summaryrefslogtreecommitdiff
path: root/source/know/concept/sturm-liouville-theory/index.md
blob: 117d8937d47b322d77320ef68506d2f185f30b6b (plain)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
---
title: "Sturm-Liouville theory"
sort_title: "Sturm-Liouville theory"
date: 2021-02-23
categories:
- Mathematics
- Physics
layout: "concept"
---

**Sturm-Liouville theory** extends
the concept of Hermitian matrix eigenvalue problems
to linear second-order ordinary differential equations.

It states that, given suitable boundary conditions,
any such equation can be rewritten using the **Sturm-Liouville operator**,
and that the corresponding eigenvalue problem,
known as a **Sturm-Liouville problem**,
will give real eigenvalues and a complete set of eigenfunctions.



## General operator

Consider the most general form of a second-order linear
differential operator $$\hat{L}$$, where $$p_0(x)$$, $$p_1(x)$$, and $$p_2(x)$$
are real functions of $$x \in [a,b]$$ and are nonzero for all $$x \in \,\,]a, b[$$:

$$\begin{aligned}
    \hat{L} \{u(x)\}
    \equiv p_2(x) \: u''(x) + p_1(x) \: u'(x) + p_0(x) \: u(x)
\end{aligned}$$

Analogously to matrices,
we now define its **adjoint** operator $$\hat{L}^\dagger$$ as follows:

$$\begin{aligned}
    \inprod{\hat{L}^\dagger f}{g}
    \equiv \inprod{f}{\hat{L} g}
\end{aligned}$$

What is $$\hat{L}^\dagger$$, given the above definition of $$\hat{L}$$?
We start from the inner product $$\inprod{f}{\hat{L} g}$$:

$$\begin{aligned}
    \inprod{f}{\hat{L} g}
    &= \int_a^b f^*(x) \hat{L}\{g(x)\} \dd{x}
    = \int_a^b (f^* p_2) g'' + (f^* p_1) g' + (f^* p_0) g \dd{x}
    \\
    &= \Big[ (f^* p_2) g' + (f^* p_1) g \Big]_a^b - \int_a^b (f^* p_2)' g' + (f^* p_1)' g - (f^* p_0) g \dd{x}
    \\
    &= \Big[ f^* (p_2 g' + p_1 g) - (f^* p_2)' g \Big]_a^b + \int_a^b \! \Big( (f p_2)'' - (f p_1)' + (f p_0) \Big)^* g \dd{x}
    \\
    &= \Big[ f^* \big( p_2 g' + (p_1 - p_2') g \big) - (f^*)' p_2 g \Big]_a^b + \int_a^b \Big( \hat{L}^\dagger\{f\} \Big)^* g \dd{x}
\end{aligned}$$

The newly-formed operator on $$f$$ must be $$\hat{L}^\dagger$$,
but there is an additional boundary term.
To fix this, we demand that $$p_1(x) = p_2'(x)$$
and that $$\big[ p_2 (f^* g' - (f^*)' g) \big]_a^b = 0$$, leaving:

$$\begin{aligned}
    \inprod{f}{\hat{L} g}
    &= \Big[ f^* \big( p_2 g' + (p_1 - p_2') g \big) - (f^*)' p_2 g \Big]_a^b + \inprod{\hat{L}^\dagger f}{g}
    \\
    &= \Big[ p_2 \big( f^* g' - (f^*)' g \big) \Big]_a^b + \inprod{\hat{L}^\dagger f}{g}
    \\
    &= \inprod{\hat{L}^\dagger f}{g}
\end{aligned}$$

Let us look at the expression for $$\hat{L}^\dagger$$ we just found,
with the restriction $$p_1 = p_2'$$ in mind:

$$\begin{aligned}
    \hat{L}^\dagger \{f\}
    &= (p_2 f)'' - (p_1 f)' + (p_0 f)
    \\
    &= (p_2'' f + 2 p_2' f' + p_2 f'') - (p_1' f + p_1 f') + (p_0 f)
    \\
    &= p_2 f'' + (2 p_2' - p_1) f' + (p_2'' - p_1' + p_0) f
    \\
    &= p_2 f'' + p_1 f' + p_0 f
    \\
    &= \hat{L}\{f\}
\end{aligned}$$

So $$\hat{L}$$ is **self-adjoint**, i.e. $$\hat{L}^\dagger$$ is the same as $$\hat{L}$$!
Indeed, every such second-order linear operator is self-adjoint
if it satisfies the constraints $$p_1 = p_2'$$ and $$\big[ p_2 (f^* g' - (f^*)' g) \big]_a^b = 0$$.

But what if $$p_1 \neq p_2'$$?
Let us multiply $$\hat{L}$$ by an unknown $$p(x) \neq 0$$
and divide by $$p_2(x) \neq 0$$:

$$\begin{aligned}
    \frac{p}{p_2} \hat{L} \{u\}
    = p u'' + p \frac{p_1}{p_2} u' + p \frac{p_0}{p_2} u
\end{aligned}$$

We now demand that the derivative $$p'(x)$$ of the unknown $$p(x)$$ satisfies:

$$\begin{aligned}
    p'(x)
    = p(x) \frac{p_1(x)}{p_2(x)}
    \quad \implies \quad
    \frac{p_1(x)}{p_2(x)} \dd{x}
    = \frac{1}{p(x)} \dd{p}
\end{aligned}$$

Taking the indefinite integral of this differential equation
yields an expression for $$p(x)$$:

$$\begin{aligned}
    \int \frac{p_1(x)}{p_2(x)} \dd{x}
    = \int \frac{1}{p} \dd{p}
    = \ln\!\big( p(x) \big)
    \quad \implies \quad
    \boxed{
        p(x)
        = \exp\!\bigg( \int \frac{p_1(x)}{p_2(x)} \dd{x} \bigg)
    }
\end{aligned}$$

We define an additional function $$q(x)$$
based on the last term of $$(p / p_2) \hat{L}$$ shown above:

$$\begin{aligned}
    \boxed{
        q(x)
        \equiv p(x) \frac{p_0(x)}{p_2(x)}
    }
    = \frac{p_0(x)}{p_2(x)} \exp\!\bigg( \int \frac{p_1(x)}{p_2(x)} \dd{x} \bigg)
\end{aligned}$$

When rewritten using $$p$$ and $$q$$,
the modified operator $$(p / p_2) \hat{L}$$ looks like this:

$$\begin{aligned}
    \frac{p}{p_2} \hat{L} \{u\}
    = p u'' + p' u' + q u
    = (p u')' + q u
\end{aligned}$$

This is the self-adjoint form from earlier!
So even if $$p_1 \neq p_2'$$, any second-order linear operator
with $$p_2(x) \neq 0$$ can easily be made self-adjoint.
The resulting general form is called the **Sturm-Liouville operator** $$\hat{L}_\mathrm{SL}$$,
for nonzero $$p(x)$$:

$$\begin{aligned}
    \boxed{
        \begin{aligned}
            \hat{L}_\mathrm{SL} \{u(x)\}
            &= \Big( p(x) \: u'(x) \Big)' + q(x) \: u(x)
            \\
            &= \hat{L}_\mathrm{SL}^\dagger \{u(x)\}
        \end{aligned}
    }
\end{aligned}$$

Still subject to the constraint $$\big[ p (f^* g' - (f^*)' g) \big]_a^b = 0$$
such that $$\inprod{f}{\hat{L}_\mathrm{SL} g} = \inprod{\hat{L}_\mathrm{SL}^\dagger f}{g}$$.



## Eigenvalue problem

An eigenvalue problem of $$\hat{L}_\mathrm{SL}$$
is called a **Sturm-Liouville problem** (SLP).
The goal is to find the **eigenvalues** $$\lambda$$
and corresponding **eigenfunctions** $$u(x)$$ that fulfill:

$$\begin{aligned}
    \boxed{
        \hat{L}_\mathrm{SL}\{u(x)\} = - \lambda \: w(x) \: u(x)
    }
\end{aligned}$$

Where $$w(x)$$ is a real weight function satisfying $$w(x) > 0$$ for $$x \in \,\,]a, b[$$.
By convention, the trivial solution $$u = 0$$ is not valid.
Some authors have the opposite sign for $$\lambda$$ and/or $$w$$.

In our derivation of $$\hat{L}_\mathrm{SL}$$ above,
we imposed the constraint $$\big[ p (f^* g' - (f')^* g) \big]_a^b = 0$$ to ensure that
$$\inprod{\hat{L}_\mathrm{SL}^\dagger f}{g} = \inprod{f}{\hat{L}_\mathrm{SL} g}$$.
Consequently, to have a valid SLP,
the boundary conditions (BCs) on $$u$$ must be such that,
for any two (possibly identical) eigenfunctions $$u_m$$ and $$u_n$$, we have:

$$\begin{aligned}
    \Big[ p(x) \big( u_m^*(x) \: u_n'(x) - \big(u_m'(x)\big)^* u_n(x) \big) \Big]_a^b = 0
\end{aligned}$$

There are many boundary conditions that satisfy this requirement.
Some notable ones are listed non-exhaustively below.
Verify for yourself that these work:

+ **Dirichlet BCs**: $$u(a) = u(b) = 0$$
+ **Neumann BCs**: $$u'(a) = u'(b) = 0$$
+ **Robin BCs**: $$\alpha_1 u(a) + \beta_1 u'(a) = \alpha_2 u(b) + \beta_2 u'(b) = 0$$ with $$\alpha_{1,2}, \beta_{1,2} \in \mathbb{R}$$
+ **Periodic BCs**: $$p(a) = p(b)$$, $$u(a) = u(b)$$, and $$u'(a) = u'(b)$$
+ **Legendre "BCs"**: $$p(a) = p(b) = 0$$

If this is fulfilled, Sturm-Liouville theory gives us
useful information about $$\lambda$$ and $$u$$.
By definition, the following must be satisfied
for two arbitrary eigenfunctions $$u_m$$ and $$u_n$$:

$$\begin{aligned}
    0
    &= \hat{L}_\mathrm{SL}\{u_m^*\} + \lambda_m^* w u_m^*
    \\
    &= \hat{L}_\mathrm{SL}\{u_n\} + \lambda_n w u_n
\end{aligned}$$

We multiply each by the other eigenfunction,
subtract the results, and integrate:

$$\begin{aligned}
    0
    &= \int_a^b u_m^* \big(\hat{L}_\mathrm{SL}\{u_n\} + \lambda_n w u_n\big)
    - u_n \big(\hat{L}_\mathrm{SL}\{u_m^*\} + \lambda_m^* w u_m^*\big) \dd{x}
    \\
    &= \int_a^b u_m^* \hat{L}_\mathrm{SL}\{u_n\} - u_n \hat{L}_\mathrm{SL}\{u_m^*\}
    + (\lambda_n - \lambda_m^*) u_m^* w u_n \dd{x}
    \\
    &= \inprod{u_m}{\hat{L}_\mathrm{SL} u_n} - \inprod{\hat{L}_\mathrm{SL} u_m}{u_n}
    + (\lambda_n - \lambda_m^*) \inprod{u_m}{w u_n}
\end{aligned}$$

The operator $$\hat{L}_\mathrm{SL}$$ is self-adjoint of course,
so the first two terms vanish, leaving us with:

$$\begin{aligned}
    0
    &= (\lambda_n - \lambda_m^*) \inprod{u_m}{w u_n}
\end{aligned}$$

When $$m = n$$, we get $$\inprod{u_n}{w u_n} > 0$$,
so the equation is only satisfied if $$\lambda_n^* = \lambda_n$$,
meaning the eigenvalue $$\lambda_n$$ is real for any $$n$$.
When $$m \neq n$$, then $$\lambda_n - \lambda_m^*$$
may or may not be zero depending on the degeneracy.
If there is no degeneracy, then $$\lambda_n - \lambda_m^* \neq 0$$,
meaning $$\inprod{u_m}{w u_n} = 0$$, i.e. the eigenfunctions are orthogonal.
In case of degeneracy, manual orthogonalization is needed,
which is guaranteed to be doable using the [Gram-Schmidt method](/know/concept/gram-schmidt-method/).

In conclusion, an SLP has **real eigenvalues**
and **orthogonal eigenfunctions**: for all $$m$$, $$n$$:

$$\begin{aligned}
    \boxed{
        \lambda_n \in \mathbb{R}
    }
    \qquad\qquad
    \boxed{
        \inprod{u_m}{w u_n}
        = A_n \delta_{nm}
    }
\end{aligned}$$

When solving a differential eigenvalue problem,
knowing that all eigenvalues are real is a huge simplification,
so it is always worth checking whether you are dealing with an SLP.

Another useful fact:
it turns out that SLPs always have an infinite number of *discrete* eigenvalues.
Furthermore, there always exists a *lowest* eigenvalue $$\lambda_0 > -\infty$$,
called the **ground state**.



## Complete basis

Not only are an SLP's eigenfunctions orthogonal,
they also form a **complete basis**, meaning any well-behaved $$f(x)$$
can be expanded as a **generalized Fourier series** with coefficients $$a_n$$:

$$\begin{aligned}
    \boxed{
        f(x)
        = \sum_{n = 0}^\infty a_n u_n(x)
        \quad \mathrm{for} \: x \in \,\,]a, b[
    }
\end{aligned}$$

This series converges faster if $$f$$ satisfies the same BCs as $$u_n$$;
in that case the expansion is also valid for the inclusive interval $$x \in [a, b]$$.

To find an expression for the coefficients $$a_n$$,
we multiply the above generalized Fourier series by $$u_m^* w$$
and integrate it to get inner products on both sides:

$$\begin{aligned}
    u_m^* w f
    &= \sum_{n = 0}^\infty a_n u_m^* w u_n
    \\
    \int_a^b u_m^* w f \dd{x}
    &= \int_a^b \bigg( \sum_{n = 0}^\infty a_n u_m^* w u_n \bigg) \dd{x}
    \\
    \inprod{u_m}{w f}
    &= \sum_{n = 0}^\infty a_n \inprod{u_m}{w u_n}
\end{aligned}$$

Because the eigenfunctions of an SLP are mutually orthogonal,
the summation disappears:

$$\begin{aligned}
    \inprod{u_m}{w f}
    &= \sum_{n = 0}^\infty a_n \inprod{u_m}{w u_n}
    = \sum_{n = 0}^\infty a_n A_n \delta_{nm}
    = a_m A_m
\end{aligned}$$

After isolating this for $$a_m$$, we see that
the coefficients are given by the projection of the target
function $$f$$ onto the normalized eigenfunctions $$u_m / A_m$$:

$$\begin{aligned}
    \boxed{
        a_n
        = \frac{\inprod{u_n}{w f}}{A_n}
        = \frac{\inprod{u_n}{w f}}{\inprod{u_n}{w u_n}}
    }
\end{aligned}$$

As a final remark, we can see something interesting
by rearranging the generalized Fourier series
after inserting the expression for $$a_n$$:

$$\begin{aligned}
    f(x)
    &= \sum_{n = 0}^\infty \frac{1}{A_n} \inprod{u_n}{w f} u_n(x)
    \\
    &= \int_a^b \bigg(\sum_{n = 0}^\infty \frac{1}{A_n} u_n^*(\xi) \: w(\xi) \: f(\xi) \: u_n(x) \bigg) \dd{\xi}
    \\
    &= \int_a^b f(\xi) \bigg(\sum_{n = 0}^\infty \frac{1}{A_n} u_n^*(\xi) \: w(\xi) \: u_n(x) \bigg) \dd{\xi}
\end{aligned}$$

Upon closer inspection, the parenthesized summation
must be the [Dirac delta function](/know/concept/dirac-delta-function/) $$\delta(x)$$
for the integral to work out.
In fact, this is the underlying requirement for completeness:

$$\begin{aligned}
    \boxed{
        \sum_{n = 0}^\infty \frac{1}{A_n} u_n^*(\xi) \: w(\xi) \: u_n(x) = \delta(x - \xi)
    }
\end{aligned}$$



## References
1.  O. Bang,
    *Applied mathematics for physicists: lecture notes*, 2019,
    unpublished.