16
Partial differential equations

In Chapters 14 and 15 we discussed ordinary differential equations and their solutions. These are equations that contain a dependent variable y, which is a function of a single variable x, and derivatives of y with respect to x. In this chapter, we extend the discussion to similar equations that involve functions of two or more variables x1x2, …, xn. These are called partial differential equations (PDEs) because the functional form analogous to (14.2) in general contains partial differentials with respect to several variables, including mixed derivatives. If we consider a function u of just two variables x1 = x and x2 = y, then examples of partial differential equations are

and

where f(x, y) is an arbitrary function of x and y.

By analogy with the definitions in Chapter 14, the degree of the equation is defined as the power to which the highest order derivative is raised after the equation is rationalised, if necessary; and the order of a partial differential equation is the order of the highest derivative in the equation. Thus (16.1a) is first-order and (16.1b) and (16.1c) are second-order equations. In addition, all three equations are linear, in the sense that they contain only u and its derivatives to first degree; products between them are absent. Non-linear equations, such as

numbered Display Equation

will not be discussed.

A linear equation is said to be homogeneous if each term contains either the dependent variable or one of its derivatives, so that if u is a solution, so is λu, where λ is a constant. Thus (16.1a) and (16.1c) are homogeneous, while (16.1b) is inhomogeneous. An important property of linear homogeneous equations is that, if u1 and u2 are solutions, then so is any arbitrary linear combination of them (αu1 + βu2), where α and β are arbitrary constants. This is called the superposition principle and is widely used in finding the general solution of some PDEs using a method known as ‘separation of variables’ that we will discuss in Sections 16.2 and 16.3.

One important difference between PDEs and ODEs needs to be mentioned. As discussed in Chapter 14, a linear combination of two or more solutions for an nth order linear homogeneous ODE will in general have n arbitrary constants, which can only be determined by suitable boundary conditions, for example, the value of the function y(x) can be specified for n values of x to determine the constants. The solution of a PDE, however, usually contains a number of arbitrary functions. For example, it is easily verified by direct substitution that the simple first-order PDE

numbered Display Equation

has solutions of the form u(x, y) = y f(x), where f(x) is an arbitrary function of x. Similarly, the simple second-order equation

has a general solution of the form

where f(x) and g(y) are arbitrary differentiable functions, as may be found by successive integration with respect to x (at fixed y) and y (at fixed x). However, unlike the analogous situation for ODEs, it does not in general follow that an nth order PDE always contains n arbitrary functions, but it is true for linear equations with constant coefficients, which is the class of equations discussed in the rest of this chapter. As for ODEs, these arbitrary functions must be determined by imposing boundary conditions. In this case, these will have to take the form of specifying u along a continuum of points, for example along a line in the (x, y) or (x, t) plane, but the appropriate form of the boundary conditions depends on the type of PDE. This will be discussed in more detail in Section 16.6.

16.1 Some important PDEs in physics

Partial differential equations are in general far more difficult to solve than ordinary differential equations, and except for certain special types of equation, no general method of solution exists. In this chapter we will concentrate on discussing some specific second-order equations, since these include many of the most important equations in physics. These include the following, where u is a finite physical quantity that can depend on three space co-ordinates x, y, z and the time t.

  1. The wave equation

    In mechanics, u could be a displacement of a vibrating medium, or in electromagnetism, the component of an electromagnetic wave, etc., and is the speed of propagation of the associated wave.

  2. The diffusion equation

    This describes the diffusion of material particles, where ρ is the density of diffusing particles and the constant κ is called the diffusivity. It also describes heat conduction in a region that contains no heat sources or sinks, where it is referred to as the heat conduction equation. In this case, u is the temperature and κ is called the thermal diffusivity. It is given by , where k is the thermal conductivity, s is the specific heat capacity, and ρ is the density of the material.

  3. Laplace's equation

    (16.5)numbered Display Equation

    Here u could be, for example, a steady-state temperature distribution, or the electrostatic or gravitational potential in free space.

  4. Poisson's equation

    (16.6)numbered Display Equation

    The quantity u describes the same quantities as in Laplace's equation, but now in a region containing an appropriate source ρ(r). For the electrostatic potential u = φ, Poisson's equation takes the form (12.60), so that the force is proportional to the electric charge density; for the gravitational potential ψ, (12.58) and (12.59) give

    numbered Display Equation

    where ρm(r) is the mass density and G is Newton's gravitational constant.

  5. Schrödinger's equation

    This is the equation of motion for a particle of mass m in a potential V(r) in non-relativistic quantum mechanics, where ℏ is Planck's constant h divided by 2π. The quantity u is the Schrödinger wave function and is usually complex.

    An important feature of all these equations is that the spatial derivatives enter via the scalar differential operator

    numbered Display Equation

    which reflects the rotational invariance of the laws of physics. In particular, they do not include mixed partial derivatives of the forms , etc. Because of this, and subject to the forms of ρ(r) and V(r), these equations can be solved by the method of separation of variables, in which the problem is converted into one of solving a set of ordinary differential equations, each in a single variable. This method is discussed in the next two sections; other methods of solution will be treated in later sections.

16.2 Separation of variables: Cartesian co-ordinates

In this method, given a PDE for a function u(x, y, z, t), we seek solutions of the form

in which u is given as the product of functions of single variables. Here we have used the convention of denoting the single-variable functions by an upper-case letter and its argument by the corresponding lower-case letter. There could also be common parameters that occur in each function, but each depends on only one independent variable. We then try to rewrite the original PDE in the form of four separate ODEs for each of the functions XYZT using a procedure to be explained below. If this is possible, then the original PDE is said to be separable, and one can seek solutions to the individual ODEs using the methods discussed in Chapters 14 and 15. If it is not possible, then the equation is not separable, and other methods must be used.

If a function can be written in the form (16.8), it is also said to be separable. Thus xy2z3tan (at) is separable in all four variables, and is said to be completely separable; whereas (x2 + y− 2)z3sin (θt) is only separable in z and t and is therefore partially separable; but xz + ayt cannot be separated in any variable and so is inseparable. Obviously the individual solutions initially found by separation of variables are, by construction, completely separable functions. However, this is not as restrictive as it might seem, since, as we shall see, the method leads to many such solutions, usually an infinite number, and the general solution, which is rarely separable, can usually be constructed from them. This is particularly easy in the important case of linear homogeneous equations, when any linear combination of a given set of solutions is itself a solution, and for such equations it is also known as the Fourier method of solution.

In this section, we shall introduce the method using Cartesian co-ordinates, leaving the important but more complicated case of polar co-ordinates until Section 16.3. In both cases, we will use the physical equations introduced in the previous section to do this, applying boundary conditions and constraints appropriate to the function u being a physical quantity.

16.2.1 The wave equation in one spatial dimension

To illustrate the general method, we will solve the wave equation (16.3a) for a single spatial variable x, when it reduces to

This equation could describe, for example, a vibrating string undergoing small transverse displacements u(x, t), where the wave velocity is . For definiteness, we will assume the string is clamped at x = 0 and x = L, so that the boundary conditions include the constraints

at all times t.

To solve (16.3b), we assume a separable form

which when substituted into (16.3b) gives,

Here the primes indicate differentiation with respect to the single variable appropriate to the symbol for the function, that is, and . If we now divide (16.11) by u = XT, we have

(16.12)numbered Display Equation

This equation is of a very special form, where each term that appears is a function of only one variable; is a function of x only and is a function of t only. It can therefore only be satisfied if each term is equal to the same constant, called the separation constant, which will be denoted by − k2 for later convenience. Thus we can write

and

where .

The first of these equations, together with the boundary conditions X(0) = X(L) = 0, which follow from (16.9), is just the eigenvalue problem that was solved earlier in Section 15.2. There we found that solutions only exist for

and are

(16.15)numbered Display Equation

The corresponding values of ω are then , and the general solution to (16.13b) is therefore

(16.16)numbered Display Equation

where An and Bn are arbitrary constants, so that we arrive at a set of solutions

where n = 1, 2, … is any positive integer. Each of these solutions is a ‘normal mode’ in which u(x, t) oscillates with a single angular frequency for all x-values1. At this point, we recall that the wave equation is linear and homogeneous, so that any linear combinations of solutions of the form (16.17) is also a solution of the original equation. The general solution is a linear combination of the normal modes, that is,

Finally, to determine the arbitrary constants An and Bn and obtain a unique solution requires imposing further boundary conditions that have to be specified according to the problem. For the case of a vibrating string, these might be that the string is released from rest at t = 0 from its initial state u(x, 0). The first of these conditions, that is,

numbered Display Equation

leads to the result An = 0 (all n) on substituting into (16.18), so that the solution becomes

In addition, the coefficients Bn can be determined, given the initial displacement u(x, 0), since from (16.19)

numbered Display Equation

which is just the Fourier expansion of the initial configuration of u(x, 0) extended, as an odd function, to the range − L < x < L (cf. Section 13.1.4). To invert it, we multiply both sides by and integrate from 0 to L using the orthonormality relation

This gives

and the solution (16.19) is completely determined.

images

Figure 16.1

16.2.2 The wave equation in three spatial dimensions

The method of Section 16.2.1 is easily extended to more than one spatial dimension. To illustrate this, we shall consider the wave equation in three dimensions with 0 < xyz < L, that is, for waves confined to a cubic box. We then find the form of the waves in this box by solving the equation

assuming that u vanishes at the walls of the box, that is,

In this case, substituting (16.8) into (16.22) and dividing by a = XYZT, gives

numbered Display Equation

As before, since the left-hand side is independent of t and the right-hand side depends solely on t, both sides must equal a constant, which we will denote by − k2. We thus obtain

and

numbered Display Equation

where we have transferred the term in Z to the right-hand side. Since the left-hand side of this equation is independent of Z and the right-hand side depends solely on Z, both sides again must be a constant, which we write as − k2 + k2z. We then obtain

and

numbered Display Equation

Taking to the right-hand side and repeating the argument then gives

(16.24c)numbered Display Equation

and

(16.24d)numbered Display Equation

where

(16.25)numbered Display Equation

We have now converted the wave equation (16.22) into four ODEs (16.24a)–(16.24d). Furthermore, (16.24b)–(16.24d) have the same form as (16.13a) with the same boundary conditions, and (16.24a) is identical to (16.13a). Solving in the same way, one finds that solutions satisfying the boundary conditions (16.23) only exist for

in analogy to (16.14), where nxnynz are positive integers; and the form of the solution is

in analogy to (16.17), where

(16.28)numbered Display Equation

and the arbitrary constants A and B can also depend on nxnynz.

The solutions (16.27) are the normal modes for waves confined to a cubic box and the general solution is a linear combination

(16.29)numbered Display Equation

These results apply to any type of wave confined to a cubic box, for example, sound waves or electromagnetic waves, and are easily generalised to any rectangular box with sides LxLyLz.

images

Figure 16.2 Diagrams used in the derivation of the density-of-states formula (16.30), where only the (kx, ky) plane is shown. The diagrams are not to the same scale, and in practical application the spacing between the points in the left-hand diagram is normally extremely small compared to typical k-values.

16.2.3 The diffusion equation in one spatial dimension

We next consider the diffusion equation (16.4), which is again second order in spatial derivatives but, in contrast to the wave equation, only first order in the time derivative. To focus on this difference, we shall just consider it in one spatial dimension. The generalisation to three spatial dimensions is very similar to that discussed in detail for the wave equation and is left as an exercise for the reader.

In one spatial dimension, the diffusion equation becomes

and, as an example, we will find solutions subject to the boundary conditions

(16.32)numbered Display Equation

where L is a constant and f(x) is a given function that must vanish at the end points 0 and L for the boundary conditions to be consistent. We start by again assuming a solution of the separable form (16.10), and after substituting into (16.31) and dividing by u we have

numbered Display Equation

Each side now has to be a constant, which we will denote by α. For α ≥ 0, it may easily be shown that the only solution consistent with the first boundary condition is u(x, t) = 0. However, for α ≡ −λ2 < 0, the solutions for X(x) and T(t) are

numbered Display Equation

where A, B and C are constants. The solution is therefore

where the constant C has been absorbed into A and B.

Using the first boundary condition with x = 0 gives

numbered Display Equation

for all t, and hence A = 0. Similarly putting x = L gives

numbered Display Equation

and hence only gives non-trivial solutions provided

numbered Display Equation

Thus the general solution is the superposition

As in the corresponding result (16.19) for the wave equation, at any fixed time t the solution takes the form of a Fourier sine series, but now each term is associated with an exponential rather than oscillating time dependence.

Finally, imposing the boundary condition at t = 0 gives

numbered Display Equation

Multiplying by and integrating from 0 to L using the orthonormality relation (16.20) then gives

(16.35)numbered Display Equation

which can be evaluated for any initial distribution u(x, 0) = f(x) to give the coefficients in (16.34).

16.3 Separation of variables: polar co-ordinates

In this section, we shall explore separation of variables using polar co-ordinates, rather than Cartesian co-ordinates. This is particularly useful for systems with circular, cylindrical or spherical symmetry. Throughout the discussion, which will again proceed via examples, we shall have in mind that the dependent variable is a physical quantity, and therefore will emphasise solutions that are finite and single-valued.

16.3.1 Plane-polar co-ordinates

Plane-polar co-ordinates were introduced in Section 2.2.1, where from Figure 2.4 we see that the co-ordinates (r, θ) and (r, θ + 2π) represent the same point in the plane. Hence if u(r, θ) represents a measurable quantity, it must not only be finite and continuous, but must also satisfy the boundary condition

which for separable solutions

implies

To illustrate the consequences of this, we will consider solutions of the Laplace and wave equations in two spatial dimensions.

In plane–polar co-ordinates, Laplace's equation is given by [cf. (7.29)]

(16.38a)numbered Display Equation

or equivalently,

Substituting (16.37) into (16.38b) and multiplying by r2 gives

(16.39)numbered Display Equation

where the left-hand side is independent of θ and the right-hand side is independent of r, so that both must be constant. Denoting the constant by m2, we then have

and

In solving these equations, it will be convenient to treat the cases m = 0 and m ≠ 0 separately.

  1. m = 0 In this case (16.40a) has the trivial solution Θ = A + Bθ, where A, B are arbitrary constants. Similarly, (16.40b) reduces to

    numbered Display Equation

    with the general solution

    numbered Display Equation

    Hence the general solution is

    numbered Display Equation

    which reduces to

    numbered Display Equation

    on imposing the boundary condition (16.36a) and absorbing the constant A into the arbitrary constants C and D.

  2. m ≠ 0 In this case (16.40a) has the general solution

    numbered Display Equation

    while the corresponding general solution of (16.40b) is

    numbered Display Equation

    as is easily verified by substitution. In addition, we note that m must be an integer if the boundary condition (16.36b) is to be satisfied, so that for m ≠ 0, the separable solutions of Laplace's equation are:

    (16.41b)numbered Display Equation

    where AmBmCmDm are arbitrary constants. Finally, the general solution is given by

    (16.42)numbered Display Equation

    where the various constants must be determined from boundary conditions.

The angular dependence of the separable solutions (16.41) is characteristic of the separable solutions of many other equations with circular symmetry, including the wave equation (16.3); the diffusion equation (16.4a); and the Schrödinger equation (16.7), provided the potential V(r) = V(r) is independent of angle. To illustrate this, we consider the wave equation in two spatial dimensions, which is given by

in planar co-ordinates. Assuming a separable solution

(16.44)numbered Display Equation

gives

numbered Display Equation

after dividing by u, and hence

(16.45a)numbered Display Equation

and

on separating off the right-hand side and denoting the separation constant by − k2. After multiplying by r2, (16.46) can also be separated in a similar way to Laplace's equation to give

and

Here, equation (16.45b) is identical to (16.40a) and has the same solution, while on expanding the differential in (16.45c) and substituting z = kr, we obtain

numbered Display Equation

which is Bessel's equation (15.53a) and has the general solution

where Cm and Dm are arbitrary constants and Jm and Nm are the Bessel functions discussed in Section 15.4.1. In particular, Nm(kr) is singular at r = 0, so if we require u to be finite at r = 0, (16.47a) reduces to

Finally, the general solution of (16.45) is

(16.48)numbered Display Equation

where the angular frequency . Hence the separable solutions that are finite at r = 0 are

(16.49a)numbered Display Equation

and

(16.49b)numbered Display Equation

where m = 1, 2, … must be an integer if the boundary condition (16.36b) is to be satisfied, and where Cm in (16.47b) has been absorbed into the other constants. In particular, the angular dependence of the separable solutions is the same as for Laplace's equation and is characteristic of the separable solutions of many equations with circular symmetry.

16.3.2 Spherical polar co-ordinates

The above discussion of plane polar co-ordinates is easily extended to the spherical polar co-ordinates (r, θ, φ) discussed in Section 11.3.1. However in this case, (r, θ, φ) and (r, θ, φ + 2π) correspond to the same point in space, so that if u is a physical quantity, we must impose the boundary condition

in addition to requiring that u is finite and continuous.

Again, we shall illustrate this by considering Laplace's equation (16.2). We start by using the result given in Table 12.1 for in spherical polar co-ordinates to express (16.2) as

and then substitute the decomposition

(16.52)numbered Display Equation

into (16.51). After dividing through by u and multiplying by r2, we obtain the equation

(16.53)numbered Display Equation

The first two terms are functions of both r and θ, whereas the third term is a function of φ alone. Since the three variables are independent, this means that the third term must be a constant, and so we can set

The rest of the equation can then be manipulated into the form

numbered Display Equation

Again we use the fact that each side is a function of a single independent variable and hence must be a constant, which we will denote by l(l + 1) for later convenience. We thus have the two equations

and

The radial equation (16.55) may be written as

numbered Display Equation

and has the general solution

where Al and Bl are arbitrary constants, and the possible values of l are restricted by consideration of the angular dependence, as we shall soon see.

We start with the φ-dependence. The only solutions of (16.54) compatible with the boundary conditions (16.50) are

(16.58)numbered Display Equation

Alternatively, and equivalently, we can choose a set of solutions

where the coefficients Am are arbitrary constants.

We next turn to the θ dependence. On setting μ = cos θ and expanding the first term in (16.56), it becomes

This is the associated Legendre equation, which has solutions2 that are finite for μ = cos θ = −1 (i.e. on the negative z-axis) only if l = 0, 1, 2, … and the m values are restricted to − lml. They are called associated Legendre polynomials and denoted Pml(cos θ), so that we write

and from (16.57), (16.59) and (16.61), we find that the finite separable solutions which satisfy the boundary conditions (16.50) are

where l = 0, 1, 2, …, − lml and AlmBlm are arbitrary constants. The explicit forms of the first few polynomials are given in Example 15.10. For m = 0, when (16.60) reduces to Legendre's equation (15.23), they reduce to the Legendre polynomials Pl(cos θ) of order l discussed in detail in Section 15.3.1. Hence for m = 0 the solutions (16.62) reduce to

which are independent of the azimuthal angle φ and unchanged by rotations about the z-axis. For l = m = 0, since P0(cos θ) = 1, (16.63) reduces to

(16.64)numbered Display Equation

which is the most general spherically symmetric solution to Laplace's equation. The most general solution, without symmetry constraints, is obtained by taking linear combinations of the separable solutions (16.63) and is

Finally, we note that the angular dependence of the separable solutions (16.62) is not specific to Laplace's equation, but is shared by the separable solutions of other important equations that have spherical symmetry. For example, the separable solutions of the wave equation and the diffusion equation also take the form

for appropriate choices of R(r) and T(t), as the reader may verify.

In practice, it is common to rewrite equations like (16.62) and (16.65) in terms of the so-called spherical harmonics defined by

(16.67a)numbered Display Equation

where the constant is chosen so that the normalisation condition

(16.67b)numbered Display Equation

is satisfied. With this convention, the first few spherical harmonics are given by

numbered Display Equation

16.3.3 Cylindrical polar co-ordinates

We conclude the discussion of polar co-ordinates by considering cylindrical polar co-ordinates defined in Section 11.3.1. In this case, the co-ordinates (ρ, φ, z) and (ρ, φ + 2π, z) represent the same point, so the boundary condition corresponding to (16.50) is

We shall again take as our example the Laplace equation, which in cylindrical polar co-ordinates takes the form

where we have used the expression for given in Table 12.1. Assuming a separable solution

then gives

(16.72)numbered Display Equation

on substituting into (16.70) and dividing by u. In this equation, the first two terms are independent of z, while the third term depends only on z, and so must be a constant. Hence, separating the third term, we obtain

and

numbered Display Equation

where we have taken the separation constant to be k2. The second of these equations is not separable as it stands, but multiplying through by ρ2 gives

numbered Display Equation

which is separable. Separating the third term and denoting the separation constant by m2 then gives

and

where we have expanded the ρ-derivative.

It remains to solve the three ODEs (16.73a)–(16.73c). For any k, the general solution of the first of these is

(16.74)numbered Display Equation

where Ak and Bk are arbitrary constants. If we impose the boundary conditions (16.69), the general solutions of (16.73b) are

(16.75)numbered Display Equation

where m ≥ 0 is an integer, and CmDm are arbitrary constants. Finally, on setting η = kρ in (16.73c) we obtain

numbered Display Equation

This is Bessel's equation (15.53), with general solutions of the form (15.70), i.e.

(16.76)numbered Display Equation

where Em and Fm are arbitrary constants and Jm(kρ) and Nm(kρ) are Bessel functions of the first and second kind respectively. These Bessel functions were discussed in Section 15.4.1, where we saw that Nm(kρ) was singular at kρ = 0. If we require solutions that are finite at ρ = 0, we must therefore set Fm = 0 in (16.71). Hence those separable solutions (16.66) that are both finite and single-valued are

and

where we have absorbed C0 and Em into the other constants. Since Laplace's equation is homogeneous, more general solutions may then be formed by linear superposition of the solutions (16.77a) and (16.77b), where in applications the possible values of k and the various constants must be determined by boundary conditions, as we shall illustrate by an example.

*16.4 The wave equation: d'Alembert's solution

In the next three sections, we shall consider other methods of solution of PDEs, mainly applied to functions of two variables. We start with the wave equation (16.3b), which we will solve by introducing the new variables

(16.78)numbered Display Equation

On changing variables using (7.24), we obtain

numbered Display Equation

and

numbered Display Equation

so that the wave equation becomes

with the general solution [cf. (16.2a) and (16.2b)]

where f and g are arbitrary differentiable functions.

Equation (16.80) is the general solution of the wave equation and each of the two terms has a simple interpretation. Let us suppose

Then we have

numbered Display Equation

for all t, so that the solution moves as illustrated in Figure 16.3 for a simple choice of u(x, 0). It represents a travelling wave moving in the positive x-direction with speed . Thus, for example, a simple harmonic wave with wavelength λ, wave number and angular frequency , can be written in the form

numbered Display Equation

where A and α are arbitrary constants. Similarly, is a wave travelling in the minus x-direction and (16.80) shows that any other non-trivial solution of the wave equation may be written as a sum of a wave travelling to the right and one travelling to the left.

images

Figure 16.3 A travelling wave corresponding to the solution (16.81) shown at two arbitrary times t1 and t2 > t1 for an arbitrary choice of u(x, 0), where .

At this point, we digress briefly to indicate how this description can be extended to three dimensions. Denoting a point in space by its position vector r, one easily shows that the functions

(16.82)numbered Display Equation

are solutions of the three-dimensional wave equation, where the unit vector indicates a chosen direction in space. Since the equation of a plane perpendicular to the unit vector is , where c is a constant, then at any fixed time t, the functions f and g are constant over the whole plane. They are therefore plane waves travelling in the positive and negative directions, respectively. This may be more familiar if we note that a simple harmonic wave in the direction analogous to (16.80) is given by

numbered Display Equation

where A and α are again constants and the wave vector .

We now return to the one-dimensional case and consider its solution subject to the initial conditions

(16.83a)numbered Display Equation

and

(16.83b)numbered Display Equation

where α(x) and β(x) are given functions. Substituting the general solution (16.80) into these boundary conditions then gives

and

where

numbered Display Equation

and similarly for g′(x). We now integrate (16.84b) to get

numbered Display Equation

where the integration constant c will depend on the arbitrarily chosen lower limit a of the integration. From this equation, together with (16.84a), we obtain

numbered Display Equation

and

numbered Display Equation

and hence

numbered Display Equation

and

numbered Display Equation

Finally, adding we find

(16.85)numbered Display Equation

This is d'Alembert's solution to the wave equation (16.3b) subject to the boundary conditions (16.83). It is unique and independent of the intermediate constants c and a introduced in its derivation. Furthermore, at any point x = x0, u(x0, t) is dependent only on the initial values u(x, 0) and in the range , and is independent of u and outside this range. This embodies the idea of ‘causality’ for the wave equation, since it is just the range from within which a signal emitted at t = 0 and travelling with speed can reach x0 in a time less than or equal to t.

*16.5 Euler Equations

The method used in the previous section to obtain (16.80) as the general solution of the wave equation can be extended to solve any equation of the form

where A, B, C, are given constants and x, y are any variables, not necessarily Cartesian co-ordinates. Such equations are often called Euler's equations.4 To solve them we introduce the new variables

where λ1 and λ2 are constants. We then try to find values of λ1 and λ2 such that (16.86) reduces to an equation of the form (16.79), with a general solution

analogous to (16.80) for the wave equation.

To see whether this is possible, we change variables using (7.24) to obtain

numbered Display Equation

Substituting these expressions into (16.86) gives

If we now choose λi (i = 1, 2) to be the roots of

then (16.89) reduces to

So, provided the term in square brackets does not vanish,

numbered Display Equation

and the solution (16.88) follows by successive integrations.

The condition that the square bracket in (16.91) does not vanish is easily found by noting from Equation (2.7) that

numbered Display Equation

so that

numbered Display Equation

that is, the equation must be such that ACB2.

If AC = B2, the square bracket in (16.91) does vanish and (16.90) has only one solution, which is a repeated root given by . In this case, we choose, , λ2 = 0, and substituting these in (16.87) we find that the first and third terms vanish and the equation reduces to

numbered Display Equation

Direct integration then gives

(16.92)numbered Display Equation

where f and g are again arbitrary functions of ξ. The solution of the PDE when AC = B2 is therefore

where f and g are arbitrary functions of ξ.

To summarise, we have to distinguish between the cases AC = B2, when the general solution is given by (16.93); and ACB2, when the general solution is given by (16.88), where λ1, λ2 are the roots of (16.90).

*16.6 Boundary conditions and uniqueness

So far, we have focussed on problems that can be solved exactly. However, it is often not possible to do this in practice, and then one must resort to numerical methods to find approximate solutions of a given PDE that satisfy specific boundary conditions. In these cases, especially, it is very useful to know in advance what boundary conditions result in a unique, stable solution of the PDE, where by stable we mean that very small changes in the boundary conditions do not lead to very large changes in the solution. Here we will simply state the main results without proof, since the derivations are often difficult.5

The boundary conditions take the form of information about the dependent variable u specified on a continuous boundary, which may be open, like a plane in three-dimensional space, or closed, like the surface of a sphere. The main types of boundary conditions are classified as follows:

  • Dirichlet.

    The value of u is specified at each point of the boundary.

  • Neumann

    The value of the normal derivative , where is the unit normal to the boundary, is specified at each point of the boundary.

  • Cauchy

    The values of both u and are specified at each point of the boundary.

     The next step is to classify PDEs into three types, called elliptic, hyperbolic and parabolic. In doing so, we shall focus on second-order linear equations that contain only the derivatives

    with constant coefficients, where is replaced by ∂2u/∂x2 if there is only a single spatial variable. Thus, we consider PDEs of the form

    (16.94b)numbered Display Equation

    where A, B, C and D are constants and ρ is a given function. This form includes many equations of physical interest, including all those listed in Section 6.1. They are then classified according to which of the partial derivatives (16.94a) occur, and by the relative sign of their coefficients. We consider each in turn.

  • Elliptic equations

    These are defined as those containing and ∂2u/∂t2 with coefficients of the same sign, that is, AB > 0; or just with no time derivatives. The latter are of most interest and include Laplace's equation, Poisson's equation and the Helmholtz equation

    (16.95)numbered Display Equation

    both with a source (ρ ≠ 0) and without (ρ = 0), where k2 is a positive real constant.

     Elliptic equations have the property that if either Dirichlet or Neumann boundary conditions are applied on a closed boundary, then the equation has a unique and stable solution within the boundary. Consequently, if Cauchy boundary conditions are applied, the equation in general has no solutions, and the equation is said to be over-constrained. The closed boundary may be finite, as illustrated for the Laplace equation in Examples 16.7 and 16.8, which used Dirichlet conditions on a finite closed surface; or the surface may be at infinity, as illustrated for Poisson's equation in Example 16.6. Alternatively, if one wishes to determine the function outside a finite closed boundary, then either Dirichlet or Neumann conditions must be applied both on the finite boundary and at infinity.

  • Hyperbolic equations

    These are defined as those containing and ∂2u/∂t2 with coefficients of the opposite sign, that is, AB < 0. They may in principle also contain terms in ∂u/∂t, but in practice such terms are usually absent in physical applications. Examples of hyperbolic equations are the wave equations (16.3a) and (16.3b) and the Klein-Gordon equation.

    numbered Display Equation

    which plays an important role in relativistic quantum mechanics, where c is the speed of light and m is a particle mass.

     Hyperbolic equations have unique and stable solutions if Cauchy boundary conditions are applied on an open boundary. In physical applications, this is usually taken to correspond to a constant time, which can always be chosen to be t = 0. One thus has to specify u(rt = 0) and the time derivative of u(r, t) at t = 0 to obtain a unique and stable solution, as illustrated in the wave equation in Section 16.2.1 [cf. (16.19)] and Example 16.2 for standing waves constrained to vanish at x = 0, L; and for travelling waves in one dimension in Section 16.4 [cf. (16.83, 85)].

  • Parabolic equations

    These are defined as those containing terms in and ∂u/∂t, but not terms in ∂2u/∂t2, that is, AB = 0. Examples are the diffusion equation (16.4) and the Schrödinger equation (16.7). In this case, unique and stable solutions are obtained if Dirichlet or Neumann conditions are imposed on an open boundary. This is almost always chosen to be constant time t = t0, and unique and stable solutions for t > t0 are obtained given either u(rt0) or . This is illustrated for the case of Dirichlet boundary conditions for the examples given in Section 16.2.3 and Section 16.6.1 below.

     Finally, we stress again that we have only considered equations containing the partial derivatives (16.89), since this covers many of the most important PDEs in physical applications. The discussion can, however, be extended to all linear second-order PDEs with constant coefficients.6

*16.6.1 Laplace transforms

In Section 14.2.4, we introduced Laplace transforms (14.44) and showed how they could be used to obtain solutions of ODEs that automatically incorporated given boundary conditions. This method can be extended in principle to PDEs in which the boundary conditions are given at an initial time t = 0, although, as in ODEs, it may be difficult to perform the inverse Laplace transform required to obtain the final solution.

To illustrate this, we shall consider the bounded solution of the diffusion equation in one spatial dimension (16.31) in the range 0 < x < ∞ for times t > 0, with boundary conditions

This could, for example, describe the temperature distribution due to heat flow along a long rod with one end at x = 0, if the rod is initially at zero temperature u = 0, but is in contact at t > 0 with a heat bath of constant temperature u0 at the end x = 0, and the sides of the bar are perfectly lagged so that heat flow from the sides can be neglected.

To solve this problem, we take the Laplace transform of both sides of (16.31) with respect to time t from (14.44) and (14.45a). We then have, with an appropriate change of notation:

and

where we have used (16.96) to set u(x, 0) = 0 in (16.97b). Hence (16.31) becomes

(16.98)numbered Display Equation

with the general solution

numbered Display Equation

where α = (p/κ)1/2 > 0 and A and B are arbitrary constants. If u(x, t) is bounded as x → ∞, which is an obvious requirement if it represents a temperature distribution, then it follows that F(x, p) must also be bounded as x → ∞, so that A = 0. The value of B is then found by imposing the boundary condition u(0, t) = u0, which from (16.97a) gives

numbered Display Equation

Hence,

and the final solution is given by the inverse transform

At this point, we remind the reader that, as discussed in Section (14.2.4), finding inverse Laplace transforms is difficult and often impossible to do in closed form. One frequently has to resort to tables of such transforms like that of Table 14.1, or the more extensive tables available in the literature.7 In the case above, the required inverse transform can be expressed in terms of the error function

and the associated complementary error function

The error function is normalised so that it tends to unity as t → ∞, since (cf. Example 11.11)

numbered Display Equation

The behaviour of both (16.100a) and (16.100b) is shown in Figure 16.4. The relevance of this becomes clear on taking the Laplace transform of , which can be shown to be [cf. Example 16.11]

images

Figure 16.4 The error function (16.100a) and the complementary error function (16.100b).

This obviously implies

(16.101b)numbered Display Equation

which, together with (16.99a) and (16.99b), gives

(16.102)numbered Display Equation

as the final solution. Hence u(x, t) → u0 as t → ∞, but more slowly as x increases, which is what one intuitively expects if u represents the temperature of a long bar, as in this example.

Problems 16

  1.   16.1 A function u(x, y, z) satisfies the Helmholtz equation

    numbered Display Equation

    in the range 0 ≤ xL, where k is a constant. Find the values of k2 such that u satisfies the boundary conditions

    numbered Display Equation

    and give the corresponding solutions.

  2.   16.2 In quantum mechanics, the wave function u(x, t) of a particle of mass m moving freely in one dimension is described by the Schrödinger equation

    numbered Display Equation

    where and h is Planck's constant. Show that separable solutions of the form

    numbered Display Equation

    exist, where E is an arbitrary real constant. What are the possible values of E if u satisfies the periodic boundary conditions

    numbered Display Equation

    for any x?

  3.   16.3 A rectangular plate with sides of length a and b is oriented so that 0 ≤ xa, 0 ≤ yb. The edges corresponding to x = 0, x = ay = 0 are each kept at temperature zero, and the other edge has a temperature distribution along its length given by , where u0 is a constant. Find an expression for the temperature at an arbitrary point on the plate. Note the integral

    numbered Display Equation
  4.   16.4 A thin rectangular plate, defined by 0 ≤ xa, 0 ≤ yb, is clamped along its perimeter. By solving the two-dimensional wave equation with velocity , show that its vibrational modes are given by

    numbered Display Equation

    where

    numbered Display Equation

    with

    numbered Display Equation

    and anm and bnm are arbitrary constants. Hence show that if the plate is released from rest with an initial profile

    numbered Display Equation

    its subsequent motion is described by

    numbered Display Equation
  5.   16.5 A thin insulated rod of length L, with ends at x = 0 and x = L, has an initial temperature along its length given by

    numbered Display Equation

    where u0 is a constant. If the ends of the rod are kept at temperature zero, find an expression for u(x, t) for t > 0.

  6.   16.6 Find the function u(x, y) that describes the steady-state distribution of temperature through a two-dimensional semi-infinite slab, where 0 ≤ xd and 0 ≤ y < ∞, if the long edges of the slab are kept at zero temperature. Assume that for 0 < x < d and y = 0, u(x, y) = f(x), where f(x) is a given function (the form of which would have to be such as to satisfy the boundary condition u(0, y) = u(d, y) = 0) and u(x, y) → 0 as y → ∞ for 0 ≤ xd.

  7.   16.7 Find the single-valued solution u(r, θ) of the two-dimensional Laplace equation within a circle of radius R, subject to the boundary condition u(R, θ) = f(θ), where f(θ) is an arbitrary positive function of θ.

  8.   16.8 Show that a spherically symmetric potential u that obeys the Laplace equation and vanishes at infinity may be written , where a is a constant.

  9.   16.9 A neutral conducting sphere of radius a is centred at the origin and is exposed to a uniform electric field E in the z-direction. Find the electrostatic potential u satisfying Laplace's equation outside the sphere if the potential on the sphere is set, by convention, to zero.

  10.  16.10 Show that the differential equation

    numbered Display Equation

    has separable solutions of the form u(r, θ, φ) = R(r)Θ(θ)Φ(φ), where r, θ, φ are spherical polar co-ordinates, and f, g and h are arbitrary functions.

  11.  16.11 On substituting into the Schrödinger equation (16.7), one obtains the so-called time independent Schrödinger equation

    Show that for spherical potentials V(r) = V(r), this equation has separable solutions of the form

    where Ylm are spherical harmonics; and find the ODE satisfied by the radial function R(r).

  12.  16.12 Verify that

    numbered Display Equation

    where κ is the diffusivity, is a solution of the diffusion equation in spherical co-ordinates.

  13.  16.13 A sphere of radius R has the surface of its upper hemisphere held at a constant temperature TU, and the surface of its lower hemisphere held at a constant temperature TL. Assuming it is in thermal equilibrium, find an expansion for the temperature u(r, θ) within the sphere, accurate to terms of order .

  14.  16.14 Find separable solutions of the Helmholtz equation

    numbered Display Equation

    in cylindrical polar co-ordinates, when k2 > 0, if u is single-valued, finite and tends exponentially to zero as z → ∞.

  15.  16.15 A solid semi-infinite cylinder of unit radius is in thermal equilibrium. Show that the temperature distribution u(ρ, φ, z) in the cylinder, subject to the boundary conditions (1) u = ρsin φ on the base z = 0, and (2) u = 0 on the curved surface, is

    numbered Display Equation

    where Jν is a Bessel function of the first kind of order ν, and kn are the zeros of J1(k).

  16.  16.16 If in question 16.15 the base is kept at a constant temperature u0, then the resulting temperature distribution is

    numbered Display Equation

    Use this expansion to calculate the value of the temperature at and z = 1 to three decimal places if u0 = 50. [Note: the positions of the zeros of the Bessel functions are given in Table 15.1 and values of the Bessel function Jn(x) may be found from a number of widely available sources, for example, the function BESSEL(x, n) in a Microsoft Excel spreadsheet, or the website www.wolframalpha.com.]

  17. *16.17 Find the solution to the equation

    numbered Display Equation

    subject to the initial conditions

    numbered Display Equation
  18. *16.18 Find the general form of the solution to the following equations:

    numbered Display Equation
  19. *16.19 Solve the equation

    numbered Display Equation

    subject to the boundary conditions u(0, y) = y2, u(x, 0) = sin x.

  20. *16.20 A thin insulated metal rod lies horizontally in the semi-infinite region x ≥ 0 and is initially at zero temperature. At time t > 0, the end at x = 0 is placed in contact with a heat bath with fixed temperature u0. If F(x, p) ≡ L[u(x, t)] is the Laplace transform of u(x, t) show that the distribution of temperature along the rod at time t may be written as

    numbered Display Equation

    where L− 1 denotes an inverse Laplace transform and κ is the thermal diffusivity.

  21. *16.21 Show that the solution of the equation

    numbered Display Equation

    for t > 0 and 0 < x < a, subject to the boundary conditions

    and

    may be written

    numbered Display Equation

    where L− 1 is the inverse Laplace transform.

  22. *16.22 Show that the solution of the wave equation with unit velocity for t > 0 and 0 < x < a, where a is a constant, subject to the boundary conditions

    and

    where f is a constant, may be written

    numbered Display Equation

    where L− 1 is the inverse Laplace transform.

Notes

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset