFj<br />

4µ(1 − ν)<br />

and simplify the equation (2.4.46) to the form<br />

ui,i = φ,ii +2ψi,i + xiψi,kk − 4(1 − ν)ψi,i. (2.4.45)<br />

ui,kk = φ,ikk +(xjψj),ikk − 4(1 − ν)ψi,kk<br />

Also by differentiating (2.4.45) one can establish that<br />

=(φ,kk),i +((xjψj),kk) ,i − 4(1 − ν)ψi,kk<br />

=[φ,kk +2ψj,j + xjψj,kk] ,i − 4(1 − ν)ψi,kk.<br />

and ψj,kk =<br />

ϱFj<br />

4µ(1 − ν) ,<br />

(2.4.46)<br />

ui,kk =2ψj,ji − 4(1 − ν)ψi,kk. (2.4.47)<br />

uj,ji =(φ,jj),i +2ψj,ji +(xjψj,kk),i − 4(1 − ν)ψj,ji<br />

=<br />

<br />

−ϱxjFj<br />

4µ(1 − ν)<br />

= −2(1 − 2ν)ψj,ji.<br />

+2ψj,ji +<br />

,i<br />

<br />

ϱxjFj<br />

4µ(1 − ν)<br />

<br />

,i<br />

− 4(1 − ν)ψj,ji<br />

(2.4.48)<br />

259


260<br />

Finally, from the equations (2.4.47) and (2.4.48) we obtain the desired result that<br />

ui,kk + 1<br />

1 − 2ν uj,ji + ϱFi<br />

µ =0.<br />

Consequently, the equation (2.4.39) is a solution of equation (2.4.38).<br />

As a special case of the above theorem, note that when the body forces are zero, the equations (2.4.40)<br />

become<br />

∇ 2 φ =0 and ∇ 2 ψ = 0.<br />

In this case, we find that equation (2.4.39) is a solution of equation (2.4.38) provided φ and each component of<br />

ψ are harmonic functions. The Papkovich-Neuber potentials are used together with complex variable theory<br />

to solve various two-dimensional elastostatic problems of elasticity. Note also that the Papkovich-Neuber<br />

potentials are not unique as different combinations of φ and ψ can produce the same value for u.<br />

Compatibility Equations<br />

If we know or can derive the displacement field ui,i =1, 2, 3 we can then calculate the components of<br />

the strain tensor<br />

eij = 1<br />

2 (ui,j + uj,i). (2.4.49)<br />

Knowing the strain components, the stress is found using the constitutive relations.<br />

Consider the converse problem where the strain tensor is given or implied due to the assigned stress<br />

field and we are asked to determine the displacement field ui,i=1, 2, 3. Is this a realistic request? Is it even<br />

possible to solve for three displacements given six strain components? It turns out that certain mathematical<br />

restrictions must be placed upon the strain components in order that the inverse problem have a solution.<br />

These mathematical restrictions are known as compatibility equations. That is, we cannot arbitrarily assign<br />

six strain components eij and expect to find a displacement field ui,i=1, 2, 3 with three components which<br />

satisfies the strain relation as given in equation (2.4.49).<br />

EXAMPLE 2.4-3. Suppose we are given the two partial differential equations,<br />

∂u<br />

∂u<br />

= x + y and<br />

∂x ∂y = x3 .<br />

Can we solve for u = u(x, y)? The answer to this question is “no”, because the given equations are inconsistent.<br />

The inconsistency is illustrated if we calculate the mixed second derivatives from each equation. We<br />

find from the first equation that ∂2u ∂x∂y = 1 and from the second equation we calculate ∂2u ∂y∂x =3x2 . These<br />

mixed second partial derivatives are unequal for all x different from √ 3/3. In general, if we have two first<br />

order partial differential equations ∂u<br />

∂u<br />

= f(x, y) and = g(x, y), then for consistency (integrability of<br />

∂x ∂y<br />

the equations) we require that the mixed partial derivatives<br />

∂2u ∂f<br />

=<br />

∂x∂y ∂y = ∂2u ∂g<br />

=<br />

∂y∂x ∂x<br />

be equal to one another for all x and y values over the domain for which the solution is desired. This is an<br />

example of a compatibility equation.


A similar situation occurs in two dimensions for a material in a state of strain where ezz = ezx = ezy =0,<br />

called plane strain. In this case, are we allowed to arbitrarily assign values to the strains exx,eyy and exy and<br />

from these strains determine the displacement field u = u(x, y) andv = v(x, y) inthex− and y−directions?<br />

Let us try to answer this question. Assume a state of plane strain where ezz = ezx = ezy =0. Further, let<br />

us assign 3 arbitrary functional values f,g,h such that<br />

exx = ∂u<br />

∂x = f(x, y), exy = 1<br />

<br />

∂u ∂v<br />

+ = g(x, y),<br />

2 ∂y ∂x<br />

eyy = ∂v<br />

= h(x, y).<br />

∂y<br />

We must now decide whether these equations are consistent. That is, will we be able to solve for the<br />

displacement field u = u(x, y)andv = v(x, y)? To answer this question, let us derive a compatibility equation<br />

(integrability condition). From the given equations we can calculate the following partial derivatives<br />

∂2exx ∂y2 = ∂3u ∂x∂y2 = ∂2f ∂y2 ∂2eyy ∂x2 = ∂3v ∂y∂x2 = ∂2h ∂x2 2 ∂2 exy<br />

∂x∂y = ∂3 u<br />

∂x∂y 2 + ∂3 v<br />

∂y∂x 2 =2 ∂2 g<br />

∂x∂y .<br />

This last equation gives us the compatibility equation<br />

or the functions g, f, h must satisfy the relation<br />

2 ∂2exy ∂x∂y = ∂2exx ∂y2 + ∂2eyy ∂x2 2 ∂2g ∂x∂y = ∂2f ∂y2 + ∂2h .<br />

∂x2 Cartesian Derivation of Compatibility Equations<br />

If the displacement field ui,i=1, 2, 3 is known we can derive the strain and rotation tensors<br />

eij = 1<br />

2 (ui,j + uj,i) and ωij = 1<br />

2 (ui,j − uj,i). (2.4.50)<br />

Now work backwards. Assume the strain and rotation tensors are given and ask the question, “Is it possible<br />

to solve for the displacement field ui,i=1, 2, 3?” If we view the equation (2.4.50) as a system of equations<br />

with unknowns eij,ωij and ui and if by some means we can eliminate the unknowns ωij and ui then we<br />

will be left with equations which must be satisfied by the strains eij. These equations are known as the<br />

compatibility equations and they represent conditions which the strain components must satisfy in order<br />

that a displacement function exist and the equations (2.4.37) are satisfied. Let us see if we can operate upon<br />

the equations (2.4.50) to eliminate the quantities ui and ωij and hence derive the compatibility equations.<br />

Addition of the equations (2.4.50) produces<br />

ui,j = ∂ui<br />

∂xj<br />

= eij + ωij. (2.4.51)<br />

261


262<br />

Differentiate this expression with respect to xk and verify the result<br />

∂ 2 ui<br />

∂xj∂xk<br />

= ∂eij<br />

∂xk<br />

+ ∂ωij<br />

.<br />

∂xk<br />

(2.4.52)<br />

We further assume that the displacement field is continuous so that the mixed partial derivatives are equal<br />

and<br />

∂2ui ∂xj∂xk<br />

= ∂2ui .<br />

∂xk∂xj<br />

(2.4.53)<br />

Interchanging j and k in equation (2.4.52) gives us<br />

∂ 2 ui<br />

∂xk∂xj<br />

= ∂eik<br />

∂xj<br />

+ ∂ωik<br />

.<br />

∂xj<br />

(2.4.54)<br />

Equating the second derivatives from equations (2.4.54) and (2.4.52) and rearranging terms produces the<br />

result<br />

∂eij<br />

∂xk<br />

− ∂eik<br />

∂xj<br />

= ∂ωik<br />

∂xj<br />

− ∂ωij<br />

∂xk<br />

(2.4.55)<br />

form<br />

Making the observation that ωij satisfies ∂ωik<br />

∂eij<br />

∂xj<br />

− ∂ωij<br />

∂xk<br />

= ∂ωjk<br />

∂xi<br />

, the equation (2.4.55) simplifies to the<br />

∂xk<br />

− ∂eik<br />

∂xj<br />

= ∂ωjk<br />

.<br />

∂xi<br />

(2.4.56)<br />

The term involving ωjk can be eliminated by using the mixed partial derivative relation<br />

∂ 2 ωjk<br />

=<br />

∂xi∂xm<br />

∂2ωjk . (2.4.57)<br />

∂xm∂xi<br />

To derive the compatibility equations we differentiate equation (2.4.56) with respect to xm and then<br />

interchanging the indices i and m and substitute the results into equation (2.4.57). This will produce the<br />

compatibility equations<br />

∂2eij +<br />

∂xm∂xk<br />

∂2emk −<br />

∂xi∂xj<br />

∂2eik −<br />

∂xm∂xj<br />

∂2emj =0. (2.4.58)<br />

∂xi∂xk<br />

This is a set of 81 partial differential equations which must be satisfied by the strain components. Fortunately,<br />

due to symmetry considerations only 6 of these 81 equations are distinct. These 6 distinct equations are<br />

known as the St. Venant’s compatibility equations and can be written as<br />

∂ 2 e11<br />

∂x2∂x3<br />

∂ 2 e22<br />

∂x1∂x3<br />

∂ 2 e33<br />

∂x1∂x2<br />

= ∂2e<strong>12</strong> −<br />

∂x1∂x3<br />

∂2e23 ∂x1 2 + ∂2e31 ∂x1∂x2<br />

= ∂2e23 −<br />

∂x2∂x1<br />

∂2e31 ∂x2 2 + ∂2e<strong>12</strong> ∂x2∂x3<br />

= ∂2e31 −<br />

∂x3∂x2<br />

∂2e<strong>12</strong> ∂x3 2 + ∂2e23 ∂x3∂x1<br />

2 ∂2e<strong>12</strong> =<br />

∂x1∂x2<br />

∂2e11 ∂x2 2 + ∂2e22 ∂x1 2<br />

2 ∂2e23 =<br />

∂x2∂x3<br />

∂2e22 ∂x3 2 + ∂2e33 ∂x2 2<br />

2 ∂2e31 =<br />

∂x3∂x1<br />

∂2e33 ∂x1 2 + ∂2e11 .<br />

∂x3<br />

2<br />

Observe that the fourth compatibility equation is the same as that derived in the example 2.4-3.<br />

These compatibility equations can also be expressed in the indicial form<br />

(2.4.59)<br />

eij,km + emk,ji − eik,jm − emj,ki =0. (2.4.60)


Compatibility Equations in Terms of Stress<br />

In the generalized Hooke’s law, equation (2.4.29), we can solve for the strain in terms of stress. This<br />

in turn will give rise to a representation of the compatibility equations in terms of stress. The resulting<br />

equations are known as the Beltrami-Michell equations. Utilizing the strain-stress relation<br />

eij = 1+ν<br />

E σij − ν<br />

E σkkδij<br />

we substitute for the strain in the equations (2.4.60) and rearrange terms to produce the result<br />

σij,km + σmk,ji − σik,jm − σmj,ki =<br />

ν<br />

1+ν [δijσnn,km + δmkσnn,ji − δikσnn,jm − δmjσnn,ki] .<br />

(2.4.61)<br />

Now only 6 of these 81 equations are linearly independent. It can be shown that the 6 linearly independent<br />

equations are equivalent to the equations obtained by setting k = m and summing over the repeated indices.<br />

We then obtain the equations<br />

σij,mm + σmm,ij − (σim,m) ,j − (σmj,m) ,i = ν<br />

1+ν [δijσnn,mm + σnn,ij] .<br />

Employing the equilibrium equation σij,i + ϱbj = 0 the above result can be written in the form<br />

σij,mm + 1<br />

1+ν σkk,ij − ν<br />

1+ν δijσnn,mm = −(ϱbi),j − (ϱbj),i<br />

or<br />

∇ 2 σij + 1<br />

1+ν σkk,ij − ν<br />

1+ν δijσnn,mm = −(ϱbi),j − (ϱbj),i.<br />

This result can be further simplified by observing that a contraction on the indices k and i in equation<br />

(2.4.61) followed by a contraction on the indices m and j produces the result<br />

σij,ij =<br />

1 − ν<br />

1+ν σnn,jj.<br />

Consequently, the Beltrami-Michell equations can be written in the form<br />

∇ 2 σij + 1<br />

1+ν σpp,ij = − ν<br />

1 − ν δij(ϱbk) ,k − (ϱbi) ,j − (ϱbj) ,i. (2.4.62)<br />

Their derivation is left as an exercise. The Beltrami-Michell equations together with the linear momentum<br />

(equilibrium) equations σij,i + ϱbj = 0 represent 9 equations in six unknown stresses. This combinations<br />

of equations is difficult to handle. An easier combination of equations in terms of stress functions will be<br />

developed shortly.<br />

The Navier equations with boundary conditions are difficult to solve in general. Let us take the momentum<br />

equations (2.4.27(a)), the strain relations (2.4.28) and constitutive equations (Hooke’s law) (2.4.29)<br />

and make simplifying assumptions so that a more tractable systems results.<br />

263


264<br />

Plane Strain<br />

The plane strain assumption usually is applied in situations where there is a cylindrical shaped body<br />

whose axis is parallel to the z axis and loads are applied along the z−direction. In any x-y plane we assume<br />

that the surface tractions and body forces are independent of z. We set all strains with a subscript z equal<br />

to zero. Further, all solutions for the stresses, strains and displacements are assumed to be only functions<br />

of x and y and independent of z. Note that in plane strain the stress σzz is different from zero.<br />

In Cartesian coordinates the strain tensor is expressible in terms of its physical components which can<br />

be represented in the matrix form<br />

⎛<br />

⎝ e11<br />

⎞ ⎛<br />

e<strong>12</strong> e13<br />

⎠ = ⎝ exx<br />

⎞<br />

exy exz<br />

⎠ .<br />

e21 e22 e23<br />

e31 e32 e33<br />

eyx eyy eyz<br />

ezx ezy ezz<br />

If we assume that all strains which contain a subscript z are zero and the remaining strain components are<br />

functions of only x and y, we obtain a state of plane strain. For a state of plane strain, the stress components<br />

are obtained from the constitutive equations. The condition of plane strain reduces the constitutive equations<br />

to the form:<br />

exx = 1<br />

E [σxx − ν(σyy + σzz)]<br />

eyy = 1<br />

E [σyy − ν(σzz + σxx)]<br />

0= 1<br />

E [σzz − ν(σxx + σyy)]<br />

exy = eyx = 1+ν<br />

E σxy<br />

ezy = eyz = 1+ν<br />

E σyz =0<br />

ezx = exz = 1+ν<br />

E σxz<br />

E<br />

σxx =<br />

(1 + ν)(1 − 2ν)<br />

=0<br />

[(1 − ν)exx + νeyy]<br />

E<br />

σyy =<br />

(1 + ν)(1 − 2ν) [(1 − ν)eyy + νexx]<br />

E<br />

σzz =<br />

(1 + ν)(1 − 2ν) [ν(eyy + exx)]<br />

σxy = E<br />

1+ν exy<br />

(2.4.63)<br />

σxz =0<br />

σyz =0<br />

where σxx, σyy, σzz, σxy, σxz, σyz are the physical components of the stress. The above constitutive<br />

equations imply that for a state of plane strain we will have<br />

σzz = ν(σxx + σyy)<br />

exx = 1+ν<br />

E [(1 − ν)σxx − νσyy]<br />

eyy = 1+ν<br />

E [(1 − ν)σyy − νσxx]<br />

exy = 1+ν<br />

E σxy.<br />

Also under these conditions the compatibility equations reduce to<br />

∂ 2 exx<br />

∂y 2 + ∂2 eyy<br />

∂x 2 =2∂2 exy<br />

∂x∂y .


Plane Stress<br />

An assumption of plane stress is usually applied to thin flat plates. The plate thinness is assumed to be<br />

in the z−direction and loads are applied perpendicular to z. Under these conditions all stress components<br />

with a subscript z are assumed to be zero. The remaining stress components are then treated as functions<br />

of x and y.<br />

In Cartesian coordinates the stress tensor is expressible in terms of its physical components and can be<br />

represented by the matrix ⎛<br />

⎝ σ11<br />

⎞ ⎛<br />

⎞<br />

σ<strong>12</strong> σ13<br />

⎠ =<br />

⎠ .<br />

σ21 σ22 σ23<br />

σ31 σ32 σ33<br />

⎝ σxx σxy σxz<br />

σyx σyy σyz<br />

σzx σzy σzz<br />

If we assume that all the stresses with a subscript z are zero and the remaining stresses are only functions of<br />

x and y we obtain a state of plane stress. The constitutive equations simplify if we assume a state of plane<br />

stress. These simplified equations are<br />

exx = 1<br />

E σxx − ν<br />

E σyy<br />

eyy = 1<br />

E σyy − ν<br />

E σxx<br />

ezz = − ν<br />

E (σxx + σyy)<br />

exy = 1+ν<br />

E σxy<br />

σxx =<br />

exz =0<br />

E<br />

1 − ν2 [exx + νeyy]<br />

σyy = E<br />

1 − ν2 [eyy + νexx]<br />

σzz =0=(1−ν)ezz + ν(exx + eyy)<br />

σxy = E<br />

1+ν exy<br />

(2.4.64)<br />

σyz =0<br />

eyz =0.<br />

σxz =0<br />

For a state of plane stress the compatibility equations reduce to<br />

and the three additional equations<br />

∂2ezz =0,<br />

∂x2 ∂2exx ∂y2 + ∂2eyy ∂x2 =2∂2 exy<br />

∂x∂y<br />

∂2ezz =0,<br />

∂y2 ∂ 2 ezz<br />

∂x∂y =0.<br />

These three additional equations complicate the plane stress problem.<br />

Airy Stress Function<br />

(2.4.65)<br />

In Cartesian coordinates we examine the equilibrium equations (2.4.25(b)) under the conditions of plane<br />

strain. In terms of physical components we find that these equations reduce to<br />

∂σxx<br />

∂x<br />

∂σxy<br />

+<br />

∂y + ϱbx =0,<br />

∂σyx ∂σyy<br />

+<br />

∂x ∂y + ϱby =0,<br />

∂σzz<br />

∂z =0.<br />

The last equation is satisfied since σzz is a function of x and y. If we further assume that the body forces<br />

are conservative and derivable from a potential function V by the operation ϱ b = −grad V or ϱbi = −V ,i<br />

we can express the above equilibrium equations in the form:<br />

∂σxx<br />

∂x<br />

∂σyx<br />

∂x<br />

+ ∂σxy<br />

∂y<br />

+ ∂σyy<br />

∂y<br />

− ∂V<br />

∂x =0<br />

− ∂V<br />

∂y =0<br />

(2.4.66)<br />

265


266<br />

We will consider these equations together with the compatibility equations (2.4.65). The equations<br />

(2.4.66) will be automatically satisfied if we introduce a scalar function φ = φ(x, y) and assume that the<br />

stresses are derivable from this function and the potential function V according to the rules:<br />

σxx = ∂2φ ∂y2 + V σxy = − ∂2φ σyy =<br />

∂x∂y<br />

∂2φ + V. (2.4.67)<br />

∂x2 The function φ = φ(x, y) is called the Airy stress function after the English astronomer and mathematician<br />

Sir George Airy (1801–1892). Since the equations (2.4.67) satisfy the equilibrium equations we need only<br />

consider the compatibility equation(s).<br />

For a state of plane strain we substitute the relations (2.4.63) into the compatibility equation (2.4.65)<br />

and write the compatibility equation in terms of stresses. We then substitute the relations (2.4.67) and<br />

express the compatibility equation in terms of the Airy stress function φ. These substitutions are left as<br />

exercises. After all these substitutions the compatibility equation, for a state of plane strain, reduces to the<br />

form<br />

∂4φ ∂x4 +2 ∂4φ ∂x2∂y2 + ∂4 2 φ 1 − 2ν ∂ V<br />

+<br />

∂y4 1 − ν ∂x2 + ∂2V ∂y2 <br />

=0. (2.4.68)<br />

In the special case where there are no body forces we have V = 0 and equation (2.4.68) is further simplified<br />

to the biharmonic equation.<br />

∇ 4 φ = ∂4φ ∂x4 +2 ∂4φ ∂x2∂y2 + ∂4φ =0. (2.4.69)<br />

∂y4 In polar coordinates the biharmonic equation is written<br />

∇ 4 φ = ∇ 2 (∇ 2 2 ∂ 1 ∂ 1<br />

φ)= + +<br />

∂r2 r ∂r r2 ∂2 ∂θ2 2 ∂ φ 1 ∂φ 1<br />

+ +<br />

∂r2 r ∂r r2 ∂2φ ∂θ2 <br />

=0.<br />

For conditions of plane stress, we can again introduce an Airy stress function using the equations (2.4.67).<br />

However, an exact solution of the plane stress problem which satisfies all the compatibility equations is<br />

difficult to obtain. By removing the assumptions that σxx,σyy,σxy are independent of z, and neglecting<br />

body forces, it can be shown that for symmetrically distributed external loads the stress function φ can be<br />

represented in the form<br />

φ = ψ − νz2<br />

2(1 + ν) ∇2ψ (2.4.70)<br />

where ψ is a solution of the biharmonic equation ∇4ψ =0. Observe that if z is very small, (the condition<br />

of a thin plate), then equation (2.4.70) gives the approximation φ ≈ ψ. Under these conditions, we obtain<br />

the approximate solution by using only the compatibility equation (2.4.65) together with the stress function<br />

defined by equations (2.4.67) with V =0. Note that the solution we obtain from equation (2.4.69) does not<br />

satisfy all the compatibility equations, however, it does give an excellent first approximation to the solution<br />

in the case where the plate is very thin.<br />

In general, for plane strain or plane stress problems, the equation (2.4.68) or (2.4.69) must be solved for<br />

the Airy stress function φ which is defined over some region R. In addition to specifying a region of the x, y<br />

plane, there are certain boundary conditions which must be satisfied. The boundary conditions specified for<br />

the stress will translate through the equations (2.4.67) to boundary conditions being specified for φ. In the<br />

special case where there are no body forces, both the problems for plane stress and plane strain are governed<br />

by the biharmonic differential equation with appropriate boundary conditions.


EXAMPLE 2.4-4 Assume there exist a state of plane strain with zero body forces. For F11,F<strong>12</strong>,F22<br />

constants, consider the function defined by<br />

φ = φ(x, y) = 1 <br />

F22 x<br />

2<br />

2 − 2F<strong>12</strong> xy + F11 y 2 .<br />

This function is an Airy stress function because it satisfies the biharmonic equation ∇4φ = 0. The resulting<br />

stress field is<br />

σxx = ∂2φ = F11 σyy =<br />

∂y2 ∂2φ = F22 σxy = −<br />

∂x2 ∂2φ = F<strong>12</strong>.<br />

∂x∂y<br />

This example, corresponds to stresses on an infinite flat plate and illustrates a situation where all the stress<br />

components are constants for all values of x and y. In this case, we have σzz = ν(F11+F22). The corresponding<br />

strain field is obtained from the constitutive equations. We find these strains are<br />

exx = 1+ν<br />

E [(1 − ν)F11 − νF22] eyy = 1+ν<br />

E [(1 − ν)F22 − νF11] exy = 1+ν<br />

E F<strong>12</strong>.<br />

The displacement field is found to be<br />

u = u(x, y) = 1+ν<br />

E [(1 − ν)F11<br />

<br />

1+ν<br />

− νF22] x + F<strong>12</strong>y + c1y + c2<br />

E<br />

v = v(x, y) = 1+ν<br />

E [(1 − ν)F22<br />

<br />

1+ν<br />

− νF11] y + F<strong>12</strong>x − c1x + c3,<br />

E<br />

with c1,c2,c3 constants, and is obtained by integrating the strain displacement equations given in Exercise<br />

2.3, problem 2.<br />

EXAMPLE 2.4-5. A special case from the previous example is obtained by setting F22 = F<strong>12</strong> =0.<br />

This is the situation of an infinite plate with only tension in the x−direction. In this special case we have<br />

φ = 1<br />

2 F11y 2 . Changing to polar coordinates we write<br />

φ = φ(r, θ) = F11<br />

2 r2 sin 2 θ = F11<br />

4 r2 (1 − cos 2θ).<br />

The Exercise 2.4, problem 20, suggests we utilize the Airy equations in polar coordinates and calculate the<br />

stresses<br />

σrr = 1 ∂φ 1<br />

+<br />

r ∂r r2 ∂2φ ∂θ2 = F11 cos 2 θ = F11<br />

2<br />

σθθ = ∂2 φ<br />

σrθ = 1<br />

r 2<br />

∂r2 = F11 sin 2 θ = F11<br />

2<br />

∂φ<br />

∂θ<br />

− 1<br />

r<br />

∂2φ ∂r∂θ<br />

= − F11<br />

2<br />

(1 − cos 2θ)<br />

sin 2θ.<br />

(1 + cos 2θ)<br />

267


268<br />

EXAMPLE 2.4-6. We now consider an infinite plate with a circular hole x2 + y2 = a2 which is traction<br />

free. Assume the plate has boundary conditions at infinity defined by σxx = F11, σyy =0, σxy =0. Find<br />

the stress field.<br />

Solution:<br />

The traction boundary condition at r = a is ti = σminm or<br />

t1 = σ11n1 + σ<strong>12</strong>n2 and t2 = σ<strong>12</strong>n1 + σ22n2.<br />

For polar coordinates we have n1 = nr =1,n2 = nθ = 0 and so the traction free boundary conditions at<br />

the surface of the hole are written σrr|r=a =0 and σrθ|r=a =0. The results from the previous example<br />

are used as the boundary conditions at infinity.<br />

Our problem is now to solve for the Airy stress function φ = φ(r, θ) which is a solution of the biharmonic<br />

equation. The previous example 2.4-5 and the form of the boundary conditions at infinity suggests that we<br />

assume a solution to the biharmonic equation which has the form φ = φ(r, θ) =f1(r)+f2(r)cos2θ, where<br />

f1,f2 are unknown functions to be determined. Substituting the assumed solution into the biharmonic<br />

equation produces the equation<br />

<br />

2 d 1 d<br />

+ f<br />

dr2 r dr<br />

′′<br />

1 + 1<br />

r f ′ 2 d 1 d 4<br />

1 + + −<br />

dr2 r dr r2 <br />

f ′′<br />

2 + 1<br />

r f ′ 2 − 4 f2<br />

r2 <br />

cos 2θ =0.<br />

We therefore require that f1,f2 be chosen to satisfy the equations<br />

<br />

2 d 1 d<br />

+ f<br />

dr2 r dr<br />

′′ 1<br />

1 +<br />

r f ′ 2 d 1 d<br />

1 =0 +<br />

dr2 r dr<br />

or r 4 f (iv)<br />

1<br />

+2r3 f ′′′<br />

1 − r2 f ′′<br />

1 + rf ′ 1 =0<br />

r 4 f (iv)<br />

2<br />

4<br />

−<br />

r2 <br />

f ′′ 1<br />

2 +<br />

r f ′ f2<br />

2 − 4<br />

r2 <br />

=0<br />

+2r3 f ′′′<br />

2 − 9r2 f ′′<br />

2 +9rf ′ 2 =0<br />

These equations are Cauchy type equations. Their solutions are obtained by assuming a solution of the form<br />

f1 = r λ and f2 = r m and then solving for the constants λ and m. We find the general solutions of the above<br />

equations are<br />

f1 = c1r 2 ln r + c2r 2 + c3 ln r + c4 and f2 = c5r 2 + c6r 4 + c7<br />

+ c8.<br />

r2 The constants ci,i =1,...,8 are now determined from the boundary conditions. The constant c4 can be<br />

arbitrary since the derivative of a constant is zero. The remaining constants are determined from the stress<br />

conditions. Using the results from Exercise 2.4, problem 20, we calculate the stresses<br />

σrr = c1(1 + 2 ln r)+2c2 + c3<br />

r<br />

σθθ = c1(3 + 2 ln r)+2c2 − c3<br />

r<br />

<br />

σrθ =<br />

2c5 +6c6r 2 − 6 c7 c8<br />

− 2<br />

r4 r2 <br />

− 2c5 +6 2 c7<br />

+4c8<br />

r4 r2 <br />

cos 2θ<br />

<br />

+ 2c5 +<strong>12</strong>c6r 2 2 +6 c7<br />

r4 <br />

cos 2θ<br />

<br />

sin 2θ.


The stresses are to remain bounded for all values of r and consequently we require c1 and c6 to be zero<br />

to avoid infinite stresses for large values of r. The stress σrr|r=a =0requiresthat<br />

2c2 + c3<br />

a2 =0 and 2c5 +6 c7<br />

+4c8 =0.<br />

a4 a2 The stress σrθ|r=a =0requiresthat<br />

2c5 − 6 c7 c8<br />

− 2 =0.<br />

a4 a2 In the limit as r →∞we require that the stresses must satisfy the boundary conditions from the previous<br />

example 2.4-5. This leads to the equations 2c2 = F11<br />

2 and 2c5 = − F11<br />

. Solving the above system of equations<br />

2<br />

produces the Airy stress function<br />

φ = φ(r, θ) = F11<br />

4<br />

and the corresponding stress field is<br />

F11<br />

+<br />

4 r2 − a2<br />

2 F11<br />

<br />

F11a<br />

ln r + c4 +<br />

2<br />

F11<br />

−<br />

2 4 r2 −<br />

σrr = F11<br />

2<br />

σrθ = − F11<br />

σθθ = F11<br />

2<br />

2<br />

<br />

1 − a2<br />

<br />

+ F11<br />

<br />

2<br />

r2 1+3 a4<br />

r r2 <br />

1 − 3 a4<br />

+2a2<br />

r4 r2 <br />

sin 2θ<br />

<br />

1+ a2<br />

r2 <br />

− F11<br />

<br />

1+3<br />

2<br />

a4<br />

r4 <br />

cos 2θ.<br />

4 − 4 a2<br />

<br />

cos 2θ<br />

F11a 4<br />

4r 2<br />

<br />

cos 2θ<br />

There is a maximum stress σθθ =3F11 at θ = π/2, 3π/2 and a minimum stress σθθ = −F11 at θ =0,π.<br />

The effect of the circular hole has been to magnify the applied stress. The factor of 3 is known as a stress<br />

concentration factor. In general, sharp corners and unusually shaped boundaries produce much higher stress<br />

concentration factors than rounded boundaries.<br />

EXAMPLE 2.4-7. Consider an infinite cylindrical tube, with inner radius R1 and the outer radius R0,<br />

which is subjected to an internal pressure P1 andanexternalpressureP0asillustrated in the figure 2.4-7.<br />

Find the stress and displacement fields.<br />

Solution: Let ur,uθ,uz denote the displacement field. We assume that uθ =0anduz = 0 since the<br />

cylindrical surface r equal to a constant does not move in the θ or z directions. The displacement ur = ur(r)<br />

is assumed to depend only upon the radial distance r. Under these conditions the Navier equations become<br />

(λ +2µ) d<br />

<br />

1 d<br />

dr r dr (rur)<br />

<br />

=0.<br />

r c2<br />

This equation has the solution ur = c1 +<br />

2 r<br />

and the strain components are found from the relations<br />

err = dur<br />

dr , eθθ = ur<br />

r , ezz = erθ = erz = ezθ =0.<br />

The stresses are determined from Hooke’s law (the constitutive equations) and we write<br />

σij = λδijΘ+2µeij,<br />

269


270<br />

where<br />

Θ= ∂ur<br />

∂r<br />

is the dilatation. These stresses are found to be<br />

+ ur<br />

r<br />

= 1<br />

r<br />

∂<br />

∂r (rur)<br />

σrr =(λ + µ)c1 − 2µ<br />

r 2 c2 σθθ =(λ + µ)c1 + 2µ<br />

r 2 c2 σzz = λc1 σrθ = σrz = σzθ =0.<br />

We now apply the boundary conditions<br />

<br />

σrr|r=R1nr = −<br />

(λ + µ)c1 − 2µ<br />

R2 c2<br />

1<br />

Solving for the constants c1 and c2 we find<br />

This produces the displacement field<br />

ur =<br />

and stress fields<br />

<br />

<br />

=+P1 and σrr|r=R0nr =<br />

c1 = R2 1P1 − R2 0P0 (λ + µ)(R2 0 − R2 1 ) , c2 = R2 1R2 0 (P1 − P0)<br />

2µ(R2 0 − R2 1<br />

(λ + µ)c1 − 2µ<br />

R2 c2<br />

0<br />

) .<br />

<br />

= −P0.<br />

R2 1P1<br />

2(R2 0 − R2 1 )<br />

<br />

r<br />

λ + µ + R2 <br />

0 R<br />

−<br />

µr<br />

2 0P0<br />

2(R2 0 − R2 1 )<br />

<br />

r<br />

λ + µ + R2 <br />

1<br />

, uθ =0, uz =0,<br />

µr<br />

σrr = R2 1P1 R2 0 − R2 <br />

1 −<br />

1<br />

R2 0<br />

r2 σθθ = R2 1P1 R2 0 − R2 1<br />

σzz =<br />

<br />

− R2 0 P0<br />

<br />

1 − R2 <br />

1<br />

R2 0 − R2 1 r2 <br />

1+ R2 0<br />

r2 <br />

− R2 0P0 R2 0 − R2 <br />

1+<br />

1<br />

R2 1<br />

r2 <br />

λ<br />

λ + µ<br />

<br />

2 R1P1 − R2 0P0 σrz = σzθ = σrθ =0<br />

R 2 0 − R2 1<br />

EXAMPLE 2.4-8. By making simplifying assumptions the Navier equations can be reduced to a more<br />

tractable form. For example, we can reduce the Navier equations to a one dimensional problem by making<br />

the following assumptions<br />

1. Cartesian coordinates x1 = x, x2 = y, x3 = z<br />

2. u1 = u1(x, t), u2 = u3 =0.<br />

3. There are no body forces.<br />

4. Initial conditions of u1(x, 0) = 0 and<br />

∂u1(x, 0)<br />

=0<br />

∂t<br />

5. Boundary conditions of the displacement type u1(0,t)=f(t),<br />

where f(t) is a specified function. These assumptions reduce the Navier equations to the single one dimensional<br />

wave equation<br />

∂2u1 ∂t2 = α2 ∂2u1 ∂x2 , α2 λ +2µ<br />

= .<br />

ρ<br />

The solution of this equation is<br />

<br />

f(t − x/α),<br />

u1(x, t) =<br />

0,<br />

x ≤ αt<br />

x > αt .


The solution represents a longitudinal elastic wave propagating in the x−direction with speed α. The stress<br />

wave associated with this displacement is determined from the constitutive equations. We find<br />

This produces the stress wave<br />

σxx =<br />

σxx =(λ + µ)exx =(λ + µ) ∂u1<br />

∂x .<br />

− (λ+µ)<br />

α f ′ (t − x/α), x ≤ αt<br />

0, x > αt .<br />

Here there is a discontinuity in the stress wave front at x = αt.<br />

Summary of Basic Equations of Elasticity<br />

The equilibrium equations for a continuum have been shown to have the form σ ij<br />

,j + ϱbi =0, where<br />

bi are the body forces per unit mass and σij is the stress tensor. In addition to the above equations we<br />

have the constitutive equations σij = λekkδij +2µeij which is a generalized Hooke’s law relating stress to<br />

strain for a linear elastic isotropic material. The strain tensor is related to the displacement field ui by<br />

the strain equations eij = 1<br />

2 (ui,j + uj,i) . These equations can be combined to obtain the Navier equations<br />

µui,jj +(λ + µ)uj,ji + ϱbi =0.<br />

The above equations must be satisfied at all interior points of the material body. A boundary value<br />

problem results when conditions on the displacement of the boundary are specified. That is, the Navier<br />

equations must be solved subject to the prescribed displacement boundary conditions. If conditions on<br />

the stress at the boundary are specified, then these prescribed stresses are called surface tractions and<br />

must satisfy the relations t i (n) = σ ij nj, where ni is a unit outward normal vector to the boundary. For<br />

surface tractions, we need to use the compatibility equations combined with the constitutive equations and<br />

equilibrium equations. This gives rise to the Beltrami-Michell equations of compatibility<br />

σij,kk + 1<br />

1+ν σkk,ij + ϱ(bi,j + bj,i)+ ν<br />

1 − ν ϱbk,k =0.<br />

Here we must solve for the stress components throughout the continuum where the above equations hold<br />

subject to the surface traction boundary conditions. Note that if an elasticity problem is formed in terms of<br />

the displacement functions, then the compatibility equations can be ignored.<br />

For mixed boundary value problems we must solve a system of equations consisting of the equilibrium<br />

equations, constitutive equations, and strain displacement equations. We must solve these equations subject<br />

to conditions where the displacements ui are prescribed on some portion(s) of the boundary and stresses are<br />

prescribed on the remaining portion(s) of the boundary. Mixed boundary value problems are more difficult<br />

to solve.<br />

For elastodynamic problems, the equilibrium equations are replaced by equations of motion. In this<br />

case we need a set of initial conditions as well as boundary conditions before attempting to solve our basic<br />

system of equations.<br />

271


272<br />

EXERCISE 2.4<br />

◮ 1. Verify the generalized Hooke’s law constitutive equations for hexagonal materials.<br />

In the following problems the Young’s modulus E, Poisson’s ratio ν, the shear modulus or modulus<br />

of rigidity µ (sometimes denoted by G in Engineering texts), Lame’s constant λ and the bulk modulus of<br />

elasticity k are assumed to satisfy the equations (2.4.19), (2.4.24) and (2.4.25). Show that these relations<br />

imply the additional relations given in the problems 2 through 6.<br />

◮ 2.<br />

◮ 3.<br />

◮ 4.<br />

◮ 5.<br />

◮ 6.<br />

µ(3λ +2µ)<br />

E =<br />

µ + λ<br />

λ(1 + ν)(1 − 2ν)<br />

E =<br />

ν<br />

3k − E<br />

ν =<br />

6k<br />

λ<br />

ν =<br />

2(µ + λ)<br />

ν =<br />

ν =<br />

E =<br />

9k(k − λ)<br />

3k − λ<br />

E =2µ(1 + ν)<br />

(E + λ) 2 +8λ 2 − (E + λ)<br />

3k − 2µ<br />

2(µ +3k)<br />

<br />

(E + λ) 2 +8λ2 +(E +3λ)<br />

k =<br />

6<br />

2µ +3λ<br />

k =<br />

3<br />

3(k − λ)<br />

µ =<br />

2<br />

λ(1 − 2ν)<br />

µ =<br />

2ν<br />

λ = 3kν<br />

1+ν<br />

µ(2µ − E)<br />

λ =<br />

E − 3µ<br />

3k(1 − 2ν)<br />

µ =<br />

2(1 + ν)<br />

µ = 3Ek<br />

9k − E<br />

4λ<br />

E<br />

k =<br />

3(1 − 2ν)<br />

µE<br />

k =<br />

3(3µ − E)<br />

µ =<br />

µ =<br />

3k − 2µ<br />

λ =<br />

3<br />

3k(3k − E)<br />

λ =<br />

9k − E<br />

E = 9kµ<br />

µ +3k<br />

E =3(1− 2ν)k<br />

E − 2µ<br />

ν =<br />

2µ<br />

ν =<br />

λ<br />

3k − λ<br />

2µ(1 + ν)<br />

k =<br />

3(1 − 2ν)<br />

λ(1 + ν)<br />

k =<br />

3ν<br />

(E + λ) 2 +8λ 2 +(E − 3λ)<br />

E<br />

2(1 + ν)<br />

4<br />

νE<br />

λ =<br />

(1 + ν)(1 − 2ν)<br />

λ = 2µν<br />

1 − 2ν<br />

◮ 7. The previous exercises 2 through 6 imply that the generalized Hooke’s law<br />

σij =2µeij + λδijekk<br />

is expressible in a variety of forms. From the set of constants (µ,λ,ν,E,k) we can select any two constants<br />

and then express Hooke’s law in terms of these constants.<br />

(a) Express the above Hooke’s law in terms of the constants E and ν.<br />

(b) Express the above Hooke’s law in terms of the constants k and E.<br />

(c) Express the above Hooke’s law in terms of physical components. Hint: The quantity ekk is an invariant<br />

hence all you need to know is how second order tensors are represented in terms of physical components.<br />

See also problems 10,11,<strong>12</strong>.


◮ 8. Verify the equations defining the stress for plane strain in Cartesian coordinates are<br />

E<br />

σxx =<br />

(1 + ν)(1 − 2ν) [(1 − ν)exx + νeyy]<br />

E<br />

σyy =<br />

(1 + ν)(1 − 2ν) [(1 − ν)eyy + νexx]<br />

Eν<br />

σzz =<br />

(1 + ν)(1 − 2ν) [exx + eyy]<br />

σxy = E<br />

1+ν exy<br />

σyz = σxz =0<br />

◮ 9. Verify the equations defining the stress for plane strain in polar coordinates are<br />

E<br />

σrr =<br />

(1 + ν)(1 − 2ν) [(1 − ν)err + νeθθ]<br />

E<br />

σθθ =<br />

(1 + ν)(1 − 2ν) [(1 − ν)eθθ + νerr]<br />

νE<br />

σzz =<br />

(1 + ν)(1 − 2ν) [err + eθθ]<br />

σrθ = E<br />

1+ν erθ<br />

σrz = σθz =0<br />

◮ 10. Write out the independent components of Hooke’s generalized law for strain in terms of stress, and<br />

stress in terms of strain, in Cartesian coordinates. Express your results using the parameters ν and E.<br />

(Assume a linear elastic, homogeneous, isotropic material.)<br />

◮ 11. Write out the independent components of Hooke’s generalized law for strain in terms of stress, and<br />

stress in terms of strain, in cylindrical coordinates. Express your results using the parameters ν and E.<br />

(Assume a linear elastic, homogeneous, isotropic material.)<br />

◮ <strong>12</strong>. Write out the independent components of Hooke’s generalized law for strain in terms of stress, and<br />

stress in terms of strain in spherical coordinates. Express your results using the parameters ν and E. (Assume<br />

a linear elastic, homogeneous, isotropic material.)<br />

◮ 13. For a linear elastic, homogeneous, isotropic material assume there exists a state of plane strain in<br />

Cartesian coordinates. Verify the equilibrium equations are<br />

Hint: See problem 14, Exercise 2.3.<br />

∂σxx<br />

∂x<br />

∂σyx<br />

∂x<br />

+ ∂σxy<br />

∂y + ϱbx =0<br />

+ ∂σyy<br />

∂y + ϱby =0<br />

∂σzz<br />

∂z + ϱbz =0<br />

273


274<br />

◮ 14 . For a linear elastic, homogeneous, isotropic material assume there exists a state of plane strain in<br />

polar coordinates. Verify the equilibrium equations are<br />

Hint: See problem 15, Exercise 2.3.<br />

∂σrr<br />

∂r<br />

1 ∂σrθ 1<br />

+ +<br />

r ∂θ r (σrr − σθθ)+ϱbr =0<br />

∂σrθ<br />

∂r<br />

1 ∂σθθ 2<br />

+ +<br />

r ∂θ r σrθ + ϱbθ =0<br />

∂σzz<br />

∂z + ϱbz =0<br />

◮ 15. For a linear elastic, homogeneous, isotropic material assume there exists a state of plane stress in<br />

Cartesian coordinates. Verify the equilibrium equations are<br />

∂σxx<br />

∂x<br />

∂σyx<br />

∂x<br />

+ ∂σxy<br />

∂y + ϱbx =0<br />

+ ∂σyy<br />

∂y + ϱby =0<br />

◮ 16. Determine the compatibility equations in terms of the Airy stress function φ when there exists a state<br />

of plane stress. Assume the body forces are derivable from a potential function V.<br />

◮ 17. For a linear elastic, homogeneous, isotropic material assume there exists a state of plane stress in<br />

polar coordinates. Verify the equilibrium equations are<br />

∂σrr<br />

∂r<br />

1 ∂σrθ 1<br />

+ +<br />

r ∂θ r (σrr − σθθ)+ϱbr =0<br />

∂σrθ<br />

∂r<br />

1 ∂σθθ 2<br />

+ +<br />

r ∂θ r σrθ + ϱbθ =0


◮ 18. Figure 2.4-4 illustrates the state of equilibrium on an element in polar coordinates assumed to be of<br />

unit length in the z-direction. Verify the stresses given in the figure and then sum the forces in the r and θ<br />

directions to derive the same equilibrium laws developed in the previous exercise.<br />

Figure 2.4-4. Polar element in equilibrium.<br />

Hint: Resolve the stresses into components in the r and θ directions. Use the results that sin dθ<br />

2<br />

cos dθ<br />

2<br />

≈ dθ<br />

2 and<br />

≈ 1 for small values of dθ. Sum forces and then divide by rdr dθ and take the limit as dr → 0and<br />

dθ → 0.<br />

◮ 19. Express each of the physical components of plane stress in polar coordinates, σrr, σθθ, andσrθ<br />

in terms of the physical components of stress in Cartesian coordinates σxx, σyy, σxy. Hint: Consider the<br />

∂x<br />

transformation law σij = σab<br />

a<br />

∂xi ∂xb j .<br />

∂x<br />

◮ 20. Use the results from problem 19 and assume the stresses are derivable from the relations<br />

σxx = V + ∂2φ ∂y2 , σxy = − ∂2φ ∂x∂y , σyy = V + ∂2φ ∂x2 where V is a potential function and φ is the Airy stress function. Show that upon changing to polar<br />

coordinates the Airy equations for stress become<br />

σrr = V + 1 ∂φ 1<br />

+<br />

r ∂r r2 ∂2φ ∂θ2 , σrθ = 1<br />

r2 ∂φ<br />

∂θ<br />

− 1<br />

r<br />

∂2φ ∂r∂θ , σθθ = V + ∂2φ .<br />

∂r2 ◮ 21. Verify that the Airy stress equations in polar coordinates, given in problem 20, satisfy the equilibrium<br />

equations in polar coordinates derived in problem 17.<br />

275


276<br />

◮ 22. In Cartesian coordinates show that the traction boundary conditions, equations (2.3.11), can be<br />

written in terms of the constants λ and µ as<br />

<br />

<br />

∂u1 ∂u1 ∂u2<br />

T1 = λn1ekk + µ 2n1 + n2 +<br />

∂x1 ∂x2 ∂x1 <br />

∂u1 ∂u3<br />

+ n3 +<br />

∂x3 ∂x1 <br />

<br />

∂u2 ∂u1<br />

T2 = λn2ekk + µ n1 +<br />

∂x1 ∂x2 <br />

<br />

∂u2 ∂u2 ∂u3<br />

+2n2 + n3 +<br />

∂x2 ∂x3 ∂x2 <br />

<br />

∂u3 ∂u1<br />

T3 = λn3ekk + µ n1 +<br />

∂x1 ∂x3 <br />

∂u3 ∂u2<br />

+ n2 +<br />

∂x2 ∂x3 <br />

∂u3<br />

+2n3<br />

∂x3 <br />

where (n1,n2,n3) are the direction cosines of the unit normal to the surface, u1,u2,u3 are the components<br />

of the displacements and T1,T2,T3 are the surface tractions.<br />

◮ 23. Consider an infinite plane subject to tension in the x−direction only. Assume a state of plane strain<br />

and let σxx = T with σxy = σyy =0. Find the strain components exx, eyy and exy. Also find the displacement<br />

field u = u(x, y) andv = v(x, y).<br />

◮ 24. Consider an infinite plane subject to tension in the y-direction only. Assume a state of plane strain<br />

and let σyy = T with σxx = σxy =0. Find the strain components exx, eyy and exy. Also find the displacement<br />

field u = u(x, y) andv = v(x, y).<br />

◮ 25. Consider an infinite plane subject to tension in both the x and y directions. Assume a state of plane<br />

strain and let σxx = T , σyy = T and σxy =0. Find the strain components exx,eyy and exy. Also find the<br />

displacement field u = u(x, y) andv = v(x, y).<br />

◮ 26. An infinite cylindrical rod of radius R0 has an external pressure P0 as illustrated in figure 2.5-5. Find<br />

the stress and displacement fields.<br />

Figure 2.4-5. External pressure on a rod.


Figure 2.4-6. Internal pressure on circular hole.<br />

Figure 2.4-7. Tube with internal and external pressure.<br />

◮ 27. An infinite plane has a circular hole of radius R1 with an internal pressure P1 as illustrated in the<br />

figure 2.4-6. Find the stress and displacement fields.<br />

◮ 28. A tube of inner radius R1 and outer radius R0 has an internal pressure of P1 andanexternalpressure<br />

of P0 as illustrated in the figure 2.4-7. Verify the stress and displacement fields derived in example 2.4-7.<br />

◮ 29. Use Cartesian tensors and combine the equations of equilibrium σij,j + ϱbi =0, Hooke’s law σij =<br />

λekkδij +2µeij and the strain tensor eij = 1<br />

2 (ui,j + uj,i) and derive the Navier equations of equilibrium<br />

where Θ = e11 + e22 + e33 is the dilatation.<br />

σij,j + ϱbi =(λ + µ) ∂Θ<br />

∂x i + µ ∂2 ui<br />

∂x k ∂x k + ϱbi =0,<br />

◮ 30. Show the Navier equations in problem 29 can be written in the tensor form<br />

or the vector form<br />

µui,jj +(λ + µ)uj,ji + ϱbi =0<br />

µ∇ 2 u +(λ + µ)∇ (∇·u)+ϱ b = 0.<br />

277


278<br />

◮ 31. Show that in an orthogonal coordinate system the components of ∇(∇·u) can be expressed in terms<br />

of physical components by the relation<br />

[∇ (∇·u)] i = 1 ∂<br />

∂xi <br />

1<br />

<br />

∂(h2h3u(1))<br />

∂x1 + ∂(h1h3u(2))<br />

∂x2 + ∂(h1h2u(3))<br />

∂x3 <br />

hi<br />

h1h2h3<br />

◮ 32. Show that in orthogonal coordinates the components of ∇ 2 u can be written<br />

∇ 2 u <br />

i = gjk ui,jk = Ai<br />

and in terms of physical components one can write<br />

3 1<br />

hiA(i) =<br />

h<br />

j=1<br />

2 ⎡<br />

⎣<br />

j<br />

∂2 (hiu(i))<br />

∂xj∂xj 3<br />

<br />

m ∂(hmu(m))<br />

− 2<br />

ij ∂x<br />

m=1<br />

j<br />

3<br />

<br />

m ∂(hiu(i))<br />

−<br />

jj ∂x<br />

m=1<br />

m<br />

3<br />

<br />

∂<br />

− hmu(m)<br />

∂xj 3<br />

3<br />

⎤<br />

m m p<br />

m p<br />

−<br />

−<br />

⎦<br />

ij ip jj jp ij<br />

m=1<br />

◮ 33. Use the results in problem 32 to show in Cartesian coordinates the physical components of [∇ 2 u]i = Ai<br />

can be represented<br />

p=1<br />

2<br />

∇ u · ê1 = A(1) = ∂2u ∂x2 + ∂2u ∂y2 + ∂2u ∂z2 2<br />

∇ u · ê2 = A(2) = ∂2v ∂x2 + ∂2v ∂y2 + ∂2v ∂z2 2<br />

∇ u · ê3 = A(3) = ∂2w ∂x2 + ∂2w ∂y2 + ∂2w ∂z2 where (u, v, w) are the components of the displacement vector u.<br />

◮ 34. Use the results in problem 32 to show in cylindrical coordinates the physical components of [∇ 2 u]i = Ai<br />

can be represented<br />

p=1<br />

2<br />

∇ u · êr = A(1) = ∇ 2 ur − 1<br />

r2 ur − 2<br />

r2 ∂uθ<br />

∂θ<br />

2<br />

∇ u · êθ = A(2) = ∇ 2 uθ + 2<br />

r2 ∂ur 1<br />

− uθ<br />

∂θ r2 2<br />

∇ u · êz = A(3) = ∇ 2 uz<br />

∂α 1<br />

+<br />

∂r r2 ∂2α ∂θ2 + ∂2α ∂z2 where ur,uθ,uz are the physical components of u and ∇ 2 α = ∂2α 1<br />

+<br />

∂r2 r<br />

◮ 35. Use the results in problem 32 to show in spherical coordinates the physical components of [∇2u]i = Ai<br />

can be represented<br />

2<br />

∇ u · êρ = A(1) = ∇ 2 uρ − 2<br />

ρ2 uρ − 2<br />

ρ2 ∂uθ<br />

∂θ<br />

2<br />

∇ u · êθ = A(2) = ∇ 2 uθ + 2<br />

ρ2 ∂uρ<br />

∂θ −<br />

1<br />

ρ2 sin θ uθ − 2cosθ<br />

ρ2 sin 2 θ<br />

2<br />

∇ u · êφ = A(3) = ∇ 2 uφ −<br />

1<br />

ρ 2 sin 2 θ uφ +<br />

where uρ,uθ,uφ are the physical components of u and where<br />

∇ 2 α = ∂2α 2 ∂α 1<br />

+ +<br />

∂ρ2 ρ ∂ρ ρ2 ∂2α cot θ<br />

+<br />

∂θ2 ρ2 ∂α<br />

∂θ +<br />

2cotθ<br />

−<br />

ρ2 uθ<br />

2<br />

−<br />

ρ2 ∂uφ<br />

sin θ ∂φ<br />

∂uθ<br />

∂φ<br />

2<br />

ρ2 ∂uρ 2cosθ<br />

+<br />

sin θ ∂φ ρ2 sin 2 ∂uθ<br />

θ ∂φ<br />

1<br />

ρ2 sin 2 ∂<br />

θ<br />

2α ∂φ2


◮ 36. Combine the results from problems 30,31,32 and 33 and write the Navier equations of equilibrium<br />

in Cartesian coordinates. Alternatively, write the stress-strain relations (2.4.29(b)) in terms of physical<br />

components and then use these results, together with the results from Exercise 2.3, problems 2 and 14, to<br />

derive the Navier equations.<br />

◮ 37. Combine the results from problems 30,31,32 and 34 and write the Navier equations of equilibrium<br />

in cylindrical coordinates. Alternatively, write the stress-strain relations (2.4.29(b)) in terms of physical<br />

components and then use these results, together with the results from Exercise 2.3, problems 3 and 15, to<br />

derive the Navier equations.<br />

◮ 38. Combine the results from problems 30,31,32 and 35 and write the Navier equations of equilibrium<br />

in spherical coordinates. Alternatively, write the stress-strain relations (2.4.29(b)) in terms of physical<br />

components and then use these results, together with the results from Exercise 2.3, problems 4 and 16, to<br />

derive the Navier equations.<br />

◮ 39. Assume ϱ b = −grad V and let φ denote the Airy stress function defined by<br />

σxx =V + ∂2 φ<br />

∂y 2<br />

σyy =V + ∂2 φ<br />

∂x 2<br />

σxy = − ∂2φ ∂x∂y<br />

(a) Show that for conditions of plane strain the equilibrium equations in two dimensions are satisfied by the<br />

above definitions. (b) Express the compatibility equation<br />

in terms of φ and V and show that<br />

∂2exx ∂y2 + ∂2eyy ∂x2 =2∂2 exy<br />

∂x∂y<br />

∇ 4 1 − 2ν<br />

φ +<br />

1 − ν ∇2V =0.<br />

◮ 40. Consider the case where the body forces are conservative and derivable from a scalar potential function<br />

such that ϱbi = −V,i. Show that under conditions of plane strain in rectangular Cartesian coordinates the<br />

compatibility equation e11,22 + e22,11 =2e<strong>12</strong>,<strong>12</strong> can be reduced to the form ∇ 2 σii = 1<br />

1 − ν ∇2 V ,i =1, 2<br />

involving the stresses and the potential. Hint: Differentiate the equilibrium equations.<br />

◮ 41. Use the relation σ i j =2µe i j + λe m mδ i j and solve for the strain in terms of the stress.<br />

◮ 42. Derive the equation (2.4.26) from the equation (2.4.23).<br />

◮ 43. In two dimensions assume that the body forces are derivable from a potential function V and<br />

ϱbi = −gijV ,j. Also assume that the stress is derivable from the Airy stress function and the potential<br />

function by employing the relations σ ij = ɛ im ɛ jn um,n + g ij V i,j,m,n =1, 2whereum = φ ,m and<br />

ɛ pq is the two dimensional epsilon permutation symbol and all indices have the range 1,2.<br />

(a) Show that ɛimɛjn (φm) ,nj =0.<br />

(b) Show that σ ij<br />

,j = −ϱbi .<br />

(c) Verify the stress laws for cylindrical and Cartesian coordinates given in problem 20 by using the above<br />

expression for σ ij . Hint: Expand the contravariant derivative and convert all terms to physical compo-<br />

nents. Also recall that ɛ ij = 1<br />

√ g e ij .<br />

279


280<br />

◮ 44. Consider a material with body forces per unit volume ρF i ,i=1, 2, 3 and surface tractions denoted by<br />

σ r = σ rj nj, where nj is a unit surface normal. Further, let δui denote a small displacement vector associated<br />

with a small variation in the strain δeij.<br />

<br />

(a) Show the work done during a small variation in strain is δW = δWB + δWS where δWB = ρF<br />

V<br />

i <br />

δui dτ<br />

is a volume integral representing the work done by the body forces and δWS = σ<br />

S<br />

r δur dS is a surface<br />

integral representing the work done by the surface forces.<br />

(b) Using the Gauss divergence theorem show that the work done can be represented as<br />

δW = 1<br />

<br />

2<br />

c ijmn δ[emneij] dτ or W = 1<br />

<br />

2<br />

σ ij eij dτ.<br />

V<br />

The scalar quantity 1<br />

2 σij eij is called the strain energy density or strain energy per unit volume.<br />

Hint: Interchange subscripts, add terms and calculate 2W = <br />

V σij [δui,j + δuj,i] dτ.<br />

◮ 45. Consider a spherical shell subjected to an internal pressure pi and external pressure po. Letadenote the inner radius and b the outer radius of the spherical shell. Find the displacement and stress fields in<br />

spherical coordinates (ρ, θ, φ).<br />

Hint: Assume symmetry in the θ and φ directions and let the physical components of displacements satisfy<br />

the relations uρ = uρ(ρ), uθ = uφ =0.<br />

◮ 46. (a) Verify the average normal stress is proportional to the dilatation, where the proportionality<br />

constant is the bulk modulus of elasticity. i.e. Show that 1<br />

3σi E<br />

i = 1−2ν<br />

V<br />

1<br />

3eii = keii where k is the bulk modulus<br />

of elasticity.<br />

(b) Define the quantities of strain deviation and stress deviation in terms of the average normal stress<br />

s = 1<br />

3σi 1<br />

i and average cubic dilatation e = 3eii as follows<br />

strain deviator ε i j = eij − eδi j<br />

stress deviator s i j = σ i j − sδ i j<br />

Show that zero results when a contraction is performed on the stress and strain deviators. (The above<br />

definitions are used to split the strain tensor into two parts. One part represents pure dilatation and<br />

the other part represents pure distortion.)<br />

(c) Show that (1 − 2ν)s = Ee or s =(3λ +2µ)e<br />

(d) Express Hooke’s law in terms of the strain and stress deviator and show<br />

which simplifies to s i j =2µε i j.<br />

E(ε i j + eδ i j)=(1+ν)s i j +(1− 2ν)sδ i j<br />

◮ 47. Show the strain energy density (problem 44) can be written in terms of the stress and strain deviators<br />

(problem 46) and<br />

W = 1<br />

<br />

σ<br />

2 V<br />

ij eij dτ = 1<br />

<br />

(3se + s<br />

2 V<br />

ij εij) dτ<br />

and from Hooke’s law<br />

W = 3<br />

<br />

((3λ +2µ)e<br />

2 V<br />

2 + 2µ<br />

3 εijεij) dτ.


◮ 48. Find the stress σrr,σrθ and σθθ in an infinite plate with a small circular hole, which is traction free,<br />

when the plate is subjected to a pure shearing force F<strong>12</strong>. Determine the maximum stress.<br />

◮ 49. Show that in terms of E and ν<br />

C1111 =<br />

E(1 − ν)<br />

(1 + ν)(1 − 2ν)<br />

C1<strong>12</strong>2 =<br />

◮ 50. Show that in Cartesian coordinates the quantity<br />

Eν<br />

(1 + ν)(1 − 2ν)<br />

C<strong>12</strong><strong>12</strong> =<br />

S = σxxσyy + σyyσzz + σzzσxx − (σxy) 2 − (σyz) 2 − (σxz) 2<br />

E<br />

2(1 + ν)<br />

is a stress invariant. Hint: First verify that in tensor form S = 1<br />

2 (σiiσjj − σijσij).<br />

◮ 51. Show that in Cartesian coordinates for a state of plane strain where the displacements are given by<br />

u = u(x, y),v = v(x, y) andw = 0, the stress components must satisfy the equations<br />

∂σxx<br />

∂x<br />

∂σyx<br />

∂x<br />

+ ∂σxy<br />

∂y + ϱbx =0<br />

+ ∂σyy<br />

∂y + ϱby =0<br />

∇ 2 (σxx + σyy) = −ϱ<br />

1 − ν<br />

∂bx<br />

∂x<br />

<br />

∂by<br />

+<br />

∂y<br />

◮ 52. Show that in Cartesian coordinates for a state of plane stress where σxx = σxx(x, y), σyy = σyy(x, y),<br />

σxy = σxy(x, y) andσxz = σyz = σzz = 0 the stress components must satisfy<br />

∂σxx<br />

∂x<br />

∂σyx<br />

∂x<br />

+ ∂σxy<br />

∂y + ϱbx =0<br />

+ ∂σyy<br />

∂y + ϱby =0<br />

∇ 2 (σxx + σyy) =− ϱ(ν +1)<br />

∂bx<br />

∂x<br />

<br />

∂by<br />

+<br />

∂y<br />

281


282<br />

§2.5 CONTINUUM MECHANICS (FLUIDS)<br />

Let us consider a fluid medium and use Cartesian tensors to derive the mathematical equations that<br />

describe how a fluid behaves. A fluid continuum, like a solid continuum, is characterized by equations<br />

describing:<br />

1. Conservation of linear momentum<br />

2. Conservation of angular momentum σij = σji.<br />

3. Conservation of mass (continuity equation)<br />

∂ϱ ∂ϱ<br />

+ vi + ϱ<br />

∂t ∂xi<br />

∂vi<br />

∂xi<br />

σij,j + ϱbi = ϱ ˙vi<br />

(2.5.1)<br />

=0 or Dϱ<br />

Dt + ϱ∇· V =0. (2.5.2)<br />

In the above equations vi,i=1, 2, 3 is a velocity field, ϱ is the density of the fluid, σij is the stress tensor<br />

and bj is an external force per unit mass. In the cgs system of units of measurement, the above quantities<br />

have dimensions<br />

[˙vj] =cm/sec 2 , [bj] =dynes/g, [σij] =dyne/cm 2 , [ϱ] =g/cm 3 . (2.5.3)<br />

The displacement field ui,i =1, 2, 3 can be represented in terms of the velocity field vi,i =1, 2, 3, by<br />

the relation<br />

t<br />

ui = vi dt. (2.5.4)<br />

The strain tensor components of the medium can then be represented in terms of the velocity field as<br />

where<br />

eij = 1<br />

2 (ui,j + uj,i) =<br />

t<br />

0<br />

0<br />

1<br />

2 (vi,j + vj,i) dt =<br />

t<br />

0<br />

Dij dt, (2.5.5)<br />

Dij = 1<br />

2 (vi,j + vj,i) (2.5.6)<br />

is called the rate of deformation tensor , velocity strain tensor, orrate of strain tensor.<br />

Note the difference in the equations describing a solid continuum compared with those for a fluid<br />

continuum. In describing a solid continuum we were primarily interested in calculating the displacement<br />

field ui,i =1, 2, 3 when the continuum was subjected to external forces. In describing a fluid medium, we<br />

calculate the velocity field vi,i =1, 2, 3 when the continuum is subjected to external forces. We therefore<br />

replace the strain tensor relations by the velocity strain tensor relations in all future considerations concerning<br />

the study of fluid motion.<br />

Constitutive Equations for Fluids<br />

In addition to the above basic equations, we will need a set of constitutive equations which describe the<br />

material properties of the fluid. Toward this purpose consider an arbitrary point within the fluid medium<br />

and pass an imaginary plane through the point. The orientation of the plane is determined by a unit normal<br />

ni , i =1, 2, 3 to the planar surface. For a fluid at rest we wish to determine the stress vector t (n)<br />

i acting<br />

on the plane element passing through the selected point P. We desire to express t (n)<br />

i in terms of the stress<br />

tensor σij. The superscript (n) on the stress vector is to remind you that the stress acting on the planar<br />

element depends upon the orientation of the plane through the point.


is colinear with the normal vector to the surface passing through<br />

the selected point. It is also assumed that for fluid elements at rest, there are no shear forces acting on the<br />

planar element through an arbitrary point and therefore the stress tensor σij should be independent of the<br />

orientation of the plane. That is, we desire for the stress vector σij to be an isotropic tensor. This requires<br />

We make the assumption that t (n)<br />

i<br />

σij to have a specific form. To find this specific form we let σij denote the stress components in a general<br />

coordinate system x i , i =1, 2, 3andletσij denote the components of stress in a barred coordinate system<br />

x i ,i=1, 2, 3. Since σij is a tensor, it must satisfy the transformation law<br />

σmn = σij<br />

∂xi ∂xm ∂xj n , i,j,m,n=1, 2, 3. (2.5.7)<br />

∂x<br />

We desire for the stress tensor σij to be an invariant under an arbitrary rotation of axes. Consider<br />

therefore the special coordinate transformations illustrated in the figures 2.5-1(a) and (b).<br />

Figure 2.5-1. Coordinate transformations due to rotations<br />

For the transformation equations given in figure 2.5-1(a), the stress tensor in the barred system of<br />

coordinates is<br />

σ11 = σ22 σ21 = σ32 σ31 = σ<strong>12</strong><br />

σ<strong>12</strong> = σ23 σ22 = σ33 σ32 = σ13<br />

σ13 = σ21 σ23 = σ31 σ33 = σ11.<br />

(2.5.8)<br />

If σij is to be isotropic, we desire that σ11 = σ11, σ22 = σ22 and σ33 = σ33. If the equations (2.5.8) are<br />

to produce these results, we require that σ11, σ22 and σ33 must be equal. We denote these common values<br />

by (−p). In particular, the equations (2.5.8) show that if σ11 = σ11, σ22 = σ22 and σ33 = σ33, then we must<br />

require that σ11 = σ22 = σ33 = −p. If σ<strong>12</strong> = σ<strong>12</strong> and σ23 = σ23, then we also require that σ<strong>12</strong> = σ23 = σ31.<br />

We note that if σ13 = σ13 and σ32 = σ32, then we require that σ21 = σ32 = σ13. If the equations (2.5.7) are<br />

expanded using the transformation given in figure 2.5-1(b), we obtain the additional requirements that<br />

σ11 = σ22 σ21 = −σ<strong>12</strong> σ31 = σ32<br />

σ<strong>12</strong> = −σ21 σ22 = σ11 σ32 = −σ31<br />

σ13 = σ23 σ23 = −σ13 σ33 = σ33.<br />

(2.5.9)<br />

283


284<br />

Analysis of these equations implies that if σij is to be isotropic, then σ21 = σ21 = −σ<strong>12</strong> = −σ21<br />

or σ21 = 0 which implies σ<strong>12</strong> = σ23 = σ31 = σ21 = σ32 = σ13 =0. (2.5.10)<br />

The above analysis demonstrates that if the stress tensor σij is to be isotropic, it must have the form<br />

Use the traction condition (2.3.11), and express the stress vector as<br />

σij = −pδij. (2.5.11)<br />

t (n)<br />

j = σijni = −pnj. (2.5.<strong>12</strong>)<br />

This equation is interpreted as representing the stress vector at a point on a surface with outward unit<br />

normal ni, wherepis the pressure (hydrostatic pressure) stress magnitude assumed to be positive. The<br />

negative sign in equation (2.5.<strong>12</strong>) denotes a compressive stress.<br />

Imagine a submerged object in a fluid medium. We further imagine the object to be covered with unit<br />

normal vectors emanating from each point on its surface. The equation (2.5.<strong>12</strong>) shows that the hydrostatic<br />

pressure always acts on the object in a compressive manner. A force results from the stress vector acting on<br />

the object. The direction of the force is opposite to the direction of the unit outward normal vectors. It is<br />

a compressive force at each point on the surface of the object.<br />

The above considerations were for a fluid at rest (hydrostatics). For a fluid in motion (hydrodynamics)<br />

a different set of assumptions must be made. Hydrodynamical experiments show that the shear stress<br />

components are not zero and so we assume a stress tensor having the form<br />

σij = −pδij + τij, i,j =1, 2, 3, (2.5.13)<br />

where τij is called the viscous stress tensor. Note that all real fluids are both viscous and compressible.<br />

Definition: (Viscous/inviscid fluid) If the viscous stress tensor<br />

τij is zero for all i, j, then the fluid is called an inviscid, nonviscous,<br />

ideal or perfect fluid. The fluid is called viscous when τij<br />

is different from zero.<br />

In these notes it is assumed that the equation (2.5.13) represents the basic form for constitutive equations<br />

describing fluid motion.


Viscosity<br />

Figure 2.5-2. Viscosity experiment.<br />

Most fluids are characterized by the fact that they cannot resist shearing stresses. That is, if you put a<br />

shearing stress on the fluid, the fluid gives way and flows. Consider the experiment illustrated in the figure<br />

2.5-2 which illustrates a fluid moving between two parallel plane surfaces. Let S denote the distance between<br />

the two planes. Now keep the lower surface fixed or stationary and move the upper surface parallel to the<br />

lower surface with a constant velocity V0. If you measure the force F required to maintain the constant<br />

velocity of the upper surface, you discover that the force F varies directly as the area A of the surface and<br />

the ratio V0/S. This is expressed in the form<br />

F<br />

A<br />

V0<br />

= µ∗ . (2.5.14)<br />

S<br />

The constant µ ∗ is a proportionality constant called the coefficient of viscosity. The viscosity usually depends<br />

upon temperature, but throughout our discussions we will assume the temperature is constant. A dimensional<br />

analysis of the equation (2.5.14) implies that the basic dimension of the viscosity is [µ ∗ ]=ML −1 T −1 . For<br />

example, [µ ∗ ]=gm/(cmsec) in the cgs system of units. The viscosity is usually measured in units of<br />

centipoise where one centipoise represents one-hundredth of a poise, where the unit of 1poise= 1gram<br />

per centimeter per second. The result of the above experiment shows that the stress is proportional to the<br />

change in velocity with change in distance or gradient of the velocity.<br />

Linear Viscous Fluids<br />

The above experiment with viscosity suggest that the viscous stress tensor τij is dependent upon both<br />

the gradient of the fluid velocity and the density of the fluid.<br />

In Cartesian coordinates, the simplest model suggested by the above experiment is that the viscous<br />

stress tensor τij is proportional to the velocity gradient vi,j and so we write<br />

τik = cikmpvm,p, (2.5.15)<br />

where cikmp is a proportionality constant which is dependent upon the fluid density.<br />

The viscous stress tensor must be independent of any reference frame, and hence we assume that the<br />

proportionality constants cikmp can be represented by an isotropic tensor. Recall that an isotropic tensor<br />

has the basic form<br />

cikmp = λ ∗ δikδmp + µ ∗ (δimδkp + δipδkm)+ν ∗ (δimδkp − δipδkm) (2.5.16)<br />

285


286<br />

where λ ∗ ,µ ∗ and ν ∗ are constants. Examining the results from equations (2.5.11) and (2.5.13) we find that if<br />

the viscous stress is symmetric, then τij = τji. This requires ν∗ be chosen as zero. Consequently, the viscous<br />

stress tensor reduces to the form<br />

τik = λ ∗ δikvp,p + µ ∗ (vk,i + vi,k). (2.5.17)<br />

The coefficient µ ∗ is called the first coefficient of viscosity and the coefficient λ ∗ is called the second coefficient<br />

of viscosity. Sometimes it is convenient to define<br />

ζ = λ ∗ + 2<br />

3 µ∗<br />

(2.5.18)<br />

as “another second coefficient of viscosity,” or “bulk coefficient of viscosity.” The condition of zero bulk<br />

viscosity is known as Stokes hypothesis. Many fluids problems assume the Stoke’s hypothesis. This requires<br />

that the bulk coefficient be zero or very small. Under these circumstances the second coefficient of viscosity<br />

is related to the first coefficient of viscosity by the relation λ ∗ = − 2<br />

3 µ∗ . In the study of shock waves and<br />

acoustic waves the Stoke’s hypothesis is not applicable.<br />

There are many tables and empirical formulas where the viscosity of different types of fluids or gases<br />

can be obtained. For example, in the study of the kinetic theory of gases the viscosity can be calculated<br />

from the Sutherland formula µ ∗ 3/2 C1gT<br />

= where C1,C2 are constants for a specific gas. These constants<br />

T + C2<br />

can be found in certain tables. The quantity g is the gravitational constant and T is the temperature in<br />

degrees Rankine ( o R = 460 + o F ). Many other empirical formulas like the above exist. Also many graphs<br />

and tabular values of viscosity can be found. The table 5.1lists the approximate values of the viscosity of<br />

some selected fluids and gases.<br />

Table 5.1<br />

Viscosity of selected fluids and gases<br />

in units of gram<br />

cm−sec =Poise<br />

at Atmospheric Pressure.<br />

Substance 0 ◦ C 20 ◦ C 60 ◦ C 100 ◦ C<br />

Water 0.01798 0.01002 0.00469 0.00284<br />

Alcohol 0.01773<br />

Ethyl Alcohol 0.0<strong>12</strong> 0.00592<br />

Glycol 0.199 0.0495 0.0199<br />

Mercury 0.017 0.0157 0.013 0.0100<br />

Air 1.708(10 −4 ) 2.175(10 −4 )<br />

Helium 1.86(10 −4 ) 1.94(10 −4 ) 2.28(10 −4 )<br />

Nitrogen 1.658(10 −4 ) 1.74(10 −4 ) 1.92(10 −4 ) 2.09(10 −4 )<br />

The viscous stress tensor given in equation (2.5.17) may also be expressed in terms of the rate of<br />

deformation tensor defined by equation (2.5.6). This representation is<br />

τij = λ ∗ δijDkk +2µ ∗ Dij, (2.5.19)<br />

where 2Dij = vi,j + vj,i and Dkk = D11 + D22 + D33 = v1,1 + v2,2 + v3,3 = vi,i = Θ is the rate of change<br />

of the dilatation considered earlier. In Cartesian form, with velocity components u, v, w, the viscous stress


tensor components are<br />

τxx =(λ ∗ +2µ ∗ ) ∂u<br />

<br />

∂v ∂w<br />

+ λ∗ +<br />

∂x ∂y ∂z<br />

τyy =(λ ∗ +2µ ∗ ) ∂v<br />

+ λ∗<br />

∂y<br />

τzz =(λ ∗ +2µ ∗ ) ∂w<br />

∂z<br />

+ λ∗<br />

<br />

∂u ∂w<br />

+<br />

∂x ∂z<br />

<br />

∂u ∂v<br />

+<br />

∂x ∂y<br />

τyx = τxy =µ ∗<br />

<br />

∂u ∂v<br />

+<br />

∂y ∂x<br />

τzx = τxz =µ ∗<br />

<br />

∂w ∂u<br />

+<br />

∂x ∂z<br />

τzy = τyz =µ ∗<br />

<br />

∂v ∂w<br />

+<br />

∂z ∂y<br />

In cylindrical form, with velocity components vr,vθ,vz, the viscous stess tensor components are<br />

∗ ∂vr<br />

τrr =2µ<br />

where ∇· V = 1<br />

r<br />

∂r + λ∗∇· V<br />

τθθ =2µ ∗<br />

<br />

1 ∂vθ vr<br />

+ + λ<br />

r ∂θ r<br />

∗ ∇·V<br />

∗ ∂vz<br />

τzz =2µ<br />

∂z + λ∗∇· V<br />

∂<br />

∂r<br />

1 ∂vθ ∂vz<br />

(rvr)+ +<br />

r ∂θ ∂z<br />

τθr = τrθ =µ ∗<br />

<br />

1 ∂vr ∂vθ vθ<br />

+ −<br />

r ∂θ ∂r r<br />

τrz = τzr =µ ∗<br />

<br />

∂vr ∂vz<br />

+<br />

∂z ∂r<br />

τzθ = τθz =µ ∗<br />

<br />

1 ∂vz ∂vθ<br />

+<br />

r ∂θ ∂z<br />

In spherical coordinates, with velocity components vρ,vθ,vφ, the viscous stress tensor components have the<br />

form<br />

∗ ∂vρ<br />

τρρ =2µ<br />

∂ρ + λ∗∇· V<br />

τθθ =2µ ∗<br />

<br />

1 ∂vθ vρ<br />

+ + λ<br />

ρ ∂θ ρ<br />

∗ ∇· V<br />

τφφ =2µ ∗<br />

<br />

1 ∂vφ vρ<br />

+<br />

ρ sin θ ∂φ ρ + vθ<br />

<br />

cot θ<br />

+ λ<br />

ρ<br />

∗ ∇· V<br />

where ∇· V = 1<br />

ρ2 ∂ <br />

2 1 ∂<br />

1 ∂vφ<br />

ρ vρ + (sin θvθ)+<br />

∂ρ<br />

ρ sin θ ∂θ ρ sin θ ∂φ<br />

τρθ = τθρ =µ ∗<br />

<br />

ρ ∂<br />

<br />

vθ<br />

+<br />

∂ρ ρ<br />

1<br />

<br />

∂vρ<br />

ρ ∂θ<br />

τφρ = τρφ =µ ∗<br />

<br />

1 ∂vr ∂ vθ<br />

+ ρ<br />

ρ sin θ ∂φ ∂ρ ρ<br />

τθφ = τφθ =µ ∗<br />

<br />

sin θ ∂ vφ<br />

ρ ∂θ sin θ<br />

<br />

<br />

+ 1<br />

ρ sin θ<br />

Note that the viscous stress tensor is a linear function of the rate of deformation tensor Dij. Such a<br />

fluidiscalledaNewtonian fluid. In cases where the viscous stress tensor is a nonlinear function of Dij the<br />

fluid is called non-Newtonian.<br />

Definition: (Newtonian Fluid) If the viscous stress tensor τij<br />

is expressible as a linear function of the rate of deformation tensor<br />

Dij, the fluid is called a Newtonian fluid. Otherwise, the fluid is<br />

called a non-Newtonian fluid.<br />

Important note: Do not assume an arbitrary form for the constitutive equations unless there is experimental<br />

evidence to support your assumption. A constitutive equation is a very important step in the<br />

modeling processes as it describes the material you are working with. One cannot arbitrarily assign a form<br />

to the viscous stress and expect the mathematical equations to describe the correct fluid behavior. The form<br />

of the viscous stress is an important part of the modeling process and by assigning different forms to the<br />

viscous stress tensor then various types of materials can be modeled. We restrict our study in these notes<br />

to Newtonian fluids.<br />

In Cartesian coordinates the rate of deformation-stress constitutive equations for a Newtonian fluid can<br />

be written as<br />

σij = −pδij + λ ∗ δijDkk +2µ ∗ Dij<br />

<br />

∂vθ<br />

∂φ<br />

(2.5.20)<br />

287


288<br />

which can also be written in the alternative form<br />

σij = −pδij + λ ∗ δijvk,k + µ ∗ (vi,j + vj,i) (2.5.21)<br />

involving the gradient of the velocity.<br />

Upon transforming from a Cartesian coordinate system yi ,i = 1, 2, 3 to a more general system of<br />

coordinates xi ,i=1, 2, 3, we write<br />

∂y<br />

σmn = σij<br />

i<br />

∂xm ∂yj n . (2.5.22)<br />

∂x<br />

Now using the divergence from equation (2.1.3) and substituting equation (2.5.21) into equation (2.5.22) we<br />

obtain a more general expression for the constitutive equation. Performing the indicated substitutions there<br />

results<br />

σmn = −pδij + λ ∗ δijv k ,k + µ∗ (vi,j + vj,i) ∂y i<br />

∂x m<br />

σmn = −pg mn + λ ∗ g mn v k ,k + µ∗ (vm,n + vn,m).<br />

Dropping the bar notation, the stress-velocity strain relationships in the general coordinates x i ,i=1, 2, 3, is<br />

Summary<br />

∂y j<br />

∂x n<br />

σmn = −pgmn + λ ∗ gmng ik vi,k + µ ∗ (vm,n + vn,m). (2.5.23)<br />

The basic equations which describe the motion of a Newtonian fluid are :<br />

Continuity equation (Conservation of mass)<br />

∂ϱ<br />

∂t + ϱv i<br />

Dϱ<br />

=0, or ,i Dt + ϱ∇· V =0 1equation. (2.5.24)<br />

Conservation of linear momentum σ ij<br />

,j + ϱbi = ϱ ˙v i , 3equations<br />

or in vector form ϱ D V<br />

Dt = ϱ b + ∇·σ = ϱ b −∇p + ∇·τ (2.5.25)<br />

where σ = 3 3 i=1 j=1 (−pδij + τij)êi êj and τ = 3 3 i=1 j=1 τij êi êj are second order tensors. Conservation<br />

of angular momentum σ ij = σ ji , (Reduces the set of equations (2.5.23) to 6 equations.) Rate of<br />

deformation tensor (Velocity strain tensor)<br />

Constitutive equations<br />

Dij = 1<br />

2 (vi,j + vj,i) , 6equations. (2.5.26)<br />

σmn = −pgmn + λ ∗ gmng ik vi,k + µ ∗ (vm,n + vn,m), 6equations. (2.5.27)


In the cgs system of units the above quantities have the following units of measurements in Cartesian<br />

coordinates<br />

vi is the velocity field ,i=1, 2, 3, [vi] =cm/sec<br />

b i<br />

σij is the stress tensor, i,j=1, 2, 3, [σij] = dyne/cm 2<br />

ϱ is the fluid density [ϱ] =gm/cm 3<br />

is the external body forces per unit mass [b i ] = dyne/gm<br />

Dij is the rate of deformation tensor [Dij] =sec −1<br />

λ ∗ ,µ ∗<br />

p is the pressure [p] = dyne/cm 2<br />

are coefficients of viscosity [λ ∗ ]=[µ ∗ ]=Poise<br />

where 1Poise = 1gm/cm sec<br />

If we assume the external body forces per unit mass are known, then the equations (2.5.24), (2.5.25),<br />

(2.5.26), and (2.5.27) represent 16 equations in the 16 unknowns<br />

ϱ, v1,v2,v3,σ11,σ<strong>12</strong>,σ13,σ22,σ23,σ33,D11,D<strong>12</strong>,D13,D22,D23,D33.<br />

Navier-Stokes-Duhem Equations of Fluid Motion<br />

Substituting the stress tensor from equation (2.5.27) into the linear momentum equation (2.5.25), and<br />

assuming that the viscosity coefficients are constants, we obtain the Navier-Stokes-Duhem equations for fluid<br />

motion. In Cartesian coordinates these equations can be represented in any of the equivalent forms<br />

ϱ ˙vi = ϱbi − p,jδij +(λ ∗ + µ ∗ )vk,ki + µ ∗ vi,jj<br />

ϱ ∂vi<br />

∂t + ϱvjvi,j = ϱbi +(−pδij + τij) ,j<br />

∂ϱvi<br />

∂t +(ϱvivj + pδij − τij) ,j = ϱbi<br />

ϱ Dv<br />

Dt = ϱ b −∇p +(λ ∗ + µ ∗ )∇ (∇·v)+µ ∗ ∇ 2 v<br />

(2.5.28)<br />

where Dv ∂v<br />

= +(v ·∇) v is the material derivative, substantial derivative or convective derivative. This<br />

Dt ∂t<br />

derivative is represented as<br />

˙vi = ∂vi<br />

∂t<br />

∂vi<br />

+<br />

∂xj dxj dt<br />

= ∂vi<br />

∂t<br />

+ ∂vi<br />

∂x j vj = ∂vi<br />

∂t + vi,jv j . (2.5.29)<br />

In the vector form of equations (2.5.28), the terms on the right-hand side of the equation represent force<br />

terms. The term ϱb represents external body forces per unit volume. If these forces are derivable from a<br />

potential function φ, then the external forces are conservative and can be represented in the form −ϱ∇ φ.<br />

The term −∇ p is the gradient of the pressure and represents a force per unit volume due to hydrostatic<br />

pressure. The above statement is verified in the exercises that follow this section. The remaining terms can<br />

be written<br />

fviscous =(λ ∗ + µ ∗ )∇ (∇·v)+µ ∗ ∇ 2 v (2.5.30)<br />

289


290<br />

and are given the physical interpretation of an internal force per unit volume. These internal forces arise<br />

from the shearing stresses in the moving fluid. If fviscous is zero the vector equation in (2.5.28) is called<br />

Euler’s equation.<br />

If the viscosity coefficients are nonconstant, then the Navier-Stokes equations can be written in the<br />

Cartesian form<br />

ϱ[ ∂vi ∂vi<br />

+ vj ]=ϱbi +<br />

∂t ∂xj<br />

∂<br />

<br />

−pδij + λ<br />

∂xj<br />

∗ ∂vk<br />

δij + µ<br />

∂xk<br />

∗<br />

<br />

∂vi<br />

+<br />

∂xj<br />

∂vj<br />

<br />

∂xi<br />

=ϱbi − ∂p<br />

+<br />

∂xi<br />

∂<br />

<br />

∗ ∂vk<br />

λ +<br />

∂xi ∂xk<br />

∂<br />

∂xj <br />

µ ∗<br />

<br />

∂vi<br />

+<br />

∂xj<br />

∂vj<br />

<br />

∂xi<br />

which can also be written in terms of the bulk coefficient of viscosity ζ = λ ∗ + 2<br />

3 µ∗ as<br />

ϱ[ ∂vi<br />

∂t<br />

∂vi<br />

+ vj ]=ϱbi −<br />

∂xj<br />

∂p<br />

∂xi<br />

=ϱbi − ∂p<br />

+<br />

∂xi<br />

∂<br />

∂xi<br />

+ ∂<br />

<br />

(ζ −<br />

∂xi<br />

2<br />

These equations form the basics of viscous flow theory.<br />

3 µ∗ ) ∂vk<br />

<br />

+<br />

∂xk<br />

∂<br />

∂xj <br />

ζ ∂vk<br />

<br />

+<br />

∂xk<br />

∂<br />

∂xj <br />

µ ∗<br />

<br />

∂vi<br />

∂xj<br />

<br />

µ ∗<br />

<br />

∂vi<br />

+<br />

∂xj<br />

∂vj<br />

<br />

∂xi<br />

+ ∂vj<br />

−<br />

∂xi<br />

2<br />

3 δij<br />

<br />

∂vk<br />

∂xk<br />

In the case of orthogonal coordinates, where g (i)(i) = h 2 i (no summation) and gij =0fori = j, general<br />

expressions for the Navier-Stokes equations in terms of the physical components v(1),v(2),v(3) are:<br />

Navier-Stokes-Duhem equations for compressible fluid in terms of physical components: (i = j = k)<br />

<br />

∂v(i) v(1) ∂v(i)<br />

ϱ + +<br />

∂t h1 ∂x1<br />

v(2) ∂v(i)<br />

+<br />

h2 ∂x2<br />

v(3) ∂v(i)<br />

h3 ∂x3<br />

− v(j)<br />

<br />

v(j)<br />

hihj<br />

∂hj<br />

− v(i)<br />

∂xi<br />

∂hi<br />

<br />

+<br />

∂xj<br />

v(k)<br />

<br />

v(i)<br />

hihk<br />

∂hi<br />

− v(k)<br />

∂xk<br />

∂hk<br />

<br />

∂xi<br />

<br />

=<br />

ϱ b(i)<br />

−<br />

hi<br />

1 ∂p<br />

+<br />

hi ∂xi<br />

1 ∂ <br />

∗<br />

λ ∇· <br />

µ<br />

V +<br />

hi ∂xi<br />

∗<br />

<br />

hj ∂ v(j)<br />

+<br />

hihj hi ∂xi hj<br />

hi<br />

<br />

∂ v(i)<br />

hj ∂xj hi<br />

∂hi<br />

∂hj<br />

+ µ∗<br />

<br />

hi ∂ v(i)<br />

+<br />

hihk hk ∂xk hi<br />

hk<br />

<br />

∂ v(k) ∂hi<br />

−<br />

hi ∂xi hk ∂xk<br />

2µ∗<br />

<br />

1 ∂v(j)<br />

+<br />

hihj hj ∂xj<br />

v(k) ∂hj<br />

+<br />

hjhk ∂xk<br />

v(i)<br />

<br />

∂hj<br />

hihj ∂xi<br />

− 2µ∗<br />

<br />

1 ∂v(k)<br />

+<br />

hihk hk ∂xk<br />

v(i) ∂hk<br />

+<br />

hihk ∂xi<br />

v(k)<br />

<br />

∂hk ∂hk 1 ∂<br />

+<br />

2µ<br />

hkhj ∂xi ∂xi hihjhk ∂xi<br />

∗ <br />

1 ∂v(i)<br />

hjhk<br />

+<br />

hi ∂xi<br />

v(j) ∂hi<br />

+<br />

hihj ∂hj<br />

v(k)<br />

<br />

∂hi<br />

hihk ∂xk<br />

+ ∂<br />

<br />

µ<br />

∂xj<br />

∗ <br />

hj ∂ v(j)<br />

hihk<br />

+<br />

hi ∂xi hj<br />

hi<br />

<br />

∂ v(i)<br />

hj ∂xj hi<br />

<br />

+ ∂<br />

<br />

µ<br />

∂xk<br />

∗ <br />

hi ∂ v(i)<br />

hihj<br />

+<br />

hk ∂xk hi<br />

hk<br />

<br />

∂ v(k)<br />

hi ∂xi hk<br />

<br />

(2.5.31)<br />

where ∇·v is found in equation (2.1.4).<br />

In the above equation, cyclic values are assigned to i, j and k. That is, for the x1 components assign<br />

the values i =1,j =2,k =3;forthex2components assign the values i =2,j =3,k = 1; and for the x3<br />

components assign the values i =3,j =1,k =2.<br />

The tables 5.2, 5.3 and 5.4 show the expanded form of the Navier-Stokes equations in Cartesian, cylindrical<br />

and spherical coordinates respectively.


ϱ<br />

ϱ DVx<br />

Dt<br />

ϱ DVy<br />

Dt<br />

ϱ DVz<br />

Dt<br />

<br />

∂p ∂ ∗ ∂Vx<br />

=ϱbx − + 2µ<br />

∂x ∂x ∂x + λ∗∇· <br />

V + ∂<br />

<br />

µ<br />

∂y<br />

∗<br />

<br />

∂Vx<br />

∂y<br />

<br />

∂p ∂<br />

=ϱby − + µ<br />

∂y ∂x<br />

∗<br />

<br />

∂Vy<br />

∂x<br />

<br />

∂p ∂<br />

=ϱbz − + µ<br />

∂z ∂x<br />

∗<br />

<br />

∂Vz<br />

∂x<br />

where D<br />

Dt ()=∂()<br />

∂() ∂() ∂()<br />

+ Vx + Vy + Vz<br />

∂t ∂x ∂y ∂z<br />

DVr<br />

Dt<br />

<br />

DVθ<br />

ϱ<br />

Dt<br />

and ∇· V = ∂Vx<br />

∂x<br />

+ ∂Vy<br />

∂y<br />

+ ∂Vz<br />

∂z<br />

<br />

∂Vy<br />

+ +<br />

∂x<br />

∂<br />

<br />

µ<br />

∂z<br />

∗<br />

<br />

∂Vx<br />

∂z<br />

<br />

∂Vx<br />

+ +<br />

∂y<br />

∂<br />

<br />

∗ ∂Vy<br />

2µ<br />

∂y ∂y + λ∗∇· <br />

V + ∂<br />

<br />

µ<br />

∂z<br />

∗<br />

<br />

∂Vy<br />

∂z<br />

<br />

∂Vx<br />

+ +<br />

∂z<br />

∂<br />

<br />

µ<br />

∂y<br />

∗<br />

<br />

∂Vz<br />

∂y<br />

<br />

∂Vz<br />

+<br />

∂x<br />

<br />

∂Vz<br />

+<br />

∂y<br />

<br />

∂Vy<br />

+ +<br />

∂z<br />

∂<br />

<br />

∗ ∂Vz<br />

2µ<br />

∂z ∂z + λ∗∇· <br />

V<br />

Table 5.2 Navier-Stokes equations for compressible fluids in Cartesian coordinates.<br />

<br />

V 2<br />

θ<br />

− =ϱbr −<br />

r<br />

∂p<br />

∂r<br />

<br />

VrVθ<br />

+<br />

r<br />

ϱ DVz<br />

Dt<br />

+ ∂<br />

∂z<br />

<br />

∂ ∗ ∂Vr<br />

+ 2µ<br />

∂r ∂r + λ∗∇· <br />

V + 1<br />

<br />

∂<br />

µ<br />

r ∂θ<br />

∗<br />

<br />

1 ∂Vr<br />

r ∂θ<br />

<br />

µ ∗<br />

<br />

∂Vr ∂Vz<br />

+ +<br />

∂z ∂r<br />

2µ∗<br />

<br />

∂Vr 1 ∂Vθ Vr<br />

− −<br />

r ∂r r ∂θ r<br />

=ϱbθ − 1<br />

<br />

∂p ∂<br />

+ µ<br />

r ∂θ ∂r<br />

∗<br />

<br />

1 ∂Vr ∂Vθ Vθ<br />

+ − +<br />

r ∂θ ∂r r<br />

1<br />

<br />

∂<br />

2µ<br />

r ∂θ<br />

∗<br />

+ ∂<br />

<br />

µ<br />

∂z<br />

∗<br />

<br />

1 ∂Vz ∂Vθ<br />

+ +<br />

r ∂θ ∂z<br />

2µ∗<br />

<br />

1 ∂Vr ∂Vθ Vθ<br />

+ −<br />

r r ∂θ ∂r r<br />

=ϱbz − ∂p<br />

∂z<br />

+ 1<br />

r<br />

∂<br />

∂r<br />

<br />

µ ∗ <br />

∂Vr<br />

r<br />

∂z<br />

where D<br />

Dt ()=∂()<br />

∂() Vθ ∂() ∂()<br />

+ Vr + + Vz<br />

∂t ∂r r ∂θ ∂z<br />

and ∇·V = 1 ∂(rVr)<br />

r ∂r<br />

1 ∂Vθ ∂Vz<br />

+ +<br />

r ∂θ ∂z<br />

<br />

∂Vz<br />

+<br />

∂r<br />

+ 1<br />

r<br />

∂<br />

∂θ<br />

<br />

µ ∗<br />

<br />

1 ∂Vz<br />

r ∂θ<br />

+ ∂Vθ<br />

∂r<br />

<br />

1 ∂Vθ<br />

r ∂θ<br />

<br />

Vθ<br />

−<br />

r<br />

<br />

Vr<br />

+ + λ<br />

r<br />

∗ ∇· <br />

V<br />

(2.5.31a)<br />

<br />

∂Vθ<br />

+ +<br />

∂z<br />

∂<br />

<br />

∗ ∂Vz<br />

2µ<br />

∂z ∂z + λ∗∇· <br />

V<br />

Table 5.3 Navier-Stokes equations for compressible fluids in cylindrical coordinates.<br />

(2.5.31b)<br />

291


292<br />

Observe that for incompressible flow Dϱ<br />

Dt = 0 which implies ∇· V =0. Therefore, the assumptions<br />

of constant viscosity and incompressibility of the flow will simplify the above equations. If on the other<br />

hand the viscosity is temperature dependent and the flow is compressible, then one should add to the above<br />

equations the continuity equation, an energy equation and an equation of state. The energy equation comes<br />

from the first law of thermodynamics applied to a control volume within the fluid and will be considered<br />

in the sections ahead. The equation of state is a relation between thermodynamic variables which is added<br />

so that the number of equations equals the number of unknowns. Such a system of equations is known as<br />

a closed system. An example of an equation of state is the ideal gas law where pressure p is related to gas<br />

density ϱ and temperature T by the relation p = ϱRT where R is the universal gas constant.<br />

<br />

DVρ V 2<br />

θ<br />

ϱ −<br />

Dt + V 2 <br />

φ<br />

= ϱbρ −<br />

ρ<br />

∂p<br />

<br />

∂ ∗ ∂Vρ<br />

+ 2µ<br />

∂ρ ∂ρ ∂ρ + λ∗∇· <br />

V + 1<br />

<br />

∂<br />

µ<br />

ρ ∂θ<br />

∗ ρ ∂<br />

<br />

Vθ<br />

+<br />

∂ρ ρ<br />

µ∗<br />

<br />

∂Vρ<br />

ρ ∂θ<br />

+ 1<br />

<br />

∂ µ ∗ ∂Vρ<br />

ρ sin θ ∂φ ρ sin θ ∂φ + µ∗ρ ∂<br />

<br />

Vφ<br />

∂ρ ρ<br />

+ µ∗<br />

<br />

4<br />

ρ<br />

∂Vρ 2 ∂Vθ 4Vρ 2 ∂Vφ<br />

− − −<br />

∂ρ ρ ∂θ ρ ρ sin θ ∂φ − 2Vθ cot θ<br />

+ ρ cot θ<br />

ρ<br />

∂<br />

<br />

Vθ<br />

+<br />

∂ρ ρ<br />

cot θ<br />

<br />

∂Vρ<br />

ρ ∂θ<br />

<br />

DVθ VρVθ<br />

ϱ +<br />

Dt ρ − V 2 <br />

φ cot θ<br />

= ϱbθ −<br />

ρ<br />

1<br />

<br />

∂p ∂<br />

+ µ<br />

ρ ∂θ ∂ρ<br />

∗ ρ ∂<br />

<br />

Vθ<br />

+<br />

∂ρ ρ<br />

µ∗<br />

<br />

∂Vρ<br />

ρ ∂θ<br />

+ 1<br />

<br />

∂ 2µ ∗ <br />

∂Vθ<br />

+ Vρ + λ<br />

ρ ∂θ ρ ∂θ ∗ <br />

∇·V<br />

+ 1<br />

<br />

∂ µ ∗ <br />

sin θ ∂ Vφ<br />

+<br />

ρ sin θ ∂φ ρ ∂θ sin θ<br />

µ∗<br />

<br />

∂Vθ<br />

ρ sin θ ∂φ<br />

+ µ∗<br />

<br />

1 ∂Vθ 1 ∂Vφ<br />

2cotθ −<br />

ρ<br />

ρ ∂θ ρ sin θ ∂φ − Vθ<br />

<br />

cot θ<br />

+3 ρ<br />

ρ<br />

∂<br />

<br />

Vθ<br />

+<br />

∂ρ ρ<br />

1<br />

<br />

∂Vρ<br />

ρ ∂θ<br />

<br />

DVφ VθVφ<br />

ϱ +<br />

Dt ρ + VθVφ<br />

<br />

cot θ<br />

= ϱbφ −<br />

ρ<br />

1<br />

<br />

∂p ∂ µ ∗ ∂Vρ<br />

+<br />

ρ sin θ ∂φ ∂ρ ρ sin θ ∂φ + µ∗ρ ∂<br />

<br />

Vφ<br />

∂ρ ρ<br />

+ 1<br />

<br />

∂ µ ∗ <br />

sin θ ∂ Vφ<br />

+<br />

ρ ∂θ ρ ∂θ sin θ<br />

µ∗<br />

<br />

∂Vθ<br />

ρ sin θ ∂φ<br />

+ 1<br />

<br />

∂ 2µ ∗ <br />

1 ∂Vφ<br />

ρ sin θ ∂φ ρ sin θ ∂φ + Vρ + Vθ<br />

<br />

cot θ + λ ∗ ∇· <br />

V<br />

+ µ∗<br />

<br />

<br />

3 ∂Vρ ∂ Vφ<br />

sin θ ∂ Vφ<br />

+3ρ +2cotθ<br />

+<br />

ρ ρ sin θ ∂φ ∂ρ ρ<br />

ρ ∂θ sin θ<br />

1<br />

<br />

∂Vθ<br />

ρ sin θ ∂φ<br />

where D<br />

Dt ()=∂()<br />

∂() Vθ ∂() Vφ ∂()<br />

+ Vρ + +<br />

∂t ∂ρ ρ ∂θ ρ sin θ ∂φ<br />

and ∇· V = 1<br />

ρ2 ∂(ρ2Vρ) 1 ∂Vθ sin θ<br />

+ +<br />

∂ρ ρ sin θ ∂θ<br />

1 ∂Vφ<br />

ρ sin θ ∂φ<br />

Table 5.4 Navier-Stokes equations for compressible fluids in spherical coordinates.<br />

(2.5.31c)


We now consider various special cases of the Navier-Stokes-Duhem equations.<br />

Special Case 1: Assume that b is a conservative force such that b = −∇ φ. Also assume that the viscous<br />

force terms are zero. Consider steady flow ( ∂v<br />

∂t<br />

= 0) and show that equation (2.5.28) reduces to the equation<br />

(v ·∇) v = −1<br />

∇ p −∇φ ϱ is constant. (2.5.32)<br />

ϱ<br />

Employing the vector identity<br />

(v ·∇) v =(∇×v) × v + 1<br />

∇(v · v), (2.5.33)<br />

2<br />

we take the dot product of equation (2.5.32) with the vector v. Noting that v · [(∇×v) × v] =0 weobtain<br />

<br />

p 1<br />

v ·∇ + φ +<br />

ϱ 2 v2<br />

<br />

=0. (2.5.34)<br />

This equation shows that for steady flow we will have<br />

p 1<br />

+ φ +<br />

ϱ 2 v2 = constant (2.5.35)<br />

along a streamline. This result is known as Bernoulli’s theorem. In the special case where φ = gh is a<br />

force due to gravity, the equation (2.5.35) reduces to p v2<br />

+ + gh = constant. This equation is known as<br />

ϱ 2<br />

Bernoulli’s equation. It is a conservation of energy statement which has many applications in fluids.<br />

Special Case 2: Assume that b = −∇ φ is conservative and define the quantity Ω by<br />

Ω =∇×v =curlv ω = 1<br />

Ω (2.5.36)<br />

2<br />

as the vorticity vector associated with the fluid flow and observe that its magnitude is equivalent to twice<br />

the angular velocity of a fluid particle. Then using the identity from equation (2.5.33) we can write the<br />

Navier-Stokes-Duhem equations in terms of the vorticity vector. We obtain the hydrodynamic equations<br />

∂v<br />

∂t + Ω × v + 1<br />

2 ∇ v2 = − 1<br />

1<br />

∇ p −∇φ +<br />

ϱ ϱ fviscous, (2.5.37)<br />

where fviscous is defined by equation (2.5.30). In the special case of nonviscous flow this further reduces to<br />

the Euler equation<br />

∂v<br />

∂t + Ω × v + 1<br />

2 ∇ v2 = − 1<br />

∇ p −∇φ.<br />

ϱ<br />

If the density ϱ is a function of the pressure only it is customary to introduce the function<br />

then the Euler equation becomes<br />

P =<br />

p<br />

c<br />

dp<br />

ϱ<br />

so that ∇P = dP 1<br />

∇p =<br />

dp ϱ ∇p<br />

∂v<br />

∂t + Ω × v = −∇(P + φ + 1<br />

2 v2 ).<br />

Some examples of vorticies are smoke rings, hurricanes, tornadoes, and some sun spots. You can create<br />

a vortex by letting water stand in a sink and then remove the plug. Watch the water and you will see that<br />

a rotation or vortex begins to occur. Vortices are associated with circulating motion.<br />

293


294<br />

Pick an arbitrary simple closed curve C and place it in the fluid flow and define the line integral<br />

K = v · êt ds, where ds is an element of arc length along the curve C, v is the vector field defining the<br />

C<br />

velocity, and êt is a unit tangent vector to the curve C. The integral K is called the circulation of the fluid<br />

around the closed curve C. The circulation is the summation of the tangential components of the velocity<br />

field along the curve C. The local vorticity at a point is defined as the limit<br />

Circulation around C<br />

lim<br />

Area→0 Area inside C<br />

= circulation per unit area.<br />

By Stokes theorem, if curlv = 0, then the fluid is called irrotational and the circulation is zero. Otherwise<br />

the fluid is rotational and possesses vorticity.<br />

If we are only interested in the velocity field we can eliminate the pressure by taking the curl of both<br />

sides of the equation (2.5.37). If we further assume that the fluid is incompressible we obtain the special<br />

equations<br />

∇·v = 0 Incompressible fluid, ϱ is constant.<br />

Ω =curlv Definition of vorticity vector.<br />

∂ Ω<br />

∂t + ∇×( Ω × v) = µ∗<br />

ϱ ∇2 (2.5.38)<br />

Ω Results because curl of gradient is zero.<br />

Note that when Ω is identically zero, we have irrotational motion and the above equations reduce to the<br />

Cauchy-Riemann equations. Note also that if the term ∇×( Ω × v) is neglected, then the last equation in<br />

equation (2.5.38) reduces to a diffusion equation. This suggests that the vorticity diffuses through the fluid<br />

once it is created.<br />

Vorticity can be caused by a rigid rotation or by shear flow. For example, in cylindrical coordinates let<br />

V = rω êθ, with r, ω constants, denote a rotational motion, then curl V = ∇× V =2ωêz, which shows the<br />

vorticity is twice the rotation vector. Shear can also produce vorticity. For example, consider the velocity<br />

field V = y ê1 with y ≥ 0. Observe that this type of flow produces shear because | V | increases as y increases.<br />

For this flow field we have curl V = ∇× V = − ê3. The right-hand rule tells us that if an imaginary paddle<br />

wheel is placed in the flow it would rotate clockwise because of the shear effects.<br />

Scaled Variables<br />

In the Navier-Stokes-Duhem equations for fluid flow we make the assumption that the external body<br />

forces are derivable from a potential function φ and write b = −∇ φ [dyne/gm] Wealsowanttowritethe<br />

Navier-Stokes equations in terms of scaled variables<br />

v = v<br />

v0<br />

p = p<br />

p0<br />

ϱ = ϱ<br />

ϱ0<br />

t = t<br />

τ<br />

φ = φ<br />

gL ,<br />

x = x<br />

L<br />

y = y<br />

L<br />

z = z<br />

L<br />

which can be referred to as the barred system of dimensionless variables. Dimensionless variables are introduced<br />

by scaling each variable associated with a set of equations by an appropriate constant term called a<br />

characteristic constant associated with that variable. Usually the characteristic constants are chosen from<br />

various parameters used in the formulation of the set of equations. The characteristic constants assigned to<br />

each variable are not unique and so problems can be scaled in a variety of ways. The characteristic constants


assigned to each variable are scales, of the appropriate dimension, which act as reference quantities which<br />

reflect the order of magnitude changes expected of that variable over a certain range or area of interest<br />

associated with the problem. An inappropriate magnitude selected for a characteristic constant can result<br />

in a scaling where significant information concerning the problem can be lost. This is analogous to selecting<br />

an inappropriate mesh size in a numerical method. The numerical method might give you an answer but<br />

details of the answer might be lost.<br />

In the above scaling of the variables occurring in the Navier-Stokes equations we let v0 denote some<br />

characteristic speed, p0 a characteristic pressure, ϱ0 a characteristic density, L a characteristic length, g the<br />

acceleration of gravity and τ a characteristic time (for example τ = L/v0), then the barred variables v, p,<br />

ϱ,φ, t, x, y and z are dimensionless. Define the barred gradient operator by<br />

∇ = ∂<br />

∂x ê1 + ∂<br />

∂y ê2 + ∂<br />

∂z ê3<br />

where all derivatives are with respect to the barred variables. The above change of variables reduces the<br />

Navier-Stokes-Duhem equations<br />

to the form<br />

ϱ ∂v<br />

∂t + ϱ(v ·∇) v = −ϱ∇φ −∇p +(λ∗ + µ ∗ )∇ (∇·v)+µ ∗ ∇ 2 v, (2.5.39)<br />

<br />

ϱ0v0<br />

ϱ<br />

τ<br />

∂v<br />

∂t +<br />

<br />

ϱ0v2 <br />

0<br />

ϱ<br />

L<br />

v · ∇ v = −ϱ0gϱ∇ φ −<br />

<br />

p0<br />

∇p<br />

L<br />

+ (λ∗ + µ ∗ )<br />

L2 v0∇ ∇·v +<br />

∗ µ v0<br />

L2 <br />

∇ 2 v.<br />

(2.5.40)<br />

Now if each term in the equation (2.5.40) is divided by the coefficient ϱ0v2 0 /L, we obtain the equation<br />

Sϱ ∂v<br />

∂t + ϱ v · ∇ v = −1<br />

F<br />

which has the dimensionless coefficients<br />

E = p0<br />

ϱ0v 2 0<br />

= Euler number<br />

F = v2 0<br />

= Froude number, g is acceleration of gravity<br />

gL<br />

∗ λ<br />

ϱ∇ φ − E∇p +<br />

µ ∗ +1<br />

<br />

1<br />

R ∇ ∇·v + 1<br />

R ∇2v (2.5.41)<br />

R = ϱ0V0L<br />

µ ∗<br />

S = L<br />

τv0<br />

= Reynolds number<br />

= Strouhal number.<br />

Dropping the bars over the symbols, we write the dimensionless equation using the above coefficients.<br />

The scaled equation is found to have the form<br />

Sϱ ∂v<br />

∗<br />

1<br />

λ<br />

+ ϱ(v ·∇)v = − ϱ∇φ − E∇p +<br />

∂t F µ ∗ +1<br />

<br />

1<br />

1<br />

∇ (∇·v)+<br />

R R ∇2v (2.5.42)<br />

295


296<br />

Boundary Conditions<br />

Fluids problems can be classified as internal flows or external flows. An example of an internal flow<br />

problem is that of fluid moving through a converging-diverging nozzle. An example of an external flow<br />

problem is fluid flow around the boundary of an aircraft. For both types of problems there is some sort of<br />

boundary which influences how the fluid behaves. In these types of problems the fluid is assumed to adhere<br />

to a boundary. Let rb denote the position vector to a point on a boundary associated with a moving fluid,<br />

and let r denote the position vector to a general point in the fluid. Define v(r) as the velocity of the fluid at<br />

the point r and define v(rb) as the known velocity of the boundary. The boundary might be moving within<br />

the fluid or it could be fixed in which case the velocity at all points on the boundary is zero. We define the<br />

boundary condition associated with a moving fluid as an adherence boundary condition.<br />

Definition: (Adherence Boundary Condition)<br />

An adherence boundary condition associated with a fluid in motion<br />

is defined as the limit lim v(r) =v(rb) whererbis the position<br />

r→rb<br />

vector to a point on the boundary.<br />

Sometimes, when no finite boundaries are present, it is necessary to impose conditions on the components<br />

of the velocity far from the origin. Such conditions are referred to as boundary conditions at infinity.<br />

Summary and Additional Considerations<br />

Throughout the development of the basic equations of continuum mechanics we have neglected thermodynamical<br />

and electromagnetic effects. The inclusion of thermodynamics and electromagnetic fields adds<br />

additional terms to the basic equations of a continua. These basic equations describing a continuum are:<br />

Conservation of mass<br />

The conservation of mass is a statement that the total mass of a body is unchanged during its motion.<br />

This is represented by the continuity equation<br />

where ϱ is the mass density and vk is the velocity.<br />

Conservation of linear momentum<br />

∂ϱ<br />

∂t +(ϱvk ),k =0 or Dϱ<br />

Dt + ϱ∇· V =0<br />

The conservation of linear momentum requires that the time rate of change of linear momentum equal<br />

the resultant of all forces acting on the body. In symbols, we write<br />

where Dvi<br />

Dt<br />

= ∂vi<br />

∂t<br />

<br />

D<br />

ϱv<br />

Dt V<br />

i <br />

dτ = F<br />

S<br />

i (s) ni<br />

<br />

dS + ϱF<br />

V<br />

i (b)<br />

dτ +<br />

n<br />

α=1<br />

F i (α)<br />

(2.5.43)<br />

∂vi + ∂xk vk is the material derivative, F i (s) are the surface forces per unit area, F i (b) are the<br />

represents isolated external forces. Here S represents the surface and<br />

body forces per unit mass and F i (α)<br />

V represents the volume of the control volume. The right-hand side of this conservation law represents the<br />

resultant force coming from the consideration of all surface forces and body forces acting on a control volume.


Surface forces acting upon the control volume involve such things as pressures and viscous forces, while body<br />

forces are due to such things as gravitational, magnetic and electric fields.<br />

Conservation of angular momentum<br />

The conservation of angular momentum states that the time rate of change of angular momentum<br />

(moment of linear momentum) must equal the total moment of all forces and couples acting upon the body.<br />

In symbols,<br />

<br />

D<br />

ϱeijkx<br />

Dt V<br />

j v k <br />

dτ = eijkx<br />

S<br />

j F k <br />

(s) dS + ϱeijkx<br />

V<br />

j F k (b)<br />

where M i (α) represents concentrated couples and F k (α)<br />

Conservation of energy<br />

dτ +<br />

n<br />

represents isolated forces.<br />

(eijkx<br />

α=1<br />

j<br />

(α) F k (α) + M i (α)<br />

) (2.5.44)<br />

The conservation of energy law requires that the time rate of change of kinetic energy plus internal<br />

energies is equal to the sum of the rate of work from all forces and couples plus a summation of all external<br />

energies that enter or leave a control volume per unit of time. The energy equation results from the first law<br />

of thermodynamics and can be written<br />

D<br />

Dt (E + K) = ˙W + ˙Qh<br />

(2.5.45)<br />

where E is the internal energy, K is the kinetic energy, W˙ is the rate of work associated with surface and<br />

body forces, and ˙ Qh is the input heat rate from surface and internal effects.<br />

<br />

Let e denote the internal specific energy density within a control volume, then E = ϱe dτ represents<br />

V<br />

the total internal energy of the control volume. The kinetic energy of the control volume is expressed as<br />

K = 1<br />

<br />

ϱgijv<br />

2 V<br />

i v j dτ where vi is the velocity, ϱ is the density and dτ is a volume element. The energy (rate<br />

of work) associated with the body and surface forces is represented<br />

<br />

˙W = gijF<br />

S<br />

i (s) vj <br />

dS + ϱgijF<br />

V<br />

i (b) vj n<br />

dτ + (gijF i (α) vj + gijM i (α) ωj )<br />

where ωj is the angular velocity of the point xi (α) , F i (α) are isolated forces, and M i (α) are isolated couples.<br />

Two external energy sources due to thermal sources are heat flow qi and rate of internal heat production ∂Q<br />

∂t<br />

per unit volume. The conservation of energy can thus be represented<br />

<br />

D<br />

ϱ(e +<br />

Dt V<br />

1<br />

2 gijv i v j <br />

) dτ = (gijF<br />

S<br />

i (s) vj − qin i <br />

) dS + (ϱgijF<br />

V<br />

i (b) vj + ∂Q<br />

) dτ<br />

∂t<br />

+<br />

n<br />

α=1<br />

(gijF<br />

α=1<br />

i (α) vj + gijM i (α) ωj + U (α))<br />

(2.5.46)<br />

where U (α) represents all other energies resulting from thermal, mechanical, electric, magnetic or chemical<br />

sources which influx the control volume and D/Dt is the material derivative.<br />

In equation (2.5.46) the left hand side is the material derivative of an integral of the total energy<br />

et = ϱ(e + 1<br />

2 gijv i v j ) over the control volume. Material derivatives are not like ordinary derivatives and so<br />

297


298<br />

we cannot interchange the order of differentiation and integration in this term. Here we must use the result<br />

that<br />

<br />

D<br />

∂et<br />

et dτ =<br />

Dt V<br />

V ∂t + ∇·(et <br />

V ) dτ.<br />

To prove this result we consider a more general problem. Let A denote the amount of some quantity per<br />

unit mass. The quantity A can be a scalar, vector or tensor. The total amount of this quantity inside the<br />

control volume is A = <br />

ϱA dτ and therefore the rate of change of this quantity is<br />

V<br />

∂A<br />

∂t =<br />

<br />

∂(ϱA)<br />

dτ =<br />

V ∂t<br />

D<br />

<br />

ϱA dτ − ϱA<br />

Dt V<br />

S<br />

V · ˆndS,<br />

which represents the rate of change of material within the control volume plus the influx into the control<br />

volume. The minus sign is because ˆn is always a unit outward normal. By converting the surface integral to<br />

a volume integral, by the Gauss divergence theorem, and rearranging terms we find that<br />

<br />

D<br />

∂(ϱA)<br />

ϱA dτ =<br />

Dt V<br />

V ∂t + ∇·(ϱA <br />

V ) dτ.<br />

In equation (2.5.46) we neglect all isolated external forces and substitute F i (s) = σijnj, F i (b) = bi where<br />

σij = −pδij + τij. We then replace all surface integrals by volume integrals and find that the conservation of<br />

energy can be represented in the form<br />

∂et<br />

∂t + ∇·(et V )=∇(σ · V ) −∇·q + ϱb · V + ∂Q<br />

(2.5.47)<br />

∂t<br />

where et = ϱe + ϱ(v2 1 + v2 2 + v2 3 )/2 is the total energy and σ = 3 3 i=1 j=1 σij êi êj is the second order stress<br />

tensor. Here<br />

σ · V = −p 3<br />

3<br />

3<br />

V + τ1jvj ê1 + τ2jvj ê2 + τ3jvj ê3 = −p V + τ · V<br />

j=1<br />

j=1<br />

j=1<br />

and τij = µ ∗ (vi,j + vj,i)+λ∗δijvk,k is the viscous stress tensor. Using the identities<br />

ϱ D(et/ϱ)<br />

Dt<br />

∂et<br />

=<br />

∂t + ∇·(et V ) and ϱ D(et/ϱ)<br />

Dt<br />

together with the momentum equation (2.5.25) dotted with V as<br />

ϱ D V<br />

Dt · V = ϱ b · V −∇p · V +(∇·τ ) · V<br />

the energy equation (2.5.47) can then be represented in the form<br />

where Φ is the dissipation function and can be represented<br />

= ϱDe<br />

Dt + ϱD(V 2 /2)<br />

Dt<br />

ϱ De<br />

Dt + p(∇· V )=−∇ · q + ∂Q<br />

+Φ (2.5.48)<br />

∂t<br />

Φ=(τijvi) ,j − viτij,j = ∇·(τ · V ) − (∇·τ ) · V.<br />

As an exercise it can be shown that the dissipation function can also be represented as Φ = 2µ ∗ DijDij +λ ∗ Θ 2<br />

where Θ is the dilatation. The heat flow vector is determined from the Fourier law of heat conduction in


terms of the temperature T as q = −κ∇ T ,whereκis the thermal conductivity. Consequently, the energy<br />

equation can be written as<br />

ϱ De<br />

Dt + p(∇· V )= ∂Q<br />

+Φ+∇(k∇T ). (2.5.49)<br />

∂t<br />

In Cartesian coordinates (x, y, z) weuse<br />

D ∂ ∂<br />

= + Vx<br />

Dt ∂t ∂x<br />

∇· V = ∂Vx ∂Vy<br />

+<br />

∂x<br />

∇·(κ∇T )= ∂<br />

∂x<br />

In cylindrical coordinates (r, θ, z)<br />

and in spherical coordinates (ρ, θ, φ)<br />

∂y<br />

<br />

κ ∂T<br />

∂x<br />

∂ ∂<br />

+ Vy + Vz<br />

∂y ∂z<br />

∂Vz<br />

+<br />

∂z<br />

<br />

+ ∂<br />

<br />

κ<br />

∂y<br />

∂T<br />

<br />

+<br />

∂y<br />

∂<br />

<br />

κ<br />

∂z<br />

∂T<br />

<br />

∂z<br />

D ∂ ∂ Vθ ∂ ∂<br />

= + Vr + + Vz<br />

Dt ∂t ∂r r ∂θ ∂z<br />

∇·V = 1 ∂ 1<br />

(rVr)+<br />

r ∂r r2 ∂Vθ ∂Vz<br />

+<br />

∂θ ∂z<br />

∇·(κ∇T )= 1<br />

<br />

∂<br />

rκ<br />

r ∂r<br />

∂T<br />

<br />

+<br />

∂r<br />

1<br />

r2 <br />

∂<br />

κ<br />

∂θ<br />

∂T<br />

<br />

+<br />

∂θ<br />

∂<br />

<br />

κ<br />

∂z<br />

∂T<br />

<br />

∂z<br />

∂ Vθ ∂ Vφ ∂<br />

+ Vρ +<br />

∂ρ ρ ∂θ ρ sin θ ∂φ<br />

∂<br />

1 ∂<br />

(ρVρ)+<br />

∂ρ ρ sin θ ∂θ (Vθ sin θ)+ 1 ∂Vφ<br />

ρ sin θ ∂φ<br />

∇·(κ∇T )= 1<br />

ρ2 <br />

∂<br />

ρ<br />

∂ρ<br />

2 κ ∂T<br />

<br />

1<br />

+<br />

∂ρ ρ2 <br />

∂<br />

κ sin θ<br />

sin θ ∂θ<br />

∂T<br />

<br />

+<br />

∂θ<br />

D ∂<br />

=<br />

Dt ∂t<br />

∇· V = 1<br />

ρ2 1<br />

ρ 2 sin 2 θ<br />

<br />

∂<br />

κ<br />

∂φ<br />

∂T<br />

<br />

∂φ<br />

The combination of terms h = e + p/ϱ is known as enthalpy and at times is used to express the energy<br />

equation in the form<br />

ϱ Dh Dp ∂Q<br />

= + −∇·q +Φ.<br />

Dt Dt ∂t<br />

The derivation of this equation is left as an exercise.<br />

Conservative Systems<br />

Let Q denote some physical quantity per unit volume. Here Q can be either a scalar, vector or tensor<br />

field. Place within this field an imaginary simple closed surface S which encloses a volume V. The total<br />

amount of Q within the surface is given by <br />

V Qdτ and the rate of change of this amount with respect<br />

to time is ∂<br />

<br />

∂t Qdτ. The total amount of Q within S changes due to sources (or sinks) within the volume<br />

and by transport processes. Transport processes introduce a quantity J, called current, which represents a<br />

flow per unit area across the surface S. The inward flux of material into the volume is denoted <br />

S − J · ˆndσ<br />

(ˆn is a unit outward normal.) The sources (or sinks) SQ denotes a generation (or loss) of material per unit<br />

volume so that <br />

V SQ dτ denotes addition (or loss) of material to the volume. For a fixed volume we then<br />

have the material balance <br />

<br />

<br />

∂Q<br />

dτ = − J · ˆndσ+ SQ dτ.<br />

V ∂t S<br />

V<br />

299


300<br />

Using the divergence theorem of Gauss one can derive the general conservation law<br />

∂Q<br />

∂t + ∇· J = SQ<br />

(2.5.50)<br />

The continuity equation and energy equations are examples of a scalar conservation law in the special case<br />

where SQ =0. In Cartesian coordinates, we can represent the continuity equation by letting<br />

Q = ϱ and J = ϱ V = ϱ(Vx ê1 + Vy ê2 + Vz ê3) (2.5.51)<br />

The energy equation conservation law is represented by selecting Q = et and neglecting the rate of internal<br />

heat energy we let<br />

<br />

3<br />

<br />

J = (et + p)v1 − viτxi + qx ê1+<br />

i=1<br />

<br />

3<br />

<br />

(et + p)v2 − viτyi + qy ê2+<br />

(2.5.52)<br />

i=1<br />

<br />

3<br />

(et + p)v3 −<br />

<br />

ê3.<br />

i=1<br />

viτzi + qz<br />

In a general orthogonal system of coordinates (x1,x2,x3) the equation (2.5.50) is written<br />

∂<br />

∂t ((h1h2h3Q)) + ∂<br />

((h2h3J1)) +<br />

∂x1<br />

∂<br />

((h1h3J2)) +<br />

∂x2<br />

∂<br />

((h1h2J3)) = 0,<br />

∂x3<br />

where h1,h2,h3 are scale factors obtained from the transformation equations to the general orthogonal<br />

coordinates.<br />

The momentum equations are examples of a vector conservation law having the form<br />

where a is a vector and T is a second order symmetric tensor T =<br />

∂a<br />

∂t + ∇·(T )=ϱ b (2.5.53)<br />

3<br />

k=1 j=1<br />

3<br />

Tjk êj êk. In Cartesian coordinates<br />

we let a = ϱ(Vx ê1 + Vy ê2 + Vz ê3) andTij = ϱvivj + pδij − τij. In general coordinates (x1,x2,x3) the<br />

momentum equations result by selecting a = ϱ V and Tij = ϱvivj + pδij − τij. In a general orthogonal system<br />

the conservation law (2.5.53) has the general form<br />

∂<br />

∂t ((h1h2h3a)) + ∂<br />

<br />

<br />

(h2h3T · ê1) +<br />

∂x1<br />

∂<br />

<br />

<br />

(h1h3T · ê2) +<br />

∂x2<br />

∂<br />

<br />

<br />

(h1h2T · ê3) = ϱ<br />

∂x3<br />

b. (2.5.54)<br />

Neglecting body forces and internal heat production, the continuity, momentum and energy equations<br />

can be expressed in the strong conservative form<br />

where<br />

∂U<br />

∂t<br />

+ ∂E<br />

∂x<br />

+ ∂F<br />

∂y<br />

+ ∂G<br />

∂z<br />

⎡<br />

ρ<br />

⎤<br />

⎢ ρVx ⎥<br />

⎢ ⎥<br />

U = ⎢ ρVy ⎥<br />

⎣ ⎦<br />

ρVz<br />

et<br />

=0 (2.5.55)<br />

(2.5.56)


⎡<br />

⎢<br />

E = ⎢<br />

⎣<br />

ρVx<br />

ρV 2 x + p − τxx<br />

ρVxVy − τxy<br />

ρVxVz − τxz<br />

⎡<br />

(et + p)Vx − Vxτxx − Vyτxy − Vzτxz + qx<br />

ρVy<br />

⎢<br />

F = ⎢<br />

⎣<br />

ρVxVy − τxy<br />

ρV 2<br />

⎤<br />

y + p − τyy<br />

ρVyVz − τyz<br />

⎥<br />

⎦<br />

⎡<br />

(et + p)Vy − Vxτyx − Vyτyy − Vzτyz + qy<br />

ρVz<br />

⎤<br />

⎢<br />

G = ⎢<br />

⎣<br />

ρVxVz − τxz<br />

ρVyVz − τyz<br />

+ p − τzz<br />

⎥<br />

⎦<br />

ρV 2<br />

z<br />

(et + p)Vz − Vxτzx − Vyτzy − Vzτzz + qz<br />

where the shear stresses are τij = µ ∗ (Vi,j + Vj,i)+δijλ ∗ Vk,k for i, j, k =1, 2, 3.<br />

Computational Coordinates<br />

⎤<br />

⎥<br />

⎦<br />

(2.5.57)<br />

(2.5.58)<br />

(2.5.59)<br />

To transform the conservative system (2.5.55) from a physical (x, y, z) domain to a computational (ξ,η,ζ)<br />

domain requires that a general change of variables take place. Consider the following general transformation<br />

of the independent variables<br />

ξ = ξ(x, y, z) η = η(x, y, z) ζ = ζ(x, y, z) (2.5.60)<br />

with Jacobian different from zero. The chain rule for changing variables in equation (2.5.55) requires the<br />

operators<br />

∂( ) ∂( )<br />

=<br />

∂x ∂ξ ξx + ∂()<br />

∂η ηx + ∂()<br />

∂ζ ζx<br />

∂( ) ∂( )<br />

=<br />

∂y ∂ξ ξy + ∂()<br />

∂η ηy + ∂()<br />

∂ζ ζy<br />

∂( ) ∂( )<br />

=<br />

∂z ∂ξ ξz + ∂()<br />

∂η ηz + ∂()<br />

∂ζ ζz<br />

(2.5.61)<br />

The partial derivatives in these equations occur in the differential expressions<br />

dξ =ξx dx + ξy dy + ξz dz<br />

dη =ηx dx + ηy dy + ηz dz<br />

dζ =ζx dx + ζy dy + ζz dz<br />

or<br />

⎡<br />

⎣ dξ<br />

⎤ ⎡<br />

dη ⎦ =<br />

dζ<br />

In a similar mannaer from the inverse transformation equations<br />

we can write the differentials<br />

⎣ ξx ξy ξz<br />

ηx ηy ηz<br />

ζx ζy ζz<br />

⎤ ⎡<br />

⎦ ⎣ dx<br />

⎤<br />

dy ⎦ (2.5.62)<br />

dz<br />

x = x(ξ,η,ζ) y = y(ξ,η,ζ) z = z(ξ,η,ζ) (2.5.63)<br />

dx =xξ dξ + xη dη + xζ dζ<br />

dy =yξ dξ + yη dη + yζ dζ<br />

dz =zξ dξ + zζ dζ + zζ dζ<br />

or<br />

⎡<br />

⎣ dx<br />

⎤ ⎡<br />

dy ⎦ =<br />

dz<br />

⎣ xξ xη xζ<br />

yξ yη yζ<br />

zξ zη zζ<br />

⎤ ⎡<br />

⎦ ⎣ dξ<br />

⎤<br />

dη ⎦ (2.5.64)<br />

dζ<br />

301


302<br />

The transformations (2.5.62) and (2.5.64) are inverses of each other and so we can write<br />

⎡<br />

⎤ ⎡<br />

⎤−1<br />

⎦ =<br />

⎦<br />

⎣ ξx ξy ξz<br />

ηx ηy ηz<br />

ζx ζy ζz<br />

⎣ xξ xη xζ<br />

yξ yη yζ<br />

zξ zη zζ<br />

⎡<br />

=J ⎣ yηzζ − yζzη<br />

−(yξzζ − yζzξ)<br />

−(xηzζ − xζzη)<br />

xξzζ − xζzξ<br />

⎤<br />

xηyζ − xζyη<br />

−(xξyζ − xζyξ) ⎦<br />

yξzη − yηzξ −(xξzη − xηzξ) xξyη − xηyξ<br />

By comparing like elements in equation (2.5.65) we obtain the relations<br />

ξx =J(yηzζ − yζzη)<br />

ξy = − J(xηzζ − xζzη)<br />

ξz =J(xηyζ − xζyη)<br />

ηx = − J(yξzζ − yζzξ)<br />

ηy =J(xξzζ − zζzξ)<br />

ηz = − J(xξyζ − xζyξ)<br />

ζx =J(yξzη − yηzξ)<br />

ζy = − J(xξzη − xηzξ)<br />

ζz =J(xξyη − xηyξ)<br />

The equations (2.5.55) can now be written in terms of the new variables (ξ,η,ζ) as<br />

∂U<br />

∂t<br />

(2.5.65)<br />

(2.5.66)<br />

+ ∂E<br />

∂ξ ξx + ∂E<br />

∂η ηx + ∂E<br />

∂ζ ζx + ∂F<br />

∂ξ ξy + ∂F<br />

∂η ηy + ∂F<br />

∂ζ ζy + ∂G<br />

∂ξ ξz + ∂G<br />

∂η ηz + ∂G<br />

∂ζ ζz =0 (2.5.67)<br />

Now divide each term by the Jacobian J and write the equation (2.5.67) in the form<br />

∂<br />

∂t<br />

<br />

U<br />

+<br />

J<br />

∂<br />

∂ξ<br />

<br />

Eξx + Fξy + Gξz<br />

J<br />

+ ∂<br />

<br />

Eηx + Fηy + Gηz<br />

∂η J<br />

+ ∂<br />

<br />

Eζx + Fζy + Gζz<br />

∂ζ J<br />

<br />

∂ ξx<br />

− E<br />

+<br />

∂ξ J<br />

∂<br />

<br />

ηx<br />

+<br />

∂η J<br />

∂<br />

<br />

ζx<br />

∂ζ J<br />

<br />

∂ ξy<br />

− F<br />

+<br />

∂ξ J<br />

∂<br />

<br />

ηy<br />

+<br />

∂η J<br />

∂<br />

<br />

ζy<br />

∂ζ J<br />

<br />

∂ ξz<br />

− G<br />

+<br />

∂ξ J<br />

∂<br />

<br />

ηz<br />

+<br />

∂η J<br />

∂<br />

<br />

ζz<br />

=0<br />

∂ζ J<br />

(2.5.68)<br />

Using the relations given in equation (2.5.66) one can show that the curly bracketed terms above are all zero<br />

and so the transformed equations (2.5.55) can also be written in the conservative form<br />

where<br />

∂ U<br />

∂t + ∂ E<br />

∂ξ + ∂ F<br />

∂η + ∂ G<br />

=0 (2.5.69)<br />

∂ζ<br />

U = U<br />

J<br />

E = Eξx + Fξy + Gξz<br />

J<br />

F = Eηx + Fηy + Gηz<br />

J<br />

G = Eζx + Fζy + Gζz<br />

J<br />

(2.5.70)


Fourier law of heat conduction<br />

The Fourier law of heat conduction can be written qi = −κT,i for isotropic material and qi = −κijT,j<br />

for anisotropic material. The Prandtl number is a nondimensional constant defined as Pr = cpµ∗<br />

κ<br />

the heat flow terms can be represented in Cartesian coordinates as<br />

∗ cpµ ∂T<br />

qx = −<br />

Pr ∂x<br />

∗ cpµ ∂T<br />

qy = −<br />

Pr ∂y<br />

∗ cpµ ∂T<br />

qz = −<br />

Pr ∂z<br />

Now one can employ the equation of state relations P = ϱe(γ − 1), cp = γR<br />

γ−1 , cpT = γRT<br />

γ−1<br />

so that<br />

and write the<br />

above equations in the alternate forms<br />

µ<br />

qx = −<br />

∗ <br />

∂ γP<br />

Pr(γ − 1) ∂x ϱ<br />

µ<br />

qy = −<br />

∗ <br />

∂ γP<br />

Pr(γ − 1) ∂y ϱ<br />

µ<br />

qz = −<br />

∗ <br />

∂ γP<br />

Pr(γ − 1) ∂z ϱ<br />

<br />

γP<br />

The speed of sound is given by a =<br />

ϱ = γRT and so one can substitute a2 in place of the ratio γP<br />

in the above equations.<br />

ϱ<br />

Equilibrium and Nonequilibrium Thermodynamics<br />

High temperature gas flows require special considerations. In particular, the specific heat for monotonic<br />

and diatomic gases are different and are in general a function of temperature. The energy of a gas can be<br />

written as e = et + er + ev + ee + en where et represents translational energy, er is rotational energy, ev is<br />

vibrational energy, ee is electronic energy, and en is nuclear energy. The gases follow a Boltzmann distribution<br />

for each degree of freedom and consequently at very high temperatures the rotational, translational and<br />

vibrational degrees of freedom can each have their own temperature. Under these conditions the gas is said<br />

to be in a state of nonequilibrium. In such a situation one needs additional energy equations. The energy<br />

equation developed in these notes is for equilibrium thermodynamics where the rotational, translational and<br />

vibrational temperatures are the same.<br />

Equation of state<br />

It is assumed that an equation of state such as the universal gas law or perfect gas law pV = nRT<br />

holds which relates pressure p [N/m 2 ], volume V [m 3 ], amount of gas n [mol],and temperature T [K] where<br />

R [J/mol − K] is the universal molar gas constant. If the ideal gas law is represented in the form p = ϱRT<br />

where ϱ [Kg/m3 ] is the gas density, then the universal gas constant must be expressed in units of [J/Kg−K]<br />

(See Appendix A). Many gases deviate from this ideal behavior. In order to account for the intermolecular<br />

forces associated with high density gases, an empirical equation of state of the form<br />

M1 <br />

p = ρRT +<br />

n=1<br />

βnρ n+r1 + e −γ1ρ−γ2ρ2<br />

M2 <br />

cnρ n+r2<br />

involving constants M1,M2,βn,cn,r1,r2,γ1,γ2 is often used. For a perfect gas the relations<br />

e = cvT γ = cp<br />

cv<br />

cv = R<br />

γ − 1<br />

n=1<br />

cp = γR<br />

γ − 1<br />

h = cpT<br />

hold, where R is the universal gas constant, cv is the specific heat at constant volume, cp is the specific<br />

heat at constant pressure, γ is the ratio of specific heats and h is the enthalpy. For cv and cp constants the<br />

relations p =(γ− 1)ϱe and RT =(γ− 1)e can be verified.<br />

303


304<br />

EXAMPLE 2.5-1. (One-dimensional fluid flow)<br />

Construct an x-axis running along the center line of a long cylinder with cross sectional area A. Consider<br />

the motion of a gas driven by a piston and moving with velocity v1 = u in the x-direction. From an Eulerian<br />

point of view we imagine a control volume fixed within the cylinder and assume zero body forces. We require<br />

the following equations be satisfied.<br />

Conservation of mass ∂ϱ<br />

∂t +div(ϱ V ) = 0 which in one-dimension reduces to ∂ϱ ∂<br />

+ (ϱu) =0.<br />

∂t ∂x<br />

Conservation of momentum, equation (2.5.28) reduces to ∂ ∂ 2<br />

(ϱu)+ ϱu<br />

∂t ∂x<br />

+ ∂p<br />

∂x =0.<br />

Conservation of energy, equation (2.5.48) in the absence of heat flow and internal heat production,<br />

∂e ∂e<br />

becomes in one dimension ϱ + u + p<br />

∂t ∂x<br />

∂u<br />

=0. Using the conservation of mass relation this<br />

∂x<br />

equation can be written in the form ∂ ∂<br />

(ϱe)+<br />

∂t ∂x (ϱeu)+p∂u<br />

∂x =0.<br />

In contrast, from a Lagrangian point of view we let the control volume move with the flow and consider<br />

advection terms. This gives the following three equations which can then be compared with the above<br />

Eulerian equations of motion.<br />

Conservation of mass d<br />

Dϱ<br />

(ϱJ) = 0 which in one-dimension is equivalent to + ϱ∂u<br />

dt Dt ∂x =0.<br />

Conservation of momentum, equation (2.5.25) in one-dimension ϱ Du ∂p<br />

+<br />

Dt ∂x =0.<br />

Conservation of energy, equation (2.5.48) in one-dimension ϱ De<br />

+ p∂u =0. In the above equations<br />

Dt ∂x<br />

D() ∂ ∂<br />

Dt = ∂t ()+u ∂x (). The Lagrangian viewpoint gives three equations in the three unknowns ρ, u, e.<br />

In both the Eulerian and Lagrangian equations the pressure p represents the total pressure p = pg + pv<br />

where pg is the gas pressure and pv is the viscous pressure which causes loss of kinetic energy. The gas pressure<br />

is a function of ϱ, e and is determined from the ideal gas law pg = ϱRT = ϱ(cp − cv)T = ϱ( cp<br />

cv − 1)cvT or<br />

pg = ϱ(γ − 1)e. Some kind of assumption is usually made to represent the viscous pressure pv as a function<br />

of e, u. The above equations are then subjected to boundary and initial conditions and are usually solved<br />

numerically.<br />

Entropy inequality<br />

Energy transfer is not always reversible. Many energy transfer processes are irreversible. The second<br />

law of thermodynamics allows energy transfer to be reversible only in special circumstances. In general,<br />

the second law of thermodynamics can be written as an entropy inequality, known as the Clausius-Duhem<br />

inequality. This inequality states that the time rate of change of the total entropy is greater than or equal to<br />

the total entropy change occurring across the surface and within the body of a control volume. The Clausius-<br />

Duhem inequality places restrictions on the constitutive equations. This inequality can be expressed in the<br />

form<br />

<br />

<br />

D<br />

ϱs dτ ≥ s<br />

Dt V<br />

S<br />

<br />

Rate of entropy increase<br />

i <br />

n<br />

ni dS + ρb dτ + B (α)<br />

V<br />

α=1<br />

<br />

Entropy input rate into control volume<br />

where s is the specific entropy density, s i is an entropy flux, b is an entropy source and B (α) are isolated<br />

entropy sources. Irreversible processes are characterized by the use of the inequality sign while for reversible


Figure 2.5-3. Interaction of various fields.<br />

processes the equality sign holds. The Clausius-Duhem inequality is assumed to hold for all independent<br />

thermodynamical processes.<br />

If in addition there are electric and magnetic fields to consider, then these fields place additional forces<br />

upon the material continuum and we must add all forces and moments due to these effects. In particular we<br />

must add the following equations<br />

Gauss’s law for magnetism ∇· B =0<br />

Gauss’s law for electricity ∇· D = ϱe<br />

Faraday’s law ∇× E = − ∂ B<br />

∂t<br />

Ampere’s law ∇× H = J + ∂ D<br />

∂t<br />

1<br />

√ g<br />

1<br />

√ g<br />

∂<br />

∂x i (√ gB i )=0.<br />

∂<br />

∂x i (√ gD i )=ϱe.<br />

ɛ ijk Ek,j = − ∂Bi<br />

∂t .<br />

ɛ ijk Hk,j = J i + ∂Di<br />

∂t .<br />

where ϱe is the charge density, J i is the current density, Di = ɛ j<br />

i Ej + Pi is the electric displacement vector,<br />

Hi is the magnetic field, Bi = µ j<br />

i Hj + Mi is the magnetic induction, Ei is the electric field, Mi is the<br />

magnetization vector and Pi is the polarization vector. Taking the divergence of Ampere’s law produces the<br />

law of conservation of charge which requires that<br />

∂ϱe<br />

∂t + ∇· J =0<br />

∂ϱe<br />

∂t<br />

+ 1<br />

√ g<br />

∂<br />

∂x i (√ gJ i )=0.<br />

The figure 2.5-3 is constructed to suggest some of the interactions that can occur between various<br />

variables which define the continuum. Pyroelectric effects occur when a change in temperature causes<br />

changes in the electrical properties of a material. Temperature changes can also change the mechanical<br />

properties of materials. Similarly, piezoelectric effects occur when a change in either stress or strain causes<br />

changes in the electrical properties of materials. Photoelectric effects are said to occur if changes in electric<br />

or mechanical properties effect the refractive index of a material. Such changes can be studied by modifying<br />

the constitutive equations to include the effects being considered.<br />

From figure 2.5-3 we see that there can exist a relationship between the displacement field Di and<br />

electric field Ei. When this relationship is linear we can write Di = ɛjiEj and Ej = βjnDn, whereɛji are<br />

305


306<br />

dielectric constants and βjn are dielectric impermabilities. Similarly, when linear piezoelectric effects exist<br />

we can write linear relations between stress and electric fields such as σij = −gkijEk and Ei = −eijkσjk,<br />

where gkij and eijk are called piezoelectric constants. If there is a linear relation between strain and an<br />

electric fields, this is another type of piezoelectric effect whereby eij = dijkEk and Ek = −hijkejk, where<br />

dijk and hijk are another set of piezoelectric constants. Similarly, entropy changes can cause pyroelectric<br />

effects. Piezooptical effects (photoelasticity) occurs when mechanical stresses change the optical properties of<br />

the material. Electrical and heat effects can also change the optical properties of materials. Piezoresistivity<br />

occurs when mechanical stresses change the electric resistivity of materials. Electric field changes can cause<br />

variations in temperature, another pyroelectric effect. When temperature effects the entropy of a material<br />

this is known as a heat capacity effect. When stresses effect the entropy in a material this is called a<br />

piezocaloric effect. Some examples of the representation of these additional effects are as follows. The<br />

piezoelectric effects are represented by equations of the form<br />

σij = −hmijDm Di = dijkσjk eij = gkijDk Di = eijkejk<br />

where hmij, dijk, gkij and eijk are piezoelectric constants.<br />

Knowledge of the material or electric interaction can be used to help modify the constitutive equations.<br />

For example, the constitutive equations can be modified to included temperature effects by expressing the<br />

constitutive equations in the form<br />

σij = cijklekl − βij∆T and eij = sijklσkl + αij∆T<br />

where for isotropic materials the coefficients αij and βij are constants. As another example, if the strain is<br />

modified by both temperature and an electric field, then the constitutive equations would take on the form<br />

eij = sijklσkl + αij∆T + dmijEm.<br />

Note that these additional effects are additive under conditions of small changes. That is, we may use the<br />

principal of superposition to calculate these additive effects.<br />

If the electric field and electric displacement are replaced by a magnetic field and magnetic flux, then<br />

piezomagnetic relations can be found to exist between the variables involved. One should consult a handbook<br />

to determine the order of magnitude of the various piezoelectric and piezomagnetic effects. For a large<br />

majority of materials these effects are small and can be neglected when the field strengths are weak.<br />

The Boltzmann Transport Equation<br />

The modeling of the transport of particle beams through matter, such as the motion of energetic protons<br />

or neutrons through bulk material, can be approached using ideas from the classical kinetic theory of gases.<br />

Kinetic theory is widely used to explain phenomena in such areas as: statistical mechanics, fluids, plasma<br />

physics, biological response to high-energy radiation, high-energy ion transport and various types of radiation<br />

shielding. The problem is basically one of describing the behavior of a system of interacting particles and their<br />

distribution in space, time and energy. The average particle behavior can be described by the Boltzmann<br />

equation which is essentially a continuity equation in a six-dimensional phase space (x, y, z, Vx,Vy,Vz). We


will be interested in examining how the particles in a volume element of phase space change with time. We<br />

introduce the following notation:<br />

(i) r the position vector of a typical particle of phase space and dτ = dxdydz the corresponding spatial<br />

volume element at this position.<br />

(ii) V the velocity vector associated with a typical particle of phase space and dτv<br />

corresponding velocity volume element.<br />

(iii)<br />

= dVxdVydVz the<br />

Ω a unit vector in the direction of the velocity V = v Ω.<br />

(iv) E = 1<br />

2 mv2 kinetic energy of particle.<br />

(v) d Ω is a solid angle about the direction Ωanddτ dE d Ω is a volume element of phase space involving the<br />

solid angle about the direction Ω.<br />

(vi) n = n(r, E, Ω,t) the number of particles in phase space per unit volume at position r per unit velocity<br />

at position V per unit energy in the solid angle d Ωattimetand N = N(r, E, Ω,t)=vn(r, E, Ω,t)<br />

the number of particles per unit volume per unit energy in the solid angle d Ωattimet. The quantity<br />

N(r, E, Ω,t)dτ dE d Ω represents the number of particles in a volume element around the position r with<br />

energy between E and E + dE having direction Ω in the solid angle d Ωattimet.<br />

(vii) φ(r, E, Ω,t)=vN(r, E, Ω,t) is the particle flux (number of particles/cm2 − Mev − sec).<br />

(viii) Σ(E ′ → E, Ω ′ → Ω) a scattering cross-section which represents the fraction of particles with energy E ′<br />

and direction Ω ′ that scatter into the energy range between E and E + dE having direction Ωinthe<br />

solid angle d Ω per particle flux.<br />

(ix) Σs(E,r) fractional number of particles scattered out of volume element of phase space per unit volume<br />

per flux.<br />

(x) Σa(E,r) fractional number of particles absorbed in a unit volume of phase space per unit volume per<br />

flux.<br />

Consider a particle at time t having a position r in phase space as illustrated in the figure 2.5-4. This<br />

particle has a velocity V in a direction Ω and has an energy E. In terms of dτ = dx dy dz, ΩandEan element of volume of phase space can be denoted dτdEd Ω, where d Ω=d Ω(θ, ψ) =sinθdθdψ is a solid angle<br />

about the direction Ω.<br />

The Boltzmann transport equation represents the rate of change of particle density in a volume element<br />

dτ dE d Ω of phase space and is written<br />

d<br />

dt N(r, E, Ω,t) dτ dE d Ω=DCN(r, E, Ω,t) (2.5.71)<br />

where DC is a collision operator representing gains and losses of particles to the volume element of phase<br />

space due to scattering and absorption processes. The gains to the volume element are due to any sources<br />

S(r, E, Ω,t) per unit volume of phase space, with units of number of particles/sec per volume of phase space,<br />

together with any scattering of particles into the volume element of phase space. That is particles entering<br />

the volume element of phase space with energy E, which experience a collision, leave with some energy<br />

E − ∆E and thus will be lost from our volume element. Particles entering with energies E ′ >Emay,<br />

307


308<br />

Figure 2.5-4. Volume element and solid angle about position r.<br />

depending upon the cross-sections, exit with energy E ′ − ∆E = E and thus will contribute a gain to the<br />

volume element. In terms of the flux φ the gains due to scattering into the volume element are denoted by<br />

<br />

d Ω ′<br />

<br />

dE ′ Σ(E ′ → E, Ω ′ → Ω)φ(r, E ′ , Ω,t) dτ dE d Ω<br />

and represents the particles at position r experiencing a scattering collision with a particle of energy E ′ and<br />

direction Ω ′ which causes the particle to end up with energy between E and E + dE and direction Ωind Ω.<br />

The summations are over all possible initial energies.<br />

In terms of φ the losses are due to those particles leaving the volume element because of scattering and<br />

are<br />

Σs(E,r)φ(r, E, Ω,t)dτ dE d Ω.<br />

The particles which are lost due to absorption processes are<br />

Σa(E,r)φ(r, E, Ω,t) dτ dE d Ω.<br />

The total change to the number of particles in an element of phase space per unit of time is obtained by<br />

summing all gains and losses. This total change is<br />

<br />

dN<br />

dτ dE dΩ =<br />

dt<br />

d Ω ′<br />

<br />

dE ′ Σ(E ′ → E, Ω ′ → Ω)φ(r, E ′ , Ω,t) dτ dE d Ω<br />

− Σs(E,r)φ(r, E, Ω,t)dτ dE dΩ<br />

− Σa(E,r)φ(r, E, Ω,t) dτ dE d Ω<br />

(2.5.72)<br />

The rate of change dN<br />

dt<br />

+ S(r, E, Ω,t)dτ dE d Ω.<br />

on the left-hand side of equation (2.5.72) expands to<br />

dN<br />

dt<br />

∂N ∂N dx ∂N dy ∂N dz<br />

= + + +<br />

∂t ∂x dt ∂y dt ∂z dt<br />

+ ∂N dVx ∂N dVy ∂N dVz<br />

+ +<br />

∂Vx dt ∂Vy dt ∂Vz dt


which can be written as<br />

where d V<br />

dt = F<br />

m<br />

dN<br />

dt<br />

= ∂N<br />

∂t + V ·∇rN + F<br />

m ·∇ V N (2.5.73)<br />

represents any forces acting upon the particles. The Boltzmann equation can then be<br />

expressed as<br />

∂N<br />

∂t + V ·∇rN + F<br />

m ·∇ V N =Gains−Losses. (2.5.74)<br />

If the right-hand side of the equation (2.5.74) is zero, the equation is known as the Liouville equation. In<br />

the special case where the velocities are constant and do not change with time the above equation (2.5.74)<br />

can be written in terms of the flux φ and has the form<br />

<br />

1 ∂<br />

v ∂t + <br />

Ω ·∇r +Σs(E,r)+Σa(E,r) φ(r, E, Ω,t)=DCφ (2.5.75)<br />

where<br />

<br />

DCφ =<br />

d Ω ′<br />

<br />

dE ′ Σ(E ′ → E, Ω ′ → Ω)φ(r, E ′ , Ω ′ ,t)+S(r, E, Ω,t).<br />

The above equation represents the Boltzmann transport equation in the case where all the particles are<br />

the same. In the case of atomic collisions of particles one must take into consideration the generation of<br />

secondary particles resulting from the collisions.<br />

Let there be a number of particles of type j in a volume element of phase space. For example j = p<br />

(protons) and j = n (neutrons). We consider steady state conditions and define the quantities<br />

(i) φj(r, E, Ω) as the flux of the particles of type j.<br />

(ii) σjk( Ω, Ω ′ ,E,E ′ ) the collision cross-section representing processes where particles of type k moving in<br />

direction Ω ′ with energy E ′ produce a type j particle moving in the direction ΩwithenergyE.<br />

(iii) σj(E) =Σs(E,r)+Σa(E,r) the cross-section for type j particles.<br />

The steady state form of the equation (2.5.64) can then be written as<br />

Ω ·∇φj(r, E, Ω)+σj(E)φj(r, E, Ω)<br />

= <br />

<br />

k<br />

σjk( Ω, Ω ′ ,E,E ′ )φk(r, E ′ , Ω ′ )d Ω ′ dE ′<br />

(2.5.76)<br />

where the summation is over all particles k = j.<br />

The Boltzmann transport equation can be represented in many different forms. These various forms<br />

are dependent upon the assumptions made during the derivation, the type of particles, and collision crosssections.<br />

In general the collision cross-sections are dependent upon three components.<br />

(1) Elastic collisions. Here the nucleus is not excited by the collision but energy is transferred by projectile<br />

recoil.<br />

(2) Inelastic collisions. Here some particles are raised to a higher energy state but the excitation energy is<br />

not sufficient to produce any particle emissions due to the collision.<br />

(3) Non-elastic collisions. Here the nucleus is left in an excited state due to the collision processes and<br />

some of its nucleons (protons or neutrons) are ejected. The remaining nucleons interact to form a stable<br />

structure and usually produce a distribution of low energy particles which is isotropic in character.<br />

309


310<br />

Various assumptions can be made concerning the particle flux. The resulting form of Boltzmann’s<br />

equation must be modified to reflect these additional assumptions. As an example, we consider modifications<br />

to Boltzmann’s equation in order to describe the motion of a massive ion moving into a region filled with a<br />

homogeneous material. Here it is assumed that the mean-free path for nuclear collisions is large in comparison<br />

with the mean-free path for ion interaction with electrons. In addition, the following assumptions are made<br />

(i) All collision interactions are non-elastic.<br />

(ii) The secondary particles produced have the same direction as the original particle. This is called the<br />

straight-ahead approximation.<br />

(iii) Secondary particles never have kinetic energies greater than the original projectile that produced them.<br />

(iv) A charged particle will eventually transfer all of its kinetic energy and stop in the media. This stopping<br />

distance is called the range of the projectile. The stopping power Sj(E) = dE<br />

dx represents the energy<br />

loss per unit length traveled in the media and determines the range by the relation dRj 1<br />

dE = Sj(E) or<br />

Rj(E) = E dE<br />

0<br />

′<br />

Sj(E ′ ) . Using the above assumptions Wilson, et.al.1 show that the steady state linearized<br />

Boltzmann equation for homogeneous materials takes on the form<br />

Ω ·∇φj(r, E, Ω) − 1 ∂<br />

Aj ∂E (Sj(E)φj(r, E, Ω)) + σj(E)φj(r, E, Ω)<br />

= <br />

<br />

k=j<br />

dE ′ d Ω ′ σjk( Ω, Ω ′ ,E,E ′ )φk(r, E ′ , Ω ′ )<br />

(2.5.77)<br />

where Aj is the atomic mass of the ion of type j and φj(r, E, Ω) is the flux of ions of type j moving in<br />

the direction ΩwithenergyE.<br />

Observe that in most cases the left-hand side of the Boltzmann equation represents the time rate of<br />

change of a distribution type function in a phase space while the right-hand side of the Boltzmann equation<br />

represents the time rate of change of this distribution function within a volume element of phase space due<br />

to scattering and absorption collision processes.<br />

Boltzmann Equation for gases<br />

Consider the Boltzmann equation in terms of a particle distribution function f(r, V,t)whichcanbe<br />

written as <br />

∂<br />

∂t + V ·∇r + F<br />

m ·∇<br />

<br />

V f(r, V,t)=DCf(r, V,t) (2.5.78)<br />

for a single species of gas particles where there is only scattering and no absorption of the particles. An<br />

element of volume in phase space (x, y, z, Vx,Vy,Vz) can be thought of as a volume element dτ = dxdydz<br />

for the spatial elements together with a volume element dτv = dVxdVydVz for the velocity elements. These<br />

elements are centered at position r and velocity V at time t. In phase space a constant velocity V1 can be<br />

thought of as a sphere since V 2<br />

1<br />

= V 2<br />

x<br />

2 2 + Vy + Vz . The phase space volume element dτdτv changes with time<br />

since the position r and velocity V change with time. The position vector r changes because of velocity<br />

1John W. Wilson, Lawrence W. Townsend, Walter Schimmerling, Govind S. Khandelwal, Ferdous Kahn,<br />

John E. Nealy, Francis A. Cucinotta, Lisa C. Simonsen, Judy L. Shinn, and John W. Norbury, Transport<br />

Methods and Interactions for Space Radiations, NASA Reference Publication <strong>12</strong>57, December 1991.


and the velocity vector changes because of the acceleration F<br />

m .Heref(r, V,t)dτdτv represents the expected<br />

number of particles in the phase space element dτdτv at time t.<br />

Assume there are no collisions, then each of the gas particles in a volume element of phase space centered<br />

at position r and velocity V1 move during a time interval dt to a phase space element centered at position<br />

r + V1dt and V1 + F<br />

mdt. If there were no loss or gains of particles, then the number of particles must be<br />

conserved and so these gas particles must move smoothly from one element of phase space to another without<br />

any gains or losses of particles. Because of scattering collisions in dτ many of the gas particles move into or<br />

out of the velocity range V1 to V1 + d V1. These collision scattering processes are denoted by the collision<br />

operator DCf(r, V,t) in the Boltzmann equation.<br />

Consider two identical gas particles which experience a binary collision. Imagine that particle 1with<br />

velocity V1 collides with particle 2 having velocity V2. Denote by σ(V1 → V ′<br />

1 , V2 → V ′<br />

2 ) dτV1dτV2 the<br />

conditional probability that particle 1is scattered from velocity V1 to between V ′<br />

1 and V ′<br />

1 + d V ′<br />

1 and the<br />

struck particle 2 is scattered from velocity V2 to between V ′<br />

2 and V ′<br />

2 + d V ′<br />

2 . We will be interested in collisions<br />

ofthetype( V ′<br />

1 , V ′<br />

2 ) → ( V1, V2) for a fixed value of V1 as this would represent the number of particles scattered<br />

into dτV1. Also of interest are collisions of the type ( V1, V2) → ( V ′<br />

1, V ′<br />

2) for a fixed value V1 as this represents<br />

particles scattered out of dτV1. Imagine a gas particle in dτ with velocity V ′<br />

1 subjected to a beam of particles<br />

with velocities V ′<br />

2. The incident flux on the element dτdτV ′<br />

1 is | V ′<br />

1 − V ′<br />

2|f(r, V ′<br />

2,t)dτV ′<br />

2<br />

and hence<br />

σ( V1 → V ′<br />

1, V2 → V ′<br />

2) dτV1dτV2dt | V ′<br />

1 − V ′<br />

2|f(r, V ′<br />

2,t) dτV ′<br />

2<br />

(2.5.79)<br />

represents the number of collisions, in the time interval dt, which scatter from V ′<br />

1 to between V1 and V1 + d V1<br />

as well as scattering V ′<br />

2 to between V2 and V2 + d V2. Multiply equation (2.5.79) by the density of particles<br />

in the element dτdτV ′<br />

1 and integrate over all possible initial velocities V ′<br />

1 , V ′<br />

2 and final velocities V2 not equal<br />

to V1. This gives the number of particles in dτ which are scattered into dτV1dt as<br />

<br />

Nsin = dτdτV1dt<br />

dτV2dτV ′<br />

2<br />

dτV ′<br />

1 σ( V ′<br />

1 → V1, V ′<br />

2 → V2)| V ′<br />

1 − V ′<br />

2 |f(r, V ′<br />

1 ,t)f(r, V ′<br />

2<br />

In a similar manner the number of particles in dτ which are scattered out of dτV1dt is<br />

Nsout = dτdτV1dtf(r, <br />

V1,t) dτV2<br />

Let<br />

dτV ′<br />

2<br />

,t). (2.5.80)<br />

dτV ′<br />

1 σ( V ′<br />

1 → V1, V ′<br />

2 → V2)| V2 − V1|f(r, V2,t). (2.5.81)<br />

W ( V ′<br />

1 → V1, V ′<br />

2 → V2) =| V1 − V2| σ( V ′<br />

1 → V1, V ′<br />

2 → V2) (2.5.82)<br />

define a symmetric scattering kernel and use the relation DCf(r, V,t)=Nsin − Nsout to represent the<br />

Boltzmann equation for gas particles in the form<br />

<br />

∂<br />

∂t + V ·∇r + F<br />

m ·∇ <br />

V<br />

f(r, V1,t)=<br />

<br />

(2.5.83)<br />

dτ V ′ 1<br />

dτ V ′ 2<br />

dτV2 W ( V1 → V ′<br />

1 , V2 → V ′<br />

2 ) f(r, V ′<br />

1 ,t)f(r, V ′<br />

2 ,t) − f(r, V1,t)f(r, V2,t) .<br />

Take the moment of the Boltzmann equation (2.5.83) with respect to an arbitrary function φ( V1). That<br />

is, multiply equation (2.5.83) by φ( V1) and then integrate over all elements of velocity space dτV1. Define<br />

the following averages and terminology:<br />

311


3<strong>12</strong><br />

• The particle density per unit volume<br />

<br />

n = n(r, t) = dτV f(r, V,t)=<br />

where ρ = nm is the mass density.<br />

• The mean velocity<br />

V1 = V = 1<br />

+∞<br />

n<br />

−∞<br />

For any quantity Q = Q(V1) define the barred quantity<br />

Q = Q(r, t) = 1<br />

<br />

n(r, t)<br />

+∞<br />

−∞<br />

f(r, V,t)dVxdVydVz<br />

V1f(r, V1,t)dV1xdV1ydV1z<br />

Q( V )f(r, V,t) dτV = 1<br />

+∞<br />

n<br />

−∞<br />

(2.5.84)<br />

Q( V )f(r, V,t)dVxdVydVz. (2.5.85)<br />

Further, assume that F<br />

m is independent of V , then the moment of equation (2.5.83) produces the result<br />

∂ <br />

nφ +<br />

∂t<br />

3<br />

i=1<br />

∂<br />

∂xi <br />

nV1iφ − n<br />

3<br />

i=1<br />

Fi<br />

m<br />

∂φ<br />

∂V1i<br />

=0 (2.5.86)<br />

known as the Maxwell transfer equation. The first term in equation (2.5.86) follows from the integrals<br />

∂f(r, V1,t)<br />

∂t<br />

φ( V1)dτV1 = ∂<br />

<br />

∂t<br />

f(r, V1,t)φ( V1) dτV1 = ∂<br />

(nφ) (2.5.87)<br />

∂t<br />

where differentiation and integration have been interchanged. The second term in equation (2.5.86) follows<br />

from the integral<br />

<br />

V1∇rf φ( <br />

3<br />

∂f<br />

V1)dτV1 = V1i φdτV1<br />

∂xi i=1<br />

3 ∂<br />

=<br />

∂xi <br />

<br />

V1iφf dτV1<br />

(2.5.88)<br />

=<br />

∂V1i<br />

i=1<br />

3<br />

i=1<br />

∂<br />

∂xi <br />

nV1iφ .<br />

The third term in equation (2.5.86) is obtained from the following integral where integration by parts is<br />

employed<br />

<br />

F<br />

m ∇<br />

<br />

fφdτV1 =<br />

V1<br />

3<br />

<br />

Fi ∂f<br />

φdτV1<br />

m ∂V1i<br />

i=1<br />

+∞ 3<br />

<br />

Fi ∂f<br />

= φ<br />

dV1xdV1ydV1y<br />

m ∂V1i<br />

−∞ i=1<br />

<br />

∂ Fi<br />

= −<br />

∂V1i m φ<br />

<br />

fdτV1<br />

= −n ∂<br />

<br />

Fi<br />

m φ<br />

<br />

= − Fi ∂φ<br />

m<br />

(2.5.89)<br />

∂V1i


since Fi does not depend upon V1 and f(r, V,t) equals zero for Vi equal to ±∞. The right-hand side of<br />

equation (2.5.86) represents the integral of (DCf)φ over velocity space. This integral is zero because of<br />

the symmetries associated with the right-hand side of equation (2.5.83). Physically, the integral of (Dcf)φ<br />

over velocity space must be zero since collisions with only scattering terms cannot increase or decrease the<br />

number of particles per cubic centimeter in any element of phase space.<br />

In equation (2.5.86) we write the velocities V1i in terms of the mean velocities (u, v, w) and random<br />

velocities (Ur,Vr,Wr) with<br />

V11 = Ur + u, V<strong>12</strong> = Vr + v, V13 = Wr + w (2.5.90)<br />

or V1 = Vr + V with V1 = Vr + V = V since Vr = 0 (i.e. the average random velocity is zero.) For<br />

future reference we write equation (2.5.86) in terms of these random velocities and the material derivative.<br />

Substitution of the velocities from equation (2.5.90) in equation (2.5.86) gives<br />

or<br />

∂(nφ)<br />

∂t<br />

Observe that<br />

∂<br />

<br />

+ n(Ur + u)φ +<br />

∂x<br />

∂<br />

<br />

n(Vr + v)φ +<br />

∂y<br />

∂<br />

<br />

<br />

n(Wr + w)φ<br />

∂z<br />

∂(nφ) ∂ ∂ ∂ <br />

+ nuφ + nvφ + nwφ<br />

∂t ∂x ∂y ∂z<br />

+ ∂ <br />

nUrφ<br />

∂x<br />

+ ∂ <br />

nVrφ<br />

∂y<br />

+ ∂ <br />

nWrφ<br />

∂z<br />

− n<br />

nuφ =<br />

+∞<br />

−∞<br />

− n<br />

3<br />

i=1<br />

Fi<br />

m<br />

3<br />

i=1<br />

∂φ<br />

Fi<br />

m<br />

∂V1i<br />

∂φ<br />

∂V1i<br />

=0.<br />

=0 (2.5.91)<br />

(2.5.92)<br />

uφf(r, V,t)dVxdVydVz = nuφ (2.5.93)<br />

and similarly nvφ = nvφ, nwφ = nwφ. This enables the equation (2.5.92) to be written in the form<br />

n ∂φ<br />

∂t<br />

+ nu∂φ<br />

∂x<br />

<br />

∂n<br />

+ φ<br />

∂t<br />

∂φ ∂φ<br />

+ nv + nw<br />

∂y ∂z<br />

∂ ∂ ∂<br />

+ (nu)+ (nv)+<br />

∂x ∂y ∂z (nw)<br />

+ ∂ <br />

nUrφ<br />

∂x<br />

+ ∂ <br />

nVrφ<br />

∂y<br />

+ ∂ <br />

nWrφ<br />

∂z<br />

− n<br />

<br />

3<br />

i=1<br />

Fi<br />

m<br />

∂φ<br />

∂V1i<br />

=0.<br />

(2.5.94)<br />

The middle bracketed sum in equation (2.5.94) is recognized as the continuity equation when multiplied by<br />

m and hence is zero. The moment equation (2.5.86) now has the form<br />

n Dφ<br />

Dt<br />

∂ <br />

+ nUrφ<br />

∂x<br />

+ ∂ <br />

nVrφ<br />

∂y<br />

+ ∂ <br />

nWrφ<br />

∂z<br />

− n<br />

3<br />

i=1<br />

Fi<br />

m<br />

∂φ<br />

∂V1i<br />

=0. (2.5.95)<br />

Note that from the equations (2.5.86) or (2.5.95) one can derive the basic equations of fluid flow from<br />

continuum mechanics developed earlier. We consider the following special cases of the Maxwell transfer<br />

equation.<br />

313


314<br />

(i) In the special case φ = m the equation (2.5.86) reduces to the continuity equation for fluids. That is,<br />

equation (2.5.86) becomes<br />

∂<br />

∂t (nm)+∇·(nm V1) =0 (2.5.96)<br />

which is the continuity equation<br />

∂ρ<br />

∂t + ∇·(ρ V )=0 (2.5.97)<br />

where ρ is the mass density and V is the mean velocity defined earlier.<br />

(ii) In the special case φ = m V1 is momentum, the equation (2.5.86) reduces to the momentum equation<br />

for fluids. To show this, we write equation (2.5.86) in terms of the dyadic V1 V1 in the form<br />

∂<br />

<br />

nm<br />

∂t<br />

<br />

V1 + ∇·(nm V1 V1) − n F =0 (2.5.98)<br />

or<br />

∂<br />

<br />

ρ(<br />

∂t<br />

Vr + <br />

V ) + ∇·(ρ( Vr + V )( Vr + V )) − n F =0. (2.5.99)<br />

Let σ = −ρ Vr Vr denote a stress tensor which is due to the random motions of the gas particles and<br />

write equation (2.5.99) in the form<br />

ρ ∂ V<br />

∂t + V ∂ρ<br />

∂t + ρ V (∇· V )+ V (∇·(ρ V )) −∇·σ − n F =0. (2.5.100)<br />

The term <br />

∂ρ<br />

V ∂t + ∇·(ρ <br />

V ) = 0 because of the continuity equation and so equation (2.5.100) reduces<br />

to the momentum equation <br />

∂<br />

ρ<br />

V<br />

∂t + V ∇· <br />

V = n F + ∇·σ. (2.5.101)<br />

For F = q E + q V × B + mb,whereqischarge, E and B are electric and magnetic fields, and b is a<br />

body force per unit mass, together with<br />

σ =<br />

3<br />

i=1 j=1<br />

the equation (2.5.101) becomes the momentum equation<br />

3<br />

(−pδij + τij)eiej<br />

(2.5.102)<br />

ρ D V<br />

Dt = ρ b −∇p + ∇·τ + nq( E + V × B). (2.5.103)<br />

In the special case were E and B vanish, the equation (2.5.103) reduces to the previous momentum<br />

equation (2.5.25) .<br />

(iii) In the special case φ = m<br />

2 V1 · V1 = m 2<br />

2 (V11 +V 2<br />

<strong>12</strong> +V 2<br />

13) is the particle kinetic energy, the equation (2.5.86)<br />

simplifies to the energy equation of fluid mechanics. To show this we substitute φ into equation (2.5.95)<br />

and simplify. Note that<br />

φ = m<br />

<br />

(Ur + u)<br />

2<br />

2 + (Vr + v) 2 + (Wr + w) 2<br />

<br />

φ = m<br />

<br />

U<br />

2<br />

2 r + V 2<br />

r + W 2 r + u2 + v 2 + w 2<br />

(2.5.104)


since uUr = vVr = wWr =0. Let V 2 = u 2 + v 2 + w 2 and C 2 r = U 2 r<br />

(2.5.104) in the form<br />

Also note that<br />

and that<br />

nUrφ = nm<br />

2<br />

= nm<br />

2<br />

nVrφ = nm<br />

2<br />

nWrφ = nm<br />

2<br />

φ = m<br />

2<br />

+ V 2<br />

r + W 2 r<br />

and write equation<br />

<br />

C 2 r + V 2<br />

. (2.5.105)<br />

<br />

Ur(Ur + u) 2 + Ur(Vr + v) 2 + Ur(Wr + w) 2<br />

<br />

<br />

UrC 2 r<br />

2 + uU 2 <br />

r + vUrVr + wUrWr<br />

<br />

VrC 2 r + uVrUr + vV 2<br />

r + wVrWr<br />

<br />

WrC 2 r + uWrUr + vWrVr + wW 2 r<br />

<br />

<br />

(2.5.106)<br />

(2.5.107)<br />

(2.5.108)<br />

are similar results.<br />

We use ∂<br />

∂V1i (φ) =mV1i together with the previous results substituted into the equation (2.5.95), and<br />

find that the Maxwell transport equation can be expressed in the form<br />

ρ D<br />

<br />

C<br />

Dt<br />

2 <br />

2<br />

r V<br />

+ = −<br />

2 2<br />

∂<br />

<br />

ρ[uU<br />

∂x<br />

2 <br />

r + vUrVr + wUrWr]<br />

− ∂<br />

<br />

ρ[uVrUr + vV<br />

∂y<br />

2<br />

<br />

r + wVrWr]<br />

− ∂<br />

<br />

ρ[uWrUr + vWrVr + wW<br />

∂z<br />

2 r ]<br />

<br />

− ∂<br />

<br />

∂x<br />

ρ UrC 2 <br />

r<br />

2<br />

− ∂<br />

<br />

∂y<br />

ρ VrC 2 <br />

r<br />

2<br />

− ∂<br />

<br />

∂z<br />

ρ WrC 2 <br />

r<br />

2<br />

+ n F · (2.5.109)<br />

V.<br />

Compare the equation (2.5.109) with the energy equation (2.5.48)<br />

ρ De<br />

2 D V<br />

+ ρ = ∇(σ ·<br />

Dt Dt 2<br />

V ) −∇·q + ρb · V (2.5.110)<br />

where the internal heat energy has been set equal to zero. Let e = C2 r<br />

2<br />

random motion of the gas particles, F = mb,andlet ∇·q = − ∂<br />

∂x<br />

= − ∂<br />

∂x<br />

<br />

ρ UrC 2 r<br />

2<br />

<br />

k ∂T<br />

∂x<br />

<br />

− ∂<br />

<br />

∂y<br />

<br />

− ∂<br />

<br />

k<br />

∂y<br />

∂T<br />

∂y<br />

<br />

−<br />

2<br />

∂<br />

<br />

∂z<br />

<br />

− ∂<br />

<br />

k<br />

∂z<br />

∂T<br />

<br />

∂z<br />

ρ VrC 2 r<br />

denote the internal energy due to<br />

ρ WrC 2 r<br />

2<br />

<br />

(2.5.111)<br />

represent the heat conduction terms due to the transport of particle energy mC2 r<br />

2 by way of the random<br />

particle motion. The remaining terms are related to the rate of change of work and surface stresses giving<br />

− ∂<br />

<br />

ρ[uU<br />

∂x<br />

2 r + vUrVr<br />

<br />

+ wUrWr]<br />

− ∂<br />

<br />

ρ[uVrUr + vV<br />

∂y<br />

2<br />

<br />

r + wVrWr]<br />

− ∂<br />

<br />

ρ[uWrUr + vWrVr + wW<br />

∂z<br />

2 r ]<br />

<br />

= ∂<br />

∂x (uσxx + vσxy + wσxz)<br />

= ∂<br />

∂y (uσyx + vσyy + wσyz)<br />

= ∂<br />

∂z (uσzx + vσzy + wσzz) .<br />

(2.5.1<strong>12</strong>)<br />

315


316<br />

This gives the stress relations due to random particle motion<br />

σxx = − ρU 2 r<br />

σxy = − ρUrVr<br />

σxz = − ρUrWr<br />

σyx = − ρVrUr<br />

σyy = − ρV 2<br />

r<br />

σyz = − ρVrWr<br />

σzx = − ρWrUr<br />

σzy = − ρWrVr<br />

σzz = − ρW 2 r .<br />

(2.5.113)<br />

The Boltzmann equation is a basic macroscopic model used for the study of individual particle motion<br />

where one takes into account the distribution of particles in both space, time and energy. The Boltzmann<br />

equation for gases assumes only binary collisions as three-body or multi-body collisions are assumed to<br />

rarely occur. Another assumption used in the development of the Boltzmann equation is that the actual<br />

time of collision is thought to be small in comparison with the time between collisions. The basic problem<br />

associated with the Boltzmann equation is to find a velocity distribution, subject to either boundary and/or<br />

initial conditions, which describes a given gas flow.<br />

The continuum equations involve trying to obtain the macroscopic variables of density, mean velocity,<br />

stress, temperature and pressure which occur in the basic equations of continuum mechanics considered<br />

earlier. Note that the moments of the Boltzmann equation, derived for gases, also produced these same<br />

continuum equations and so they are valid for gases as well as liquids.<br />

In certain situations one can assume that the gases approximate a Maxwellian distribution<br />

f(r, <br />

m<br />

3/2 <br />

V,t) ≈ n(r, t)<br />

exp −<br />

2πkT<br />

m<br />

2kT V · <br />

V<br />

(2.5.114)<br />

thereby enabling the calculation of the pressure tensor and temperature from statistical considerations.<br />

In general, one can say that the Boltzmann integral-differential equation and the Maxwell transfer<br />

equation are two important formulations in the kinetic theory of gases. The Maxwell transfer equation<br />

depends upon some gas-particle property φ which is assumed to be a function of the gas-particle velocity.<br />

The Boltzmann equation depends upon a gas-particle velocity distribution function f which depends upon<br />

position r, velocity V and time t. These formulations represent two distinct and important viewpoints<br />

considered in the kinetic theory of gases.


EXERCISE 2.5<br />

◮ 1. Let p = p(x, y, z), [dyne/cm2 ] denote the pressure at a point (x, y, z) inafluidmediumatrest<br />

(hydrostatics), and let ∆V denote an element of fluid volume situated at this point as illustrated in the<br />

figure 2.5-5.<br />

Figure 2.5-5. Pressure acting on a volume element.<br />

(a) Show that the force acting on the face ABCD is p(x, y, z)∆y∆z ê1.<br />

(b) Show that the force acting on the face EFGH is<br />

<br />

−p(x +∆x, y, z)∆y∆z ê1 = − p(x, y, z)+ ∂p<br />

∂x ∆x + ∂2p ∂x2 (∆x) 2 <br />

+ ··· ∆y∆z ê1.<br />

2!<br />

(c) In part (b) neglect terms with powers of ∆x greater than or equal to 2 and show that the resultant force<br />

in the x-direction is − ∂p<br />

∆x∆y∆z ê1.<br />

∂x<br />

(d) What has been done in the x-direction can also be done in the y and z-directions. Show that the<br />

resultant forces in these directions are − ∂p<br />

∂y ∆x∆y∆z ê2 and − ∂p<br />

∂z ∆x∆y∆z ê3. (e) Show that −∇p =<br />

<br />

∂p<br />

−<br />

∂x ê1 + ∂p<br />

∂y ê2 + ∂p<br />

∂z ê3<br />

<br />

is the force per unit volume acting at the point (x, y, z) of the fluid medium.<br />

◮ 2. Follow the example of exercise 1above but use cylindrical coordinates and find the force per unit volume<br />

at a point (r, θ, z). Hint: An element of volume in cylindrical coordinates is given by ∆V = r∆r∆θ∆z.<br />

◮ 3. Follow the example of exercise 1above but use spherical coordinates and find the force per unit volume<br />

at a point (ρ, θ, φ). Hint: An element of volume in spherical coordinates is ∆V = ρ2 sin θ∆ρ∆θ∆φ.<br />

◮ 4. Show that if the density ϱ = ϱ(x, y, z, t) is a constant, then v r ,r =0.<br />

◮ 5. Assume that λ∗ and µ ∗ are zero. Such a fluid is called a nonviscous or perfect fluid. (a) Show the<br />

Cartesian equations describing conservation of linear momentum are<br />

∂u ∂u ∂u ∂u<br />

+ u + v + w<br />

∂t ∂x ∂y ∂z = bx − 1 ∂p<br />

ϱ ∂x<br />

∂v ∂v ∂v ∂v<br />

+ u + v + w<br />

∂t ∂x ∂y ∂z = by − 1 ∂p<br />

ϱ ∂y<br />

∂w ∂w ∂w ∂w<br />

+ u + v + w<br />

∂t ∂x ∂y ∂z = bz − 1 ∂p<br />

ϱ ∂z<br />

where (u, v, w) are the physical components of the fluid velocity. (b) Show that the continuity equation can<br />

be written<br />

∂ϱ ∂ ∂ ∂<br />

+ (ϱu)+ (ϱv)+ (ϱw) =0<br />

∂t ∂x ∂y ∂z<br />

317


318<br />

◮ 6. Assume λ∗ = µ ∗ = 0 so that the fluid is ideal or nonviscous. Use the results given in problem 5 and<br />

make the following additional assumptions:<br />

• The density is constant and so the fluid is incompressible.<br />

• The body forces are zero.<br />

• Steady state flow exists.<br />

• Only two dimensional flow in the x-yplane is considered such that u = u(x, y), v = v(x, y) and<br />

w =0. (a) Employ the above assumptions and simplify the equations in problem 5 and verify the<br />

results<br />

u ∂u ∂u 1 ∂p<br />

+ v +<br />

∂x ∂y ϱ ∂x =0<br />

u ∂v ∂v 1 ∂p<br />

+ v +<br />

∂x ∂y ϱ ∂y =0<br />

∂u ∂v<br />

+<br />

∂x ∂y =0<br />

(b) Make the additional assumption that the flow is irrotational and show that this assumption<br />

produces the results<br />

∂v ∂u<br />

− =0 and<br />

∂x ∂y<br />

1<br />

2<br />

2 2<br />

u + v + 1<br />

p = constant.<br />

ϱ<br />

(c) Point out the Cauchy-Riemann equations and Bernoulli’s equation in the above set of equations.<br />

◮ 7. Assume the body forces are derivable from a potential function φ such that bi = −φ,i. Show that for an<br />

ideal fluid with constant density the equations of fluid motion can be written in either of the forms<br />

∂v r<br />

∂t + vr ,sv s = − 1<br />

ϱ grm p,m − g rm φ,m or<br />

∂vr<br />

∂t + vr,sv s = − 1<br />

ϱ p,r − φ,r<br />

◮ 8. The vector identities ∇ 2 v = ∇ (∇·v) −∇×(∇×v) and (v ·∇) v = 1<br />

∇ (v · v) − v × (∇×v) are<br />

2<br />

used to express the Navier-Stokes-Duhem equations in alternate forms involving the vorticity Ω = ∇×v.<br />

(a) Use Cartesian tensor notation and derive the above identities. (b) Show the second identity can be written<br />

in generalized coordinates as v j v m ,j = g mj v k vk,j − ɛ mnp ɛ ijk gpivnvk,j. Hint: Show that ∂v2<br />

∂xj =2vkvk,j. ◮ 9. Use problem 8 and show that the results in problem 7 can be written<br />

or<br />

∂vr ∂t − ɛrnp rm ∂<br />

Ωpvn = −g<br />

∂vi<br />

∂t − ɛijkv j Ω k = − ∂<br />

∂x i<br />

∂x m<br />

<br />

p v2<br />

+ φ +<br />

ϱ 2<br />

<br />

p v2<br />

+ φ +<br />

ϱ 2<br />

◮ 10. In terms of physical components, show that in generalized orthogonal coordinates, for i = j, therate<br />

of deformation tensor Dij can be written D(ij) = 1<br />

<br />

hi ∂<br />

2 hj ∂xj <br />

v(i)<br />

+<br />

hi<br />

hj ∂<br />

hi ∂xi <br />

v(j)<br />

, no summations<br />

hj<br />

and for i = j there results D(ii) = 1 ∂v(i) v(i)<br />

−<br />

hi ∂xi h2 3 ∂hi 1<br />

+ v(k)<br />

i ∂xi hihk<br />

k=1<br />

∂hi<br />

, no summations. (Hint: See<br />

∂xk Problem 17 Exercise 2.1.)


Figure 2.5-6. Plane Couette flow<br />

◮ 11. Find the physical components of the rate of deformation tensor Dij in Cartesian coordinates. (Hint:<br />

See problem 10.)<br />

◮ <strong>12</strong>. Find the physical components of the rate of deformation tensor in cylindrical coordinates. (Hint: See<br />

problem 10.)<br />

◮ 13. (Plane Couette flow)<br />

in the figure 2.5-6.<br />

Assume a viscous fluid with constant density is between two plates as illustrated<br />

(a) Define ν = µ∗<br />

ϱ as the kinematic viscosity and show the equations of fluid motion can be written<br />

∂v i<br />

∂t + vi ,s vs = − 1<br />

ϱ gim p,m + νg jm v i ,mj + gij bj, i =1, 2, 3<br />

(b) Let v =(u, v, w) denote the physical components of the fluid flow and make the following assumptions<br />

• u = u(y), v= w =0<br />

• Steady state flow exists<br />

• The top plate, with area A, isadistanceℓabovethe bottom plate. The bottom plate is fixed and<br />

a constant force F is applied to the top plate to keep it moving with a velocity u0 = u(ℓ).<br />

• p and ϱ are constants<br />

• The body force components are zero.<br />

Find the velocity u = u(y)<br />

(c) Show the tangential stress exerted by the moving fluid is F<br />

A = σ21<br />

∗ u0<br />

= σxy = σyx = µ . This<br />

ℓ<br />

example illustrates that the stress is proportional to u0 and inversely proportional to ℓ.<br />

◮ 14. In the continuity equation make the change of variables<br />

t = t ϱ<br />

, ϱ = , v =<br />

τ ϱ0<br />

v<br />

v0<br />

, x = x<br />

L<br />

y z<br />

, y = , z =<br />

L L<br />

and write the continuity equation in terms of the barred variables and the Strouhal parameter.<br />

◮ 15. (Plane Poiseuille flow) Consider two flat plates parallel to one another as illustrated in the figure<br />

2.5-7. One plate is at y = 0 and the other plate is at y =2ℓ. Letv =(u, v, w) denote the physical components<br />

of the fluid velocity and make the following assumptions concerning the flow The body forces are zero. The<br />

derivative ∂p<br />

∂x = −p0 is a constant and ∂p ∂p<br />

= =0. The velocity in the x-direction is a function of y only<br />

∂y ∂z<br />

319


320<br />

Figure 2.5-7. Plane Poiseuille flow<br />

with u = u(y) andv = w = 0 with boundary values u(0) = u(2ℓ) =0. The density is constant and ν = µ ∗ /ϱ<br />

is the kinematic viscosity.<br />

(a) Show the equation of fluid motion is ν d2u p0<br />

+ =0, u(0) = u(2ℓ) =0<br />

dy2 ϱ<br />

(b) Find the velocity u = u(y) and find the maximum velocity in the x-direction. (c) Let M denote the<br />

mass flow rate across the plane x = x0 = constant, ,where0≤ y ≤ 2ℓ, and 0 ≤ z ≤ 1.<br />

Show that M = 2<br />

3µ ∗ ϱp0ℓ 3 . Note that as µ ∗ increases, M decreases.<br />

◮ 16. The heat equation (or diffusion equation) can be expressed div ( k grad u)+H = ∂(δcu)<br />

, where c is the<br />

∂t<br />

specific heat [cal/gm C], δ is the volume density [gm/cm 3 ], H is the rate of heat generation [cal/sec cm 3 ], u<br />

is the temperature [C], k is the thermal conductivity [cal/sec cm C]. Assume constant thermal conductivity,<br />

volume density and specific heat and express the boundary value problem<br />

k ∂2u = δc∂u,<br />

∂x2 ∂t<br />

0


Figure 2.5-8. Rayleigh impulsive flow<br />

where erf and erfc are the error function and complimentary error function respectively. Pick a point on the<br />

line y = y0 =2 √ ν and plot the velocity as a function of time. How does the viscosity effect the velocity of<br />

the fluid along the line y = y0?<br />

◮ 19. Simplify the Navier-Stokes-Duhem equations using the assumption that there is incompressible and<br />

irrotational flow.<br />

◮ 20. Let ζ = λ∗ + 2<br />

3 µ∗ and show the constitutive equations (2.5.21) for fluid motion can be written in the<br />

form<br />

σij = −pδij + µ ∗<br />

<br />

vi,j + vj,i − 2<br />

3 δijvk,k<br />

<br />

+ ζδijvk,k.<br />

◮ 21. (a) Write out the Navier-Stokes-Duhem equation for two dimensional flow in the x-y direction under<br />

the assumptions that<br />

• λ ∗ + 2<br />

3 µ∗ = 0 (This condition is referred to as Stoke’s flow.)<br />

• The fluid is incompressible<br />

• There is a gravitational force b = −g∇ h Hint: Express your answer as two scalar equations<br />

involving the variables v1,v2,h,g,ϱ,p,t,µ ∗ plus the continuity equation. (b) In part (a) eliminate<br />

the pressure and body force terms by cross differentiation and subtraction. (i.e. take the derivative<br />

of one equation with respect to x and take the derivative of the other equation with respect to y<br />

and then eliminate any common terms.) (c) Assume that ω = ω ê3 where ω = 1<br />

<br />

∂v2 ∂v1<br />

− and<br />

2 ∂x ∂y<br />

derive the vorticity-transport equation<br />

dω<br />

dt = ν∇2 ω where<br />

dω<br />

dt<br />

= ∂ω<br />

∂t<br />

∂ω<br />

+ v1<br />

∂x<br />

∂ω<br />

+ v2<br />

∂y .<br />

Hint: The continuity equation makes certain terms zero. (d) Define a stream function ψ = ψ(x, y)<br />

satisfying v1 = ∂ψ<br />

∂y<br />

and v2 = − ∂ψ<br />

and show the continuity equation is identically satisfied.<br />

∂x<br />

Show also that ω = − 1<br />

2 ∇2 ψ and that<br />

∇ 4 ψ = 1<br />

ν<br />

If ν is very large, show that ∇ 4 ψ ≈ 0.<br />

∂∇ 2 ψ<br />

∂t<br />

∂ψ ∂∇<br />

+<br />

∂y<br />

2ψ ∂ψ<br />

−<br />

∂x ∂x<br />

∂∇2 <br />

ψ<br />

.<br />

∂y<br />

321


322<br />

◮ 22. In generalized orthogonal coordinates, show that the physical components of the rate of deformation<br />

stress can be written, for i = j<br />

σ(ij) =µ ∗<br />

<br />

hi ∂<br />

hj ∂xj <br />

v(i)<br />

+<br />

hi<br />

hj ∂<br />

hi ∂xi <br />

v(j)<br />

,<br />

hj<br />

no summation,<br />

and for i = j = k<br />

σ(ii) =−p +2µ ∗<br />

<br />

1<br />

hi<br />

∂v(i) 1<br />

+<br />

∂xi hihj<br />

+ λ∗<br />

<br />

∂<br />

h1h2h3 ∂x1 {h2h3v(1)} + ∂<br />

v(j) ∂hi 1<br />

+ v(k)<br />

∂xj hihk<br />

∂hi<br />

∂xk <br />

∂x2 {h1h3v(2)} + ∂<br />

<br />

{h1h2v(3)} , no summation<br />

∂x3 ◮ 23. Find the physical components for the rate of deformation stress in Cartesian coordinates. Hint: See<br />

problem 22.<br />

◮ 24. Find the physical components for the rate of deformations stress in cylindrical coordinates. Hint: See<br />

problem 22.<br />

◮ 25. Verify the Navier-Stokes equations for an incompressible fluid can be written ˙vi = − 1<br />

ϱ p,i + νvi,mm + bi<br />

where ν = µ∗<br />

ϱ is called the kinematic viscosity.<br />

◮ 26. Verify the Navier-Stokes equations for a compressible fluid with zero bulk viscosity can be written<br />

˙vi = − 1<br />

ϱ p,i + ν<br />

3 vm,mi + νvi,mm + bi with ν = µ∗<br />

ϱ the kinematic viscosity.<br />

◮ 27. The constitutive equation for a certain non-Newtonian Stokesian fluid is σij = −pδij +βDij +γDikDkj.<br />

Assume that β and γ are constants (a) Verify that σij,j = −p,i + βDij,j + γ(DikDkj,j + Dik,jDkj)<br />

(b) Write out the Cauchy equations of motion in Cartesian coordinates. (See page 236).<br />

◮ 28. Let the constitutive equations relating stress and strain for a solid material take into account thermal<br />

stresses due to a temperature T . The constitutive equations have the form eij = 1+ν<br />

E σij − ν<br />

E σkk δij +αT δij<br />

where α is a coefficient of linear expansion for the material and T is the absolute temperature. Solve for the<br />

stress in terms of strains.<br />

◮ 29. Derive equation (2.5.53) and then show that when the bulk coefficient of viscosity is zero, the Navier-<br />

Stokes equations, in Cartesian coordinates, can be written in the conservation form<br />

∂(ϱu)<br />

∂t + ∂(ϱu2 + p − τxx)<br />

+<br />

∂x<br />

∂(ϱuv − τxy)<br />

+<br />

∂y<br />

∂(ϱuw − τxz)<br />

∂z<br />

∂(ϱv) ∂(ϱuv − τxy)<br />

+ +<br />

∂t ∂x<br />

∂(ϱv2 + p − τyy)<br />

+<br />

∂y<br />

∂(ϱvw − τyz)<br />

∂z<br />

∂(ϱw) ∂(ϱuw − τxz)<br />

+ +<br />

∂t ∂x<br />

∂(ϱvw − τyz)<br />

+<br />

∂y<br />

∂(ϱw2 + p − τzz)<br />

∂z<br />

= ϱbx<br />

= ϱby<br />

= ϱbz<br />

where v1 = u,v2 = v,v3 = w and τij = µ ∗ (vi,j + vj,i − 2<br />

3 δijvk,k). Hint: Alternatively, consider 2.5.29 and use<br />

the continuity equation.


◮ 30. Show that for a perfect gas, where λ ∗ = − 2<br />

3 µ∗ and η = µ ∗ is a function of position, the vector form<br />

of equation (2.5.25) is<br />

ϱ Dv<br />

Dt = ϱ b −∇p + 4<br />

3 ∇(η∇·v)+∇(v ·∇η) − v ∇2 η +(∇η) × (∇×v) − (∇·v)∇η −∇×(∇×(ηv))<br />

◮ 31. Derive the energy equation ϱ Dh Dp ∂Q<br />

= + −∇·q +Φ. Hint: Use the continuity equation.<br />

Dt Dt ∂t<br />

◮ 32. Show that in Cartesian coordinates the Navier-Stokes equations of motion for a compressible fluid<br />

can be written<br />

ρ Du<br />

Dt =ρbx − ∂p<br />

<br />

∂ ∗ ∂u<br />

+ 2µ<br />

∂x ∂x ∂x + λ∗∇· <br />

V + ∂<br />

<br />

µ<br />

∂y<br />

∗ ( ∂u ∂v<br />

+<br />

∂y ∂x )<br />

<br />

+ ∂<br />

<br />

µ<br />

∂z<br />

∗ ( ∂w ∂u<br />

+<br />

∂x ∂z )<br />

<br />

ρ Dv<br />

Dt =ρby − ∂p<br />

<br />

∂ ∗ ∂v<br />

+ 2µ<br />

∂y ∂y ∂y + λ∗∇· <br />

V + ∂<br />

<br />

µ<br />

∂z<br />

∗ ( ∂v ∂w<br />

+<br />

∂z ∂y )<br />

<br />

+ ∂<br />

<br />

µ<br />

∂x<br />

∗ ( ∂w ∂w<br />

+<br />

∂y ∂x )<br />

<br />

ρ Dv<br />

Dt =ρbz − ∂p<br />

<br />

∂ ∗ ∂w<br />

+ 2µ<br />

∂z ∂z ∂z + λ∗∇· <br />

V + ∂<br />

<br />

µ<br />

∂x<br />

∗ ( ∂w ∂u<br />

+<br />

∂x ∂z )<br />

<br />

+ ∂<br />

<br />

µ<br />

∂y<br />

∗ ( ∂v ∂w<br />

+<br />

∂z ∂y )<br />

<br />

where (Vx,Vy,Vz) =(u, v, w).<br />

◮ 33. Show that in cylindrical coordinates the Navier-Stokes equations of motion for a compressible fluid<br />

can be written<br />

2 DVr Vθ ϱ − =ϱbr −<br />

Dt r<br />

∂p<br />

<br />

∂ ∗ ∂Vr<br />

+ 2µ<br />

∂r ∂r ∂r + λ∗∇· <br />

V + 1<br />

<br />

∂<br />

µ<br />

r ∂θ<br />

∗ ( 1 ∂Vr ∂Vθ Vθ<br />

+ −<br />

r ∂θ ∂r r )<br />

<br />

+ ∂<br />

<br />

µ<br />

∂z<br />

∗ ( ∂Vr ∂Vz<br />

+<br />

∂z ∂r )<br />

<br />

+ 2µ∗<br />

r (∂Vr<br />

1 ∂Vθ Vr<br />

− −<br />

∂r r ∂θ r )<br />

<br />

DVθ VrVθ<br />

ϱ + =ϱbθ −<br />

Dt r<br />

1<br />

<br />

∂p 1 ∂<br />

+ 2µ<br />

r ∂θ r ∂θ<br />

∗ ( 1 ∂Vθ Vr<br />

+<br />

r ∂θ r )+λ∗∇· <br />

V + ∂<br />

<br />

µ<br />

∂z<br />

∗ ( 1 ∂Vz ∂Vθ<br />

+<br />

r ∂θ ∂z )<br />

<br />

+ ∂<br />

<br />

µ<br />

∂r<br />

∗ ( 1 ∂Vr ∂Vθ Vθ<br />

+ −<br />

r ∂θ ∂r r )<br />

<br />

+ 2µ∗<br />

r (1<br />

∂Vr ∂Vθ Vθ<br />

+ −<br />

r ∂θ ∂r r )<br />

ϱ DVz<br />

Dt =ϱbz − ∂p<br />

<br />

∂ ∗ ∂Vz<br />

+ 2µ<br />

∂z ∂z ∂z + λ∗∇· <br />

V + 1<br />

<br />

∂<br />

µ<br />

r ∂r<br />

∗ r( ∂Vr ∂Vz<br />

+<br />

∂z ∂r )<br />

<br />

+ 1<br />

<br />

∂<br />

µ<br />

r ∂θ<br />

∗ ( 1 ∂Vz ∂Vθ<br />

+<br />

r ∂θ ∂z )<br />

<br />

◮ 34. Show that the dissipation function Φ can be written as Φ = 2µ ∗ DijDij + λ ∗ Θ 2 .<br />

◮ 35. Verify the identities:<br />

(a) ϱ D<br />

Dt (et/ϱ) = ∂et<br />

∂t + ∇·(et V ) (b) ϱ D<br />

Dt (et/ϱ) =ϱ De D 2<br />

+ ϱ V /2 .<br />

Dt Dt<br />

◮ 36. Show that the conservation law for heat flow is given by<br />

∂T<br />

∂t<br />

+ ∇·(Tv − κ∇T )=SQ<br />

where κ is the thermal conductivity of the material, T is the temperature, Jadvection = Tv,<br />

Jconduction = −κ∇T and SQ is a source term. Note that in a solid material there is no flow and so v =0and<br />

323


324<br />

the above equation reduces to the heat equation. Assign units of measurements to each term in the above<br />

equation and make sure the equation is dimensionally homogeneous.<br />

◮ 37. Show that in spherical coordinates the Navier-Stokes equations of motion for a compressible fluid can<br />

be written<br />

ϱ( DVρ<br />

2 Vθ −<br />

Dt + V 2 φ<br />

)=ϱbρ −<br />

ρ<br />

∂p<br />

<br />

∂ ∗ ∂Vρ<br />

+ 2µ<br />

∂ρ ∂ρ ∂ρ + λ∗∇· <br />

V + 1<br />

<br />

∂<br />

µ<br />

ρ ∂θ<br />

∗ (ρ ∂ 1 ∂Vρ<br />

(Vθ/ρ)+<br />

∂ρ ρ ∂θ )<br />

<br />

+ 1<br />

<br />

∂<br />

µ<br />

ρ sin θ ∂φ<br />

∗ 1 ∂Vρ ∂<br />

( + ρ<br />

ρ sin θ ∂φ ∂ρ (Vφ/ρ))<br />

<br />

+ µ∗ ∂Vρ 2 ∂Vθ 4Vρ 2 ∂Vφ<br />

(4 − − −<br />

ρ ∂ρ ρ ∂θ ρ ρ sin θ ∂φ − 2Vθ cot θ<br />

+ ρ cot θ<br />

ρ<br />

∂ cot θ ∂Vρ<br />

(Vθ/ρ)+<br />

∂ρ ρ ∂θ )<br />

ϱ( DVθ VρVθ<br />

+<br />

Dt ρ − V 2 φ cot θ<br />

)=ϱbθ −<br />

ρ<br />

1<br />

∗<br />

∂p 1 ∂ 2µ<br />

+<br />

ρ ∂θ ρ ∂θ ρ (∂Vθ<br />

∂θ + Vρ)+λ ∗ ∇· <br />

V<br />

+ 1<br />

<br />

∂<br />

µ<br />

ρ sin θ ∂φ<br />

∗ sin θ ∂<br />

(<br />

ρ ∂θ (Vφ/ sin θ)+ 1 ∂Vθ<br />

ρ sin θ ∂φ )<br />

<br />

+ ∂<br />

<br />

µ<br />

∂ρ<br />

∗ (ρ ∂ 1 ∂Vρ<br />

(Vθ/ρ)+<br />

∂ρ ρ ∂θ )<br />

<br />

+ µ∗<br />

<br />

1 ∂Vθ 1 ∂Vφ<br />

2 −<br />

ρ ρ ∂θ ρ sin θ ∂φ − Vθ<br />

<br />

cot θ<br />

cot θ +3 ρ<br />

ρ<br />

∂<br />

<br />

1 ∂Vρ<br />

(Vθ/ρ)+<br />

∂ρ ρ ∂θ<br />

<br />

DVφ VφVρ<br />

ϱ +<br />

Dt ρ + VθVφ<br />

<br />

cot θ<br />

= ϱbφ −<br />

ρ<br />

1<br />

<br />

∂p ∂<br />

+ µ<br />

ρ sin θ ∂φ ∂ρ<br />

∗<br />

<br />

1 ∂Vρ ∂<br />

+ ρ<br />

ρ sin θ ∂φ ∂ρ (Vφ/ρ)<br />

<br />

+ 1<br />

∗ ∂ 2µ 1 ∂Vφ<br />

ρ sin θ ∂φ ρ sin θ ∂φ + Vρ<br />

<br />

+ Vθ cot θ + λ ∗ ∇· <br />

V<br />

+ 1<br />

<br />

∂<br />

µ<br />

ρ ∂θ<br />

∗<br />

<br />

sin θ ∂<br />

ρ ∂θ (Vφ/ sin θ)+ 1<br />

<br />

∂Vθ<br />

ρ sin θ ∂φ<br />

+ µ∗<br />

<br />

1 ∂Vρ ∂<br />

3<br />

+ ρ<br />

ρ ρ sin θ ∂φ ∂ρ (Vφ/ρ)<br />

<br />

sin θ ∂<br />

+2cotθ<br />

ρ ∂θ (Vφ/ sin θ)+ 1<br />

<br />

∂Vθ<br />

ρ sin θ ∂φ<br />

◮ 38. Verify all the equations (2.5.28).<br />

◮ 39. Use the conservation of energy equation (2.5.47) together with the momentum equation (2.5.25) to<br />

derive the equation (2.5.48).<br />

◮ 40. Verify the equation (2.5.55).<br />

◮ 41. Consider nonviscous flow and write the 3 linear momentum equations and the continuity equation<br />

and make the following assumptions: (i) The density ϱ is constant. (ii) Body forces are zero. (iii) Steady<br />

state flow only. (iv) Consider only two dimensional flow with non-zero velocity components u = u(x, y) and<br />

v = v(x, y). Show that there results the system of equations<br />

u ∂u ∂u 1 ∂P<br />

∂v 1 ∂P<br />

+ v + =0, u∂v + v +<br />

∂x ∂y ϱ ∂x ∂x ∂y ϱ ∂y =0,<br />

∂u ∂v<br />

+<br />

∂x ∂y =0.<br />

Recognize that the last equation in the above set as one of the Cauchy-Riemann equations that f(z) =u−iv<br />

be an analytic function of a complex variable. Further assume that the fluid flow is irrotational so that<br />

∂v ∂u<br />

1 2 2<br />

− =0. Show that this implies that u + v<br />

∂x ∂y 2<br />

+ P<br />

= Constant. If in addition u and v are derivable<br />

ϱ<br />

from a potential function φ(x, y), such that u = ∂φ<br />

∂φ<br />

∂x and v = ∂y , then show that φ is a harmonic function.<br />

By constructing the conjugate harmonic function ψ(x, y) the complex potential F (z) =φ(x, y)+iψ(x, y) is<br />

such that F ′ (z) =u(x, y) − iv(x, y) andF ′ (z) gives the velocity. The family of curves φ(x, y) =constantare<br />

called equipotential curves and the family of curves ψ(x, y) = constant are called streamlines. Show that<br />

these families are an orthogonal family of curves.


§2.6 ELECTRIC AND MAGNETIC FIELDS<br />

Introduction<br />

In electromagnetic theory the mks system of units and the Gaussian system of units are the ones most<br />

often encountered. In this section the equations will be given in the mks system of units. If you want the<br />

equations in the Gaussian system of units make the replacements given in the column 3 of Table 1.<br />

MKS<br />

symbol<br />

Table 1. MKS AND GAUSSIAN UNITS<br />

MKS<br />

units<br />

Replacement<br />

symbol<br />

GAUSSIAN<br />

units<br />

E (Electric field) volt/m E statvolt/cm<br />

B (Magnetic field) weber/m 2 B<br />

c<br />

D (Displacement field) coulomb/m 2 D<br />

4π<br />

H (Auxiliary Magnetic field) ampere/m c H<br />

4π<br />

gauss<br />

statcoulomb/cm 2<br />

oersted<br />

J (Current density) ampere/m 2 J statampere/cm 2<br />

A (Vector potential) weber/m<br />

A<br />

c<br />

gauss-cm<br />

V (Electric potential) volt V statvolt<br />

ɛ (Dielectric constant)<br />

µ (Magnetic permeability)<br />

Electrostatics<br />

A basic problem in electrostatic theory is to determine the force F on a charge Q placed a distance r<br />

from another charge q. The solution to this problem is Coulomb’s law<br />

F = 1 qQ<br />

er<br />

(2.6.1)<br />

4πɛ0 r2 where q, Q aremeasuredincoulombs,ɛ0 =8.85 × 10 −<strong>12</strong> coulomb 2 /N · m 2 is called the permittivity in a<br />

vacuum, r is in meters, [ F ] has units of Newtons and er is a unit vector pointing from q to Q if q, Q have<br />

thesamesignorpointingfromQto q if q, Q are of opposite sign. The quantity E = F/Q is called the<br />

electric field produced by the charges. In the special case Q =1,wehave E = F and so Q = 1 is called<br />

a test charge. This tells us that the electric field at a point P can be viewed as the force per unit charge<br />

exerted on a test charge Q placed at the point P. The test charge Q is always positive and so is repulsed if<br />

q is positive and attracted if q is negative.<br />

The electric field associated with many charges is obtained by the principal of superposition. For<br />

example, let q1,q2,...,qn denote n-charges having respectively the distances r1,r2,...,rn from a test charge<br />

Q placed at a point P. The force exerted on Q is<br />

F = F1 + F2 + ···+ Fn<br />

F = 1<br />

<br />

q1Q<br />

4πɛ0 r2 er1 +<br />

1<br />

q2Q<br />

r2 er2 + ···+<br />

2<br />

qnQ<br />

r2 <br />

ern<br />

n<br />

(2.6.2)<br />

n<br />

or E = E(P <br />

F 1 qi<br />

)= = eri<br />

Q 4πɛ0<br />

r<br />

i=1<br />

2 i<br />

ɛ<br />

4π<br />

4πµ<br />

c 2<br />

325


326<br />

where E = E(P ) is the electric field associated with the system of charges. The equation (2.6.2) can be generalized<br />

to other situations by defining other types of charge distributions. We introduce a line charge density<br />

λ ∗ ,(coulomb/m), a surface charge density µ ∗ ,(coulomb/m 2 ), a volume charge density ρ ∗ ,(coulomb/m 3 ),<br />

then we can calculate the electric field associated with these other types of charge distributions. For example,<br />

if there is a charge distribution λ ∗ = λ ∗ (s) along a curve C, wheres is an arc length parameter, then we<br />

would have<br />

E(P )= 1<br />

<br />

er<br />

4πɛ0 C r2 λ∗ds (2.6.3)<br />

as the electric field at a point P due to this charge distribution. The integral in equation (2.6.3) being a<br />

line integral along the curve C and where ds is an element of arc length. Here equation (2.6.3) represents a<br />

continuous summation of the charges along the curve C. For a continuous charge distribution over a surface<br />

S, the electric field at a point P is<br />

E(P )= 1<br />

<br />

er<br />

4πɛ0 S r2 µ∗dσ (2.6.4)<br />

where dσ represents an element of surface area on S. Similarly, if ρ∗ represents a continuous charge distribution<br />

throughout a volume V , then the electric field is represented<br />

E(P )= 1<br />

4πɛ0<br />

<br />

V<br />

er<br />

r 2 ρ∗ dτ (2.6.5)<br />

where dτ is an element of volume. In the equations (2.6.3), (2.6.4), (2.6.5) we let (x, y, z) denote the position<br />

of the test charge and let (x ′ ,y ′ ,z ′ ) denote a point on the line, on the surface or within the volume, then<br />

r =(x − x ′ ) e1 +(y − y ′ ) e2 +(z − z ′ ) e3<br />

(2.6.6)<br />

represents the distance from the point P to an element of charge λ ∗ ds, µ ∗ dσ or ρ ∗ dτ with r = |r| and er = r<br />

r .<br />

If the electric field is conservative, then ∇× E = 0, and so it is derivable from a potential function V<br />

by taking the negative of the gradient of V and<br />

E = −∇V. (2.6.7)<br />

For these conditions note that ∇V · dr = − E · dr is an exact differential so that the potential function can<br />

be represented by the line integral<br />

P<br />

V = V(P )=− E · dr<br />

α<br />

(2.6.8)<br />

where α is some reference point (usually infinity, where V(∞) = 0). For a conservative electric field the line<br />

integral will be independent of the path connecting any two points a and b so that<br />

b a b<br />

V(b) −V(a) =− E · dr − − E · dr = − E · dr =<br />

α<br />

α<br />

a<br />

b<br />

a<br />

∇V · dr. (2.6.9)<br />

Let α = ∞ in equation (2.6.8), then the potential function associated with a point charge moving in<br />

the radial direction er is<br />

r<br />

V(r) =− E · dr =<br />

∞<br />

−q<br />

r<br />

1 q 1<br />

dr =<br />

4πɛ0 ∞ r2 4πɛ0 r |r q<br />

∞ =<br />

4πɛ0r .


By superposition, the potential at a point P for a continuous volume distribution of charges is given by<br />

V(P )= 1<br />

<br />

ρ<br />

4πɛ0 V<br />

∗<br />

<br />

1 µ<br />

dτ and for a surface distribution of charges V(P )=<br />

r 4πɛ0 S<br />

∗<br />

dσ and for a line<br />

r<br />

distribution of charges V(P )= 1<br />

<br />

λ<br />

4πɛ0 C<br />

∗<br />

ds; and for a discrete distribution of point charges<br />

r<br />

V(P )= 1<br />

N qi<br />

. When the potential functions are defined from a common reference point, then the<br />

4πɛ0 ri<br />

i=1<br />

principal of superposition applies.<br />

The potential function V is related to the work done W in moving a charge within the electric field.<br />

The work done in moving a test charge Q from point a to point b is an integral of the force times distance<br />

moved. The electric force on a test charge Q is F = Q E and so the force F = −Q E is in opposition to this<br />

force as you move the test charge. The work done is<br />

W =<br />

b<br />

a<br />

F · dr =<br />

b<br />

a<br />

−Q E · dr = Q<br />

b<br />

a<br />

∇V · dr = Q[V(b) −V(a)]. (2.6.10)<br />

The work done is independent of the path joining the two points and depends only on the end points and<br />

the change in the potential. If one moves Q from infinity to point b, then the above becomes W = QV (b).<br />

An electric field E = E(P ) is a vector field which can be represented graphically by constructing vectors<br />

at various selected points in the space. Such a plot is called a vector field plot. A field line associated with<br />

a vector field is a curve such that the tangent vector to a point on the curve has the same direction as the<br />

vector field at that point. Field lines are used as an aid for visualization of an electric field and vector fields<br />

in general. The tangent to a field line at a point has the same direction as the vector field E at that point.<br />

For example, in two dimensions let r = x e1 + y e2 denote the position vector to a point on a field line. The<br />

tangent vector to this point has the direction dr = dx e1 + dy e2. If E = E(x, y) =−N(x, y) e1 + M(x, y) e2<br />

is the vector field constructed at the same point, then E and dr must be colinear. Thus, for each point (x, y)<br />

onafieldlinewerequirethatdr = K E for some constant K. Equating like components we find that the<br />

field lines must satisfy the differential relation.<br />

dx<br />

−N(x, y) =<br />

dy<br />

M(x, y) =K<br />

or M(x, y) dx + N(x, y) dy =0.<br />

(2.6.11)<br />

In two dimensions, the family of equipotential curves V(x, y) =C1 =constant, are orthogonal to the family<br />

of field lines and are described by solutions of the differential equation<br />

N(x, y) dx − M(x, y) dy =0<br />

obtained from equation (2.6.11) by taking the negative reciprocal of the slope. The field lines are perpendicular<br />

to the equipotential curves because at each point on the curve V = C1 we have ∇V being perpendicular<br />

to the curve V = C1 and so it is colinear with E at this same point. Field lines associated with electric<br />

fields are called electric lines of force. The density of the field lines drawn per unit cross sectional area are<br />

proportional to the magnitude of the vector field through that area.<br />

327


328<br />

Figure 2.6-1. Electric forces due to a positive charge at (−a, 0) and negative charge at (a, 0).<br />

EXAMPLE 2.6-1.<br />

Find the field lines and equipotential curves associated with a positive charge q located at the point<br />

(−a, 0) and a negative charge −q located at the point (a, 0).<br />

Solution: With reference to the figure 2.6-1, the total electric force E on a test charge Q =1place<br />

at a general point (x, y) is, by superposition, the sum of the forces from each of the isolated charges and is<br />

E = E1 + E2. The electric force vectors due to each individual charge are<br />

where k = 1<br />

4πɛ0<br />

E1 = kq(x + a) e1 + kqy e2<br />

r 3 1<br />

E2 = −kq(x − a) e1 − kqy e2<br />

r 3 2<br />

is a constant. This gives<br />

E = E1 + <br />

kq(x + a)<br />

E2 =<br />

−<br />

This determines the differential equation of the field lines<br />

r 3 1<br />

kq(x+a)<br />

r 3 1<br />

dx<br />

− kq(x−a)<br />

r 3 2<br />

To solve this differential equation we make the substitutions<br />

x + a<br />

cos θ1 =<br />

r1<br />

with r 2 1 =(x + a) 2 + y 2<br />

with r 2 2 =(x − a)2 + y 2<br />

kq(x − a)<br />

r3 <br />

kqy<br />

e1 +<br />

2<br />

r3 −<br />

1<br />

kqy<br />

r3 <br />

e2.<br />

2<br />

=<br />

kqy<br />

r 3 1<br />

dy<br />

− kqy<br />

r 3 2<br />

x − a<br />

and cos θ2 =<br />

(2.6.<strong>12</strong>)<br />

. (2.6.13)<br />

r2<br />

(2.6.14)


Figure 2.6-2. Lines of electric force between two opposite sign charges.<br />

as suggested by the geometry from figure 2.6-1. From the equations (2.6.<strong>12</strong>) and (2.6.14) we obtain the<br />

relations<br />

which implies that<br />

− sin θ1 dθ1 = r1dx − (x + a) dr1<br />

r 2 1<br />

2r1dr1 =2(x + a) dx +2ydy<br />

− sin θ2 dθ2 = r2 dx − (x − a)dr2<br />

r 2 2<br />

2r2 dr2 =2(x − a) dx +2ydy<br />

− sin θ1 dθ1 = −<br />

− sin θ2 dθ2 = −<br />

(x + a)ydy<br />

r3 1<br />

(x − a)ydy<br />

r3 2<br />

+ y2 dx<br />

r 3 1<br />

+ y2 dx<br />

r 3 2<br />

(2.6.15)<br />

Now compare the results from equation (2.6.15) with the differential equation (2.6.13) and determine that<br />

y is an integrating factor of equation (2.6.13) . This shows that the differential equation (2.6.13) can be<br />

written in the much simpler form of the exact differential equation<br />

− sin θ1 dθ1 +sinθ2 dθ2 =0 (2.6.16)<br />

in terms of the variables θ1 and θ2. The equation (2.6.16) is easily integrated to obtain<br />

cos θ1 − cos θ2 = C (2.6.17)<br />

where C is a constant of integration. In terms of x, y the solution can be written<br />

These field lines are illustrated in the figure 2.6-2.<br />

x + a<br />

<br />

(x + a) 2 + y2 −<br />

x − a<br />

= C. (2.6.18)<br />

(x − a) 2 + y2 329


330<br />

The differential equation for the equipotential curves is obtained by taking the negative reciprocal of<br />

the slope of the field lines. This gives<br />

dy<br />

dx =<br />

r 3 1<br />

kq(x−a)<br />

r3 2<br />

kqy<br />

r3 1<br />

− kq(x+a)<br />

r 3 1<br />

− kqy<br />

r 3 2<br />

This result can be written in the form<br />

<br />

(x + a)dx + ydy (x − a)dx + ydy<br />

−<br />

+<br />

=0<br />

which simplifies to the easily integrable form<br />

− dr1<br />

r 2 1<br />

+ dr2<br />

r 2 2<br />

in terms of the new variables r1 and r2. An integration produces the equipotential curves<br />

or<br />

1<br />

(x + a) 2 + y 2 −<br />

=0<br />

r 3 2<br />

1<br />

r1<br />

.<br />

− 1<br />

r2<br />

=C2<br />

1<br />

(x − a) 2 + y 2 =C2.<br />

The potential function for this problem can be interpreted as a superposition of the potential functions<br />

V1 = − kq<br />

and V2 = kq<br />

associated with the isolated point charges at the points (−a, 0) and (a, 0).<br />

r1<br />

r2<br />

Observe that the electric lines of force move from positive charges to negative charges and they do not<br />

cross one another. Where field lines are close together the field is strong and where the lines are far apart<br />

the field is weak. If the field lines are almost parallel and equidistant from one another the field is said to be<br />

uniform. The arrows on the field lines show the direction of the electric field E. If one moves along a field<br />

line in the direction of the arrows the electric potential is decreasing and they cross the equipotential curves<br />

at right angles. Also, when the electric field is conservative we will have ∇× E =0.<br />

In three dimensions the situation is analogous to what has been done in two dimensions. If the electric<br />

field is E = E(x, y, z) =P (x, y, z) e1 + Q(x, y, z) e2 + R(x, y, z) e3 and r = x e1 + y e2 + z e3 is the position<br />

vector to a variable point (x, y, z) on a field line, then at this point dr and E must be colinear so that<br />

dr = K E for some constant K. Equating like coefficients gives the system of equations<br />

dx<br />

P (x, y, z) =<br />

dy<br />

Q(x, y, z) =<br />

dz<br />

= K. (2.6.19)<br />

R(x, y, z)<br />

From this system of equations one must try to obtain two independent integrals, call them u1(x, y, z) =c1<br />

and u2(x, y, z) =c2. These integrals represent one-parameter families of surfaces. When any two of these<br />

surfaces intersect, the result is a curve which represents a field line associated with the vector field E. These<br />

type of field lines in three dimensions are more difficult to illustrate.<br />

The electric flux φE of an electric field E over a surface S is defined as the summation of the normal<br />

component of E over the surface and is represented<br />

<br />

φE = E · ˆn dσ with units of Nm2<br />

(2.6.20)<br />

C<br />

S


where ˆn is a unit normal to the surface. The flux φE can be thought of as being proportional to the number<br />

of electric field lines passing through an element of surface area. If the surface is a closed surface we have<br />

by the divergence theorem of Gauss<br />

<br />

φE = ∇· <br />

Edτ = E · ˆn dσ<br />

where V is the volume enclosed by S.<br />

V<br />

Gauss Law<br />

Let dσ denote an element of surface area on a surface S. A cone is formed if all points on the boundary<br />

of dσ are connected by straight lines to the origin. The cone need not be a right circular cone. The situation<br />

is illustrated in the figure 2.6-3.<br />

Figure 2.6-3. Solid angle subtended by element of area.<br />

We let r denote a position vector from the origin to a point on the boundary of dσ and let ˆn denote a<br />

unit outward normal to the surface at this point. We then have ˆn · r = r cos θ where r = |r| and θ is the<br />

angle between the vectors ˆn and r. Construct a sphere, centered at the origin, having radius r. This sphere<br />

intersects the cone in an element of area dΩ. The solid angle subtended by dσ is defined as dω = dΩ<br />

. Note<br />

r2 that this is equivalent to constructing a unit sphere at the origin which intersect the cone in an element of<br />

area dω. Solid angles are measured in steradians. The total solid angle about a point equals the area of the<br />

sphere divided by its radius squared or 4π steradians. The element of area dΩ is the projection of dσ on the<br />

ˆn · r<br />

ˆn · r dΩ<br />

constructed sphere and dΩ =dσ cos θ = dσ so that dω = dσ = . Observe that sometimes the<br />

r r3 r2 dot product ˆn · r is negative, the sign depending upon which of the normals to the surface is constructed.<br />

(i.e. the inner or outer normal.)<br />

The Gauss law for electrostatics in a vacuum states that the flux through any surface enclosing many<br />

charges is the total charge enclosed by the surface divided by ɛ0. The Gauss law is written<br />

<br />

Qe<br />

E<br />

for charges inside S<br />

ɛ0<br />

· ˆn dσ =<br />

(2.6.21)<br />

0 for charges outside S<br />

S<br />

S<br />

331


332<br />

where Qe represents the total charge enclosed by the surface S with ˆn the unit outward normal to the surface.<br />

The proof of Gauss’s theorem follows. Consider a single charge q within the closed surface S. The electric<br />

field at a point on the surface S due to the charge q within S is represented E = 1 q<br />

4πɛ0 r2 er and so the flux<br />

integral is<br />

<br />

<br />

φE = E<br />

q er · ˆn<br />

· ˆn dσ =<br />

S<br />

S 4πɛ0 r2 dσ = q<br />

<br />

dΩ q<br />

= (2.6.22)<br />

4πɛ0 S r2 ɛ0<br />

since er · ˆn<br />

r2 cos θdσ<br />

=<br />

r2 = dΩ<br />

<br />

= dω and dω =4π. By superposition of the charges, we obtain a similar<br />

r2 S<br />

n<br />

result for each of the charges within the surface. Adding these results gives Qe = qi. For a continuous<br />

<br />

i=1<br />

distribution of charge inside the volume we can write Qe = ρ<br />

V<br />

∗ dτ, whereρ ∗ is the charge distribution<br />

per unit volume. Note that charges outside of the closed surface do not contribute to the total flux across<br />

the surface. This is because the field lines go in one side of the surface and go out the other side. In this<br />

case E · ˆn dσ = 0 for charges outside the surface. Also the position of the charge or charges within the<br />

S<br />

volume does not effect the Gauss law.<br />

The equation (2.6.21) is the Gauss law in integral form. We can put this law in differential form as<br />

follows. Using the Gauss divergence theorem we can write for an arbitrary volume that<br />

<br />

<br />

E · ˆn dσ = ∇·<br />

S<br />

V<br />

<br />

ρ<br />

Edτ =<br />

V<br />

∗<br />

dτ =<br />

ɛ0<br />

Qe<br />

=<br />

ɛ0<br />

1<br />

<br />

ρ<br />

ɛ0 V<br />

∗ dτ<br />

which for an arbitrary volume implies<br />

∇· E = ρ∗<br />

. (2.6.23)<br />

The equations (2.6.23) and (2.6.7) can be combined so that the Gauss law can also be written in the form<br />

∇ 2 V = − ρ∗<br />

ɛ0<br />

which is called Poisson’s equation.<br />

EXAMPLE 2.6-2<br />

Find the electric field associated with an infinite plane sheet of positive charge.<br />

Solution: Assume there exists a uniform surface charge µ ∗ and draw a circle at some point on the plane<br />

surface. Now move the circle perpendicular to the surface to form a small cylinder which extends equal<br />

distances above and below the plane surface. We calculate the electric flux over this small cylinder in the<br />

limit as the height of the cylinder goes to zero. The charge inside the cylinder is µ ∗ A where A is the area of<br />

the circle. We find that the Gauss law requires that<br />

<br />

E · ˆn dσ = Qe<br />

=<br />

ɛ0<br />

µ∗A (2.6.24)<br />

ɛ0<br />

S<br />

where ˆn is the outward normal to the cylinder as we move over the surface S. By the symmetry of the<br />

situation the electric force vector is uniform and must point away from both sides to the plane surface in the<br />

direction of the normals to both sides of the surface. Denote the plane surface normals by en and − en and<br />

assume that E = β en on one side of the surface and E = −β en on the other side of the surface for some<br />

constant β. Substituting this result into the equation (2.6.24) produces<br />

<br />

E · ˆn dσ =2βA (2.6.25)<br />

S<br />

ɛ0


since only the ends of the cylinder contribute to the above surface integral. On the sides of the cylinder we<br />

will have ˆn ·±en = 0 and so the surface integral over the sides of the cylinder is zero. By equating the<br />

results from equations (2.6.24) and (2.6.25) we obtain the result that β = µ∗<br />

and consequently we can write<br />

2ɛ0<br />

E = µ∗<br />

en where en represents one of the normals to the surface.<br />

2ɛ0<br />

Note an electric field will always undergo a jump discontinuity when crossing a surface charge µ ∗ . As in<br />

theaboveexamplewehave Eup = µ∗<br />

en and<br />

2ɛ0<br />

Edown = − µ∗<br />

2ɛ en so that the difference is<br />

Eup − Edown = µ∗<br />

It is this difference which causes the jump discontinuity.<br />

ɛ0<br />

en or E i n (1)<br />

i + Ein (2) µ∗<br />

i + =0. (2.6.26)<br />

ɛ0<br />

EXAMPLE 2.6-3.<br />

Calculate the electric field associated with a uniformly charged sphere of radius a.<br />

Solution: We proceed as in the previous example. Let µ ∗ denote the uniform charge distribution over the<br />

surface of the sphere and let er denote the unit normal to the sphere. The total charge then is written as<br />

q = µ<br />

Sa<br />

∗ dσ =4πa 2 µ ∗ . If we construct a sphere of radius r>aaround the charged sphere, then we have<br />

by the Gauss theorem <br />

E · er dσ = Qe<br />

ɛ0<br />

= q<br />

.<br />

ɛ0<br />

(2.6.27)<br />

Sr<br />

Again, we can assume symmetry for E and assume that it points radially outward in the direction of the<br />

surface normal er and has the form E = β er for some constant β. Substituting this value for E into the<br />

equation (2.6.27) we find that<br />

<br />

<br />

E · er dσ = β dσ =4πβr 2 = q<br />

. (2.6.28)<br />

Sr<br />

This gives E = 1 q<br />

4πɛ0 r2 er where er is the outward normal to the sphere. This shows that the electric field<br />

outside the sphere is the same as if all the charge were situated at the origin.<br />

Sr<br />

For S a piecewise closed surface enclosing a volume V and F i = F i (x 1 ,x 2 ,x 3 ) i =1, 2, 3, a continuous<br />

vector field with continuous derivatives the Gauss divergence theorem enables us to replace a flux integral<br />

of F i over S by a volume integral of the divergence of F i over the volume V such that<br />

<br />

S<br />

F i <br />

ni dσ =<br />

V<br />

F i ,i dτ or<br />

<br />

S<br />

ɛ0<br />

<br />

F · ˆn dσ =<br />

V<br />

div Fdτ. (2.6.29)<br />

If V contains a simple closed surface Σ where F i is discontinuous we must modify the above Gauss divergence<br />

theorem.<br />

EXAMPLE 2.6-4.<br />

We examine the modification of the Gauss divergence theorem for spheres in order to illustrate the<br />

concepts. Let V have surface area S which encloses a surface Σ. Consider the figure 2.6-4 where the volume<br />

V enclosed by S and containing Σ has been cut in half.<br />

333


334<br />

Figure 2.6-4. Sphere S containing sphere Σ.<br />

Applying the Gauss divergence theorem to the top half of figure 2.6-4 gives<br />

<br />

F i n T <br />

<br />

<br />

i dσ +<br />

dσ +<br />

dσ =<br />

ST<br />

Sb1<br />

F i n bT<br />

i<br />

ΣT<br />

F i n ΣT<br />

i<br />

VT<br />

F i ,i dτ (2.6.30)<br />

where the ni are the unit outward normals to the respective surfaces ST , Sb1 and ΣT . Applying the Gauss<br />

divergence theorem to the bottom half of the sphere in figure 2.6-4 gives<br />

<br />

<br />

<br />

<br />

dσ +<br />

dσ +<br />

dσ =<br />

F<br />

SB<br />

i n B i<br />

Sb2<br />

F i n bB<br />

i<br />

ΣB<br />

F i n ΣB<br />

i<br />

VB<br />

F i ,i dτ (2.6.31)<br />

Observe that the unit normals to the surfaces Sb1 and Sb2 are equal and opposite in sign so that adding the<br />

equations (2.6.30) and (2.6.31) we obtain<br />

<br />

F<br />

S<br />

i <br />

ni dσ + F<br />

Σ<br />

i n (1)<br />

<br />

i dσ =<br />

F<br />

VT +VB<br />

i ,i dτ (2.6.32)


where S = ST + SB is the total surface area of the outside sphere and Σ = ΣT +ΣB is the total surface area<br />

of the inside sphere, and n (1)<br />

i is the inward normal to the sphere Σ when the top and bottom volumes are<br />

combined. Applying the Gauss divergence theorem to just the isolated small sphere Σ we find<br />

<br />

F<br />

Σ<br />

i n (2)<br />

<br />

i dσ = F<br />

VΣ<br />

i ,i dτ (2.6.33)<br />

where n (2)<br />

i<br />

is the outward normal to Σ. By adding the equations (2.6.33) and (2.6.32) we find that<br />

<br />

F<br />

S<br />

i <br />

ni dσ + F<br />

Σ<br />

i n (1)<br />

i + F i n (2)<br />

<br />

i dσ = F<br />

V<br />

i ,i dτ (2.6.34)<br />

where V = VT + VB + VΣ. The equation (2.6.34) can also be written as<br />

<br />

F<br />

S<br />

i <br />

ni dσ = F<br />

V<br />

i <br />

,i dτ − F<br />

Σ<br />

i n (1)<br />

i + F i n (2)<br />

<br />

i dσ. (2.6.35)<br />

In the case that V contains a surface Σ the total electric charge inside S is<br />

<br />

Qe =<br />

V<br />

ρ ∗ <br />

dτ + µ<br />

Σ<br />

∗ dσ (2.6.36)<br />

where µ ∗ is the surface charge density on Σ and ρ ∗ is the volume charge density throughout V. The Gauss<br />

theorem requires that <br />

S<br />

E i ni dσ = Qe<br />

ɛ0<br />

= 1<br />

<br />

ρ<br />

ɛ0 V<br />

∗ dτ + 1<br />

<br />

µ<br />

ɛ0 Σ<br />

∗ dσ. (2.6.37)<br />

In the case of a jump discontinuity across the surface Σ we use the results of equation (2.6.34) and write<br />

<br />

E<br />

S<br />

i <br />

ni dσ = E<br />

V<br />

i <br />

,i dτ − E<br />

Σ<br />

i n (1)<br />

i + Ein (2)<br />

<br />

i dσ. (2.6.38)<br />

Subtracting the equation (2.6.37) from the equation (2.6.38) gives<br />

<br />

E<br />

V<br />

i ,i − ρ∗<br />

<br />

dτ − E<br />

ɛ0<br />

Σ<br />

i n (1)<br />

i + Ein (2)<br />

<br />

µ∗<br />

i + dσ =0. (2.6.39)<br />

ɛ0<br />

For arbitrary surfaces S and Σ, this equation implies the differential form of the Gauss law<br />

E i ,i = ρ∗<br />

Further, on the surface Σ, where there is a surface charge distribution we have<br />

ɛ0<br />

. (2.6.40)<br />

E i n (1)<br />

i + Ein (2) µ∗<br />

i + =0 (2.6.41)<br />

ɛ0<br />

which shows the electric field undergoes a discontinuity when you cross a surface charge µ ∗ .<br />

335


336<br />

Electrostatic Fields in Materials<br />

When charges are introduced into materials it spreads itself throughout the material. Materials in<br />

which the spreading occurs quickly are called conductors, while materials in which the spreading takes a<br />

long time are called nonconductors or dielectrics. Another electrical property of materials is the ability to<br />

hold local charges which do not come into contact with other charges. This property is called induction.<br />

For example, consider a single atom within the material. It has a positively charged nucleus and negatively<br />

charged electron cloud surrounding it. When this atom experiences an electric field E the negative cloud<br />

moves opposite to E while the positively charged nucleus moves in the direction of E.If E is large enough it<br />

can ionize the atom by pulling the electrons away from the nucleus. For moderately sized electric fields the<br />

atom achieves an equilibrium position where the positive and negative charges are offset. In this situation<br />

the atom is said to be polarized and have a dipole moment p.<br />

Definition: When a pair of charges +q and −q are separated by a distance 2 d the electric dipole<br />

moment is defined by p =2 dq, wherephasdimensionsof[Cm]. In the special case where d has the same direction as E and the material is symmetric we say that p<br />

is proportional to E and write p = α E,whereαis called the atomic polarizability. If in a material subject<br />

to an electric field their results many such dipoles throughout the material then the dielectric is said to be<br />

polarized. The vector quantity P is introduced to represent this effect. The vector P is called the polarization<br />

vector having units of [C/m2 ], and represents an average dipole moment per unit volume of material. The<br />

vectors Pi and Ei are related through the displacement vector Di such that<br />

For an anisotropic material (crystal)<br />

where ɛ j<br />

i<br />

is called the dielectric tensor and αj<br />

i<br />

Pi = Di − ɛ0Ei. (2.6.42)<br />

Di = ɛ j<br />

i Ej and Pi = α j<br />

i Ej<br />

is called the electric susceptibility tensor. Consequently,<br />

Pi = α j<br />

i Ej = ɛ j<br />

i Ej − ɛ0Ei =(ɛ j j<br />

i − ɛ0δi )Ej so that α j<br />

i = ɛj<br />

i<br />

(2.6.43)<br />

j<br />

− ɛ0δi . (2.6.44)<br />

A dielectric material is called homogeneous if the electric force and displacement vector are the same for any<br />

two points within the medium. This requires that the electric force and displacement vectors be constant<br />

parallel vector fields. It is left as an exercise to show that the condition for homogeneity is that ɛ j<br />

i,k =0.<br />

A dielectric material is called isotropic if the electric force vector and displacement vector have the same<br />

direction. This requires that ɛ j<br />

i = ɛδi j where δi j is the Kronecker delta. The term ɛ = ɛ0Ke is called the<br />

dielectric constant of the medium. The constant ɛ0 =8.85(10) −<strong>12</strong> coul 2 /N · m2 is the permittivity of free<br />

space and the quantity ke = ɛ<br />

ɛ0 is called the relative dielectric constant (relative to ɛ0). For free space ke =1.<br />

Similarly for an isotropic material we have α j<br />

i<br />

linear medium the vectors P , D and E are related by<br />

= ɛ0αeδ j<br />

i where αe is called the electric susceptibility. For a<br />

Di = ɛ0Ei + Pi = ɛ0Ei + ɛ0αeEi = ɛ0(1 + αe)Ei = ɛ0KeEi = ɛEi<br />

(2.6.45)


where Ke =1+αe is the relative dielectric constant. The equation (2.6.45) are constitutive equations for<br />

dielectric materials.<br />

The effect of polarization is to produce regions of bound charges ρb within the material and bound<br />

surface charges µb together with free charges ρf which are not a result of the polarization. Within dielectrics<br />

we have ∇· P = ρb for bound volume charges and P · en = µb for bound surface charges, where en is a<br />

unit normal to the bounding surface of the volume. In these circumstances the expression for the potential<br />

function is written<br />

V = 1<br />

and the Gauss law becomes<br />

<br />

4πɛ0 V<br />

<br />

ρb 1 µb<br />

dτ + dσ<br />

r 4πɛ0 S r<br />

(2.6.46)<br />

ɛ0∇· E = ρ ∗ = ρb + ρf = −∇ · P + ρf or ∇(ɛ0 E + P )=ρf . (2.6.47)<br />

Since D = ɛ0 E + P the Gauss law can also be written in the form<br />

∇· D = ρf or D i ,i = ρf . (2.6.48)<br />

When no confusion arises we replace ρf by ρ. In integral form the Gauss law for dielectrics is written<br />

<br />

D · ˆn dσ = Qfe<br />

(2.6.49)<br />

where Qfe is the total free charge density within the enclosing surface.<br />

S<br />

Magnetostatics<br />

A stationary charge generates an electric field E while a moving charge generates a magnetic field B.<br />

Magnetic field lines associated with a steady current moving in a wire form closed loops as illustrated in the<br />

figure 2.6-5.<br />

Figure 2.6-5. Magnetic field lines.<br />

The direction of the magnetic force is determined by the right hand rule where the thumb of the right<br />

hand points in the direction of the current flow and the fingers of the right hand curl around in the direction<br />

of the magnetic field B. The force on a test charge Q moving with velocity V in a magnetic field is<br />

Fm = Q( V × B). (2.6.50)<br />

The total electromagnetic force acting on Q is the electric force plus the magnetic force and is<br />

<br />

F = Q E +( V × B) <br />

(2.6.51)<br />

337


338<br />

which is known as the Lorentz force law. The magnetic force due to a line charge density λ∗ moving along<br />

acurveCis the line integral<br />

<br />

Fmag = λ ∗ ds( V × <br />

B)= I × Bds. (2.6.52)<br />

C<br />

Similarly, for a moving surface charge density moving on a surface<br />

<br />

Fmag = µ ∗ dσ( V × <br />

B)= K × Bdσ (2.6.53)<br />

and for a moving volume charge density<br />

<br />

Fmag =<br />

S<br />

V<br />

C<br />

S<br />

ρ ∗ dτ( V × <br />

B)=<br />

V<br />

J × Bdτ (2.6.54)<br />

where the quantities I = λ∗V , K = µ ∗ V and J = ρ∗ V are respectively the current, the current per unit<br />

length, and current per unit area.<br />

A conductor is any material where the charge is free to move. The flow of charge is governed by Ohm’s<br />

law. Ohm’s law states that the current density vector Ji is a linear function of the electric intensity or<br />

Ji = σimEm, whereσim is the conductivity tensor of the material. For homogeneous, isotropic conductors<br />

σim = σδim so that Ji = σEi where σ is the conductivity and 1/σ is called the resistivity.<br />

Surround a charge density ρ∗ with an arbitrary simple closed surface S having volume V and calculate<br />

the flux of the current density across the surface. We find by the divergence theorem<br />

<br />

<br />

J · ˆn dσ = ∇· Jdτ. (2.6.55)<br />

S<br />

If charge is to be conserved, the current flow out of the volume through the surface must equal the loss due<br />

to the time rate of change of charge within the surface which implies<br />

<br />

<br />

J · ˆn dσ = ∇·<br />

S<br />

V<br />

Jdτ = − d<br />

<br />

ρ<br />

dt V<br />

∗ <br />

dτ = −<br />

V<br />

∂ρ∗ dτ<br />

∂t<br />

(2.6.56)<br />

or <br />

∇·<br />

V<br />

J + ∂ρ∗<br />

<br />

dτ =0.<br />

∂t<br />

(2.6.57)<br />

This implies that for an arbitrary volume we must have<br />

V<br />

∇· J = − ∂ρ∗<br />

. (2.6.58)<br />

∂t<br />

Note that equation (2.6.58) has the same form as the continuity equation (2.3.73) for mass conservation and<br />

so it is also called a continuity equation for charge conservation. For magnetostatics there exists steady line<br />

currents or stationary current so ∂ρ∗<br />

∂t = 0. This requires that ∇· J =0.


Figure 2.6-6. Magnetic field around wire.<br />

Biot-Savart Law<br />

The Biot-Savart law for magnetostatics describes the magnetic field at a point P due to a steady line<br />

current moving along a curve C and is<br />

B(P )= µ0<br />

<br />

I × er<br />

4π C r2 ds (2.6.59)<br />

with units [N/amp · m] and where the integration is in the direction of the current flow. In the Biot-Savart<br />

law we have the constant µ0 =4π × 10−7 N/amp2 which is called the permeability of free space, I = I et is<br />

the current flowing in the direction of the unit tangent vector et to the curve C, er is a unit vector directed<br />

from a point on the curve C toward the point P and r is the distance from a point on the curve to the<br />

general point P. Note that for a steady current to exist along the curve the magnitude of I must be the<br />

same everywhere along the curve. Hence, this term can be brought out in front of the integral. For surface<br />

currents K and volume currents J the Biot-Savart law is written<br />

B(P )= µ0<br />

<br />

K × er<br />

4π S r2 dσ<br />

<br />

µ0 J<br />

and B(P <br />

× er<br />

)=<br />

4π r2 dτ.<br />

EXAMPLE 2.6-5.<br />

Calculate the magnetic field B adistancehperpendicular to a wire carrying a constant current I.<br />

Solution: The magnetic field circles around the wire. For the geometry of the figure 2.6-6, the magnetic<br />

field points out of the page. We can write<br />

I × er = I et × er = Iê sin α<br />

where ê is a unit vector tangent to the circle of radius h which encircles the wire and cuts the wire perpendicularly.<br />

V<br />

339


340<br />

For this problem the Biot-Savart law is<br />

B(P )= µ0I<br />

4π<br />

In terms of θ we find from the geometry of figure 2.6-6<br />

Therefore,<br />

<br />

ê<br />

ds.<br />

r2 tan θ = s<br />

h with ds = h sec2 θdθ and cos θ = h<br />

r .<br />

B(P )= µ0<br />

π<br />

θ2<br />

But, α = π/2+θ so that sin α =cosθ and consequently<br />

B(P )= µ0Iê<br />

4πh<br />

θ2<br />

θ1<br />

θ1<br />

Iê sin αhsec 2 θ<br />

h 2 / cos 2 θ<br />

dθ.<br />

cos θdθ= µ0Iê<br />

4πh (sin θ2 − sin θ1).<br />

For a long straight wire θ1 →−π/2 andθ2 → π/2 to give the magnetic field B(P )= µ0Iê<br />

2πh .<br />

For volume currents the Biot-Savart law is<br />

B(P )= µ0<br />

<br />

4π V<br />

and consequently (see exercises)<br />

J × er<br />

r 2 dτ (2.6.60)<br />

∇· B =0. (2.6.61)<br />

Recall the divergence of an electric field is ∇· E = ρ∗<br />

is known as the Gauss’s law for electric fields and so<br />

ɛ0<br />

in analogy the divergence ∇· B = 0 is sometimes referred to as Gauss’s law for magnetic fields. If ∇· B =0,<br />

then there exists a vector field A such that B = ∇× A. The vector field A is called the vector potential of<br />

B. Note that ∇· B = ∇·(∇ × A)=0. Also the vector potential A is not unique since B is also derivable<br />

from the vector potential A + ∇φ where φ is an arbitrary continuous and differentiable scalar.<br />

Ampere’s Law<br />

Ampere’s law is associated with the work done in moving around a simple closed path. For example,<br />

consider the previous example 2.6-5. In this example the integral of B around a circular path of radius h<br />

which is centered at some point on the wire can be associated with the work done in moving around this<br />

path. The summation of force times distance is<br />

<br />

○ <br />

B · dr = ○ B · ê ds = µ0I<br />

<br />

○ ds = µ0I (2.6.62)<br />

2πh<br />

C<br />

C<br />

<br />

where now dr = ê ds is a tangent vector to the circle encircling the wire and ○ ds =2πh is the distance<br />

C<br />

around this circle. The equation (2.6.62) holds not only for circles, but for any simple closed curve around<br />

the wire. Using the Stoke’s theorem we have<br />

<br />

○ <br />

B · dr = (∇× <br />

B) · en dσ = µ0I = µ0 J · en dσ (2.6.63)<br />

C<br />

S<br />

C<br />

S


where J · en dσ is the total flux (current) passing through the surface which is created by encircling<br />

S<br />

some curve about the wire. Equating like terms in equation (2.6.63) gives the differential form of Ampere’s<br />

law<br />

∇× B = µ0 J. (2.6.64)<br />

Magnetostatics in Materials<br />

Similar to what happens when charges are introduced into materials we have magnetic fields whenever<br />

there are moving charges within materials. For example, when electrons move around an atom tiny current<br />

loops are formed. These current loops create what are called magnetic dipole moments m throughout the<br />

material. When a magnetic field B is applied to a material medium there is a net alignment of the magnetic<br />

dipoles. The quantity M, called the magnetization vector is introduced. Here M is associated with a<br />

dielectric medium and has the units [amp/m] and represents an average magnetic dipole moment per unit<br />

volume and is analogous to the polarization vector P used in electrostatics. The magnetization vector M<br />

acts a lot like the previous polarization vector in that it produces bound volume currents Jb and surface<br />

currents Kb where ∇× M = Jb is a volume current density throughout some volume and M × en = Kb is a<br />

surface current on the boundary of this volume.<br />

From electrostatics note that the time derivative of ɛ0 ∂ E<br />

∂t has the same units as current density. The<br />

total current in a magnetized material is then Jt = Jb + Jf + ɛ0 ∂ E<br />

∂t where Jb is the bound current, Jf is the<br />

free current and ɛ0 ∂ E<br />

∂t is the induced current. Ampere’s law, equation (2.6.64), in magnetized materials then<br />

becomes<br />

∇× B = µ0 Jt = µ0( Jb + ∂<br />

Jf + ɛ0<br />

E<br />

∂t )=µ0 ∂<br />

J + µ0ɛ0<br />

E<br />

∂t<br />

(2.6.65)<br />

where J = Jb + Jf . The term ɛ0 ∂ E<br />

∂t is referred to as a displacement current or as a Maxwell correction to<br />

the field equation. This term implies that a changing electric field induces a magnetic field.<br />

An auxiliary magnet field H defined by<br />

Hi = 1<br />

Bi − Mi<br />

µ0<br />

(2.6.66)<br />

is introduced which relates the magnetic force vector B and magnetization vector M. This is another constitutive<br />

equation which describes material properties. For an anisotropic material (crystal)<br />

Bi = µ j<br />

i Hj and Mi = χ j<br />

i Hj<br />

(2.6.67)<br />

where µ j<br />

i is called the magnetic permeability tensor and χji<br />

is called the magnetic permeability tensor. Both<br />

of these quantities are dimensionless. For an isotropic material<br />

µ j<br />

i = µδj i where µ = µ0km. (2.6.68)<br />

Here µ0 =4π × 10 −7 N/amp 2 is the permeability of free space and km = µ<br />

µ0<br />

coefficient. Similarly, for an isotropic material we have χ j<br />

i<br />

is the relative permeability<br />

= χmδ j<br />

i where χm is called the magnetic sus-<br />

ceptibility coefficient and is dimensionless. The magnetic susceptibility coefficient has positive values for<br />

341


342<br />

materials called paramagnets and negative values for materials called diamagnets. For a linear medium the<br />

quantities B, M and H are related by<br />

Bi = µ0(Hi + Mi) =µ0Hi + µ0χmHi = µ0(1 + χm)Hi = µ0kmHi = µHi<br />

(2.6.69)<br />

where µ = µ0km = µ0(1 + χm) is called the permeability of the material.<br />

Note: The auxiliary magnetic vector H for magnetostatics in materials plays a role similar to the<br />

displacement vector D for electrostatics in materials. Be careful in using electromagnetic equations from<br />

different texts as many authors interchange the roles of B and H. Some authors call H the magnetic field.<br />

However, the quantity B should be the fundamental quantity. 1<br />

Electrodynamics<br />

In the nonstatic case of electrodynamics there is an additional quantity Jp = ∂ P<br />

∂t<br />

current which satisfies<br />

∇· Jp = ∇· ∂ P<br />

∂t<br />

and the current density has three parts<br />

∂<br />

=<br />

∂t ∇· P = − ∂ρb<br />

∂t<br />

J = Jb + Jf + Jp = ∇× M + Jf + ∂ P<br />

∂t<br />

called the polarization<br />

(2.6.70)<br />

(2.6.71)<br />

consisting of bound, free and polarization currents.<br />

Faraday’s law states that a changing magnetic field creates an electric field. In particular, the electromagnetic<br />

force induced in a closed loop circuit C is proportional to the rate of change of flux of the magnetic<br />

field associated with any surface S connected with C. Faraday’s law states<br />

<br />

○ E · dr = − ∂<br />

<br />

B · en dσ.<br />

∂t<br />

C<br />

Using the Stoke’s theorem, we find<br />

<br />

(∇×<br />

S<br />

<br />

∂<br />

E) · en dσ = −<br />

S<br />

B<br />

∂t · en dσ.<br />

The above equation must hold for an arbitrary surface and loop. Equating like terms we obtain the differential<br />

form of Faraday’s law<br />

∇× E = − ∂ B<br />

. (2.6.72)<br />

∂t<br />

This is the first electromagnetic field equation of Maxwell.<br />

Ampere’s law, equation (2.6.65), written in terms of the total current from equation (2.6.71) , becomes<br />

which can also be written as<br />

∇× B = µ0(∇× M + Jf + ∂ P<br />

∂t )+µ0ɛ0<br />

∂ E<br />

∂t<br />

∇×( 1<br />

µ0<br />

S<br />

B − M)= Jf + ∂<br />

∂t ( P + ɛ0 E)<br />

1 D.J. Griffiths, Introduction to Electrodynamics, Prentice Hall, 1981. P.232.<br />

(2.6.73)


or<br />

∇× H = Jf + ∂ D<br />

. (2.6.74)<br />

∂t<br />

This is Maxwell’s second electromagnetic field equation.<br />

To the equations (2.6.74) and (2.6.73) we add the Gauss’s law for magnetization, equation (2.6.61) and<br />

Gauss’s law for electrostatics, equation (2.6.48). These four equations produce the Maxwell’s equations of<br />

electrodynamics and are now summarized. The general form of Maxwell’s equations involve the quantities<br />

for i =1, 2, 3. Therearealsothequantities<br />

Ei, Electric force vector, [Ei] =Newton/coulomb<br />

Bi, Magnetic force vector, [Bi] =Weber/m 2<br />

Hi, Auxilary magnetic force vector, [Hi] =ampere/m<br />

Di, Displacement vector, [Di] =coulomb/m 2<br />

Ji, Free current density, [Ji] =ampere/m 2<br />

Pi, Polarization vector, [Pi] =coulomb/m 2<br />

Mi, Magnetization vector, [Mi] =ampere/m<br />

ϱ, representing the free charge density, with units [ϱ] =coulomb/m 3<br />

ɛ0, Permittivity of free space, [ɛ0] =farads/m or coulomb 2 /Newton · m 2<br />

µ0, Permeability of free space, [µ0] = henrys/m or kg· m/coulomb 2<br />

In addition, there arises the material parameters:<br />

µ i j, magnetic permeability tensor, which is dimensionless<br />

ɛ i j , dielectric tensor, which is dimensionless<br />

α i j , electric susceptibility tensor, which is dimensionless<br />

χ i j, magnetic susceptibility tensor, which is dimensionless<br />

These parameters are used to express variations in the electric field Ei and magnetic field Bi when<br />

acting in a material medium. In particular, Pi,Di,Mi and Hi are defined from the equations<br />

Di =ɛ j<br />

i Ej = ɛ0Ei + Pi<br />

The above quantities obey the following laws:<br />

ɛ i j = ɛ0δ i j + α j<br />

i<br />

Bi =µ j<br />

i Hj = µ0Hi + µ0Mi, µ i j = µ0(δ i j + χi j )<br />

Pi =α j<br />

i Ej, and Mi = χ j<br />

i Hj for i =1, 2, 3.<br />

Faraday’s Law This law states the line integral of the electromagnetic force around a loop is proportional<br />

to the rate of flux of magnetic induction through the loop. This gives rise to the first electromagnetic field<br />

equation:<br />

∇× E = − ∂ B<br />

∂t<br />

or ɛ ijk Ek,j = − ∂Bi<br />

.<br />

∂t<br />

(2.6.75)<br />

.<br />

343


344<br />

Ampere’s Law This law states the line integral of the magnetic force vector around a closed loop is<br />

proportional to the sum of the current through the loop and the rate of flux of the displacement vector<br />

through the loop. This produces the second electromagnetic field equation:<br />

∇× H = Jf + ∂ D<br />

∂t<br />

or ɛ ijk Hk,j = J i f + ∂Di<br />

. (2.6.76)<br />

∂t<br />

Gauss’s Law for Electricity This law states that the flux of the electric force vector through a closed<br />

surface is proportional to the total charge enclosed by the surface. This results in the third electromagnetic<br />

field equation:<br />

∇· D = ρf or D i ,i = ρf or<br />

1 ∂<br />

√<br />

g ∂xi √ i<br />

gD = ρf . (2.6.77)<br />

Gauss’s Law for Magnetism This law states the magnetic flux through any closed volume is zero. This<br />

produces the fourth electromagnetic field equation:<br />

∇· B =0 or B i ,i =0 or<br />

1 ∂<br />

√<br />

g ∂xi √ i<br />

gB =0. (2.6.78)<br />

When no confusion arises it is convenient to drop the subscript f from the above Maxwell equations.<br />

Special expanded forms of the above Maxwell equations are given on the pages 176 to 179.<br />

Electromagnetic Stress and Energy<br />

Let V denote the volume of some simple closed surface S. Let us calculate the rate at which electromagnetic<br />

energy is lost from this volume. This represents the energy flow per unit volume. Begin with the<br />

first two Maxwell’s equations in Cartesian form<br />

ɛijkEk,j = − ∂Bi<br />

∂t<br />

(2.6.79)<br />

ɛijkHk,j =Ji + ∂Di<br />

.<br />

∂t<br />

(2.6.80)<br />

Now multiply equation (2.6.79) by Hi and equation (2.6.80) by Ei. This gives two terms with dimensions of<br />

energy per unit volume per unit of time which we write<br />

ɛijkEk,jHi = − ∂Bi<br />

∂t Hi<br />

Subtracting equation (2.6.82) from equation (2.6.81) we find<br />

(2.6.81)<br />

ɛijkHk,jEi =JiEi + ∂Di<br />

∂t Ei. (2.6.82)<br />

ɛijk(Ek,jHi − Hk,jEi) =− JiEi − ∂Di<br />

∂t Ei − ∂Bi<br />

∂t Hi<br />

ɛijk [(EkHi),j − EkHi,j + Hi,jEk] =− JiEi − ∂Di<br />

∂t Ei − ∂Bi<br />

∂t Hi<br />

Observe that ɛjki(EkHi),j is the same as ɛijk(EjHk),i so that the above simplifies to<br />

ɛijk(EjHk),i + JiEi = − ∂Di<br />

∂t Ei − ∂Bi<br />

∂t Hi. (2.6.83)


Now integrate equation (2.6.83) over a volume and apply Gauss’s divergence theorem to obtain<br />

<br />

<br />

<br />

ɛijkEjHkni dσ + JiEi dτ = − (<br />

S<br />

V<br />

V<br />

∂Di<br />

∂t Ei + ∂Bi<br />

∂t Hi) dτ. (2.6.84)<br />

The first term in equation (2.6.84) represents the outward flow of energy across the surface enclosing the<br />

volume. The second term in equation (2.6.84) represents the loss by Joule heating and the right-hand side<br />

is the rate of decrease of stored electric and magnetic energy. The equation (2.6.84) is known as Poynting’s<br />

theorem and can be written in the vector form<br />

<br />

( E × <br />

H) · ˆn dσ = (− E · ∂ D<br />

∂t − H · ∂ B<br />

∂t − E · J) dτ. (2.6.85)<br />

For later use we define the quantity<br />

S<br />

V<br />

Si = ɛijkEjHk or S = E × H [Watts/m 2 ] (2.6.86)<br />

as Poynting’s energy flux vector and note that Si is perpendicular to both Ei and Hi and represents units<br />

of energy density per unit time which crosses a unit surface area within the electromagnetic field.<br />

Electromagnetic Stress Tensor<br />

Instead of calculating energy flow per unit volume, let us calculate force per unit volume. Consider a<br />

region containing charges and currents but is free from dielectrics and magnetic materials. To obtain terms<br />

with units of force per unit volume we take the cross product of equation (2.6.79) with Di and the cross<br />

product of equation (2.6.80) with Bi and subtract to obtain<br />

−ɛirsɛijk(Ek,jDs + Hk,jBs) =ɛrisJiBs + ɛris<br />

which simplifies using the e − δ identity to<br />

which further simplifies to<br />

−(δrjδsk − δrkδsj)(Ek,jDs + Hk,jBs) =ɛrisJiBs + ɛris<br />

<br />

∂Di<br />

∂t Bs + ∂Bs<br />

∂t Di<br />

<br />

∂<br />

∂t (DiBs)<br />

−Es,rDs + Er,sDs − Hs,rBs + Hr,sBs = ɛrisJiBs + ∂<br />

∂t (ɛrisDiBs). (2.6.87)<br />

Observe that the first two terms in the equation (2.6.87) can be written<br />

whichcanbeexpressedintheform<br />

Er,sDs − Es,rDs =Er,sDs − ɛ0Es,rEs<br />

=(ErDs),s − ErDs,s − ɛ0( 1<br />

2 EsEs),r<br />

=(ErDs),s − ρEr − 1<br />

2 (EjDjδsr),s<br />

=(ErDs − 1<br />

2 EjDjδrs),s − ρEr<br />

Er,sDs − Es,rDs = T E rs,s<br />

− ρEr<br />

345


346<br />

where<br />

T E rs = ErDs − 1<br />

2 EjDjδrs<br />

is called the electric stress tensor. In matrix form the stress tensor is written<br />

⎡<br />

⎣ E1D1 − 1<br />

2EjDj E1D2 E1D3<br />

T E rs =<br />

E2D1 E2D2 − 1<br />

2EjDj E2D3<br />

E3D1 E3D2 E3D3 − 1<br />

2EjDj (2.6.88)<br />

⎤<br />

⎦ . (2.6.89)<br />

By performing similar calculations we can transform the third and fourth terms in the equation (2.6.87) and<br />

obtain<br />

where<br />

Hr,sBs − Hs,rBs = T M rs,s<br />

T M rs = HrBS − 1<br />

2 HjBjδrs<br />

is the magnetic stress tensor. In matrix form the magnetic stress tensor is written<br />

T M rs =<br />

⎡<br />

⎣ B1H1 − 1<br />

The total electromagnetic stress tensor is<br />

Then the equation (2.6.87) can be written in the form<br />

or<br />

2BjHj B1H2 B1H3<br />

B2H1 B2H2 − 1<br />

2BjHj B2H3<br />

B3H1 B3H2 B3H3 − 1<br />

2BjHj (2.6.90)<br />

(2.6.91)<br />

⎤<br />

⎦ . (2.6.92)<br />

Trs = T E rs + T M rs . (2.6.93)<br />

Trs,s − ρEr = ɛrisJiBs + ∂<br />

∂t (ɛrisDiBs)<br />

ρEr + ɛrisJiBS = Trs,s − ∂<br />

∂t (ɛrisDiBs). (2.6.94)<br />

For free space Di = ɛ0Ei and Bi = µ0Hi so that the last term of equation (2.6.94) can be written in terms<br />

of the Poynting vector as<br />

∂Sr<br />

µ0ɛ0<br />

∂t<br />

= ∂<br />

∂t (ɛrisDiBs). (2.6.95)<br />

Now integrate the equation (2.6.94) over the volume to obtain the total electromagnetic force<br />

<br />

<br />

<br />

<br />

∂Sr<br />

ρEr dτ + ɛrisJiBs dτ = Trs,s dτ − µ0ɛ0<br />

V<br />

V<br />

V<br />

V ∂t dτ.<br />

Applying the divergence theorem of Gauss gives<br />

<br />

<br />

<br />

<br />

ρEr dτ + ɛrisJiBs dτ = Trsns dσ − µ0ɛ0<br />

V<br />

V<br />

S<br />

V<br />

∂Sr<br />

∂t<br />

dτ. (2.6.96)<br />

The left side of the equation (2.6.96) represents the forces acting on charges and currents contained within<br />

the volume element. If the electric and magnetic fields do not vary with time, then the last term on the<br />

right is zero. In this case the forces can be expressed as an integral of the electromagnetic stress tensor.


EXERCISE 2.6<br />

◮ 1. Find the field lines and equipotential curves associated with a positive charge q located at (−a, 0) and<br />

a positive charge q located at (a, 0). The field lines are illustrated in the figure 2.6-7.<br />

Figure 2.6-7. Lines of electric force between two charges of the same sign.<br />

◮ 2. Calculate the lines of force and equipotential curves associated with the electric field<br />

E = E(x, y) =2y e1 +2x e2. Sketch the lines of force and equipotential curves. Put arrows on the lines of<br />

force to show direction of the field lines.<br />

◮ 3. A right circular cone is defined by<br />

x = u sin θ0 cos φ, y = u sin θ0 sin φ, z = u cos θ0<br />

with 0 ≤ φ ≤ 2π and u ≥ 0. Show the solid angle subtended by this cone is Ω = A<br />

r2 =2π(1 − cos θ0).<br />

◮ 4. Acharge+qis located at the point (0,a) and a charge −q is located at the point (0, −a). Show that<br />

the electric force E at the position (x, 0), where x>ais E = 1 −2aq<br />

4πɛ0 (a2 + x2 e2.<br />

) 3/2<br />

◮ 5. Let the circle x2 + y2 = a2 carry a line charge λ∗ . Show the electric field at the point (0, 0,z)is<br />

E = 1 λ<br />

4πɛ0<br />

∗az(2π) e3<br />

(a2 + z2 .<br />

) 3/2<br />

◮ 6. Use superposition to find the electric field associated with two infinite parallel plane sheets each<br />

carrying an equal but opposite sign surface charge density µ ∗ . Find the field between the planes and outside<br />

of each plane. Hint: Fields are of magnitude ± µ∗<br />

and perpendicular to plates.<br />

2ɛ0<br />

◮ 7. For a volume current J the Biot-Savart law gives B = µ0<br />

<br />

J × er<br />

4π V r2 dτ. Show that ∇· B =0.<br />

Hint: Let er = r<br />

r and consider ∇·( J × r<br />

). Then use numbers 13 and 10 of the appendix C. Also note that<br />

r3 ∇× J = 0 because J does not depend upon position.<br />

347


348<br />

◮ 8. A homogeneous dielectric is defined by Di and Ei having parallel vector fields. Show that for a<br />

homogeneous dielectric ɛ j<br />

i,k =0.<br />

◮ 9. Show that for a homogeneous, isotropic dielectric medium that ɛ is a constant.<br />

◮ 10. Show that for a homogeneous, isotropic linear dielectric in Cartesian coordinates<br />

Pi,i = αe<br />

ρf .<br />

1+αe<br />

◮ 11. Verify the Maxwell’s equations in Gaussian units for a charge free isotropic homogeneous dielectric.<br />

∇· E = 1<br />

ɛ ∇· D =0<br />

∇· B =µ∇ H =0<br />

∇× E = − 1 ∂<br />

c<br />

B µ ∂<br />

= −<br />

∂t c<br />

H<br />

∂t<br />

∇× H = 1 ∂<br />

c<br />

D 4π<br />

+<br />

∂t c J = ɛ ∂<br />

c<br />

E<br />

∂t<br />

+ 4π<br />

c σ E<br />

◮ <strong>12</strong>.<br />

charge.<br />

Verify the Maxwell’s equations in Gaussian units for an isotropic homogeneous dielectric with a<br />

∇· D =4πρ<br />

∇· ∇×<br />

B =0<br />

E = − 1 ∂<br />

c<br />

B<br />

∂t<br />

∇× H = 4π<br />

c J + 1 ∂<br />

c<br />

D<br />

∂t<br />

◮ 13. For a volume charge ρ in an element of volume dτ located at a point (ξ,η,ζ) Coulombs law is<br />

E(x, y, z) = 1<br />

<br />

4πɛ0<br />

ρ<br />

erdτ<br />

r2 (a) Show that r 2 =(x− ξ) 2 +(y − η) 2 +(z − ζ) 2 .<br />

(b) Show that er = 1<br />

r ((x − ξ) e1 +(y − η) e2 +(z − ζ) e3) .<br />

(c) Show that<br />

E(x, y, z) = 1<br />

<br />

4πɛ0 V [(x − ξ) 2 +(y − η) 2 +(z − ζ) 2 ]<br />

(d) Show that the potential function for E is V = 1<br />

<br />

<br />

(x − ξ) e1 +(y − η) e2 +(z − ζ) e3<br />

1<br />

ρdξdηdζ =<br />

3/2 4πɛ0<br />

V<br />

V<br />

∇<br />

<br />

er<br />

r2 <br />

ρdξdηdζ<br />

ρ(ξ,η,ζ)<br />

[(x − ξ) 2 +(y − η) 2 +(z − ζ) 2 dξdηdζ<br />

] 1/2<br />

4πɛ0 V<br />

(e) Show that E = −∇V.<br />

(f) Show that ∇ 2 V = − ρ<br />

Hint: Note that the integrand is zero everywhere except at the point where<br />

ɛ<br />

(ξ,η,ζ) = (x, y, z). Consider the integral split into two regions. One region being a small sphere<br />

about the point (x, y, z) in the limit as the radius of this sphere approaches zero. Observe the identity<br />

er<br />

∇ (x,y,z)<br />

r2 <br />

<br />

er<br />

= −∇(ξ,η,ζ)<br />

r2 <br />

enables one to employ the Gauss divergence theorem to obtain a<br />

surface integral. Use a mean value theorem to show − ρ<br />

<br />

er ρ<br />

· ˆndS = 4π since ˆn = − er.<br />

4πɛ0 S r2 4πɛ0<br />

◮ 14. Show that for a point charge in space ρ∗ = qδ(x − x0)δ(y − y0)δ(z − z0), where δ is the Dirac delta<br />

function, the equation (2.6.5) can be reduced to the equation (2.6.1).<br />

◮ 15.<br />

(a) Show the electric field E = 1<br />

r2 er is irrotational. Here er = r<br />

r is a unit vector in the direction of r.<br />

(b) Find the potential function V such that E = −∇V which satisfies V(r0) =0forr0 > 0.


◮ 16.<br />

(a) If E is a conservative electric field such that E = −∇V, then show that E is irrotational and satisfies<br />

∇× E =curl E =0.<br />

(b) If ∇× E =curl E = 0, show that E is conservative. (i.e. Show E = −∇V.)<br />

Hint: The work done on a test charge Q = 1 along the straight line segments from (x0,y0,z0) to<br />

(x, y0,z0) andthenfrom(x, y0,z0) to(x, y, z0) and finally from (x, y, z0) to(x, y, z) can be written<br />

x<br />

y<br />

z<br />

V = V(x, y, z) =− E1(x, y0,z0) dx − E2(x, y, z0) dy − E3(x, y, z) dz.<br />

Now note that<br />

x0<br />

∂V<br />

∂y = −E2(x,<br />

z<br />

∂E3(x, y, z)<br />

y, z0) −<br />

dz<br />

z0 ∂y<br />

and from ∇× E = 0 we find ∂E3 ∂E2<br />

∂V<br />

= , which implies<br />

∂y ∂z ∂y = −E2(x, y, z). Similar results are obtained<br />

for ∂V ∂V<br />

and<br />

∂x ∂z . Hence show −∇V = E.<br />

◮ 17.<br />

(a) Show that if ∇· B = 0, then there exists some vector field A such that B = ∇× A.<br />

The vector field A is called the vector potential of B.<br />

Hint: Let 1<br />

A(x, y, z) = s B(sx, sy, sz) × rdswhere r = x e1 + y e2 + z e3<br />

1<br />

dBi<br />

and integrate<br />

0 ds s2 ds by parts.<br />

(b) Show that ∇·(∇× A)=0.<br />

◮ 18. Use Faraday’s law and Ampere’s law to show<br />

0<br />

g im (E j<br />

,j ),m − g jm E i ,mj<br />

y0<br />

∂<br />

= −µ0<br />

∂t<br />

<br />

J i ∂E<br />

+ ɛ0<br />

i <br />

∂t<br />

◮ 19. Assume that J = σ E where σ is the conductivity. Show that for ρ = 0 Maxwell’s equations produce<br />

µ0σ ∂ E<br />

∂t<br />

and µ0σ ∂ B<br />

∂t<br />

∂<br />

+ µ0ɛ0<br />

2E ∂t2 =∇2E + µ0ɛ0<br />

∂ 2 B<br />

∂t 2 =∇2 B.<br />

Here both E and B satisfy the same equation which is known as the telegrapher’s equation.<br />

◮ 20. Show that Maxwell’s equations (2.6.75) through (2.6.78) for the electric field under electrostatic<br />

conditions reduce to<br />

∇× E =0<br />

∇· D =ρf<br />

Now E is irrotational so that E = −∇V. Show that ∇ 2 V = − ρf<br />

ɛ .<br />

z0<br />

349


350<br />

◮ 21. Show that Maxwell’s equations (2.6.75) through (2.6.78) for the magnetic field under magnetostatic<br />

conditions reduce to ∇× H = J and ∇· B =0. The divergence of B being zero implies B can be derived<br />

from a vector potential function A such that B = ∇× A.Here A is not unique, see problem 24. If we select<br />

A such that ∇· A = 0 then show for a homogeneous, isotropic material, free of any permanent magnets, that<br />

∇ 2 A = −µ J.<br />

◮ 22. Show that under nonsteady state conditions of electrodynamics the Faraday law from Maxwell’s<br />

equations (2.6.75) through (2.6.78) does not allow one to set E = −∇V. Why is this? Observe that<br />

∇· B =0sowecanwrite B = ∇× A for some vector potential <br />

A. Using this vector potential show that<br />

Faraday’s law can be written ∇× E + ∂ <br />

A<br />

= 0. This shows that the quantity inside the parenthesis is<br />

∂t<br />

conservative and so we can write E + ∂ A<br />

= −∇V for some scalar potential V. The representation<br />

∂t<br />

E = −∇V − ∂ A<br />

∂t<br />

is a more general representation of the electric potential. Observe that for steady state conditions ∂ A<br />

∂t =0<br />

so that this potential representation reduces to the previous one for electrostatics.<br />

◮ 23. Using the potential formulation E = −∇V − ∂ A<br />

derived in problem 22, show that in a vacuum<br />

∂t<br />

(a) Gauss law can be written ∇ 2 V + ∂∇· A ρ<br />

= −<br />

∂t ɛ0<br />

(b) Ampere’s law can be written<br />

<br />

∇× ∇× <br />

A = µ0 <br />

∂V ∂<br />

J − µ0ɛ0∇ − µ0ɛ0<br />

∂t<br />

2A ∂t2 (c) Show the result in part (b) can also be expressed in the form<br />

<br />

∇ 2 A ∂<br />

− µ0ɛ0<br />

<br />

A<br />

−∇ ∇·<br />

∂t<br />

<br />

∂V<br />

A + µ0ɛ0 = −µ0<br />

∂t<br />

J<br />

◮ 24. The Maxwell equations in a vacuum have the form<br />

∇× E = − ∂ B<br />

∂t<br />

∇× H = ∂ D<br />

∂t + ρ V ∇· D = ρ ∇· B =0<br />

where D = ɛ0 E, B = µ0 H with ɛ0 and µ0 constants satisfying ɛ0 µ0 =1/c2 where c is the speed of light.<br />

Introduce the vector potential A and scalar potential V defined by B = ∇× A and E <br />

∂<br />

= − A<br />

∂t −∇V.<br />

Note that the vector potential is not unique. For example, given ψ as a scalar potential we can write<br />

B = ∇× A = ∇×( A + ∇ ψ), since the curl of a gradient is zero. Therefore, it is customary to impose some<br />

kind of additional requirement on the potentials. These additional conditions are such that E and B are<br />

not changed. One such condition is that A and V satisfy ∇· A + 1<br />

c2 ∂V<br />

=0. This relation is known as the<br />

∂t<br />

Lorentz relation or Lorentz gauge. Find the Maxwell’s equations in a vacuum in terms of A and V and show<br />

that <br />

∇ 2 − 1<br />

c2 ∂2 ∂t2 <br />

V = − ρ<br />

ɛ0<br />

and<br />

<br />

∇ 2 − 1<br />

c2 ∂2 ∂t2 <br />

A = −µ0ρ V.


◮ 25. In a vacuum show that E and B satisfy<br />

∇ 2 E <br />

1<br />

=<br />

c2 ∂2E ∂t2 ∇ 2 B <br />

1<br />

=<br />

c2 ∂2B ∂t2 ∇· E =0 ∇ B =0<br />

◮ 26.<br />

(a) Show that the wave equations in problem 25 have solutions in the form of waves traveling in the<br />

x- direction given by<br />

E = E(x, t) = E0e i(kx±ωt)<br />

and<br />

B = B(x, t) = B0e i(kx±ωt)<br />

where E0 and B0 are constants. Note that wave functions of the form u = Ae i(kx±ωt) are called plane<br />

harmonic waves. Sometimes they are called monochromatic waves. Here i2 = −1 is an imaginary unit.<br />

Euler’s identity shows that the real and imaginary parts of these type wave functions have the form<br />

A cos(kx ± ωt) and A sin(kx ± ωt).<br />

These represent plane waves. The constant A is the amplitude of the wave , ω is the angular frequency,<br />

and k/2π is called the wave number. The motion is a simple harmonic motion both in time and space.<br />

That is, at a fixed point x the motion is simple harmonic in time and at a fixed time t, themotionis<br />

harmonic in space. By examining each term in the sine and cosine terms we find that x has dimensions of<br />

length, k has dimension of reciprocal length, t hasdimensionsoftimeandωhas dimensions of reciprocal<br />

time or angular velocity. The quantity c = ω/k is the wave velocity. The value λ =2π/k has dimension<br />

of length and is called the wavelength and 1/λ is called the wave number. The wave number represents<br />

the number of waves per unit of distance along the x-axis. The period of the wave is T = λ/c =2π/ω<br />

and the frequency is f =1/T. The frequency represents the number of waves which pass a fixed point<br />

in a unit of time.<br />

(b) Show that ω =2πf<br />

(c) Show that c = fλ<br />

(d) Is the wave motion u =sin(kx − ωt)+sin(kx + ωt) a traveling wave? Explain.<br />

(e) Show that in general the wave equation ∇ 2 φ = 1<br />

c2 ∂2φ have solutions in the form of waves traveling in<br />

∂t2 either the +x or −x direction given by<br />

φ = φ(x, t) =f(x + ct)+g(x − ct)<br />

where f and g are arbitrary twice differentiable functions.<br />

(f) Assume a plane electromagnetic wave is moving in the +x direction. Show that the electric field is in<br />

the xy−plane and the magnetic field is in the xz−plane.<br />

Hint: Assume solutions Ex = g1(x − ct), Ey = g2(x − ct),Ez = g3(x − ct),Bx = g4(x − ct),<br />

By = g5(x − ct),Bz = g6(x − ct) wheregi,i =1, ..., 6 are arbitrary functions. Then show that Ex<br />

does not satisfy ∇· E = 0 which implies g1 must be independent of x and so not a wave function. Do<br />

the same for the components of B. Since both ∇· E = ∇· B =0thenEx = Bx = 0. Such waves<br />

are called transverse waves because the electric and magnetic fields are perpendicular to the direction<br />

of propagation. Faraday’s law implies that the E and B waves must be in phase and be mutually<br />

perpendicular to each other.<br />

351


352<br />

BIBLIOGRAPHY<br />

• Abramowitz, M. and Stegun, I.A., Handbook of Mathematical Functions, 10thed,<br />

New York:Dover, 1972.<br />

• Akivis, M.A., Goldberg, V.V., An Introduction to Linear Algebra and Tensors, New York:Dover, 1972.<br />

• Aris, Rutherford, Vectors, Tensors, and the Basic Equations of Fluid Mechanics,<br />

Englewood Cliffs, N.J.:Prentice-Hall, 1962.<br />

• Atkin, R.J., Fox, N., An Introduction to the Theory of Elasticity,<br />

London:Longman Group Limited, 1980.<br />

• Bishop, R.L., Goldberg, S.I.,Tensor Analysis on Manifolds, New York:Dover, 1968.<br />

• Borisenko, A.I., Tarapov, I.E., Vector and Tensor Analysis with Applications, New York:Dover, 1968.<br />

• Chorlton, F., Vector and Tensor Methods, Chichester,England:Ellis Horwood Ltd, 1976.<br />

• Dodson, C.T.J., Poston, T., Tensor Geometry, London:Pittman Publishing Co., 1979.<br />

• Eisenhart, L.P., Riemannian Geometry, Princeton, N.J.:Univ. Princeton Press, 1960.<br />

• Eringen, A.C., Mechanics of Continua, Huntington, N.Y.:Robert E. Krieger, 1980.<br />

• D.J. Griffiths, Introduction to Electrodynamics, Prentice Hall, 1981.<br />

• Flügge, W., Tensor Analysis and Continuum Mechanics, New York:Springer-Verlag, 1972.<br />

• Fung, Y.C., A First Course in Continuum Mechanics, Englewood Cliffs,N.J.:Prentice-Hall, 1969.<br />

• Goodbody, A.M., Cartesian Tensors, Chichester, England:Ellis Horwood Ltd, 1982.<br />

• Hay, G.E., Vector and Tensor Analysis, New York:Dover, 1953.<br />

• Hughes, W.F., Gaylord, E.W., Basic Equations of Engineering Science, New York:McGraw-Hill, 1964.<br />

• Jeffreys, H., Cartesian Tensors, Cambridge, England:Cambridge Univ. Press, 1974.<br />

• Lass, H., Vector and Tensor Analysis, New York:McGraw-Hill, 1950.<br />

• Levi-Civita, T., The Absolute Differential Calculus, London:Blackie and Son Limited, 1954.<br />

• Lovelock, D., Rund, H. ,Tensors, Differential Forms, and Variational Principles, New York:Dover, 1989.<br />

• Malvern, L.E., Introduction to the Mechanics of a Continuous Media,<br />

Englewood Cliffs, N.J.:Prentice-Hall, 1969.<br />

• McConnell, A.J., Application of Tensor Analysis, New York:Dover, 1947.<br />

• Newell, H.E., Vector Analysis, New York:McGraw Hill, 1955.<br />

• Schouten, J.A., Tensor Analysis for Physicists,New York:Dover, 1989.<br />

• Scipio, L.A., Principles of Continua with Applications, New York:John Wiley and Sons, 1967.<br />

• Sokolnikoff, I.S., Tensor Analysis, New York:John Wiley and Sons, 1958.<br />

• Spiegel, M.R., Vector Analysis, New York:Schaum Outline Series, 1959.<br />

• Synge, J.L., Schild, A., Tensor Calculus, Toronto:Univ. Toronto Press, 1956.<br />

Bibliography


Prefixes.<br />

Basic Units.<br />

APPENDIX A<br />

UNITS OF MEASUREMENT<br />

The following units, abbreviations and prefixes are from the<br />

Système International d’Unitès (designated SI in all Languages.)<br />

Abreviations<br />

Prefix Multiplication factor Symbol<br />

tera 10 <strong>12</strong> T<br />

giga 10 9 G<br />

mega 10 6 M<br />

kilo 10 3 K<br />

hecto 10 2 h<br />

deka 10 da<br />

deci 10 −1 d<br />

centi 10 −2 c<br />

milli 10 −3 m<br />

micro 10 −6 µ<br />

nano 10 −9 n<br />

pico 10 −<strong>12</strong> p<br />

Basic units of measurement<br />

Unit Name Symbol<br />

Length meter m<br />

Mass kilogram kg<br />

Time second s<br />

Electric current ampere A<br />

Temperature degree Kelvin ◦ K<br />

Luminous intensity candela cd<br />

Supplementary units<br />

Unit Name Symbol<br />

Plane angle radian rad<br />

Solid angle steradian sr<br />

353


354<br />

DERIVED UNITS<br />

Name Units Symbol<br />

Area square meter m 2<br />

Volume cubic meter m 3<br />

Frequency hertz Hz (s −1 )<br />

Density kilogram per cubic meter kg/m 3<br />

Velocity meter per second m/s<br />

Angular velocity radian per second rad/s<br />

Acceleration meter per second squared m/s 2<br />

Angular acceleration radian per second squared rad/s 2<br />

Force newton N (kg· m/s 2 )<br />

Pressure newton per square meter N/m 2<br />

Kinematic viscosity square meter per second m 2 /s<br />

Dynamic viscosity newton second per square meter N · s/m 2<br />

Work, energy, quantity of heat joule J (N· m)<br />

Power watt W (J/s)<br />

Electric charge coulomb C (A· s)<br />

Voltage, Potential difference volt V (W/A)<br />

Electromotive force volt V (W/A)<br />

Electric force field volt per meter V/m<br />

Electric resistance ohm Ω (V/A)<br />

Electric capacitance farad F (A· s/V)<br />

Magnetic flux weber Wb (V · s)<br />

Inductance henry H (V· s/A)<br />

Magnetic flux density tesla T (Wb/m 2 )<br />

Magnetic field strength ampere per meter A/m<br />

Magnetomotive force ampere A<br />

Physical constants.<br />

4arctan1=π =3.14159 26535 89793 23846 2643 ...<br />

<br />

lim 1+<br />

n→∞<br />

1<br />

n = e =2.71828 18284 59045 23536 0287 ...<br />

n<br />

Euler’s constant γ =0.5772156649 01532 86060 65<strong>12</strong> ...<br />

<br />

γ = lim 1+<br />

n→∞<br />

1<br />

<br />

1 1<br />

+ + ···+ − log n<br />

2 3 n<br />

speed of light in vacuum = 2.997925(10) 8 ms −1<br />

electron charge = 1.60210(10) −19 C<br />

Avogadro’s constant = 6.02252(10) 23 mol −1<br />

Plank’s constant = 6.6256(10) −34 Js<br />

Universal gas constant = 8.3143 JK −1 mol −1 = 8314.3 JKg −1 K −1<br />

Boltzmann constant = 1.38054(10) −23 JK −1<br />

Stefan–Boltzmann constant = 5.6697(10) −8 Wm −2 K −4<br />

Gravitational constant = 6.67(10) −11 Nm 2 kg −2


APPENDIX B<br />

CHRISTOFFEL SYMBOLS OF SECOND KIND<br />

1. Cylindrical coordinates (r, θ, z) =(x 1 ,x 2 ,x 3 )<br />

x = r cos θ<br />

y = r sin θ<br />

z = z<br />

r ≥ 0<br />

0 ≤ θ ≤ 2π<br />

−∞


356<br />

3. Parabolic cylindrical coordinates (ξ,η,z) =(x 1 ,x 2 ,x 3 )<br />

x = ξη<br />

y = 1<br />

2 (ξ2 − η 2 )<br />

z = z<br />

−∞


5. Elliptic cylindrical coordinates (ξ,η,z) =(x 1 ,x 2 ,x 3 )<br />

x =coshξ cos η<br />

y =sinhξ sin η<br />

z = z<br />

ξ ≥ 0<br />

0 ≤ η ≤ 2π<br />

−∞


358<br />

7. Bipolar coordinates (u, v, z) =(x1 ,x2 ,x3 )<br />

a sinh v<br />

x =<br />

,<br />

cosh v − cos u<br />

0 ≤ u


9. Prolate spheroidal coordinates (u, v, φ) =(x 1 ,x 2 ,x 3 )<br />

x = a sinh u sin v cos φ, u ≥ 0<br />

y = a sinh u sin v sin φ, 0 ≤ v ≤ π<br />

h 2 1 = h2 2<br />

h 2 2 = a2 (sinh 2 u +sin 2 v)<br />

z = a cosh u cos v, 0 ≤ φ


360<br />

11. Toroidal coordinates (u, v, φ) =(x 1 ,x 2 ,x 3 )<br />

a sinh v cos φ<br />

x = ,<br />

cosh v − cos u<br />

0 ≤ u


<strong>12</strong>. Confocal ellipsoidal coordinates (u, v, w) =(x 1 ,x 2 ,x 3 )<br />

x 2 = (a2 − u)(a2 − v)(a2 − w)<br />

(a2 − b2 )(a2 − c2 , u < c<br />

)<br />

2


362<br />

APPENDIX C<br />

VECTOR IDENTITIES<br />

The following identities assume that A, B, C, D are differentiable vector functions of position while<br />

f,f1,f2 are differentiable scalar functions of position.<br />

1.<br />

2.<br />

A · ( B × C)= B · ( C × A)= C · ( A × B)<br />

A × ( B × C)= B( A · C) − C( A · B)<br />

3. ( A × B) · ( C × D)=( A · C)( B · D) − ( A · D)( B · C)<br />

4.<br />

A × ( B × C)+ B × ( C × A)+ C × ( A × B)=0<br />

5. ( A × B) × ( C × D)= B( A · C × D) − A( B · C × D)<br />

= C( A · B × C) − D( A · B × C)<br />

6. ( A × B) · ( B × C) × ( C × A)=( A · B × C) 2<br />

7. ∇(f1 + f2) =∇f1 + ∇f2<br />

8. ∇·( A + B)=∇· A + ∇· B<br />

9. ∇×( A + B)=∇× A + ∇× B<br />

10. ∇(f A)=(∇f) · A + f∇· A<br />

11. ∇(f1f2) =f1∇f2 + f2∇f1<br />

<strong>12</strong>. ∇×(f A)=)∇f) × A + f(∇× A)<br />

13. ∇·( A × B)= B · (∇× A) − A · (∇× B)<br />

14. ( A ·∇) <br />

|<br />

A = ∇<br />

A| 2<br />

<br />

−<br />

2<br />

A × (∇× A)<br />

15. ∇( A · B)=( B ·∇) A +( A ·∇) B + B × (∇× A)+ A × (∇× B)<br />

16. ∇×( A × B)=( B ·∇) A − B(∇· A) − ( A ·∇) B + A(∇· B)<br />

17. ∇·(∇f) =∇ 2 f<br />

18. ∇×(∇f) =0<br />

19. ∇·(∇× A)=0<br />

20. ∇×(∇× A)=∇(∇· A) −∇ 2 A


A<br />

Absolute differentiation <strong>12</strong>0<br />

Absolute scalar field 43<br />

Absolute tensor 45,46,47,48<br />

Acceleration <strong>12</strong>1, 190, 192<br />

Action integral 198<br />

Addition of systems 6, 51<br />

Addition of tensors 6, 51<br />

Adherence boundary condition 294<br />

Aelotropic material 245<br />

Affine transformation 86, 107<br />

Airy stress function 264<br />

Almansi strain tensor 229<br />

Alternating tensor 6,7<br />

Ampere’s law 176,301,337,341<br />

Angle between vectors 80, 82<br />

Angular momentum 218, 287<br />

Angular velocity 86,87,201,203<br />

Arc length60, 67, 133<br />

Associated tensors 79<br />

Auxiliary Magnetic field 338<br />

Axis of symmetry 247<br />

B<br />

Basic equations elasticity 236, 253, 270<br />

Basic equations for a continuum 236<br />

Basic equations of fluids 281, 287<br />

Basis vectors 1,2,37,48<br />

Beltrami 262<br />

Bernoulli’s Theorem 292<br />

Biharmonic equation 186, 265<br />

Bilinear form 97<br />

Binormal vector 130<br />

Biot-Savart law 336<br />

Bipolar coordinates 73<br />

Boltzmann equation 302,306<br />

Boundary conditions 257, 294<br />

Bulk modulus 251<br />

Bulk coefficient of viscosity 285<br />

C<br />

Cartesian coordinates 19,20,42, 67, 83<br />

Cartesian tensors 84, 87, 226<br />

INDEX 363<br />

Cauchy stress law 216<br />

Cauchy-Riemann equations 293,321<br />

Charge density 323<br />

Christoffel symbols 108,110,111<br />

Circulation 293<br />

Codazzi equations 139<br />

Coefficient of viscosity 285<br />

Cofactors 25, 26, 32<br />

Compatibility equations 259, 260, 262<br />

Completely skew symmetric system 31<br />

Compound pendulum 195,209<br />

Compressible material 231<br />

Conic sections 151<br />

Conical coordinates 74<br />

Conjugate dyad 49<br />

Conjugate metric tensor 36, 77<br />

Conservation of angular momentum 218, 295<br />

Conservation of energy 295<br />

Conservation of linear momentum 217, 295<br />

Conservation of mass 233, 295<br />

Conservative system 191, 298<br />

Conservative electric field 323<br />

Constitutive equations 242, 251,281, 287<br />

Continuity equation 106,234, 287, 335<br />

Contraction 6, 52<br />

Contravariant components 36, 44<br />

Contravariant tensor 45<br />

Coordinate curves 37, 67<br />

Coordinate surfaces 37, 67<br />

Coordinate transformations 37<br />

Coulomb law 322<br />

Covariant components 36, 47<br />

Covariant differentiation 113,114,117<br />

Covariant tensor 46<br />

Cross product 11<br />

Curl 21, 173<br />

Curvature 130, 131, 134, 149<br />

Curvature tensor 134, 145<br />

Curvilinear coordinates 66, 81<br />

Cylindrical coordinates 18, 42, 69


364 INDEX<br />

D<br />

Deformation 222<br />

Derivative of tensor 108<br />

Derivatives and indicial notation 18, 31<br />

Determinant 10, 25, 32, 33<br />

Dielectric tensor 333<br />

Differential geometry <strong>12</strong>9<br />

Diffusion equation 303<br />

Dilatation 232<br />

Direction cosines 85<br />

Displacement vector 333<br />

Dissipation function 297<br />

Distribution function 302<br />

Divergence 21, 172<br />

Divergence theorem 24<br />

Dot product 5<br />

Double dot product 50, 62<br />

Dual tensor 100<br />

Dummyindex4,5<br />

Dyads 48,62,63<br />

Dynamics 187<br />

E<br />

e Permutation symbol 6, 7, <strong>12</strong><br />

e-δ identity <strong>12</strong><br />

Eigenvalues 179,189<br />

Eigenvectors 179,186<br />

Einstein tensor 156<br />

Elastic constants 248<br />

Elastic stiffness 242<br />

Elasticity 211,213<br />

Electrostatic field 322,333<br />

Electric flux 327<br />

Electric units 322<br />

Electrodynamics 339<br />

Electromagnetic energy 341<br />

Electromagnetic stress 341,342<br />

Elliptic coordinates 72<br />

Elliptical cylindrical coordinates 71<br />

Enthalpy 298<br />

Entropy 300<br />

Epsilon permutation symbol 83<br />

Equation of state 300<br />

Equilibrium equations 273,300<br />

Elastic constants 243,248<br />

Equipotential curves 325<br />

Eulernumber294<br />

Euler-Lagrange equations 192<br />

Eulerian angles 201, 209<br />

Eulerian form 287<br />

Eulerian system 227<br />

Eulers equations of motion 204<br />

F<br />

Faraday’s law 176,301, 340<br />

Field lines 324, 327<br />

Field electric 322<br />

First fundamental form 133,143<br />

Fourier law 297, 299<br />

Free indices 3<br />

Frenet-Serret formulas 131, 188<br />

Froude number 294<br />

Fluids 281<br />

G<br />

Gas law 300<br />

Gauss divergence theorem 24, 330<br />

Gauss equations 138<br />

Gauss’s law for electricity 176,301,328<br />

Gauss’s law for magnetism 176,301,341<br />

Gaussian curvature 137,139, 149<br />

Geodesics 140, 146<br />

Geodesic curvature 135, 140<br />

General tensor 48<br />

Generalized e − δ identity 84, 104<br />

Generalized Hooke’s law 242<br />

Generalized Kronecker delta 13, 31<br />

Generalized stress strain 242<br />

Geometry in Riemannian Space 80<br />

Gradient 20, 171<br />

Gradient basis 37<br />

Green’s theorem 24<br />

Group properties 41, 54<br />

Generalized velocity <strong>12</strong>1<br />

Generalized acceleration <strong>12</strong>1


H<br />

Hamiltonian 208<br />

Heat equation 316<br />

Hexagonal material 247<br />

Higher order tensors 47, 93<br />

Hooke’s law 2<strong>12</strong>, 242, 252<br />

Hydrodynamic equations 283<br />

I<br />

Ideal fluid 283<br />

Idemfactor 50<br />

Incompressible material 231<br />

Index notation 1, 2, 14<br />

Indicial notation 1, 2, 14,24<br />

Inner product 52<br />

Inertia 30<br />

Integral theorems 24<br />

Intrinsic derivative <strong>12</strong>0<br />

Invariant 43<br />

Inviscid fluid 283<br />

Isotropic material 248<br />

Isotropic tensor 104<br />

J<br />

Jacobian 17, 30, 40, 101, <strong>12</strong>7<br />

Jump discontinuity 330<br />

K<br />

Kronecker delta 3, 8, 13, 31, 76<br />

Kinetic energy 201<br />

Kinematic viscosity 302<br />

L<br />

Lagrange’s equation of motion 191, 196<br />

Lagrangian 209<br />

Laplacian 174<br />

Linear form 96<br />

Linear momentum 209, 287<br />

Linear transformation 86<br />

Linear viscous fluids 284<br />

Lorentz transformation 57<br />

Lame’s constants 251<br />

INDEX 365<br />

M<br />

Magnitude of vector 80<br />

Magnetostatics 334,338<br />

Magnetic field 334<br />

Magnetization vector 337<br />

Magnetic permeability 337<br />

Material derivative 234, 288<br />

Material symmetry 244, 246<br />

Maxwell equations 176, 339<br />

Maxwell transfer equation 308<br />

Maximum, minimum curvature 130, 140<br />

Mean curvature 137, 148<br />

Metric tensor 36, 65<br />

Meusnier’s Theorem 150<br />

Mixed tensor 49<br />

Mohr’s circle 185<br />

Moment of inertia 30, 184, 200<br />

Momentum 217, 218<br />

Multilinear forms 96, 98<br />

Multiplication of tensors 6, 51<br />

N<br />

Navier’s equations 254, 257<br />

Navier-Stokes equations 288, 290<br />

Newtonian fluids 286<br />

Nonviscous fluid 283<br />

Normal curvature 135, 136<br />

Normal plane 188<br />

Normal stress 214<br />

Normal vector 130, 132<br />

Notation for physical components 92<br />

O<br />

Oblate Spheroidal coordinates 75<br />

Oblique coordinates 60<br />

Oblique cylindrical coordinates 102<br />

Order 2<br />

Orthogonal coordinates 78, 86<br />

Orthotropic material 246<br />

Outer product 6, 51<br />

Osculating plane 188


366 INDEX<br />

P<br />

Parallel vector field <strong>12</strong>2<br />

Pappovich-Neuber solution 263<br />

Parabolic coordinates 70<br />

Parabolic cylindrical coordinates 69<br />

Particle motion 190<br />

Pendulum system 197, 210<br />

Perfect gas 283, 299<br />

Permutations 6<br />

Phase space 302<br />

Physical components 88, 91,93<br />

Piezoelectric 300<br />

Pitch,roll, Yaw 209<br />

Plane Couette flow 315<br />

Plane Poiseuille flow 316<br />

Plane strain 263<br />

Plane stress 264<br />

Poisson’s equation 329<br />

Poisson’s ratio 2<strong>12</strong><br />

Polar element 273<br />

Polarization vector 333<br />

Polyads 48<br />

Potential energy 191<br />

Potential function 323<br />

Poynting’s vector 341<br />

Pressure 283<br />

Principal axes 183<br />

Projection 35<br />

Prolated Spheroidal coordinates 74<br />

Pully system 194, 207<br />

Q<br />

Quotient law 53<br />

R<br />

Radius of curvature 130, 136<br />

Range convention 2, 3<br />

Rate of deformation 281, 286<br />

Rate of strain 281<br />

Rayleighimplusive flow 317<br />

Reciprocal basis 35, 38<br />

Relative scalar <strong>12</strong>7<br />

Relative tensor 50, <strong>12</strong>1<br />

Relative motion 202<br />

Relativity 151<br />

Relative motion 155<br />

Reynolds number 294<br />

Ricci’s theorem 119<br />

Riemann Christoffel tensor 116, <strong>12</strong>9,139, 147<br />

Riemann space 80<br />

Rectifying plane 188<br />

Rigid body rotation 199<br />

Rotation of axes 85, 87, 107<br />

Rules for indices 2<br />

S<br />

Scalar 40, 43<br />

Scalar invariant 43, 62, 105<br />

Scalar potential 191<br />

Scaled variables 293<br />

Second fundamental form 135, 145<br />

Second order tensor 47<br />

Shearing stresses 214<br />

Simple pulley system 193<br />

Simple pendulum 194<br />

Skew symmetric system 3, 31<br />

Skewed coordinates 60, 102<br />

Solid angle 328<br />

Space curves 130<br />

Special tensors 65<br />

Spherical coordinates 18, 43, 56, 69, 103,194<br />

Stokes flow 318<br />

Stokes hypothesis 285<br />

Stokes theorem 24<br />

Straight line 60<br />

Strain 218, 225, 228<br />

Strain deviator 279


Stress 214<br />

Stress deviator 279<br />

Strong conservative form 298<br />

Strouhal number 294<br />

St Venant 258<br />

Subscripts 2<br />

Subtraction of tensors 51, 62<br />

Summation convention 4, 9<br />

Superscripts 2<br />

Surface 62, 131<br />

Surface area 59<br />

Surface curvature 149<br />

Surface metric <strong>12</strong>5, 133<br />

Susceptibility tensor 333<br />

Sutherland formula 285<br />

Symmetric system 3, 31, 51, 101<br />

Symmetry 243<br />

System 2, 31<br />

T<br />

Tangential basis 37<br />

Tangent vector 130<br />

Tensor and vector forms 40, 150<br />

Tensor derivative 141<br />

Tensor general 48<br />

Tensor notation 92, 160<br />

Tensor operations 6, 51, 175<br />

Test charge 322<br />

Thermodynamics 299<br />

Third fundamental form 146<br />

Third order systems 31<br />

Toroidal coordinates 75, 103<br />

Torus <strong>12</strong>4<br />

Transformation equations 17, 37, 86<br />

Transitive property 45,46<br />

Translation of coordinates 84<br />

Transport equation 302<br />

Transposition 6<br />

Triad 50<br />

Trilinear form 98<br />

Triple scalar product 15<br />

INDEX 367<br />

U<br />

Unit binormal 131, 192<br />

Unit normal 131, 191<br />

Unit tangent 131, 191<br />

Unit vector 81, 105<br />

V<br />

Vector identities 15, 20, 315<br />

Vector transformation 45, 47<br />

Vector operators 20, 175<br />

Vector potential 188<br />

Velocity 95, <strong>12</strong>1, 190, 193<br />

Velocity strain tensor 281<br />

Viscosity 285<br />

Viscosity table 285<br />

Viscous fluid 283<br />

Viscous forces 288<br />

Viscous stress tensor 285<br />

Vorticity 107, 292<br />

W<br />

Wave equation 255, 269<br />

Weighted tensor 48, <strong>12</strong>7<br />

Weingarten’s equation 138, 153<br />

Work 191, 279<br />

Work done 324<br />

Y<br />

Young’s modulus 2<strong>12</strong>

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!