1. Introduction
The Sturm–Liouville equation has been studied in almost every possible detail, and the several impressive properties of its solutions have been elucidated. Although a long-time established classical in the theory of differential equations, its mathematical facets, as well as applications, are diverse and still studied in recent times [
1,
2,
3], with extensions also going beyond the original formulation of the problem [
4,
5,
6,
7,
8]. A straightforward view on the classical theory is offered by the theory of integral equations. Nevertheless, we still deem it interesting to highlight some simple ideas and results that help to further clarify the simplicity of the theory. In particular, via the Implicit Function Theorem, we obtain a result that reinforces Sturm’s statement about the upper and lower bounds for the number of zeros of the solutions. The validity of the equipartition of the energy is stated for
. Its extension to
complex is set in
Appendix B.
The eigenfunctions
, satisfying the conditions
, are similar, locally, for large
, to
, with
,
x in the neighbourhood of
; see
Section 3.
For the eigenvalues problem, with the exception of the non-compact sets present in
Section 7, we restricted ourselves everywhere to the eigenfunctions
,
, satisfying null boundary conditions, i.e.,
. The treatment of the other two classical boundary conditions for the eigenvalue problem is straightforward.
The Sturm–Louville equation can be reduced, via a Liouville transform, to
, with
[
9]. However, the transform is not very simple, also changing the independent variable, and the results, in turn, must be translated back to the original variables. This simplification does not compromise the generality of the treatment entirely. This is very useful for the theory, but in this way, we lose some information about the original problem, since the solutions do not maintain track of the evolution of the original function. It is not difficult to understand that, with
, the results will tend asymptotically to those of
. We prefer to maintain
, and make only a simple transform that reduces the equation to
, i.e., an anharmonic oscillator with a non-constant elastic parameter. We prove what can be seen and easily tested in numerical simulations. Then, having exact results, there comes the possibility to define rigorously the behaviour of a set of anharmonic oscillators. It is surprising that, even if the Sturm–Liouville equation has been extensively studied also in the complex variable, no use has been made of the holomorphic theory to obtain the closure of the set of eigenfunctions in the space of square integrable functions
. We shall use the following notation for intervals in the real line:
(making explicit references solely to the cases
or
). In
Section 7,
and
will refer to the square integrable functions:
To recall the definition of the closure property, we proceed to briefly illustrate our findings. The solution
is a twice continuously differentiable function in the variable
x, i.e.,
is
, and holomorphic in the parameter
; see Theorem 1. The procedure, particularly simple, is the following. From the condition
valid for all the eigenfunctions, we deduce, through a Phragmén–Lindelöf theorem, that
is identically verified for
, and then it follows that the function
has to be the null function. The latter implication is precisely the definition of the closure of the set of functions
in
. In the case of the simplest eigenvalue problem, computations are easily made and Liouville’s theorem is enough to deduce the closure of the set of functions
.
The closure of the Hermite and Laguerre polynomials in non-compact sets, with convenient weight functions, is well established. We give an alternative proof of it that does not use the generating functions of such polynomials, nor the Weierstrass Theorem for compact sets, as in the celebrated proof attributed to J. von Neumann.
Also, the possibility to define general integral transforms is considered, with the conclusion that, apart from the Fourier and Hankel transforms, this possibility does not exist.
At various points, in addition to the p-integrable functions mentioned above, we shall use the following notations for real-valued functions on intervals I in the real line .
denotes an n-th times continuously differentiable function, .
For the particular case of the solutions , where is a complex parameter, we shall also consider the case
denotes a bounded variation function, . In this paper, we shall only refer to such functions when the interval is .
Absolutely continuous functions and Lipschitz continuous functions (shortly Lipschitz functions) shall also be intended for , .
In fact, with the exception of
Section 7 and
Section 8, when omitted, it will always be
.
2. Present Analysis
Many classical books consider, in detail, the properties of the solutions of the Sturm–Liouville equation, for example [
1,
10]. The equation
is considered, with the usual conditions:
The operator
results to be Hermitian in the space of functions
and null at the boundary.
The eigenvalue problem for the equation
with the three classical boundary conditions is similar to the elementary case
so nothing substantially new emerges in the apparently more general case of Equation (
2).
Nonetheless, one may observe that a general linear equation of second order can always be reduced, with a change in variables [
11] (p. 16), to the form
The condition
is not truly restrictive, since if
changes into
, then the eigenvalues
change into
.
We start simplifying the Sturm–Liouville equation by a simple change in variables
. Choosing
we obtain, again using the old variables,
with the new functions
and
still greater than zero. Here,
is a monotonic increasing function of
x that takes the interval
into
. So, nothing changes substantially using this new variable.
To be definite, for the eigenvalues and eigenfunctions, we restrict ourselves to the interval
, with null boundary conditions. We know that any solution of Equation (
5) not identically zero has only zeros of first order. We have the following:
Proposition 1. Any solution of Equation (5) cannot have infinite zeros in the interval . Proof. An accumulation point of zeros would be a zero of second order. □
Proposition 2. If , then the solutions cannot have other zeros in the interval if it is not identically zero.
Proof. In fact, and would have the same sign between two consecutive zeros of , which is impossible. □
Proposition 3. If , then the solution will be a regular bell between two consecutive zeros.
Proof. In fact is zero only where is zero, and the local maxima of will be at points where (since will be a monotonic function between consecutive zeros). □
For completeness, we report on the following theorem, with the aim to obtain also bounds for the solutions.
Theorem 1. The solutions of Equation (5) are in x and analytic for Remark 1. Then, must hold and the coefficients turn out to be a closed set (see Theorem 8).
Proof of Theorem 1. Equation (
5), with
, leads us to the integral equation
with the definition
. Iterating the integral Equation (
6), and setting
, the solution is given by
where, for
,
is the iterated kernel, dominated by
having defined
. The last integral, containing
, goes to zero in
m, in any compact set of the complex plane. The solution is thus given by the series
uniformly convergent in any such set. This proves the holomorphy of
in the whole complex plane
. □
From Equation (
6),
and, from Gronwall’s inequality,
In
Appendix B, an exponential lower bound is also found for
and any
, Equation (
A11).
If we consider the integral equation
we obtain the general bound
(notice that, for the particular case
real and
, better bounds can be easily found)
Now, we can also rewrite the solution in the following form:
Substituting it into Equation (
5), we have the infinite system of differential equations for the coefficients
The system (
10) and (
11) can be solved recursively.
Note that the homogeneous equation for Equation (
11) is always Equation (
10). The integral equation connected to the differential Equation (
11), with initial conditions
is
The convergence of the series (
9) for every value of
is straightforward. Further, note that the signs of
are alternating for
: if
, then
gives
and, equally, if
, then
. This shows a first indication of the connection of
to
(and to
for
), as will be well clarified in what follows. For the definition of
k see
Section 3. The series expansion (
9) and Equations (
10) and (
11) also provide, in fact, the possibility to recover, for example, the Bessel functions (not reported here).
We already noted that the zeros in x of have to be simple, otherwise is identically zero. We have the following Lemma:
Lemma 1. If , the followingholds, implying that, if , thenThat is, a zero of of the first order with respect to x is a zero of first order also with respect to λ. Furthermore, if then Proof. From Equation (
5), written for
and
, and
, we deduce
and
Subtracting and adding
, and passing to the limit
, the statement follows. □
Remark 2. Equation (13) is valid also for complex λ. It is useful for us to underline the results with the following corollaries.
Corollary 1. At the zeros of , the derivativesare inversely proportional, have the same sign, and their product has the same behavior of . Corollary 2. The norm of the eigenfunctions is given by For example, for the Bessel functions
, which are in the lucky situation of being analytical functions of
, with
, noting that the derivation is with respect to
, we recover
with
the eigenvalues of
for
.
From Equation (
16), we also recover the Orthogonality of eigenfunctions of Hermitian operators.
Having established these results, we now proceed to state a qualitative property we stress in this paper.
Proposition 4. All the zeros and all stationary points (and then every point) of the solutions of the Sturm–Liouville equation for are strictly decreasing functions of λ.
Proof of Proposition 4. If
is a zero of
, from the Implicit Function Theorem, being
in
, we can deduce that
is identically zero in a small neighbourhood of
, with
a regular function of
. Then
From the previous Corollary 1, we have that
and
have the same sign at the zeros of
, so
is negative and the result follows. This result is obvious in the elementary case
.
This proposition reinforces Sturm’s result, which gives lower and upper bounds for the number of zeros in the interval .
For the stationary points of
i.e.,
, we have that, while
is positive,
is always negative for
. So, applying the Implicit Function Theorem to
, since
is different from zero, we have
□
In the rest of the paper, when there is no possible confusion, shall mean .
For , the solution with is always greater than zero. Then, from the previous proposition, also considering Sturm’s Theorem, when increasing there exists a first value such that , for . Then, there exists a value such that vanishes at the boundaries and has just one zero inside the interval . So, going on, we find a set of infinite eigenvalues , such that has exactly zeros inside the interval . If it were , then should have ∞ zeros inside the interval , which is impossible. Still in the light of the previous proposition, we have the following property:
Proposition 5. There cannot be two different eigenfunctions of the Sturm–Liouville equation satisfying the same null boundary conditions, and having the same number of zeros inside the interval (0,1).
That is, the eigenfunctions are uniquely determined, but for a constant factor, by the number of their zeros.
The existence of eigenvalues
is easily deduced, as it is well known, also from the theory of integral equations, and from direct methods of calculus of variations [
12,
13,
14,
15,
16,
17], considering that the eigenvalue
is the minimum of the integral
in the space of normalized functions, with
and orthogonal to the eigenfunctions
If we impose the condition
, little changes conceptually, since between two consecutive zeros
there exists a unique point with
, and this point shifts to the left increasing
(see Proposition 4). So, there exists
such that
. If we consider the condition
the function
behaves, varying
, like the tangent function, so the condition (
21) can be satisfied, for any value of
and
, for an infinite set of values of
.
3. Some Preliminary Results
With
and
, we can give detailed information about the asymptotic behavior of
. For
and
, from the integral equation, equivalent to the differential Equation (
5),
it is easy to obtain, via the Gronwall inequality, the asymptotic estimate
and then
Here and in the following, we shall use to denote that there exists a such that, for any , then , with , positive numbers and the intended for all x in . Similarly, the same symbol shall be attributed to similar estimates of quantities independent of x and whose dependence on k may be explicit or implicit. Additionally, if and it turns out that , then we shall write .
From Equation (
22) it also follows that
The same conclusion (
23) is obtained by defining
meaning, from Sturm’s Theorem, that the square of the frequency of
is contained between these two values. For
, we have, for the two
,
that is, asymptotically, the eigenvalues
are given by
giving again for
the result (
23), which we can now write as
This substantial identification of
with
suggests directly, as we shall see, that the theorem for the point convergence of the Fourier series is still valid, including obviously the completeness of the set (
26).
For
and
, the analysis is more complex. We can write the integral equation, for any
k,
The comparison of Equations (
22) and (
27) shows the great simplification of considering
. Naturally, the integral Equation (
27) can be useful only for
, where
and
,
, with the square bracket in the integral controlled by
. Here and in the following, by
, we mean
greater than any arbitrarily large number. From Proposition 7 and from Proposition A3 of
Appendix C, it turns out that the two terms at the second member of (
27) are of the same order.
Now, the single oscillation depends only on the local values of
and
. Thus, for
, when writing the integral equation in any given interval
, where
, since the interval
is at most of the order of
, the extreme values of
have to be considered only on small intervals. Therefore, the local value of
is dominated by a constant for
, a Lipschitz function. Then, in the general case
, we have
with the first term now prevailing with respect to the integral, meaning that we cannot have that a single
approximates
, but solely that it is well behaved in any small interval, where
In any case,
since
is bounded by a constant
as can be seen without the need to invoke the expression (
28), but rather just by integrating the differential equation and considering the next Equation (
34). The same bounds can be directly obtained in
Appendix B, Equation (
A16), for small
. Furthermore, we have the following Lemma:
Lemma 2. Any derivation of with respect to x produces a factor in the numerator, as can be seen in Equations (27) and (28) and confirmed by the asymptotic relation . Conversely, the same equations entail that any derivation with respect to λ gives a factor in the denominator (consider also Lemma 1 and Corollary 1). Given
strictly contained between two consecutive zeros
, the derivatives of
, a regular bell, are small corrections to the derivatives of
around
itself. We wrote the integral Equation (
28) to stress that locally, and asymptotically, the solution is like that of the elementary case
, but with a different frequency for any single oscillation.
Equation (
28) lets us also estimate the eigenvalues. Letting
, we can estimate
from (
28), and then, from Equations (
25) and (
26), evaluated at
, it has to be
M connected to the Lipschitz constant. So, asymptotically,
, i.e.
Naturally, for
,
holds, in agreement with (
25).
It is easy to prove directly from the differential equation the following Lemma:
Lemma 3. The solutions of the equationinitially decrease by increasing λ: if , holds for , where is the first zero of . In fact, the inequality holds in absolute value up to the intersection of the two solutions, from which point, for the needs of the following corollary, the reasoning can be reiterated. Proof. From Equation (
16),
showing that
is a decreasing function, since
, starting from
. □
Corollary 3. If large, then .
Proof. Take large, and , where is sufficiently small such that all the intersections of the two solutions are close to their zeros. From the previous Lemma, starting from these intersections, it follows that locally . The assertion is then true, for any , just going with n steps from to . But now, the number of oscillations is different, and the statement is true only for the absolute maximum. □
Remark 3. It can happen, for large and a little larger, that if has a flat minimum, the oscillation with can take a larger part in that interval, and the inequality can be locally reversed.
These results can be understood from Equations (
27) and (
28). Obviously we can state also the following:
Corollary 4. For and the same λ,
The surprising fact is that, using the apparently worse integral equation,
where there is no information about the oscillatory character of the solutions for
, we very easily obtain exact results, identical to those valid for the elementary case. First of all, we have that
just like
Furthermore, from
taking
as a stationary point of
, i.e.,
, we obtain, for any
,
So, if
is another stationary point, we have
In particular, for the first stationary point
, similarly to
we now have
and then, for any semi-oscillation
, and any
, we have that the mean value is exactly zero, with the weight function
. Observe that, if
,
is monotonically increasing with respect to
and
x, and there are no stationary points.
From the previous formulas, we again easily obtain that
for
. We have
where
is the last stationary point before
x. Now
, so the integral is of the order
, with the local estimate
Proposition 6. From (33), since , it follows that, while , we have That is, by virtue of the oscillatory character of the solutions, the integral goes to zero more rapidly by a factor
. The simplest example where we can observe the consequences of this oscillation is
with
and
. In fact, Equation (
34) is valid also for
with
a monotonic or bounded variation function. We state here the following proposition, whose proof will be postponed to
Appendix C.
Proposition 7. If is a monotonic, or a bounded variation function, then Knowing that for monotonic functions the derivative exists almost everywhere, and that such derivative is in , this proposition corresponds to the usual integration by parts, with functions and trigonometric functions. From the density of bounded variation functions in , we have then that the Riemann–Lebesgue Theorem is valid in general.
Theorem 3. If , then .
We conclude this section by writing some additional simple, more useful, results ahead. We have
From a mechanical point of view, with
, what is now a total energy is no longer a constant of the motion, as we would expect.
is increasing or decreasing according to the sign of
From (
35), it follows that
also asymptotically for both signs of
. The result descends directly from Equation (
36) for
while, for
it follows from the equations
and
From the same Equations (
36) and (
38), changing the conditions on the sign of
, it always follows that
and for the maxima of
, taken at points where
, we have
Obviously, for
, inf
means
, and for
it means
. In conclusion, Equations (
39) and (
40) give
In the same way, we obtain
The solutions of our equations exactly satisfy what we still call “equipartition”, like in the case of the harmonic oscillator, with in the place of .
Theorem 4 (Equipartition).
It always holds thatwhich, for the harmonic oscillator , and for , reduces to Proof. It is enough to integrate
□
Remark 4. This continues to be approximately valid also for “complex potentials”, for small , as we will see in Appendix B. Using this equipartition, from Equation (
36), and
the stationary point nearest to 1, which for
satisfies
, it follows that
from which
leading to the final inequality
Since
then
Finally, from (
39), we obtain
then:
Proposition 8. The normalized solutions are uniformly bounded for .
4. The Closure Property
We now pass to the closure of the set of eigenfunctions in the space , i.e., we have to show that any function , orthogonal to all the eigenfunctions, is the null function, which is expected, since we have an eigenfunction for any and is a separable pace. In this section, is considered variable on the whole complex plane.
To get acquainted with the procedure, we first consider the classical problem of
We show that the entire holomorphic function
is identically zero for
. We cannot use the Carlson Theorem [
18] (p. 185), which says that if a holomorphic function
is such that
for every
and if, for
, it holds that
, then
. In fact, we now have the opposite inequality:
, for any
and
.
The bound is essential, e.g., barely not satisfied by .
In fact, in the present case, it is easy to obtain the expected result through the Liouville Theorem. We can appreciate that
is an entire holomorphic function, since
is a zero of the first order, and it is bounded for all
. Indeed, for
, we have
clearly bounded for
. The same will be true for
, uniformly in
n. It is enough to subtract
from the numerator of
and apply the same reasoning. The boundedness of
for
is clear from (
45). Then, from Liouville’s Theorem,
Take now
, and consider the Riemann–Lebesgue Theorem. The conclusion then is that
has to be zero:
giving now
a.e. as a consequence of the inverse sine Fourier transform of
, with
for
.
Remark 5. Equation (47) means, obviously, that also the integrals are identically zero for any n. Remark 6. The integral is not a counterexample, since it is zero only for even numbers.
For the eigenfunctions
, we cannot refer directly to Liouville’s Theorem but, considering also the bounds we found in
Appendix B, we can refer to the following formulation of the Phragmén–Lindelöf theorem [
18] (p. 177).
Theorem 5. If is holomorphic in the semi-plane , if it is bounded on the real axis andthen is bounded in the upper semi-plane. The same can be stated for the lower semi-plane, if is holomorphic on the full complex plane and the previous condition is satisfied. So the first step to the closure is the following theorem.
Theorem 6. Given the entire holomorphic functionif , then is identically zero: Proof. Proceeding as previously, define the function
which again results to be an entire holomorphic function, because the zeros in
of
are again simple. We have first to prove its boundedness on the real axis. The limitation on the compact set
is obvious. For any
n,
and
uniformly in
n from Equation (
17), from the bounds on the derivatives stated in Lemma 2, and finally from Equation (
42). We have to control that the conditions of Phragmén–Lindelöf Theorem are satisfied for
and then for
.
We know that
is represented locally by Equations (
27) and (
28), with the integral of the order of
, so everything goes as for Equation (
44), and we can conclude that
is uniformly bounded for
. For
is a function of exponential type, increasing in
x, and then
so the limitation of
along the full real axis is established. Also the limitation of
for
, and for any
, is obvious: Equation (
A9) in
Appendix B says that, for any
,
is strictly increasing, so again Equation (
50) is valid.
It remains to control
for
. At the end of
Appendix B, e.g., Equation (
A22), there is a bound of
, for
and
, slightly increasing in
. So we have a much better limitation than the one needed for the Phragmén–Lindelöf theorem, and
is bounded on the upper semi-plane. The same result is also true for the lower semi-plane, giving us the possibility now to use the Liouville Theorem
the constant depending on
. It is easy to see now that
has to be zero. Reasoning as in Equation (
46), taking
such that
,
, we have, from Equation (
40),
. On the other hand, from Theorem 3,
valid for any
. Then the conclusion
follows. □
Theorem 7. If fromit follows that is the null function. Proof. Now, we cannot infer directly the statement from the inverse sine Fourier transform. Here is a direct proof. Take
and
. Denote with
and
, respectively, the positive and negative parts of
in the interval
. Since
, we can write
Apply to these integrals the mean value theorem. Then
The function
is of an exponential type for
, so it is easy to understand that, for example,
, if
for any small
. This means that the points
approach, the extremum of the set where the function is different from zero. We need the following result. Knowing that
and
are positive increasing functions for
, see Equation (
30), then for
from which, for
,
Suppose then that
is the biggest one of those limits; divide all the terms of Equation (
52) by
, and take the limit
. The conclusion is that
. The same can subsequently be obtained for
. If it were
, it would directly be
. The choice of
x is arbitrary, so
The conclusion is
□
Remark 7. We know that if for a large class of functions , then it is . For that, the class of functions for example, is enough, , as we just proved.
The previous Theorem 7 can be rephrased in the following form:
Theorem 8. The set of coefficients of the series expansion (9) of form a closed set in the space . Proof. If the integrals of the coefficients satisfy the conditions
then the holomorphic function
is identically zero, (and vice versa) and,
being a zero function for Theorem 7, the conclusion follows. Lebesgue’s Theorem lets us interchange the integration with the sum of the series. □
In the next section, we will obtain the same result as Theorems 6 and 7 with a very different method, in the particular case of .
5. An Interesting Byproduct
Considering the integral Equation (
6), we can rewrite Equation (
51), valid for
, as
That is,
or
having defined
The conclusion is
with the implicit definition of the functions. Note that, with
,
. Easy conclusions follow if we suppose that
. In this case it has to be, for
complex,
That is, we are led to the eigenvalue problem
a little different from the classical Sturm–Liouville eigenvalue problem. We can think that asymptotically, there will not be a big difference.
It is interesting to underline that with and , we have the following:
Proposition 9. For normalized solutions, , showing the orthogonality of the derivatives with weight 1. See the discussion following the proof of Theorem 10. From this, we again have , .
The interesting result is that, from Theorem 6,
we obtain
and
, that is
Then, applying the result found for
to
, one has
Note that
. We now show that
is the null function.
We are led to consider the recursive sequence, starting from
,
which satisfies the conditions
,
, and the differential equation
The sequence converges uniformly to zero, since
, where
.
We have the following result:
Theorem 9. If, for , the sequencestarting from , satisfies the conditions (i) , and (ii) one of the functions of the sequence has a finite number of zeros, then the function is a zero function.
Remark 8. Observe that the condition (ii) is connected to the one needed for the validity of the Hankel integral transform [19]. Remark 9. Even if is identically zero in an interval , the function has at most one zero in that interval, becauseis constant for . Proof of Theorem 9. Starting from , if some has a finite number of zeros, there is an iterate without zeros, because, just looking at the differential equation and considering the convexity of the functions, we have that the number of zeros decreases in the iterates. Between the last zero of and 1, has no zeros. Furthermore, the number of further zeros of is less or equal than the number of its changes in convexity, and the latter number is equal to the number of zeros of the odd order of . That is, has at least one less zero than So, if has no zeros and has to satisfy condition (i), it has to be identically zero and, tracing back, all the previous functions have to be zero. □
The conclusion is the alternative proof of the following
Proposition 10. The sequence of eigenfunctions of the Sturm–Liouville equation, with , is closed in the space satisfying the hypothesis (ii) of the theorem.
6. The Uniform Convergence of the Fourier Series
We give a direct proof of the following result, stronger than the completeness property and similar to the property of the trigonometric series. In this section, we take normalized solutions , giving and uniformly bounded for .
Theorem 10. If is an absolutely continuous function, with and if (we already know that ), then the Fourier series converges uniformly to .
See the theorem in [
20], for the classic case.
Proof. We add at the moment the condition that
is a BV function. In
Appendix D, the proof will be given with the more general condition on
. Then
The last integral goes asymptotically as
according to Proposition A4 in
Appendix C, where it is the unnormalized
to be uniformly bounded, here replaced by
. So we have the absolute, uniform convergence for every
x
for the uniform boundedness of normalized
, and it is uniformly
, for the closure of the set of eigenfunctions. □
It is interesting to observe that the functions
are, asymptotically, “nearly” orthonormal, and exactly orthonormal if
. In fact, we have
The last terms are small, but there is nevertheless a problem of convergence: the sum in
m of
gives 1 for any
n, but for the last terms, the convergence has to be proved.
The closure (and completeness) of the set
in
can be stated independently through the methods of the calculus of variations [
17] (p. 160). Here, we observe that the space of
functions, null at the border, orthogonal to all the eigenfunctions
, is void. Otherwise, we could solve the problem of the minimum of the functional (
20) in this space [
13] (p. 198), and obtain an eigenfunction
that has to have a definite (finite) number of zeros inside the interval
, and then has to be already in the set
, showing that the space of
function, null at the border, orthogonal to all
, is void. Now, if
is an arbitrary
function, null at the border, in the integral
we can substitute to
its uniformly convergent Fourier series. So, if we have
then the given integral is identically zero for any
function
null at the border, and then
is the null function [
21]. Observe that in this way, we did not give an alternative proof of completeness, since the uniform convergence comes from Theorem 10, where the closure was needed.
8. A Final Remark About the Possibility of General Integral Transform
To conclude our reasoning, we ask ourselves if it is possible to define in general an integral transform like the Hankel transform. The idea could be to repeat what is true for the Bessel functions. In that case, the eigenvalues and the eigenfunctions are defined in the interval
. Then, consider the orthogonal condition given, from Formula (
18), by
where
are the eigenvalues of the Bessel functions of order
n. In this relation, we can turn to the continuous values
, and can expect, correctly in this case, that the symbol of Kronecker turns into the Dirac function, also via considering that the Bessel functions are analytic in the variable
, so we can extend the integration in
r to
. Obviously, in this “transition”, we need to introduce a factor of conversion from the symbol of Kronecker to the
function. The conclusion is that we have
and the symmetric relation
taking us to the Hankel transform
with the anti-transform
In this case,
k and
r are the respective factors of conversion.
Now we could be tempted to follow the same line of reasoning in more generality. From the orthogonal relation (and
indicating the 2-norm in
)
we could try to go to
with
a suitable factor of conversion. For example, from
we could try to go from the discrete variables
to the continuous
k, obtaining
and a similar expression taking the sum with respect to
n. Naturally, there lies the difficulty to go from
to
at the same time. But there are in fact more stringent difficulties. One is the asymmetry between
and
(or
). Now, contrary to the Bessel functions and trigonometric functions, which depend on the product
or
, there is no possibility, in general, to go from
to
(or
). Still more importantly, to pass from (
58) to (
59), the integral (
58) has to diverge for
going to
k. This is possible only if, remaining with a finite interval of integration, the integrand becomes singular in the limit, as in the case of Poisson kernel. This is not the case for Equation (
58).
So, unfortunately, the conclusion is that, except for the Fourier Transform and the Hankel transform, with the trigonometric functions and the Bessel functions being analytic in the product of varibles, there is no possibility to define in general other integral transforms, starting from the solutions of the Sturm–Liouville equation in the interval .
9. Discussion
We saw that, apart from some technical details, the steps to obtain some properties of the Sturm–Liouville equation are very simple. First of all, the use of the Implicit Function Theorem elucidates, in a clear way, the dependence of the zeros of the solutions of Equation (
5) on the positive parameter
. Furthermore, the complex analysis takes us directly to the closure property in
of the set of eigenfunctions of the Sturm–Liouville equation. In fact, retrospectively, it is natural to expect that the function
under the conditions
would be an entire holomorphic function bounded in the complex plane
, and then conclude that
gives
.
A non-trivial output of our procedure is that the closure we obtained in , and the uniform convergence of the Fourier series in the general case , can be deduced without referring to the Weierstrass Theorem. Furthermore, an analogue of the Riemann–Lebesgue Theorem and the equipartition of the energy are valid for the functions .
The principal conclusion of the paper is that we will not be wrong in identifying asymptotically, locally, the eigenfunctions of with for , , x in the neighbourhood of . Notable is the fact that some results for , with non-constant, are exactly the same as those for .
In this paper, we studied in detail only a finite interval, with the first classical boundary conditions, and non-singular equations, while, e.g., in the equations connected to Mathematical Physics singular equations, or unlimited intervals, do occur. But for these equations, we have polynomial solutions. Further, we showed in a simple way how the Laguerre and Hermite polynomials, with appropriate weight functions, constitute closed sets for not bounded intervals, without referring to their generating functions or to the Weierstrass Theorem for compact sets.
Finally, we concluded negatively on the possibility of defining, in general, an integral transform, apart from the well-known Fourier and Hankel transforms.