5.3 Special Functions

INTRODUCTION

The two differential equations

(1)

(2)

often arise in the advanced studies of applied mathematics, physics, and engineering. Equation (1) is called Bessel’s equation of order ν and is named after the noted German astronomer and mathematician Friedrich Wilhelm Bessel (1784–1846), who was the first person to determine the accurate distance from the Sun to another star. Bessel first encountered a special form of equation (1) in his study of elliptical planetary motion and eventually carried out a systematic study of the properties of the solutions of the general equation. Differential equation (2) is known as Legendre’s equation of order n and is named after the French mathematician Adrien-Marie Legendre (1752–1833). When we solve (1), we will assume that ν ≥ 0; whereas in (2), we will consider only the case where n is a nonnegative integer.

A stamp of Friedrich Wilhelm.

Friedrich Wilhelm Bessel

© PjrStamps/Alamy Stock Photo

5.3.1 Bessel Functions

The Solution

Because x = 0 is a regular singular point of Bessel’s equation, we know that there exists at least one solution of the form . Substituting the last expression into (1) then gives

(3)

From (3) we see that the indicial equation is r2ν2 = 0 so that the indicial roots are r1 = ν and r2 = −ν. When r1 = ν, (3) becomes

Therefore, by the usual argument we can write (1 + 2ν)c1 = 0 and

(k + 2)(k + 2 + 2ν)ck + 2 + ck = 0

or ck+2 = ,     k = 0, 1, 2, . . . . (4)

The choice c1 = 0 in (4) implies c3 = c5 = c7 = … = 0, so for k = 0, 2, 4, … we find, after letting k + 2 = 2n, n = 1, 2, 3, . . ., that

. (5)

Thus (6)

It is standard practice to choose c0 to be a specific value—namely,

c0 =

where Γ(1 + ν) is the gamma function. See Appendix A. Since this latter function possesses the convenient property Γ(1 + α) = αΓ(α), we can reduce the indicated product in the denominator of (6) to one term. For example,

Hence we can write (6) as

for n = 0, 1, 2, . . . .

Bessel Functions of the First Kind

The series solution is usually denoted by Jν(x):

(7)

If ν ≥ 0, the series converges at least on the interval [0, ∞). Also, for the second exponent r2 = −ν we obtain, in exactly the same manner,

(8)

The functions Jν(x) and Jν(x) are called Bessel functions of the first kind of order ν and −ν, respectively. Depending on the value of ν, (8) may contain negative powers of x and hence converge on the interval (0, ∞).*

Now some care must be taken in writing the general solution of (1). When ν = 0, it is apparent that (7) and (8) are the same. If ν > 0 and r1r2 = ν − (−ν) = 2ν is not a positive integer, it follows from Case I of Section 5.2 that Jν(x) and Jν(x) are linearly independent solutions of (1) on (0, ∞), and so the general solution on the interval is y = c1Jν(x) + c2Jν(x). But we also know from Case II of Section 5.2 that when r1r2 = 2ν is a positive integer, a second series solution of (1) may exist. In this second case we distinguish two possibilities. When ν = m = positive integer, Jm(x) defined by (8) and Jm(x) are not linearly independent solutions. It can be shown that Jm is a constant multiple of Jm (see Property (i) on page 295). In addition, r1r2 = 2ν can be a positive integer when ν is half an odd positive integer. It can be shown in this latter event that Jν(x) and Jν(x) are linearly independent. In other words, the general solution of (1) on (0, ∞) is

(9)

The graphs of y = J0(x) (blue) and y = J1(x) (red) are given in FIGURE 5.3.1.

Five curves are graphed on an x y plane. The curves follow an oscillatory pattern. The curve labeled J subscript 0 starts at the point (0, 1), goes down and to the right through the positive x axis, and reaches the approximate point (4, negative 0.4). Then it goes up and to the right through the positive x axis, reaches the approximate point (7, 0.28), again goes down and to the right, and ends at the approximate point (8, 0.16). The curve labeled J subscript 1 starts at the point (0, 0), goes up and to the right, and reaches the approximate point (2, 0.6). Then it goes down and to the right through the positive x axis, reaches the approximate point (5.5, negative 0.32), again goes up and to the right through the positive x axis, and ends at the approximate point (8, 0.2). The third curve starts at the point (0, 0), goes up and to the right, and reaches the approximate point (3.5, 0.48). Then it goes down and to the right through the positive x axis, reaches the approximate point (6.5, negative 0.28), again goes up and to the right, and ends at the approximate point (8, negative 0.12). The fourth curve starts at the approximate point (1, 0), goes up and to the right, and reaches the approximate point (4.25, 0.44). Then it goes down and to the right through the positive x axis, and ends at the approximate point (8, negative 0.3). The fifth curve starts at the approximate point (1.5, 0), goes up and to the right, and reaches the approximate point (5.5, 0.4). Then it goes down and to the right through the positive x axis, and ends at the approximate point (8, negative 0.1).

FIGURE 5.3.1 Bessel functions of the first kind for n = 0, 1, 2, 3, 4

EXAMPLE 1 General Solution: ν Not an Integer

By identifying ν2 = and ν = we can see from (9) that the general solution of the equation x2y″ + xy′ + (x2)y = 0 on (0, ∞) is .

Bessel Functions of the Second Kind

If ν ≠ integer, the function defined by the linear combination

(10)

and the function Jν(x) are linearly independent solutions of (1). Thus another form of the general solution of (1) is y = c1Jν(x) + c2Yν(x), provided ν ≠ integer. As νm, m an integer, (10) has the indeterminate form 0/0. However, it can be shown by L’Hôpital’s rule that limνm Yν(x) exists. Moreover, the function

and Jm(x) are linearly independent solutions of x2y″ + xy′ + (x2m2)y = 0. Hence for any value of ν the general solution of (1) on the interval (0, ∞) can be written as

(11)

Yν(x) is called the Bessel function of the second kind of order ν. FIGURE 5.3.2 shows the graphs of Y0(x) (blue) and Y1(x) (red). Because the Bessel function of the second kind of order ν was first introduced in 1867 by the German mathematician Carl Gottfried Neumann (1832–1925), it is also referred to as the Neumann function of order ν. The Neumann function is often denoted by Nν

Five curves are graphed on an x y plane. The curve labeled Y subscript 0 starts at the approximate point (0, negative 2.4), goes sharply up and to the right through the positive x axis, and reaches the approximate point (2, 0.5). Then it goes down and to the right through the positive x axis, reaches the approximate point (5.5, negative 0.5), again goes up and to the right through the positive x axis, and ends at the approximate point (8, 0.2). The curve labeled Y subscript 1 starts at the approximate point (0.2, negative 3), goes sharply up and to the right through the positive x axis, and reaches the approximate point (4, 0.4). Then it goes down and to the right through the positive x axis, reaches the approximate point (5, negative 0.3), again goes up and to the right, and ends at the approximate point (8, negative 0.1). The third curve starts at the approximate point (0.75, negative 3), goes sharply up and to the right through the positive x axis, and reaches the approximate point (5, 0.4). Then it goes down and to the right through the positive x axis, and ends at the approximate point (8, negative 0.25). The fourth curve starts at the approximate point (1.25, negative 3), goes sharply up and to the right through the positive x axis, and reaches the approximate point (6, 0.3). Then it goes down and to the right, and ends at the approximate point (8, 0). The fifth curve starts at the point (2, negative 3), goes sharply up and to the right through the positive x axis, and reaches the approximate point (7, 0.3). Then it goes slightly down and to the right, and ends at the approximate point (8, 0.25).

FIGURE 5.3.2 Bessel functions of the second kind for n = 0, 1, 2, 3, 4

EXAMPLE 2 General Solution: ν an Integer

By identifying ν2 = 9 and ν = 3 we see from (11) that the general solution of the equation x2y″ + xy′ + (x2 − 9)y = 0 on (0, ∞) is .

DEs Solvable in Terms of Bessel Functions

Sometimes it is possible to transform a differential equation into equation (1) by means of a change of variable. We can then express the solution of the original equation in terms of Bessel functions. For example, if we let t = αx, α > 0, in

(12)

then by the Chain Rule,

Accordingly (12) becomes

The last equation is Bessel’s equation of order ν with solution y = c1Jν(t) + c2Yν(t). By resubstituting t = αx in the last expression we find that the general solution of (12) on the interval (0, ∞) is

(13)

Equation (12), called the parametric Bessel equation of order ν, and its general solution (13) are very important in the study of certain boundary-value problems involving partial differential equations that are expressed in cylindrical coordinates.

Modified Bessel Functions

Another equation that bears a resemblance to (1) is the modified Bessel equation of order ν,

(14)

This DE can be solved in the manner just illustrated for (12). This time if we let t = ix, where i2 = −1, then (14) becomes

Since solutions of the last DE are Jν(t) and Yν(t), complex-valued solutions of equation (14) are Jν(ix) and Yν(ix). A real-valued solution, called the modified Bessel function of the first kind of order ν, is defined in terms of Jν(ix):

See Problem 21 in Exercises 5.3. The general solution of (14) is

(15)

When ν is an integer n the functions In(x) and I−n (x) are not linearly independent on the interval (0, ∞). So, analogous to (10), we define the modified Bessel function of the second kind of order ν ≠ integer to be

(16)

and for integral ν = n,

Because Iν and Kν are linearly independent on the interval (0, ∞) for any value of ν, the general solution of (14) is

(17)

The graphs of I0(x) (blue) and I1(x) (red) are given in FIGURE 5.3.3 and the graphs K0(x) (blue) and K1(x) (red) are shown in FIGURE 5.3.4. Unlike the Bessel functions of the first and second kinds, the graphs of the modified Bessel functions of the first kind and second kind are not oscillatory. Moreover, the graphs in Figures 5.3.3 and 5.3.4 illustrate the fact that the modified Bessel functions In(x) and Kn(x), n = 0, 1, 2, . . . have no real zeros in the interval (0, ∞). Also, note that Kn(x) → ∞ as x → 0+.

Five curves are graphed on an x y plane. The curve labeled I subscript 0 starts at the point (0, 1), goes sharply up and to the right, and ends at the approximate point (2.5, 4). The curve labeled I subscript 1 starts at the point (0, 0), goes sharply up and to the right, and ends at the approximate point (2.75, 4). The third curve starts at the point (0, 0), goes sharply up and to the right, and ends at the approximate point (3.5, 4). The fourth curve starts at the point (0, 0), goes slightly to the right to the approximate point (1, 0), then goes sharply up and to the right, and ends at the approximate point (4.2, 4). The fifth curve starts at the point (0, 0), goes slightly to the right to the approximate point (1.75, 0), then goes sharply up and to the right, and ends at the approximate point (4.8, 4).

FIGURE 5.3.3 Modified Bessel function of the first kind for n = 0, 1, 2, 3, 4

Five curves are graphed on an x y plane. The curve labeled K subscript 0 starts at the approximate point (0.1, 4), goes sharply down and to the right to the approximate point (2, 0.1), then goes to the right, and ends at the approximate point (8, 0). The curve labeled K subscript 1 starts at the approximate point (0.25, 4), goes sharply down and to the right to the approximate point (2, 0.2), then goes to the right, and ends at the approximate point (8, 0). The third curve starts at the approximate point (0.75, 4), goes sharply down and to the right to the approximate point (2, 0.3), then goes to the right, and ends at the approximate point (8, 0). The fourth curve starts at the approximate point (1.25, 4), goes sharply down and to the right to the approximate point (2, 0.4), then goes to the right, and ends at the approximate point (8, 0). The fifth curve starts at the approximate point (1.75, 4), goes sharply down and to the right to the approximate point (3, 0.5), then goes to the right, and ends at the approximate point (8, 0).

FIGURE 5.3.4 Modified Bessel function of the second kind for n = 0, 1, 2, 3, 4

Proceeding as we did in (12) and (13), we see that the general solution of the parametric form of the modified Bessel equation of order ν

(18)

on the interval (0, ∞) is

(19)

EXAMPLE 3 Parametric Modified Bessel Equation

By identifying α2 = 25, ν2 = 4, α = 5, and ν = 2 it follows from (19) that the general solution of the equation on (0, ∞) is .

Yet another equation, important because many differential equations fit into its form by appropriate choices of the parameters, is

(20)

Although we shall not supply the details, the general solution of (20),

(21)

can be found by means of a change in both the independent and the dependent variables: z = bxc, y(x) = w(z). If p is not an integer, then Yp in (21) can be replaced by Jp.

EXAMPLE 4 Using (20)

Find the general solution of xy″ + 3y′ + 9y = 0 on (0, ∞).

SOLUTION

By writing the given DE as

we can make the following identifications with (20):

The first and third equations imply a = −1 and c = . With these values the second and fourth equations are satisfied by taking b = 6 and p = 2. From (21) we find that the general solution of the given DE on the interval (0, ∞) is

Aging Spring and Bessel Functions

In Section 3.8 we saw that a mathematical model for the free undamped motion of a mass m on an aging spring is given by We are now in a position to find the general solution of this differential equation.

EXAMPLE 5 Aging Spring Revisited

Use the substitution to solve the differential equation .

SOLUTION

By the Chain Rule,

Thus

Multiplying the last equation by then gives

or

The last equation is recognized as (1) with and where the symbols x and s play the roles of y and x, respectively. Hence from (11), the general solution of the last differential equation is Finally, if we substitute s, then the general solution of is seen to be

The other model discussed in Section 5.1 of a spring whose characteristics change with time was mx″ + ktx = 0. By dividing through by m we see that the equation x″ + (k/m) tx = 0 is Airy’s equation, y″ + α2xy = 0. See the discussion following Example 2 in Section 5.1. The general solution of Airy’s differential equation can also be written in terms of Bessel functions. See Problems 34, 35, and 46 in Exercises 5.3.

Properties

We list below a few of the more useful properties of Bessel functions of the first and second kinds of order m, m = 0, 1, 2, . . . :

Note that Property (ii) indicates that Jm(x) is an even function if m is an even integer and an odd function if m is an odd integer. The graphs of Y0(x) and Y1(x) in Figure 5.3.2 illustrate Property (iv): Ym(x) is unbounded at the origin. This last fact is not obvious from (10). The solutions of the Bessel equation of order 0 can be obtained using the solutions y1(x) in (21) and y2(x) in (22) of Section 5.2.It can be shown that (21) of Section 5.2 is y1(x) = J0(x), whereas (22) of that section is

The Bessel function of the second kind of order 0, Y0(x), is then defined to be the linear combination for x > 0. That is,

where γ = 0.57721566 . . . is Euler’s constant. Because of the presence of the logarithmic term, it is apparent that Y0(x) is discontinuous at x = 0.

Numerical Values

The first five nonnegative zeros of J0(x), J1(x), Y0(x), and Y1(x) are given in Table 5.3.1. Some additional functional values of these four functions are given in Table 5.3.2.

TABLE 5.3.1 Zeros of J0, J1, Y0, and Y1
The table captioned zeros of J subscript 0 (x), J subscript 1 (x), Y subscript 0 (x), and Y subscript 1 (x) has four columns and five rows. The column headings are as follows: Column 1, J subscript 0 (x). Column 2, J subscript 1 (x). Column 3, Y subscript 0 (x). Column 4, Y subscript 1 (x). The row entries are as follows: Row 1. J subscript 0 (x), 2.4048. J subscript 1 (x), 0.0000. Y subscript 0 (x), 0.8936. Y subscript 1 (x), 2.1971. Row 2. J subscript 0 (x), 5.5201. J subscript 1 (x), 3.8317. Y subscript 0 (x), 3.9577. Y subscript 1 (x), 5.4297. Row 3. J subscript 0 (x), 8.6537. J subscript 1 (x), 7.0156. Y subscript 0 (x), 7.0861. Y subscript 1 (x), 8.5960. Row 4. J subscript 0 (x), 11.7915. J subscript 1 (x), 10.1735. Y subscript 0 (x), 10.2223. Y subscript 1 (x), 11.7492. Row 5. J subscript 0 (x), 14.9309. J subscript 1 (x), 13.3237. Y subscript 0 (x), 13.3611. Y subscript 1 (x), 14.8974.
TABLE 5.3.2 Numerical Values of J0, J1, Y0, and Y1
The table captioned Numerical Values of J subscript 0 (x), J subscript 1 (x), Y subscript 0 (x), and Y subscript 1 (x) has five columns and 16 rows. The column headings are as follows: Column 1, x. Column 2, J subscript 0 (x). Column 3, J subscript 1 (x). Column 4, Y subscript 0 (x). Column 5, Y subscript 1 (x). The row entries are as follows: Row 1. x, 0. J subscript 0 (x), 1.0000. J subscript 1 (x), 0.0000. Y subscript 0 (x), ‘blank’. Y subscript 1 (x), ‘blank’. Row 2. x, 1. J subscript 0 (x), 0.7652. J subscript 1 (x), 0.4401. Y subscript 0 (x), 0.0883. Y subscript 1 (x), negative 0.7812. Row 3. x, 2. J subscript 0 (x), 0.2239. J subscript 1 (x), 0.5767. Y subscript 0 (x), 0.5104. Y subscript 1 (x), negative 0.1070. Row 4. x, 3. J subscript 0 (x), negative 0.2601. J subscript 1 (x), 0.3391. Y subscript 0 (x), 0.3769. Y subscript 1 (x), 0.3247. Row 5. x, 4. J subscript 0 (x), negative 0.3971. J subscript 1 (x), negative 0.0660. Y subscript 0 (x), negative 0.0169. Y subscript 1 (x), 0.3979. Row 6. x, 5. J subscript 0 (x), negative 0.1776. J subscript 1 (x), negative 0.3276. Y subscript 0 (x), negative 0.3085. Y subscript 1 (x), 0.1479. Row 7. x, 6. J subscript 0 (x), 0.1506. J subscript 1 (x), negative 0.2767. Y subscript 0 (x), negative 0.2882. Y subscript 1 (x), negative 0.1750. Row 8. x, 7. J subscript 0 (x), 0.3001. J subscript 1 (x), negative 0.0047. Y subscript 0 (x), negative 0.0259. Y subscript 1 (x), negative 0.3027. Row 9. x, 8. J subscript 0 (x), 0.1717. J subscript 1 (x), 0.2346. Y subscript 0 (x), 0.2235. Y subscript 1 (x), negative 0.1581. Row 10. x, 9. J subscript 0 (x), negative 0.0903. J subscript 1 (x), 0.2453. Y subscript 0 (x), 0.2499. Y subscript 1 (x), 0.1043. Row 11. x, 10. J subscript 0 (x), negative 0.2459. J subscript 1 (x), 0.0435. Y subscript 0 (x), 0.0557. Y subscript 1 (x), 0.2490. Row 12. x, 11. J subscript 0 (x), negative 0.1712. J subscript 1 (x), negative 0.1768. Y subscript 0 (x), negative 0.1688. Y subscript 1 (x), 0.1637. Row 13. x, 12. J subscript 0 (x), 0.0477. J subscript 1 (x), negative 0.2234. Y subscript 0 (x), negative 0.2252. Y subscript 1 (x), negative 0.0571. Row 14. x, 13. J subscript 0 (x), 0.2069. J subscript 1 (x), negative 0.0703. Y subscript 0 (x), negative 0.0782. Y subscript 1 (x), negative 0.2101. Row 15. x, 14. J subscript 0 (x), 0.1711. J subscript 1 (x), 0.1334. Y subscript 0 (x), 0.1272. Y subscript 1 (x), negative 0.1666. Row 16. x, 15. J subscript 0 (x), negative 0.0142. J subscript 1 (x), 0.2051. Y subscript 0 (x), 0.2055. Y subscript 1 (x), 0.0211.

Differential Recurrence Relation

Recurrence formulas that relate Bessel functions of different orders are important in theory and in applications. In the next example we derive a differential recurrence relation.

EXAMPLE 6 Derivation Using the Series Definition

Derive the formula     xJ′ν(x) = νJν(x) − xJν+1(x).

SOLUTION

It follows from (7) that

The result in Example 6 can be written in an alternative form. Dividing by x gives

This last expression is recognized as a linear first-order differential equation in Jν(x). Multiplying both sides of the equality by the integrating factor xν then yields

(22)

It can be shown in a similar manner that

(23)

See Problem 27 in Exercises 5.3. The differential recurrence relations (22) and (23) are also valid for the Bessel function of the second kind Yν(x). Observe that when ν = 0 it follows from (22) that

(24)

Results similar to (24) also hold for the modified Bessel functions of the first and second kind of order ν = 0:

(25)

See Problems 43 and 45 in Exercises 5.3 for applications of these derivatives.

Bessel Functions of Half-Integral Order

When the order ν is half an odd integer, that is, . . ., Bessel functions of the first and second kinds can be expressed in terms of the elementary functions sin x, cos x, and powers of x. To see this let’s consider the case when From (7) we have

See Appendix A.

In view of the properties the gamma function, and the fact that the values of Γ(1 + + n) for n = 0, n = 1, n = 2, and n = 3 are, respectively,

In general,

Hence,

The infinite series in the last line is the Maclaurin series for sin x, and so we have shown that

(26)

We leave it as an exercise to show that

(27)

See Problems 31 and 32 in Exercises 5.3.

If n is an integer, then the order is half an odd integer. Because and , we see from (10) that

(28)

For n = 0 and n = −1 in the last formula, we get, in turn, Y1/2(x) = −J−1/2 (x) and Y−1/2(x) = J1/2 (x). In view of (26) and (27) these results are the same as

(29)

and (30)

Spherical Bessel Functions

Bessel functions of half-integral order are used to define two more important functions:

     and      (31)

The function jn(x) is called the spherical Bessel function of the first kind and yn(x) is the spherical Bessel function of the second kind. For example, by using (26) and (29) we see that for the expressions in (31) become

and            

The graphs of Jn(x) and yy(x) for n ≥ 0 are very similar to those given in Figures 5.3.1 and 5.3.2; that is, both functions are oscillatory, and yn(x) becomes unbounded as The graphs of J0(x) (blue) and y0(x) (red) are given in FIGURE 5.3.5. See Problems 39 and 40 in Exercises 5.3.

Two curves are graphed on an x y plane. The curve labeled j subscript 0 follows an oscillatory pattern. It starts at the point (0, 1), goes down and to the right through the positive x axis, and reaches the approximate point (4, negative 0.25). Then, it goes up and to the right through the positive x axis, reaches the approximate point (8, 0.2), again goes down and to the right through the positive x axis, and ends at the approximate point (10, negative 0.1). The curve labeled y subscript 0 starts at the approximate point (0.4, negative 3), goes sharply up and to the right through the positive x axis, and reaches the approximate point (3, 0.4). Then it goes down and to the right through the positive x axis, reaches the approximate point (6, negative 0.2), then goes up and to the right through the positive x axis, and ends at the approximate point (10, 0.1).

FIGURE 5.3.5 Spherical Bessel functions j0(x) and y0(x)

Spherical Bessel functions arise in the solution of a special partial differential equation expressed in spherical coordinates. See Problems 41 and 42 in Exercises 5.3 and Problem 14 in Exercises 14.3.

5.3.2 Legendre Functions

The Solution

Since x = 0 is an ordinary point of Legendre’s equation (2), we substitute the series , shift summation indices, and combine series to get

which implies that n(n + 1)c0 + 2c2 = 0

(n − 1)(n + 2)c1 + 6c3 = 0

(j + 2)(j + 1)cj+2 + (nj)(n + j + 1)cj = 0

or (32)

Letting j take on the values 2, 3, 4, . . ., recurrence relation (32) yields

c4 = −

and so on. Thus for at least |x| < 1 we obtain two linearly independent power series solutions:

(33)

Notice that if n is an even integer, the first series terminates, whereas y2(x) is an infinite series. For example, if n = 4, then

Similarly, when n is an odd integer, the series for y2(x) terminates with xn; that is, when n is a nonnegative integer, we obtain an nth-degree polynomial solution of Legendre’s equation.

Since we know that a constant multiple of a solution of Legendre’s equation is also a solution, it is traditional to choose specific values for c0 or c1, depending on whether n is an even or odd positive integer, respectively. For n = 0 we choose c0 = 1, and for n = 2, 4, 6, . . .

c1 = (−1)n/2 ;

whereas for n = 1 we choose c0 = 1, and for n = 3, 5, 7, …,

c1 = (−1)(n−1)/2 .

For example, when n = 4 we have

y1(x) = (−1)4/2 (35x4 − 30x2 + 3).

Legendre Polynomials

These specific nth-degree polynomial solutions are called Legendre polynomials and are denoted by Pn(x). From the series for y1(x) and y2(x) and from the above choices of c0 and c1 we find that the first six Legendre polynomials are

(34)

Remember, P0(x), P1(x), P2(x), P3(x), …, are, in turn, particular solutions of the differential equations

(35)

n = 3: (1 − x2)y″ − 2xy′ + 12y = 0

⁝ ⁝

The graphs, on the interval [−1, 1], of the six Legendre polynomials in (34) are given in FIGURE 5.3.6.

Two straight lines and four curves are graphed on an x y plane. The line labeled P subscript 0 is horizontal and starts at the point (negative 1, 1), goes to the right, and ends at the point (1, 1). The line labeled P subscript 1 starts at the point (negative 1, negative 1), goes up and to the right through the point (0, 0), and ends at the point (1, 1). The first curve is in the form of semicircle, and starts at the point (negative 1, 1), goes down and to the right through the positive x axis, and reaches the point (0, negative 0.5). Then it goes up and to the right through the positive x axis, and ends at the point (1, 1). The second curve is in the form of a wave, and starts at the point (negative 1, negative 1), goes up and to the right through the positive x axis, and reaches the point (negative 0.5, 0.4375). Then it goes down and to the right through the point (0, 0), reaches the point (0.5, negative 0.4375), again goes up and to the right through the positive x axis, and ends at the point (1, 1). The third curve follows an oscillatory pattern, and starts at the point (negative 1, 1), goes down and to the right through the positive x axis, and reaches a point in the third quadrant. Then it goes up and to the right through the positive x axis, reaches the point (0, 0.375), again goes down and to the right through the positive x axis, reaches a point in the fourth quadrant, again goes up and to the right through the positive x axis, and ends at the point (1, 1). The fourth curve follows an oscillatory pattern, and starts at the point (negative 1, negative 1), goes up and to the right through the positive x axis, and reaches a point in the second quadrant. Then it goes down and to the right through the positive x axis, reaches a point in the third quadrant, again goes up and to the right through the point (0, 0), and reaches a point in the first quadrant. Then, it goes down and to the right through the positive x axis, reaches a point in the fourth quadrant, again goes up and to the right through the positive x axis, and ends at the point (1, 1).

FIGURE 5.3.6 Legendre polynomials for n = 0, 1, 2, 3, 4, 5

Properties

You are encouraged to verify the following properties for the Legendre polynomials in (34):

(35)

Property (i) indicates, as is apparent in Figure 5.3.6, that Pn(x) is an even or odd function according to whether n is even or odd.

Recurrence Relation

Recurrence relations that relate Legendre polynomials of different degrees are also important in some aspects of their applications. We state, without proof, the following three-term recurrence relation

(36)

which is valid for k = 1, 2, 3, … . In (34) we listed the first six Legendre polynomials. If, say, we wish to find P6(x), we can use (36) with k = 5. This relation expresses P6(x) in terms of the known P4(x) and P5(x). See Problem 51 in Exercises 5.3.

Another formula, although not a recurrence relation, can generate the Legendre polynomials by differentiation:

(37)

The formula given in (37) first appeared in 1815 in the doctoral dissertation of the French mathematician Benjamin Olinde Rodrigues (1795–1851). So naturally, (37) is called Rodrigues’ formula. There are several such derivative formulas for generating polynomial solutions of differential equations with variable coefficients. See Problems 29 and 30 in Chapter 5 in Review.

REMARKS

Although we have assumed that the parameter n in Legendre’s differential equation

(1 − x2)y″ − 2xy′ + n(n + 1)y = 0

represented a nonnegative integer, in a more general setting n can represent any real number. If n is not a nonnegative integer, then both Legendre functions y1(x) and y2(x) given in (33) are infinite series convergent on the open interval (–1, 1) and divergent (unbounded) at x = ±1. If n is a nonnegative integer, then as we have just seen one of the Legendre functions in (33) is a polynomial and the other is an infinite series convergent for −1 < x < 1. You should be aware of the fact that Legendre’s equation possesses solutions that are bounded on the closed interval [–1, 1] only in the case when n = 0, 1, 2, … . More to the point, the only Legendre functions that are bounded on the closed interval [–1, 1] are the Legendre polynomials Pn(x) or constant multiples of these polynomials. See Problem 53 in Exercises 5.3 and Problem 24 in Chapter 5 in Review.

5.3 Exercises Answers to selected odd-numbered problems begin on page ANS-13.

5.3.1 Bessel Functions

In Problems 1–6, use (1) to find the general solution of the given differential equation on (0, ∞).

  1. x2 y″ + xy′ + (x2)y = 0
  2. x2 y″ + xy′ + (x2 − 1)y = 0
  3. 4x2 y″ + 4xy′ + (4x2 − 25)y = 0
  4. 16x2 y″ + 16xy′ + (16x2 − 1)y = 0
  5. xy″ + y′ + xy = 0

In Problems 7 and 8, use (12) to find the general solution of the given differential equation on the interval (0, ∞).

In Problems 9 and 10, use (18) to find the general solution of the given differential equation on (0, ∞).

In Problems 11 and 12, use the indicated change of variable to find the general solution of the given differential equation on the interval (0, ∞).

In Problems 13–20, use (20) to find the general solution of the given differential equation on the interval (0, ∞).

  1. Use the series in (7) to verify that Iν (x) = iνJν (ix) is a real function.
  2. Assume that b in equation (20) can be pure imaginary; that is, b = βi, β > 0, i2 = −1. Use this assumption to express the general solution of the given differential equation in terms of the modified Bessel functions In and Kn.

    (a) y″x2 y = 0

    (b) xy″ + y′ − 7x3y = 0

In Problems 23–26, first use (20) to express the general solution of the given differential equation in terms of Bessel functions. Then use (26) and (27) to express the general solution in terms of elementary functions.

  1. y″ + y = 0
  2. x2 y″ + 4xy′ + (x2 + 2)y = 0
  3. 16x2 y″ + 32xy′ + (x4 − 12)y = 0
  4. 4x2 y″ − 4xy′ + (16x2 + 3)y = 0
  5. (a) Proceed as in Example 6 to show that

    [Hint: Write 2n + ν = 2(n + ν) − ν.]

    (b) Use the result in part (a) to derive (23).

  6. Use the formula obtained in Example 6 along with part (a) of Problem 27 to derive the recurrence relation

    2ν Jν (x) = x Jν + 1(x) + xJν1(x).

In Problems 29 and 30, use (22) or (23) to obtain the given result.

  1. J′0(x) = J1(x) = −J1(x)
    1. Proceed as on pages 296–297 to derive the elementary form of J−1/2(x) given in (27).
    2. Use along with (26) and (27) in the recurrence relation in Problem 28 to express J−3/2(x) in terms of sin x, cos x, and powers of x.
    3. Use a graphing utility to plot the graph of J−3/2(x).
    1. Use the recurrence relation in Problem 28 to express J3/2(x), J5/2(x), and J7/2(x) in terms of sin x, cos x, and powers of x.
    2. Use a graphing utility to plot the graphs of J3/2(x), J5/2(x), and J7/2(x) in the same coordinate plane.
  2. Use the solution of the aging spring equation , obtained in Example 5, to discuss the behavior of as in the three cases

    (a)

    (b)

    (c)

  3. Show that y = x1/2w(αx3/2) is a solution of the given form of Airy’s differential equation whenever w is a solution of the indicated Bessel’s equation. [Hint: After differentiating, substituting, and simplifying, then let t = αx3/2.]

    (a) y″ + α2xy = 0, x > 0; t2w″ + tw′ + (t2)w = 0, t > 0

    (b) y″α2xy = 0, x > 0; t2w″ + tw′ − (t2 + )w = 0, t > 0

  4. Use the result in parts (a) and (b) of Problem 34 to express the general solution on (0, ∞) of each of the two forms of Airy’s equation in terms of Bessel functions.
  5. Use Table 5.3.1 to find the first three positive eigenvalues and corresponding eigenfunctions of the boundary-value problem

    xy″ + y′ + λxy = 0,

    y(x), y′(x) bounded as x → 01, y(2) = 0.

    [Hint: By identifying λ = α2, the DE is the parametric Bessel equation of order zero.]

    1. Use (20) to show that the general solution of the differential equation xy″ + λy = 0 on the interval (0, ∞) is

    2. Verify by direct substitution that y = is a particular solution of the DE in the case λ = 1.
  6. Use Iν(x) = iνJν(ix) along with (7) and (8) to show that:

    (a) I1/2 (x) =

    (b) I−1/2(x) =

    (c) Use (16) to express K1/2(x) in terms of elementary functions.

    1. Use the first formula in (31) to find the spherical Bessel functions j1(x), j2(x), and j3(x).
    2. Use a graphing utility to plot the graphs of j1(x), j2(x), and j3(x) in the same coordinate plane.
    1. Use the second formula in (31) to find the spherical Bessel functions y1(x), y2(x), and y3(x).
    2. Use a graphing utility to plot the graphs of y1(x), y2(x), and y3(x) in the same coordinate plane.
  7. If n is an integer, use the substitution to show that the differential equation

    (38)

    becomes

    (39)

    1. In Problem 41, find the general solution of the differential equation in (39) on the interval (0, ∞).
    2. Use part (a) to find the general solution of the differential equation in (38) on the interval (0, ∞).
    3. Use part (b) to express the general solution of (38) in terms of the spherical Bessel functions of the first and second kind defined in (31).

Mathematical Model

  1. Cooling Fin A cooling fin is an outward projection from a mechanical or electronic device from which heat can be radiated away from the device into the surrounding medium (such as air). See the accompanying photo. An annular, or ring-shaped, cooling fin is normally used on cylindrical surfaces such as a circular heating pipe. See FIGURE 5.3.7. In the latter case, let r denote the radial distance measured from the center line of the pipe and T(r) the temperature within the fin defined for r0rr1. It can be shown that T(r) satisfies the differential equation

    ,

    where a2 is a constant and Tm is the constant air temperature. Suppose r0 = 1, r1 = 3, and Tm = 70. Use the substitution w(r) = T(r) − 70 to show that the solution of the given differential equation subject to the boundary conditions

    is ,

    where I0(x) and K0(x) are the modified Bessel functions of the first and second kind. You will also have to use the derivatives given in (25) on page 296.

    An annular cooling fin is used on a cylindrical pipe. The radius of the pipe is labeled r subscript 0. The radius of the cooling fin is labeled r subscript 1.

    FIGURE 5.3.7 Annular cooling fin

    A photo of Cooling fins on a motorcycle engine.

    Cooling fins on a motorcycle engine

    © GDragan/iStock/Thinkstock

  2. Solve the differential equation in Problem 43 when the boundary conditions are

    .

Computer Lab Assignments

  1. (a) Use the general solution given in Example 5 to solve the IVP

    Also use J′0(x) = −J1(x) and Y′0(x) = −Y1(x) along with Table 5.3.1 or a CAS to evaluate coefficients.

    (b) Use a CAS to graph the solution obtained in part (a) for 0 ≤ t < ∞.

  2. (a) Use the general solution obtained in Problem 35 to solve the IVP

    Use a CAS to evaluate coefficients.

    (b) Use a CAS to graph the solution obtained in part (a) for 0 ≤ t ≤ 200.

  3. Column Bending Under Its Own Weight A uniform thin column of length L, positioned vertically with one end embedded in the ground, will deflect, or bend away, from the vertical under the influence of its own weight when its length or height exceeds a certain critical value. It can be shown that the angular deflection θ(x) of the column from the vertical at a point P(x) is a solution of the boundary-value problem

    where E is Young’s modulus, I is the cross-sectional moment of inertia, δ is the constant linear density, and x is the distance along the column measured from its base. See FIGURE 5.3.8.

    A uniform thin column of length L is placed vertically with one end embedded in the ground. The distance along the column measured from its base is labeled x. x equals 0 at base. A vertical line is drawn from the base. The column is deflected toward the right from the vertical line. A point P(x) is marked on the column. The angle of deflection of the column from the vertical line at the point P is labeled theta.

    FIGURE 5.3.8 Column in Problem 47

    The column will bend only for those values of L for which the boundary-value problem has a nontrivial solution.

    1. Restate the boundary-value problem by making the change of variables t = Lx. Then use the results of a problem earlier in this exercise set to express the general solution of the differential equation in terms of Bessel functions.
    2. Use the general solution found in part (a) to find a solution of the BVP and an equation that defines the critical length L, that is, the smallest value of L for which the column will start to bend.
    3. With the aid of a CAS, find the critical length L of a solid steel rod of radius r = 0.05 in., δg = 0.28 A lb/in., E = 2.6 × 107 lb/in.2 , A = πr2, and I = πr4.
  4. Buckling of a Thin Vertical Column In Example 4 of Section 3.9 we saw that when a constant vertical compressive force, or load, P was applied to a thin column of uniform cross section and hinged at both ends, the deflection y(x) is a solution of the boundary-value problem:

    1. If the bending stiffness factor EI is proportional to x, then EI(x) = kx, where k is a constant of proportionality. If EI(L) = kL = M is the maximum stiffness factor, then k = M/L and so EI(x) = Mx/L. Use the information in Problem 37 to find a solution of

      if it is known that is not zero at x = 0.

    2. Use Table 5.3.1 to find the Euler load P1 for the column.
    3. Use a CAS to graph the first buckling mode y1(x) corresponding to the Euler load P1. For simplicity assume that c1 = 1 and L = 1.
  5. Pendulum of Varying Length For the simple pendulum described on page 194 of Section 3.11, suppose that the rod holding the mass m at one end is replaced by a flexible wire and that the wire is fed by a pulley through the horizontal support at point O in FIGURE 5.3.9. In this manner, while it is in motion in a vertical plane, the mass m can be raised or lowered. In other words, the length l(t) of the pendulum varies with time. Under the same assumptions leading to equation (6) in Section 3.11, it follows from (1) in Chapter 3 in Review that the differential equation for the displacement angle θ(t) is now

    1. If l increases at a constant rate v and l(0) = l0, then show that a linearization of the foregoing differential equation is

      (40)

    2. Make the change of variables x = (l0 + vt)/v and show that (40) becomes

    3. Use part (b) and (20) to express the general solution of equation (40) in terms of Bessel functions.
    4. Use the general solution obtained in part (c) to solve the initial-value problem consisting of equation (40) and the initial conditions θ(0) = θ0, θ′(0) = 0. [Hints: To simplify calculations use a further change of variable u = Also, recall (22) holdsfor both J1(u) and y1(u). Finally, the identity

      J1(u)y2(u) − J2(u)y1(u) = −

      will be helpful.]

    5. Use a CAS to graph the solution θ(t) of the IVP in part (d) when l0 = 1 ft, θ0 = radian, and v = ft/s. Experiment with the graph using different time intervals such as [0, 10], [0, 30], and so on.
    6. What do the graphs indicate about the displacement angle θ(t) as the length l of the wire increases with time?
A pendulum has a flexible wire connected to a pulley at the top. The wire is fed by the pulley through a horizontal support at point O. The wire from the horizontal support has a mass m hanging at its bottom and a length of l(t). It is deflected to the right from the vertical from point O. The angle of deflection of the wire from the vertical at the point O is labeled theta.

FIGURE 5.3.9 Pendulum of varying length in Problem 49

5.3.2 Legendre Functions

    1. Use the explicit solutions y1(x) and y2(x) of Legendre’s equation given in (33) and the appropriate choice of c0 and c1 to find the Legendre polynomials P6(x) and P7(x).
    2. Write the differential equations for which P6(x) and P7(x) are particular solutions.
  1. Use the recurrence relation (36) and P0(x) = 1, P1(x) = x, to generate the next six Legendre polynomials.
  2. Show that the differential equation

    can be transformed into Legendre’s equation by means of the substitution x = cos θ.

  3. Find the first three positive values of λ for which the problem

    (1 − x2)y″ − 2xy′ + λy = 0,

    y(0) = 0, y(x), y′(x) bounded on [−1, 1]

    has nontrivial solutions.

  4. The differential equation

    is known as the associated Legendre equation. When m = 0 this equation reduces to Legendre’s equation (2). A solution of the associated equation is

    where Pn(x), n = 0, 1, 2, . . . are the Legendre polynomials given in (34). The solutions for . . ., are called associated Legendre functions.

    1. Find the associated Legendre functions
    2. What can you say about when m is an even nonnegative integer?
    3. What can you say about when m is a nonnegative integer and
    4. Verify that satisfies the associated Legendre equation when and

Computer Lab Assignments

  1. For purposes of this problem, ignore the list of Legendre polynomials given on page 299 and the graphs given in Figure 5.3.6. Use Rodrigues’ formula (37) to generate the Legendre polynomials P1(x), P2(x), . . . , P7(x). Use a CAS to carry out the differentiations and simplifications.
  2. Use a CAS to graph P1(x), P2(x), . . . , P7(x) on the closed interval [−1, 1].
  3. Use a root-finding application to find the zeros of P1(x), P2(x), . . . , P7(x). If the Legendre polynomials are built-in functions of your CAS, find the zeros of Legendre polynomials of higher degree. Form a conjecture about the location of the zeros of any Legendre polynomial Pn(x), and then investigate to see whether it is true.

 

*When we replace x by |x|, the series given in (7) and (8) converge for 0 < |x| < ∞.