Solving Partial Differential Equations in Cylindrical8-2 Solving Laplace’s Equation in Cylindrical Coordinates 9.. Solving Partial Differential Equations in Spherical 9-2 The Solution to
Trang 4Differential Equations
Trang 5www.TheSolutionManual.com
Trang 6Mathematical Physics
with Partial Differential
Equations
James R Kirkwood
Sweet Briar College
AMSTERDAM • BOSTON • HEIDELBERG • LONDON NEW YORK • OXFORD • PARIS • SAN DIEGO SAN FRANCISCO • SINGAPORE • SYDNEY • TOKYO
Trang 7© 2013 Elsevier Inc All rights reserved
No part of this publication may be reproduced or transmitted in any form or by any means, electronic or
mechanical, including photocopying, recording, or any information storage and retrieval system, without
permission in writing from the Publisher Details on how to seek permission, further information about the
Publisher’s permissions policies, and our arrangements with organizations such as the Copyright Clearance
Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions
This book and the individual contributions contained in it are protected under copyright by the
Publisher (other than as may be noted herein).
Notices
Knowledge and best practice in this field are constantly changing As new research and experience
broaden our understanding, changes in research methods, professional practices, or medical treatment
may become necessary.
Practitioners and researchers must always rely on their own experience and knowledge in evaluating and
using any information, methods, compounds, or experiments described herein In using such information
or methods they should be mindful of their own safety and the safety of others, including parties for
whom they have a professional responsibility.
To the fullest extent of the law, neither the Publisher, nor the authors, contributors, or editors, assume
any liability for any injury and/or damage to persons or property as a matter of products liability,
negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas
contained in the material herein.
Library of Congress Cataloging-in-Publication Data
A catalogue record for this book is available from the British Library.
For information on all Academic Press publications
visit our website at www.elsevierdirect.com
Printed in the United States of America
12 13 14 15 10 9 8 7 6 5 4 3 2 1
Trang 9Integrals of Scalar Functions Over Surfaces 83
Trang 104-4 Fourier Sine and Cosine Series 208
7-4 Solving the Wave Equation in Two Dimensions in Cartesian
Trang 118 Solving Partial Differential Equations in Cylindrical
8-2 Solving Laplace’s Equation in Cylindrical Coordinates
9 Solving Partial Differential Equations in Spherical
9-2 The Solution to Bessel’s Equation in
Trang 1211-2 Solving Differential Equations Using the Laplace Transform 356
Trang 13www.TheSolutionManual.com
Trang 14The major purposes of this book are to present partial differential equations
(PDEs) and vector analysis at an introductory level As such, it could be
con-sidered a beginning text in mathematical physics It is also designed to provide
a bridge from undergraduate mathematics to the first graduate mathematics
course in physics, applied mathematics, or engineering In these disciplines, it
is not unusual for such a graduate course to cover topics from linear algebra,
ordinary and partial differential equations, advanced calculus, vector analysis,
complex analysis, and probability and statistics at a highly accelerated pace
In this text we study in detail, but at an introductory level, a reduced list
of topics important to the disciplines above In partial differential equations,
we consider Green’s functions, the Fourier and Laplace transforms, and how
these are used to solve PDEs We also study using separation of variables to
solve PDEs in great detail Our approach is to examine the three prototypical
second-order PDEs—Laplace’s equation, the heat equation, and the wave
equation—and solve each equation with each method The premise is that in
doing so, the reader will become adept at each method and comfortable with
each equation
The other prominent area of the text is vector analysis While the usual
topics are discussed, an emphasis is placed on understanding concepts rather
than formulas For example, we view the curl and gradient as properties of a
vector field rather than simply as equations A significant—but optional—
portion of this area deals with curvilinear coordinates to reinforce the idea of
conversion of coordinate systems
Reasonable prerequisites for the course are a course in multivariable
cal-culus, familiarity with ordinary differential equations including the ability to
solve a second-order boundary problem with constant coefficients, and some
experience with linear algebra
In dealing with ordinary differential equations, we emphasize the linear
operator approach That is, we consider the problem as being an eigenvalue/
eigenvector problem for a self-adjoint operator In addition to eliminating
some tedious computations regarding orthogonality, this serves as a unifying
theme and an introduction to more advanced mathematics
The level of the text generally lies between that of the classic encyclopedic
texts of Boas and Kreysig and the newer text by McQuarrie, and the partial
differential equations books of Weinberg and Pinsky Topics such as Fourier
series are developed in a mathematically rigorous manner The section on
completeness of eigenfunctions of a Sturm-Liouville problem is considerably
xi
Trang 15more advanced than the rest of the text, and can be omitted if one wishes to
merely accept the result
The text can be used as a self-contained reference as well as an introductory
text There was a concerted effort to avoid situations where filling in details
of an argument would be a challenge This is done in part so that the text
could serve as a source for students in subsequent courses who felt “I know I’m
supposed to know how to derive this, but I don’t.” A couple of such examples
are the fundamental solution of Laplace’s equation and the spectrum of the
Laplacian
I want to give special thanks to Patricia Osborn of Elsevier Publishing
whose encouragement prompted me to turn a collection of disjointed notes
into what I hope is a readable and cohesive text, and also to Gene Wayne of
Boston University who provided valuable suggestions
James Radford Kirkwood
Trang 161-1 SELF-ADJOINT OPERATORS
The purpose of this text is to study some of the important equations and techniques
of mathematical physics It is a fortuitous fact that many of the most important
such equations are linear, and we can apply the well-developed theory of linear
operators We assume knowledge of basic linear algebra but review some
defini-tions, theorems, and examples that will be important to us
Definition: A linear operator (or linear function), from a vector space V to a
vector space W, is a functionL : V-W for which
Lða1 ^v11 a2 ^v2Þ5 a1Lð^v1Þ1 a2Lð^v2Þfor all ^v1; ^v2AVand scalars a1and a2
One of the most important linear operators for us will be
L½y 5 a0ðxÞyðxÞ1 a1ðxÞy0ðxÞ1 a2ðxÞyvðxÞ;
where a0(x), a1(x), and a2(x) are continuous functions
Definition: If L : V-V is a linear operator, then a nonzero vector ^v is an
eigenvector ofL with eigenvalue λ if Lð^vÞ 5 λ^v
Note that ^0 cannot be an eigenvector, but 0 can be an eigenvalue
Example: For L 5 d
dx, we haveLðeax
Þ5 d
dxðe
ax
Þ5 aeax;
so eaxis an eigenvector ofL with eigenvalue a
An extremely important example is forL 5 d2
dx2
Lðsin nxÞ 5 d2
dx2ðsin nxÞ5 2n2sin nx; and
1
Mathematical Physics with Partial Differential Equations
© 2013 Elsevier Inc All rights reserved.
Trang 17Lðcos nxÞ 5 d2
dx2ðcos nxÞ5 2 n2cos nx:
We leave it as exercise 1 to show that if ^v is an eigenvector of L with
eigenvalueλ, then a^v is also an eigenvector of L with eigenvalue λ
Definition: An inner product (also called a dot product) on a vector space V
with scalar field F (which is the real numbersℝ or the complex numbers ℂ)
is a function h , i : V3 V-F such that for all f, g, h A V, and a A F,
ha f; gi 5 ah f ; gi;
hf; agi 5 ahf ; gi; where x 1 iy 5 x 2 iy;
hf1 g; hi 5 h f ; hi 1 hg; hi;
hf; gi 5 hg; f i;
and h f, fi$ 0 with equality if and only if f 5 0
A vector space with an inner product is called an inner product space
If V5 ℝn, the usual inner product for ^a 5 ða1; ; anÞ; ^b5 ðb1; ; bnÞis
^a; ^b5 a1b11 ? 1 anbn:
If the vector space isℂn, then we must modify the definition, because, for
example, under this definition, if ^a 5 ði; iÞ, then
h^a; ^ai 5 i21 i25 22:
Thus, onℂn, for ^a 5 ða1; ; anÞ; ^b5 ðb1; ; bnÞ, we define
h^a; ^bi 5 a1b11 ? 1 anbn:
We use the notation h f, fi5 :f:2, which is interpreted as the square of
the length of f, and:f 2 g: is the distance from f to g
We will be working primarily with vector spaces consisting of functions
that satisfy some property such as continuity or differentiability In this
set-ting, one usually defines the inner product using an integral A common
inner product is
hf; gi 5
ðb a
f ðxÞgðxÞdx;where a or b may be finite or infinite There might be a problem with some
vector spaces in that h f, fi5 0 with f 6¼ 0 This problem can be overcome by
a minor modification of the vector space or by restricting the functions to
being continuous, and will not affect our work We leave it as exercise 4 to
show that the function defined above is an inner product
Trang 18On some occasions it will be advantageous to modify the inner product
above with a weight function w(x) If w(x)$ 0 on [a, b], then
hf; giw5
ðb a
f ðxÞwðxÞgðxÞdx
is also an inner product as we show in exercise 5
Definition: A linear operator A on the inner product space V is self-adjoint if
hAf, gi5 h f, Agi for all f, g A V
Self-adjoint operators are prominent in mathematical physics One
exam-ple is the Hamiltonian operator It is a fact (Stone’s Theorem) that energy is
conserved if and only if the Hamiltonian is self-adjoint Another example is
shown below Part of the significance of this example is due to Newton’s
law F5 ma
Example: The operator d 2
dx 2is self-adjoint on the inner product space
V5 ff jf has a continuous second derivative and is periodic on ½a; bg;
with inner product
hf; gi 5
ðb a
f ðxÞgvðxÞdx:
To do this, we integrate by parts twice Consider the integral on the left Let
u5 gðxÞ; du5 g0ðxÞ;
dv5 f vðxÞ; v 5 f0ðxÞ;so
fvðxÞgðxÞdx 5 2
ðb a
f0ðxÞg0ðxÞdx:Integrating the integral on the right by parts with
u5 g0ðxÞ; du 5 gvðxÞ;
dv5 f0ðxÞ; v 5 f ðxÞ;
Trang 19f ðxÞgvðxÞdx
5
ðb a
f ðxÞgvðxÞdx:
Notice that if [a, b] is of length 2π, then {sin (nx) cos (nx) j n A ℤ} is a
subset of V
We next prove two important facts about self-adjoint operators
Theorem: If L : V-V is a self-adjoint operator, then
a The eigenvalues of L are real
b The eigenvectors of L with different eigenvalues are orthogonal; that is,
their inner product is 0
Proof: (a) Suppose that f is an eigenvector of L with eigenvalue λ Then
hLf ; f i 5 hλf ; f i 5 λhf ; f iand
hf; Lf i 5 hf ; λf i 5 λhf ; f i:
SinceL is self-adjoint,
hLf ; f i 5 hf ; Lf i so λhf ; f i 5 λhf ; f iand since h f, fi 6¼ 0, we haveλ 5 λ , so λ is real
(b) Suppose
Lf 5 λ1f and Lg 5 λ2g with λ16¼λ2:Then
hLf ; gi 5 hλ1f; gi 5 λ1hf; giand
hLf ; gi 5 hf ; Lgi 5 h f ; λ2gi5 λ2hf; gi:
So
λ1hf; gi 5 λ2hf; giand thus h f; gi 5 0 because λ16¼λ2:
Trang 20ð2 π 0
sin ðnxÞ sin ðmxÞ dx5 0; if m 6¼ n
for m and n integers This is because sin (nx) and cos (mx) are eigenfunctions
of the self-adjoint operator d 2
dx 2with the inner product defined above with ferent eigenvalues
dif-We will use the technique of the example above to prove the
orthogonal-ity of functions, such as Bessel functions and Legendre polynomials, without
having to resort to tedious calculations
Fourier Coefficients
We now describe how to determine the representation of a given vector with
respect to a given basis That is, if f ^b1; ^b2; g is a basis for the vector space
V, and if ^vAV, we want to find scalars a1, a2, for which
^v 5 a1 ^b11 a2 ^b21 ?:
If the basis satisfies the characteristic below, then this is easy
Definition: If f ^b1; ^b2; g is a set of vectors from an inner product space for
which
h ^bi; ^bji5 0 if i 6¼ j;
then f ^b1; ^b2; g is called an orthogonal set If, in addition,
h ^bi; ^bii5 1 for all i;
then f ^b1; ^b2; g is called an orthonormal set A basis that is an orthogonal
(orthonormal) set is called an orthogonal (orthonormal) basis
Theorem: If f ^b1; ^b2; g is an orthogonal basis for the inner product space
h^v; ^bki5 ha1 ^b11 a2 ^b21 ? 1 a1^bk1 ?; ^bki
5 a1h ^b1; ^bki1 ? 1 akh ^bk; ^bki1 ? 5 akh ^bk; ^bki:Thus,
ak5 h^v; ^bki
h ^bk; ^bki5 h^v; ^bki
: ^bk:2:Note that if f ^b ; ^b ; g is an orthonormal basis, then a 5 h^v; ^b i
Trang 21Definition: The constants {a1, a2, } in the theorem above are called the
Fourier coefficients of ^v with respect to the basis f ^b1; ^b2; g
Fourier coefficients are important because they provide the best
approxi-mation to a vector by a subset of an orthogonal basis in the sense of the
fol-lowing theorem
Theorem: Suppose ^v is a vector in an inner product space V, and
ℬ 5 f ^b1; ^b2; g is an orthogonal basis for V Let {c1, c2, .} be the Fourier
coefficients of ^v with respect to ℬ Then
Proof: We assume the constants are real, and the basis is orthonormal to
simplify the notation We have
Trang 22with equality if and only if ci5 difor all i5 1, , n.
Note that from equation(2)we have Bessel’s inequality,
Xn
i 51
ci2# h^v; ^vi 5 :^v:2
:
Example: In this example, we demonstrate an application of eigenvalues
and eigenfunctions (eigenvectors) to solve a problem in mechanics
Suppose that we have a body of mass m1 attached to a spring of which
the spring constant is k1 See Figure 1-1-1 We assume that the surface is
frictionless If x1 is the displacement of the spring from equilibrium, then,
according to Hooke’s law, the spring creates a force ^F5 2x1k1 Then,
^F 5 m1
d2x1
dt2 5 2x1k1:Now consider the coupled system shown in Figure 1-1-2 We use the
convention that force is positive if it pushes a body to the right We suppose
that no spring is under no tension if the masses are at points a and b
Suppose that the masses are at points x1and x2
Trang 23Force on mass m1:
a Force due to spring 1: If x1 a, then spring 1 is stretched an amount
x12 a and pulls m1to the left If the spring constant of spring 1 is k1,
then the force on m1due to spring 1 is
F1;15 2k1ðx12 aÞ:
b Force due to spring 3: If x22 x1, b 2 a, then spring 3 is compressed an
amount (b2 a) 2 (x22 x1) If the spring constant of spring 3 is k3 then
spring 3 pushes the body m1to the left with force
F1;35 2k3½ðb2 aÞ 2 ðx22 x1Þ5 2k3½ðb2 x2Þ1 ðx12 aÞ:
Thus, the total force on m1is
F15 F1 ;11 F1 ;35 2 k1ðx12 aÞ 2 k3½ðb2 x2Þ1 ðx12 aÞ: ð3Þ
Force on mass m2:
a Force due to spring 2: If x2, b, then spring 2 is stretched an amount
b2 x2 and pulls m2to the right If the spring constant of spring 2 is k2,
then the force on m2due to spring 2 is
F2 ;25 k2ðb2 x2Þ:
b Force due to spring 3: If x22 x1, b 2 a, then spring 3 is compressed an
amount (b2 a) 2 (x22 x1) If the spring constant of spring 3 is k3, then
spring 3 pushes the body m2to the right with force
Trang 241CC:
Suppose thatλ is an eigenvalue for A, so that
Then we would have
t:The eigenvalues of A are those values of λ for which det(A 2 λI) 5 0
With the values we have, this would best be done with a CAS; however, if
we set the value of each mass to be m, and the value of each spring constant
to be k, then the matrix A is
22km
kmk
m
0BB
1CCand
detðA2λIÞ5
22k
mk
:
Trang 25Thus, the eigenvalues for A areλ 5 2k
m and5 23k
m.Before continuing, we note there is an alternate method to calculate the
equations of motion If V is the potential energy of the system, then
kmk
m
0BB
1CC
m, suppose that
22km
kmk
m
0BB
1CC
mz122k
mz25 23k
mz2;
Trang 26zðtÞ5 C1
11
ei
ffiffik m
p
t1 C2
121
ei
ffiffiffi3k m
p
t;where zðtÞ5 z1ðtÞ
z2ðtÞ
.The linear operators that we will use in the text will usually be differen-
tial operators A typical example of which is
L½y5 yvðxÞ 1 pðxÞy0ðxÞ1 qðxÞ:
We will often use the Principle of Superposition, which states that if
y1(x) and y2(x) are solutions to
L½y5 yvðxÞ 1 pðxÞy0ðxÞ1 qðxÞ 5 0;
1 Show that if ^v is an eigenvector for A with eigenvalue λ, then for any
scalar a, the vector a^v is an eigenvector of A with eigenvalue λ
2 For h , i an inner product, show that h f, g 1 hi 5 h f, gi 1 h f, hi
3 Show that the following is a linear operator,
L½y5 a0ðxÞyðxÞ1 a1ðxÞy0ðxÞ1 a2ðxÞyvðxÞ;
where a0(x), a1(x), and a2(x) are continuous functions
4 Show that the following function is an inner product,
hf; gi 5
ðb a
f ðxÞgðxÞdx;where f(x) and g(x) are continuous functions on [a, b]
5 Show that the following function is an inner product,
hf; giw5
ðb a
f ðxÞwðxÞgðxÞdx;where f(x), w(x), and g(x) are continuous functions on [a, b] and w(x) 0
What if w(x), 0?
Trang 276 In ℝ2orℝ3the angleθ between the vectors ^u and ^v is determined by
cosθ 5 ^uU^v: ^u::^v::
b Find the angle between ^u and ^v
c What is the distance between ^u and ^v traveling along the surface of
9 Show that if f^x1; ; ^xngis a basis for the vector space V, then every vector
in V can be written as a linear combination of^x1; ; ^xnin exactly one way
10 Show that if f^x1; ; ^xngis an orthogonal basis for the vector space V, and
What if f^x1; ; ^xngis an orthonormal basis?
11 Suppose that f^x1; ; ^xngis a basis for the vector space V and T : V- V
is a linear transformation for which Tð^xÞ 5 ^0 if and only if ^x 5 ^0
a Show that T is a one-to-one function
b Show that fTð^x Þ; ; Tð^x Þgis a basis for V
Trang 2812 For a function f(t), determine which of the following are linear
13 Suppose that f^x1; ^x2gis an orthonormal basis for V and T : V-V is a
lin-ear transformation for which fTð^x1Þ; Tð^x2Þgis also an orthonormal basis
for V Show that for any vector ^x; :Tð^xÞ: 5 :^x: (This is also true for
the case of any finite basis.)
14 Recall that if T : V-V is a linear transformation and V is an
n-dimensional vector space, then T can be represented as multiplication
by an n3 n matrix The matrix depends on the choice of the basis In
particular, if f^x1; ; ^xngis a basis for the vector space V and if
Tð^xiÞ5 a1i^x11 ? 1 ani^xn;then
Show that if T : V-V is a linear transformation and there is a basis of V,
f^x1; ; ^xng, consisting of eigenvectors of T, so that Tð^xiÞ5 λi^xi, then
the matrix of T with respect to this basis is the diagonal matrix
This is important because computations with diagonal matrices are
par-ticularly simple
15 A linear transformation U : V-V is called an orthogonal transformation
(or a unitary transformation if the field is the complex numbers rather
than the real numbers) if:U ^x: 5 :^x: for every ^x ε V
a Show that if U is an orthogonal transformation, then hU^x; U ^yi 5 h^x; ^yi
for every pair of vectors ^x; ^y ε V Hint: Use hUð^x 1 ^yÞ; Uð^x 1 ^yÞi 5
h^x 1 ^y; ^x 1 ^yi and expand both sides of the equation
Trang 29b What is the physical interpretation of part (a.) in the case that V is
c What is the dimension of T(V)? Find a basis for T(V)
17 The kernel of a linear transformation T : V-V is fxAVjTðxÞ 5 ^0g
a Show that the kernel of T is a vector space
b Find the kernel of T in the case where V is the n 1 1-dimensional
space of real polynomials in x of degree less than or equal to n and
T5 d
dx
1-2 CURVILINEAR COORDINATES
Many problems have a symmetry associated with them, and finding the
solu-tions to such problems, as well as interpreting the solution, can often be
simpli-fied if we work in a coordinate system that takes advantage of the symmetry In
this section we describe the methods of transforming some important functions
to other coordinate systems The most common coordinate systems besides
Cartesian coordinates are cylindrical and spherical coordinates, but the methods
we develop are applicable to other systems as well In our discussion, we
include some of the less common systems The less common systems will not
be used in later sections but are included to reinforce the techniques of the
transformations
In Figure 1-2-1a we give a diagram of how cylindrical coordinates are
defined, and in Figure 1-2-1b we do the same for spherical coordinates
We note that while the convention we use for spherical coordinates
is common, it is not universal Some sources reverse the roles of θ
Y y
θ
FIGURE 1-2-1a
Trang 30Our approach will be to describe the general case of converting from
Cartesian coordinates (x, y, z) to a system of coordinates (u1, u2, u3)
After making a statement that holds in the general case, to visualize that
statement, we demonstrate how the statement applies to cylindrical
coordinates
General Case: We start with Cartesian coordinates (x, y, z) and select
the group of variables u1, u2, u3so that each of x, y, z is expressible in terms
of u1, u2, u3; that is, we have
x5 xðu1; u2; u3Þ; y 5 yðu1; u2; u3Þ; z 5 zðu1; u2; u3Þ:
Cylindrical Case: The variables in cylindrical coordinates are r, θ, and z
The relations are
x5 r cos θ; y 5 r sin θ; z 5 z; 0 # r , N 0 # θ , 2π; 2N , z , N:
We write the vector ^r 5 x^i1 y^j1 z ^k in terms of u1, u2, u3; that is,
^r 5 xðu1; u2; u3Þ ^i1 yðu1; u2; u3Þ ^j1 zðu1; u2; u3Þ ^k:
In cylindrical coordinates, this is
^r 5 r cos θ ^i1 r sin θ ^j1 z ^k:
For some of our relations to be viable, the coordinates (u1, u2, u3) must
be orthogonal This means that the pairs of surfaces ui5 constant and
uj5 constant must meet at right angles In the case of cylindrical
coordi-nates, the surface r5 constant is shown in Figure 1-2-2a, the surface
θ 5 constant is shown in Figure 1-2-2b, and the surface z5 constant is
shown inFigure 1-2-2c Each pair does indeed meet at right angles It is also
possible to determine that the coordinates are orthogonal by analytical
meth-ods, as we now describe
In the general case, the vector @^r
@u 1 will be tangent to the u1curve, which
is the intersection of the u25 constant and u35 constant surfaces Similar
relations hold for @^r
P′
Z
Y y
FIGURE 1-2-1b
Trang 31FIGURE 1-2-2a
z
y x
u Surface of constant u
Trang 32We can show that a system of coordinates forms an orthogonal
coordi-nate system by showing that the vectors @^r
@u i are orthogonal; that is, by ing their inner product is zero In the cylindrical case,
5 hcos θ ^i1 sin θ ^j; 2 r sin θ ^i1 r cos θ ^ji
5 2r cos θ sin θ 1 r cos θ sin θ 5 0;
mutu-setting
^ei5 @u@^ri:@^r
h2 5 2r sin θ ^i1 r cos θ ^j
r 5 2sin θ ^i1 cos θ ^j;
^e35 ^ez5 @u@^r3
h 5 ^k:
Trang 33Back to the general case, we have
d^r 5 h1^e1du11 h2^e2du21 h3^e3du35 ^erdr1 r ^eθdθ 1 ^ezdz:
Volume Integrals
We now describe how to convert volume integrals to other coordinate
systems
General Case: Our aim is to determine an expression for an incremental
volume element dV in a general coordinate system The volume of the
paral-lelepiped formed by three non-coplanar vectors ^A; ^B, and ^C is j ^AUð ^B 3 ^CÞj
(See exercise 1.) For the Cartesian case, we compute an incremental volume
since^e1Uð^e23 ^e3Þ51 because f^e1; ^e2; ^e3gis an orthonormal system
Another way to do this computation is to use
Trang 34The determinant in equation (1) is called the Jacobian of x, y, z with
respect to u1, u2, u3, and is denoted @ðx;y;zÞ
We now compute dV for cylindrical coordinates We demonstrate two
methods First, we use
dV5 h1h2h3du1du2du3;where u15 r, u25 θ, u35 z so that du15 dr, du25 dθ, u35 dz We have pre-
viously found that h15 1, h25 r, h35 1, so
There are different ways that equation(2) is expressed in other sources
One other way is
Trang 35Example: Evaluate ðð
A
ex21y 2dxdy;where A is the circle x21 y2# 9, by changing to polar coordinates
In Cartesian coordinates,
ðð
A
ex21y 2dxdy5
The region A in polar coordinates is 0 # r # 3, 0 # θ # 2π So, in this
case, equation(2)says
ðð
A
ex21y 2dxdy
er2rdθdr
5 2π
ð3 0
We seek a coordinate system (u, v) so that the transformed region of
inte-gration will be a rectangle a# u # b, c # v # d The graph of the region A
v5xyy x
5 x2; so x 5
ffiffiffiuv
r:
We compute the Jacobian We have
ffiffiffiuv
r
; @y
@u5
12
ffiffiffivu
r
; @y
@v5
12
ffiffiffiuv
r:
Trang 362 ffiffiffiffiffiuv
2v
ffiffiffiuvs
12
ffiffiffivu
s12
ffiffiffiuvs
ð4
v 51
u22v2 1 u22
Trang 37The Gradient
Next, we determine the gradient of a function f, denoted rf In Cartesian
coordinates,
rf5@f@x^i1 @@yf ^j1 @@zf ^k:
To compute rf in the general case, we set rf5 A^e11 B^e21 C ^e3 and
write df in two different ways
First, df5 rf Ud ^r and using that d ^r 5 h1^e1du11 h2^e2du21 h3^e3du3 we
Trang 38The Laplacian
The final function that we consider in this section is the Laplacian, one of the
most important operators in mathematics and physics The Laplacian of the
function f, denoted Δf (some authors use r2f ) in Cartesian coordinates, is
for cylindrical coordinates with i5 2 We have
^e25 ^eθ5 @θ@ ðr cosθ ^i1 r sin θ ^j1 z ^kÞ 5 2r sin θ ^i1 r cos θ ^j
Trang 39since h15 h35 1 Now h25 r and u15 r so @h 2
5 2 ^e15 2 ðr cos θ^i1 r sin θ^jÞ;
so the formula holds in this case
In fact, it will be simpler to derive expressions for r2f and r3 F, after
we have studied the Divergence Theorem and Stokes’ Theorem in Section 2-3
For now, we simply state the results:
Trang 40Spherical Coordinates
One of the most commonly used coordinate systems is spherical coordinates
(In the previous examples, we used cylindrical coordinates because the
com-putations are simpler.)
The transformations are
Other Curvilinear Systems
The most common curvilinear systems are cylindrical and spherical
coordi-nates, but there are several lesser-known systems, some of which we
dis-cuss here We will not use these examples in the sequel but present these
results to develop familiarity and facility in the computations The list we
give is not complete Other examples are given in Morse and Feshbach
(1953) and Arfken (1970)
The first example is elliptic cylindrical coordinates We give the
transfor-mations and
G Demonstrate that the system is orthogonal
G Compute the scaling factors
G Determine the orthonormal basis