12.5 Sturm–Liouville Problem

INTRODUCTION

For convenience we present here a brief review of some of the ordinary differential equations that will be of importance in the sections and chapters that follow.

Linear equations General solutions
Cauchy–Euler equation General solutions, x > 0
Parametric Bessel equation (ν = 0) General solution, x > 0
Legendre’s equation (n = 0, 1, 2, …) Particular solutions are polynomials


Regarding the two forms of the general solution of y″α2y = 0, we will, in the future, employ the following informal rule:

This rule will be useful in Chapters 13 and 14.

Use the exponential form y = c1eαx + c2eαx when the domain of x is an infinite or semi-infinite interval; use the hyperbolic form y = c1 cosh αx + c2 sinh αx when the domain of x is a finite interval.

Eigenvalues and Eigenfunctions

Orthogonal functions arise in the solution of differential equations. More to the point, an orthogonal set of functions can be generated by solving a two-point boundary-value problem involving a linear second-order differential equation containing a parameter λ. In Example 2 of Section 3.9 we saw that the boundary-value problem

y″ + λy = 0, y(0) = 0, y(L) = 0, (1)

possessed nontrivial solutions only when the parameter λ took on the values λn = n2π2/L2, n = 1, 2, 3, … called eigenvalues. The corresponding nontrivial solutions y = c2 sin(nπx/L) or simply y = sin(nπx/L) are called the eigenfunctions of the problem. For example, for (1) we have

For our purposes in this chapter it is important to recognize the set of functions generated by this BVP; that is, {sin(nπx/L)}, n = 1, 2, 3, …, is the orthogonal set of functions on the interval [0, L] used as the basis for the Fourier sine series.

EXAMPLE 1 Eigenvalues and Eigenfunctions

It is left as an exercise to show, by considering the three possible cases for the parameter λ (zero, negative, or positive; that is, λ = 0, λ = –α2 < 0, α > 0, and λ = α2 > 0, α > 0), that the eigenvalues and eigenfunctions for the boundary-value problem

y″ + λy = 0, y′(0) = 0, y′(L) = 0(2)

are, respectively, λn = = n2π2/L2, n = 0, 1, 2, …, and y = c1 cos (nπx/L), c1 ≠ 0. In contrast to (1), λ0 = 0 is an eigenvalue for this BVP and y = 1 is the corresponding eigenfunction. The latter comes from solving y″ = 0 subject to the same boundary conditions y′(0) = 0, y′(L) = 0. Note also that y = 1 can be incorporated into the family y = cos (nπx/L) by permitting n = 0. The set {cos (nπx/L)}, n = 0, 1, 2, 3, …, is orthogonal on the interval [0, L]. See Problem 3 in Exercises 12.5.

Regular Sturm–Liouville Problem

The problems in (1) and (2) are special cases of an important general two-point boundary-value problem. Let p, q, r, and r′ be real-valued functions continuous on an interval [a, b], and let r(x) > 0 and p(x) > 0 for every x in the interval. Then

Solve: [r(x)y′] + (q(x) + λp(x))y = 0(3)

Subject to: A1y(a) + B1y′(a) = 0(4)

A2 y(b) + B2 y′(b) = 0(5)

is said to be a regular Sturm–Liouville problem. Jacques Charles François Sturm (1803–1855) was born in Switzerland but spent his adult life as a professor of mathematics and mechanics in Paris, France. Sturm’s name is one of 72 names of scientists, engineers, and mathematicians engraved on the Eiffel Tower. The French mathematician Joseph Liouville (1809–1882) was also a professor in Paris. In addition to his work with Sturm on boundary-value problems, an important theorem in complex analysis bears his name (Liouville’s theorem). The small impact crater Liouville on the moon is named after him.

The coefficients in the boundary-conditions (4) and (5) are assumed to be real and independent of λ. In addition, A1 and B1 are not both zero, and A2 and B2 are not both zero. The boundary-value problems in (1) and (2) are regular Sturm–Liouville problems. From (1) we can identify r(x) = 1, q(x) = 0, and p(x) = 1 in the differential equation (3); in boundary condition (4) we identify a = 0, A1 = 1, B1 = 0, and in (5), b = L, A2 = 1, B2 = 0. From (2) the identifications would be a = 0, A1 = 0, B1 = 1 in (4), and b = L, A2 = 0, B2 = 1 in (5).

The differential equation (3) is linear and homogeneous. The boundary conditions in (4) and (5), both a linear combination of y and y′ equal to zero at a point, are also called homogeneous. A boundary condition such as A2 y(b) + B2 y′(b) = C2, where C2 is a nonzero constant, is nonhomogeneous. Naturally, a boundary-value problem that consists of a homogeneous linear differential equation and homogeneous boundary conditions is said to be homogeneous; otherwise it is nonhomogeneous. The boundary conditions (4) and (5) are said to be separated because each condition involves only a single boundary point. Boundary conditions are referred to as mixed if each condition involves both boundary points x = a and x = b. For example, the periodic boundary conditions y(a) = y(b), y′(a) = y′(b) are mixed boundary conditions.

Because a regular Sturm–Liouville problem is a homogeneous BVP, it always possesses the trivial solution y = 0. However, this solution is of no interest to us. As in Example 1, in solving such a problem we seek numbers λ (eigenvalues) and nontrivial solutions y that depend on λ (eigenfunctions).

Properties

Theorem 12.5.1 is a list of some of the more important of the many properties of the regular Sturm–Liouville problem. We shall prove only the last property.

THEOREM 12.5.1 Properties of the Regular Sturm–Liouville Problem

  1. There exist an infinite number of real eigenvalues that can be arranged in increasing order λ1 < λ2 < λ3 < < λn < such that λn → ∞ as n → ∞.
  2. For each eigenvalue there is only one eigenfunction (except for nonzero constant multiples).
  3. Eigenfunctions corresponding to different eigenvalues are linearly independent.
  4. The set of eigenfunctions corresponding to the set of eigenvalues is orthogonal with respect to the weight function p(x) on the interval [a, b].
PROOF OF (d):

Let ym and yn be eigenfunctions corresponding to eigenvalues λm and λn, respectively. Then

(6)

(7)

Multiplying (6) by yn and (7) by ym and subtracting the two equations gives

Integrating this last result by parts from x = a to x = b then yields

(8)

Now the eigenfunctions ym and yn must both satisfy the boundary conditions (4) and (5). In particular, from (4) we have

A1ym(a) + B1(a) = 0

A1yn(a) + B1(a) = 0.

For this system to be satisfied by A1 and B1, not both zero, the determinant of the coefficients must be zero:

ym(a)(a) – yn(a)(a) = 0.

A similar argument applied to (5) also gives

ym(b)(b) – yn(b)(b) = 0.

Using these last two results in (8) shows that both members of the right-hand side are zero. Hence we have established the orthogonality relation

p(x)ym(x)yn(x) dx = 0, λmλn. (9)

It can also be proved that the orthogonal set of eigenfunctions {y1(x), y2(x), y3(x), …} of a regular Sturm–Liouville problem is complete on [a, b]. See page 684.

EXAMPLE 2 A Regular Sturm–Liouville Problem

Solve the boundary-value problem

y″ + λy = 0, y(0) = 0, y(1) + y′(1) = 0.(10)

SOLUTION

You should verify that for λ = 0 and for λ = –α2 < 0, where α > 0, the BVP in (10) possesses only the trivial solution y = 0. For λ = α2 > 0, α > 0, the general solution of the differential equation y″ + α2y = 0 is y = c1cos αx + c2sin αx. Now the condition y(0) = 0 implies c1 = 0 in this solution and so we are left with y = c2 sin αx. The second boundary condition y(1) + y′(1) = 0 is satisfied if

Choosing c2 ≠ 0, we see that the last equation is equivalent to

(11)

If we let x = α in (11), then FIGURE 12.5.1 shows the plausibility that there exists an infinite number of roots of the equation tan x = –x, namely, the x-coordinates of the points where the graph of y = –x intersects the branches of the graph of y = tan x. The eigenvalues of problem (10) are then , where αn, n = 1, 2, 3, …, are the consecutive positive roots α1, α2, α3, … of (11). With the aid of a CAS it is easily shown that, to four rounded decimal places, α1 = 2.0288, α2 = 4.9132, α3 = 7.9787, and α4 = 11.0855, and the corresponding solutions are y1 = sin 2.0288x, y2 = sin 4.9132x, y3 = sin 7.9787x, and y4 = sin 11.0855x. In general, the eigenfunctions of the problem are {sin αn x}, n = 1, 2, 3, ….

Two functions y equals tan x and y equals negative x are graphed on an x y plane. The function y equals tan x is graphed as curves in five parts. The first part starts from the third quadrant, goes up and to the right through the origin, and exits the top of the first quadrant. The second part starts from the fourth quadrant, goes up and to the right through the positive x axis, and exits the top of the first quadrant. The third part starts from the fourth quadrant, goes up and to the right through the positive x axis, and exits the top of the first quadrant. The fourth part starts from the bottom of the fourth quadrant, goes up and to the right through the positive x axis, and exits the top of the first quadrant. The fifth part starts from the bottom right of the fourth quadrant, goes up and to the right through the positive x axis, and exits the top of the first quadrant. The function y equals negative x is a straight line that starts from the left of the second quadrant, goes down and to the right through the origin, intersects all the curves, and exits the bottom right of the fourth quadrant. The points of intersection are marked with a black dot and these points correspond to the x value of 0, x subscript 1, x subscript 2, x subscript 3, x subscript 4, respectively.

FIGURE 12.5.1 Positive roots of tan x = –x in Example 2

With identifications r(x) = 1, q(x) = 0, p(x) = 1, A1 = 1, B1 = 0, A2 = 1, and B2 = 1 we see that (10) is a regular Sturm–Liouville problem. Thus {sin αn x}, n = 1, 2, 3, … is an orthogonal set with respect to the weight function p(x) = 1 on the interval [0, 1].

In some circumstances we can prove the orthogonality of the solutions of (3) without the necessity of specifying a boundary condition at x = a and at x = b.

Singular Sturm–Liouville Problem

There are several other important conditions under which we seek nontrivial solutions of the differential equation (3):

  • r(a) = 0 and a boundary condition of the type given in (5) is specified at x = b;(12)
  • r(b) = 0 and a boundary condition of the type given in (4) is specified at x = a;(13)
  • r(a) = r(b) = 0 and no boundary condition is specified at either x = a or at x = b;(14)
  • r(a) = r(b) and boundary conditions y(a) = y(b), y′(a) = y′(b).(15)

The differential equation (3) along with one of conditions (12) or (13) is said to be a singular boundary-value problem. Equation (3) with the conditions specified in (15) is said to be a periodic boundary-value problem because the boundary conditions are periodic. Observe that if, say, r(a) = 0, then x = a may be a singular point of the differential equation, and consequently a solution of (3) may become unbounded as xa. However, we see from (8) that if r(a) = 0, then no boundary condition is required at x = a to prove orthogonality of the eigenfunctions provided these solutions are bounded at that point. This latter requirement guarantees the existence of the integrals involved. By assuming the solutions of (3) are bounded on the closed interval [a, b] we can see from inspection of (8) that

  • If r(a) = 0, then the orthogonality relation (9) holds with no boundary condition at x = a;(16)
  • If r(b) = 0, then the orthogonality relation (9) holds with no boundary condition at x = b*;(17)
  • If r(a) = r(b) = 0, then the orthogonality relation (9) holds with no boundary conditions specified at either x = a or x = b;(18)
  • If r(a) = r(b), then the orthogonality relation (9) holds with the periodic boundary conditions y(a) = y(b), y′(a) = y′(b).(19)

Self-Adjoint Form

If we carry out the differentiation , the differential equation in (3) is the same as

(20)

For example, Legendre’s differential equation is exactly of the form given in (20) with and . In other words, another way of writing Legendre’s DE is

(21)

But if you compare other second-order DEs (say, Bessel’s equation, Cauchy–Euler equations, and DEs with constant coefficients) you might believe, given the coefficient of y′ is the derivative of the coefficient of y″, that few other second-order DEs have the form given in (3). On the contrary, if the coefficients are continuous and a(x) ≠ 0 for all x in some interval, then any second-order differential equation

a(x)y″ + b(x)y′ + (c(x) + λd(x))y = 0(22)

can be recast into the so-called self-adjoint form (3). To see this, we proceed as in Section 2.3 where we rewrote a linear first-order equation in the form by dividing the equation by a1(x) and then multiplying by the integrating factor where, assuming no common factors, P(x) = a0(x)/a1(x). So first, we divide (22) by a(x). The first two terms are then where, for emphasis, we have written Y = y′. Second, we multiply this equation by the integrating factor , where a(x) and b(x) are assumed to have no common factors

In summary, by dividing (22) by a(x) and then multiplying by we get

(23)

Equation (23) is the desired form given in (20) and is the same as (3):

For example, to express 3y″ + 6y′ + λy = 0 in self-adjoint form, we write y″ + 2y′ + λ y = 0 and then multiply by . The resulting equation is

Note.

It is certainly not necessary to put a second-order differential equation (22) into the self-adjoint form (3) in order to solve the DE. For our purposes we use the form given in (3) to determine the weight function p(x) needed in the orthogonality relation (9). The next two examples illustrate orthogonality relations for Bessel functions and for Legendre polynomials.

EXAMPLE 3 Parametric Bessel Equation

In Section 5.3 we saw that the general solution of the parametric Bessel differential equation x2y″ + xy′ + (α2x2n2)y = 0, n = 0, 1, 2, … is y = c1Jn(αx) + c2Yn(αx). After dividing the parametric Bessel equation by the lead coefficient x2 and multiplying the resulting equation by the integrating factor , we obtain the self-adjoint form

where we identify r(x) = x, q(x) = –n2/x, p(x) = x, and λ = α2. Now r(0) = 0, and of the two solutions Jn(αx) and Yn(α x) only Jn(α x) is bounded at x = 0. Thus in view of (16) above, the set {Jn(αi x)}, i = 1, 2, 3, …, is orthogonal with respect to the weight function p(x) = x on an interval [0, b]. The orthogonality relation is

x Jn(αi x)Jn(αj x) dx = 0, λiλj, (24)

provided the αi, and hence the eigenvalues λi = , i = 1, 2, 3, …, are defined by means of a boundary condition at x = b of the type given in (5):

A2Jn(λb) + B2α(αb) = 0.(25)

The extra factor of α in (25) comes from the Chain Rule:

For any choice of A2 and B2, not both zero, it is known that (25) has an infinite number of roots xi = αi b. The eigenvalues are then λi = = (xi /b)2. More will be said about eigenvalues in the next chapter.

EXAMPLE 4 Legendre’s Equation

From the result given in (21) we can identify q(x) = 0, p(x) = 1, and λ = n(n + 1). Recall from Section 5.3 when n = 0, 1, 2, … Legendre’s DE possesses polynomial solutions Pn(x). Now we can put the observation that r(–1) = r(1) = 0 together with the fact that the Legendre polynomials Pn(x) are the only solutions of (21) that are bounded on the closed interval [–1, 1], to conclude from (18) that the set {Pn(x)}, n = 0, 1, 2, …, is orthogonal with respect to the weight function p(x) = 1 on [–1, 1]. The orthogonality relation is

Pm(x)Pn(x) dx = 0, mn.

REMARKS

(i) A Sturm–Liouville problem is also considered to be singular when the interval under consideration is infinite. See Problems 9 and 10 in Exercises 12.5.

(ii) Even when the conditions on the coefficients p, q, r, and r′ are as assumed in the regular Sturm–Liouville problem, if the boundary conditions are periodic, then property (b) of Theorem 12.5.1 does not hold. You are asked to show in Problem 4 of Exercises 12.5 that corresponding to each eigenvalue of the BVP

there exist two linearly independent eigenfunctions.

12.5 Exercises Answers to selected odd-numbered problems begin on page ANS-32.

In Problems 1 and 2, find the eigenfunctions and the equation that defines the eigenvalues for the given boundary-value problem. Use a CAS to approximate the first four eigenvalues λ1, λ2, λ3, and λ4. Give the eigenfunctions corresponding to these approximations.

  1. y″ + λy = 0, y′(0) = 0, y(1) + y′(1) = 0
  2. y″ + λy = 0, y(0) + y′(0) = 0, y(1) = 0
  3. Consider y″ + λy = 0 subject to y′(0) = 0, y′(L) = 0. Show that the eigenfunctions are

    .

    This set, which is orthogonal on [0, L], is the basis for the Fourier cosine series.

  4. Consider y″ + λy = 0 subject to the periodic boundary conditions y(–L) = y(L), y′(–L) = y′(L). Show that the eigenfunctions are

    This set, which is orthogonal on [–L, L], is the basis for the Fourier series.

  5. Find the square norm of each eigenfunction in Problem 1.
  6. Show that for the eigenfunctions in Example 2,

    1. Find the eigenvalues and eigenfunctions of the boundary-value problem

      x2y″ + xy′ + λy = 0, y(1) = 0, y(5) = 0.

    2. Put the differential equation in self-adjoint form.
    3. Give an orthogonality relation.
    1. Find the eigenvalues and eigenfunctions of the boundary-value problem

      y″ + y′ + λy = 0, y(0) = 0, y(2) = 0.

    2. Put the differential equation in self-adjoint form.
    3. Give an orthogonality relation.
  7. Laguerre’s differential equation

    has polynomial solutions Ln(x). Put the equation in self-adjoint form and give an orthogonality relation on (0, ).

  8. Hermite’s differential equation

    has polynomial solutions Hn(x). Put the equation in self-adjoint form and give an orthogonality relation on (–, ).

  9. Consider the regular Sturm–Liouville problem:

    1. Find the eigenvalues and eigenfunctions of the boundary-value problem. [Hint: Let x = tan θ and then use the Chain Rule.]
    2. Give an orthogonality relation.
    1. Find the eigenfunctions and the equation that defines the eigenvalues for the boundary-value problem

      y is bounded at x = 0, y(3) = 0.

    2. Use Table 5.3.1 of Section 5.3 to find the approximate values of the first four eigenvalues λ1, λ2, λ3, and λ4.

Discussion Problem

  1. Consider the special case of the regular Sturm–Liouville problem on the interval [a, b]:

    Is λ = 0 an eigenvalue of the problem? Defend your answer.

Computer Lab Assignments

    1. Give an orthogonality relation for the Sturm–Liouville problem in Problem 1.
    2. Use a CAS as an aid in verifying the orthogonality relation for the eigenfunctions y1 and y2 that correspond to the first two eigenvalues λ1 and λ2, respectively.
    1. Give an orthogonality relation for the Sturm–Liouville problem in Problem 2.
    2. Use a CAS as an aid in verifying the orthogonality relation for the eigenfunctions y1 and y2 that correspond to the first two eigenvalues λ1 and λ2, respectively.

 

*Conditions (16) and (17) are equivalent to choosing A1 = 0, B1 = 0 in (4), and A2 = 0, B2 = 0 in (5), respectively.