Series SN Mathematical HVDC
Series SN Mathematical HVDC
іс
Series Sn I
Mathematical
hvdcs.
NUNC COCNOSCO EX PARTE
TRENT UNIVERSITY
LIBRARY
Digitized by the Internet Archive
in 2019 with funding from
Kahle/Austin Foundation
https://archive.org/details/partialdifferent0001tikh
PARTIAL DIFFERENTIAL EQUATIONS
OF MATHEMATICAL PHYSICS
Volume I
HOLDEN-DAY SERIES IN MATHEMATICAL PHYSICS
Volume I
Translated by S. Radding
HOLDEN-DAY, INC.
San Francisco, London, Amsterdam
1964
© Copyright 1964 by Holden-Day, Inc., 728 Montgomery Street,
San Francisco, California. All rights reserved. No part of this
book may be reproduced in any form, by mimeograph or any
other means, without permission in writing from the publisher.
ONULP
EDITOR’S PREFACE
This text reflects the authors’ unique approach to the study of the basic
types of partial differential equations of mathematical physics. The system¬
atic presentation of the material offers the reader a natural entrde to the
subject. Each of the basic types of equations which are to be studied is
motivated by its physical origins. The derivation of an equation from the
physics to its final mathematical structure is very instructive to the student.
The authors have gone to great length to make clear the meaning of a
solution to an initial value or boundary-value problem. Various methods of
solving such problems are treated in great detail, as are the questions of
existence and uniqueness of solutions. Thus, the student gains an apprecia¬
tion of the theoretical foundations of the subject and simultaneously acquires
the manipulative skills for solving such problems.
The exercises which accompany each chapter have been selected to test
the student’s ability both to formulate the correct mathematical statement
of the problem and to apply the appropriate method for its solution. The
applications treated by the authors are non-trivial and are completely worked
out in detail.
The present volume covers the two dimensional class of partial differential
equations of mathematical physics and is well suited as a basic text for both
the undergraduate and graduate level at the university. The second volume
will cover the three dimensional counterparts of the present volume and
contain an additional chapter on the special functions which arise in mathe¬
matical physics.
In view of the dearth of texts having the scope and depth of the present
one, Holden-Day felt the need for a careful and articulate translation of
Tychonov and Samarski’s valuable contribution to the literature in the field.
The translator and editor exercised great care in making the English edition
read smoothly and correctly. Several misprints and other flaws have been
corrected, and we are pleased to make this excellent text available to the
English-reading scientific community.
Julius J. Brandstatter
Stanford Research Institute
FROM THE PREFACE TO THE RUSSIAN EDITION
A. Tychonov
A. Samarski
PREFACE TO THE GERMAN EDITION
APPENDIX
Tables of error integrals and some cylindrical functions 363
Literature references of the editor 371
1
The equation has an analogous form for more than two independent variables.
A partial differential equation of the second order is called linear with
respect to the highest derivative, if it has the form
where the coefficients an, «12, and a22 are functions of x and y.
If the coefficients are not only functions of x and y, but also F is a function
of x, у, M, Mx, and My у then it is called a quasilinear differential equation.
The equation is called linear if it is linear in the higher derivatives
Mxx, Mxy, Myy, also in the function u(x, y), and in the first derivatives ux, му\
where an, a12, a22, blt b2, c, and / are functions which depend only on x and
y. If the coefficients are independent of x and y, then Eq. (1-1.2) is a linear
differential equation with constant coefficients. Equation (1-1.2) is called
homogeneous if f(x, y) = 0.
With the aid of a unique inverse transformation
Ux UtJ^x “Ь Uxi7Jx
Uy - U£$y у
‘2
Uyy - U££^ у + 2u&t£y7)y ~b UrjXqTJ у ~b U^yy + UxqTjyy .
If these expressions are inserted into (1-1.1) an equation results of the form
and F is independent of the partial derivatives of the second order of w(£, >?)
with respect to f and rj. If the initial equation is linear, that is,
2 If the transformation of the variables is linear, then F = F, since the second deri¬
vatives of f and 7] vanish in formula (1-1.3), and in the expression for F none of the
transformations of the second derivatives appears in the preceding sums.
1-1. DIFFERENTIAL EQUATIONS WITH TWO INDEPENDENT VARIABLES 3
is valid for all x, у of the region in which z = <p(x, y) is defined and <py(x, у) Ф 0.
The relation <p(x,y) = C is a general integral of Eq. (1-1.6) if the function у
defined implicitly by <p(x, у) = C satisfies Eq. (1-1.6). If, namely, y=f(x,C)
is this function, then it satisfies
so that у = f(x, C) satisfies Eq. (1-1.6). The expression in the brackets vanishes,
not only for у = f(x, C), but for all values of x, y.
Proof of the second lemma. Let <p(x,y) = C be a general integral of Eq. (1-1.6).
We can show that for each point (x, y)
is valid. Let (x0,y0) be any given point. If it can be shown that Eq. (1-1.7)
is satisfied at this point, it follows that Eq. (1-1.7) is valid at all points, since
(х0,Уо) is arbitrarily chosen. The function <p(x,y) then represents a solution
of Eq. (1-1.7). We now construct through (x0, y0) an integral curve of Eq.
(1-1.6) in which we set <p(x0,yo) = C0 and consider the curve y=f(x,C0).
Obviously, y0 = f(x0,Co). For all points of this curve the following equation
is valid,
t/x CL л
The sign of the expression under the root determines the type of the dif¬
ferential equation
anuxx + 2al2uxy + a22uyy + F = 0 . (1-1.1)
3 This relationship between Eqs. (1-1.5) and (1-1.6) is the equivalent of the well-known
relation between a linear partial differential equation of the first order and a system of
ordinary differential equations. This can be shown if the left side of Eq. (1-1.5) is
represented as the product of two linear differential expressions.
[See V. I. Smirnov: Course in Higher Mathematics, Part II, 2d ed., Berlin, 1958, p. 62,
and V. V. Stepanov: Textbook of Differential Equations, Berlin, 1956, p. 328 (Translated
from Russian).
4 This terminology is taken from the theory of curves of the second order.
1-1. DIFFERENTIAL EQUATIONS WITH TWO INDEPENDENT VARIABLES 5
and reduce Eq. (1-1.4) after division by the coefficient of иcv to the form
F
U(i! = Ф(£, у], и, U(, Ur)) with Ф =— ЙГ2 Ф 0
2 a.
This is the so-called canonical form for an equation of the hyperbolic type.5
Frequently, a second canonical form is used. If we set
£ = a + ft f] = a — /3
i.e.,
a =
£ + у n _ £—у
2 ‘ 2
where a and /3 are the new variables, then
uaa — ирв = Фі Фі = 4Ф
5 The introduction of the new variables f and rj through the functions <p and ф is only
possible when these functions are independent of each other. Thus it is sufficient that
the corresponding functional determinant obtained from these functions be distinct from
zero. This is the case here, since if
<Px фх
V>y фу
at any point M were zero, then for this point the columns of the determinant would be
proportional to each other; hence
Vx _ фх
<Py фу
but since
this is impossible (without loss of generality we assume an Ф 0). Thus, the independence
of functions <p and ф is demonstrated.
6 PARTIAL DIFFERENTIAL EQUATIONS OF THE SECOND ORDER
results for an equation of the parabolic type. If, in particular, щ does not
appear in this equation, then it is an ordinary differential equation with £ as
a parameter.
3. For an equation of the elliptic type, fli! - йцй22 < 0, and the right
sides of Eqs. (1-1.9) and (1-1.10) are complex conjugates of each other. Thus,
if
<f(x, у) = C
is a complex integral of the differential Eq. (1-1.9), then
<p*(x, у) = C ,
where <p* is a complex function conjugate to <p, a general integral of Eq.
(1-1.10), and a complex conjugate to (1-1.9). We introduce now complex
variables by setting
? = 4>(X, У) v = 9*(x, У) .
In this way an equation of the elliptic type, as in the case of the hyperbolic
type, is converted to another form.
In order to avoid calculations with complex variables we introduce new
real variables a and p, through
1 * *
a
9 + 9 9-9
(8 =
2 2i
such that
£ = a + ij3 rj — a — i[3 .
Thus we obtain
«11 fx T + «22 ^у
= (fluff! + 2al2axay + a22ay) — (anfi + 2anpxfiy + a22fil)
T 2і(&ц (Xx fix “b «12{oLx fly -}“ OCyflx) T «22 CCу fiy) — 0 ,
6 Such a transformation is valid only if the coefficients of Eq. (1-1.1) are analytic
functions. Namely, if aL — ana22 < 0, then Eqs. (1-1.9) and (1-1.10) are complex; con¬
sequently, the function у takes on complex values. We can only speak of the solutions
of such equations when the coefficients of aik(x, у) are defined for complex values of y.
For the conversion of the differential equation of the elliptic type to canonical form we
shall limit ourselves to equations with analytic coefficients.
1-2. DIFFERENTIAL EQUATIONS WITH MANY INDEPENDENT VARIABLES 7
We shall consider now the linear differential equation with real coefficients.
?4 = Ы*1, *2 , • * •, Xn) k = 1, • • •, n .
Then
2 2 akiUtn, + 2 bkU(k + cu + f — 0
4= 1 i = l 4=1
with
n «П
aki = 2 2 ацсцкап
t=l i =l
_ n n n
bk 2 biajk T 2 2 aij($k)xiXi »
»=1 i=l J = 1 J
2 2 аЬуіУі, (1-2.2)
<=іj=i
У>= 2 aikrjk
4=1
8 PARTIAL DIFFERENTIAL EQUATIONS OF THE SECOND ORDER
where
flkl — ^ dijOCikOCjl
i= 1 j=1
The coefficients of the principal parts of the equation transform like the
coefficients of a quadratic form under a linear transformation. As is well
known, with the help of a suitable linear transformation the coefficient matrix
(a°j) of a quadratic form can be transformed into a diagonal matrix in which
This is called the transformation of the quadratic form to the normal form.
According to a theorem of Sylvester, the so-called inertia rule for quadratic
forms, the number of coefficients of a°u distinct from zero is equal to the
rank of the coefficient matrix, and the number of negative coefficients is
invariant.
We call Eq. (1-2.1) at the point M0 of the elliptic type if all n coefficients
of a°i are different from zero and have equal signs; of the hyperbolic (or of
normal-hyperbolic) type, if, likewise, all а°ц Ф 0, n — 1 coefficients of а°ц have
the same sign, and one coefficient is different from the other in the signs;
of the ultrahyperbolic type if m coefficients of a°u are equal and n — m have
opposite signs (m > 1, n — m > 1); and of the parabolic type if at least one
of the coefficients a°u vanishes.
We choose the new independent variables so that at the point M0
_ д£к о
ОСік ^ ОС-ік у
ot\
where aik is the coefficient of the transformation which converts the quadratic
form (1-2.2) to the canonical form, in which we can set = 'Z.a'iV xt. Then
Eq. (1-2.1) at the point M0 can be transformed to one of the following
canonical forms:
m n
ultrahyperbolic type: I Ux{Xi = I uHH + Ф
i=l i = m-\-l
n—m
parabolic type: I (± uHH) + 0 = 0 m > 0
i= 1
lyi + a\2
V1 & 11 #2 «12 + Ѵа\г - йп О-гг
У У = C2
и = v ,
10 PARTIAL DIFFERENTIAL EQUATIONS OF THE SECOND ORDER
Mf = ex^'(vz + Xv)
Ur, = /f+M(f„ + yv)
Mff - eXf+M(fff + 2,XV( + X2v)
Wfl? = eXf+M(ff, + Xvr, + yv^ + Xyv)
tirjrj = e*+M(vr,r, + 2 yv r, + yv)
fee + vVt) + (b, + 2X) ve + (b2 + 2у) Vr, + (i2 + у + b,X + b2y + c) v + f, = 0 .
If in this equation the parameters X and у are so chosen that there are two
coefficients, say in which both of the first derivatives are made to vanish,
that is, X = —(bj2) and у — —(b2l2), then we obtain
fee + Vvv + yv + /i = 0 ,
We have already noted (1-2) that a differential equation with constant co¬
efficients in the case of several independent variables,
I I ац инч + I Ь{ uXi + cu + / = 0 ,
t—1 j = 1 i=l
Problems
uxx “b уМуу — 0
Uxx Uyy - 0
is usually called the differential equation of the vibrating string. In this
and the following chapters we shall restrict ourselves to the treatment of
linear differential equations.
FIG. 1.
Thus no elongation of a single segment of the string occurs within the scope
of our required limits of accuracy, whereas according to Hooke’s law it
follows that at each point, the tension T is independent of x, i.e.,
T — T0 = const.
For the projection of the tension on the x and и axes, indicated by Tx and
Tu, we find
But since Xi and x2 were selected arbitrarily, it follows that the tension is
not dependent on x; i.e., for all x and t values,
T = To. (2-1.2)
( w«(f, t)p(£)d£ ,
J
where p denotes the linear density of the string. We now set the change
of momentum during the time Jt = t2 — ,
14 HYPERBOLIC DIFFERENTIAL EQUATIONS
equal to the impulse of the acting force which arises from the tension T0uz
at the points Xi and x2, as well as from the external force. The latter we
assume as continuously distributed with the density p (the load) and denote it
(per unit of length) by f(x, t). Thus we obtain the equation of a transverse
vibrating string in integral form
where
ux(x, t)
u(xо + 0, t) = u(x0 — 0, t)
(2-1.8)
ux{xо + 0, t) — ux(xо 0 ,t) =~±fo(t)
1 0
must be fulfilled. The first expresses the continuity of the string, while the
second determines the magnitude of the jump of the derivative at x0 in terms
of f0(t) and the tension T0.
2 The geometric variable x chosen here is called the Lagrange coordinate. In the
Lagrange coordinates, each physical point of a rod in the course of an entire process is
characterized by one and the same geometric coordinate x. A physical point, which at
the initial time (in the equilibrium state) is at the point x, is found after an arbitrary
time t at the point with the coordinate X = x+u(x, t). If we choose any geometric point
A with the coordinate X, then at different times different physical points (with different
Lagrange coordinates x) would be found at this point. Frequently we also use the
Eulerian variables X, t, where X is the geometric coordinate. If U(X, t) denotes the
displacement of the point with the Eulerian coordinate X, then the Lagrange coordinate
is x = X — ЩХ, t). An example of the use of Eulerian coordinates is found in § 2-6.
16 HYPERBOLIC DIFFERENTIAL EQUATIONS
where f(x,t) is the density of the external force per unit of length.
If now the second derivative of the function u(x, t) exists and is con¬
tinuous, then by use of the mean-value theorem and after taking the limits*
as Ax = хг — Xi —»0 and At = t2 — tt-> 0 we obtain
where
3 In the following we shall waive the discussion of such details for a transverse
vibrating string.
4 The requirement that the vibrations be sufficiently small depends in the present
case only on the limits of applicability of Hooke’s law. Generally, T = k{x, ux)ux; we
then arrive at the quasilinear equation [k(x, ux)ux]x = putt — fix, t).
2-1. SIMPLE PROBLEMS AND BOUNDARY-VALUE PROBLEMS 17
ymv2 = -^-p{x)dx(ut)2,
The potential energy of a transverse vibrating string which has the form
u(x, t0) = u0(x) at time t = t0 is equal to the work done in transforming the
string from the equilibrium state to the state u0(x). The profile of the string
at time t is given by the function u{x, t). Thus
whereas the time dt is related by means of ut(x, t)dt. The work done is
equal to
J_ d_
T0{uxfdx + T0uxut I
2 dt
By integration over t from 0 to t we obtain
The meaning of the latter terms of this equation is easy to see; indeed,
ToUx U=0 is the magnitude of the tension at the end point x = 0; ut(0, t) is
the displacement of the end point while the integral
j ToUxUt\x^dt (2-1.15)
T,[u'o(x)Ux, (2-1.16)
i.e., it equals the potential energy of the string at time t = t0 but with op¬
posite sign. Therefore we have
as the total energy of the string. We similarly obtain the potential energy
of a longitudinal vibrating rod. Moreover, we arrive at the potential energy
of the rod starting from the formula
U= k(ux)2dx.
which is the amount of electricity necessary for the charging of the element
Ax plus the amount which is lost due to insufficient insulation, where C is
the capacity and G is the loss coefficient (both are expressed per unit of
length). We assume that the magnitude of the loss is proportional to the
voltage at those points of the conductor under consideration.
From (2-1.18), (2-1.19), and (2-1.20) we then obtain the so-called system of
telegraphic equations5
5 These equations within the structure of the theory of electromagnetic fields have
only approximate validity, since the electromagnetic vibrations in the material-filled
space are neglected.
2-1. SIMPLE PROBLEMS AND BOUNDARY-VALUE PROBLEMS 19
ix Cvt Gv — 0
(2-1.21)
vx 4" Lit T Ri = 0 .
In order to obtain only a single equation defining the function i, we dif¬
ferentiate the first of the two equations (2-1.21) with respect to x and the
second with respect to t after we have multiplied these by C. By subtrac¬
tion of the equations thus obtained we then find, under the assumption that
the coefficients are constant,
I J
X = Xy {T(x2, y, t) — T(xi, y, t)}dy = 0
yi
FIG. 3.
or
T* =[ {T(x,y2,t) — T(x,yl,t)}dx = 0 .
J
Hence according to the mean-value theorem, since the surface ABCD is arbi¬
trarily chosen, it follows that
i.e., the tension T does not vary with x and у and therefore can only be a
function of t alone.
The surface area of a membrane element at time t in the sense of our
approximation is equal to
If we now apply the mean-value theorem and take into consideration that
both Si and the time interval (G, t2) are arbitrary, then we see that the ex¬
pression in the brackets must vanish identically. In this manner we arrive
at the differential equation of a vibrating membrane:
where Fix, y, t) is the force density per unit of mass of the membrane.
In order to describe the motion of а fluid one uses three functions vdx,y, z, t),
v2ix,y,z,t), and v3ix, у, z,t), which are the components of the velocity vector v
at the point ix,v,z) at time t. Further quantities for the characterization of
a fluid motion are the density p(x,y,z,t), the pressure pix,y,z,t), and the
density of the external force given in units of mass Fix,y,z,t), in case such
forces are present.
We consider a fixed portion of a fluid occupying a region of space T, and
calculate the force acting on it. Moreover we shall neglect the friction forces
6 See V. I. Smirnov, Textbook of Higher Mathematics, 2d ed., Part II, Berlin, 1958,
p. 175; in the literature it is also called Gauss’ integral theorem.
22 HYPERBOLIC DIFFERENTIAL EQUATIONS
caused by the viscosity, i.e., we consider an ideal fluid. For the resultant
of the pressure forces we obtain the following expression in the form of a
surface integral
-JJ/>ndS, (2-1.33)
where 5 is the surface of the region T and n is the unit vector in the direc¬
tion of the outward directed normal. The formula of Green then yields
For the calculation of the acceleration of any fluid element, naturally the
displacement of the points themselves are to be considered. If x = x(t), у —
y(t), and 2 = z(t) are the paths of the points of this element, then the derivative
of the velocity with respect to time is
dv dv dv dv dv dv dv dv dv dv
—— = —- + -т-x + ——y + z= Vi + ——v2 + ——va = —- + (vp)v ,
dt dt dx dy dz dt dx dy dz dt
where
This derivative with respect to time, which takes into account the motion of
a single element of the medium (the substance), is called the substantive
derivative. The equation of motion of a fluid which describes the connection
between the acceleration of the element and the forces acting upon it is
given by
SlS4-*4lSgrad*Hi>№- (2-1'3S
The last integral represents the resultant of the external forces which act on
the region of space T. But since T was chosen arbitrarily we obtain the
equation of motion of an ideal fluid in Eulerian form:
To derive the needed continuity equation, we shall assume that at all points
of the streaming fluid region T no additional fluid is introduced, i.e., we limit
ourselves to the study of source-free streams, in which we regard a sink as
a negative source. Then the change in the amount of fluid contained in T
per unit of time is equal to the flux through the boundary S of the region
d_
^ \^pdz=— J \^pvndS. (2-1.37)
dt
The transformation of the surface integral to a volume integral yields
+ div pv \dv = 0 .
2-1. SIMPLE PROBLEMS AND BOUNDARY-VALUE PROBLEMS 23
Since this relation is valid for an arbitrarily small region, there follows the
continuity equation
P = Ap)-
Thus we obtain a system of five equations for the five unknown functions
vx, vy, v,, p, and p. If the equation of state also includes the temperature,
then a heat-conduction equation must be added. The system of equations
can be taken as the equation of state (p0 designates the initial density and p0
is the initial pressure, cp and cv are the specific heats at a fixed pressure and
a fixed volume); and (3) the vibrations of the gas are sufficiently small so
that the higher powers of the velocity, the gradients of the velocity, and the
gradients of the density can be neglected.
For the condensation of the gas we designate the relative change in density
by the quantity
P
s(x,y, z, t) (2-1.40)
po
p Pod + s) (2-1.41)
With the above assumptions the hydrodynamic equations take the form
24 HYPERBOLIC DIFFERENTIAL EQUATIONS
Vt——- grad p
P0
pt + po div I? = 0 (2-1.42)
where the dots denote the terms which are of second and higher order of
smallness. With the notation a2 = rPolpo we rewrite the system (2-1.42) in
the form
st + div v = 0 .
If we now apply the divergence operator to the first equation of (2-1.42') and
interchange the order of differentiation, we obtain
2 d2 d2
V
dy2 dz2
is the Laplace operator. Then the second expression of (2-1.42') can be used
to yield the vibration equation
(2-1.43)
or
a (sxx T Syy T Szz) — s*t •
From this and from (2-1.40) we obtain the density equation
Vt — — a2 grad s
follows
v — — grad U
(2-1.47)
st + div v = 0 ,
Equations for the pressure p and the velocity v similar to Eq. (2-1.48) can
be derived. These are usually called the equations of acoustics.
For the solution of two- and one-dimensional problems the Laplace opera¬
tor is replaced in Eq. (2-1.48) by дг/дх* + д2/ду2 or дгІдх2.
For the vibrations of a gas in a bounded region of space, specific bound¬
ary conditions must be fulfilled; that is, certain restrictions are imposed which
are to be applied directly on the sought functions at the boundary of the
gas-filled region of space. The simplest example is for the case of a gas
moving in a vessel whose walls are fixed. Since the stream velocity must
always be directed tangentially on these walls, the normal component of the
velocity must be equal to zero. This leads to the condition
or (2-1.49)
дп л-
The constant
On the other hand, there are initial conditions to be satisfied, i.e. the dis¬
placement and velocity of the string at the initial time t0, say
If the end points of the string move in a prescribed way, then the boundary
conditions assume another form:
u{0, t) = frit) ,
(2-1.50')
U(l, t) = fjt2(t) ,
where (i^t) and fi2(t) are prescribed functions of the time /. The statement
of problems for longitudinal vibrating rods or springs reads analogously.
Boundary conditions occur also in other forms. We shall consider, for
example, a longitudinal vibrating spring which is fastened at one end (sus¬
pended point) while the other is free to move. The motion of the free end
is not known and is thus the function to be determined. At the point of
suspension x = 0 there can be no displacement,
u(0,t) = 0;
2-1. SIMPLE PROBLEMS AND BOUNDARY-VALUE PROBLEMS 27
and equals zero (no external force), so that the mathematical formulation of the
conditions for the free end takes the form
ux(l, t) = 0 .
If the fixed end x = 0 moves according to a definite law //(/), whereas at x= l
the prescribed force v(t) acts, then we write
kux(l, t) = —au(l, t)
or
and amount to the requirement that и and ux are continuous. In the boundary
conditions the derivative with respect to t can also occur. If the end of a spring
undergoes a resistance from the outside, which is proportional to its velocity
(perhaps by a disc fastened to the end of the spring, and for which the plane
of the disc is perpendicular to the spring axis), the boundary conditions read:
must be fulfilled at x = l.
In the following discussion we shall limit ourselves to the consideration
of the three simplest types of boundary conditions, directing attention to an
example of boundary conditions of the first type, and indicate only incidentally
the peculiarities which occur in connection with the second and third types.
To formulate the first boundary-value problem for Eq. (2-1.52), let us seek
a function u(x, t) which satisfies the equation
и (0, t) = g^t)
и (/, t) = fJ2(t)
(2-1.60)
и (x, 0) = (p(x)
ut(x,0) - </>(x)
и (0, t) = fi(t) , t ^0
и ix,0) = <pix) , о ^ x <QO (2-1.62)
■Utix, 0) = фіх) .
The character of the process for a time interval which is sufficiently
long from the initial time t = 0, is completely defined by the boundary values
since the influence of the initial conditions, because of the friction which
occurs in every real system, vanishes with increasing f.7 Such cases occur
mostly when the system under consideration is acted on by a periodic bound¬
ary influence which persists for an indefinitely long time. The formulation
of such problems without initial conditions is as follows:
Find a solution of the considered equation for 0 й x^l and t > — oo with
the boundary conditions
This is analogous to the problem without initial conditions for the semi¬
infinite line.
Besides the fundamental boundary-value problems, we shall also consider
in what follows the limiting cases:
1. Problems for an infinite region when one or both boundaries lie at infinity.
2. Problems without initial conditions when the solutions considered are
defined in the course of an infinitely large time interval.
7 The vibration equation, taking into consideration the influence of friction which is
proportional to the velocity, has the form utt = a2uxx — aut, where a > 0. For details
of the above problem without initial conditions for a = 0, see § 3-7.
30 HYPERBOLIC DIFFERENTIAL EQUATIONS
Ui(0, t) = fA(t)
Ui(l, t) = /4(0
Ui(x, 0) = <p'(x) (2-1.65)
^(x,0 ) = ф\х).
at
Obviously, the solutions can be superposed in such a way that the function
i /4(0 , k = 1, 2
i= 1
Фт(х) = i ф\х).
i—1
We see that this superposition principle is valid not only for the aforemen¬
tioned problem, but also for every linear equation with linear auxiliary con¬
ditions. We shall use this property repeatedly in the following.
The solution of the general boundary-value problem
where uy, u2, u3, ut are solutions of the particular boundary-value problems
2-1. SIMPLE PROBLEMS AND BOUNDARY-VALUE PROBLEMS 31
2 о u>%
■32
= a * = 1,2,3 ,
dt2 dx2
(2-1.71)
d2ut
= a2^- + f{x, t) , 0 < x < l, t > 0 ,
dt2 ox
possesses only one solution which satisfies the initial and boundary condi¬
tions
Here it is assumed that the function u{x, t) and its first and second deri¬
vatives are continuous in the interval 0 x ^ l for t ^ 0, and p(x) > 0, k{x) > 0
are continuous functions.
Suppose there are two solutions
a2 v_ _ д Л dv \
_
(2-1.77)
dt2 dx \ dx )
v {x, 0) = 0
vt(x, 0) = 0
(2-1.78)
v (0, 0 = 0
v(l, 0 = 0.
Then we shall prove that the function v(x, t) defined by the conditions (2-1.78)
is identically zero.
To this end, we consider the function
dE(t)
(kvxvxt + pvtvtt)dx.
dt
An integration by parts of the first term of the sum on the right side gives
dE(t)
vt{kvx)x vt[pvtt — (kvx)x\dx = 0 ;
dt
this means E(t) = const. By taking into consideration the initial conditions
we find further that
because
v(x, 0) = C0 = 0 ,
and therefore we prove that
If, therefore, ity{x, t) and u2(x, t) are two functions which satisfy all the con¬
ditions of our theorem, then Uy(x, t) = u2(x, t).
For the second boundary-valife problem the function v = иУ — иг satisfies
the boundary conditions
so that the brackets in (2-1.80) vanish again. The rest of the proof remains
unchanged.
With regard to the third boundary-value problem, the proof of the unique¬
ness theorem requires several changes. If we again consider two solutions
иi and и2, then for their difference, v(x, t) = uy — u2, the differential Eq.
(2-1.77) results, while the boundary conditions read:
vx(0, t) - hM0, 0 = 0, hy ^ 0
(2-1.85)
Vx{l, 0 + h2vd, t) = 0 , ht ^ 0 .
34 HYPERBOLIC DIFFERENTIAL EQUATIONS
from which, with respect to the equation and the initial conditions, it follows
that
Thus the uniqueness theorem is proved also for the third boundary-value
problem.
The method of proof used here, which was based on the use of the expres¬
sion for the total energy, is often used for the proof of uniqueness theorems
in different branches of mathematical physics (for example, in electrodynamics,
elasticity theory, and hydrodynamics).
The proof of existence for these and other boundary-value problems will
be discussed later when we consider the corresponding questions.
Problems
1. Prove that the differential equation for small torsional vibrations of a rod
has the form
Ѳн = а2Ѳхх,
where Ѳ is the angle of torsion of the cross section of the rod at the abscissa
x, G is the modulus of torsion, / is the polar moment of inertia of the cross
section, and k is the moment of inertia per unit of length of the rod. Give
a physical interpretation of the first, second, and third boundary conditions
for this problem.
2. Let an absolutely flexible homogeneous cable be fastened at one end.
Under the influence of its weight, it is aligned then with the vertical axis.
Find the differential equation for small vibrations of the cable.
Solution'.
d2u 2 3 V.du~\ ... 2
IF = a-s7Vl ~ х)~эІ \ w‘tha=!''
2-1. SIMPLE PROBLEMS AND BOUNDARY-VALUE PROBLEMS 35
where u(x, t) is the displacement of the points, l is the length of the cable,
and g is the acceleration due to gravity.
3. A heavy homogeneous cable of length /, which is fastened at its upper
end (x = 0) of a vertical axis, rotates about this axis with a uniform angular
velocity w. Derive the differential equation for small vibrations in the neigh¬
borhood of the vertical state of equilibrium.
Solution:
d2u 2 d V. du 1 . 2 ... 2
= a — (/ - x,)— + ш u with a = g.
uH = a2uxx + but + cu .
u(x, 0) = <p(x),
(2-2.2)
ut(x, 0) = ф{х).
First we transform this equation to the canonical form in which the mixed
derivative is obtained. The characteristic equation
dx2 — a2dt2 = 0
we see that the equation of a vibrating string can be transformed to the form
Me, = 0 . (2-2.3)
2-2. WAVE-PROPAGATION METHOD 37
v) = f*(v)
is valid, where f*(rj) is a function of rj alone. By integration of this equation
with respect to rj for fixed £ we get
where depends only on £ and A only on rj. Conversely, any arbitrary dif¬
ferentiable functions fi and /2 which define the function w(£, rj) through (2-2.4)
represent solutions of Eq. (2-2.3). Since every solution of Eq. (2-2.3) for a
suitable choice of /, and /2 can be represented in the form (2-2.4), then (2-2.4)
yields the general solution of this equation. Consequently,
fiix) — Ш = — f </>(a)da + C ,
a J*o
AM + AM = 4>M
AM - AM = —{ <p{a)da + C
we then find
AM = -^vM
(2-2.8)
AM = y^M
j
s rx+at rx—at
8 In formula (2-2.5), /i and /2 are not uniquely defined. That is, if we subtract
from f\ any fixed number ci and add this to /2, и remains unchanged. In 2-2.8, how¬
ever, the constant C is not determined by <p and ф. It can be omitted or replaced by
another without changing the value of w. When and /2 are added, the sum differs
by ± Cl2.
38 HYPERBOLIC DIFFERENTIAL EQUATIONS
or
I (•!+at
(p(x + at) + <p(x — at)
u{x, t) + —\ <f>(a)da . (2-2.9)
2 j x—at
Formula (2-2.9), the so-called D’Alembert formula, was derived under the
assumption that a solution of the given problem exists. This formula proves
the uniqueness of the solution. If there were to exist another solution of
the differential Eq. (2-2.1) with initial conditions (2-2.2) then it would have
to be a solution of the form (2-2.9) and hence would coincide with the first
solution. It can be shown that the function u(x,t), defined by (2-2.9) under the
assumption that <p be twice differentiable and that ф is only once differenti¬
able, satisfies both Eq. (2-2.1) and also the initial conditions (2-2.2), so that
the D’Alembert method besides proving uniqueness also proves the existence
of the solution of the given problem.
2. Physical interpretation
The function u(x, t) which appears as a solution of the wave equation
with initial conditions, is the sum of two functions
u2{x, t) — ф(а)(1а.
2a J x-at
The first sum ифх, t) represents the path of propagation of the initial dis¬
placement without the initial velocity, ф(х) — 0; the second sum u2{x, t) con¬
tains the initial velocity (the initial impulse) for vanishing initial displace¬
ment. The function u{x, t) can be interpreted geometrically as a surface in
u, x, t space (Fig. 4a). The intersection of this surface with the plane
FIG. 4.
2-2. WAVE-PROPAGATION METHOD 39
t = t0 is analytically given by и = u(x, t0) and yields the profile of the string
at time t0. On the other hand, the intersection of the surface u(x,t) with
the plane x = x0 gives и = u(x0, t), which is the path of motion of the points
Xo-
The function u(x, t) given by и — f(x — at) is described in physics as a
propagating wave.* * * 9 The displaced profile defined by this function at different
times t can be easily illustrated in the following manner: We shall assume
that an observer moves parallel to the x axis with the velocity a (Figure 4b).
If the observer then is found at the initial time t = 0 at a position x = 0,
then up to time t he has moved along the path toward the right. If we
now introduce a new coordinate system through
x — at = const.
The surface и = f(x — at) is also a cylindrical surface whose generators are
parallel to the straight lines x = at. Thus the form of the cylindrical surface
is determined by the profile of the initial displacement.
In the interval (xltx%) let the function f(x) now be different from zero
and outside of this interval equal to zero. The straight lines x — at = x2
and x — at — Xi represent the forward surface and the rear surface of the
propagating wave f(x — at). These lines divide the x, t plane into three
regions I, II, and III.
The regions I and III consist of the
points {x, t) which correspond at the time
t considered to the point x of the string
that lies behind the forward propagating
wave. On the other hand, the points (x, t)
of region II correspond to those points x
through which, at time t, the forward pro¬
pagating wave travels (Figure 5).
X
It is obvious that f(x + at) represents
a wave propagating toward the left at a
FIG. 5. velocity a. For this wave a similar inter-
pretation can be given. The function ux(x, t) which describes the propagation
of the initial displacement <p(x) with vanishing initial velocity, ф(х) = 0, is
given by formula (2-2.10) as the sum of two waves propagating to the right
and left at a velocity a. Thus the initial form of both waves is characterized
by the function <p(x)/2, which is equal to one half of the original displacement.
As a first simple example we shall consider the propagation of an original
displacement which has the form of an equilateral triangle. The string
preserves this form if it lengthens in the middle of the interval (xi, x2) while
the points Xi and x2 remain fixed. Figure
6 shows the successive behavior of the
string after a time interval of amount
x2 — Xi
At =
8a
If one wishes to exhibit the behavior of the
string in the course of a sufficiently small
time interval, then one can, so to speak,
group together snapshots of the propaga¬
tion of the original length.
If we place the characteristics in the
x, t phase plane through the end points of
the intervals P{xГі,0), Q(xz, 0) (Fig. 7), the
plane is divided into six regions—regions
I and V at the time t considered in which
lengthening has yet to occur, region III in
which maximum lengthening has already
occurred, and regions II, IV, and VI in
FIG. 6. which lengthening has just occurred.
As a second example we shall investi-
gate the case in which there is no initial lengthening, but in which an
initial velocity exists. Let it be different from zero only in the interval
(xi, x2) and possess there the constant value ф0. Then we can write
u(x, 0) = <p(x) = 0
Ut(x, 0) = (p(x) = фо for Xi S x й x2,
= 0 for X > X2 or X < Xi .
The function u(x, t) is given also in this case as the sum of two waves.
Thus W{x) is the integral of ф(а) and represents the profile of a wave pro¬
pagating to the left:
V{x) - j ф{а)сіа .
0
2-2. WAVE-PROPAGATION METHOD 41
¥(x) = 0 for X 5; Xi ,
change of the momentum of the object struck during the time of the stroke.
Let the change in velocity of the point in the interval Ax be equal to v,
where v is the initial velocity. Under the assumption that the initial velocity
v is constant in Ax, then we obtain the change of momentum,
pvAx = I,
^int — v — —t— x, x + Ax ,
pAx
</>ext = 0 X, X + Ax ,
x — at = x0 — at0
x + at = x0 + at0 .
They determine two angles oa and a2,
the so-called upper and lower characteristic
angles at the point (x0, to) •
The action of a point impulse at the
fig. io. point (x0,t0) produces a lengthening which
in the interior of the above characteristic
angles equals 1/2a • 1/p and outside the interval equals zero.
Of interest to us now is the region in which the solution is uniquely
defined by the initial conditions when these are prescribed in a given interval
PQ of the lines t = 0.
Formula (2-2.9) shows that it suffices for the determination of the func¬
tion и at any point M(x, t) of the x, t phase plane (Figure 7) when the initial
conditions in the interval PQ are known. Thus, P, Q are the points of the
x axis with the coordinates x — at and x + at. The segments MP and MQ
of the characteristics passing through the point M and the segment PQ of
the x axis form a triangle MPQ called the characteristic triangle of the
point M.
If the initial conditions are not given on the entire line — oo < x < oo
but are given only in a fixed interval PQ, then these initial conditions define
2-2. WAVE-PROPAGATION METHOD 43
the solution uniquely within the characteristic triangle which has the interval
PQ as a base.
The proof of this theorem is surprisingly simple. The functions ифх, t) and
u2{x, t) are linked to the initial values by formula (2-2.9), so that
I rx+at
I <Рі(х — at) — <p2(x—at) I
у-\ \фі(ос) — </)2(a)\da,
+ 2 + 2 a J x—at
l + ^o
Every physically defined process must be capable of description through
functions which depend continuously on those initial conditions determining
the process. If the solution of a boundary-value problem depends continuous¬
ly on the initial conditions, then one also says that the boundary-value problem
is well set or the solution is stable.
If this continuous dependence did not exist, there could be two essentially
different processes corresponding to practically the same set of initial con¬
ditions (whose difference lies within the limits of the accuracy of measure¬
ment); that is, the solution would not be stable. It cannot be asserted that
such processes are determined by the initial conditions (in a physical sense).
From the above theorem, it follows that the vibrations of a string are deter-
44 HYPERBOLIC DIFFERENTIAL EQUATIONS
mined not only mathematically but also physically by the initial conditions.
We shall now consider such a problem in which the solution is not
stable. The functions resulting as solutions of the Laplace equation
uxx + uyy — 0
are defined uniquely by its initial conditions10
satisfy the Laplace equation. In uw(x, у), X plays the role of a parameter.
The initial conditions
can only be a solution of Eq. (2-2.1) provided <p{x) is once and <p(x) is twice
differentiable. Hence, it follows that the functions represented in Figures
u.
x, xz хг
11 and 12 cannot be solutions of Eq. (2-2.1) because they are not twice dif¬
ferentiable throughout. Beyond this, the assertion holds in that a solution of
the wave equation does not exist which satisfies the initial conditions (2-2.2)
when (p{x) and ф(х) do not possess the required derivatives. By repeating
10 These conditions define uniquely the solution of the Laplace equation mathemati¬
cally. The origin of the function uy(x, 0), is, of course, equivalent to the origin of the
function vx(x, 0), where v(x, y) is the harmonic function conjugate to u(x, y). Hence that
analytic function of which the function u(x, у) is the real part is defined uniquely up to
an arbitrary constant. (See §4-1, 4-5.)
2-2. WAVE-PROPAGATION METHOD 45
the reasonings which led us to formula (2-2.9) we can show that from the
existence of a solution of the wave equation its representation follows ac¬
cording to (2-2.9). If, however, cp and ф are not a sufficient number of times
differentiable, then (2-2.9) defines a function which does not satisfy Eq. (2-2.1),
i.e., no solution to the problem exists.
If, however, we change the initial conditions a little—replace them by
differentiable functions cp{x) and ф{х)—then these new initial conditions cor¬
respond to a solution of Eq. (2-2.1). Moreover, it is still to be noted that
according to the proof of the last theorem we have in fact proved the con¬
tinuous dependence of the functions (p and ф defined by formula (2-2.9)—
independent of whether these are or are not differentiable. If, therefore,
certain functions <p, ф do not correspond to a solution of the wave equation
which satisfies the conditions (2-2.2), then the functions defined by (2-2.9) are
boundary values of the solutions of the wave equation with somewhat smoother
initial conditions.
u(0, t) — 0 or ux{0, t) = 0 ,
i.e., the propagation of the initial displacement of a string with a fixed end
point x = 0 (or a free end point).
For the solutions of the wave equation which are defined for the infinite
straight line, the following two lemmas are valid.
1. If the functions <p(x) and ф(х) occurring in the initial conditions with
respect to any point x0 are odd, the corresponding solution at this point
is equal to zero.
2. If the functions <p(x) and ф(х) occurring in the initial conditions are even
with respect to any point x0, then the derivative of the corresponding
solution with respect to * at this point equals zero.
Proof of the first lemma. We select x0 as the origin of coordinates, i.e.,
x0 = 0. The conditions on the function in question are odd; therefore we have
then the first summand vanishes, since <p(x) is odd, whereas the second, the
integral of an odd function within the limits shown, is symmetrical with
respect to the origin and likewise is equal to zero.
Proof of the second lemma. The conditions on the functions in question are
even; therefore we have
then the first term of the sum vanishes since <p'(x) is odd and the second
vanishes, since ф(х) is even.11
With the help of these two lemmas the following problems can be solved:
Find a solution of Eq. (2-2.1) which satisfies the inidal conditions
и (x, 0) = <p(x)
(2-2.2)
ut(x, 0) = ф(х),
and the boundary condition
u(0, 0 = 0
(first boundary-value problem).
The functions Ф(х) and ¥(x) defined by the relations
defined with their help is defined for all x and t > 0. According to the first
11 These two lemmas are consequences of the fact that for even (or odd) initial con¬
ditions the function u(x, t), given by the formula of D’Alembert, for t > 0 is likewise
even (or odd); we leave the proof of this to the reader. Geometrically, one sees im¬
mediately that an odd continuous function, as well as the derivative of an even dif¬
ferentiable function, vanishes at x = 0.
2-2. WAVE-PROPAGATION METHOD 47
u(0, t) = 0 .
Moreover, u(x, t) for t — 0 satisfies the initial conditions
«*(0, 0 = 0,
and, in fact, we take the even continuations of <p(x), and <p(x)
u(x, t) = —1 ¥{a)da
J x-at
or
ux(0, 0 —0.
In the following we shall frequently have occasion to use the above ap-
48 HYPERBOLIC DIFFERENTIAL EQUATIONS
FIG. 13.
2-2. WAVE-PROPAGATION METHOD 49
process proceeds for x > 0 as if it were initially on the infinite straight line.
Accordingly, the reflection occurs at the fixed end, and finally a wave moves,
whose profile in this case is an isosceles trapezoid with constant velocity
toward the right.
The investigation of the reflection at a free end proceeds analogously only
if the initial conditions are continued evenly, so that the reflection of the
50 HYPERBOLIC DIFFERENTIAL EQUATIONS
wave at the free end does not proceed with a changing phase but with the
same phase.
Finally, we shall consider problems with homogeneous boundary conditions
u(0, t) = /jt(t) = 0
or
ux{0, t) = v(t) = 0 .
In the general case of nonhomogeneous boundary conditions the solution
can be written as a sum, each of whose terms satisfies only one of the stated
conditions (either the boundary or the initial condition).
We turn now to the solutions of differential equations with homogeneous
initial conditions and prescribed boundary conditions. Let
Then
and therefore
This function is defined, however, only for the region x — at -й 0 since fi(t)
is defined only for t ^ 0. In Figure 15 this region is shown as the shaded
part of the phase plane. Now in order to determine u(x, t) for all the values
FIG. 15.
2-2. WAVE-PROPAGATION METHOD 51
is defined for all values of the argument and satisfies the homogeneous initial
conditions.
The sum of these and function (2-2.12) defined earlier represent the solu¬
tion of the first boundary-value problem for the homogeneous wave equation:
= filt
<p(x + at) — ip(at — x) 1 r+ttt
+ + <p(a)da for t >— .
2a a
(2-2.13)
Analogously one can construct the solution of the second boundary-value
problem. For the third boundary-value problem see page 61. We shall
limit ourselves here to the solution of the boundary-value problem for the
homogeneous wave equation. For the solution of the nonhomogeneous wave
equation see page 61.
We shall take now a bounded interval (0, l) as a basis and begin our
investigation with the search for the solution of
Hit Cl tlxX t
U{1, 2
t) = fJt (t) ,
For these we shall seek the solution by means of the method of continua¬
tion. Hence we construct the expression (Ansatz)
u(x, t)
Ф(х + at) -f Ф(х — at)
Sx+at
¥(a)da ,
2 x—at
In order that now Ф{х) and ¥(x) satisfy the homogeneous boundary con¬
ditions, we require that Ф(х) and ¥(x) be odd with respect to the point x = 0,
x = /, i.e.,
Ф(х) = - Ф(- x) , Ф(х) = - Ф{21 - x)
¥(x)=-¥(-x), ¥{x) = - V(2l - x) .
t, /-21
' a
STT І yk
jp 4-X
тЖр
“41
/l\
■y^v4 1
ІХ X"FXi4 1 N
+ІХ7=Х
1—
i
i
1 1
1
x -01 \X- ti i ! 1 1
i 1 1 1
1 1 1
i 1 Ui , 1 I ! 1
i I i 1 1
1 ' - i i 1 ' ! 1 1
i 1 1
i i i 1 1
1 1 i i 1 i i i
i i 1 t
1 t i i i i i 1 1
i 1
1 i '
i 1 ! 1 ' 1 1
i 1 1 1
1 1 1
i i i 1 1
i
1
1
1
1
1
1 1 1 1
I 1
! !
1 ■ 1 1i >
У'Х ! i 1/NJ
FIG. 16.
signs (the phase) of the displacement (in the form of an isosceles triangle).
With the aid of this figure one can easily illustrate the profile of the string
at any arbitrary time t. Thus one can recognize at time t = 2l/a a displace¬
ment which coincides with the original displacement. The function u(x, t)
is therefore a periodic function of t with period T = 2l/a.
We shall now consider the propagation of the boundary effects. For this
purpose we shall seek the solution of the equation
_ 2
Utt ^ Mxx
is a solution. This function, however, does not satisfy the boundary condition
The reflected wave which propagates to the left and at x = / has a displace¬
ment of magnitude Д(і — l/a) is represented analytically by the equation
_l_ l — x I . 21 x
W = H[t-+ —
a a a a
The difference between the two waves, i.e.,
2 nl ” _/ 2nl
u(x, t) = 1 v(t - 1M + — (2-2.14)
n=0 a
This contains only a finite number of terms distinct from zero since the
argument with each new reflection about 2l/a is decreased whereas ДЦ) = 0
for / < 0. That the boundary conditions are satisfied we prove directly when
we set x = 0 in (2-2.14). The summand for n = 0 of the first sum is then
equal to whereas the remaining terms of the first and second sums are
cancelled pairwise for equal n values. Thus u(0, t) = цЦ).
If we now replace n by n — 1 and vary accordingly the summation limits,
then the first sum reads
2 nl 21 — x
I
n= a a
If we now set x — l, it can be seen directly that the summands of the first
and second sums mutually cancel each other.12
Formula (2-2.14) has a simple physical significance. First
12 The initial conditions can likewise be proven directly, since the arguments of all
functions for t = 0 are negative.
54 HYPERBOLIC DIFFERENTIAL EQUATIONS
6. Wave dispersion
ал 2^12^ T a22a — 0
bt — b2a = 0 (2-2.17)
c=0 .
If the differential equation for waves has solutions of arbitrary form then
we speak of a lack of wave dispersion. For this, it is necessary and sufficient
that the conditions (2-2.17) be satisfied.
From the first relationship follows the wave velocity:
a — fli2 dr a12 2
a22
2-2. WAVE-PROPAGATION METHOD 55
bt = b2 = c = 0.
$ = x - yt , у=t,
we obtain the equation
which for у = #i2/#22 coincides with the equation of the vibrating string.
In this case we have
2
Ut) tj — CL
with
For elliptic differential equations (a?2 — #n#22 < 0), waves with real velo¬
cities as solutions are not possible. For parabolic differential equations
(#*2 — #n #22 = 0), solutions in the form of waves with real velocities likewise
are rejected. There is an exception in the case of Eq. (2-2.15), which de¬
generates into an ordinary differential equation.
By taking into consideration that al2 = l/#ii • l/#22, # = #іг/#22 = l/#n/#22,
b2_
byux + b2ut (l/ #11 и* +1/ a22 #i) •
Л/ #22
The resulting equation can be written in the form
и = 0
d2u b2
+ b^- = 0 , b = -^=
de #? л/ #22
d_ /— Э /— d
— V an~z + V a22——
ds dx at
Consequently, in this case every arbitrary function of the variables
Ѵогг _ . _ _£
V = t
T/ou a
is a solution of the equation. Solutions in the form of propagating waves
are thus possible for the simplest wave equations in a moving or nonmoving
coordinate system.
In physics the concept of wave dispersion is usually introduced some¬
what differently.
Thus, consider a harmonic wave of the form
where w is the frequency, k = 2к/к is the wave number, and к is the wave
length.
The velocity with which the phase of the wave moves in space,
a = at — kx ,
is called the phase velocity and is obviously equal to
One speaks of wave dispersion if the phase velocity of the harmonic wave
is dependent on the frequency.
An impulse or a signal of an arbitrary form can be represented by a
superposition of harmonic waves of the form (*), that is, by a Fourier in¬
tegral. If the phase velocity depends on the frequency, the harmonic signals
are displaced relative to each other so that a distorted signal appears. In
this case the wave dispersion takes place in the sense of the definition on
page 54.
If the solution of the equation under consideration, which yields a wave
of arbitrary form, can be inverted, the phase velocity can be determined from
the first equation of (2-2.17) and thus does not depend on the frequency.
The concept of wave dispersion in the sense of our definition therefore
coincides with the property that the phase velocity depends on the frequency.
We shall now determine the class of Eq. (2-2.15) which permit solutions
in the form of damped waves
The function /u(t) thus satisfies an ordinary differential equation with con¬
stant coefficients and has the form
—kt
(t — e
If we set the coefficients of the last equation equal to zero, then we obtain
for the determination of a and k the relations
йи — 2 ana + a22a2 = 0
(bx — b2a) — 2k(an — a22a) = 0 (2-2.19)
a22k2 — b2k + c = 0 .
By elimination of a and k from (2-2.19) we obtain a condition for the compat¬
ibility of these three relations. The first equation shows that only the
hyperbolic differential equation allows damped waves as solutions. The
damping coefficient k is obtained from the second relation. Accordingly if
we insert k into the third equation, we arrive at the following relation be¬
tween the coefficients:
If these are satisfied, solutions of the equation exist in the form of damped
waves.
Example: The “telegraphic” equation
1 - a2CL = 0 .
, CR + LG
k= ZCL -■
The condition for the validity of Eq. (2-2.19) reads:
where Ф(х, t) is a continuous function and C„ is the arc of the boundary C*,
which approximates C (Figure 17).
Let t = tn(x) be the equation of Cn and t = t{x) be the equation of C.
Obviously then tn{x) converges uniformly towards t(x) and we can write
ь
lim \ Ф[х, tn(x)]dx Ф[х, t{x)]dx ,
™—“ Jo
FIG. 17.
j jj f(x,t)dxdt = 0 (2-2.23')
и (x, 0) = <p(x)
ut(x, 0) = ф{х).
Solution. Let <p(x) be piecewise smooth and ф(х) and f(x,t) be piecewise con¬
tinuous. C is an arbitrary piecewise smooth curve which lies in the region
/ ^ 0. We shall prove that this problem has a single solution which can
also be determined by the D’Alembert formula.
Let us assume that the function u{x, t) is a solution of our problem.
Then we shall consider the triangle ABM, whose base lies on the x axis,
13 Since dx — 0 at the vertical parts of the polygonal arc Cn> t — tn(x) represents
in this formula the equation of a horizontal part of the polygonal arc Cn.
60 HYPERBOLIC DIFFERENTIAL EQUATIONS
whose upper vertex is the point M(x, t), and whose remaining sides have been
formed from the corresponding segments of the characteristics x — at — const,
and x + at = const. (Figure 18). Within this triangle we apply formula
(2-2.23'). Since along AM the relation dx/dt = a is valid, it follows that
dt dx \dt dx J
Along MB, by contrast, dx/dt ——a is valid; we obtain, therefore,
du , . 2 du ( du ,, . du , \ ,
-—dx + a -—dt = —a -—dt + -^—dx — —adu .
dt dx \dt dx J
Consequently, the integrand along the characteristics is a total differential,
and by integration along BM and MA we immediately obtain
u(B) + u{A) B du , 1
u(M) = +
iri ——dx + ——
л dt 2a ABM
f dx dt
or
j rx+at (*x+a(t-t)
<p{x + at) + cp(x — at)
u{x, t) + Ф(№ + _ dv\ /(£, :№•
2a L-ai^
x-at J x—a U-'
(2-2.24)
where /i,/2 are piecewise smooth and /3 is piecewise continuous, satisfies Eq.
(2-2.22) and therefore also Eq. (2-2.23). The solutions of the problems con¬
sidered as examples in § 2-4 are piecewise
smooth functions and thus are accounted for
by the above theory.
We shall now turn to the first boundary-
value problem for the semi-infinite line and
seek a solution of Eq. (2-2.23) at a point
M(x, t) for t > x/a, since in the region t < x/a
(outside the characteristic x = at) the influ¬
ence of the boundary conditions still does not
enter into the phenomenon and the solution FIG. 19.
2-2. WAVE-PROPAGATION METHOD 61
or
д:+а(і —r)
For /=0, this formula, as is easily seen, coincides with formula (2-2.13).
Similarly, one obtains the solution of the second boundary-value problem,
as well as the solution of problems for a bounded interval.
For the investigation of the first boundary-value problem we saw pre¬
scribing both the initial conditions
was sufficient for the complete determination of the solution. Hence it follows
that a relation must exist which connects the functions <p, ф, p, v with v(t) —
ux(0, t). If we differentiate (2-2.25) with respect to x and set x — 0, we get
x — at = const. x + at = const.
62 HYPERBOLIC DIFFERENTIAL EQUATIONS
X = x(t)
du , . 2 du du , . 2 du ,,
——ax + a -—at + —ax + a ——at = 0
at ox DO + OB dt dx
du , . 2 du j, du , . 2 du \ j, л
-—ax + a ——dt
BA+AD dt dx dt ox /1
du du t . 2 du \ i. л
dx + a —л: + a —— ) dt = 0 ,
DO+OB dt dt dx
where the parentheses ( )t and ( )2 will indicate that the corresponding limit
values are taken from the interior of Л or zf2 respectively. By subtraction
of the first equation from the sum of the last two equations we obtain
(2-2.27)
The square brackets indicate here, as is usual, the magnitude of the jump
of the function, i.e.,
[/]=/.-/1.
We now form the derivative of the function u(x, t) with respect to t
du du
—u(x(t), t) = x' + £ — 1,2
dt dx dt
Therefore, we can choose both the value of the derivatives and the limit
values corresponding to and J2 (taken from the interior). The difference
of the right sides for i = 1 and i = 2 is
= 0.
2-2. WAVE-PROPAGATION METHOD 63
If we equate this equation with (2-2.27) and assume at least one of the jumps
[du/dt], [du/dx] to be different from zero, we see that both equations are simul¬
taneously satisfied if the determinant of the system vanishes:
r'
XU n 2 / /\2 2 a
, = (x ) — a = 0
or
x — +at + const.
Problems
= 0 for x > 2c .
t0 — 0, k = l,2,---,8.
64 HYPERBOLIC DIFFERENTIAL EQUATIONS
They are in the x, t phase plane, which is interpreted for the zones
corresponding to the different states.
2. Find the solution of 1(1) for all values of x and t (thus the formula
representing u{x, t) which differs for the different zones of the phase plane).
3. Determine the displacement at a point (x0, t0) by using the x, t phase plane
and the x, и plane in which (Figure 21) an initial displacement (<p = 0) is given
for an unbounded string and also for a semi-infinite string with a fixed (or
a free) end.
4. At one end of a long cylindrical tube filled with gas there is a piston
which moves according to an arbitrary law x — f(t) with the velocity v =
/'(/) < a. The initial displacement and the initial velocity of the gas particles
are equal to zero. Find the displacement of the gas at the plane with co¬
ordinate x. In this case assume that the piston moves with a constant velocity
c < a. What can be said about the solution of the problem if the piston at
a certain time has a velocity v > a?
5. A wave u(x, t) — f(x — at) propagates along an unbounded string. Choose
the condition of the string at the time t = 0 as the initial condition and then
solve the wave equation with the corresponding initial conditions; compare
with problem 1(1).
6. By joining two elastic rods with the characteristics
Uehal
о
V
kl, for
II
ki , Рг i for x > 0
Г p2
u{x, t)
is prescribed where / is the given function. Find the reflection and trans¬
mission coefficients that are due to the passage of the wave at the boundary
(x — 0). Investigate the conditions under which no reflection of the wave
occurs.
(b) Solve the corresponding problem when the local initial displacement
u(x, t) — f I
FIG. 22.
Mavp
Ui(x, t) [1 for x — at < 0
direct wave: 27
= 0 for x — at > 0
Mav о [l_e-<*r/*.u(*+.«>]
fl , A wave:
reflected
,
Ui{x’l)
,ч
= ~OlT
^
for x — at < 0
= 0 for x — at > 0
G_ R_
C - L '
66 HYPERBOLIC DIFFERENTIAL EQUATIONS
If
v(x, 0) = f{x)
11. Find the solution of the integral wave equation for a semi-infinite string
when the boundary conditions prescribed are of the third kind.
12. A membrane is fastened at the end x = 0 of a semi-infinite spring. It
is undergoing resistance due to the longitudinal vibrations of the spring and
is proportional to the velocity uti0, t). Investigate the course of vibrations
if the initial displacement is known and ut(x, 0) = fix) = 0.
The differential Eq. (2-3.1) is linear and homogeneous. Hence the sum of
individual solutions again is a solution of this equation. Accordingly, when
we have a sufficient number of individual solutions, the desired solution is
obtained by summing the individual solutions when each solution is multi¬
plied by a suitable coefficient.
For this purpose we shall consider the following fundamental lemma:
Find a solution of (2-3.1), which does not vanish identically, which satisfies
2-3. SEPARATION OF VARIABLES 67
X"(x) _ 1 T"(t)
(2-3.6)
X(x) a2 T{t)
Thus, if the function defined by (2-3.4) is a solution of (2-3.1), Eqs. (2-3.5)
and (2-3.6) must be identical, i.e., for all values of x, such that 0 < x < l, t > 0.
The right side of (2-3.6) is a function of t alone, while the left side is de¬
pendent only on x. If, for example, x is fixed and t changes (or conversely),
we see that the right and the left sides of (2-3.6), for changes in their argu¬
ments, remain constant:
X"(x) _ 1 T"(t)
(2-3.7)
X(x) a2 T(t)
where X is a constant which for the following investigations is appropriately
marked with a minus sign. Nothing will be assumed about the sign of X.
From the relation (2-3.7) we then obtain for the determination of X(x)
and T(t) the ordinary differential equations
u(0, t) = X(0)T(t) = 0
u(l, t) = X(l)T(t) = 0 .
Hence it follows that X(x) must satisfy the auxiliary conditions
while in particular if
exists and find the corresponding solutions. The X value for which a solution
68 HYPERBOLIC DIFFERENTIAL EQUATIONS
X(0) = C, + C2 = 0 ,
i.e.,
Since now a for X < 0 is real and positive, we have ea — e~a Ф 0. Thus
Cy = 0 , C2 = 0
and
X(x) = 0.
X(x) = ax + b ,
while the boundary conditions are
X(x) = 0.
3. For X > 0, the general solution possesses imaginary exponents and
therefore can be represented in the form
X(0) = A = 0
X(/) = D2 sin \/ll = 0 .
If X(x) does not vanish identically, then D2 Ф 0 so that
or
2-3. SEPARATION OF VARIABLES 69
Xn(x) = Dns'm~-x.
(2-3.13)
-\r / \ TtYl
•
Xn(x) = sin —j~x > (2.3.14)
which, except for arbitrary factors which we have set equal to unity in (2-3.14)
throughout, are determined uniquely. The solutions of Eq. (2-3.9) corre¬
sponding to these X values are
л 7T71 . r, . Tfyi
I \ 7m I
un{x, t) = Xnix)Tn(t) = Ancos—j-at + BnSm-j-at (sin (2-3.16)
are specific solutions of Eq. (2-3.1) which satisfy the boundary conditions
(2-3.2). They can be described by (2-3.4) as the product of two functions, of
which one is dependent only on x and the other only on t. These solutions
must satisfy the initial conditions (2-3.3) of our original problem only for
prescribed functions <p(x) and ф(х).
We turn now to the solution of the problem in the general case. Because
of the linearity and homogeneity of (2-3.1) the sum of the particular solutions
From the theory of Fourier series it is now known that an arbitrary piece-
wise continuous and piecewise differentiable function f(x), which is defined
in an interval 0 й x sS /, can be developed in a Fourier series
where14
Hence the function (2-3.17) which gives the solution to the problem under
consideration is defined completely.
The solution can be determined in the form of an infinite series (2-3.17).
If, however, series (2-3.17) diverges or the function determined by this series
is not differentiable, then, of course, it is not the solution of our differential
equation.
We shall limit ourselves at this point to the formal construction of the
solution. The conditions under which the series (2-3.17) converges and rep¬
resents a solution will be investigated in Section (2-3.3).
If F(x) is defined only in an interval (0,/) then F(x) can be continued oddly and developed
in the interval from -l to +/. This then leads to formulas (2-3.19) and (2-3.20).
2-3. SEPARATION OF VARIABLES 71
, ( л ttn , n . лп \ . лп
un(x, t) = ( zUcos ——at + Bn sm—j-fltfjsin-y-
with
Ttfl . 7ГН
un(xо, t) = an cos —-a(t + dn) sin ~j~x0
7Г n
an sin —j~x° •
Thus
When cosa>n(/ + <5„) = ±1 at the time t, the displacement reaches its maxi¬
mum value when the velocity equals zero. By contrast the displacement for
those t values for which cos wnit + dn) = 0 is valid equals zero but the
velocity is maximum. The frequencies of vibration of all points of the string
coincide and have the values
ЛП
(On —-—CL (2-3.26)
The frequencies oon are called the eigenfrequencies of the vibrating string.
For a transverse vibrating string a2 = Tip, and accordingly
лп /Т
GO (2-3.27)
n_ / V ~p '
72 HYPERBOLIC DIFFERENTIAL EQUATIONS
The energy of the nth standing wave (the nth harmonic) for a transverse
vibrating string equals
F —
=Ц.Ш)+ Tif)]dx
• 2 КМ i I 2 7ГП 7 l
ь
sin ——xdx = \ cos ——xdx — — .
/ l
By using the expressions for am and wn as well as the relation T = a2p, we
obtain
2 2 /I 2 I D2
77 pOCn^n / 2 -^T-n I Dn
bn — 1—t—I — 0)nM--- , (2-3.29)
4 4
because
• 7Г - л
sin — £ > 0.
Аг= 0,
and the lowest tone corresponding to the frequency
If the string (0, l) is touched at the point x — 1/3, it increases the pitch of
the basic tones threefold, since here only those harmonics whose nodes lie
at the point x = Z/3 remain. The formula
or (2-3.30)
for the frequency or the period of the basic vibration explains the following
rules which were first found experimentally. (1) For strings of uniform
density and uniform tension, the vibration period is proportional to the
length of the string. (2) In a prescribed length of string, the period is
inversely proportional to the square root of the tension. (3) For prescribed
lengths and tension, the period is proportional to the square root of the linear
74 HYPERBOLIC DIFFERENTIAL EQUATIONS
u(x, t) = X(x)T(t)
is equivalent to the existence of standing waves, since the profiles of such
a solution at different times are proportional to each other.
15 See V. I. Smirnov, Textbook of Higher Mathematics, 2d ed., Part II, Berlin, 1958.
2-3. SEPARATION OF VARIABLES 75
From this it then follows that u{x, t) depends continuously on the correspond¬
ing initial and boundary conditions. Here it is sufficient to prove the uniform
convergence of the series which represents the function u(x, t), since the
general term of the series is a continuous function, and a uniformly conver¬
gent series of continuous functions represents a continuous function. There¬
fore we can proceed as follows: From the well-known inequality
\un(x,t)\ й I An \ + \Bn\
we conclude that the series
2 (I An I + I Bn I) (2-3.33)
is a majorant of the series (2-3.32). If the series (2-3.33) converges, then the
series (2-3.32) converges uniformly, and u(x, t) is continuous.
Furthermore, in order to see that ut(x, t) depends continuously on the
initial conditions, the continuity of this function must be demonstrated; there¬
fore, it is sufficient to prove the uniform convergence of the series
£ du„ ПП 7ІЯ ПП
ut(x, t) ^ TT
— 1 a~T~( — AnSm-^B^at + Bn cos -^-at jsin-^f1* (2-3.34)
' n 1 dt l l l l
/ у d2un n x- 21 j-,
Л • nn j \ .nn . J. ГЬ HfL
2. n ( An cos ——at + Bn sin ——at sin —— x
'Л dx2 l ) n—\ l l l
■ИЛ nn Л . nn
2
Z d2un nn
un n ( An cos—j-at + B.sin — at \ sm-j-x
n=i dt2
Since
An - фи J Bn -фп ,
nna
where
2 f! . . . 7in , . 2 [l .... nn ,
<рп = —l <p(x) sin ——xdx , = —j ФМ sm~J~xdx •
76 HYPERBOLIC DIFFERENTIAL EQUATIONS
f nk 1 <pn 1 , k = 0,1,2
n=1
(2-3.37)
Іпк\фп\ , k =- 1,0,1
n —1
is demonstrated.
To demonstrate the convergence of (2-3.37) we use some well-known pro¬
perties of Fourier series.16
If a periodic function F(x) with period 2/ possesses continuous derivatives
up to and including the kth order, while the (k + l)th derivative is only
piecewise continuous, then the series
while for the continuity of the odd derivatives no auxiliary conditions are
imposed.
For the convergence of the series
I nk 1 <pn I , k = 0,1,2
n=1
it is sufficient to require that the initial displacement <p(x) satisfy the follow¬
ing requirements:
Condition 1. The derivatives of <p(x) are continuous up to and including
the second order, the third derivative is piecewise continuous, and moreover
Іпк\фп\, k =- 1,0,1
n= l
where Ф and ¥ are the odd continuations with respect to 0 and l of the
prescribed initial functions ф(х) and <[>(x) defined in the interval (0, l). The
functions Ф and ¥ are, as shown, periodic functions of period 21 and there¬
fore can be represented by the series
л/ \ £ , • 7ГИ
Ф(х) = Z <pn sin——*,
nri \
¥(x)=Z<pn sin——x,
n 1 i n I /
where y>„ and фп are respectively the Fourier coefficients of the functions
<p{x) and ф(х). By insertion of this series in (2-3.42) we obtain, with the aid
of the corresponding addition theorem for the trigonometric functions,
u(x, t) = Z (<Pn cos —-—at H-фп sin ——at sin ——x , (2-3.43)
n i\ l nna l ) t
Thus Conditions 1 and 2 from the method of proof, while sufficient for
the exact foundation of the method of separation of variables, depend on and
contain additional conditions, in comparison with conditions which guarantee
the existence of the solution.
For the foundation of the representation of the solution by a superposi¬
tion of standing waves, we used the first method for the proof of the con¬
vergence of the series, since it did not depend on the special form of solu¬
tion (2-3.42), which is applicable only for the solution of the simplest vibra¬
tion problems. Furthermore, this method is carried over without difficulty
to a series of other problems, although more stringent requirements are
imposed on the initial functions.
78 HYPERBOLIC DIFFERENTIAL EQUATIONS
4. Inhomogeneous equations
и (x, 0) = <p(x)
(2-3.46)
Mt( x, 0) = <p(x)
and the homogeneous boundary conditions
u(0, /) = 0
t > 0. (2-3.47)
u(l, t) — 0
Unity - (fn
(2-3.51)
Unity = фп •
These auxiliary conditions completely define the solution of Eq. (2-3.50). Now
un{t) can be represented in the form
ПП Д ЛП ,0 о C/U
+ I U, cos ——at H-фп sin ——at ) sin ——x .
, / nn t I I I • .
(2-3.54)
n=i\ / тіпа 11 l
The second sum solves the problem of a freely vibrating string with pre¬
scribed initial conditions and was completely investigated earlier. The first
sum by contrast represents the forced vibrations of the string under the
influence of an external force with homogeneous initial conditions. With
the use of (2-3.49) for fn(t) we find
with
£o ^ ^ £o + , т0 й т.й t0 + At
of a point Mo(£o,r0), is different from zero and vanishes outside this neigh¬
borhood. The function pfi£, r) represents the force density of the acting
force; the force developed in the interval (f0.fo + ^£) is therefore equal to
№ flrff •
fo
Then
Stq+Jt
F(T)dz = p\
Г то+Лт +
f(Z,z)d$dz
tq J r0
is the impulse of the force during the time dr. By applying the mean-value
theorem to the expression
pZ pZ p то+Лт p
L(U) = 0,
with the initial conditions
U{i)(0) = 0, г = 0,1, — 2
{/'"-■’(О) = 1 . (4*)
.(5*)
i.e., the differential equation is satisfied. Obviously, u(t) also satisfies the
initial conditions (2*).
For the function U(t) and formula (3*) we can easily find a clear physical
interpretation. Usually u(t) designates the displacement of a given system
and f(t) is the force acting on this system. Our system is in a state of rest
for t < 0. Let the displacement of the system be given by the non-negative
function let fe(t) be different from 0 only at the time 0 < t < e. We
denote the impulse of this force by
1= f / (z)dr .
J0
Further let ue(t) be the function corresponding to /e(/). Therefore we can
consider e as a parameter and set 1=1. Obviously then, as e—>0, lim ue(t)
£—►0
exists independent of the selection fe(t). Also one easily recognizes that this
limit value is equal to the function U(t) defined above; namely
if (/(/) = 0 for t < 0. U(t) is called the influence function of the instantaneous
impulses.
If we apply the mean-value theorem to (3*) we obtain
/(» = i m, t=i
where
Then
U{t) = I Ui(t) ,
І—1
where Ui(t) is the solution of the equation L(ui) = f{ with homogeneous initial
conditions.
If m is sufficiently large, Ui(t) can be regarded as the influence function
of the instantaneous impulse of intensity
I = fi(Ti)JT — f (гі)Тг ,
so that for u(t) we have
u(t) — f U (t — z) f (z)dz .
J 0
so that
The integral representation (3*) derived above for the solution of the
ordinary differential Eq. (1*) has the same physical significance as formula
(2-3.59), which gave the integral representation of the solution of the homo¬
geneous wave equation.
The general first boundary-value problem for the wave equation reads:
Find a solution of the differential equation
и (x, 0) = <pix)
(2-3.46)
Utix, 0) = фіх)
84 HYPERBOLIC DIFFERENTIAL EQUATIONS
w(0, t) = /^(f)
(2-3.47)
u(l, t) = fi2{t) .
First, by means of
we introduce a new function v(x, t). This then signifies the difference of the
function u(x, t) from a function U(x, t) still to be determined.
The function v(x, t) can be determined as a solution of the wave equation
= 0 and /u2(t) = 0 .
Thus it suffices to set
Hence the general boundary-value problem for u(x, t) has been reduced
to a boundary-value problem for v(x, t) with homogeneous boundary condi¬
tions which we have already solved (see Section 2-3 §4).
и (x, 0) = <p(x)
(2-3.46)
ut(x, 0) = ф(х)
и (0, t) = и,,
(2-3.47')
и (l, t) — иг .
In this case we seek the solution of the problem in the form
агй"(х) + fa(x) = 0
m(0) = ut
u(l) = u2
while v(x, t) designates the displacement from the stationary state. It is
easily seen that
u(x) = Ul + (ut - +
* ( Jo Jo a Jo jo &
If, in particular, /„ = const., then
Vu = a2vxx
with the homogeneous boundary conditions
v(0, t) = 0
v(l, t) = 0
and the initial conditions
We can easily see that for the solution of this problem it is not necessary
to know the exact analytic expression for й(х).
According to formula (2-3.17), v(x, t) has the form
where
Bn — 0
and
2 f! _
An= —— 1 u(x)Xn(x)dx .
I JO
The following method is appropriate for the calculation of these integrals.
From (2-2.8) we find
Xn(x) =-±Xi'(x).
An
We introduce this expression into the formula for An and integrate twice
by parts:
An -u'Xn + [ U"Xn(x)dx\ ,
0 J0 J
from which, by taking into consideration the differential equation and the
boundary conditions for й(х), there follows
_2_
An —
or
21 fo(x)
An -[«*(- 1)" — Ml] — 2 Xn{x)dx .
nn Tin2 ) о
In particular we find for the homogeneous equation (f0(x) = 0)
With the aid of this method it is possible to calculate the Fourier coefficients
also for the boundary conditions of the second and third type and for the
boundary-value problems of the inhomogeneous string
+ *p[x)X = 0 ’
if the eigenfunctions and the eigenvalues are known.
As shown above, the general first boundary-value problem for the equa-
2-3. SEPARATION OF VARIABLES 87
The term aut on the right side corresponds to a frictional force proportional
to the velocity.
We shall first investigate the consequences of periodic boundary influences:
u(l, t) = A cos wt or u(l, t) = В sin wt , (2-3.62)
u(0, t) =0. (2-3.63)
If
u(x, t) = un>(x, t) + iu(2){x, t)
satisfies Eq. (2-3.61) with the boundary conditions (2-3.62) and (2-3.63), then
un)(x, t) and uw{x, t), the real and imaginary parts of u(x, t), separately satisfy
the same equation (since it is linear); the condition (2-3.63) and the boundary
conditions at x = l lead to
un)(l, t) = A cos wt , uw(l, t) = A sin wt .
We therefore seek the solution of the problem
u(0, t) =0 (2-3.65)
«(/, t) = Аеш .
With the expression
88 HYPERBOLIC DIFFERENTIAL EQUATIONS
u(x, t) = X(x)eiu,t
we obtain the following problem for X(x):
X(l) = A . (2-3.68)
X(x) = C sin kx .
The condition at x = l gives
A
C (2-3.69)
sin kl
Thus
where Xi{x) and X2(x) are the real and imaginary parts of X(x).
We can then represent the sought solution in the form
k = lim k = — (2-3.71)
a 0 a
and correspondingly
j/ .4 _ л sin (w/a)x
и \x, t) — lim и
12
(x, t) = A— sin wt . (2-3.73)
a -o sin (w/a)l
The functions и (x, t) and и (x, t) are obviously solutions of the equation
litt — CL lixx
iim(0,t) = 0 йП)(0, t) = 0
A solution for a = 0 does not always exist. Thus if the frequency w of the
forced vibrations coincides with a characteristic frequency wn of the vibrat-
2-3. SEPARATION OF VARIABLES 89
then the denominator vanishes in the equation for ма> and uw and no solu¬
tion exists for the problem considered here.
This fact has a simple physical significance, namely: for w = wn there
occurs a resonance. At a given time t = t0 the amplitude grows without
bound. With the existence of friction (а Ф 0) a vibration process is possible
for each w.
If f(t) is a periodic function which can be represented by the series
where w is the lowest frequency and An,Bn are the Fourier coefficients,
then the solution of the equation for a = 0 has the form
_ / ,, Tin . . D . nn A . nn
2 ( An cos -j-at + Bn sin —j~at J sin ——x ,
where An and Bn are arbitrary constants, then the sum so obtained likewise
satisfies the equation and the boundary conditions.
In order that problem (1) for a = 0 be uniquely solvable we introduce
the additional condition on the vanishing friction:
We say a solution of the problem (2) has vanishing friction if it is the
limit value of a solution of problem (1) as a—>0. The problem for а fixed
end x — l and with a prescribed boundary condition u(0, t) = /x(t) at x = 0 is
solved in a similar manner.
The solution of the general problem without initial conditions
v(0, t) = 0 , v(l, t) = 0 .
The Fourier coefficients for v
Vn “l- OCVn “r ^ j ^ f ( )
a) a , / a2 2 (2) a /a2
+ a,n> Qn -y--и», « > 0
are the roots of the characteristic equation. There are now two possible cases:
1. The roots are real and negative.
2. The roots are complex and possess a negative real part. From this
it follows that every solution (**) of Eq. (*) either is identically zero or is
such that its absolute value as t—*—oo becomes arbitrarily large. Now,
however, | vn \ < 4M is valid for all n so that v„ = 0 for all n.
Hence we must have
and the condition which determines the magnitude of the jump in the first
derivative at the point x0 at which the concentrated force f(t) acts:
The functions Xx and X2 on the other hand must satisfy the boundary con¬
ditions
92 HYPERBOLIC DIFFERENTIAL EQUATIONS
which results from (2-3.77) and also satisfy the continuity and jump conditions
^о) w , со cd ., .
гл A
C— COS —X0 + D— COS —(l — X0) =— .
a a a a k
Thus the problem for the case f(t) = A cos cut and F(f) = Asinwf is
solved. Now if f(t) is a periodic function, say
18 The first summand of these sums corresponds to the stationary deflection which is
ami A 7:m
sin-= 0 , wn = ——a - a)m ,
a l
i.e., if the frequency spectrum of the driving force contains one of the
characteristic frequencies of the vibration.
If one of the nodes of the standing wave is found at the point x0 of
action of the force which corresponds to the free vibration with frequency
a>„, then
Here the numerators of the corresponding summands for и are equal to zero,
and no resonance occurs. However, if a displacement of the corresponding
standing wave of frequency a>m is found at the point of action of the force
which acts with the frequency a>m, then
sin-x0 — 1 ,
a
The method of separation of variables can be used not only for the wave
equation of a homogeneous string, utt = a2uxx, but also for the wave equation
of the inhomogeneous string,
where k, q, and p depend on x and are positive (k > 0, p > 0, q > 0).'9
For example, consider the first boundary-value problem for Eq. 2-3.84:
19 The case that k{x) vanishes at some point must be considered separately.
94 HYPERBOLIC DIFFERENTIAL EQUATIONS
u(0, t) = 0 , u(l, t) — 0
and is representable as a product
u(x, t) = X(x)T(t).
If we substitute this product expression into the partial differential equa¬
tion, we obtain, with consideration of the boundary conditions for X(x) and
T(t), the two ordinary differential equations
4~\тЩ
ax L ax J
-qx + ipx = о
T" + ЛТ = 0 .
For the determination of the function X(x) we are led to the following
eigenvalue problem:
Find those values of the parameter X for which the boundary-value problem
^ Xm{x)Xn{x)p(x)dx = 0 , m Ф n . (2-3.89)
X" + pX = 0 , p=-^X
fc 0
X(0) = 0 , X(l) = 0,
which we have already investigated.
2-3. SEPARATION OF VARIABLES 95
F(x) = f FnXn(x),
»=>
Fn = -i- Y F(x)Xn(x)p(x)dx
iVnJo
Nn = j Xl(x)p(x)dx . (2-3.90)
The proof for propositions 1 and 4 are usually given in the theory of
integral equations. We shall limit ourselves to the proof of propositions 2
and 3.
Before we prove these properties we shall first derive Green’s formula.
Let u(x) and v(x) be two functions which are twice differentiable in the inter¬
val a < x < b and possess continuous first derivatives in a S x й b. Next we
consider the expression
Proof of the third proposition. Let Xm(x) and Xn(x) be two eigenfunctions
with the corresponding eigenvalues 2m and Xn. Then if и = Xn{x), v = Xn(x)
are inserted in the formula (2-3.91) we obtain, by consideration of the bound¬
ary conditions (2-3.88),21
(^n Xm{x)Xn{x)p{x)dx — 0 .
11 Xm(x)Xn(x)p(x)dx = 0 , (2-3.92)
i.e., the eigenfunctions Xm(x) and Xn(x) are orthogonal to each other with
respect to the density function p(x). We shall now prove that every eigen-
k{x)Xn(x) — j \q — Xmp)Xmdx + C .
From this follows in particular the existence of the derivative X^(x) for x = 0 and x =1.
96 HYPERBOLIC DIFFERENTIAL EQUATIONS
X*(0) = 0
_X'(0)
Z(0) = 0 , Z'(0) = 0
is identically zero and, therefore, no eigenfunction can exist.
If Xn(x) is an eigenfunction corresponding to the eigenvalue Xn, then
AnXn(x), where An is an arbitrary constant, because of the linearity and the
homogeneity of the equation and the boundary conditions, is likewise an
eigenfunction corresponding to Xn.
We proved earlier that the class of eigenfunctions has been completely
exhausted. The eigenfunctions which differ from each other only by con¬
stant factors will appear as not essentially different. However, in order to
avoid ambiguity in the selection of factors the eigenfunctions will be normal¬
ized by
Nn = j Xl(x)p(x)dx = 1 .
If, at the outset, an arbitrary eigenfunction Xn(x) does not satisfy this nor¬
malizing condition, then it can be normalized by multiplication by a suitable
coefficient An:
22 This property of the first boundary-value problem therefore depends on the fact
that two linearly independent solutions of a differential equation of the second order
cannot be at one and the same point equal to 0. This assertion is based on the bound¬
ary-value problem with homogeneous boundary conditions. With other boundary con¬
ditions (for example, with X(0) = X(l), X'(0) = X'(l)) two different characteristic func¬
tions can be given which correspond to one and the same characteristic value.
2-3. SEPARATION OF VARIABLES 97
AnXn(x) = Xn(x) , An
l a
Xn{x) p(x)dx
L[Xn] = — inp(x)Xn{x).
Multiplication of both sides of this equation with Xn{x) and integration with
respect to x from 0 to l yields
Xl{x)p{x)dx = — j Xn(x)L[Xn\dx
or
Jn = — ( Xn-^7-\k(x)^^-~\dx + i q{x)Xl{x)dx ,
Jo dx L dx J Jo
1 Г! Г1
^n — nk^Cn + \ k(x)[Xn(x)fdx + j q(x)Xl(x)dx
in > 0
since by assumption k(x) > 0 and q{x) ^ 0.
We soon arrive at the calculation of the development coefficients of
F(x), but not at the proof of the development theorem itself, which we
shall use in this case. Obviously,
( p(x)F(x)Xn(x)dx
Fn = -^гт-. (2-3.94)
j p(x)Xn(x)dx
F(x) =f n~ 1
FnXn(x)
We turn again to the partial differential Eq. (2-3.84). For T(t) we obtain
the equation
without any auxiliary conditions. Since Xn is positive, its solution has the
form
The formal scheme for satisfying the initial conditions (2-3.86) depends on
the Steklov theorem (4) above. It is completely analogous for the homo¬
geneous string, and we find from the equations
the relation
Hence (pn and фп are Fourier coefficients of the functions (p{x) and ф(х) which
occur in the development of these functions with respect to the orthogonal
system of functions [Xn{x)} with the weight factor p(x).
Since we have limited ourselves here to the general scheme of separation
of variables, the conditions for the applicability of this method, with regard
to both the coefficient of the equation and the initial functions, have not been
discussed.
The basic ideas of this method are due to V. A. Steklov.23
Problems
23 Report of the Kharkov Mathematical Society, second series, Vol. 5, nos. 1 and 2
(1896); also Fundamental Problems of Mathematical Physics, Vol. 1 (1922).
2-3. SEPARATION OF VARIABLES 99
bn — v^--\--p—
<t>n
( <p(x) s\n~[/InXdx
Wn Jo
a>„ = л/аг?.п — v .
.
13 Let an isolated electrical conductor of length / with characteristic co¬
efficients L, R, C, and G = 0 be charged to a fixed potential v0. One end of
the conductor is grounded at the initial moment while the other remains
isolated during the entire time of the process. Determine the voltage dis¬
tribution in the conductor.
Solution.-.
(2 n -P l)n CR42
°>n ~ Я1
2lVLC^ Ln (2n + If
4y0 . о L
anП г. I 1 \ • > фп 2<J)n .
n(2n + 1) sin (pn R
14 . A string which is rigidly fastened at the ends vibrates under the in¬
fluence of a harmonic force which is distributed with density f(x, t) — Ф(х)
sin wt. Determine the displacement u{x, t) of the string with arbitrary initial
conditions. When is resonance possible and what solution results in the case
of resonance?
15 . Solve Problem 14 under the assumption that the vibrations occur in a
medium in which the resistance is proportional to the velocity. Find the
vibrations which comprise the principal part of the solution for / —» °o.
16 . An elastic rod of length l is in a vertical position and at its upper end
a free-falling elevator is fastened, which at the moment it is stopped, has
attained a velocity v0. What vibrations are performed by the rod if the
other end is freely movable?
17. Solve the equation
Uu = a uXz -p b и -p A
with homogeneous initial conditions and the boundary conditions
u(0, t) — 0 u(l, t) = В ,
where b, A, and В are constant.
.
18 Solve the differential equation
u(0, t) — В u(l, t) — C ,
where A, B, and C are constant.
.
19 On a homogeneous string with rigidly fixed ends x = 0 and x = l acts a
harmonic force
2-4. PROBLEMS WITH AUXILIARY CONDITIONS ON THE CHARACTERISTICS 101
u(x, 0) - —X for 0 5S x ^ c ,
c
u{l, t) = A sin wt .
We shall first consider the simplest of the problems in which the auxiliary
conditions are prescribed on the characteristics. It reads as follows:
uxy = f(x, y)
u(x, 0) = <pi{x) (2-4.1)
w(0, y) = <p2(y) .
The auxiliary conditions here are given on the straight lines x = 0 and v = 0,
which represent the characteristics of Eq. (2-4.1). We assume the functions
<Pi(x) and (p2(y) to be differentiable and related to each other by the relation¬
ship <р2(0) = <p2(0). If we integrate Eq. (2-4.1) with respect to x and then with
respect to y, we obtain
or
Therefore, the solution in this case, in which the differential equation contains
neither the first derivatives nor the sought function itself, can be represented
explicitly by the analytic expression (2-4.2). From (2-4.2) the uniqueness
and the existence of the solution of our problem follows directly.
We shall now seek the solution of the linear hyperbolic differential equation
We shall determine its solution with the help of the method of successive
approximations. For the zero-th approximation we choose the function
U0(x, y) — 0 .
Eq. (2-4.4) then gives for the successive approximations the expression
diin-i
+ a{Z, v)~ + b(Z, + c(Z
dZ 3ry
dun dlli
= + [ \a{x, rj)—in -1- + b(x, г})дНп 1 + c(x, r])un-Adr]
dx M JoL OX Of] J
(2-4.6)
ou
У) "-1 + c(£
ОУ Оу JoL Oi; dy
v) dzZ.~ dtdr; ,
drj + c(t> r])zn-i(^> 7)1
dznjx, y) _ dun+lix, y) dun(x, y)
dx dx dx
dzn
< 3HMnKn~l (x + y)n-
dy n\
2-4. PROBLEMS WITH AUXILIARY CONDITIONS ON THE CHARACTERISTICS 105
where К > 0 is a constant number whose value will be specified later. With
these estimates and the formula for the (n + l)th approximation we obtain
after a series of simplifications
К = 2L + 2 .
On the right sides of these inequalities we have, to within constant factors,
the general term in the series development of e2KLM. Thus the above estimates
show that the sequences of functions
Un = Uo + Z\ + • • • + Zn-1
By passage to the limit under the integral signs in (2-4.5) and (2-4.6) we then
find
v = ux w — uy
establish the fact that u(x, y) satisfies the integrodifferential equation
and also the original differential Eq. (2-4.3). We verify this directly by dif¬
ferentiation of (2-4.4) with respect to x and y. Moreover, as one easily sees,
u(x, y) satisfies the auxiliary conditions.
The uniqueness of the solution of the stated problem will now be de¬
monstrated. If we assume the existence of two solutions и{(х, у) and u2(x, y),
we obtain for their difference
Now if Hi is a common upper bound of the magnitudes \U\, \UX \ and | Uy\
for 0 S x S L and 0 ^ у ^ L, then by a repetition of the estimates which led
to those on the function zn(x, y), the correctness of the inequality
(x + ѵГ2 3 Hi (2 KLM)n+*
I U\ < 3tf,Mn+1/r <
(n + 2)! K2M (n + 2)!
for an arbitrary value of n follows easily. Hence it follows that
и = ve^y,
can be brought to the form
For C| = 0 we obtain the simple Eq. (2-4.1) whose solution was given by
formula 2-4.2.
By contrast, if Ci Ф 0 then we obtain for the solution of (2-4.8) an explicit
analytic representation by the method given in Section 2-5.
Problems
1. Air with a velocity v passes through a tube (x > 0) which is filled with a
moist fluid. Let v(x, t) be the concentration of the moisture in the saturated
2-5. SOLUTION OF GENERAL LINEAR HYPERBOLIC DIFFERENTIAL EQUATIONS 107
substance and u(x, t) be the concentration of the free vapor. What equations
are satisfied by the functions u(x, t) and v(x, t) describing the drying process
when (a) the process proceeds isothermally; and (b) the dry isotherm has the
form и = у ■ v where у is the isothermal constant (see also Section 2-6, Applica¬
tion 5)?
2. Hot water with velocity v flows through a tube (x > 0). Let и be the
temperature of the water in the tube, v the temperature of the walls of the
tube, and u0 the temperature of the environment. Derive the equations for
и and v by neglecting the temperature distribution in the individual cross
sections of the tube and the walls. Assume further that a temperature
gradient exists at the boundaries of the water-tube wall and the wall environ¬
ment, and that the flow of heat follows Newton’s law (see Chapter 3-1).
L[u\ = uxx — uyy + a{x, y)ux + b(x, y)uy -I- c(x, y)u , (2-5.1)
where a(x, y), b(x, y), c(x, y) are differentiable functions, be a linear differential
operator corresponding to a linear hyperbolic differential equation. We
multiply L[u\ by a function v and write the individual summands in the form
rr i wr i 3K (2-5.2)
vL[u] — uM[v] = —— + —
dx oy
where
if the difference
vL[u] — uM[v]
is equal to the sum of the partial derivatives of the expressions H and К
with respect to x and y.
In particular, if L[u\ = M\u\, then the operator L is called self-adjoint.
For the double integral, the difference vL[u\ — uM[v\, where и and v are
twice differentiable functions throughout a region G which is bounded by a
piecewise smooth curve C, we have
We shall apply Formula (2-5.6) for the solution of the following problem.
Find a solution of a linear hyperbolic differential equation
L[u] = uxx — Uyy + a(x, y)ux + b(x, y)uy + c{x, y)u — — f{x, y) , (2-5.7)
1 {vL[u] — uM\v])d£dr)
M J PQ
We perform the first two integrals which are taken along the characteristics
MQ and MP, and by taking into consideration the relations
ds
dt; =— dr] on the characteristic QM
l/2
2-5. SOLUTION OF GENERAL LINEAR HYPERBOLIC DIFFERENTIAL EQUATIONS 109
ds
d£ = drj on the characteristic MP
l/2
(where ds is the element of the arc along QM and MP), and the formulas
(2-5.4) and (2-5.5), we obtain
dv a +b \ j
(Hdrj — Kd$) = — \ d(uv) -f
ds 17Tvrs
and correspondingly
(Ь-а)/(»ѵТ) is
v = e *° on MP
C (6 + o)/(2V2) ds
v = e s° on MQ
The stated problem is solved by this formula since the expression which
stands under the integral extended along PQ contains only functions which
are known along C. The function v has already been defined above while
the functions
и \c = <p{x)
ip'(x) - <P(x)f'(x)V\ + fn(x)
ux — us cos (x, s) + un cos (x, n) =
1 +/'*(*)
cp,(x)f'(x) + ф(х)Ѵ l+f'\x)
Uy — Us cos (y, S) + Un cos (y, n) =
1 +f'\x)
can be calculated from the initial conditions.
If the initial conditions along the arc PQ are known, the function и in
the triangle PMQ can be completely determined according to formula (2-5.10)
when f{x, y) is known in this region.25
Formula (2-5.10), obtained under the assumption of the existence of a
solution, defines the solution from the initial conditions and the right side of
Eq. (2-5.7). Most important, it also proves the uniqueness of the solution
(compare it with the D’Alembert formula, Section 2-2).
It can also be shown that the function и defined by (2-5.10) satisfies the
conditions of the problems.
In order to clarify the physical meaning of the Riemann function v(M, M')
we shall first determine the solution of the inhomogeneous equation
25 If the characteristics intersect the curve C at the two points P and Mi (Figure 26),
the value u(M\) cannot be arbitrarily given. On the contrary, it is determined accord¬
ing to formula (2-5.10) by the initial value on the arc PQi and the value of f{x,y) in
APMiQi.
2-5. SOLUTION OF GENERAL LINEAR HYPERBOLIC DIFFERENTIAL EQUATIONS 111
J j/i(MW = l. (2-5.12)
then denotes the impulse of the force. Hence, because of (2-5.11) we conclude
that v(M, Mi) is the influence function of the unit impulse acting at Mx.
The function v(M, Mx) = v(x, y, £,rj) can be defined as a function of the
parameter M{x,y), which with respect to the coordinates of the point
Mi satisfies the equation
M,e.„[y] = 0 (2-5.16)
with the auxiliary conditions (2-5.9').
We shall now consider the function
и = u(M, Mi) ,
which depends on the parameter M,(f, rj) and with respect to the coordinates
x, у of the point M satisfies the equation
L и. у)\м\ — 0 (2-5.17)
with the auxiliary conditions
du b — a
и on the characteristic MxQx ,
8s 2i/2
(2-5.18)
^ on the characteristic МХРХ,
ds 2\/ 2
u(Mi, Mi) = 1 .
{vL[u\ — uL[v])dt;dy
=
pi
{Hdrj - Kd£) +
I .if Je,
+[Qi+ p
J Mi JPi
= 0
If we use formulas (2-5.4) and (2-5.5) for К and H and the conditions (2-5.9'),
the first two integrals on the right side can be easily calculated
l“l Гя!
{Hdrj — Kd£) = \ [— (vu)(drj — {uv)vd£\ + j v[{2 u(dr] + 2 uvd£) + (audrj — bud£)]
ds
{Hdrj — Kd$) = {uv)Ml — (иѵ)ол , d£ = dr] = —y=
ЛГ, V b
2 (uv)M = 2 (uv)Ml
or
since
{V)M, = {V)M = 1 •
The influence function of the unit impulse acting at the point Mi can also
be defined as the solution of the equation
f
uyy = uxx + fy{x, t) , у = at, fy = »
a
u{x, 0) = (p(x) ,
Uy - фу{х) , фі=— .
a
As in (2-5.10) a strip of the axis у — 0 occurs as the arc PQ. The operator
L - uxx Uyy
because of
M — L — uxx — Uy,
is self-adjoint.
Because a — 0 and b = 0, v equals one on the characteristics MP and MQ.
Hence it follows
v(M, M') = 1
for every point M' which lies interior to the triangle PMQ.
Now in our case,
dr] = 0 for PQ .
Thus we obtain
u(P) + u(Q)
u(M) — —\ Uvds + -7Г /(6, ѵШѵ
PMQ
U = ueKl^y (2-5.24)
v = 1 on the characteristic MP ,
(2-5.28)
v — 1 on the characteristic MQ .
v = v(z) (2-5.29)
with
2 = V (X — f)2 — (y — 7]f
or (2-5.30)
z2 = (x — £)2 — (y — r])2 .
On the characteristics MP and MQ, z is equal to zero so that v(0) = 1.
The left side of (2-5.27) can be brought to the form
zzx = x — $
ZZy = — (y - v)
zzxx + z\ — 1
I 2 1
ZZyy I Zy — -L •
v" +— v1 + ctv = 0
z
with the condition v(0) = 1. A solution of this equation is the Bessel func¬
tion of the zero order (see Appendix)
v(z) = Мл/Тіг)
or
For the determination of U(x,y) we shall now use formula (2-5.10) which in
our case reads
U(P) + U(Q)
U(M) =
2
+ -ІГ
2
(vUvdZ - Uvr,d$) , drj = 0 . (2-5.32)
' 2 )i
- Vix -f)! - Л
uxx — uyy + cu = 0.
116 HYPERBOLIC DIFFERENTIAL EQUATIONS
«<*, y) = JO +ЛІ1 + » +
1
UXx 2 H
a
with the initial conditions
и (x, 0) = <p(x)
ut{x, 0) = ф(х)
ф(х) — аф(х) = auy{x, 0) .
Problems
1. Solve Problem 1 from Section 2-4 under the assumption that at the initial
moment the concentration of the liquid in the entire tube is constant and dry
air streams in through an opening.
2. Solve Problem 2 from Section 2-4 under the assumption that the initial
temperature of the system equals u0 while the temperature at the ends of the
tube during the entire time maintains a constant value v0 > u0.
3. Find a solution of the telegraphic system of equations (see Eq. (2-1.21)),
ix T Cvt T Gv — 0
vx T Lit T Ri — 0
i(x, 0) = <p(x),
v{x, 0) = ф(х).
Hint: First reduce the equation system to an equation of the second order
for one of the functions i(x, t) or v(x, t), then find, for example,
i(x, 0) = <p(x),
di
+ -J-i) = - \ф\х) - j-<p(x) = фо(х)
at (=0 L L, /1=о L* L,
Then use formula (2-5.35). Investigate the solution of the telegraphic equation
2-6. APPLICATIONS TO CHAPTER 2 117
which is obtained for small G and R, formula (2-5.35). Consider the limiting
case G—>0, /?—>0 and obtain from (2-5.35) the D’Alembert formula for the
solution of the equation of the vibrating string.
A vibrating string produces vibrations in the air which the human ear
perceives as the tone emitted from the string. The tone strength is deter¬
mined by the energy or the amplitude of the vibrations, the tone pitch by
the vibration period, and the tone color by the energy relation between the
lowest tone and the overtones.26
We shall not go into the physiological process of sound perception and
the propagation of sound in air; on the contrary the acoustic effect produced
by a string is characterized by the energy, the period, and the distribution
of the energy in the overtones.
Ordinarily in musical instruments transverse vibrations of the strings
are generated. We can distinguish three types of string instruments, depend¬
ing on whether the strings are stroked, plucked, or struck. Strings which
are struck (for example, in a piano) are thus given a fixed initial velocity,
but undergo no initial displacement. With the plucked instruments (for ex¬
ample, a harp or guitar) the strings are made to vibrate from a fixed initial
displacement without initial velocity.
The free vibrations of a string excited in an arbitrary manner can be
represented in the form (see Section 2-3)
2 hi2 • лпс
sin —j— ,
, A
bn — 0 (2-6.1)
an —
n n c(l — c)
provided that the initial displacement has the form of a triangle with the
height h at the point x — c (Figure 29). For the energy of the nth harmonic
vibration we find
, .. 2 К ^ 1 . ППС . nn . , ПП r o\
u(x, t) =- 2—sin—-—• sin — x • sin wnt , wn = —a (Z-b.3)
nap n=i nil l
4v0l 1 • nnc . nn ~ . nn . .
u{x, t) = 2 —5- sin - sin — 0 sin — X sin Wnt
n a n=1 n2 l l l
(see Section 2-3, Problem 3) while the energy of the individual harmonic
vibrations is given by
n
V0 COS for I x — с I <5
0) 0 ~2
at
0 for I x — с I >8 .
nn - • nn
. „ cos — 8 • sin — c
u(x, t) —
8 v08 j J_ /
2 - sin у л; • sinaij
n2a *-1 n 1
I
{see Section 2-3, Problem 3). The energy of the nth harmonic vibration is
2-6. APPLICATIONS TO CHAPTER 2 119
2 югд . 2 nnc
cos —-— sm
l
(3) A hammer which strikes a string is not ideally rigid. Then the
vibrations are no longer determined by the initial velocity but by a time-
varying force. Thus we arrive at an inhomogeneous equation with the
function
F(x, t) =
(T-, X— С
Г о COS—-— •
0
7Г . nt
— sin —
Z т
for I x — с I < d, 0 ^ t йт ,
in the right-hand member. The solution of this equation for t > r reads
2. Vibrations of rods
27 When, for example, the basic frequency (the first harmonic) corresponds to 440
vibrations per second —A in the center octave—then at the sevenfold frequency the
A, G of the fourth octave is produced. The interval A-G, the so-called minor seventh,
is unpleasant to the ear and is felt as a dissonance.
120 HYPERBOLIC DIFFERENTIAL EQUATIONS
fixed at its ends as in a vise. The determination of the form of the vibra¬
tions of a tuning fork and its frequencies leads to the equation of a transverse
vibrating rod,
g2v d*y
+ a‘ (2-6.5)
dt2 ox
From this equation we are led to several problems in vibrations of rods, the
calculation of the stability of rolling waves, and investigations of the vibra¬
tion of vessels.28
To give an elementary derivation of (2-6.5) we shall consider a rectan¬
gular rod of length l (0 ^ x S l) of height h and of width b (Figure 30). Let
dy dy
dcp dx
dx dx x+dx
is valid. The section of the rod which is at a distance rj from the rod axis
у = 0 changes its length by an amount rjdy> (Figure 31). Therefore, according
to Hooke’s law the tension acting along the section is equal to
(2-6.6)
where
dM = Fdx .
(2-6.7)
where p is the density of the rod and 5 is the area of the cross section
(here we neglect the rotary motion arising from the distortion), we obtain
the equation of a transverse vibrating rod:
(2-6.5)
(2-6.8)
At the free end the moment of flexure (2-6.6) and the tangential force
(2-6.7) must be equal to zero. This yields
In order to completely define the motion of the rod even the initial
conditions (the initial displacement and initial velocity) must be given in the
form
(2-6.10)
122 HYPERBOLIC DIFFERENTIAL EQUATIONS
T"(t) Yu\x)
a2T(t) ~ Y(x) ~ Л ’
Y^-XY = 0 (2-6.12)
dY d2Y d3Y
Y = 0 , = 0 , = o,
x=0 dx x=0 dx2 x=l dx3
sinh2 y/ll — sin2 у/XI = cosh2 y/J 1 + 2 cosh y/X l cos fyX l + cos2 y/11.
The roots of Eq. (2-6.14) can be calculated without difficulty29, as, for example,
graphically we find
A*! = 1.875 ,
[iz - 4.694 ,
Иг = 7.854 ,
*» For the calculation of the roots of Eq. (2-6.14) see Rayleigh, op. cit., fn. 26.
2-6. APPLICATIONS TO CHAPTER 2 123
The last formula accurately gives the value of pn for n = 3 to the third
decimal place and accurately for n = 7 up to the sixth place.
Now we shall consider the frequency of vibration of a tuning fork. The
equation
T" + a2i.nT = 0
4 = 6.267 , 4 = 17-548
Pi Pi
the second eigentone is more than two and one-half octaves higher than the
base tone, i.e., higher than the sixth harmonic of the string with equal base
tones, while the third eigenvibration is more than four octaves higher than
the base tone. If the tuning fork, for example, has as a basic frequency of
440 vibrations per second (usually one takes for A' the small A of the
first octave) then the following eigenfrequency of the tuning fork equals
2757.5 vibrations per second (between C"" = 2637.3 and = 2794.0, thus
between the E and F of the fourth octave of uniformly tempered scale) while
the third eigenfrequency of 7721.1 vibrations per second already is higher
than those frequencies used in music.
If the tuning fork is set to vibrating by a blow, higher frequencies ap¬
pear in addition to the first, which explains the initial metallic sound. The
higher harmonics are nevertheless quickly damped so that the tuning fork
soon rings out with the pure basic tone.
Fi = -MiUuiXi, t) ,
124 HYPERBOLIC DIFFERENTIAL EQUATIONS
then we obtain
where p is the density of the string. Let ue(x, t) be a solution of this equation.
By integration of Eq. (2-6.18) with respect to x from xx — e to Xi + £ and
passage to the limit as s—>0, the condition (2-6.17) results for the function
u(x, t) = lim Ue(x, t). We shall not go into a more precise investigation of the
8—0
passage to the limit.
The complete formulation of our problem then reads:
Find a solution of the equation
u{0, t) = 0,1
(2-6.20)
u(l, t) = 0, J
u(Xi — 0, t) = u{Xi + 0, t)
i (2-6.21)
MiUtt{Xi, t) = kux I
T" + IT = 0 (2-6.24)
and
4-(kX') + ЛРХ= 0,
dx (2-6.25)
X(0) - 0 X(l) = 0 .
2-6. APPLICATIONS TO CHAPTER 2 125
X (Xi — 0) = X (хі + 0)
МіХШТ" = kX' \:\t°0T.
By consideration of (2-6.24) we can write this relation in the form
Xi(x) , X2(x), • • •
of the problem
j Xm(x)X„(x)p(x)dx = 0 , m Ф n . (2-6.28)
We now distribute each mass Mi with uniform density <5 in the interval
X, — e < x < Xi + e, where e > 0 is an arbitrarily small number. Hence we
arrive at a corresponding vibration problem for the inhomogeneous string
with density pe(x). Let and Xn(x) be the eigenvalues and eigenfunctions
respectively of this problem. The eigenfunctions must satisfy the orthogo¬
nality relation,
j Xem{x)Xen(x)pe(x)dx = 0 . (2-6.29)
We multiply the first relation with Xn(x), the second with Xm(x), and
subtract the resulting equations from each other. If we integrate this dif¬
ference over (0, Xi), {Xi, хг), • ■ • (хк, l) and add the results, we find
Xj(Xi — 0) — Xj(Xi + 0) . /о с. оn
, , i — m, n . (l-b.A7
kX j(Xi + 0) — kX j(Xi — 0) = —MiXjXj(Xi)
We shall not at this point go into the proof of the existence of infinitely
many eigenvalues and eigenfunctions, the positiveness of the eigenvalues,
and the corresponding development theorem. The boundary-value problem
considered above, as also the problem investigated in Section 2-3, leads to an
2-6. APPLICATIONS TO CHAPTER 2 127
fix) = І fnXnix)
71 = 1
\ f (x)Xn(x)p(x)dx + I Mif{Xi)Xn{Xi)
fn = ^-. (2-6.33)
Nn
Muuil, t) = —kuxfl, t) .
X'r! + XnXn = 0
sing/ lnx
Xn(x) =
sin VYnl
where is fixed by
The orthogonality condition for the functions Xn(x) assumes the form
Xn(x)Xn(x)p(x)dx + MXn(l)Xm(l) = 0 .
)
Nn = j Xl(x)p(x)dx + MX 1(1) .
N =— + — + ^—21
i /ТіТГ 2n + 1 n
V An — ry
2 l
(2) M = со. The end x — I is rigidly fixed: u(l, t) = 0. Here the eigen¬
values can be calculated from
Of interest to us now are the eigenvalues for small M {M —>0) and large
M (M—> oo).
(a) M is small. In order to find the corrections to the eigenvalues
we set
VJn = VW + eM , (2-6.35)
К = a (2-6.36)
2-6. APPLICATIONS TO CHAPTER 2 129
i.e., the eigenvalues of the weighted string increase as M—*■ 0 and approaches
the eigenvalues of the string with a free end.
(b) M is large. Now we set
л/~Ап = л/.Af' + £— •
Vfff ’
where the terms which contain 1/M2 and higher powers of 1/M can be
neglected. Thus, there results
i.e., the eigenfrequencies with increasing loads becomes smaller and approaches
uniformly the eigenvalues of the string with rigidly fastened ends.
о
+
dv . dv dp
momentum equation: (2-6.39)
pTt pVTx - ~x
+
(2-6.41)
where the first member signifies the kinetic energy and the second, the
internal energy. Obviously, here e is the internal energy per unit of mass.
For an ideal gas, e = cvT where cv is the specific heat at constant volume
and T is the temperature. For the energy change in a unit of time we
therefore have
_a_
(2-6.42)
dt
If we carry out the differentiation in the first summand, then on the basis of
Eq. (2-6.38) and (2-6.39) we obtain
d / pv2 \ 2 dp dv v2 d . .
~~ + pv (2-6.43)
dt \ 2 ) dt dt YTX {"v)
To transform the derivative (d/dt){pe) we apply the first law of thermodynamics
which expresses the law of conservation of energy:
dQ = de + pdz . (2-6.44)
Thus dQ is the amount of heat which the system obtains from the outside
or gives to the outside, and pdz is the work which must be performed in
order to change the volume by an amount dz (z = 1/p is the specific volume).
If the process proceeds adiabatically (i.e., no heat exchange occurs with
the surrounding medium) then
dQ = 0 ,
and
Jr (pe) = (2-6.47)
at at
where
w = e +— (2-6.48)
P
is the heat function (enthalpy), or the heat content per unit of mass.
The derivative діѵ/dx on the basis of the expressions (2-6.46) and (2-6.48)
satisfies the equation
2-6. APPLICATIONS TO CHAPTER 2 131
d_
(2-6.50)
dt
d_ *2
'2/£t + P£)dx=-Pv(T w
dt 11
On the left side is the energy change per unit of time in the interval (x,_,
x2)\ on the right is the amount of energy which in a unit of time flows out
from the considered volume.
If the heat conduction can not be ignored, the conservation of energy
equation assumes the form
d_
(2-6.51)
dt
Further we write the continuity equation, the equation of motion, and the
conservation of energy relation in the following form,
(2-6.38')
!=->
(2-6.39')
\ d r■ fv‘ , Vi
/ “ dx[ fv(j + w) J (2-6.50)
In the x, t plane we consider the “spur” x = a{t) of the shock front. Let
AC be any arc of x = a(t). The points A and C must therefore have the
coordinates x,, tx, or x2 = Xi + Ax, t2 = tl + dt. Further, let the points В
and D be situated so that the sides of the rectangle ABCD run parallel to
the coordinate axes.
132 HYPERBOLIC DIFFERENTIAL EQUATIONS
Here the left side is the change of mass in the interval (xt, x2) during the
period of time (t,, t2); on the right side is the amount of fluid (the gas)
which flows out from (x,, x2) during the time (1,, t2). If the functions p and
pv are continuous throughout and differentiable in ABCD, Eqs. (2-6.52) and
(2-6.37') are equivalent. In the present case, however, this does not occur.
We apply the mean-value theorem to each individual summand, with the
result
Here x*, ***, t*, t** signify the corresponding mean values of x and t.
If we pass to the limit as dx —»0 (x2 —»■ xj and At —>0 (t2—> td and denote
the value of the function above the curve x = a(t), behind the front of the
shock wave, by the index 1 and the value of the function below this curve,
before the front, by the index 2, we obtain
where
TT da ,. Ax
U — — = lim —
at jt-oAt
is the velocity of the shock wave.
In the coordinate system moving with the shock wave we denote by
и, — U — v,, u2 = U — v2
the velocity of the elements before and behind the front. Therefore (2-6.53)
can be written in the form
This equation signifies the continuity of the matter flowing across the fronts
of the shock wave.
The law of conservation of momentum in integral form reads
x2 Г *2
[ipv)t2 — (pv)tjdx = — 1 [(p + рѵ2)Хг — (p + pv2)Xl\dt ,
xi Jq
where the right side represents the sum of the acting forces produced by the
impulses (pressure forces) and the flow of the momentum. As Ax—>0 and
At -> 0 we obtain
U[(pv)2 — (pv),] = —(p + pv2), + (p + pv2)2
or
the conservation law for the flow of the momentum across a front; and for
the conservation of energy on the front we obtain the equation
pv
£%- + P*) U~ + joe ) U — — piVA — + W ) + p2Vz(— + W
PiUi P2U2
or according to (2-6.53)
(2-6.55)
Thus, on the front of a shock wave, the following equations (the Hugoniot
conditions) must be satisfied:
. u\ . u\
Wi + ~2 — wi 3—ту • (2-6.55)
From both of the first Eqs. (2-6.53') and (2-6.54) we can express щ and uz
through p and p:
.fi — Pz Pi ~ Pi ni — Pi Pi~ Pi
Pi P1 pz Pz pi P2
and
1
£i — e2 (pi — Pi)(pi + Pz) •
2 pipi
For an ideal gas we have
Cp — У t P
p = RpT , s = cvT , w = cPT =
cP — cv у — 1 p
that is,
(2-6.56)
134 HYPERBOLIC DIFFERENTIAL EQUATIONS
pi (r + Dpi — (r — Dpi
From this, the quantities ti, pit t2, p2 can be calculated if the remaining three
are known.
A shock wave can move continuously with respect to the gas from
positions of high pressure to positions of lower pressure: p2~+pi (law of
Zemplen). Hence it follows that the density of gas behind the front is larger
than that before the front.
Formula (2-6.57) expresses the dependence between p2 and p2 for prescribed
pi and pi. For a given pi and piy p2 = p2{p2) is a monotone increasing function
which for p2lpi —> oo (shock wave of larger amplitude) tends towards a finite
limit value
Pi _ Г + 1 (2-6.59)
Pi Г ~ 1 ’
This formula gives the maximum jump in the density which can occur on
the front of the shock wave. For diatomic gases у — 7/5, and the maximum
density jump is then equal to 6:
Ti = 6
Pi
From Eqs. (2-6.53'), (2-6.54), and (2-6.57), we find for pi = 0,
Ml = m2 = (r- D2 Pi
2(r + 1) Pi
If the motion of the shock wave with respect to the gas is considered as
static (Vi = 0), the velocity of propagation of the shock wave equals
U = Jl±l.h,
2 pi
piston and the front of the shock wave a region 2 forms, in which the gas
moves with the velocity of the piston. Before the front (region 1) the gas
is found in the original state: p = plt p = pi (v = 0).
With the conditions (2-6.53'), (2-6.54), and (2-6.55) for the front, we can
easily determine its velocity, the magnitude of the jump, the density, and
the pressure.
To this end we introduce the nondimensional quantities
where cx = т/rPilpi is the velocity of sound before the front (in the non-
excited region 1). Then the equation of state can be written in the form
Since, obviously, a, < 1, (p2 > p{), the smallest root is given by
рг — Pi
1+(xTi-?)+x/^CT
16 C (2-6.68)
1 +
(r - 1)^
2c!
(2-6.70)
136 HYPERBOLIC DIFFERENTIAL EQUATIONS
Since the velocity of the shock wave is constant there, we find for the
position of the front at time t
x = «(/) = „ + c, /1 } t . (2-6.71)
In the limiting case v/ci » 1 (shock wave of greater intensity) we find from
the formulas (2-6.68) through (2-6.70) the limiting relations
r + 1 7 + 1 rir + 1)
Рг — p 1 U— v , Pi — Pi
r- 1 ’ 2 2 c\
obtained earlier.
If v/ci € 1 (shock wave of smaller intensity) then the term ѵг!с\ can be
neglected:
рг — Pi
U = Ci + (r + 1)
4
pt = 0 and pt = 0
because pi = pi = vt = 0 .
For an adiabatic process the equation of state of an ideal gas reads
P = Po(fJ ‘ (2'6'72)
For the solution of the problem we assume that
dj_= ]_ df
dt t ^ d£ ’ dx t d£
where f — p, v, or p, and substitute the found expression in (2-6.38) and (2-6.39)
then we obtain
, g..dp dv
(v - £)-£ = -P-Ji
dt; d£
(2-6.73)
. dv dp
lv-()Trd7-
Now if we multiply the first equation by (v — £) and add the product to
the second, we obtain
2 dp
(v-$)
d$
or
dp
= (v- $)2 .
dp
Hence
°-!=ЧТг±С'
where r is the velocity of sound for the adiabatic process.
138 HYPERBOLIC DIFFERENTIAL EQUATIONS
v — $ = —c . (2-6.74)
dv_c_
(2-6.75)
dp p
or
dv_1_
dp pc
cл = y—
P
P
and after integration of Eq. (2-6.75)
(2-6.76)
Г- 1
P — (Ool 1 + (2-6.77)
2 c0
Here we denote by
c0
Г Pa
the velocity of sound for v = 0 (in the static gas). For (2-6.76) we can also
write
Y- 1 ^,T/IHI
P — poll + (2-6.79)
2 Co
x
--Co (2-6.80)
r +i Vt
which gives the dependence of the velocity n on т and t. Accordingly, if
we insert expression (2-6.80) for v into (2-6.77) and (2-6.79), we obtain the
explicit dependence of the quantities p and p on x and t. Hence, all the
2-6. APPLICATIONS TO CHAPTER 2 139
Vi =-- c0 . (2-6.81)
T —1
From this it follows that the flow velocity of the gas in vacuum is finite.
For diatomic gases у — 7/5 and, therefore,
Vi = —5c0 •
We can also obtain the expression (2-6.81) for the velocity of the forward
front x = xx(t) from the relation
r*z
l pdx ~ poXz — poCot , (2-6.82)
J xi
which is the equation of conservation of mass. If, namely, we introduce
f = -
C t ’
we obtain
1 pd£ — p0c0 ■
J'1
Here we substitute the expression for p from (2-6.77). With
1 +
T + 1 co
it follows that
Гл 2 r-i
= (2-6.83)
'лі Г + 1
with
У — 1 _
li = 1 + +-І- ■ -—^ ,
V\ — C0
Ъ = 1
r + 1 Co
i.e.,
Л = 0 ,
where
2c0
Vi =
r- 1
The problem of the flow of a gas in a vacuum is therefore solved.
140 HYPERBOLIC DIFFERENTIAL EQUATIONS
~ = p{u - y) , (2-6.86)
ot
where ft is the so-called kinetic coefficient and у is the concentration of the
gas which exists in equilibrium with the amount of gas absorbed.
The quantities a and у are linked to each other by a relation
a = f(y) , (2-6.87)
f(y) =
u0 + py ’
then we speak of a Langmuir isotherm. A simpler form of the function /
corresponds to the so-called Henry isotherm, which is accurate in regions of
low concentration:
1
(2-6.88)
du da
(2-6.85)
= Hi+ Tt
~ = P(u - ya) (2-6.89)
at
determine the functions u(x, t) and a(x, t) which satisfy the auxiliary condi¬
tions
a(x, 0) = 0
(2-6.90)
u(x, 0) = 0
u(0, t) = Uo (2-6.91)
= p{u- у a) (2-6.89)
a(x, 0) = 0 , u(0, t) — u0 .
If we eliminate the function a(x, t) by introducing the second equation into
the first equation and differentiating with respect to t, then
35 For the system, Eqs. (2-6.85') and (2-6.89) satisfy a single condition, since the axis
t = 0 here is characteristic.
142 HYPERBOLIC DIFFERENTIAL EQUATIONS
uxt — — ut + = 0 (2-6.92)
v
a(x, 0) = 0 (2-6.90)
T — t — -
V
t = T + — e = x
d_ d_
dt ' дт
so that Eq. (2-6.85) has the form
du _ da
(2-6.85")
d£ dx
The initial conditions (2-6.90) and Eqs. (2-6.85) and (2-6.89) read
£: Sj: о.
The problem is therefore to determine the function и in the region between the straight
<2-6-90'>
lines t — 0 and the f axis corresponding to the initial conditions (2-6.90') (Cauchy pro¬
blem). Obviously, in this region u(x, t) = 0 but also a = 0). Further, from (2-6.85')
and (2-6.89) it is evident that the function u{x, t) for г = 0 is discontinuous, while a(x, t)
remains continuous. Therefore, и for r = 0, as was shown above, is defined by Eq.
(2-6.85') for a{x, 0) = 0. As shown on pages 141-142 (see formulas (2-6.93) and (2-6.95)),
when we define the value u(x, 0) and a(x, 0) we obtain for u(x, t) and a(x, t), problems
with auxiliary conditions on the characteristics.
2-6. APPLICATIONS TO CHAPTER 2 143
a— f(y) .
If we introduce the dimensionless variables
„ _xfi
X\ — -
_tfi
t1 — и - z = у v =■
г Щ M0 МоГ
then the system of Eqs. (2-6.85), (2-6.89), (2-6.90), and (2-6.91) becomes
eg |
CO
II
dxi dti
(2-6.97)
dv .
—— = (u — z)
dti
m(0, t) = 1 (2-6.99)
v(x, 0) = 0 . (2-6.100)
z =/л\ѵ) = F(v)
v = /і(м) or и = F(v) .
From (2-6.99) it follows
The condition (2-6.102) states that as i-+oo (f—>—00) saturation must occur
throughout.
If we insert (2-6.101) into (2-6.97), we obtain
1
(2-6.107)
P(f) ^=-00
Ml)
or (in the dimensionless quantities)
a = Г- (2-6.107')
do
From (2-6.105) and (2-6.106) we find
(2-6.108)
o<p — t(<p)
<0(9) = € — £0 i (2-6.109)
2-6. APPLICATIONS TO CHAPTER 2 145
where w(p) is an integral of the left side and f0 is the integration constant.
The sought function p(£) is defined by this relation to within the unknown
constant f0 :
P = - fo) (2-6.110)
We want now to determine whether the function oTl is defined at all and
whether the functions p and ф as £ -> oo and f —> — oo for the inverse function
satisfy the required conditions. Therefore we show that
da>
—a < 0 (2-6.112)
dp ^9 — f i \p)
i.e. £ — £o = «(у)
op — fAp) P — f i \p) •
The first summand represents the abscissa of points on the line corresponding
to the ordinate p which runs from the origin of the coordinates to the point
(1, /i(l)), (Figure 33). Since by hypothesis the curve p = /1(2) lies above these
lines, we have
and consequently
op — /1 \p) > 0 .
Moreover,
^ — fo = o)(p) — 00 f°r P — 0
? — f 0 = o){p) = — 00 for P = /i(l) •
ф — op 9 = 1 for f = — 00
/i(l)
1
ф = op = p = 0 for f = 00 .
Л(і)
146 HYPERBOLIC DIFFERENTIAL EQUATIONS
The conditions (2-6.102) and (2-6.103) are therefore satisfied. Besides this it is
shown that our system of equations permits a solution in the form of a pro¬
pagating wave which still contains an arbitrary constant ?0-
In order to determine £0 we integrate the first equation of (2-6.97) from
0 to t0 and 0 to x0:
We now assume that the solution of our problem for large values of t
approximates the functions й and v which we have found above as propagat¬
ing waves.
If then we determine ?0 according to the condition
then this is just the f0-value which corresponds to the functions й(х, t) and
v(x, t).
We form our integral in the following manner:
=Г
J—<TtQ—tQ
<o~\№ = Г аГ\Ж J£j
®_1(0) = <p* .
Then it is easy to see that
«r\Qdt: + w_1(C№
fi Jfl
“_1(^ p
Г-Сіа»"‘(С.) + a>(ip)d<p + (2-6.115)
is valid when (p = a> ‘(£) which is the inverse function corresponding to £ = w(tp)
(Figure 34). Hence it follows that instead of the conditions (2-6.114)
( <o~\OdC - to0 =
— -|(<rfo + 6o )<p(—oto — ?o) +
54>{-<Tto
-<r<0—fo)
u)((p)d(p J* — /о —> 0
о
<
J -<rtQ-fo
(2-6.115')
As t —* oo,
= [o — E'^V’o)] W — <Po) + • • •,
where
_(Up_
= d£ . (2-6.116)
[<J — F'(<p0)] (<p — <fo) +
The dots here denote the terms of higher order with respect to (<p — cp0).
From the requirement 3 for /, it follows that
1
F'(<p0) > <j
/,(1) '
If now in (2-6.115') we still let t0~> oo and bearing in mind (2-6.115") and
(2-6.115'"), we obtain
£o = (2-6.118)
_(Up_
= dt (2-6.119)
ay — <p/( 1 - pip)
where a — l//i(l) = 1 + p is the wave velocity. From (2-6.119) we find
£ — £o = o>(<p)
with
148 HYPERBOLIC DIFFERENTIAL EQUATIONS
If (p varies from 0 to /i(l), then obviously w(<p) varies from —00 to + 00.
We choose A so that
i.e.,
* 1 1 1
w(<p ) = 0 for
'—2Л<1,“2 l + p
A =-
a
and
12 16 20 24 2Ѳ
FIG. 35.
2-6. APPLICATIONS TO CHAPTER 2 149
6. Physical analogies
d2x , л
a + bx — 0 ,
dl _ c ЗУ
+ GV
dx dt
(2-6.121)
— = L — + RI
dx dt
where C, G, L, R are coefficients of capacity, the loss, the induction, and the
resistance of the system. If we can neglect R and G, the voltage V and the
current density / satisfy the ordinary wave equations
d2V _ d2V
=о
dx2 dt2
ЭѴ _ Trll = 0
dx2 dt2
dl =cdV
dx dt (2-6.122)
dV = L dl
dx dt
On the other hand, in the investigation of the propagation of sound in a
unique direction, we arrive—for example, for the motion of air in a tube—
150 HYPERBOLIC DIFFERENTIAL EQUATIONS
at the equations
dp _ dv
dx P dt
(2-6.123)
dv _ 1 dp
dx t dt
where v is the velocity of the vibrating particles, p is the density, p is the
pressure, and г = p0y is the elastic coefficient of air.
The similarity of Eqs. (2-6.122) and (2-6.123) point to the analogy between
the acoustic and the electrical quantities. The potential difference corresponds
to the pressure, the current density to the displacement velocity of the
particles. Further, the induction of the electrical system corresponds to the
density which determines the inertia properties of the gas, and the capacity
corresponds to the quantity 1/r, i.e., the reciprocal of the elastic coefficient.
This analogy can also be determined from the expressions for the kinetic
and potential energy of the electrical and acoustical systems.
By a glance at Eqs. (2-6.121) we can, by analogy with the corresponding
electrical quantities, introduce an acoustic resistance and loss. The magnitude
of the acoustic resistance is then to be considered if, in the motion, the
friction of the gas on the side of the vessel plays an essential role. By
analogy with electrical resistance, which is defined as the ratio of the voltage
to the current density, we define the acoustic resistance as the ratio of the
pressure to the current in the medium which is proportional to the displace¬
ment velocity of the gas particles:
uv
In those cases in which the motion of a gas in a porous medium is considered,
one has to introduce a quantity which corresponds to the loss in an electrical
system. This quantity (which we designate by P) is called the porosity and
is defined per unit volume for those materials which are filled with air.
The mechanical analogy of telegraphic equations are the equations of a
longitudinal vibrating rod. These can be written, similar to Eqs. (2-6.122), as
dv 1 dT
dx k dt
dT dv
dx
where T is the tension of the rod, v is the velocity of the vibrating points,
p is the density, and k is the elastic coefficient of the rod.
By comparison of these equations with Eqs. (2-6.122) we can define further
analogies, this time between mechanical and electrical quantities. If the
electrical voltage can correspond to the tension of the rod and the current
density to the motion velocity of the particles, then the reciprocal of the
elastic coefficient is known to correspond to the capacity and the density to
the induction.
2-6. APPLICATIONS TO CHAPTER 2 151
Variables
Voltage V Pressure, p Tension, T
Current density, / Particle velocity, v Particle velocity, x
Charge, e Displacement, и Displacement, x
Parameter
Inductance, L Inertia (density), p Mass density, pm
Capacity, C Acoustic capacity, Rigidity, elastic modulus
CA = 1/r CM = l/k
Resistance, R Acoustic resistance, RA Mechanical resistance, RM
The above table is based on the results of acoustic problems about the
character of the phenomena and gives an insight for the solution of the
problem.
Thus the problem of motion of air in a porous material for simple
harmonic waves leads to the equations38
— i(Dpmu + ru = —grad/)
Ap + (r — iwpm)p = 0
pc
where и is the spatial velocity of air through the pores, p is the pressure,
p is the density, pm is the effective density of air in the pores (pm can be
larger than p, since the material particles and the air can also vibrate in the
pores), P is the porosity, c is the velocity, w is the frequency of sound, and
r is the flow resistance. The latter can be characterized by prescribing the
pressure in the material. If we put r—RA, pm = LA, yPIpc2 = CA, then the
above equations take the form
CalA + CarJ£- = Ap .
ot ot
These equations are completely analogous to the equations for the propaga¬
tion of electrical vibrations in a conductor. Therefore, by analogy with the
wave impedance in conductors,
Z=
/ R T iu)L
G + iwC
G = 0
37 See, for example, H. F. Olson, Dynamic Analogies, New York, Toronto, London,
1944.
39 W. Furduev, Electroacoustics, Moscow, 1948.
152 HYPERBOLIC DIFFERENTIAL EQUATIONS
Z=
uxx — uy = 0
is usually denoted as the heat-conduction equation.
U(x) = Ul + X . (3-1.1)
equation we formulate first the physical laws which govern heat conduction.
1. The Fourier law If the temperature of a body is distributed non-
uniformly, then there arises in the body a heat flow in the direction of
temperature decrease.
The amount of heat which flows through the cross section at x during
an interval of time (t, t+dt) is equal to
dQ — qSdt , (3-1.3)
where
is the density of the heat flow. This is equal to the amount of heat flowing
through an area of 1 cm2 per unit of time. This law represents a generali¬
zation of formula (3-1.2). It can also be written in the integral form
Г12 ди
Q=-S\ k—(x,t)dt. (3-1.5)
Jq OX
Then Q is the amount of heat flowing in the time interval (tit t2) through
the cross section at x. If the rod is inhomogeneous, k is a function of x.
2. The amount of heat which must be added to a homogeneous body
in order to raise its temperature by an amount Au, is equal to
Q — cm Au — cpV Au , (3-1.6)
where c is the specific heat capacity, m is the mass of the body, p is its
density, and V is the volume of the body.
If the temperature change differs at different places of the rod or the
rod can be treated as inhomogeneous, then
*2
Q = cpSAu(x)dx . (3-1.7)
q
3. Within a rod, heat can arise or vanish (for example, by the flow of
an electric current, because of chemical reaction, etc.). The occurrence of
heat can be completely described by a function F(x, t), which is a measure
of the heat source at the point x at time f.39 The effect of these sources
on an element of the rod (x, x + dx) during an interval of time (t, t + dt)
induces an amount of heat given by
By integration we find
which is the amount of heat which exists in the interval (xlt x2) of the rod
in the time interval (t%, t2).
We obtain the equation of heat conduction from the law of conservation
of energy for an interval (xt, x2) in a time interval (tt, t2) and with the use
of formulas (3-1.5), (3-1.7), and (3-1.9) for the energy balance, we can write
,du.
-[ k —dudx {x,,
I ,
r)
,
— k—(x, t)
dx >+П“
F($, t) d£ dr
(3-1.10)
= 1 cp[u(£, t2) - «(f, ti)]d£ .
and hence with the help of the mean-value theorem of differential calculus
where t3, tt, tb and x3, x,, xb are suitable intermediate values in the intervals
(tу, t2) and (xi, x2).
If we divide the last equation by AxAt, we find
d_ du
+ F(x, t) (3-1.13)
dx cpm
All these considerations hold for arbitrary intervals {xi, x2) and (A, t2). Ac¬
cordingly if Xi, x2 tend towards x and , t2 tend towards t, there results
£(* ^
the so-called equation of heat conduction.
We shall next consider a special case.
1. If the rod is homogeneous then k, c, p can be taken as constant, and
the heat-conduction equation takes the form
ut = агихх + f(x, t)
a‘ = ±, /(*,,)= «£-'>■
cp cp
40 Through these restrictions on the function u(x, t), in general, we will lose a class
of solutions, namely those which satisfy the integral equation but not the differential
equation. In the case of the heat conduction equation, however, these requirements do
not exclude any possible solutions. That is, we can show that a function which satisfies
Equation (3-1.10) must be differentiable.
156 PARABOLIC DIFFERENTIAL EQUATIONS
ut = a2uxx. (3-1.14')
F0 = h{u — Ѳ) ,
F -- F^x, t) — h(u — Ѳ) ,
2. Diffusion equation
41 Since with our approximation the temperature distribution inside an individual cross
section was not considered, the influence of the surface sources is equivalent to the in¬
fluence of the heat sources in the interior.
3-1. SIMPLE PROBLEMS WHICH LEAD TO PARABOLIC DIFFERENTIAL EQUATIONS 157
scribed by a function u{x, t), which gives the concentration in the cross sec¬
tion at x at time t.
According to the Nernst law, the mass of the gas passing through the
cross section at x during the time interval (t, t + At) equals
dQ = - D^(x,t)Sdt , (3-1.16)
ox
where D is the diffusion coefficient and 5 is the area of the tube cross sec¬
tion.
From the definition of concentration,
Q = uV
results for the amount of gas found in the volume V. Here we know that
the change of mass of the gas in the interval (xi, x2) of the tube equals
c{x) Au-Sdx
*1
when the concentration changes by an amount An. Here c(x) denotes the
coefficient of porosity.42
The condition for the conservation of mass in the interval (xi, x2) during
the time interval (ti, t2) reads
д_ ( и ди\ _ c du
(3-1.17)
йж\ dx) dt ’
ut — auxx
with
Ul — Dllxx •
dQ — — k — u(x, у, z, t)dadt .
on
Here k is the heat conductivity of the body43 and n is the normal to the
surface element da in the direction of the heat flow. As is known,
ди ди , ч , du , \ , ди , , , лт
— = — cos (n, x) + — cos (и, у) + — cos (n, z) = grad u-N ,
on ox oy oz
so that we can write
dQ — — k grad u-Ndadt
where N is the exterior normal. Hence it follows that the heat flow per unit
of time and area is equal to
qn = q-N (3-1.18)
Qi = - k -~-dt da
or
Finally, if we denote the density of the heat sources by F(x, y, z, t), then
for the amount of heat freed in the volume V in the time interval (t{, t2)
we obtain
Q2 — Q3 — (Qi . (3-1.22)
Under the assumption that the function u, in the region considered, with
respect to the variables x, y, z, is twice continuously differentiable, and with
respect to t, is once continuously differentiable, we now form the relations
(3-1.19), (3-1.20), and (3-1.21). By applying Green’s theorem44
s Jr
and the mean-value theorem of integral and differential calculus for functions
of several variables, we obtain
Cp -І7 {x2, Уг, г2, ti) = - div q(Xi, Уі, zlt t3) + F(x3 ,y3,z3, t3) . (3-1.23)
at
Here all the values of the argument lie within the region considered, i.e.,
they are the coordinates of certain interior points of V for a certain value
of time in the interval (ti, t2).
Equation (3-1.23) holds for any volume V within the body. If we shrink
the volume to the point with coordinates x, y, z, and carry out the passage
to the limit tlf t2-+t, we obtain45
(3-1.25)
where
4 = jL.u JL + A
dx2 з/ a*2
is the Laplace operator.
= (3-1.28)
ax
We arrive at this condition when the heat flow Q{1, t) occurring at the end
of the rod is given by
Q(l, = t) .
ox
From this there results {du{l, t)dx) = v(t), where v(t) is a known function; then
we obtain
Q(L t)
v{t) =
k
3. At one end a linear relation exists between the derivative and the
function given by
3-1. SIMPLE PROBLEMS WHICH LEAD TO PARABOLIC DIFFERENTIAL EQUATIONS 161
-fl-(l,t)=-t[u(l,t)-0(t)\, (3-1.29)
ox
u(Xi — 0, t) = u{Xi + 0, t)
have an initial-value problem (the Cauchy problem) for the temperature dis¬
tribution in an infinite straight line, i.e.:
Find a solution of the heat-conduction equation in the region — oo < x < oo,
t ^ t0, which satisfies the equation
u{0, t) = fiAt)
(3-1.31)
u(l, t) = //,(/) .
According to the nature of the boundary conditions, other types of problems
without initial conditions are also possible. Of great importance is the prob¬
lem without initial conditions for a one-sided bounded rod (l = oo): Find a
solution of the heat-conduction equation for 0<x<°o,t>—oo which satisfies
the condition
boundary conditions,
0 йх<>1, to£t£T.
where <p(x), fii(t) and fi2(t) are continuous functions which satisfy the transi¬
tion conditions
which are necessary for the continuity of u(x, t) in the closed region.
We consider now the x, t phase plane (Figure 37). In our problem we
164 PARABOLIC DIFFERENTIAL EQUATIONS
seek a function u{x, t), which is defined in the interior of a rectangle ABCD.
This region is already determined by the statement of the problem, since
the course of the heat propagation in the rod 0 й x ^ / during the time in¬
terval t й t = T, in which the heat behavior
of the boundary is known, was already in¬
vestigated. Let t0 = 0; we assume that u(x, t)
D C
satisfies the heat-conduction equation only
for 0 < x < l, 0 < / S Г, i.e., not for t = 0
(the side AB) or for x = 0, x = / (the sides
AD and BC). For / = 0, as well as x = 0
В ..x and x = l, the value of this function is given
directly by the initial and boundary condi-
fig. 37. tions. To require that the heat-conduction
equation, for example, be satisfied also for
/ = 0 would imply that the derivative <p" = uxx(x, 0) in this equation exists.
Therefore, the generality of the physical phenomena to be investigated is
limited, and thus the basic functions which do not satisfy this requirement
are eliminated from consideration. The condition (3-1.3) loses its meaning
when it is not required that u(x, t) in the region 0 ^ x ^ l, 0 ^ t ^ T (i.e., in
the closed rectangle ABCD) be continuous or this requirement must be replaced
by another appropriate assumption.46 To understand the significance of this
requirement we consider the function v(x, t) defined by the following condi¬
tions:
v(x, t) = C , 0 < x < l, 0 < t й T
v(x, 0) = cp(x) , 0 й x й /
v(0, t) = цМ) , v(l, t) = [і,Ц) , 0 йТ
where C is an arbitrary constant. The function v obviously satisfies both
condition (3-1.2) and the boundary conditions. However, this function in no
case describes the course of the heat distribution in the rod with an initial
temperature <p(x) Ф C and boundary temperatures ^,(/) Ф C and f*2(t) Ф C, since
it is discontinuous for t = 0, x = 0, x = /.
The continuity of u{x, t) for 0<x<l,0<t<T directly follows in that
u(x, t) satisfies the differential equation. Therefore, the requirement that
u(x, t) be continuous in 0 x й /, 0 ^ / й T, is based essentially only on those
points at which the boundary and the initial values are prescribed. In the
following, by a solution of the equation which satisfies the boundary condi¬
tions, we shall always mean a function which satisfies the requirements (3-1.1),
(3-1.2), and (3-1.3) and hence not repeat these each time, unless there are
special conditions.
Correspondingly, this is the case for other boundary-value problems, in
particular for problems of an infinite rod and problems without initial condi¬
tions.
The solutions of this equation have the following properties which will
be denoted as the principle of the maximum.
A function u(x, t) defined and continuous in the closed region 0 ь / T,
0 x й I and satisfying the heat-conduct ion equation
Ut = a2uxx (3-1.35)
^(Xo,to)^0. (3-1.37)
ot
By comparison of the signs on the left and right sides of (3-1.35) it follows
that both sides can be different. These considerations, however, still do not
prove the correctness of our theorem; since the right and the left sides can
simultaneously equal zero, it would signify no contradiction. We bring forth
this consideration simply to emphasize the fundamental concepts of our proof.
For the completion of the proof we shall seek more than one point (xi, ti)
at which d2u/dx2 ^ 0 and du/dt > 0. Therefore, we consider the auxiliary
function
v(x, t) = u(x, t) + k{to — t) , (3-1.38)
where k is a constant. Obviously then
df dV
= 0, > о
dx dx2 x=x0
are sufficient. If, therefore, at the point xo the function /(x) has a maximum value, then
(a) /'(xo) = 0, and (6) /"(xo) > 0 cannot hold; therefore /"(xo) = 0.
50 Obviously, du/dt = 0, in case to < T, whereas for to = T, then du/dt — 0 must hold.
3-1. SIMPLE PROBLEMS WHICH LEAD TO PARABOLIC DIFFERENTIAL EQUATIONS 167
We now select k > 0 so that kT < e/2, i.e., let k < e/2T; then the maximum
of v(x, t) for 1 = 0 or for x = 0, x = / does not exceed the value M+e/2,
i.e.,
since for this argument the first summand of (3-1.38) is not larger than M,
and the second is not larger than e/2.
Now, v(x, t) is a continuous function. Thus a point (jq, ti) exists at which
it assumes its maximum. Then we have
By comparison of the signs on the right and the left sides in (3-1.35) at the
point (xi, ti) we conclude that Eq. (3-1.35) at the point (xlt ti) cannot be satis¬
fied, since the quantities on the right and left sides have different signs.
Therefore, the first part of our proposition is proved. The statement for the
minimum can be proved analogously, and it is sufficient to apply the first
part to Mi = — u.
then necessarily51
Ui(X, t) = иг(х, t) .
For the proof of this theorem we consider the function
51 Previously this theorem was refined and the continuity requirement at t = 0 was
dropped.
168 PARABOLIC DIFFERENTIAL EQUATIONS
О йх^І, О^ійТ
are continuous, their difference v(x, t) in the same region is continuous. Further,
v(x, t) as the difference of two solutions of the heat-conduction equation for
0<x<l,t>0 is similarly a solution of the heat-conduction equation in that
region. Consequently, the principle of the maximum can also be applied to
this function, and the maximum and the minimum of v(x, t) for t — 0 or x = 0
or x = l is assumed. According to the hypothesis we obtain
v(x, t) = 0 ,
i.e.
Ui{x, t) = u2(x, t) ,
from which the uniqueness of the solution of the first boundary-value problem
follows.
We shall now prove a series of direct conclusions from the principle of
the maximum. In the following discussion We shall refer to “the solution
of the heat-conduction equation,” instead of enumerating the properties of
the function in detail which also satisfy the initial and boundary conditions.
1. If two solutions Ui(x, t) and u2{x, t) of the heat-conduction equation
satisfy the conditions
then
uy{x, t) 5S u2{x, t)
0<x<l, 0<tST
3. If, for two solutions Ui{x, t) and u2(x, t) of the heat conduction equa¬
tion, the inequality
I Ui(x, t) — u2(x, t) I -й e
for all x, t in
Ъйх-йі, ОйійТ
is satisfied.
This assertion results from conclusion (2), when we apply the heat-con¬
duction equation to the solutions
u(x, t) = — e
u{x, t) = Hi(x, t) — u2(x, t)
й(х, t) = e .
Let these be given by functions <p*(x), and /it{t) which differ by less than
e from the functions <p{x), Ці(/), and pi2(t):
However, the function Ui(x, t) according to conclusion (3) differs by less than
e from the function u(x, t):
I u(x, t) — Mi(x, t) I g £ .
\ x\ й L ,
where L is a parameter which is permitted to increase unboundedly, and a
function
V{x, 0) ^ I v(x, 0) I = 0
V(± L, t)^2M^ y(± L, t) .
3-2. THE METHOD OF SEPARATION OF VARIABLES 171
m (x2
+ «** ) = v(x> *) =
L2 v2
+ a2t (3-1.41)
If we now consider any fixed pair of values (x, t) and let L increase un¬
boundedly, we obtain
v(x, 0 = 0,
and thus our theorem is proved.
We turn now to the solution of the first boundary-value problem for the
heat-conduction equation
ut = a2uxx , (3-2.4)
lit — CL Uxx у
If we put (3-2.6) into (3-2.4) and divide both sides of the equation so ob¬
tained by a2XT, then we obtain
±Г X^_
Л , (3-2.7)
a2 T X
where X = const., since the left side of the equation depends only upon t and
the right only on x.
Hence, we obtain the equations
which has already been investigated for the solution of the wave equation
(Section 2-3 § 1), where it was shown that only for the value of the parameters
Xn = (jj , n = 1, 2, 3, • • • (3-2.11)
are particular integrals of Eq. (3-2.4) which satisfy the homogeneous boundary
conditions (3-2.5).
To solve Problem 1, we formally construct the series
The function u{x, t) satisfies the boundary conditions since these are satisfied
by every member of the series. If the initial conditions are to be satisfied
also, then we must have
i.e., Cn are the Fourier coefficients of the function <p{x) when these are de-
3-2. THE METHOD OF SERARATION OF VARIABLES 173
The coefficients Cn of the series (3-2.15) must now have the value defined
by (3-2.17). To show that this series then satisfies all the conditions of
Problem 1, we must prove that the function defined by (3-2.15) is differenti¬
able, satisfies Eq. (3-2.4) in the region 0 < x < l, t > 0, and at the boundary
points of this region (for t = 0, x = 0 and x = l) is continuous.
Because of the linearity of Eq. (3-2.4), a series consisting of particular
integrals, according to the superposition principle, is also a solution if this
series converges and is termwise twice differentiable with respect to x and
once with respect to t (see the auxiliary theorem in Section 2-3 § 3). We
show next that the series arising by termwise differentiation
У dun У d2un
and
»=i dt »=i dx2
In the following we shall formulate the additional conditions which the func¬
tion <p(x) must satisfy. First, let <p(x) be bounded | <p{x) \ < M\ then
from which
dtin
< 2M(-j) for t^t ,
dt
d Un
< 2M(-j) пге~^п/1)2аЧ for t^t .
dx2
In general, we find
2 k+l
dk+lun < 2 M(y) пгк+Іагке-<.пп/1) 2e2t for t si t
dtkdxl
an = Nnqe-Un,l)2a2i (3-2.15')
(n 4- 1 \q o~Стг/i) za2(nz+2n+l) t
&n+1 — lim ( Ц-) e~{*/l) a {2n+l)t = 0 .
lim lim iHJZ-LL-X.- = 1І
lim
П—*oo 71—oo Yfi g-(jT/I) 2a2n27
n
174 PARABOLIC DIFFERENTIAL EQUATIONS
Hence, it follows that the series (3-2.15) for t ^ t > 0 is arbitrarily often term-
wise differentiable. Further, we conclude from the superposition principle
that the function defined by this series satisfies Eq. (3-2.4). Since / is arbi¬
trary it holds for all t > 0. Consequently, the series (3-2.15) for t > 0 represents
a function which is arbitrarily often differentiable and satisfies Eq. (3-2.4).52
If the function <p(x) is continuous, possesses piecewise continuous deriva¬
tives, and satisfies the conditions $o(0) = 0 and <p{l) = 0, then the series
2. Green’s function
= i r^tW^in^r?
» = l|_ l Jo
/
dp[-e^«n/Ms-m™x
J /
= (T4
JоL i
ii r("/1,!*21 sin^x-sin^fly>(f)J?
/ / J
.
Summation and integration can be interchanged for t > 0, since the series
for t > 0 in the brackets converges uniformly with respect to f.54
We now set
With the use of this function G(x, t), u{x, t) can be written in the form
52 In proving that the series (3-2.15) satisfies the equation щ — a2uzx for t > 0, only
the boundedness of the Fourier coefficients Cn was used. This, however, is the case for
bounded <p(x).
53 See Section 2-3 § 3.
54 The series Ian, where an is taken according to formula (3-2.15'), represents a
majorant corresponding to the series, standing for q = 0, in the brackets.
3-2. THE METHOD OF SEPARATION OF VARIABLES 175
cp[+tM$)de = Q , (3-2.20)
Jf-e
since the left side of this equation precisely represents the amount of heat
producing the temperature change of amount ^8(f). The temperature distri¬
bution in this case is given by formula (3-2.19)
Now we let e tend toward 0. From the continuity of G(x, f, t) for t > 0
as well as Eq. (3-2.20) we obtain next by use of the mean-value theorem for
fixed x, t
(3-2.21')
ме(x, t) ^ 0
We have already proved this theorem under the assumption that y>(f)
possesses a piecewise continuous derivative.
We shall now prove the theorem without this assumption. Therefore
we consider a sequence of piecewise continuous differentiable functions <pn(x),
<pn{0) = <pn(l) = 0, which converges uniformly towards <pix), since (pn(x), for
example, can be chosen as the function which represents the step function
which coincides with <p(x) at the points l-k/n, k = 0, 1, 2, • • •, n. The func¬
tions un{x, t) defined by (3-2.19) through <pn(x) then satisfy all the assumptions
of the theorem, since the <pn{x) are piecewise differentiable. The functions
ujx, t) converge uniformly towards a continuous limit function uix, t). To
each e —> 0, therefore, we can find an nie) such that
о
(3-2.5)
II
II
and the initial condition
u(x, 0) = <p(x) (3-2.2)
X
"V L>
хк~д
4J К
V*5 V.
fv
(У
,_<5
1 J
1 1
у - 7
FIG. 38.
° = 2n + 3 AN
so that
Therefore, the passage to the limit under the integral signs is permissible,
and if a function u(x, t) exists which satisfies the assumptions of our theorem,
it can be represented by the formula (3-2.19), from which also the unique de¬
termination of such a function is proved.
We shall now show that formula (3-2.19) represents a bounded solution
of Eq. (3-2.4) which satisfies the conditions (3-2.2) for an arbitrary piecewise
continuous function <p(x) and is continuous everywhere that <p(x) is continuous.
We shall prove this theorem in two steps. First, to show that it is true
in case <p(x) is a linear function
<p(x) = cx , (3-2.2')
, n—1 l
cl-= a— i.e., a = (n — l)c .
n n
Un(X, 0) — фп{х) •
un(x, t) ^ wn+1(x, t) .
The function U0(x) = cx is a continuous solution of the heat-conduction equa¬
tion. It follows from the principle of the maximum that necessarily
Un(x, t) ^ U0(x) ,
since this inequality is valid for x = 0, x = l, and / = 0. The sequence {un(x, f)}
is therefore monotonic, nondecreasing, and is bounded above by the function
U0(x)', consequently, it converges. We can now easily recognize the validity
of the relations
and pass to the limit under the integral sign. On the basis of footnote 52
this function satisfies Eq. (3-2.4) and the homogeneous boundary conditions
(3-2.5) for t > 0. Moreover, it is continuous at t — 0 and 0 ^ x < 1, as we
now prove. Let x0 < l. We choose n such that x0 < 1(1 — l/и). In this case,
(pn(xo) = Uo(x0). If we bear in mind that
and
tion is also bounded since it does not exceed U0(x). Therefore, the theorem
for (p(x) — cx is demonstrated.
If x is replaced by l — x, we see that the theorem holds also for
<p(x) = В + Ax
is valid, since such a function can be obtained by addition of (3-2.2') and
(3-2.2"). Further, it follows that the theorem is also true for an arbitrary
continuous function without the assumption ^(0) = <p(l) — 0. Every function
of this type can therefore be represented in the form
where the sum in the brackets represents a linear function and <p(x) is a
continuous function which vanishes at the ends of the interval: <p(0) = <p(l) = 0.
However, we have already seen that the theorem is true for both summands;
therefore, it is also true for <p(x).
We turn now to the proof of the theorem for an arbitrary piecewise con¬
tinuous function <p{x). In this case, formula (3-2.19) also determines a solution
which satisfies Eq. (3-2.4) and the homogeneous boundary conditions (3-2.5).
Let Xo denote any point of continuity of the function <p(x). To every
positive s, a d{e) can be found such that | u(x, t) — cp(x0) | < e holds when
I x — x0 I < d(e) and t < 8(e). In order to understand this we first note that
because of the continuity of the function <p(x) at the point x0, an ^(e) exists
such that
(a)
1 > Ф)
21
IIV
for 1 ^ - Xo
A
1 * - Xo 1
(b)
'2.
2.
3
ІІЛ
for 1 *0 1
O)
X -
In the interval | x — x01 > y(e), Ф and <p must satisfy only the requirements
(a) and (b), but are otherwise arbitrary. On the basis of the inequality (3-2.23)
we already have
182 PARABOLIC DIFFERENTIAL EQUATIONS
Because of the continuity of ф(х) and <p(x), й(х, t) and u(x, t) are also continuous
at the point x0, i.e., a 5(e) exists such that
u(x, 0) = 0 (3-2.26)
о о
s'
II
(3-2.5)
a
II
The solution u(x, t) of this problem is sought in the form of a Fourier series
in terms of the functions sin (nn/l)x:
3-2. THE METHOD OF SEPARATION OF VARIABLES 183
with
Ш = fie,t)sm?jZdS . (3-2.28)
/Jo /
This equation will be satisfied if all the development coefficients are equal
to 0, i.e., if
uni0) = 0 . (3-2.30)
If we solve the ordinary differential Eq. (3-2.29) with the homogeneous initial
conditions (3-2.30)56 then we find
uix, t)= І IT
*=iLJo
g-(,rn/!>2a2((-r>y(r) I .
Jsln
nn
l
X . (3-2.32)
where
F(f, z) = Cjo/(f, z)
denotes the density of the heat source. The total amount of heat which is
set free in the interval (0, l) during the time of the action of the source (i.e.,
during Az) is then equal to
Sr0+ar r
cpf(Z,z)dtdz. (3-2.35)
Г0 Jfo
If we use the mean-value theorem, we arrive at the expression
fr0-Mrf f0-Mf _ Г)
= \ \ G(x, f, t — z)f($, z)d£dz = G(x, f, t — z)—- = u{x, t)
J ro J fo CP
where
For the case of a continuous distribution of heat sources in the entire region
0 ^ £ й 1,0 ^z S t we obtain
3-2. THE METHOD OF SEPARATION OF VARIABLES 185
which represents the difference between u{x, t) and a known function U(x, t).
We define v(x, t) as the solution of the equation
vt - a2vxx = f(x, t)
f(x, t) =f(x, t) — [Ut — агихх\
with the auxiliary conditions
lit — d i'ixx
u(x, 0) = <p(x)
u(0, t) = u0
u(l, t) = Ui .
The function v(x, t), determined by the initial condition and the homoge¬
neous boundary conditions, is then easily found by separation of variables.
Problems
1. Derive the equation for the heating of a homogeneous thin wire due to
a constant electrical current when heat exchange with the environment takes
place on the surface of the wire.
Solution:
Ut = a2uxx — hu -f q ,
where h and q are fixed constants.
ut = Duxx — wux ,
dz2 ~ ~ ~дГ ’
11. Solve the problem of the cooling of a homogeneous rod whose lateral
surface is heat-insulated; its initial temperature is u{x, 0) = <p(x), and at the
ends a heat exchange occurs with the environment whose temperature is 0.
Consider the special case <p(x) = u0.
12. Solve Problem 11 under the assumption that the temperature of the
environment is equal to u0.
13. Solve Problem 11 under the assumption that on the lateral surface a
temperature exchange occurs with the environment whose temperature (a)
equals 0, (b) is constant and equals .
14. What temperature does a rod assume when one end is heat-insulated
and through the other end passes a heat current changing harmonically with
time, if the heat exchange on the lateral surface is neglected?
15. Solve Problem 11 now assuming that the temperature at one end of the
rod is 0 while the temperature at the second end of the rod changes harmoni¬
cally with time.
16. Let a rod (0, l) be composed of two homogeneous pieces of equal cross
section which touch at the point x = x0 and are characterized by , ki and
a2, k2 respectively.What temperature is assumed in such a rod when the
temperature at one end of the rod, x = 0, tends continuously toward the value
0 while the temperature at the other end changes sinusoidally with time?
17. The left end of the combined rods in Problem 16 has a constant tem¬
perature of 0; the right end, by contrast, has the temperature u(l, t) = uy.
Let the initial temperature of the rod be 0. Determine the temperature of
the rod (taking into consideration only the first term in the series develop¬
ment).
18. Find the temperature u(x, t) of a rod whose initial temperature is equal
to 0 when the boundary conditions have the form
Gt(x, f, t) = 4
t
I
n=l
e Un'l)at sin-if-*, sin ^4 (3-3.1)
X=X~Y' f=f-T-
The Green’s function of an instantaneous point-forming source of intensity
Q = cp which is found at the point f' of the interval (—//2, l\2) then has the
form
2 1 \ ■ КП (. I
Gl(x', t) (3-3.1')
/ »=i l 2 )-Sm—(f +T
We transform the product of both sine functions. If n is even, i.e., n = 2m,
then
2кт ( / . I 2кт (
(. / \ . 2лm /
2ю . 2-кт
sin ——I x + — X • Sin ;—$ .
l V ' 2)
sin
l (f +Y) = sm-
If n is odd, then n = 2m + 1, and
Consequently,
2 -ыпп>2<А nn , xn
+fr« ° ' cos —-—x cos —-—£' (3-3.1")
/ n=1 L l
where 2" ranges over the even and 2; over the odd n.
We shall next determine the limit value of the first sum as /—»oo. It
can be written in the form
4-
I
f " <Гх"а2( sin /U' • sin
n=0
' = —
71
f " fVInW
u=0
(3-3.3)
with
2 2 О-г
j , пк
/І(Л) = g-* “ e sin Ля' sin Л£', ЛЛ =—— and An = —j— .
The sum (3-3.3) suggests the corresponding sums for the integral of the func-
190 PARABOLIC DIFFERENTIAL EQUATIONS
tion /і(Л) over the interval 0 ^ Д < oo. As l—>oo then AX —>0. By passage to
the limit we obtain for the integral60
with
60 In formula (3-3.4) we obtain an improper integral on the left side as the limit value
of the integral sum is extended throughout the infinite interval (0, oo). For the justification
of this passage to the limit, one must show that it does not contradict the usual definition
{>№
exists.
We show that the limit value
C^k
lim \ f(X)dX
oo J 0
exists, independent of the manner in which ^4 approaches infinity. First, we shall con¬
sider the case цк = Xk- Then
and correspondingly,
j. i k-\-h 1
(**) k+hf(X)dX ^ V f(Xi)(Xi+i - до,
JAjt i=k
where /(Xi) and /(Xi) denote the largest and smallest values of /(Д) in the interval (Xi, Ді+і).
From the convergence of the integral sums with respect to the chosen sequence of Xk,
it then follows that for every e an N(e) exists such that for к > N(e),
holds. From this and from the expressions (*) and (**) we obtain
'k + h
f(X)dX < e for к > N(e).
(Next page)
3-3. PROBLEMS FOR THE INFINITE STRAIGHT LINE 191
The integrals for fx and /2 are formed by splitting the interval (0, l) into
equal subintervals of length AX = 2n/I. In this way it can be seen that the
intermediate point is the right end point of the interval in the first sum and
the middle point in the second sum.
Summarizing the results, we obtain
'^k+l !'Ki-
lf{X)dX - \ k+hf(X)dX < e for к > N(e) ,
where Xk and Xk+h are those points of the subdivision that lie nearest to the points
and fik+i.
This proves the existence of the improper integral
f(X)dX .
s;
We show now the existence of a limit value
\ f(X)dX - Zf(X*)AX
Jo i=о
The convergence of this sum is easily shown by using the D’Alembert criterion in which
we investigate the ratio of the fth term of the (г—l)th term
(2г—1) (J\)2azt
g —(i —1)гЛЛ2а 2f
192 PARABOLIC DIFFERENTIAL EQUATIONS
v 4_£ .
- a a t
sin Лх sin Л&/Л f — \ e
1 Г" x*e*< cos Лх COS Л&//1
=-r
JT Jo tt Jo
’•-"•cos X(x-S)dX .
7Г о
Green’s function for the infinite straight line therefore has the form
G (x, £,« = —(
7Г JoVxV‘ cos X(x - $W . (3-3.7)
which depends on both parameters a and ft. To this end we fix a and denote
the integral by /(/3). Here, obviously, in order to calculate the derivative
we must differentiate under the integral sign:
dl
dp -І e~K aX sin xpdX .
dl . 1 _\2«
JL[
2« Jo
c * cos XpdX = ——2^—f(P)
a
Zl = _ jL
I 2a ’
whose solution is
because
l/ n
e z dz
s0 ^
Hence we obtain
І „-(х-()г/4а2и-10) dx
c/>\ G(x, £, t — t0)dx = —(4=
V* j 2Va\t- U)
Q
Г c-\la Q = cp
Vx
since
Г e~*da =
J -
(a =
V 2Va\t
e
- U)
, da
dx
2 Va\t ■ t.) /
Therefore, the amount of heat on the straight line does not change with time.
The function G(x, £, t — t0) depends only on the time through the argument
0 = a2(t — to), and therefore we can write
r— 1 1 ,, C 07*n (3-3.10)
U “ W7 VO
Figure 40 shows the graphic representation of G with respect to x for
different 0 values. Almost the entire area which is bounded by this curve
lies in the interval
(£ — s, £ + s),
where e is an arbitrarily small number when 0 dl{t. t„) is sufficiently small.
The magnitude of this area when multiplied by cp coincides with the amount
194 PARABOLIC DIFFERENTIAL EQUATIONS
G x=$
1 _ 1
2\/к -[/f '
The temperature, at this point in which the heat source is found, therefore
becomes unbounded for sufficiently small Ѳ.
If x Ф i.e., x = t; + h with h Ф 0, then G is represented by the product
of two factors:
GA
lim GWf - 0.
0-0
_1_1_1 1 - \Ѳ~Ь/2
lim -кг/*Ѳ
■=lim
Ѵѳ
= 0 .
e-o [ 2г/тГ j 2V 7Г 0-0 (/г2/40Ѵ2/4й
or
(3'ЗЛ4,)
=27гГ.77Ге"“'',Ѵ“'Ѵ(?№
represents a bounded solution of the heat-conduction equation which for t = 0
at all points of continuity of this function u(x, t) is continuously connected
with <p(x).
Therefore, we shall prove the following lemma (generalized superposition
principle). If the function U(x, t, a) for each fixed value of the parameter a,
and with respect to the variables x and t satisfies a linear differential equation
L(U) — 0,
then
62 See V. I. Smirnov, Textbook of Higher Mathematics, 2d. ed. Part II, Ch. Ill, §3,
no. 80, Berlin, 1958.
3-3. PROBLEMS FOR THE INFINITE STRAIGHT LINE 197
in which a and b once more are finite, similarly can be differentiated under the
integral sign if fix, a) satisfies the above stated conditions and tpia) is a
bounded (or absolutely integrable) function. If the limits of integration are
infinite then in this case we must require the uniform convergence of those
integrals which are obtained by differentiation of the integrands with respect
to the parameter. This observation holds also for multiple integrals which
depend on parameters.
For linear differential equations Liu) = 0 the superposition principle holds;
that is, the function
is also a solution of Liu) = 0. The proven lemma, similar to the one above,
gives the conditions under which the limit value of the sum (3-3.15), i.e., here
If” 1 -U-e)2/4a2<
uix, t) I < M—7=1 -y—
2ук J-oo Va t
= M-4=\
1
e~a2da = M , a
6—x
Vic Waft ’
since
e~a da =у/яГ .
Further, we shall prove that the integral in (3-3.14') for t > 0 satisfies
the heat-conduction equation. Therefore, it is sufficient to show that the
derivatives of this integral for t > 0 can be calculated by differentiation under
the integral sign.
In the case of finite integration limits, this is certainly permissible since
the derivatives of the function
198 PARABOLIC DIFFERENTIAL EQUATIONS
1 g-(x-f)2/4a2t
2V7 n a2t
are proven when we have shown that the integral standing on the right side
converges uniformly; in particular, for the proof of the differentiability at
the point (x0, to) it is sufficient to show the uniform convergence of the integral
in a fixed region
ti й t0 й U , I x — x01 й x
of that point.
Sufficient for the uniform convergence of an integral (an analogous crite¬
rion holds for the uniform convergence of a series) is the existence of a
positive function F(£) which does not depend on the parameters x and t and
which majorizes the function
such that
•- 1 1 g — JCo I + X.
~F(€)de = g-(|{-l0l-J,2/«“2<2
d$
x, 2Vк 2[аги}ъ/г
“ 1 _ + 2x _
*j-7 2\/я 2[аН$/ів
£l = I £ — X0 I — X,
3-3. PROBLEMS FOR THE INFINITE STRAIGHT LINE 199
and this integral converges, since a factor of the form (a£ + b)e cj2 appears
under the integral sign. Therefore
ax
= Г
J-oo ox
• (3-3.17)
By a completely analogous method, we can also prove that all the above
derivatives can be calculated by differentiation under the integral sign.
Therefore, we have proved that the function (3-3.14') satisfies the heat-con¬
duction equation.
We turn now to an important property of the integral (3-3.14), and indeed
we shall show that at all the points of continuity x0 of the function (p(x),
the following limit
is valid.
Let <p(x) be continuous at x0. Then we have to show that
holds when
On the basis of the assumed continuity of <p(x) at x0, an rj(e) exists with
for
I X — Xo I < У] •
We now split up the integration interval and represent u(x, t) as the sum of
three summands:
-и-еі2/4Л
<p(WZ + l_p..*+l {-...*
U[X,t) 2v/*L~-l/^ 2і/л- Jxj 2l/TZ Jx2
= иг(х, t) + uz(x, t) + u3(x, t), (3-3.19)
where
Xi = x0 — v and x2 — x0 + v • (3-3.20)
(pjx oj_r2
u2(x, t)
2i/?r )XlVaH
g — x
da = (3-3.21)
2\/a2t 2\/a2t
Now if I x — x01 < y, then the upper limit is positive and the lower, negative,
and as / —»0 the upper limit tends towards + oo and the lower limit toward
— oo. It then follows, however, that
lim Ii = <p(x0).
t—0
when
We shall now prove that the remaining integrals, that is, I2, их and щ, can
be made arbitrarily small. First we estimate I2 as follows:
ЛІ ^ 1 Г2 i_ g-lx-t)2/*a2t j ^ _ ^o) I #
2р//7Г Jxj
VaH
From (3-3.20) it is seen that for
Xi < f < x2
the inequality
I f - x01 < r)
1
e~*2da < e~a2da = 1 ,
г/я
1 1===e U |l/4a Af
«з(х, t) I = < e a2da—*0
2у/ ,VaH і/тг
Uj—x)/2 v^a2«
U{(X, t) I = e a da—* 0
2і/ 7Г
As *—►*<>, then дг2 — x > 0 and x, — x < 0. As / —>0, the lower limit in the
last member of (3-3.24) tends toward +oo and the upper limit in (3-3.25)
tends toward — oo. Consequently, there exists a <S2 such that
holds when
With the use of the estimates (3-3.24) and (3-3.25) obtained above, we find
<T + T + T + T = £’ p-3-27*
provided
1 p 1 е~и-()г/* a*lt-t0)
u{x, t) <p(Z)dt . (3-3.14")
2i/?Tj - Va\t - to)
The uniqueness of this solution for a continuous function <p(x) results from
the theorem proven in 3-2§3. If the initial function <p(x) possesses a finite
number of points of discontinuity, then (3-3.14') represents a bounded solution
of Eq. (3-3.1), which is continuous everywhere with the exception of the
points of discontinuity of <p(x).63
We shall consider as an example the following problem: Find a solution
of the heat-conduction equation when the initial temperatures (at t = t0 — 0)
for x > 0 and x < 0 are constant but have different values, i.e.,
63 With the aid of the method described in 3-3 §3, we know that the function u(x, t)
is uniquely defined by the enumerated properties.
202 PARABOLIC DIFFERENTIAL EQUATIONS
r° x-()‘/4a2t_d{__ _T\_f _
-U-f>‘/4a2*
Тг
i/2Va 2t
Г Tx
e~a da + Ц"
l/ff J- і/яЛ J— x/2^a%t
7\ + Г2 , Тг-ТгГ^
+ і/ 7Г
e a da (3-3.28)
since
—7=1
1 f-
e da =—7=4 e
1 f° -«S 1 fl-»2J 1 1 f*-.2,
da-7=1 e da— — — —17=1Г e da
V n J_» V К )-*, V К Jo 2 Vx Jo
1/7Г Ji
and
* .2
x
e da, 2 =
г/jTj-* 2VaH *
In particular,
2 =■
M(*’/} " 2 (X + 1/* Le " ^) ’ 2i/a2/
holds for
T2 = 0 , Ti = 1.
1 1 r* 2
/w-2+tt
2 f1 _e*
Ф(г) - —7=1 e da ,
vn Jo
the so-called error integral, occurs often and has been tabulated many times
in probability calculations.64
For arbitrary Тг and T2, formula (3-3.28) can also be written in the form
Here we recognize that the temperature at the point x = 0 during the entire
time is constant, and therefore is equal to half of the sum from the right
and left sides of the initial value, since Ф(0) = 0.
Finally, the solution of the inhomogeneous equation
u(x, 0) = 0
obviously can be expressed by the formula
This results from the meaning of the Green’s function G(x, f, t) (see Section
3-2 §4). At this point, however, we shall not investigate this formula and
the conditions imposed on f(x,t) for its applicability.
As we have already noted in 3-1 §4, when we are interested in the dis¬
tribution of heat in the neighborhood of one end of a rod, while the influence
at the other end of the rod is unessential, we make the assumption the latter
end lies at infinity. This then leads to the problem of determining those
solutions of the heat-conduction equation
ut — a2uxx (3-3.11)
on the semi-infinite line x > 0 for t > 0 which satisfies the initial condition
where M is a constant. From this it follows that the initial function <p(x)
must also satisfy the condition
I <p(x) I < M .
= (3'3-31»
u(x, t) =
1 1
■ /-^\ —7==- e--U-f)2/4a2(
2\/k J_oo з/аЧ
Фта
vanishes for x — 0:
u{0, t) = 0 .
Here, it must be assumed that the integral defining u(x, t) converges; this is
the case when ф(х) is bounded. The integrand in
1 _f~ 1
u(0, t) -- е-иг,іаг"ф{!;Ш
2-[/л j_oo i/ a21
consequently
u(0, t) = 0 ,
which proves our lemma.
2. If ф(х) is an even function, i.e.,
the derivative of the function u(x, t) from (3-3.31) for x = 0 and all t > 0 is
equal to 0:
3-3. PROBLEMS FOR THE INFINITE STRAIGHT LINE 205
-f^(0, /) = 0 .
That is,
1 f+-U-g) g-U-f)2/4a2t
= 0,
2\/n J_»2(a2t)3/2 x=0
This function can be defined by use of the first lemma and an initial func¬
tion ¥ (x), which for x > 0 coincides with <p(x) and for x < 0 represents the
odd continuation of <p(x), i.e.,
where the substitution $'= — f is introduced in the first integral and the
relation
¥(£) = -&-£) = -?(£')
,-U-{)2/4a2t _ -U + f)2/4a2«
{e }<p($)d$ (3-3.32)
Ui(x’t} ~ гѵАо i/m
1 f~ 1_
Мі(дг, /) (3-3.32')
2г/^Г]о 1/аН
are fulfilled.
Since the initial condition is prescribed not for f = 0 but for t = t0, in
place of (3-3.32) we obtain
If we now split up the integral into two summands and introduce the variables
$ — X £+X
a = ai =
2Va\t-U) ’ 2Va\t - to)
we obtain
or
fi(t) = Цо — const.
The function
К 2T/4-J (M-34)
is the sought function, since it satisfies the same equation and the conditions
where
then with the help of the function /n(t) the solution of the second boundary-
value problem can be represented in the form
/ V dU(x,t — z)
U{X, t) = I Щ-K-r-- Az + fin-iU(x, t — tn-i) with ti й т< ^ ti+i.
t=0 ot
(3-3.38)
j -^rU, t — z)fi{z)dt ,
Obviously, the sought solution u2(x, t) of the second problem must satisfy
ax ?—*2/4a2t
3/2
2i/V [a2t]
1
■2a'^-(x, 0,0 = 2 G = -Д=
dx d£ f=0 2f/ я т/azt
Consequently the sought solution in the case of an arbitrary function fi(t)
can be represented in the form65
a2 f *_x_
u2(x, t) = g-x2/4a2( t-r)^T)dz
2т/гг Jt0 [a\t — r)]3/2
or
65 This representation of the solution of the first boundary-value problem with homo¬
geneous initial conditions was mentioned here so that it can be compared to an alterna¬
tive solution in Chapter 5, Section 4.
3-3. PROBLEMS FOR THE INFINITE STRAIGHT LINE 209
used. Further, we never used the analytical expression of the function U(x, t)
but only the property that it satisfies the boundary and initial conditions
for t > 0
for t < 0
u(x, t) -
s ‘dU_
0 dt
(x, t — т)ц(т)сіт ,
U(0, o=i.
The principle formulated here, the so-called Duhamel principle, shows
that the constant boundary value causes the principal difficulty in the solution
of boundary-value problems. If a boundary-value problem for a constant
boundary value has already been solved, then one can immediately determine
the solution of the same problem with variable boundary conditions by use
of formula (3-3.41). This principle is often used for the solution of several
boundary-value problems in which the solution is determined only for a con¬
stant boundary value, and it is assumed that the solution for a variable
boundary condition fi(t) then can be represented by formula (3-3.41).
The sum
Ui(X, t) + u2(x, t)
gives the solution of the first boundary-value problem for the semi-infinite
line in the case of the homogeneous equation.
By the use of formula (3-3.29) and the method of odd continuation, it
follows easily that the solution of the inhomogeneous equation
when the initial and the boundary conditions are homogeneous (u(0, t) = 0),
can be represented by the formula
The sum
210 PARABOLIC DIFFERENTIAL EQUATIONS
I u(x, t) \ < M
I MO I < M.
It will be shown later that the function u(x, t) is uniquely defined. Often
we deal with the boundary condition
Fourier has treated this problem, and it was applied to determine the tem¬
perature variation of the earth.66
We write the boundary condition in the form
MO = Аеш . (3-4.2')
From the linearity of the heat-conduction equation it follows that the real
and imaginary parts of a complex solution of the heat-conduction equation
separately satisfy this equation.
If one also finds a solution of the heat-conduction equation which satisfies
the condition (3-4.2'), then its real part satisfies condition (3-4.2) and its imagi¬
nary part satisfies the condition
Ut — d tiXx у
Awt (3-4.3)
u{0, t) = Ae'
a = —t /3 , /3 = iw ,
a
from which there results
a —
CO (1 + i)
o>2 л/ 2
Hence we have
u(x, t) — Ae±у,(ш/2а2)д:+<(±',(ш/2“2))с+“() (3-4.5)
(о
u(x, t) — Ле~Ѵ(ш/2а2)Іcos cot (3-4.7)
2 a2
’• = /¥=/ik(1+i)
212 PARABOLIC DIFFERENTIAL EQUATIONS
where Xi and Хг are the real and imaginary parts of X{x). Therefore, for
й(х, t) we obtain the representation
By separating out the real parts of й(х, t), we finally find the solution of the
original problem without initial conditions in the form
In fact, we have
£ X i ^ X
«1 - ——7===- and a2 = ——. =- .
2l/aHt -t0) 2i/a2(t - t0)
From this (3-4.16) follows, since as t0-*co both x and t remain fixed. If in
(3-4.15) the variables * and t remain fixed and t„ tends towards — oo, then
u(x, t) is equal to the limit value of the first summand. We therefore obtain
_x_
u(x, t) = _«!_[ (x2/4a2(f—r)
p(x)dx, (3-4.17)
2V7) [a2(t-r)f2
which shows that there cannot be two different solutions of our problem.
We can also prove that formula (3-4.17), for every bounded piecewise con¬
tinuous function represents the solution of the corresponding problems.
We can proceed analogously for problems without initial conditions for a
finite interval 0 ^ x ■й /. Here if the boundedness of the solution is not re¬
quired these problems would not be uniquely solvable, since the function
Problems
1. Determine the Green’s function for (a) a one-sided bounded rod for bound¬
ary conditions of the first and second type when no heat exchange results
on the lateral surface; (b) a two-sided unbounded rod when there is heat
exchange on the lateral surface; (c) a one-sided bounded rod in the presence
of a heat exchange on the lateral surface and with boundary conditions of
the first or second type.
2. Find the Green’s function for a one-sided bounded rod with heat insulated
lateral surface for the third boundary value problem, where the boundary
conditions have the form
-f--hu(0,t) = f(t).
dx
Solution:
_1_ _|_ g-(x+()2/ta2lt-T)
G(x, £,t — t)
2Vruil{t
g-hz-(a+t+z)2/ia2 (
}•
3. Solve the heat-conduction equation for (a), (b), and (c) of Problem 1 when
(1) a heat source Q = Q(t) develops at the point x = £0 (Q = Q0 = const.); (2)
an initial temperature distribution u(x, 0) = <p(x) is prescribed, where
for 0 < x < l
outside of the interval (0,/) ;
214 PARABOLIC DIFFERENTIAL EQUATIONS
(3) the heat sources are distributed throughout the entire rod with the density
f(x, t) while the initial temperature is equal to zero, where f(x, t) = q0 = const,
(stationary sources).
4. A one-sided bounded rod with heat-insulated lateral surface is uniformly
heated up to a temperature
«(0, 0 = 0, t > 0.
Determine the temperature u(x, t) of the rod and use the tables for the error
integral
2 f1 2
<P(z) = —7=1 e a da
V к Jo
0 йх ^ l for
16a2 2 a2
exists; (b) a heat flow Q{t) = В sin wt occurs; (c) Newtonian heat exchange
with the environment exists, whose temperature changes according to the
law v(t) = C sin cot.
10. By using the method of reflections construct the Green’s function for a
two-sided bounded rod with heat-insulated lateral surfaces for boundary con¬
ditions of the first and second types.
11. Let a two-sided unbounded rod be comprised of two homogeneous rods
which are brought together at a point x = 0 and are characterized by ax, kx
and a2, kz respectively. Let the initial temperature be
for x < 0
u(x, 0) = <p(x) = j^1
for x > 0 .
Determine the temperature u(x, t) in the rod when the lateral surface is heat-
insulated.
1. Temperature waves
ato
*-4.
ІІЛ
(3-5.1)
H
J3
8
II
1
Cu
On the basis of this solution the temperature waves in the earth can be
described in the following manner: When the temperature of the surface
changes periodically over a long period of time, then a temperature fluctuation
with the same period is developed in the earth. Therefore:
1. The amplitude of the distribution decreases exponentially with the
depth: A(x) = Ae~v^iu2^, i.e., an arithmetic increase in the depth corre¬
sponds to a geometric decrease of the amplitude (Fourier’s first law).
2. The temperature distribution in the earth takes place with a phase
displacement. The time <5 between the occurrence of the temperature maximum
(minimum) in the earth and the corresponding time point on the earth surface
is proportional to the depth:
5=/ 2^X
(Fourier’s second law).
3. The depth of penetration of the temperature in the earth depends on
the amplitude of the temperature distribution on the surface. The relative
change of the temperature amplitude is equal to
A{x) _ _ ''(ш/2аг) *
A
This formula shows that the depth of penetration of the temperature is
smaller when the period is smaller. For two temperature distributions with
periods Ti or T2, the corresponding depths xt and x2 in which the relative
temperature changes coincide, are connected by the relation
(Fourier’s third law). For example, it shows the comparison between the
daily and the yearly variations (T2 = 3657\):
x2 = і/365жі = 19.1 Xi
that is, the depth of penetration of the yearly distributions with equal am¬
plitude on the surface is 19.1 times as great as the depth of penetration of
the daily distribution.
For example we shall give the results of observations of the yearly tem¬
perature distributions of the station Gosch (Amur):68
depth has diminished to 13.3 per cent of its value on the earth’s surface,
which is equal to 19.5 degrees.
On the basis of these values one can determine the thermal conductivity
of the earth:
1n Mx) ПГ .» _ (ox2
A ~ V2tf ’ 2 Inг(А(х)1А) ■
From this we find the thermal conductivity,
а2 = 4-l(T3 —.
sec
КП
= Ane~'/unlTai)x cos x 2m-t + A
Taz T
or
Val{xi) + blix2)
This formula shows the following: If one measures the temperature
changes at any two points Xi and хг during an entire period, then by deter-
69 Special Physical Practices, vol. I, Prob. 35, Moscow, 1945; V. I. Iveronova, Physical
Practice, Prob. 23, p. 117, Berlin, 1957.
218 PARABOLIC DIFFERENTIAL EQUATIONS
For the estimation of the state of temperature in the earth’s interior some
hint can be obtained from observations on the earth’s surface. The impor¬
tance of this lies in the fact that the daily and the yearly temperature varia¬
tions take place only in a relatively thin layer of the surface (approximately
10—20 m for the yearly variations), while the temperature below this layer
changes very slowly in the course of time.
It has been observed in ravines and caves which lie 2—3 km beneath the
earth’s surface that the temperature increases with increasing depth on the
average of 3°C per 100 m.
The first test, carried out at the end of the 1800’s, to give a theoretical
explanation for the observed geometric gradient, met with insurmountable
difficulties.70 It was concluded that earlier the earth must have been radiat¬
ing heat and that it cooled gradually. The initial temperature characteriz¬
ing this cooling process must have been of the order of magnitude of T0 —
1200°C (the melting temperature of rock); the surface temperature is of the
order of magnitude of 0°. Further, the surface temperature may not have
changed essentially (not more than 100°) from the period when vegetable or
animal organisms first occurred on the earth. The cooling process, considered
in this sense as purely quantitative, leads then to the solution of the heat-
conduction equation
du _ 2 d2u
dt dz
in the half space 0 < z < со with the initial and boundary conditions
u(z, 0) = T0
u(0, t) — 0 .
The solution of this problem was treated in Section 3-3; we obtained it
in the form
2[г'г
u(z, t) — Tf da .
°і/іГ)о
The gradient of this function at z — 0 is given by
du_T0 ^-(z2/4a2t) Ta _
dz z=o л/л Vazt z=o Vл i/a2t ’
If we substitute in this expression the value of the geothermal gradient у =
3 • КГ4 degrees/cm, T0 — 1200°C, and the value a* = 0.006 cm2/sec (the average
value of the experimentally found coefficients of the thermal conductivity
of granite and basalt), then for the time duration of the cooling process we
obtain the value / = 0.85 • 1015 sec = 27 million years. Such an estimate of
the age of the earth is, however, incompatible with the geological facts. The
approximate nature of the theory being considered—that is, disregarding the
earth’s curvature, the variability of the thermal conductivity, and the in¬
accurate value of To—naturally cannot strongly influence the order of mag¬
nitude of the value found for the age of the earth, which according to modern
investigation has been estimated at around 2 • 109 years.
The physical scheme of the course of the temperature in the earth can
be represented in a different manner, using radioactive decay. The radioactive
elements distributed in the earth’s crust generate by their decay a certain
amount of heat which naturally contributes to the warming of the earth.
Therefore, the corresponding heat-conduction equation must have the form
du 2 d2u r
~dt a +
where A represents the volume density of the heat source. On the basis of
numerous measurements of the radioactivity of rocks and the amount of heat
generated by their radioactive decay we usually choose the value
A = 1.3 • 10"12—^— .
cm sec
Q = -|-7ГІ?3Л .
220 PARABOLIC DIFFERENTIAL EQUATIONS
where R = 6. 3 • 103 km is the earth’s radius and k = 0.004 is the average heat
conductivity of the soil rocks.
Accordingly, under the assumption that the distribution of the radioactive
elements are constant and the earth is not warmed by the radioactive decay,
the calculated geothermal coefficient exceeds the observed value
у = 3 • 1СГ4 degree/cm
by two orders of magnitude.
We now drop the hypothesis that the radioactive elements are distributed
uniformly and assume instead that they lie in a layer of thickness H of the
earth’s crust. Without considering the curvature of the earth we then obtain
for the determination of the stationary temperature the equation
d2u _ A
for 0 ^ г ^ H
dz2 k
= 0 for z > H
u(0) = 0 , = 0.
The solution of this problem is obviously given by
0 йг^Н
-у) ■
_ A H2
z^H
~ k 2 '
du _ AH
dz г=о k ’
у = 3 • 1СГ4 degree/cm ,
3-5. APPLICATIONS TO CHAPTER 3 221
then we find
H— « 106 cm = 10 km .
A
We shall now estimate how the assumption that the temperature is station¬
ary affects the size of the geothermal coefficient. To this end we consider
the solution of the heat-conduction equation
dw
+f
~dt
f=— , 0 йгйН
cp
= 0, z > H
w(z, 0) = 0
w(0, /) = 0 .
In this case G is the Green’s function for the one-sided bounded straight line:
For the value of the gradient of the function w(z, t) for z = 0, by considera¬
tion of the value of / there results therefore
dw
e-ii/^t-T)d^dz
dz Cp2\/ к Jo Jol/[tf2(/ — t)]
A ft 1 ГЙ2/ 4a2(t-r)
——7=1 —. e-«dadz
cpv л JoVa\t — z) Jo
[1 - e~a'l"4\de , 0 = t —z.
cpVn Jol/«2^
Therefore
dw A Г 2 Vt H f „-СТ2 da \
Л \ ^ 2 f f
dz z=0 cp\/n 4 « a )o0 0 J
where
a=
H
(T0 —
я da a2 dd
2yVtf ’ 2\/aH ’ a2 ~~ H VaM
Now
222 PARABOLIC DIFFERENTIAL EQUATIONS
о
e^'da e a da ,
Oo ao
so that
dw
—Y-t [i - <r"2/<‘2'] + нЛЛШ _e~^da\ . (3-5.4)
dz cpa l. \/n V К Jff/Fa2! J
Further,
dw
lim-
(dz
then cpa2 = k, the limit value of the first summand in the curved brackets
vanishes and the limit value of the second summand is equal to H.
We now calculate the difference between the limit value of dw/dz obtained
above and the value of dw/dz for
H 10e 1
<T0 : 0.025.
2УаЧ 2г/6 • ІО-3 • 6 • ІО16 2 • 19
i.e., dw/dz |z=0 differs by about 4 per cent from its limit value as / —»oo.
We can calculate the function w{z, t) for z > 0 and prove that for z Ш H,
when one introduces the age of the earth for t, the function still deviates
considerably from its limit value71 (although, as we see at the surface, the
gradient is practically equal to its limit value).
These considerations are all only estimates; nevertheless we must conclude
from the fact that the rate of the radioactive decay does not change with
the temperatures and pressures obtainable, that the concentration of the radio¬
active elements with increasing depth decreases rapidly when the value of
A for the upper layers of the earth’s crust, as it has been determined by
numerous measurements, is taken as a basis. A physical theory which would
allow us to derive a law for the decrease of the concentration of radioactive
elements with increasing depth has not yet been developed.
The method of analogy has been shown to be very useful for the solu¬
tion of a series of problems of heat conduction. As an example, therefore,
we shall consider the following two problems.
1. The Green's function for the infinite straight line. The heat-conduction
equation obviously remains unchanged by the substitution
That is, when the linear scale of the rod changes &-fold, and the time scale
changes &2-fold.
We seek next the solution of the heat-conduction equation
ut — a2uxx (3-5.6)
for x > 0
u( x, (3-5.7)
for x < 0 .
Also, the initial condition (3-5.7) is not changed by the above scale changes.
Therefore, for the function u(x, t), the equation
(3-5.9)
then we obtain
x
(3-5.11)
2 yT •
d2u _ d2 f 1 du _ x ■ Up d f _ _ z_ d f
dx2 U° dz2 4/ ’ dt 4ti/2 dz U° 21 dz
If we introduce this expression into Eq. (3-5.6) and divide it by u0l4/, then it
follows that
>d2f (3-5.12)
— 2
dz2 dz
with the auxiliary conditions
which correspond to the initial condition for the function u(x, t).
By integration of Eq. (3-5.12) we obtain
Ct „ fz/a
a2^r=-2z, f = Ce-^2 , f=c\ e~i la2d£ = Ci \ e--?2JC-
J —oo J —oo
224 PARABOLIC DIFFERENTIAL EQUATIONS
Therefore, the lower limit is so chosen that the first condition of (3-5.13) is
fulfilled. The second will similarly be fulfilled if we set
Therefore
Ф(г) = -%=[
VK Jo
is the error integral. If the initial value has the general form
uо for X > X
u{x, 0) (3-5.15)
0 for X < X
then
x — X
u(x, t) = у j^l + ф(- (3-5.16)
2\/a4 )]
We turn now to the solution of the second auxiliary problem for which
the initial conditions have the form
In this case
u(x, t) = Y
Q — cp{x2 — Xi)u0.
If Q = cp, then
u(x, t) =- (3-5.18)
x2 — Xi
The Green’s function then obviously represents the limit value of the func¬
tion u(x, t) as (хг — ДГі)—>0 in the passage to a point source.
The corresponding passage to the limit in formula (3-5.18) gives
u(x, t) — — (3-5.19)
since on the right side of (3-5.18) the difference quotient arises whose limit
value is the derivative in (3-5.19).
Now if we perform the differentiation we find
3-5. APPLICATIONS TO CHAPTER 3 225
и^‘)=-^77-Ѵше~{-ч"''л • (3-5'20)
д Г,. . ди 1 du
ITJ = cpir (3-5.21)
u(0, t) = Mi (3-5.22)
u{x, 0) = m2 . (3-5.23)
Also in this case a substitution of the form (3-5.5) changes neither the Eq.
(3-5.21) nor the auxiliary conditions (3-5.22) and (3-5.23).
From this it follows that
(3-5.24)
"fe') = /(w) = /(2)’ 2 = w-
By using this expression we obtain for / the equation
(3-5.25)
3 Гп I i ди 1 du_
1*11 + ІІГ (3-5.27)
Now if we set
x
u(x, t) = f(z) ,
2ѴГ '
FIG. 42.
must also be fulfilled where kx and k2 are the heat conductivity coefficients
of the first and the second states while X is the heat of fusion.
If now in this expression we let Jt tend to zero, then the auxiliary con¬
dition on the transition surface takes the form
(3-5.31)
This condition is valid for the solidification process (for > 0 and d$/dt > 0)
and also for the melting process (for < 0 and dtldt < 0); the direction in
which the process proceeds can be determined from the signs on the left side.
Now we shall consider the freezing process of water where the melting
temperature is zero. Let the half space x ^ 0 be filled with water. This
amount of water then is bounded on one side by the plane x = 0. At the
initial moment t = 0, the water has the constant temperature c > 0. If the
surface x = 0 steadily maintains the constant temperature ci < 0, then the
transition surface x — f penetrates, with time, into the fluid.
According to the above, the problem of the temperature distribution in
the freezing water and the determination of the velocity of propagation of
the transition surface leads to the equations
dui d2ul
for 0 < x < £
~df dx2
(3.5.32)
ди 2 32m2
for £ < X < OO
dt dx2
with the auxiliary conditions
Ml = Ci for X = 0
(3-5.33)
u2 — c for t = 0 .
Mi = m2 = 0 for x = £ (3-5.34)
Зм i , Зм 2 dt,
ki (3-5.31)
dx i*={ dx ^if
must be fulfilled where ki, a\ or k2, al are the coefficients of the heat and
thermal conductivity of the solid and liquid states.
For the solution of the problem we make the Ansatz
Mi = Ai + ВіФ\
2міі/1
— л‘+
where At, , A2, and B2 are constants to be determined while Ф is the error
integral
2
228 PARABOLIC DIFFERENTIAL EQUATIONS
If we consider the conditions (3-5.33) and (3-5.34) then we obtain from (3-5.33)
Ai = Ci, A% + B2 — c
and from (3-5.34)
А' + В'І2^7Г) = °
A° + В’Ч^7Г) = 0 •
These conditions must hold for arbitrary values of t. This is possible only
when
f = aVT (3-5.35)
Al Cl ’ Bl ФМЩ
(3-5.36)
л _ сФ(а/2а2) B _ c
1 — Ф(а/2а2) ’ 2 '1— Ф(а/2а2) ’
For the determination of a the requirement (3-5.31) leads to:
^1clg~0,2/4a ■ k2ce~a2/*a 1 _ .
(3-5.37)
ахФ{а!2аі) a2[l — Ф(а/2а2)] ^ 2
Ci
A — B2 — 0 ,
2 A i — Ci f Bi (3-5.36')
Ф{а/2аі)
and
1 е~Р2
where D is determined by
73 For an asymptotic representation of the function 1 — Ф(г) as z -> oo see the Ap¬
pendix.
3-5. APPLICATIONS TO CHAPTER 3 229
D= < 0 .
kiCi
One can easily prove that all the conditions of the problem remain un¬
changed when the length scale changes &-fold and the time scale &2-fold.
Therefore the solution depends only on the argument xj\/T, i.e.,
u(x-t)=fwij-
Hence, in particular, it follows that the motion of the zero isotherm can be
described by the equation
f = aq/1 ,
when a is the value of the argument, for which
f(a) = 0
is valid.
For the determination of the of the function f the following conditions
result:
dfj_
a i d2fi 2z for 0 < z < a
dz2 dz
-«d'A dft
^2 2
i
2z for a < z < oo
dz2 dz
/i(0) = Ci , /2(00 )-c , fi(a) = /2(a) = 0
Therefore,
f(z) =
Ш
fi{z) — A i + ВіФ\
for0
< z < a
For the determination of the constants Al, Bi, Az, Д,, the requirements
(3-5.33) and (3-5.34) must again be used, from which (3-5.36) results. For the
determination of a the condition (3-5.37) is used. Therefore the respective
analytical parts of both methods of solution coincide.
These considerations show that the solidification problem can also be
solved in the cases in which the latent heat is freed not for a fixed tempe¬
rature but in a certain temperature interval. In a similar manner the solution
is possible when not one but many critical temperatures are given. Such
cases can occur for changes of state in the transition from one crystalline
structure to another (for example, for the recrystallization of steel).
W(M, t; Mo, t0) = I W(M, t\ P, Ѳ) W(P, Ѳ\ Mo, t0)dVP , U < Ѳ < t. (3-5.41)
We shall now show that, under certain conditions which the function
W(M, t\ Mo, to) must satisfy, the solution of the Einstein-Kolmogoroff equation
satisfies a parabolic differential equation. Therefore, we shall consider the
case in which the position of the point M is described by a single coordinate
x and assume that the function W{x,t; x0, t0) satisfies the following conditions:
(a) lim —-— = lim — [(# — ?) W{x, t + r; £, t)df = A(x, t). (3-5.42)
r-0 г T J
If the particles during the time z are displaced from the position £ to
the position x, then (x — £/r) is the average velocity of the particles. Con¬
sequently, requirement (a) means that the ordered motion of the particles
takes place with a finite velocity.
(b) lim (x ~ ^)Z- = lim — ( (x - 6)2 W(x, t + r; £, t)d£ = 2B(x, t). (3-5.43)
r-o r : ]
The quantity (x — f)2 does not depend on the direction of the displace¬
ment of the point x with respect to the point £. The average value of the
square of the distance during the time r,
\[ф"'(х) \ <A,
and take into consideration
we obtain
1
'"(£*)(* - ?)3 W(Z, t; Xo, /0) W(X, t + r,$, tmdx
T
^ — fI * — f |3 W(x, t + r; $, t)dx = AAl—£11 .
T J Г
d(AW) d\BW)
dx dx2 ]
dx — 0 .
Since this relation must hold for an arbitrary function ф(х), the Einstein-
Kolmogoroff differential equation
dW d(AW) d\BW)
dt dx dx2
where
a = — A + Bx
P = — Ax + Bxx = ax.
From Eq. (3-5.46) we recognize that the quantity В has the physical signi¬
ficance of a diffusion coefficient. If the process considered is homogeneous
in time and space, i.e., the function W depends only on the difference f =
x — x0 and Ѳ — t — t0, then the coefficients A and В are not dependent on x
and t and are constant. Eq. (3-5.45) is then a differential equation with con¬
stant coefficients:
dW л dW , „ d2W
—T — — /І-г -О -5“ . (3-5.47)
dt dx dx
If W depends only on \x — f|, i.e., the probability for a right-and left-sided
displacement at the same distance from the point f, coincide, then obviously
A must be equal to zero. Analytically this follows from the formula (3-5.42),
since the integrand is an odd function.
In this case, Eq. (3-5.45) is transformed into the simple heat-conduction
equation
dW „d2W
(3-5.48)
dt dx2 '
6. The d function
This obviously represents the potential of the masses which are distributed
with the density pn. For n—>°o we obtain
234 PARABOLIC DIFFERENTIAL EQUATIONS
These results obviously do not depend on the choice of the sequence {,»„},
and although the sequence {un} now approaches 1/r, the sequence {p*} pos¬
sesses no limit value in the class of the considered piecewise continuous func¬
tions. The “limit form’’ corresponding to the sequence {/>„} is called the
Dirac <5 function and is designated by 5(M, M0).
The basic property which defines the 8 function is the formal relation
SSS/<M-M,/WWr"={'o(W £ m. IT <M-51)
where f(M) is an arbitrary continuous function of the point M. From the
fact that the functions pn, for n—> oo in each region which does not contain
the point M0, converge uniformly to zero and become unbounded in the neigh¬
borhoods of 5 of the point M0, one often defines the 8 function formally by the
relations
and
for M0 e T
8(M, M0)dr (3-5.53)
-f. for Mo £ T .
lim f fix)un(x)dx
n-ooja
ітЫ/l) (x—xq)
8n(x, x0) I e (3-5.55')
21 —n
when this expression is understood in the sense of the above explained weak
convergence; and in the same sense the relation
holds, where {(pn{x)} is a complete orthonormal system in the interval (a, by,
similarly
We shall now show that for the calculation of the integrals which con¬
tain the d function, the series (3-5.57) can be used and can be integrated
term by term. To this end we consider a function g(x) which can be developed
in a Fourier series, and the integral
j g(x)d(x0, x)dx .
/ \ Oo , /- я-m . = • 7Гm \
g{x) = -7г + I \Qm cos ——X + gm sin - x (3-5.59')
2 m=i\ I I J
with
If!
Qo 9(Xo)dx0
l
d(x, x0)g{x)dx = g{x0) , — / < Xo < I
-i
ut = a2uxx (3-5.61)
u{x, 0) = <p(x) (3-5.62)
u(0, t) = u(l, t) = 0 . (3-5.63)
Here a prescribed function cp(x) corresponds to a uniquely determined solution
u{x, t) = L[(p{x)]
of the problem. We assume that the operator L can be represented in the
form
where each summand must satisfy the heat-conduction equation. Hence there
238 PARABOLIC DIFFERENTIAL EQUATIONS
follows
An{t) = Bne.
From the initial condition there also follows
n 2 . m
B, = TS'n—ДГ..
This agrees with the representation of the Green’s function given in Section
5-3. The solution of the problem (3-5.61)-(3-5.63) is given by the formula
(3-5.64), where G(x,x0, t) is the function defined by (3-5.67).
In a similar manner, we can arrive at an expression for the Green’s func¬
tion for an infinite straight line. The function G is defined here by the
conditions
Ut — a2uxx = 0, — со < x < 00 (3-5.68)
u(x, 0) = <p(x) = d(x — x0) . (3-5.69)
Since here the 8 function should be defined by a Fourier integral
1
8(x — x0) = —\ cos X(x — x0)dX ,
n Jo
Under which conditions the formulas derived by use of the <5 function are
valid, is a question that must be investigated separately.
As an example, we shall now treat the inhomogeneous equation
8{x — f) = — ( cos fШ
7Г Jo
If we put this expression into Eq. (3-5.75), we obtain for u\(t) the equation
V Cp
240 PARABOLIC DIFFERENTIAL EQUATIONS
Therefore,
where
_1_ -(*-«>*/4a2(t-l0>
G(x, £, t — t0)
2\/na2(t - to)
represents the Green’s function. Similar methods for the construction of the
Green’s function have often been used in theoretical physics.75
du = 0 .
A function и is said to be harmonic in a region T if in this region the
function and its derivatives up to the second order are continuous and satisfy
Laplace’s equation.
For the investigation of the properties of harmonic functions, different
mathematical methods have been constructed which can be successfully ap¬
plied to hyperbolic and parabolic differential equations.
1
ut = a du ,
, 2
a =— .
k
cp
du = 0 . (4-1.1)
du = f, /=£ (4-1.2)
k
results, where F is the density of the heat sources and k is the coefficient
of heat conductivity. The inhomogeneous Laplace Eq. (4-1.2) is usually called
the Poisson differential equation.
242 ELLIPTIC DIFFERENTIAL EQUATIONS
Au = — f(x, y, z) (4-1.2)
v = grad (p (4-1.3)
or in terms of components
div v = 0 . (4-1.4)
76 Obviously a stationary temperature distribution can occur only when the total heat
flow through the boundary of the region is equal to zero. Consequently, the function
/2 must still satisfy the auxiliary requirement
4-1. PROBLEMS WHICH LEAD TO LAPLACE S DIFFERENTIAL EQUATION 243
or
J<p = 0 , (4-1.5)
div j = 0 . (4-1.6)
The electrical field strength E is determined by the current density, that is,
according to Ohm’s law
E = ^j, (4-1.7)
where Л is the conductivity of the medium. Since now the process is stationary,
the curl-free electrical field must be a potential field,77 i.e., a scalar function
<p(x, y, z) exists such that
E = — grad <p . (4-1.8)
A<p — 0 , (4-1.9)
i.e., the potential of the electric field of a stationary flow satisfies Laplace’s
differential equation.
We consider further an electrical field generated by stationary charges.
Because of the time independence,
rot E = 0 (4-1.10)
E- — grad<p . (4-1.8)
we obtain
77 From the second Maxwell equation (plc)H = — rot E, it follows that rot E — 0.
244 ELLIPTIC DIFFERENTIAL EQUATIONS
div E = 4Ttp .
i.e., the electrostatic potential <p satisfies the Poisson differential equation.
If, however, no spatial charges exist (p = 0), then <p satisfies the Laplace equa¬
tion
J<p =0.
The basic boundary-value problems for the processes considered correspond
to the three types given above. We shall not consider here other boundary-
value problems which also are important for the description of definite physi¬
cal processes. One of these problems is discussed in Section 4-7.
* = (pMi. <72, q3), У = Ч>г(Ц\, <h , <h), z = <p3(qit q2, q3) , (4-1.14)
rather than use the rectilinear coordinates x, y, z. The new orthogonal curvi¬
linear coordinate system is denoted by
fi(x, y, z) = Ci , f3{x, y, z) = C2
and
A(x, y, z) = C3 . (4-1.16)
Therefore, the orthogonality conditions for the edges of the elements of the
volume can be written in the form
dip i dip! , 42
ds2 = dx2 + dy2 + dz2 = ( dqx + dq2 + -P-
dqi dq 2 dq3
/ 5^2 Э^2_
+ ( dqi + dq2 + dq3
\ dqi dq2 dq3
I 4- +
dip
dq3 (4-1.18)
V dq3
where
(4-1.20)
Only one coordinate changes along each edge of the element of volume. The
lengths of these edges are therefore given according to (4-1.19)
Given a vector field A(x, y, z), then by a known formula of vector analysis
\AndS
div A — lim —- , (4-1.23)
uM-o vM
Qi — A\ds2ds3 Aids2ds31?1 .
246 ELLIPTIC DIFFERENTIAL EQUATIONS
In a similar way, we find for the other pairs of bounding surfaces the net
flux:
Q3 = (H1H2A3)dqldq2dq3 . (4-1.26)
dq3
Now if we substitute in formula (4-1.23) the value
AndS — Qi + Q 2 + Q3
and use (4-1.22), we obtain for the divergence, in the orthogonal curvilinear
coordinates, the expression
д {НгНМ+4-іНгНгА»
1
div A (H2H3Ai (4-1.27)
НуНгНг Г—
=
Uqi dq dq3 ]•
If we assume that A has a potential, i.e.,
A — grad и , (4-1.28)
then
du _ J_ ди — J_ ^u 1 du
(4-1.29)
dSi Hi dqi ~ H2 dq2 ’ H3 dq3
If these expressions for Alt A2, and A3 are substituted in (4-1.27), then
Laplace’s differential equation becomes
Ли = div grad и
= ттіЪтЬг
HiH2H3ldqi \ Н1 dqi ) + Т-{пг¥)
(Щг-Т-) dq2 \ Н2 dq2 J + аЧтг
dq3 \ H3 I1)!
dq3 )J • I4'1-30»
ds2 = (sin d cos (p dr + r cos Ѳ cos ipdd — r sin <9 sin <p d<p)2
+ (sin d sin (pdr + rcosdsinipdd + rsin<9cosy> d<p)2 + (cos ddr — rs'mddd)2
where, after simplification
Hi = 1 , H2 = r , Нъ — r sin d .
By substituting these values for Hlt H2, H3 in (4-1.31) we arrive at Laplace’s
differential equation in spherical coordinates:
_1_ da
sini? sin d + = 0
r2 sin d dd
or finally
i-JL 1 du 1 d и
—( sin Ѳ + = 0 . (4-1.32)
r2 dr r sin Ѳ dd dd r sin2 d dip2
— — ( дм \ J_ d2u
(4-1.34)
p dp\P dp ) p2 dip2
A/О dU\ = 0,
dr\ dr J
whose general integral is given by
U=^- + C2,
r
248 ELLIPTIC DIFFERENTIAL EQUATIONS
U, = — , (4-1.35)
r
и - U(p)
and by use of (4-1.33) or (4-1.34), we find the general form of the cylindrically
or circularly symmetric solutions (in the case of two-dimensional problems):
U(p) = Ci In p + C2 .
For Ci = —1 and Cz = 0 we obtain the function
U0 = In— , (4-1.36)
P
which is also often called the fundamental solution of the two-dimensional
Laplace differential equation.
The function U0 = 1 lr satisfies the equation Ли = 0 throughout with the
exception of the point r — 0, where it becomes infinite. To within a constant
factor it coincides with the potential due to a charge e placed at the origin
of the coordinates; the potential of this field is actually
e
и —— .
r
The function In 1 Ip satisfies Laplace’s differential equation throughout with
the exception of the point p = 0, at which it becomes logarithmically infinite.
To within a factor it agrees with the potential of an infinitely long wire (see
Section 4-5 § 2), whose potential is given by
и — 2ex In —
P
where ex is the charge density per unit of length. This function plays an
important role in the theory of harmonic functions.
і^=1іт^=1іт^4±І£НѴЫ
dz 4z-o Az jz-o Az
ux = vy
(4-1.37)
liy — Vx •
These conditions are necessary and sufficient. We can derive them as fol¬
lows:
Let w = и + iv — f(z) be an analytic function. If we calculate the deriva¬
tives
dw(z) dw
wx = ux + ivx
dz dz
dw(z) _ . dw_
wy = uy + ivy
dz Zy dz
and require that the value of dw/dz coincide with both expressions, then it
follows that
dw
ux + ivx = vy — iuy
dz ’
from which the Cauchy-Riemann differential equations result directly. We
shall not prove here that this is also sufficient.
In the theory of functions it is shown that every analytic function defined
in a region G of the z plane has derivatives of all orders and can be developed
in a power series. In particular, for each such function, the real part u(x, y)
and the imaginary part v(x, y) are twice continuously differentiable with respect
to x and y.
By differentiation of the first expression of (4-1.37) with respect to x and
the second with respect to у we obtain
This maps a region G of the x, у plane onto a region G' of the u, v plane
such that each point of G uniquely corresponds to a point of G', and con¬
versely.
Now let
U = U(x, y)
be a real, twice continuously differentiable function which is defined in the
interior of the region G.
We shall investigate how the Laplace operator behaves under the trans¬
formation U = U[x(u, v), y{u, v)] = U(u, v). To this end we first calculate the
derivatives
Now if и and v are harmonic functions conjugate to each other, then the
transformation (4-1.38) is equivalent to those transformations which are brought
about by the analytic function
w = f(z) = и + iv , z — x + iy . (4-1.40)
or
Ли,ѵи=—±—2Лх,уи. (4-1.41')
I / (2) I
/ 2
rr— a or (4-1.42)
du d2u
= j+^ =°
and
2 . d dv d2v
p Др,<рѴ = p P— ) + = 0.
dp \ dp / d<p
From this, by passage to the variables p and <p, it follows that
dv _ dv dp' _ / dv
P 77 ,
^ dp P dp' dp ‘ dp
i.e., v(p', <p) satisfies the equation t<pv = 0, since
i2 л r d / I dv > d2 v
P A?'.<pV = P 77 P 77 + = 0
dp \ dp dtp2
J_ A дги 2 ди _ 1д2(ги)
Г
(4-1.45)
Эт г dr г Эг2
T“2 ^ -л 2
r dr -2
Consequently,
or
Now, however,
or
r'4Ar',e,<pV = 0 .
(4-2.1)
between the z axis and the exterior normal to the surface I. The correct¬
ness of this formula is obtained by integration with respect to z. Ordinarily
we write the Ostrogradski formula1 in the form
where
dP dQ 3R
div A =
dx dy dz
and
„ dv ^ dv D dv
P = u— , Q = u— , R — u— ,
ox dy dz
then by the use of the Ostrogradski formula we obtain the so-called first
Green’s formula
ГГ Г/ди dv du dv du_ dv \
SS,\иШт=\Утпл° J J tJ dx dy dy^ dz dz )
(4-2.3)
d d 0d , d
— = cos a— + cos /9— + cos y —
dn dx dy dz
, , du dv du dv du dv
grad и grad v = Fu ■ rv = — — + — — + -—— ,
dx ox ay ay dz dz
dv ,
JLH*=-fLSr“-r”*+Lf и —— da
dn
(4-2.3')
If now we subtract the identity (4-2.4) from (4-2.3'), we obtain the second
Green’s formula
S$L.(“J7-T *)*
(4-2.6)
du
da .
dn
On the right side of this relation the last two integrals depend only on s.
For the derivative in the direction of the exterior normal to Ze we find
d_ d_ 1
dn dr r=e ~ £2
4-2. GENERAL PROPERTIES OF HARMONIC FUNCTIONS 255
•Therefore,
where u* is a suitable average value of u(M) on Іг. Further, the third in¬
tegral can be transformed as follows
l
(4-2.8)
r on e JjJ on e
where (du/dn)* is a suitable average value of the normal derivative du/dn on
If If we substitute the expression (4-2.7) and (4-2.8) into (4-2.6) and note
that d(llr) = 0 in T — Kt, then
1 du du
—^-\du dr = do + Anи — Ane
T— К p \ r) dn r dn dn
(4-2.9)
Now if we let the radius e approach zero, we obtain:
1. lim u* = u(M0), since u{M) is a continuous function and и* is an average
8—"0
du du , du „ , du
— = -^— cos a -f —— cosj9 + —- cos?-
dn dx dy dz
1
Anu(MQ) = — (4-2.10)
Г MqP
u(M) = (4-2.11)
and t tr~(—\(P)doP,
JjJ dnp \ r up )
(4-2.12)
256 ELLIPTIC DIFFERENTIAL EQUATIONS
1
and -L (J-') |(|)cos«, +^(!)cosA- +!(t)cos»
r MP dnP \Гмр )
satisfy Laplace’s differential equation with respect to the variable points
M(x, y, z), the functions (4-2.12) by the generalized superposition principle
(see the lemma on page 63) similarly satisfy Laplace’s differential equation
with respect to x, y, z.
Hence we arrive at an important conclusion: Every harmonic function,
in the interior of the region in which it is harmonic, possesses derivatives
of all orders.79 Further we note that a harmonic function in a region T at
each point M0 in T is analytic (i.e., it can be developed in a power series).
This assertion also results from the integral representation (4-2.11).
Corresponding formulas hold also for harmonic functions of two independent
variables. Let S be any region in the x, у plane and C its boundary curve.
Further let n be the direction of the exterior normal (with respect to 5) to
this curve. Then if v = ln(l/гм0р) is introduced where Гм0р = л/{х— х0)г + {у— y0)2
is the distance from a point M0 in the interior of S, then, by similar consid¬
erations as above, instead of (4-2.10) we arrive at the expression
du
u(M0) = (4-2.12")
2k dn
(4-2.13)
79 If for a function и which is harmonic in T the condition that it and its first de¬
rivatives are continuous on the boundary E is not fulfilled, the theorem still remains
valid. By this we mean that each point M is enclosed by a region, including its boundary,
lying in the interior of T.
4-2. GENERAL PROPERTIES OF HARMONIC FUNCTIONS 257
elude that the second boundary-value problem (Au = 0 in T, du\dn = f\x) pos¬
sesses a solution only if
Ц'л-°-
This property of a harmonic function can be interpreted as a condition that
no sources exist in the region T.
2. If u(M) is a harmonic function in a region T, and M0 is a point lying
in the interior of T, then
where Ea is a spherical surface with radius a and center point M0 which lies
entirely in the region T. This property is stated by the average-value theorem,
which reads:
Theorem. The value of a harmonic function at any point M0 is equal to
the average value of the function on an arbitrary spherical surface Ea with
center at M0, provided Ea lies entirely in the region in which u(M) is harmonic.
For the proof of this proposition we apply formula (4-2.11) to a sphere
Ka with center M0 and surface Ia:
4ku(M0) =
as well as
d_ d_ 1
2
dn dr a
1
u{M0) и do ,
4k a2
80 For the proof of this theorem we have used Eq. (4-2.13). This, however, assumes
that the derivatives also exist on the spherical surface. If a function u(M), which is
continuous in the closed region T + I, satisfies the equation Au — 0 only for interior
points of T, then this conclusion would not be correct for a sphere Zai), which borders
the boundary T. However, if the theorem for each a < ao is true, then as a -* do we
obtain
Now if we assume that at least one point M exists on Zp such that the
inequality u(M) < u{M0) is valid, then obviously in the last formula the in¬
equality sign must hold, which in turn implies a contradiction; consequently
on the entire surface Zp we must have u(M) = u(M0).
If po1 is the minimal distance of the point M0 from the boundary Zp,
then u(M) = u(M0) is also valid for points belonging to TP“. Hence, because
of continuity, it also follows that at those points M* which belong to the
intersection of TP™ and Z the relation u(M*) = u(M0) is valid. Therefore our
theorem is proved; and the last conclusion shows that the maximum u(M0)
is assumed at least at one point on the boundary.
It is easily seen that u(M) = u(M0) must be valid in the entire region
when the region T is connected and if at least at one interior point M0 the
maximum is assumed.
To demonstrate the above, we select another arbitrary point Mi0) in T
and connect it with M0 by the polygonal line L (Figure 46), whose length is
designated by l. Let My be the last current
point of L through the spherical surface
TP™. At this point, then, и(Мг) = u(M0) is
still valid. Now we enclose this point by
the spherical surface Zp™, where pГ is the
minimal distance of the point Mi from the
boundary. We obtain another such point
M2 as the last current point of L through
the spherical surface Tpj". By this procedure,
we arrive after a finite number of steps
(the number p of necessary steps is certainly
not larger than 1 lpim' if p[m) denotes the
minimum distance between L and T) at a
4-2. GENERAL PROPERTIES OF HARMONIC FUNCTIONS 259
spherical surface which contains the point M(0). From this it follows that
u(M{0)) = u{M0). Because of the arbitrary choice of the point M(0) and the
continuity of u(M) in the closed region T + 2 we conclude that u(M) = u{M0)
holds everywhere (including the boundary). Of all harmonic functions, there¬
fore, only the constant functions can assume their maximum at an interior
point.
The corresponding statement is also true for the minimum.
Conclusion 1:
If и and U are continuous in T + 2 and are harmonic functions in T
for which
и tk U on 21
U — и 'Sz 0 on I .
U-u^O
at all points in the interior of T— precisely our assertion.
Conclusion 2:
If и and U are continuous in the region T + 2 and are harmonic functions
in T for which
I и I 5= U on I
then also
I и I 5І U
— U -й и -й U onJ.
-ийийU
at all points in the interior of T, or
\ u\ й U
in the interior of T.
We assume that two different solutions ux and m2 exist which are continu¬
ous in the closed region T + I, satisfy Laplace’s differential equation in the
interior of T, and on the boundary I assume one and the same value /.
Then the difference и = Ui — m2 has the following properties:
1. Au = 0 in the interior of T.
2. и is continuous in the closed region T + S.
3. и U = 0.
Consequently, u(M) represents a continuous and harmonic function in
the interior of T which on the boundary is equal to zero. As is known, a
continuous function assumes its maximum in a closed region. Therefore we
must have и = 0. That is, if и Ф 0 and for at least one point и > 0, then и
must assume its maximum at an interior point of the region, which is impos¬
sible. In exactly the same way we show that и is never negative in T.
Consequently,
и =0 .
Often one has to deal with a first boundary-value problem with discon¬
tinuous boundary conditions. Naturally, a continuous function in the closed
region cannot be a solution to this problem, and we must therefore modify
the formulation of the first boundary-value problem accordingly.
Let a piecewise continuous function f(P) be defined on a curve C which
bounds a bounded region S. We seek a function u(M) with the following
properties:
1. u(M) is harmonic in the interior of S.
2. u(M) continuously approaches the boundary values at all points where
these are continuous.
3. u(M) is bounded in the closed region 5 + C.
We note that the additional requirement of boundedness is essential only
in the neighborhood of the points of discontinuity of f(P). Therefore, the
following theorem holds:
Theorem. The solution of the first boundary-value problem with piecewise
continuous boundary conditions is uniquely determined.
Let Ui and u2 be two solutions of the stated problem. Then we have for
the difference v — — u2:
1. v is a harmonic function in the interior of 5;
2. v continuously approaches the value zero on the boundary with the excep¬
tion of the points of discontinuity of f{P) at which v can be discontinuous;
3. v is bounded in S + C; | v \ < A.
We now construct the harmonic function
U(M) = e І In— .
i= l Гі
Here e denotes any positive number, D is the set theoretic diameter of the
region 5, and r, is the distance of the zth point of discontinuity Pi under
consideration. The function U(M) is positive since all the summands are
positive.
Further we enclose each point of discontinuity Pi with a circle Ki of
radius 8 and let 8 be sufficiently small so that each summand
i D
£ ІП-
Гі
of the corresponding circle Ci is larger than A, i.e., let £ In D/8 > A. Then,
in the closed region S — U v is continuous.
t=l
According to the principle of the maximum, U in this region is a majorant
of v:
262 ELLIPTIC DIFFERENTIAL EQUATIONS
I v(M) I ^ U(M) .
For a fixed point M in S, as e —> 0 we obtain
lim U(M) = 0 .
8—0
Therefore,
v(M) = 0 ,
since v is independent of £; or
ti\ — U-2 у
5. Isolated singularities
w=и—v
then has the following properties:
1. With the exception of the point P, at which w is not defined at all, it is
harmonic everywhere in Ka.
2. It continuously approaches the boundary value zero on Ca.
3. It is bounded in the closed region Ka + Ca, (I v \ < A).
As in the proof of the previous theorems (Section 4-2 §4), we again con¬
struct the non-negative harmonic function
U(M) = e In — .
r
Here £ denotes a positive number, a is the radius of the circle Ka, and r is
81 The existence of such a function will be shown in Section 4-3. Its construction
therefore does not depend on the theorem proved here.
4-2. GENERAL PROPERTIES OF HARMONIC FUNCTIONS 263
thus
w =0 .
Therefore, the function u, with the exception of the point P, coincides with
the function v everywhere in the region S. Now if we set u(P) = v(P) then
we arrive at the function и = v, which is harmonic everywhere in S. There¬
fore our theorem is proved.
Similarly, we proceed to the proof of the theorem for the three-dimen¬
sional case in which the function U(M) = 2(l/r — 1/a) can be used as the
majorant function.
For the proof of this theorem we should assume that the function и
remains bounded in the neighborhood of the point P. The proof also remains
valid if instead of assuming the boundedness we assume only that the func¬
tion u, in the neighborhood of the point P, satisfies the inequality
then it is bounded in the neighborhood of this point, and the value u(P) can
be so defined that u(M) is also harmonic at the point P.
du A du A du A
и 1 < — , ^ 2 f < . < 2
(4-2.19)
r (be r dy dz Г
with
Therefore, v(r', Ѳ, <p), on the basis of the last theorem in Section 4-2 §5,
is bounded and harmonic for r' = r0':
■ , a , , 1 v(r', Ѳ, <p) I A r 1
I u(r, Ѳ, cp) I = -r - ^ — for r ^ r0 = —
r0
Since now for r' =0, v is harmonic, we can write:
ЩьЛ = АЛ1. у, у.
x . 1 Г дѵ дх' дѵ ду' , дѵ dz' "1 .. 0
= -— ‘V + — —— + —7’*+-*-, (4-2. 22)
r r \_dx дх ду ox dz dx J
where
/ X / / V/ / Z f
x — —r , у ——r , z — —r
r r r
4-2. GENERAL PROPERTIES OF HARMONIC FUNCTIONS 265
If we now calculate the derivatives дх'/дх, ду'/дх, dz'/дх and consider the
boundedness of the first derivatives of v in the neighborhood of the point
r' - 0, we find
du
for r OO .
dx
Corresponding estimates are obtained for du/ду and du/dz.
и |sB = const. = /о .
If Condition 4 is dropped, then both the functions ut — f0 and u2 — foR/r as
well as the functions
и = aih + ft u2 with a + P = 1
are admissible solutions of the problem. On the other hand we have the
following theorem:
Theorem. The exterior first boundary-value problem for the three-dimensional
Laplace differential equation has a uniquely determined solution.
If Ui and m2 are two solutions which satisfy Conditions 1 to 4, then their
difference и — Ui — u2 represents a solution which satisfies the corresponding
homogeneous boundary condition. Since Condition 4 for the function и is
also fulfilled then for every e > 0, an R* can be found such that
If the point M lies in a region T' (Figure 47), which lies between the surface
I and a sphere Sz (r 2; R*), then obviously u(M) < e. This follows from the
principle of the maximum applied to the region T'. Since e can be chosen
arbitrarily we conclude that u, in T' as well as in the entire region T, vanishes
identically. Therefore we have demonstrated the unique solvability of the
266 ELLIPTIC DIFFERENTIAL EQUATIONS
— лг In r/R (4-2.23)
Rl N In RJR
We can also prove the unique solvability of this problem by using the
transformation of reciprocal radii. With this transformation the curve C in
this case is transformed into the curve C', so that the region lying in the
exterior of C is transformed into the region lying in the interior of C'. Thus,
the point at infinity transforms into an isolated singular point in whose neigh¬
borhood the function v remains bounded. From the theorem of the harmonic
functions in Section 4-2 § 5, it then follows that v is harmonic at the origin
of coordinates and the solution is unique.
From these considerations it follows that a harmonic function which is
bounded at infinity tends towards a definite limit value as M—> oo.
The difference in the formulation of the exterior first boundary-value
problem for two and three variables can be clarified by the following physical
example. Given a sphere of radius R whose surface is held at a constant
temperature u0, find the stationary temperature distribution of the surrounding
space. The function и = u0R/r represents the solution of this problem which
vanishes at infinity.
We consider now the corresponding two-dimensional problem. Thus, let
a circle of radius R be given, on whose circumference a constant temperature
u\s = fo = const.
prevails. In this case и = f0 is the uniquely determined bounded solution of
the problem and no further solution exists which would be equal to zero at
infinity. Earlier we emphasized the essential difference in the relation of
harmonic functions at infinity for two and three independent variables (for
example, the relation of 1/r and In 1/r at infinity).
Also, for spatial and plane unbounded regions the principle of the maxi¬
mum remains valid. This we know from the considerations which parallel
the arguments in the proof of the uniqueness theorem. From this follows
again the continuous dependence of the solution on the boundary conditions.
du
= 0 .
дп г
Now if we set v — и into the first Green’s formula (4-2.3) and take into con¬
sideration that An — 0 and du/dn |x = 0, we obtain
dr — 0 .
However, because of the continuity of the function и and its first derivatives,
it follows that
du du _ du
dx dy dz
i.e.,
и = const. ,
which was to be proved.
The method of proof used here can also be applied to the case of an un¬
bounded region when the function under consideration is regular at infinity.
First of all, we prove the following theorem:
Theorem. In the case of an unbounded region, Green’s formula (4-2.3) is ap¬
plicable to functions which are regular at infinity.
Therefore, we consider an unbounded region T whose boundary we denote
by 1. Further we choose a sphere of radius
R and denote by TR that part of our region
which lies entirely in this sphere. The
region TR is then bounded by that part of
the spherical surface IR which belongs to
T and by the surface which forms a
portion of the surface I (Figure 49). Now
if we apply Green’s formula in the region
TR to the two functions и and v which are
regular at infinity, we obtain
ГI пн
nr du dv du
uAvdr
rjl_ dx dx + dy dy
+ dz dz J dz
(4-2.23')
dv ,
+ Lj'“irA' + u——do .
dn
By using the regularity of the functions и and v we estimate the integral
over SR:
dv ,
и-do u(vx cos a Ar vy cos + vt cos y) do
dn
3A , ЗЛ2 12тгЛ2
< =S^r4 тгД2
i?da R3 R
4-2. GENERAL PROPERTIES OF HARMONIC FUNCTIONS 269
lim f f u-^—do = 0 .
dn
The integral over TK on the right side as R—>■ oo tends toward the integral
extended throughout the entire region T. This integral exists since the in¬
tegrand, because of the regularity of и and v at infinity, approaches zero as
l/R*. If, however, the limit value of the integral over Ti, which is equal
to the integral over I, exists, then the limit value of the right side of (4-2.23')
also exists and is given by
ff ГГ du_dv_ du dv du dv ,
uAvdz = — и-da . (4-2.24)
JJt-JLsx dx dy dy dz dn
Hence, the applicability of the first and consequently also of the second Green’s
formula in the unbounded region, under the assumption that the functions
which occur are regular at infinity, is proved.
We are now in a position to prove the following uniqueness theorem:
Theorem. The second boundary-value problem for an unbounded region pos¬
sesses a uniquely determined solution which is regular at infinity.
ux — 0 , uy — 0 , uz = 0 , and и = const .
Now, however, at infinity и — 0, so that
и = 0 ,
i.e.
Hi — ІІ2 у
(ul + Uy + Uz)dz = 0
ux = uy — uz = 0 and и = const .
The function и on the surface 2 is equal to zero. We can therefore conclude
that
и = 0 and Mi = m2 .
This proof is incorrect, however, since in the method of proof the existence
of the derivatives of the sought function on the surface 2 was assumed. The
proof of uniqueness which is based on the principle of the maximum does
not contain this weakness.
We begin with the solution of the first boundary-value problem for the
circle.
Determine the function и which satisfies the equation
Ju = 0 (4-3.1)
u —f (4-3.2)
center point of the considered circle of the radius a. Eq. (4-3.1), as is known,
has the form
1 dzu
ли
4
=-
1 3
+ 2 о 2
= 0 (4-2.3)
P dP P °<P
d/dp-(p(dRldp)) _ Ф" _ ;
Rip Ф
where X = const. Hence, we obtain for Ф and R the two differential equations
ф" + ХФ = 0 (4-3.4)
and
(4-3-5»
From Eq. (4-3.4) there results
R(p) = / .
With this and (4-3.5),
nz = p or pt = ± n , n > 0
results after replacing R{p) by /Л Consequently,
contrast, R(p) = Dp~n (p = — n) must be assumed for the solution of the exterior
problem, since this solution must be bounded at infinity.
Therefore we have found the following particular solutions of our prob¬
lem:88
u(p, ip) = f pn(An cos n<p + Bnsmwp) for the interior problem
71 = 0
and
°° 1
u{p, <p) = J, — (An cos nip + BnSinncp) for the exterior problem,
71 — 0 p
We assume that / can be written as a function of the angle <p. Then the
Fourier series for / has the form
(X
f{<P) = -7Г + 1 {an COS nip + /3* sin nip) (4-3.7)
Z 71=1
with
83 The Laplace differential expression in polar coordinates (4-3.3) loses its meaning
for p = 0. Therefore, we still have to prove that Aun = 0 also for p — 0. For this proof
we shall use not polar but rectangular coordinates; the particular solutions
A - h- A — —
■n'n
Ao~ 2 ’
—
a ’
A0 = An — OCn& , Bn = P nan
for the exterior problem. In this manner, we have found a solution of the
interior first boundary-value problem for the circle in the form
where
дкип
й tnnk2M,
d<pk
Now we choose a fixed value p0 < a (for the interior problem) or oy — a2/p0 > a
(for the exterior problem) where t0 = p0/a < 1. Then,
from which because of the convergence of the series on the right we infer
the uniform convergence on the left for t й to < 1 and arbitrary k. Therefore
series (4-3.8) and (4-3.9) at each point lying inside (outside) the circle can be
differentiated arbitrarily often with respect to <p. Correspondingly, we show
that series (4-3.8) and (4-3.9) inside (outside) a circle of radius p0 < a can be
differentiated arbitrarily often with respect to p.
Now po was chosen arbitrarily under the given restrictions. Therefore,
series (4-3.8) and (4-3.9) at each interior (exterior) point of the circle are term-
wise differentiable. Consequently, the superposition principle is applicable
and the functions (4-3.8) and (4-3.9) satisfy the equation Au — O.84
For the proof we only used the fact that the Fourier coefficients of /(y>)
were bounded (formula (4-3.10)). However, this is valid for every bounded
function (indeed for every absolutely integrable function). Therefore, provided
f(<p) is a bounded function, the series (4-3.8) and (4-3.9) define functions which
satisfy the equation
Au = 0 for t < 1 .
We shall use this fact later for the generalization of the results found
in this section.
We turn now to the proof of the continuity of the solutions in the closed
region t й 1. It is easy to see that we must go into the properties of the
function f(<p) more precisely.
Because of the assumed continuity and differentiability of f(p), it follows
that f(<p) can be developed in a Fourier series and that the series
І (I а» I + I I) (4-3.11)
n—1
so that series (4-3.8) and (4-3.9) converge uniformly for t й 1. The functions
represented by them are therefore continuous on the circumference of the
circle. Finally, formula (4-3.11) shows that the function (4-3.9), as the solution
of the exterior problem, is bounded at infinity. Therefore, we have proved
84 The Laplace differential equation is also satisfied for p = 0; if, therefore, the de¬
rivatives with respect to the polar coordinates are expressed through the derivatives in
terms of the cartesian coordinates, we can easily see that the functions (4-3.8) and (4-3.9)
are differentiable for t й to arbitrarily often with respect to x and y. On the basis of
footnote 83, it follows that
Au — 0 for p =0 .
4-3. SOLUTION OF THE BOUNDARY-VALUE PROBLEMS 275
that series (4-3.8) and (4-3.9) completely satisfy the conditions of the problem.
2. Poisson’s integral
u(p, <p)
_ J_
7Г
Г m\[-І-
l 2
+ f (p. y,
n=i \ a )
I пф cos n<p + sin пф sin tup) > йф
Г
(4-3.12)
J_
71
m\
l 2
[i-
+ I (^Y'cos n(<p
»=i \ a /
— Ф) |# •
1 oo
+1 tn cos n{<p — Ф)
} ”
— I
2 n—1 = y + ' 2 n=i
4-
2 [
/l + n=2i [{teu<e~^)n + (fir^^’nlJ
teUv~^
-ii+— +
2 \ ' 1 — }
1 1 - t2
t = -P-< 1.
2 1 — 21 cos (<p — ф) + t2 ’ a
1 - P2la2 йф
u(p, <p) = j' № (p2la2) — 2(p/a) cos (<p — ф) + 1
or
a — p
u{p, <p) = 7Г , №
27Г J —; 2 — lap cos {<p — ф) + a2^
(4-3.13)
This formula for the solution of the first boundary-value problem for the
interior of a circle is the Poisson integral, and the expression under the in¬
tegral sign
„2 2
a — p
K{p, <p, а, ф) =
p — lap cos (<p — ф) + a
is called the Poisson kernel. For p < a, we have K(p, <p, а, ф) > 0, since
lap < a2 + p2 for p Ф a.
The Poisson integral was derived under the assumption that p > a\ for
p — a, expression (4-3.8) loses its meaning. However,
because the series from which we obtained the Poisson integral is a continu¬
ous function in a closed region.
Consequently the function defined by
for p < a ,
u{p, (f) = I 2k L,mpl - 2ap cos (<f-ф) + atd<P (4-3.13')
{/(?) for p — a
is a solution of the equation Au = 0 for p < a, and in the closed region, in¬
cluding the boundary p = a, it is continuous.
The solution of the exterior problem has the form
1 p — a
7Г №p ap cos (<p — ф) + a'
АФ for p > a ,
u(p, (p) — 2k J_ — 2 (4-3.14)
l/M for p = a .
Under our initial assumption that the function f(<p) is continuous and
differentiable, we proved that the solution can be represented by a series.
Finally, by means of the above-cited identity, we arrive at the Poisson in¬
tegral.
Now we shall prove that the Poisson integral represents the solution of
the first boundary-value problem when the function f(cp) is continuous only.
The Poisson integral represents the solution of the Laplace differential
equation in the region p < a (t < 1) for an arbitrary bounded function f(<p).
Therefore for p < a (t < 1) the Poisson integral is identical with the series
(4-3.8) and the equation Au — 0 is satisfied for every bounded function f(<p).
It still remains to be proved that the function u(p, <p) in our case continu¬
ously approaches the boundary values. For this purpose we choose any sequence
of continuous and differentiable functions
lim/*(£>) = f(<p) .
fc —oo
For the functions Uk{r, <p) which are solutions of the first boundary-value
problem it follows from the principle of the maximum that
85 We shall not concern ourselves how it arises; such a sequence can be chosen in
many ways.
4-3. SOLUTION OF THE BOUNDARY-VALUE PROBLEMS 277
function и = lim Uk(p, <p). The limit function u(p, <p) is continuous in the closed
к— °o
region, since all functions Uk(p, <p) which are represented by the integral
since the sequence {fk(<p)} converges uniformly towards f(tp) and therefore
the passage to the limit can be carried out under the integral sign.
The function
If11 a2 — o’
u(p, V) = -f 2 9 / . № dtp
2л ) -np —2ap cos (tp — <p) + a
is thus, for every continuous function f(<p), a solution of the Laplace differen¬
tial equation which can assume the continuous boundary values given on the
circumference of a circle.
We now show that formulas (4-3.13') and (4-3.14) for every piecewise
continuous function f(tp) represent the solution of the boundary-value problem,
i.e., that this solution in the entire region is bounded and tends continuously
to the boundary value at the points of continuity of the function f(<p).
Therefore also it is a unique solution (see Section 4-2 § 4). If <p0 denotes
any point of continuity of the function f(<p), we have to prove that a d exists
for every e > 0 such that
І/Ы-/ЫІ <y
holds when
I ip — <Po I < SQ(s) .
but are otherwise arbitrary and for which formula (4-3.13) is applicable.
Then if we define the functions U(p, <p) and u(p, <p) by the formula (4-3.13)
for / and /, then these are harmonic functions which have the functions f(<p)
and f(<p) as boundary values.
Since the Poisson kernel is positive, then
Further, from the continuity of u(p, (p) and u(p, <p) on the boundary for 9 = <p0,
it follows that a 5i(e) exists such that
for
and
for
for
or
I u{p, (p) — f((po) I < £ for I a — p I < 5(e), | <p — <p0 \ < 5(e)
and hence the continuity of u{p, <p) at the point (a, (pQ).
We shall now show the boundedness of u(p, <p). Since the Poisson kernel
is positive, then
4-4. GREEN’S FUNCTION 279
2*
a
u(p, <p) < M^- йф = M,
jo a2 + p2 — 2ap cos (<p — <p)
since the left side, as was proved earlier, represents a harmonic function
which tends continuously toward the boundary value / = 1. Such a func¬
tion, however, is identically equal to 1. Correspondingly, we [could] show
that u(p, (p) > Mi for f > Mi and therefore prove the boundedness of the
absolute value of u(p, <p).
Let u(M) be а function which with its first derivatives in a closed region
T bounded by a sufficiently smooth finite surface I, is continuous; further,
let u(M) in the interior of T possess continuous second derivatives. Then,
as was proved in Section 4-2 § 1, the integral representation
u(M0) = —
гг i ди Ди
dru (4-4.1)
4л 1J L rPMo dn ?MM 0
is valid. If u{M) is a harmonic function, the volume integral equals zero;
if u(M) satisfies the Poisson equation, the volume integral represents a
known function.
Further let v(M) be a harmonic function which nowhere possesses singu¬
larities. The second Green’s formula
then yields
where
AG = 0 .
G |r = 0 ,
i.e.
The function G defined in this manner is called the Green’s function of the
first boundary-value problem for the Laplace differential equation Au — 0 (or
also, the function of an instantaneous point source of the first boundary-value
problem for the differential equation An = 0). It permits us to give a direct
representation of the solution of the first boundary-value problem for the
equation Au = 0. From (4-4.3) follows
It must also be noted that formula (4-4.4) was derived by the use of
Green’s formula, where certain conditions concerning the functions u(M) and
G(M, M0) as well as the boundary 2 were assumed. Moreover, the expres¬
sion dG/dn occurs, whose existence on 2 is not immediately seen from the
definition of G.
In the derivation of (4-4.4) we proceed from the existence of a harmonic
function u(M) which assumes the value / on I. For the region in which
the Green’s function exists and for which the Green’s formula is applicable,
formula (4-4.4) thus yields an explicit representation of only such solutions
u(M) of the first boundary-value problem as satisfy the conditions for the
applicability of the Green’s formula. (This formula also proves the uni¬
queness of this class of solutions of the first boundary-value problem.)
A detailed investigation by A. M. Liapunov showed that under very
general assumptions formula (4-4.4) represents the solution of the first
boundary-value problem for a wide class of boundaries—the so-called Liapunov
surfaces (see Section 4-5).
We turn now once more to the definition of the Green’s function G(M, M0)
which was introduced by means of the function v(M)—itself a solution of the
first boundary-value problem for the equation Av = 0 with the boundary-values
4-4. GREEN’S FUNCTION 281
v\i = — 1/4лт. This gives the impression that here this one problem (the
determination of the solution и of the first boundary-value problem) would
lead back to another of equal difficulty (the determination of v as the solution
of the same problem). This is not the case, however, since the knowledge
of the Green’s function allows the boundary-value problem to be solved for
an arbitrary boundary condition {u\i — /), while for the determination of the
function G itself only the boundary-value problem with the particular
boundary condition {v\z = —1/4лr) is to be solved. This problem is consider¬
ably simpler, as we shall see from a series of examples.
In electrostatics, the Green’s function
For the proof, we consider two spheres Ti and T2 of radius e with Mi and
Mo as center points; thus let Mi and Mi' be two arbitrary fixed points in
T (Figure 50). By 7\ we shall designate the region which results from T by
the removal of the regions enclosed by and I2. Now if we set
и = G(M, M'o) , v = G(M, M'o')
to Te, we obtain
since because AG = 0, the left side of (4-4.5) is equal to 0, while the integral
over I vanishes because of the boundary conditions.
Now, if we let e approach zero and consider the singularity of the
Green’s function, we obtain the expression:87
51~ln-.
i 1
Ал Тммо
87 Liapunov derived this theorem in the application to the class of so-called Liapunov
surfaces.
4-4. GREEN'S FUNCTION 283
. 1 , 1
v\a = — 7Г In-•
Умм0
A method often used for the construction of the Green’s function is the
method of images in electrostatics. Its fundamental idea in the construction
of the Green’s function
arises from the fact that the induced field v can be represented by the field
of the charges which are distributed on the exterior of the surface I and
are so chosen that the condition
is satisfied. These charges are called the electrostatic image points of the
unit charge found at the point M0, which, if I were not present, would
produce the potential 1/4ят. In many cases the choice of such charges causes
no difficulty. We shall give three examples for the construction of the
Green’s function by means of the method of electrostatic images. From the
representations of the Green’s function, which will be given for the three
examples, the continuity of the first derivatives of G on S will be obvious.
As the first example we consider the Green’s function for the sphere:
Given a sphere with radius R whose center lies at the origin of coordinates
O, find the corresponding Green’s function.
For the determination of this function we place a unit charge at the
point M0 and choose the segment OMi on the radial line passing through M0
such that
PoPi — R , (4-4.6)
the triangles OPM0 and OPM\ (Figure 51); they are similar since the angle
at О is the same for both triangles and the adjacent sides are proportional:
Po R OMo R
or
R pi R OMi •
Го _ Po_ _ R_ (4-4.7)
Гі R pi
(4-4.8)
is the sought Green’s function for the sphere; thus it is a harmonic function
which at M0 has the singularity (1 4—)(1 r„) and is equal to zero on the spherical
surface.
The solution of the first boundary-value problem can then be obtained
from formula (4-4.4).
We now calculate the normal derivative
(4-4.9)
where n is the exterior normal and ri = MXM (in general, M does not lie on
the spherical surface).
For the derivatives of l/r0 and l/гі along the normal n we have
(4-4.10)
since
dr
4^- = cos (r0, n) , = cos (iq , n) . (4-4.11)
on on
_ R2 + rl - pi
COS (r0, n) (4-4.11')
2Rr0
r>2 I 2 2
R + ri — p,
COS (Гі , It) (4-4.11")
2 Rrt
p2 ,
Л -Г 2 Го 2 2,2 т-,2
Po Po _po 4" Го — /t
COS(r1; Л) I г =
г,г> R 2роГо
ZR—r
Po
R2
Pi = —
po
holds, and
Гі = — Го on I .
po
We now introduce a spherical coordinate system whose origin lies at the center
of the sphere. Now let R, Ѳ, <p be the coordinates of the point P and p0,
e0, <po the coordinates of the point M0; let у be the angle between the radius
vectors OP and OM0. Formula (4-4.12) can then be written in the form
p л2й|-я p2 _ 2
u(po, 00, <Po) = T-\ 1 fi.0,<p)-775.-np , 2\3/2-^ітіѲdPdtp (4-4.12')
4л: Jo Jo {R — ZRp0 COS у + Po) '
where88
This formula is called the Poisson integral for the spherical surface.
By the same method the Green’s function can also be constructed for
the region lying exterior to the sphere:
—У —^
88 The direction cosines of the vectors OP and OMo are therefore equal to
(sin Ѳ cos <p, sin Ѳ sin <p, cos Ѳ) or (sin Ѳо cos <p0, sin Oo sin <p0, cos do) .
Thus
cos у = cos Ѳ • cos Ѳо + sin Ѳ sin Ѳо (cos <p cos <p0 + sin <p sin <рй)
= cos Ѳ cos Ѳо + sin Ѳ sin Ѳо cos (<p — <p0) .
286 ELLIPTIC DIFFERENTIAL EQUATIONS
(4-4.14)
4 л: Vn pi r0
Here ri = MMi is the distance of a fixed point Mx lying outside of the sphere,
r0 = MM0 is the distance of the point M0 conjugate to Mi, pz is the distance
of the point Mi from the origin of coordinates, and R is the radius of the
sphere.
If we bear in mind the distinction between the direction of the normals
for the interior and the exterior problem, then we obtain
л 2it p it
Pi R2
u{pi, Ѳі, (pi) = „— ,
4л: J о J, о го* 9 VnCOSc.I
[R — 2piR у + pi\ ' Ode d<p ,
where cos у is given by formula (4-4.13) (the subscript zero is there replaced
by 1).
The Green’s function for the circle can be obtained in the same manner
as the Green’s function for the sphere. In this case the Green’s function
assumes the form
G=^-ln— + v. (4-4.15)
2л: r
If we repeat the arguments of the previous
section from formula (4-4.6) to formula
(4-4.8), we obtain for G(P, M0) the repre¬
sentation
For the solution of the first boundary-value problem the normal derivative
dG/dn on the circumference C must be calculated. The calculation proceeds
as for the sphere and yields
8G 1 R2-pl
дп о 2лR rl
Now let p, Ѳ be the polar coordinates of the points P lying on the circum¬
ference and p0, Ѳ0 the coordinates of M0. Then, first of all, we have
1 Г
2
R2 po
u(p0, Ѳ0) = — \ /«?) dff , (4-4.17)
Lll Jo R + (Oo — 2i?p0 cos (0 — 0O)
the so-called Poisson integral for the interior of a circle. Except for the
signs this formula also gives the solution of the exterior problem.
The notion of the Green’s function and formula (4) are also valid for an
unbounded region of space, provided the functions considered are regular at
infinity (see Section 4-2 § 6).
To determine the Green’s function for the half space z > 0, we place at
the point M0(xo, y0, z0) a unit charge which produces in the unbounded region
of space a field whose potential is defined by the function
We easily see that the induced field v is the field of a negative unit charge
placed at the point Mi(x0, Уо, —z0), which is the mirror image of the point
M0 in the plane z — 0 (Figure 53). The function G(M, M0) is equal to
1 1
G(M, M0)
4лг0 4лгі
where
We now calculate
3G dG
dn dz
Obviously,
dG Zp Z + Z0
=i.r.
dz 4 7Г [_ Го r\ ]■
For z = 0, we thus find
dG _ dG Zp
_£o_
u(Mo) = — ' f{P)dop ,
Jx, .3
M0P
where T0 is the surface 2 = 0 and f(P) = и |*=0. Thus we can also write
OO Г со
u(xо , y0 , Zo) =
Zo
■f(x,y)dxdy. (4-4.18)
2n J-coJ-,» {{x — Xof + (у — Уо)2 + .31З/2
Zo]
The function
2_1_
r _ V(X - I)2 + (y - >y)2 + (2 - C)2 ’
which represents the potential of the field of a unit mass (charge) placed at
the point M0(f, у, 0, is a solution of Laplace’s equation. Here £, 77, and C
are interpreted as parameters. The integral of this function with respect to
the parameters is called the potential. It has a significant meaning both in
physics and also in the development of the methods of solution for boundary
value problems.
Let а mass m0 exist at M0(f, rj, £). According to the law of gravitation,
the attracting force on a mass m placed at the point M(x, y, 2) is
F=_rmmor
(4-5.1)
F=
X = F cos a = - ™°(x-£)
r
Y = F cos /3 = — -y (y — y) (4-5.2)
r
Z = F cos у = -^т(г-С)
г
where a, [3, and у are the angles between the vector F and the coordinate
axes.
Through
F = grad и
or
4-5. POTENTIAL THEORY 289
v _ du Y _du du
dx dy dz
we introduce the so-called potential u(x, y, z) of the force field.89 In our case
we have
By superposition of the force fields, the potential field of n mass points can
be expressed by the formula
AX= -£$■(*-£) ,
r
with
x=~\l\p£vldT- m-3)
Therefore we obtain for the potential of a spatial region T at the point M
the so-called Newtonian potential
89 The potential introduced here should not be confused with the energy potential
of a force field. Potential here is understood in the sense of the force function in
mechanics.
90 More precisely stated, it is assumed that the influence of a body T of mass m
on a point lying outside a convex region T containing this body can be replaced by the
influence of a certain concentrated force with the same mass m which lies in the interior
of T.
290 ELLIPTIC DIFFERENTIAL EQUATIONS
/iAz _ pAz
1?~ _ (X2 + 22)
AX = AF cos a = — uAz-■■ .
r V{x2 + 22)3
Hence it follows that
dz 2 1
й =-
2P
X= */2
cos a da — ,
fX (X2 + 22)3/2 “ ^ X3 -*/2 x
z
— = tg
* a .
r
If P{x, y) is an arbitrary point, then the attracting force acting on a point
—»
at О obviously is in the direction of OP and its magnitude is equal to
91 A sufficient condition for differentiation under the integral sign, with respect to
a parameter, for an integral of the form
is the continuity of the derivative of F(M, P) with respect to the parameter and the
absolute integrability of <p{P). Ordinarily this theorem is formulated only for <p{P) — 1,
although the proof for this and the general case are not different.
4-5. POTENTIAL THEORY 291
where
p = ~[/x2 + yz .
1
V = 2pln (4-5.5)
Y = Fsina = —2p-~ , У
sin a = —
P
92 In the calculation of the potential of an infinitely long straight line, the potential
of the individual elements cannot be integrated directly since an improper integral
would result. The potential of an element Az is therefore equal to
Az
Ди — p-
*/ p2 + z2
so that formal integration would yield the improper integral
dz
и
F p2 + z2
292 ELLIPTIC DIFFERENTIAL EQUATIONS
X=
x-e ■d$drj
v)
(x - o2 + (y- v) (4-5.6)
y— ~2 fJsjUe,9)7(y-£=-?—
— y)+(x — t)
-d£ dr]
which we can prove by differentiation if the point lies outside the region S.
If P lies inside of S, then an additional investigation is necessary in the
proof.
3. Improper integrals
/. = [[ \Fdv en^0
JJ T— K^nJ
has a limit value independent of the choice of Ken, then the limit value is
called an improper integral of the function F(x, y, z) over the region T and
is usually denoted by
(4-5.8)
==2к - 2CL” £
47гС Г ——~—F a~\ for аФ3
Ы — a
4л:C [In r]f„, for a = 3 .
Now if we let en—>0, then it can be shown that the limit value exists for
a < 3 but not for a ^ 3.
Now let F(x, y, z) be a nonnegative function and let the following limit
value exist
/ = lim ((( F dr ,
sn-° J J Jr-Ken
F dr F dv S: Fdv
T-K„ T-Kp
»i c»2
F dr = lim F dz = I,
J J JT-Ke
1ітШ T-Kp
since the limit value of the outer integral exists and coincides with it.
294 ELLIPTIC DIFFERENTIAL EQUATIONS
4dr (4-5.8)
tJ У
is valid.
For the proof, we consider any sequence of regions K<.n which entirely
contains the singular point M0. Because of the convergence of the sequence
{/„} of the integrals of F(x, y, z), an N(e) exists for each e > 0 such that
Fdr < e
-K-tn
n\
when tii, пг > N(e) holds. On the basis of the Cauchy convergence criterion,
the condition (4-5.10') is, however, sufficient for the convergence of the
sequence
/» = Fdr
-K?n
toward a limit value
I= lim I„ = (( [Fdr .
JJrJ
4-5. POTENTIAL THEORY 295
We easily see that this limit value is not dependent on the choice of the K!n.
Therefore, the existence of the improper integral (4-5.9) is completely proved.
If, on the other hand, a nonnegative function F(x, y, z) can be found
such that F(x, y, z) 2; F(x, y, z), and if the improper integral of F diverges
over the region T, then the integral of F also diverges.
Conclusion:
If for a function F(M, P) which becomes unbounded at P = M the ine¬
quality
F(M, P)drP
is valid for each point M whose distance from M0 is less than 8(e), and for
every region TS{e) which contains the point M0 and whose diameter d ^ 8(e).
We prove now that at the point M0 a uniformly convergent integral
V = Fi + V2,
where the integral Vx is taken over 7\ and the integral V2 over T2 = G — 7\.
We shall consider later the magnitude over 7\. The following estimate is
valid:
Each summand standing on the right can be made smaller than e/3 for
sufficiently small \M0M\. Now if we choose Ti in the interior of a sphere
I VM) - Vt(M) I ^ у ,
provided
Now if we set
V(M) (4-5.12)
rMp
Х(М) = -[\[^Р-(х-Штр
JJrJ TMP
Z(M)=-\\\^(z-OdTP ,
JJrJ Tup
at the points which lie in the interior of the attracting body. The improper
integrals (4-5.12) and (4-5.13) converge when the density p(M) is bounded,
p\(M)\ < C. For the potential V(M) this is obvious since
For the components of the attractive force the assertion follows from the
inequality
Ы \x - £| C_
„2 „ ^ A a=2 < 3
since I я — ? I < r.
As an illustration of the uniform convergence of an improper integral
we show that the integrals (4-5.12) and (4-5.13) are continuous functions.
Therefore, we have to prove that the integrals (4-5.12) and (4-5.13) con¬
verge uniformly at each point M0.
First,
dr p
йc
T.MP Tmp
о^
Thus the condition for the uniform convergence of the integral V is satisfied.
By a corresponding consideration for the integral
X(M)= \p(P)^Z±dTP,
JItJ T sip
93 The integral (4-5.12) results from integral (4-5.11) for F(M, P) = 1 /гмр, AP) = p(P)-
298 ELLIPTIC DIFFERENTIAL EQUATIONS
we obtain
dr i
p(P)^AdzP й c г» 0
dzf
<c = 8n5C < e
Ts J rip ГІР 25 rip
when
d Y d(e)
8nC ‘
Therefore the potential V and the components X(M), Y(M), and Z(M) of
the attracting force in the entire space are shown to be continuous functions.94
W) = ([ \№-drp .
JJrJ Гup
If the differentiation under the integral signs is carried out for the function
V(M), then
dV _ dV dV
X= (4-5.14)
dx ’ dy ’ dz
i.e., V is the potential of a field with components X, Y, and Z.
If the pointM lies outside the region T, then
a: — <?_—(x — g)_3_1_
rip [(x — f)2 + {у — T])2 + (2 С)Г2
— дх r HP
is a continuous function with respect to the two arguments M(x, y, z) and
P(£, 7], Q. Consequently, in this case differentiation under the integral sign
can be easily carried out.
The higher derivatives can be calculated by differentiation under the
integral sign at all points exterior to T. Hence it follows on the basis of
the lemma from Chapter 3, Section 2 that the potential outside the attracting
mass satisfies the Laplace differential equation
JV = 0 .
We shall now show that the derivatives of the first order of the potential
V also can be obtained by differentiation under the integral sign when the
point M considered lies inside of the region T.
94 The uniform convergence of the integrals V(M) and X(M) was shown under the
assumption that the density p is bounded (| p \ < C). Consequently, these integrals also
are continuous at the points of discontinuity of the function p, in particular on the
boundary of the region covered with a continuous distribution of mass.
4-5. POTENTIAL THEORY 299
For the proof we use the boundedness but not the continuity of the
function p(x, y, z) (I p(x, y, z) I < C). It follows that V(x, y, z) also is continuous
at the boundary points of T. These points can be considered as points of
discontinuity of the function p(P) which vanishes outside of T.
We shall now prove that for every e > 0 a 5(e) exists such that
Гі T r
• g / л 2it p it 2
x f j
< С sin MM-VAL = 4nC5' < ~ , (4-5.15)
l*il = P——
r dr
Jo Jo Jo r 3
since \(x — Z)jr\ ^ 1 and \p \ < C. Further, we have for the third summand
with
4-(( Ui—i-r
Ax JJrJ \Гі
dz
m r — r1 ,
p-dz
ГГ 1
The sides of the triangle М0ММ, are equal to r, rir and \Ax\. Hence it
follows that
I r — Гі I ^ I Ax I ,
so that
SI ^ C
Therefore,
= 4лд' ,
and
can be satisfied. If we choose o' according to the condition (4-5.16), then the
inequality (4-5.15) also is satisfied. Now we set 7\ = К*0 and therefore also
T2 = T — Ti.
The first equation of (4-5.14) applied to the region T2 so selected means
that for every e > 0, ад" exists such that
V2(x + Ax, y, z) - Ѵг{х, у, z) £
T2
Ax T
8V
X (4-5.17)
dx
The expressions
dV
and = Z
dz
can be proved similarly.
4-5. POTENTIAL THEORY 301
dr (4-5.18)
with a=3 .
т-м
the integral sign since the point M0 lies outside of T2, e.g.,
дгѴг
dx2 dx
For the first derivative of Vi with respect to x we find
(4-5.19)
since
8_
dx
By applying the Green’s formula, the integral (4-5.19) is transformed
into
302 ELLIPTIC DIFFERENTIAL EQUATIONS
where Tf° is the spherical surface bounding 7\ and a denotes the angle
between the exterior normal to the surface l'*0 and the x axis. The first
summand is a differentiable function at the point M0, since M0 lies outside
of Tf°. Similarly, if the second summand is differentiable in the neighborhood
of the point M0, then p is differentiable in 7\. From this it follows that at
the point Mo the second derivative of the function Vi exists. Consequently,
(4.5-20)
dx\r JdZ
provided
dp_
< Cx .
dt
If we apply the mean-value theorem to the surface integral, we obtain
-k*»S4(-7)cosaA’ = ■
d_ x—6 1
3 — — cos a
dx r
and
&Н^4(т)С08Н = • <4’5-21)
,
The equation
«JLj'w)*=іігЩт)* (4-5'23)
follows from the existence of the second derivative d2V/dx2 proved above.
4-5. POTENTIAL THEORY 303
This integral was obtained by a special choice of passage to the limit, and
indeed the sphere was shrunk to the point M0.95 This is indicated in formula
(4-5.23) by the bar above the integral sign. In general, a change in the shape
of this region implies a change of the limit value; the integral (4-5.23) is
conditionally convergent (in the sense of the above definition). Accordingly,
dv — 4 лр(М0) (4-5.25)
— —4л-p(M0)
since 1 lr is a harmonic function.96
Therefore the spatial potential satisfies Poisson’s equation
AV = -Акр
inside the body, and the Laplace equation
AV = 0
Au = — f (4-5.25')
6. Surface potentials
U(M)
= l4
4л JjJL rMp dn on \r J J
МП = lim «<£!>. =
<T)/ '0 (fy dop
as the surface density of the surface I at the point P.97 The potential of
this mass load can then be represented by the surface integral
where ту and r2 are the distances of the points M from 7\ and T2, respectively.
97 If the mass with spatial density p is distributed into layers of thickness h on the
surface I and the field is considered at the points whose distances from the surface
are large in comparison with h, (h/r •€ 1)> then the consideration of the thickness
of the layer in general is unimportant. It is suitable however to consider, instead of
the spatial potential with density p, the corresponding surface potential with a surface
density /( = ph.
4-5. POTENTIAL THEORY 305
r=
Here the vector l is along the direction from the repelling mass to the at¬
tracting mass, and r is the distance of the point M(x, y, z) from a fixed inter¬
mediate point P($, 7], Q of the interval Al.
For the derivative in the l direction we obtain
where the vector r is directed along the dipole to a fixed point M, whereas
ip is the angle between the vectors l and r. Hence, the potential of the
dipole is
(4-5.27)
—>
(4-5.28)
If (p\ is the angle between the interior normal and the direction PM,
then
W= ( \^^-v{P)d(Xp .
JjJ r HP
V = [ ii{P)\n—^—ds , (4-5.29)
Jo Tup
do p
cosy
W vds ,
-L-
we consider the curve C in the x, у plane and select as the origin of
coordinates the point P. The x axis coincides with the tangent to the curve,
the у axis with the normals at this point (Figure 61). In a fixed neighbor-
4-5. POTENTIAL THEORY 307
hood of the point P, the equation of the curve can be written in the form
У = У(х) .
Now C, by assumption, has a continuous curvature, i.e., y(x) possesses a
continuous second derivative. Consequently,
where, because of the special choice of the coordinate axes, it follows that
y(x) = —хгу"{Ях)
Therefore, we have
/*■ + ** = У1 + ,■ ,
cos у _y"(9z)_
2{1 + ^J}
Further, for the curvature,
y"( 0) =K(P)
results from the expression
кy
(1 + y'2?'* ’
Therefore,
,. cos (p 1 jr. n.
lim — = —K(P)
MP—0 r 6
This expression shows the continuity of (cos <p)lr along the segments of the
curve and therefore also the existence of the potentials of the double layer
at the points of the curve C, provided v is bounded.
In the three-dimensional case, the potential of the double layer also exists
at the points of the surface considered, since the function
cosy
where r is the distance between the points P and P', A is a fixed constant,
and 0 < d ^ 1.
Now let P0 be an arbitrary point of 2. We then choose a rectangular
coordinate system whose origin is at P0 and whose z axis is in the direction
of the exterior normal. The x, у plane then coincides with the tangent
plane at this point. Because of condition (4-5.2), a p„ exists such that the
equation of the surface 2 can be written in the form98
2 = f(x, y) , (4-5.32)
provided
98 If the function f(x, y) [in the neighborhood of the point P0 has a continuous
second derivative, then the surface satisfies the Liapunov conditions. Consequently,
surfaces with continuous curvature belong to the class of Liapunov surfaces.
4-5. POTENTIAL THEORY 309
1
1 > COS у= > ѴГ (4-5.34)
V\ + zl + zl 2
and since
cos a cos В
zx --, zy —-— with < 2
cos?- cos у cos у
then
0^ x^x, 0 ^у^у.
From this follows the estimate
The estimates (4-5.34) and (4-5.36) now permit the proof of the proposition
that the potential of the double layer
2 = f(x, y) ,
where the function f{x, y) satisfies the inequalities (4-5.34) and (4-5.36).
Now if <p is the angle between the direction of the interior normal at
the point P($, 1}, 0 and the direction PP0, then we can easily obtain the
relation
310 ELLIPTIC DIFFERENTIAL EQUATIONS
£ « Г \r\
I COS (p —cos a + —cos/3 H-cos у 5S I COS a I + I COS /3 I + —
r r r r
= + Лг/>р0 T 4Лг/>/-0 = 6 Ar]>p0
and
cosy
< 6Л- 2-S о<г<l (4-5.37)
Further, we can write the integral W{M) as the sum of two integrals:
W — Wi + Wi .
Here Wi denotes the integral over the surface S'Fq containing the singular
point P0, and W2 is the integral over the remaining part of the surface,
that is, it is extended over I — Z'p0. For proof of the convergence of W it
is sufficient to show the convergence of W{, since the integrand of W2 re¬
mains finite everywhere. In polar coordinates p — i/£2 + Vz and Ѳ in the
x, у plane,
— d£dr] _ pdpdd
cos у cos у
On the basis of the estimates (4-5.34), (4-5.36), and (4-5.37) and because p < r
we have for the integrands
cosy 1 12 AC
v(P) ^ F=
r2 cos у i°2~S
This form of the majorant function F, however, guarantees, in the case
of two independent variables, the convergence of the improper integrals (see
Section 4-3).
It can easily be shown that for a Liapunov surface the potential of a
single layer
also converges at points of this surface. Let it also be noted that the
potential for a more comprehensive class of surfaces converges.
In the two-dimensional case the potential of the single and double layers
converges at points of the curve considered [see formulas (4-5.29) and (4-5.30)],
provided this potential exists on a Liapunov curve. The conditions for these
curves are analogous to Conditions 1-3 for the Liapunov surfaces.
W(M) = -^^^(P)Jsp .
JO fMP
We begin with an arbitrary element of arc ds, whose end points are P
and Pi. Through the point P we construct a circular arc of radius MP about
M. This cuts the ray MP at the point Q. Then we can write to within
terms of higher order
(Figure 62), where ds = PP and da — PQ, while da> denotes the angle by which
one observes the arc ds from M. The sign of da) coincides with the sign of
cosy>. Consequently, day > 0, provided y> (the angle between the interior
normal at the point P and the vector PM) is less than ?r/2, and day < 0, pro¬
vided <p > 7t/2. If day > 0, i.e., <p < к/2, then the point M is seen from the
“interior” side, and if day < 0 (y> > тг/2), by contrast, it is seen from the
“exterior” side of the curve C. From this it follows that the angle by which
one sees an arc PyP2, from the point M equals the angle PiMP2, which de¬
scribes the ray MP, when the point traverses the arc PP2 •
We now consider the potential W° of a double layer with constant
density v = v0 = const. The ray MP describes the angle
[ 2л: providing the point M lies in the interior of curve C
& = j 7Г providing the point M lies on the curve C
1 0 providing the point M lies in the exterior of curve C
when the point P traverses the entire curve C. From this we find for the
potential W°:
da cos <p , ,, r
-5—— = dw , (4-5.42)
r
where dw is the solid angle subtended by the element
da of the surface 2. Now let da' be an element of
the spherical surface which we obtain when we let
a cone whose vertex is at point M and whose base
is the surface element da intersect the sphere of
radius MP about M. We then obtain for the spheri¬
cal surface element
da' = da cos <p .
From this formula (4-5.42) also follows. The above
observations with regard to the signs of dw also remain valid here. This
leads to the relations.
[ 4л-ц, provided M lies inside of 2
W° = vaQ — j 2nv0 provided M lies on S
1 0 provided M lies outside of 2
in which W°, W°A denote the values of the potential W° in the interior and
in the exterior of the surface 2, while W\ is the value of the potential on 2.
We proceed now to the consideration of the potential of a double layer
with a variable density and show that at the points of continuity of the
density, relations exist which correspond to formulas (4-5.41) and (4-5.41').
Let P0 be a point of the surface 2 at which the function v(P) is con¬
tinuous. Further, let W° be the potential of the double layer with constant
density v0 = v(P0). Then the function
point P0. Let us prescribe an arbitrary number e > 0. From the continuity
of v(P) at the point P0 it follows that for an arbitrary prescribed r] > 0 a
neighborhood Ex exists of P0 on the surface I, such that
I h I < f]Bs ,
(4-5.43)
but which does not depend on the choice of the surface Ex. More details
will be given later about this constant.
If we now select у =
e/Bs, we know that for every e > 0 there actually
exists a Ei containing a P0 such that for each point M we have
I h(M) \<e.
From this, however, follows the uniform convergence of the integral I(M)
at the point P0 and therefore also the continuity at this point.
If W,(P0) and Wa{P0) are the limit values of the potentials Ws as M—
on T+ and E~, respectively (E+ and E~ denote the interior and the exterior
sides of E), then
V= f f—p(P)daP + ( \—ii{P)daP = Vi + V2 ,
Гнр JX2J r»IP
I /л{Р) I < A
and note that
da — — d£drj
cos у cos у
and
_ dz dr; _
a„s .
VAM) < 2A
V(x — f)2 + (y — j?)2 0 p
For 8 = е/8л-A,
VAM) < e ,
when MPa < 8. Consequently, at each point P0 on 2, V(M) converges uniform¬
ly and therefore is a continuous function.
On the other hand, however, the normal derivatives of a single layer
have discontinuities on 2 of the same kind as the potential of a double layer.
4-5. POTENTIAL THEORY 315
The exterior and interior normal derivatives of V, that is, dV/dnA and
dV/dni, are defined as follows. Let P0 be an arbitrary point on 2. Through
Po we draw the z axis, which then can have the direction of either the ex¬
terior or interior normal.
We now consider the derivative dV/dz at a
point on the z axis. We denote by (dV/dz)r and
(dV/dz)A the limit values of the derivative dV/dz
as the point M approaches the point P0 from 2+
or 2~, respectively. If the z axis has the direction
of the exterior (interior) normal, then this value
is called the inner (outer) limit value of the deriva¬
tive with respect to the exterior (interior) normal
at the point P0.100 FIG. 63.
Now we set
- - wm - KPo)
= £— J ^(/z cosѲ) C0S2^-~do — I j^sinffcos#
ГР0Р
where <p0 is the angle between the z axis and the vector P0P. If we note
that MPo) = t*(Po), then
fdV
— 2tt/z(P0)
\drii
+ ^Kfi(Po) (4-5.47)
\ОПі Jo
results, since by convention the 2 axis has the direction of the interior
normal. If the z axis is in the direction of the exterior normal, the sign of
cos<p changes and we obtain
2 лyz(P0 (4-5.48)
\dnAJA \dnAJ0
In the two-dimensional cases there are corresponding formulas; we have
only to replace 2л: by л.
For an arbitrary choice of v(P) the function W(M) satisfies the Laplace
differential equation interior to C. On C, W{M) is discontinuous. There¬
fore if the boundary condition is to be fulfilled, we obviously must have
W,(Po) = f(Po) •
By using formula (4-5.41) we obtain the equation
is the kernel of the integral equation. Here we are dealing with a Fredholm
integral equation of the second kind.104 For the exterior problem we obtain
the corresponding equation
where
COS фо
к list, s) ln-M = (4-5.55)
dn, r pp0) Гррп
1
u(M) In-/Lt(P)dsi
-I Г UP
since (p is the angle of P0PP' (Figure 64). The integral equation for the
function v(s) therefore assumes the form
1
v{s) - —f(s) + A, (4-5.57)
7Г
A = -ih\fs)ds-
Therefore
v(s) = -f(s) -
7Г
— 4 —
R Jo
f f(s)ds (4-5.58)
юг we easily see that K(so, s) = Ki(sf so). Such kernels are called adjoint and the
corresponding equations, adjoint integral equations.
4-5. POTENTIAL THEORY 319
1
W = -[ ———^—f(s)ds — f(s)ds) ^ds
Л Jo r 4 k2R /Jo r
1
■f(s)ds f(s)ds) • 2л (4-5.59)
л J0
=1( r 4 n2R
-7І(*1*-к>*
Now we see from AOPM (Figure 65) that
COS (p _l_27? cos <p — r _ 2Rr cos <p — r
К
r 2R ~ 2 Rr ~ 2 Rr2
r>2 2
R — pa (4-5.60)
2R[R2 + pi - 2R/>0cos(0 - Ѳ0)] '
Then
pi = R2 + r* — 2Rr cos<p .
Finally, if we substitute expression (4-5.60) for К into (4-5.59), we obtain
Poisson’s integral
1 (R2 — pl)f(0)d0
U = W(p0,0a) = (4-5.61)
2л R + po — 2Rp0 cos {0 — 0 o)
as the solution of the first boundary-value problem for the circle.
These considerations show that for every continuous p
function f(0), a harmonic function is defined by formula
(4-5.61) which continuously approaches the boundary value
№.
If / is only piecewise continuous, then the function
W, because of the properties of the potentials of the
double layer, also is everywhere continuous where / is
continuous. Further, from the boundedness of /,
I/I <C,
Г
2
1 /?2
W(p0, Oo) I < C;
(>n
■do = C ;
2л-Jo R + f>o — 2Rp0 cos(0 — 0o)
then1
r>2 2
R — po ■dO = 1 (4-5.62)
2it)o R + pi — 2Rp0cos{0 — 0o)
106 Equation (4-5.62) results from the fact that the left side represents the solution
of the first boundary-value problem for /=1 (see page 281).
320 ELLIPTIC DIFFERENTIAL EQUATIONS
2. The first boundary-value problem for the half space. Let us determine a
harmonic function which is everywhere continuous in the half space z is 0,
assumes a prescribed value f(x, v) on the boundary 2 = 0, and at infinity is
equal to zero.
We write the solution of this boundary-value problem as the potential of
a double layer:
cosy
Г2 Г3 [(x — f)2 + (у — 7])г + 22]3/z
and for the kernel of the integral equation we find
1 ( cosy \ _ 0 .
2Д r / 2= 0
v(P) = ±f(P) ,
Ка)(Р, М) = К(Р, М)
is continuous.
If
then
c
\Kl>2\ < г-Лі~аг for tt! + a2 < 2 , r = P,P2 .
Obviously, it is sufficient to prove this estimate for the case in which the
point P2 lies in the Liapunov neighborhood 1 „ of the point Px. Therefore,
instead of the integrals over , we can consider the integral over the pro¬
jection So of this neighborhood on the tangent plane at the point P!, since
^ В> 0
r(P,M)
(where p(P, M) is the distance between the pro¬
jections of P and M on the tangent plane, В is a
certain constant), and since the relation do = dSI cosy
exists between the surface element do and its pro¬
jection dS; then according to formula (4-5.34),
cos у > 1/2.
The following lemma holds for a plane region:
If
Ci
Ki I <
r]~ai
then
C
ІЛ ( \Kl(P1,M)Kl(M,Pt)dxdy ^ r2-“l-«2 *
Js0J
Let R be the diameter of the region S0. Then we split the integral I
into the integral Л, over a circle Gi of radius 2r about P,, and /2, over the
322 ELLIPTIC DIFFERENTIAL EQUATIONS
C,
2ir ГR 2-Л1-Л2 * аі + а2 < 2
1
I h I < 4СіС2 Oby—Cb 2
ridr{d(p < •
0 J 2r A
С /^л1+а>2-2 «і + а2 > 2 .
1 ( C]C2
/il < -®l-*2 rn-a, rn-^dx'dy'
In the last integral, extended over the circle G[ of radius 2r, r[ is the dis¬
tance from the center and A is the distance from the bisection point of
the radius. Consequently, this integral converges, i.e., it does not depend
on r.
From this it follows that
C,
l/il < yi-a\-<t2
Accordingly, from a certain index on, the integral, which can be represented
by the iterated kernel, is bounded and converges uniformly, i.e., they are
continuous functions of their arguments.
With the use of the first Fredholm theorem we shall now prove that
the integral Eq. (4-5.50) has exactly one solution.
We limit ourselves here to the consideration of convex curves which
contain no straight line elements. In this case, the kernel of Eq. (4-5.50) is
not negative, since
K(P0, P)dsP — dco ,
where dw is the angle formed by the arc dsP as seen from the point P0.
We consider the first interior boundary-value problem. The homogene¬
ous integral equation corresponding to Eq. (4-5.50) has the form
4-5. POTENTIAL THEORY 323
j K(s0, s) ds = n .
With the help of this expression we can write the homogeneous Eq. (4-5.63)
in the form
Now let P*(s*) be a point of the curve C at which the function | v(s) |
assumes its maximum. Then the sum v(s*) + v(s) has a constant sign. If
now in (4-5.64) we set s0 = s* and note that K(s0, s) = 0, we obtain
This result, however, contradicts the continuity at the point s*, if only
v(s*) Ф 0.
Consequently, the homogeneous Eq. (4-5.63) possesses only the zero solution
so that the inhomogeneous equation for an arbitrary /(5) has exactly one solu¬
tion.107
The exterior second boundary-value problem leads, as we have already
seen (Section 4-5 § 10), to the integral equation
whose kernel Ky (s0, s) is adjoint to the kernel К (s0, s), i.e., (s0, s) = K(s, s0)
now holds.
We shall now consider the second Fredholm theorem, which reads:
Theorem. The number of linearly independent solutions of a homogeneous
Fredholm integral equation is equal to the number of linearly independent
solutions of the adjoint equation.
According to this theorem, the solution of Eq. (4-5.54) is uniquely defined.
The exterior first boundary-value problem corresponds to the integral equation
v(s*) = v(s)
107 If straight line boundary elements are present, then the considerations are some¬
what complicated but present no special difficulties.
324 ELLIPTIC DIFFERENTIAL EQUATIONS
v(s) = const. = V0
is the single solution of the homogeneous equation. On the basis of the second
Fredholm theorem, therefore, the adjoint homogeneous equation has exactly one
solution.
The third Fredholm theorem reads:
Theorem. If a homogeneous integral equation
j f(x)<pi{x) dx = 0 .
^ f(s)ds = 0 (4-5.66)
^ f(s)h(s) ds = 0 , (4-5.67)
£f
h2 4[ A+f
h
A~fl
h \
[f(x+h)-f(x)\-[f(x)-f{x-h)]
h2
f{x + h) +f{x—h)—2f(x)
(4-6.1)
h2
Here A+f = f(x+h) — f(x) is the first right-sided and A~f — f(x)—f(x—h) is the
first left-sided difference. In the case of two independent variables, the second
differences read
108 For further details, see, for example, I. G. Petrovski, Lectures on Partial Dif¬
ferential Equations, Interscience, 1954.
326 ELLIPTIC DIFFERENTIAL EQUATIONS
Mq.jq > 0 .
Mj, k ^ MtQ, 4q
109 The degree of arbitrariness in the selection can, because of the continuity of /,
be made arbitrary small, when h is chosen sufficiently small.
4-6. THE DIFFERENCE METHOD 327
и{, к = 0
at all interior points of the region S, i.e., the homogeneous difference system has
only the trivial solution. Simultaneously, the uniqueness of the solution of
the first boundary-value problem for the difference equations corresponding to
the Laplace differential equation is proved.
If we solve the difference equation, the net function results as an approxi¬
mate solution of the original problem for the Laplace differential equation.
The difference method in the present case consists of finding, instead of the
solution of a boundary-value problem for the Laplace differential equation Au —
0 , the solution of the boundary-value problem for the corresponding difference
equation. For the basis of this method one has to prove that for sufficiently
small h, the function Uh differs arbitrarily little from u, the exact solution of the
equation Au = 0. We will not go into the proof of this here.110
Instead, we shall consider in more detail the methods for the solution of dif¬
ference equations. The solution of the boundary-value problem by means of
the difference method leads to the solution of a system of algebraic equations
with many unknowns, of which there can be hundreds and even thousands.
The solution of such systems using the methods of determinant theory gives rise
to severe technical computational difficulties. On the other hand, the method
of successive approximations is essentially more suitable.
For systems of linear algebraic equations the method of successive approxi¬
mations proceeds as follows.
We write the existing system of equations in the form
Mi — f\ — {al2u2 H-
u2 — f2 — (auUi -Valnun) (4-6.4)
First we choose numbers u\,u\,--%, unn as the zero-th approximation and sub¬
stitute these in the right side of the Eq. (4-6.4); then we obtain the first approx¬
imation Ui1,-• •, u(nl). We continue this process. The (k + l)-th approxi-
then these limit values are the sought solutions of the systems (4-6.4).
The method of successive approximations is not applicable to every system
of equations. However, it can be used to solve the first boundary-value problem
for the system of difference equations Jhu — 0. The system of algebraic
equations which correspond to the equations Jhu = 0, has the form
(A) , .. (A) , (A) , (A)
(A)_Ui +1 к ■+ Ui k + 1 + Ui-i, к "Г Ui, к-1
,
Ui,k-- ,
4
hold.
We prove that the difference sequence {Vi*l} approaches zero:
lim vl*I = 0 .
Thus, we set max v\*l = An and estimate the (n + l)-th approximation, Obvi-
ously,
(»+l) о .
Vi,k Si -r An on Cll)
4
4-6. THE DIFFERENCE METHOD 329
holds, since in this case at least one of the summands in (4-6.5) is equal to zero.
From this it follows that
v\T] Ѣ (l on Ci2\
and, in general,
on Oh •
Therefore,
An+i = ex Ant
a ^ 4v ’
i.e., lim An — 0 .
n—*oo
lim Bn = 0 .
71—*°°
Consequently,
The convergence proved here is valid for an arbitrary choice of the zero-th
approximation. However, the degree of convergence (speed of convergence,
etc.) depends strongly on the choice of the zero-th approximation.
3. Electronic integrators
In recent times several mathematical machines have been used for the
solution of systems of difference equations.
M2
The construction of these machines is based
on the analogies that exist between different Ѵг
i= 1,2,3,4
П + ji + ja + jt - 0
there results
F.+ T2 + Т3 + V<
4
with
Problems
1. Find the function и which is harmonic in a circle of radius a and on the cir¬
cumference C assumes the value
Show that the formulas derived here define the solution of the problem for every
piecewise continuous function which is prescribed on the boundary. Further,
solve the problem for the special case
Ш = Ay (b-y) , f2( x) = Be os Л = Л = o.
3. Solve the equation Ди — 1 for a circle of radius a with the boundary con¬
dition и |p=0 = 0 .
4. Solve the equation Ди = Axy for a circle of radius a with center at (0,0) for
the boundary condition и |p=0 = 0 .
5. Determine the solution of the differential equation Ди — A + B(x2 — y2) in a
circular ring a ^ p ^ b if
и = Alt = 0 .
p=a P=b
u(x, 0) =
Г for
for
x < 0
x > 0 .
(a) и |v=0 = , и |p=<p0 = <7i, и |p=tt = q2 (<7i and <?2 are constants)
(b) U Itp= о — U |ір=(р0 = 0 , U |p=a = f\ф) •
10. By the difference method solve the first boundary-value problem for Ди = 0
in a rectangle 0-йх-йа,0-й.у-йЬ, with each side of the rectangle divided into
eight equal parts. The boundary conditions read:
density v = vQ. Hint: Solve Ди — 0 in the exterior and in the interior of a sphere,
and for the matching of the solutions on the spherical surface use the condition
that the potential of a simple layer is discontinuous across the surface.
13. Solve the first boundary-value problem for a bounded circular cylinder
(p g a, 0 g z si /):
(a) On the end surfaces of the cylinder let the boundary value (first or second
type) be equal to zero, while on the lateral surface it must satisfy
U |p=a = f(z) .
(b) On the lateral surface and on one of the end surfaces let the boundary value
(first or second type) be equal to zero, while on the other end surface of the
cylinder the condition is
и = f(p) ,
Ди — — /
at large distances from a body T, we usually take the value of the potential as
equal to m/R, where m is the total mass of T and R is the distance of the center
of gravity from the exterior point M. We shall now derive an exact asymptotic
4-7. APPLICATIONS TO CHAPTER 4 333
from which
1
d
1 1
r l/l + a2 — 2ap ’ a
^ n
r
= cos Ѳ . (4-7.3)
— І ccnPn{[i) , (4-7.4)
d Г n=0
pryPy{(j.)dr + j pryP2{p)dr +
(4-7.5)
The first member here is equal to m/r, where m is the total mass of the body.
This yields a first approximation for the calculation of the potential for large r.
We now calculate the following terms in (4-7.5). The integrand of the second
member is
where
My = jj ^ pXydr = Mx , M2 = ^ pyydr = My
M3= jj I pzydz = Mz
is the moment of the first order and x, у, z are the coordinates of the center of
gravity. Therefore, the second term is as small as 1 lr2. If the origin of the
coordinates is at the center of gravity (x = 0,y = 0,z = 0), then V2 = 0 .
We consider now the third term of the development. For the integrand,
we have
F3= у jy pr\P2(p)dz
The polynomial inside the braces is harmonic, since it can be written in the
form
in which each summand satisfies the Laplace differential equation. The co¬
efficients standing in the square brackets can be expressed in terms of the
moment of inertia with respect to the coordinate axes. The moment of inertia
of a body T with respect to the x axis, as is known, is given by
Correspondingly, the moments of inertia with respect to the у and z axes are,
respectively,
В = M22 + Мц , C = Мц T M22.
m 1
V + {(x2 - у2) (B-A) + (/-z2) (C-B) + (z2 - x2) {A — C)} , (4-7.7)
2r
provided the origin of the coordinates
coincides with the center of gravity and
the coordinate axes have the direction of
the principal axes of inertia.
This asymptotic representation of the
spatial potential permits us to give the
answers to a series of questions on the
inverse problem of potential theory. This
consists of determining the characteristic
quantities of a body from its potential (or
from any of the derivatives of its potentials).
That is, from the coefficients of the develop¬
ment (4-7.6), we can calculate the mass,
the coordinates of the center of gravity, and the moment of inertia.
2. Problems of electrostatics
E = —grad (p .
div E = 4rtp
we obtain
A(p = —4np .
Therefore, the potential at points of space at which electrical charges are found
satisfies the Poisson differential equation, and at points where no charges are
found it satisfies the Laplace differential equation.
1. The basic problem of electrostatics is to determine the field which is
produced by a system of charges on a given conductor. Two different formu¬
lations of this problem are possible.
(a) Given the potential of the conductor, determine the field exterior to
the conductor and the charge density on the conductor. The mathematical
formulation here reads as follows:
Determine the function <p which satisfies the Laplace differential equation
J(p = 0 everywhere exterior to the given conducting system, vanishes at in¬
finity, and on the conducting surface assumes the prescribed value <pp.
9 = 9i •
336 ELLIPTIC DIFFERENTIAL EQUATIONS
In this case, therefore, we arrive at the first boundary-value problem for the
Laplace differential equation. Its unique solvability results from the general
theory discussed above.
(b) The inverse statement of the problem is also possible, i.e., given the
total charges on the conductor, determine its potential, the charge distribution
on its surface, and the field in the interior of the conductor. The solution of
this problem leads to the following problem:
Determine a function <p which satisfies the Laplace differential equation
A(p — 0
exterior to the given conducting system, vanishes at infinity, and on the surface
of the conductor assumes a certain constant value
<P I Si = const.
<j>* da = —Апві.
J s{ dn
<p' = фі— Ѵг
satisfies the equation
/Up' = 0
(P(pr) dz
_ . dcp'
(p -^7—ida + Z
" ,d<p' ,
Zr dn <=1
From this, because of the above conditions,112 we obtain
lim (
R-oo J Tr (Pcp'fdz = 0 .
112 As a consequence of the condition <p' !« = 0, the function <p' is regular at in¬
finity (see page 270); then
dv'
——do 0 for R -> OO .
on
4-7. APPLICATIONS TO CHAPTER 4 337
r<p' = 0
or
<p' = const.
everywhere in the region under consideration. By considering the condition
<P' |oo = 0
we finally obtain
<p' = 0 ,
and hence the uniqueness is proved.
3. From the uniqueness of the solution of the boundary-value problem for
the Laplace differential equation it follows that the potential <p of an individual
conductor is directly proportional to its total charge e, i.e.,
el<p - C .
Therefore, if the charges e and e' — me lie entirely on a single conductor, then
the corresponding potentials and <p' must satisfy the equations
A<p — 0 , A<p' — 0
and the boundary conditions
d<p j 1 Г d<p' ,
-т-da — e t Ф
—4л — me
4л J 5 dn
e — C<p
exists for a single conductor. The capacity of an individual conductor is
numerically equal to the charge when the potential of the conductor is equal to
unity. If, in addition to the given conductor, still another conductor is present,
its potential depends essentially on the shape and distribution of charges on the
other conductor; indeed, for a system of conductors we have
where and <pi are the charge and the potential, respectively, of the г th
conductor. The coefficient C<* is the “partial capacity” of the fth conductor
338 ELLIPTIC DIFFERENTIAL EQUATIONS
Ca = Cki •
For simplicity we shall consider the case of two conductors; for a larger
number of conductors the proof proceeds analogously.
Given two conductors a and b, the determination of the coefficients Саь
and Cba leads to the determination of the solutions u{1) and и<2) of the equa¬
tions Aua>=0 and Auw — 0, which satisfy the boundary conditions
,(Di ,.(1> 1 _
= 0, u ISb — 1 , иШ loo
= Cab
-i:L bnd°
,(2) I ua)\Isb -— o,
= 1, u M(2,U
_J_ ( r dum,
= вьг> = Cba
47Г „ L. ьпіа
dum 2 dun)
( )
(ull)Aum -uwAuil))dr \ —u \da .
Tr J ZR+Sa-\- Sf, dn dn
The integral on the left side of this equation vanishes. By using the
boundary conditions and the conditions at infinity we therefore obtain
duw dun)
da — da = 0
dn dn
ОГ Cab — Cba r
w = —a and a=-^~
r 4na
If the total charge e0 on the sphere is given instead of the potential <p0, then
во во во
<P о = a — (r > a) .
а
<P
4ла r
Here the sphere has the capacity
C= a ,
4-7. APPLICATIONS TO CHAPTER 4 339
A(p = 0
and the conditions
ГіГг
V Vo —-—
r2 — r 1 \Г Гг
so that the spherical condenser has the capacity
ГіГг
C =
Гг —Гг
A complicated problem is represented by the calculation of the potential
of a sphere when the second sphere is not concentric to the first. This
problem is solved with the help of the method of images. Since the analy¬
tical solution is quite extensive, we shall not go into it here.113
The first sphere can be mapped onto a plane by means of an inversion.114
We shall now show that by means of a second inversion the determination
of the potential of a plane surface and of a sphere leads to the calculation
of the potential of a system of two concentric conducting spheres.
For our purposes it is sufficient to consider, instead of a plane surface
and a sphere, a straight line E and a circle К with the center 0 and radius
-H
113 See Ph. Frank and R. v. Mises, Differential and Integral Equations of Mechanics
and Physics, Vol. II, Braunschweig, 1937, p. 713.
114 Ibid.
340 ELLIPTIC DIFFERENTIAL EQUATIONS
P (Figure 70). From the point 0 we drop the perpendicular OF on the straight
line E (let l be the length of OF) and from the point F draw the tangent
FD to the circle K. With FD as a radius we let the point F trace a circle
H, which cuts the extension of the perpendicular line OF at a point S. We
choose the point S as the center of the inversion.
By this inversion, the straight line L = SO, which cuts the circle К at
the point A, and the circle H are transformed into the straight line L and H
perpendicular to each other, while the straight line E and the circle К are
transformed into the circles Ё and K.
Since orthogonality is preserved under an inversion, the circles E and К
must be orthogonal to the straight lines L and H. This is possible, how¬
ever, only when the centers of the circles E and К coincide with the inter¬
section point В of the straight lines L and H.
Consequently, the given inversion transforms the straight line and the
circle into a pair of concentric circles.
By some simple calculations we find the radii of the circles К and Ё:
VF^-d-p) _ 1
2^/12— p2 id — p) + ~[/12 — p2] У'Л/12 — p2
Thus
Vl2-P2~{l-P)
tp — 2k In ,
Гі
and for the capacity per unit of length for the cylindrical condenser we
obtain
2 In (rjri) '
dV
= 0 , (4-7.9)
dz z= 0
which asserts that the vertical component of the current density on the surface
2 = 0 is equal to zero. This is the case since the half space 2 < 0 is non¬
conducting (air).
We now consider a point-forming electrode on the boundary of the half
space at the point A. Obviously, the field potential satisfies
ІР (4-7.10)
V=
2xR ’
where R is the distance from the source, p is the specific resistance of the
medium, and I is the current strength.
This function differs from the Green’s function in an unbounded space on
the basis of condition (4-7.9) by a factor of 2.
Now if we measure, by means of a resistance bridge, the potential dif¬
ference between two points M and N which lie on a straight line through
A, we obtain
V(M) - V(N) dV Ip
Ar dr 2 л: r2
Here r is the distance of the point 0 (the center point of the segment MN)
342 ELLIPTIC DIFFERENTIAL EQUATIONS
2кг2 dV
P = ~T~ (4-7.11)
dr
If the medium is inhomogeneous then we call the quantity p determined by
(4-7.11) the apparent resistance. We denote it by pk\ pk is not constant.
We shall now consider the problem of vertical electrical probing. Let
the layers of the earth’s crust be horizontal, and let their resistance depend
only on the depth: p = p(z).
In this case, the apparent resistance is a function of the distance r = AO.
The problem raised by the vertical electrical probing thus consists of deter¬
mining the function p(z), which yields the “electric layer’’ of the medium
with respect to a known value pk(r).
We turn next to the problem of a two-layered medium. A homogeneous
layer of density l and resistance p0 is in contact with a homogeneous medium
of resistance pt.
= FxU. (4-7.12)
Also the normal components of the current density must be continuous:
1 5F0 1 dVi
(4-7.13)
Pо dz Z— I
p1 dz
At z — 0 the potential V0 must satisfy the condition (4-7.9), while at the
point A, which we chose as the origin of a cylindrical coordinate system
(<p,r,z), it must possess a singularity of the form (4-7.10):
pj 1
L0 + Vo , (4-7.14)
2k 1/z2 + r2
By separation of variables two types of solutions for V can result, which are
bounded at r = 0:
е±ХгШг).
Here /о is the Bessel function of the zero-th order (see Appendix) and X is
the separation parameter. We write the solution in the form
Vo(r, z) = &L.
2n
_L_ + Jo\°°(A0e~Xz + B0eXz)JQ(Xr)dX
V r1 + z2
where A0, B0, Alt Bx are constants. The condition (4-7.9) relates A0 and B0.
We calculate the derivative
^ (B0 — A0)J0(Xr)XdX = 0 .
ѴЛг, z) = [“А^ЫШХ ,
and
1
Jo(Xr)e XzdX (4-7.15)
Ѵгг + 22 0
(Pi - Ро)е-ш
A0
(joi + p0) — (^>i — ро)е~ш
(4-7.16)
F”(r- -££1/“+ + ■
Here we have set
Pi Pa _ fo
Pi +"i°o
We shall now transform expression (4-7.16). Since \ k\ <1, we can write
со
кё~ш
I F • в_Ш*
1 - ке~ш
and
Г oo
о X'
oo oo r oo
V0(r, л) = ІР
1 j^n g—\(2nl+z) Ipo n —\ { 2nl —z)
2i7T n = о J q
Ja{Xr)dX + -I
2л n = 1
kne JoUr)ctt . (4-7.16')
Ipo/Z,n 1
V0(r, z) — I kn + I kn- (4-7.17)
2л \и=і \/ г2 + (z — 2nl)2 -\/r2 + (z + 2nl)2 ) '
, о f knr
dr 2л \_гг n=i [г2 + (2и/)2]3/2
knr3
+ 21
1
n— 1 [r2 + (2w/)2]3/2 ]
F(f/2)3
1 + 2 І (4-7.19)
n=. i [(f/2)2 + n2]3/2
where f = r/l and /(f) are the expressions in the square brackets. For r € l,
Pk ~ Po •
In order to estimate the behavior of pk for large r, use formula (4-7.19) and
let r—>■ °o (f —>oo). The limit value of this nth term of the sum equals kn;
hence
1 + к pl + po + (Pl — Po)
= Pol1- = pi.
— kh = P°'^~L
Pl +- 7-
Po — (Pl — Po)^
4-7. APPLICATIONS TO CHAPTER 4 345
Pk\r) Ф p?\r).
Consequently, the problem—to determine the electric layer from the apparent
resistances—appears from the mathematical side to have a uniquely deter¬
mined solution.116
Problems analogous to those of electrical exploration considered above
occur in different areas of physics and technology. Electrostatic problems
of this type arise in the construction of electronic equipment, whereas heat-
theory and hydrodynamic problems arise in different areas of technology (heat
loss of buildings, filtration of water under a retaining dam, etc.).117 The
problem of the determination of a magnetic field in an inhomogeneous medium
occurs, for example, in magnetic testing of materials. To determine an
ultimate failure in a sample—for example, an air bubble below the surface—
the metallic sample is placed between the poles of a magnet, and the magnetic
field on the surface is measured. From the perturbations of the magnetic
field, the presence of the defects is observed, and if possible their extent,
the depth at which they occurred, etc., are also determined.
For the solution of such problems several methods of analogy can be
applied. Therefore the analogy between the potential fields of different
physical processes is used.118
We consider now the potential field in an inhomogeneous medium (for
example, a stationary temperature field, a magnetic field in an inhomogeneous
115 See, for example, the excellent book by A. I. Zaborovskii, Electrical Explora¬
tion, 1943.
116 A. N. Tychonoff, “On the uniqueness of solutions of problems of electrical ex¬
ploration,’’ Doklady, 69, (6): 797 (1949).
117 N. N. Pavlovskii, Theory of Motion of Ground Waters for Hydro-technical In¬
stallations and its Fundamental Applications, 1922, Ch. XIV.
118 A. V. Lukyanov, “On the electrolytic simulation of spatial problems,” Doklady, 75
(5): (1950).
346 ELLIPTIC DIFFERENTIAL EQUATIONS
, du^l _ U2duw
1 dn dn
holds in which ki and k2 are the corresponding physical constants.
On the boundaries of similar geometric regions the corresponding numeri¬
cal values of the potential or the normal derivatives of different physical
fields are prescribed. We shall assume that the physical inhomogeneities of
this region are not different geometrically and are distributed similarly. The
ratios of the physical constants (heat conductivity, magnetic permeability,
etc.) of an arbitrary pair of corresponding inhomogeneities are also equal.
Then the values of the potential of this field at the corresponding points
coincide numerically, since the potential solutions are one and the same
mathematical problem and this problem permits only a single solution.
The direct measurement of a temperature, in magnetic or other fields,
is significantly more difficult than the measurement of a current field in
an electrolytic tank. We replace it therefore in a suitable manner by meas¬
urements in such a tank. We can also choose the units in a suitable way.
div A =C (4-7.21)
The functions В, C, and / are not arbitrarily prescribed. Often the re¬
lations
div В =0 (4-7.23)
value problem with continuous boundary values is solvable for it. We shall
solve our problem in three steps. First, we shall seek a vector Ax which
satisfies the conditions
rot Ai = 0 (4-7.25)
div Ax =C . (4-7.26)
Ai — grad (p . (4-7.27)
div Л2 = 0 . (4-7.30)
If we set
A2 = rotcp, (4-7.31)
Now we require
Аф = — В. (4-7.34)
В = grad x in Gi — G .
The condition div В — 0 yields
AX = 0 in Gi — G . (4-7.36)
ГИЯ ELLIPTIC DIFFERENTIAL EQUATIONS
The boundary condition, according to (a) and (b), has the form
= BH on S (4-7.36')
dn
= о on S, (4-7.36")
an
where B„. is the limit value of Bn from the interior side of 5. For the func¬
tion we obtain the second boundary-value problem (4-7.36)—(4-7.36").
The necessary and sufficient condition for the solvability of this problem
is that
dS B„dS = 0
S I S dn
be satisfied, because the relation
div Bdz = 0
a
is satisfied.
If we set
f IHQ) dvQ
Ф(Р) (4-7.34)
Jff, r,-Q
with P P(x, y, z), Q = Q (£, rj, 0, then (4-7.34) is obviously satisfied; the con¬
dition (4-7.33) also is satisfied.
Therefore, we calculate the derivatives
dx
d
dx i S \.Л*=i S U-Л v*+S L(a,-^s - J
follows, where cos a = cos (n, x) |$, cos at = cos (n, x) |s , and n is the direction
of the exterior normal to the surface.
For Ьфх/dx we find
The vector Аг defined by (4-7.31) satisfies (4-7.29) when the vector ф satis¬
fies conditions (4-7.33) and (4-7.34).
Therefore, the sought vector A satisfies the boundary condition (4-7.22).
Now a vector Az is to be determined which satisfies the following conditions:
rot A3 = 0 (4-7.37)
inside of G
div A3 —0 (4-7.38)
A = Ai + Ai + A3.
Au = 0 (4-7.42)
и = Ui, (4-7.43)
where S; designates the surface of the z'th conductor. If the field can be
regarded as a planar field which does not change along the z axis, then Eq.
(4-7.42) and the boundary conditions assume the form
350 ELLIPTIC DIFFERENTIAL EQUATIONS
u\a< = Ui (4-7.45)
vx = uy , vy = — ux (4-7.47)
and
From the boundary condition (4-7.45) it follows that the imaginary part of
/(2) on the contour C, is constant.
From the conditions (4-7.47) we know that
represents the equation of the lines of force,119 and the lines of equipotential
can be described by the equation
because of (4-7.48).
To solve the problem under consideration it is sufficient, therefore, to
find the conformal mapping of the 2 plane (2 = x + iy) on the w plane (w =
и + iv) for which the boundaries of the conductor are mapped into the straight
lines
и = const.
or
Im w = const.
и = u(x, y) = Im/(2).
du du
Ey=- (4-7.51)
dx 7 dy
and the calculation of the density of the surface charges along the 2 axis:
1 ✓bJTIF; 1 / du
+
/ du у
dx \dy) '
119 The equation of the lines of force results from dxlux = dy/uy. If one replaces ux
and uy according to (4-7.47) by — vy and vx, respectively, then we obtain vxdx + vydy =
dv = 0, or v(x, y) — const.
4-7. APPLICATIONS TO CHAPTER 4 351
(4-7.52)
IV = <p + Іф (4-7.53)
2= 1t(W + eW) ’
carries the lines z =± d/2, x < 0, which cuts the
z plane (z = x + iy) into the region | ф | ^ к of the 7T
-Jr2"», (4-7.54)
x = <p + ev cos ф)
Лк
(4-7.56)
у - -ту-іф + e9 sin ф) .
Zk
x и У (4-7.57)
120 See Ph. Frank and R. v. Mises, Ch. XV, §5, op. cit. fn. 113.
352 ELLIPTIC DIFFERENTIAL EQUATIONS
i.e., inside the condenser at large distances from the boundary the field is
planar; but as <p—>co
i.e., outside the condenser, at large distances from the boundary, the equi-
potential lines are circles.
If instead of w the complex potential / = (u0l2n)w is introduced so that
w = {2n/ut)f, then the relationship between z and f(z) is given by
= d(-£- + -Ц?2л//“Л .
V Mo An )
From this it follows that
dz_d_
1 + e2*//u°^
df M0
on the other hand, for f = (u0/2n)[<p ± ni) we obtain
dz Mo
—(1 - ev) or f\z)
df M0 d( l-e*) '
1
a = (4.7.59)
An And\l-ev\ ‘
OO
i
a
And ’
and as <p OO
1
And# ’
is that it can only be used in plane problems which lead to the two-dimen¬
sional Laplace differential equation Аги = 0.
Vn(x, y, z) = vx(x, y, z) cos (n, x) + Vy(x, y, z) cos {n, y) + vz(x, y, z) cos (n, z)
for all points on the surface of the body must be equal to the normal com¬
ponent of the velocity of motion of the body. If the body does not move,
then the boundary condition on the surface of the body assumes the simplest
form
Vn - o .
dtp
= 0 in case the body does not move;
dn S
dtp
— un in case the body moves with the velocity u.
dn s
A<p = 0 .
Therefore the determination of the potential for the streaming about a solid
body in an incompressible ideal fluid is reduced to the solution of the Laplace
differential equation
A<p = 0,
dtp
Un
dn s
must hold, i.e., the second boundary-value problem for the Laplace differential
equation must be solved.
If the motion is planar, then the solution of the problem can be obtained
by using function theory.
For planar motion of an incompressible fluid the continuity equation is:
dx _ dy
vx Vy
in the form
дф дф
Vx = ~ , Vy = - .
ay ox
From Eq. (4-7.60) it then follows that the left side of expression (4-7.61) is
the total differential of the function ф\
vxdy — Vydx = dф .
ф(х, у) = С
Аф = 0.
dip _ дф dip _ дф
dx dy ’ dy dx
i.e., the functions <p and ф satisfy the Cauchy-Riemann differential equations.
Consequently, the complex function
is an analytic function.
Therefore, every irrotational (rot v = 0) planar motion of a fluid has a
corresponding definite complex analytic function, and conversely every analytic
function corresponds to a definite form of fluid motion, i.e., two motions
correspond to an analytic function, since the roles of the functions <p and ф
can be interchanged.
2. In conclusion, we shall treat a concrete example of the application of
the theory of functions to the solution of problems for the streaming of a
body—here, a circular cylinder—in a planar fluid.
On a circular cylinder at rest of radius r = a, let a planar fluid stream
impinge which at infinity has the constant velocity u. In the case of a
static fluid, we would consider the motion of the cylinder moving with the
constant velocity и with respect to the stationary fluid.
As a reference system, we choose a rectangular coordinate system (x,y,z)
attached to the cylinder in which the x axis is chosen parallel to the direc¬
tion of the velocity of the cylinder.
On the surface of this moving body, the boundary condition
4-7. APPLICATIONS TO CHAPTER 4 355
дФ dy
— u~r~
ds ds
ф — uy + C .
Дф = 0
div _ dф + ■ dф
vx — ivy
dz dy dx
w = Ci In z —
Now we set
Ck — Лк + iBk
ф = ua sin Ѳ + C ,
1 . a
IV = ——Г In z — и—
2лг z
/ л п а
T~i 2
The first term in the expression for w denotes the circulation around the
356 ELLIPTIC DIFFERENTIAL EQUATIONS
a2
iv = — и— .
z
For the flow about a static cylinder, which at infinity has the velocity u, the
complex potential has the form
. ua . I
w = uz A-V ——r In z .
2 T~t
.
z Zki
On the plate a constant planar flow occurs which at infinity has a velocity
c with components и and v. By means of the analytic function
2=І(с+т)=/ю
we obtain a reversible single-valued mapping of the region exterior to the
plate in the z plane onto the region exterior to the unit circle in the C plane.
Here the point z — oo corresponds to the point С = °°» and
We shall now investigate the nature of the mapping of the condition at in¬
finity. For the complex potential
w(z) = <p + іф
we have
(dw\ .
—Г- = и — IV = Voo ,
\ (l Z J z oo
r r, m dw dw dz
MO = w[f{z)) , .
aC az a£
/ dw\
kVo k= —
V dZ Л=с * 2
The fictitious flow therefore represents the flow about a cylinder of unit
radius and at infinity has the complex velocity kvFor such a motion, the
complex potential has the form
^ _ г + т/z2 — a2 l/z2
Г 2 + t/z2 — Й2
w(z) — uz — г/t/z2 — й2 in
27гг
d<p = 0
—>— — u cos Ѳ
072 |r=a
<p = Ли grad— .
r
a3ur
7. Biharmonic equation
(4-7.62)
or + a2JJu = 0. (4-7.63)
и — 0 and (4-7.64)
dn
must be satisfied.
Moreover, the function и must satisfy the initial conditions
If external forces act on the plate which are distributed with the density
f(x,y), then the deformation of a plate whose boundaries are fixed can be
described by the equation
ДДи = /, (4-7.66)
и — 0 and = 0 . (4-7.64)
on
The equation
ДДи = 0 (4-7.66')
is called the biharmonic equation and its solutions biharmonic functions. These
functions must possess continuous derivatives up to and including the fourth
order.
The fundamental boundary-value problem for the biharmonic equation
reads:
Determine a function u{x, y), which including its first derivatives in the
closed region 5 4- C is continuous, in 5 possesses continuous derivatives up
to the fourth order, in S satisfies Eq. (4-7.66) or (4-7.66'), and on the boundary
C satisfies the boundary conditions
du
и Ip = g(s), = h(s) , (4-7.67)
dn о
where g{s) and h(s) are continuous functions of the arc length s on C.
For the solution of the above-formulated initial value problem (4-7.63)—
(4-7.65) by means of separation of variables, we usually set
If we introduce this expression into (4-7.63) and separate the variables, then
for the determination of the eigenvalues we arrive at the equation
ДДѵ — Лѵ — 0 (4-7.69)
v = 0 , —=0 on C. (4-7.70)
on
ДДи = 0
V = Ui — U2 ,
which obviously also satisfies the biharmonic Eq. (4-7.66) and the homogeneous
boundary conditions
( {AvfdS = 0
Ja
and hence
Дѵ = 0.
v = 0 and Ui = u2,
1. e., the biharmonic functions are uniquely determined by the boundary con¬
ditions (4-7.64).
2. Representation of a Biharmonic Function by Harmonic Functions. We first
prove the theorem:
Theorem. If ux and u2 are two harmonic functions in a region G, then и =
xu,. + М2 is a biharmonic function in G.
For the proof we use the identity
With
<p = X , ф = Ui
it follows that
JJ(xui + m2) = 0 .
If the region G is so chosen that each straight line parallel to the x axis
cuts the boundary of G at two points at most, then the converse of the
above-proven theorem also is valid.
For every biharmonic function и in the region G, two harmonic functions
Mi and m2 can be found such that
4-7. APPLICATIONS TO CHAPTER 4 361
u = xu, + иг.
For the proof of this statement it is sufficient obviously to consider a
function и I which satisfies the two conditions
Au, = 0 (4-7.73)
and
A(u — xui) = 0 . (4-7.74)
From condition (4-7.74) and formula (4-7.72) we find
Au = A(xui) = # (4-7.75)
bx
Eq. (4-7.75) is satisfied by the function
Г 1
u,(x, y) = \ —Au(s, y)dZ .
Jx0
Since
Au, = —r = — v(y),
л t
by
holds, and set
u, = u, + u, .
This function then obviously satisfies the conditions (4-7.73) and (4-7.74), which
was to be proved.
We shall now give another representation of a biharmonic function. Thus
we assume that the origin of coordinates lies inside of the region G and each
line from the origin touches the boundary of G at only one point. Under
these assumptions the following is valid: Every biharmonic function и in G
can be represented by two harmonic functions u, and иг in the form
и = (гг — r\)u, + иг. (4-7.76)
Here гг = хг + уг, and r0 is a given constant.
The proof here proceeds analogously to the preceding one by means of
the identity (4-7.71) and the relations
bu, au, bx . bu, by
Ar2 = 4 , — —- —+ -T—
or bx or by or
3. Solution of the Biharmrmic Equation for the Circle. Given a circle of radius
r0 about the origin of coordinates, we seek a biharmonic function which
362 ELLIPTIC DIFFERENTIAL EQUATIONS
satisfies the boundary conditions (4-7.67) for r = r0. As was shown above,
the sought function can be written in the form
и — (r2 — rl)ui + u2, (4-7.76)
where и, and u2 are harmonic functions. From the boundary conditions we
find
u2\ r=r0 = g. (4-7.77)
From this we recognize that u2 is a solution of the first boundary problem
for the Laplace differential equation. Therefore, u2 can be represented by
the Poisson integral
_ 1 f2* (rl-r2)gda
(4-7.78)
2л Jo г2 + Го — 2rr0 cos(« — Ѳ)
From the second boundary condition we obtain
9 ru (4-7.80)
r0 dr
also satisfies the Laplace differential equation and therefore can be expressed
by the Poisson integral
и 1 (r« _ M 1 Г - hda
2лг0 L
2 Jo rz + rl — 2rr0 cos (a — 0)
f2* g[r0 — r cos (a — 0)]da ~|
Jo [r2 + rl — 2 rr0 cos (a — 0)f J '
APPENDIX
In the following pages we list tables of some special functions used for
the solution of boundary-value problems of mathematical physics. These tables
consist of a compilation of the simple properties of the functions considered.
Ф(г)
2 f -.2
da
"v^ Jo*
For small values of z, it has the series development
3 5
z z
m = v~Jz 1!3 + 2!5
1 - 0(Z) = Д» е~*гda
V n )z
by means of a = z + (/9/2z), then we obtain
1. e~z
e-*2 f°° g-p-tp2/4*2, dp .
1 - Ф(г) =
T/tt Z Jo
Consequently
,. 1 - 0(Z) ,
lim-r——r
z-*oo 1 £, z
= i;
~[/ n z
ье.,
1 е~*г
1 -Ф(2)~4=— .
Vл z
If we develop e p2/4z2 in the series
в-*/.* = у lull! £1
»=o (2z)2n n\ ’
multiply this series with e~p and integrate from zero to infinity, then we
obtain the divergent series
“ (-1)" (2м)! = “ (-1Г(м + 1)(m + 2)---2m
nt0(2z)2n n\ nt0 (2z)2"
364 APPENDIX
CYLINDRICAL FUNCTIONS
1. Bessel functions
2. Neumann functions
7Г
№) =-—
nx jv'w = -/Icos(^t)+--’
+ у /і(х) (in у + с) + • • •
(n > 1)
(C — 0.577215... is the Euler constant.)
3. Hankel functions
I Ax) = J0(ix) +■
ш = /У7
= i , (x/2f (xffl
1! + (2!)2
I Ax) = — iJAix) №) = /2Хе-
_ x Г1 (*/2)2 (ХІ2У A
' 2 L Г 1-2 1-2-2-3 J
Iy(x) = (-ifMix) +■
ад = /гБ
1
2 / Г(ѵ + 1)
/ О
/ілМ -л — sin л:; J-і/Ах) = — cos x ;
г хх t 7ГХ
г , ч
/з /Ах) =
НЕ( sin*
— (-cos X /_,дМ = 1/І(-^-5іп.Л
jtt\ л: Г 7ГХ V X /
6. Recursion formulas
d \Mx) 1 _ _ At i(x)
-y-[*7v(*)] - x7v-i(*) and
ax rfx L xv J x
2v
Jy-Ax) + Jy+Ax) = — Jy(x) ,
X
Analogous formulas are also valid for other cylindrical functions with
real arguments.
For functions with an imaginary argument we have:
366 APPENDIX
Mx)Nl(x) - Ату(х)Л(х) = — ;
nx
In particular,
2_
Jo(x)NM) - N0{x)JM =
nx
8. Integral formulas
1 f*
Jn(x) = — \ e~ix sin ѵ*іпіе dip
2л J —*
In particular,
K0(x) =
j %_х7.(^) dX = v -г (г > 0) ,
+ 2‘
<p(z) = -|=Г
Vn Jo
da 0 ^ z ^ 2.8
2.4 0.9993
2.5 0.9996
2.6 0.9998
2.7 0.9999
2.8 0.9999
368 APPENDIX
TABLE 3
Albrecht, R., and Hochmuth, H. Exercises in Higher Mathematics. Part III. R. Olden-
bourg: Munich, 1956.
Bateman, H. Partial Differential Equations of Mathematical Physics. Dover Publica¬
tions: New York, 1944.
Baule, B. “Mathematics for Scientists and Engineers,” in Partial Differential Equations.
Volume VI, 5th edition. S. Hirzel: Leipzig, 1955.
Bergmann, S., and Schiffer, M. Kernel Functions and Differential Equations in Mathe¬
matical Physics. Academic Press, Inc.: New York, 1953.
Bernstein, D. L. Existence Theorems in Partial Differential Equations. 2nd printing.
Princeton University Press: Princeton, 1951.
Bieberbach. L. Introduction to the Theory of Differential Equations in the Real Domain.
Springer Verlag: Berlin, 1956.
-. Theory of Differential Equations. 3rd printing. Springer Verlag: Berlin, 1930.
Borgnis, F. E. and Papas, С. H. Boundary-value Problems in Microwave Physics-
Springer Verlag: Berlin, 1955.
Budak, В. M., Samarskii, A. A., and Tychonoff, A. N. A Collection of Problems for
Mathematical Physics. Gostechisdat: Moscow, 1956.
Collatz, L. Numerical Treatment of Differential Equations. 2nd printing. Springer
Verlag: Berlin, 1955.
-. Eigenvalue Problems with Technical Applications. Academic Verlagsgesells-
chaft: Leipzig, 1949.
-. “Numerical and Graphical Methods,” in Handbook of Physics. Volume II,
Mathematical Methods, Part II. Springer Verlag: Berlin, 1955.
First Conference on Equations with Partial Derivatives. Louvain, 1953 (CBRM). George
Thone: Liege, 1954.
Second Conference on Equations with Partial Derivatives. Bruxelles, 1954 (CBRM).
George Thone: Liege, 1955.
Contributions to the Theory of Partial Differential Equations. Princeton University
Press: Princeton, 1954.
International Conference on Linear Equations with Partial Derivatives. Trieste, 1954;
Edizioni Cremonese, Rome, 1955.
Courant, R., and Hilbert, D. Methods of Mathematical Physics. Volume I, 2nd printing;
Volume II, 1st printing. Springer, Verlag: Berlin, 1931 and 1937, respectively.
Courant, R., and Schiffer, M. Dirichlet's Principle: Conformal Mapping and Minimal
Surfaces. Interscience Publishers, Inc.: New York, 1950.
Duff, G. F. D. Partial Differential Equations. University of Toronto Press: Toronto,
1956.
Duschek, A. Lectures on Higher Mathematics. Volume III. Springer Verlag: Vienna, 1953.
372 LITERATURE REFERENCES OF THE EDITOR
Frank, Ph., and R. v. Mises. Differential and Integral Equations of Mechanics and
Physics. Part I, 2nd printing; Part II, 2nd printing. F. Vieweg: Braunschweig,
1930 and 1935, respectively.
Ford, L. R. Differential Equations. 2nd edition. McGraw-Hill Book Co., Inc.: New York,
1955.
Forsyth, A. R., and Jacobsthal, W. Textbook of Differential Equations. (Translated
from the English.) 2nd edition. F. Vieweg: Braunschweig, 1912.
Freda, H. The Method of Characteristics for the Integration of Linear Hyperbolic
Partial Differential Equations. Mem. sci. math., fasc 84. Gauthier-Villars: Paris,
1937.
Friedman, B. Principles and Techniques of Applied Mathematics. John Wiley & Sons,
Inc.: New York, 1956.
Geronimus, J. L., and Steklov, V. A. Integration of the Differential Equations of
Mathematical Physics. (Translated from the Russian.) Verlag Technik: Berlin,
1954.
Godeaux, L. Mathematical Analysis. Volume III. Sciences and Letters: Liege, 1947.
Gosse, R. The Method of Darboux for the Equation s = f(x, y, z, p, q). Mem. sci.
math., fasc 12. Gauthier-Villars: Paris, 1926.
Goursat, Ed. Lessons on the Integration of Equations with Partial Derivatives of the
Second Order in Two Independent Variables. 2 vols. A. Hermann: Paris, 1896
and 1898, respectively.
Guldberg, A. “Partial and Total Differential Equations,” in E. Pascal, Repertory of
Higher Mathematics. Volume I, Part II, 2nd printing. B. G. Teubner: Leipzig,
1927.
Gunter, N. M., and Kusmin, R. O. Collection of problems in Higher Mathematics. Volume
II. (Translated from the Russian.) VEB Deutscher Verlag der Wissenschaften:
Berlin, 1957.
Hadamard, J. Lectures on Cauchy's Problem in Linear Partial Differential Equations.
Reprinting of the 1st edition of 1923. Dover Publications: New York, 1952.
-. Cauchy's Problem and Linear Hyperbolic Partial Differential Equations.
(Revised translation of the preceding title.) Hermann et Cie: Paris, 1932.
Hahn, H., Lichtenstein, L., and Lense, J. “The theory of Integral Equations and Func¬
tions of Infinitely Many Variables and Their Application to Boundary-value
Problems for Ordinary and Partial Differential Equations,” in E. Pascal, Repertory
of Higher Mathematics. Volume I, Part III, 2nd edition. B. G. Teubner: Leipzig,
1929.
Heilbronn, G. Integration of Equations with Partial Derivatives of the Second Order
by the Method of Drach. Mem. sci. math., fasc. 129. Gauthier-Villars: Paris,
1955.
Hildebrand, F. B. Methods of Applied Mathematics. Prentice-Hall, Inc.: New York,
1952.
Hoheisel, G. Collection of Problems in Ordinary and Partial Differential Equations.
2nd edition. W. de Gruyter: Berlin, 1952.
-. Partial Differential Equations. 3rd edition. W. de Gruyter: Berlin, 1953.
Horn, J. Partial Differential Equations. 4th edition. W. de Gruyter: Berlin, 1949.
Hort, W., and Thoma, A. Differential Equations of Technology and Physics. 6th edi¬
tion. J. A. Barth: Leipzig, 1954.
Janet, M. Lessons on Systems of Equations with Partial Derivatives. Gauthier-Villars:
Paris, 1929.
-. Systems of Equations with Partial Derivatives. Mem. sci. math., fasc. 21.
Gauthier-Villars: Paris, 1927.
LITERATURE ON THE THEORY OF PARTIAL DIFFERENTIAL EQUATIONS OF HIGHER ORDER 373
John, F. Plane Waves and Spherical Means Applied to Partial Differential Equations.
Interscience Publishers: New York, 1955.
Julia, G. Exercises in Analysis. Volume IV. Gauthier-Villars: Paris, 1935.
Катке, E. Differential Equations of Real Functions. 3rd edition. Akademische Verlags-
gesellschaft: Leipzig, 1956.
Kantorowitsch, L. W., and Krylov, W. I. Approximate Methods of Higher Analysis.
(Translated from the Russian.) VEB Deutscher Verlag der Wissenschaften:
Berlin, 1956.
Krylov, A. N. On Certain Differential Equations of Mathematical Physics. 5th edition.
Gostechisdat: Moscow, 1950.
Kupradse, W. D. Boundary-value Problems of Vibration Theory and Integral Equa¬
tions. (Translated from the Russian.) VEB Deutscher Verlag der Wissenschaften:
Berlin, 1956.
Ladyzhenskaya, 0. A. Mixed Problems of the Theory of Hyperbolic Partial Differential
Equations. Gostechisdat: Moscow, 1953.
Lense, J. “Partial Differential Equations,” in Handbook of Physics. Volume I, Mathe¬
matical Methods, Part I. Springer Verlag: Berlin, 1956.
Leray, J. Hyperbolic Differential Equations. Institute for Advanced Study: Princeton,
1955.
Lewin, W. I., and Grosberg, J. I. Differential Equations of Mathematical Physics.
(Translated from the Russian.) Verlag Technik: Berlin, 1952.
Lichtenstein, L. “New Developments in the Theory of Partial Differential Equations of
the Second Order of the Elliptic Type,” in Encyclopedia of Mathematical
Sciences. Volume II, Part III, Second half. B. G. Teubner: Leipzig, 1924.
Love, A. E. H. Textbook of Elasticity. (Translated from the English.) B. G. Teubner:
Leipzig, 1907.
Madelung, E. Mathematical Aids for Physicists. 6th edition. Springer Verlag: Berlin,
1954.
Malkin, I. G. The Methods of Liapunov and Poincare in the Theory of Non-Linear
Vibrations. Gostechisdat: Moscow, 1949.
Margenau, H., and Murphy, G. M. The Mathematics of Physics and Chemistry. 2nd
edition. D. van Nostrand Co., Inc.: New York, 1956.
Miller, F. H. Partial Differential Equations. 6th reprint. John Wiley & Sons, Inc.:
New York, 1953.
Miller, K. S. Partial Differential Equations in Engineering Problems. Prentice-Hall,
Inc.: New York, 1953.
Miller, N. A First Course of Differential Equations. University Press: Oxford, 1935.
Milne, W. E. Numerical Solutions of Differential Equations. John Wiley & Sons, Inc.:
New York, 1953.
Miranda, C. Equations with Partial Derivatives of the Elliptic Type. Springer Verlag:
Berlin, 1955.
Morse, P. M., and Feshbach, H. Methods of Theoretical Physics. 2 vols. McGraw-Hill
Book Co., Inc.: New York, 1953.
Muskhelishvili, N. I. Some Fundamental Problems of the Theory of Elasticity. 4th edi¬
tion. Akademie-Verlag: Moscow, 1954. (This work contains a complete list of
references which primarily give an insight into the Soviet literature in this field.)
Page, С. H. Physical Mathematics. D. van Nostrand Co., Inc.: New York, 1955.
Panov D. J. A Collection of Formulas for the Numerical Treatment of Partial Dif¬
ferential Equations by Difference Methods. (Translated from the Russian.)
Akademie-Verlag: Berlin, 1955.
374 LITERATURE REFERENCES OF THE EDITOR
In addition to the books already enumerated, the theory is explained in the following
works.
Burkhardt, H., and Meyer, F., “Potential Theory,” in Encyclopedia of Mathematical
Sciences. Volume II, Part I, first Half. B. G. Teubner: Leipzig, 1900.
Gunter, N. M. Potential Theory and its Application to the Basic Problems of Mathe¬
matical Physics. (Translated from the Russian.) B. G. Teubner: Leipzig, 1957.
Kellogg, O. D. Foundations of Potential Theory. Springer, Verlag: Berlin, 1929.
Korn, A. Five Treatises on the Potential Theory. F. Dummler: Berlin, 1902.
Liapunov, A. M. Studies in Potential Theory. Gostechisdat: Moscow, 1949.
Lichtenstein, L. “New Developments in Potential Theory,” in Encyclopedia of Mathe¬
matical Sciences. Volume II, Part III, first Half. B. G. Teubner: Leipzig, 1909.
Neumann, E. R. Studies of the Methods of C. Neumann and G. Robin for the Solution
of the Two Boundary-value Problems of Potential Theory. B. G. Teubner:
Leipzig, 1905.
Plemelj, I. Potential Theoretic Investigations. B. G. Teubner: Leipzig, 1911.
Rothe, R. Higher Mathematics. Part VII by W. Schmeidler (also treats potential theory.)
B. G. Teubner: Stuttgart, 1956.
Sternberg, W. Potential Theory. 2 vols. W. de Gruyter & Co.: Berlin, 1926.
Sternberg, W., and Smith, Turner L. The Theory of Potential and Spherical Harmonics.
3rd edition. University of Toronto Press: Toronto, 1952.
Besides the references given above, special functions are also discussed in the fol¬
lowing works. Tables and Tabular work will not be included in the reference list. We
have used L. M. Ryshik, and I. S. Gradstein: Tables (Translated from the Russian),
VEB Deutscher Verlag der Wissenschaften, Berlin, 1957, as a reference for special values
of these functions. This book contains a list of references which gives further references
to tables.
Angelesco, A. On Some Polynomials Generalizing the Polynomials of Legendre and
Hermite and the Approximate Calculation of Multiple Integrals. Gauthier-
Villars: Paris, 1916.
Appell, P., and de Feriet, J. Kampe. Hypergeometric and Hyperspherical Functions'.
Hermite Polynomials. Gauthier-Villars: Paris, 1926.
Bateman, H. Higher Transcendental Functions. 2 vols. McGraw-Hill Book Co., Inc.:
New York, 1953 and 1955.
376 LITERATURE REFERENCES OF THE EDITOR
Bowman, F. Introduction to Bessel Functions. Longmans, Green and Co.: New York,
1938.
Goudet, G. Bessel Functions and Their Application in Physics. 2nd edition. Masson
et Cie.: Paris, 1954.
Graf, J. H., and Gubler, H. Introduction to the Theory of Bessel Functions. 2 Parts,
2nd edition. K. J. Wyss: Bern, 1900.
Gray, A., Mathews, G. B., and MacRobert, T. M. A Treatise on Bessel Functions and
Their Applications to Physics. 2nd edition. Macmillan: London, 1922.
Hilb, E. “Spherical Functions, Bessel and Related Functions,” in E. Pascal, Repertory
of Higher Mathematics. Volume III, Part III, 2nd edition. B. G. Teubner: Leipzig,
1929.
Lagrange, R. Polynomials and Legendre Functions. Mem. sci. math., fasc. 97. Gauthier -
Villars: Paris, 1939.
Lebedev, N. N. Special Functions and Their Applications. Gostechisdat: Moscow,
1953.
Lense, J. Spherical Functions. Akademische Verlagsgesellschaft: Leipzig, 1950.
Magnus, W., and Oberhettinger, F. Formulas and Theorems for the Special Functions
of Mathematical Physics. 2nd edition. Springer Verlag: Berlin, 1948.
McLachlan, N. W. Bessel Functions for Engineers. 2nd edition. Clarendon Press:
Oxford, 1955.
Meixner, J. “Special Functions of Mathematical Physics,” in Handbook of Physics.
Volume I, Mathematical Methods, Part I. Springer Verlag: Berlin, 1956.
Nielsen, N. Handbook of the Theory of Cylindrical Functions. B. G. Teubner: Leipzig,
1904.
Petiau, G. The Theory of Bessel Functions. CNRS, Masson et Cie: Paris, 1955.
Prasad, G. A Treatise on Spherical Harmonics and the Functions of Bessel and Lame.
2 vols. Mahamandal Press: Benares, 1930 and 1932.
Relton. F. E. Applied Bessel Functions. Blackie & Son, Ltd.: London, 1946.
Schafheitlin, P. The Theory of Bessel Functions. B. G. Teubner, Leipzig, 1908.
Smirnov, V. I. Textbook of Higher Mathematics (Translated from the Russian.) Part III,
Section 2, Chapter VI. VEB Deutscher Verlag der Wissenschaften: Berlin, 1955.
Sonin, N. J. Investigations of Cylindrical Functions and Special Polynomials. Gostechis¬
dat: Moscow, 1954.
Wangerin, A. “Theory of Spherical Functions and Related Functions, in Particular the
Lame and Bessel Functions,” in Encyclopedia of Mathematical Sciences. Volume
II, Part I, Second Half. B. G. Teubner: Leipzig, 1904.
Watson, G. N. A Treatise on the Theory of Bessel Functions. University Press: Cam¬
bridge, 1922. (This book contains an enlarged literature list in which all publi¬
cations on the Bessel functions prior to 1922 are given.)
Weyrich, R. Cylindrical Functions and Their Applications. B. G. Teubner: Leipzig,
1937.
Whittaker, E. T. and Watson, G. N. A Course of Modern Analysis. 4th edition. Uni¬
versity Press: Cambridge, 1952.
INDEX
«rm
ЛШ'З' 1976
MflPg i«y
t
L
btB A
w V- >»
63 0 27475 3
TRENT UNIVERSITY
иТікѢоШ.ѵ>- -
40337
QA Tikhonov, A N
401 Partial differential
T513 equations of mathematical
v.l physics
Trent
University