Sturm-Liouville Problems
Sturm-Liouville Problems
1 INTRODUCTION
1
2
Each section, except the introduction, ends with comments. Here we make some
historical remarks, mention the names of some of the major contributors, give
references for further reading, state some problems, etc. These problems reflect the
interests and knowledge, or lack thereof, of the author. No effort has been made to
classify these problems by difficulty, some may be routine, others intractible.
Although the subject of Sturm-Liouville problems is over 160 years old a surpris-
ing number of the results surveyed here are of recent origin, some were published
within the last couple of years and a few are not in print at the time of this writing.
Instructions for downloading the SLEIGN2 package, including a FORTRAN code
to compute eigenvalues and eigenfunctions for regular and singular SLP, from the
internet are given in Subsectiom 5.3. This package consists of 8 files and a number
of associated papers, all on SLP. The “readme” file has more detailed instructions
and some additional information. The code “sleign2” comes with a user friendly
interface and can be used by novices and experts alike. It can also be used to ap-
proximate the continuous (essential) spectrum of singular problems when combined
with some theoretical results. These can be found in the associated papers which
can also be downloaded.
I take this opportunity to thank Don Hinton and the Mathematics Department of
the University of Tennessee, Knoxville, particularly Phil Schaefer and John Conway,
for the invitation to give the Barrett Lectures of 1996. This forced me to try to
organize my notes and my thoughts on SLP. Also Don Hinton has been one of my
mathematical heroes from the time we were graduate students together until now.
Special thanks also to my pre-Ph.D. teachers John Neuberger, William Mahavier,
and the late John Barrett for stimulating and encouraging my interest and curiosity
about the wonderful subject of Mathematics. And to my chief post-Ph.D. teacher
and friend, Norrie Everitt, with whom I have been privileged to collaborate for
some twenty five years now.
Also I wish to thank my colleague Qingkai Kong for finding and correcting my
errors. His criticisms have significantly improved the final version of the paper in
several respects. But I am solely responsible for any remaining errors.
Last but not least I thank my wife Sandra for her help with the database for the
references and, especially, for helping with the hardware and software problems that
arose during the typing of these notes. Also for her tolerance and understanding
during this and many other Mathematics projects.
The world of Mathematics is full of wonders and of mysteries, at least as much
so as the physical world.
2.1 Introduction.
This section is devoted to the study of basic properties of first order systems.
NOTATION. An open interval is denoted by (a, b) with −∞ ≤ a < b ≤ ∞; [a, b]
always denotes the compact interval with (finite) left endpoint a and (finite) right
endpoint b. Let R denote the reals, C the complex numbers, and
For any interval J of the real line, open, closed, half open, bounded or unbounded,
by L(J) we denote the linear manifold of complex valued Lebesgue measurable
functions y defined on J for which
Z b Z Z
|y(t)| dt ≡ |y(t)| dt ≡ |y| < ∞.
a J J
The notation Lloc (J) is used to denote the linear manifold of functions y sat-
isfying y ∈ L([α, β]) for all compact intervals [α, β] ⊆ J. If J = [a, b], then
Lloc (J) = L(J). Also, we denote by ACloc (J) the collection of complex-valued
functions y which are absolutely continuous on all compact intervals [α, β] ⊆ J.
For a given set S, Mn,m (S) denotes the set of n×m matrices with entries from S.
If n = m we write Mn (S); also if m = 1 we sometimes write S n for Mn,1 (S). The
norm of a constant matrix as well as the norm of a matrix function P is denoted
by |P |. This may be taken as
X
|P | = |pij |.
DEFINITION 2.1 (Solution). Let J be any interval, open, closed, half open,
bounded or unbounded; let n, m ∈ N, let P : J → Mn (C), F : J → Mn,m (C).
By a solution of the equation Y 0 = P Y + F on J we mean a function Y from J
into Mn,m (C) which is absolutely continuous on all compact subintervals of J and
satisfies the equation a.e. on J. A matrix function is absolutely continuous if each
of its components is absolutely continuous.
THEOREM 2.2. Let J be any interval, open, closed, half open, bounded or
unbounded; let n, m ∈ N. If
and
(2.3) Y 0 = PY + F
4
Z t
(2.5) Y (t) = C + (P Y + F ), t ∈ J.
u
Conversely, every solution of the integral equation (2.5) is also a solution of the
IVP (2.3) (2.4).
Choose c in J, c 6= u. We show that (2.3), (2.4 ) has a unique solution on [u, c]
if c > u and on [c, u] if c < u . Assume c > u. Let
Rt
(2.6) kY k = sup {e−K u
|P (s)|ds
|Y (t)|, t ∈ [u, c]},
where K is a fixed positive constant K > 1. It is easy to see that with this norm B
is a Banach space. Let the operator T : B → B be defined by
Z t
(2.7) (T Y )(t) = C + (P Y + F )(s) ds, t ∈ [u, c], Y ∈ B.
u
Z t
|(T Y )(t) − (T Z)(t)| ≤ |P (s)||Y (s) − Z(s)|ds
u
and hence
Rt
Z t Rt
e−K u
|P (s)|ds
|(T Y )(t) − (T Z)(t)| ≤ kY − Zk |P (s)|e−K s
|P (r)|dr
ds
u
1
≤ kY − Zk.
K
5
Therefore
1
kT Y − T Zk ≤ kY − Zk.
K
¿From the contraction mapping principle in Banach space it follows that the
map T has a unique fixed point and therefore the IVP (2.3), (2.4) has a unique
solution on [u, c]. The proof for the case c < u is similar; in this case the norm of
B is modified to
Rt
|P (s)|
kY k = sup{eK u ds |Y (t)|, t ∈ [c, u]}.
Since there is a unique solution on every compact subinterval [u, c] and [c, u] for c ∈
J, c 6= u it follows that there is a unique solution on J. To establish the furthermore
part take the Banach space of real-valued functions and proceed similarly. This
completes the first proof.
For the second proof we constuct a solution of (2.5) by successive approximations.
Define
Z t
(2.8) Y0 (t) = C, Yn+1 (t) = C + (P Yn + F ), t ∈ J, n = 0, 1, 2, . . .
u
Z t
(2.9) p(t) = |P (s)| ds, t ∈ J; Bn (t) = maxu≤s≤t |Yn+1 (s) − Yn (s)|, u ≤ t ≤ b.
u
Then
Z t
(2.10) Yn+1 (t) − Yn (t) = P (s)[Yn (s) − Yn−1 (s)] ds, t ∈ J, n ∈ N.
u
Z t
(2.11) |Y2 (t) − Y1 (t)| ≤ B0 (t) |P (s)| ds = B0 (t) p(t) ≤ B0 (b) p(b), u ≤ t ≤ b.
u
6
Z t Z t
|Y3 (t) − Y2 (t)| ≤ |P (s)| |Y2 (s) − Y1 (s)| ds ≤ |P (s)| B0 (s) p(s) ds
u u
Z t
p2 (t)
≤ B0 (t) |P (s)| p(s) ds ≤ B0 (b)
u 2!
2
p (b)
≤ B0 (b) , u ≤ t ≤ b.
2!
pn (b)
|Yn+1 (t) − Yn (t)| ≤ B0 (b) , u ≤ t ≤ b.
n!
|Yn+k+1 (t) − Yn (t)| ≤ |Yn+k+1 (t) − Yn+k (t)| + |Yn+k (t) − Yn+k−1 (t)| +
. . . + |Yn+1 (t) − Yn (t)|
pn (b) p(b) p2 (b)
≤ B0 (b) [1 + + + ...]
n! n + 1 (n + 2)(n + 1)
2
p(b) p (b)
Choose m large enough so that n+1 ≤ 12 then (n+2)(n+1) ≤ 41 , etc. when n > m
and the term in brackets is bounded above by 2. It follows that the sequence
{Yn : n ∈ N0 } converges uniformly, say to Y, on [u, b]. From this it follows that Y
satisfies the integral equation (2.5) and hence also the IVP (2.3), (2.4) on [u, b].
To show that Y is the unique solution assume Z is another one; then Z is
continuous and therefore |Y − Z| is bounded, say by M > 0 on [u, b]. Then
Z t Z t
|Y (t) − Z(t)| = P (s)[Y (s) − Z(s)] ds ≤ M P (s) ds ≤ M p(t), u ≤ t ≤ b.
u u
pn (t) pn (b)
|Y (t) − Z(t)| ≤ M ≤M , u ≤ t ≤ b, n ∈ N.
n! n!
Therefore Y = Z on [u, b]. There is a similar proof for the case when b < u. This
completes the second proof. 2
(2.12) Y = Y (·, u, C, P, F )
THEOREM 2.3. Let J = (a, b), and assume that P ∈ Mn (Lloc (J)). If Y is an
n × m matrix solution of
(2.13) Y 0 = P Y on J,
then we have
Z t
(2.15) (det Y )(t) = (det Y )(u)exp traceP (s)ds , t ∈ J.
u
PROOF: The formula (2.15) follows from the fact that y = detY satisfies the first
order scalar equation y 0 + py = 0 where p = traceP. To prove the general case let
Y (u) = C and let rank C = r. If r = 0, then Y (t) = 0 for all t by Theorem 2.2.
For r > 0 let Ci , i = 1, . . . r be linearly independent columns of C and construct
a nonsingular n × n matrix D by adding n − r appropriate constant vectors to
Ci , i = 1, . . . , r. Denote by Z the solution of (2.13) satisfying the initial condition
Z(u) = D. Then by (2.15) rank Z(t) = n, for t ∈ J. Hence the first r columns
of Z(t), Z1 (t), ..., Zr (t) are linearly independent. From this and the uniqueness
part of Theorem 2.2 the (constant) n-vectors Y1 (t), Y2 (t), . . . , Yr (t) are linearly
independent since Zj = Yj on J. Hence rank Y (t) ≥ r, for t ∈ J. Now suppose
that rank Y (c) > r for some c in J. Then by repeating the above argument with
u replaced by c we reach the conclusion that rank Y (t) > r for all t ∈ J. But this
contradicts rank Y (u) = r and concludes the proof. 2
Y 0 = P Y + F, Y (u) = Ci , i = 1, . . . , n,
8
has a (vector) solution Yi on J, then P ∈ Mn (Lloc (J)) and F ∈ Mn,1 (Lloc (J)).
Furthermore, if each Yi is a C 1 solution, then there exist such P and F which are
continuous.
PROOF: We first prove the special case when F = 0 on J. Let Yi be a vector
solution satisfying Yi (u) = Ci and let Y be the matrix whose i − th column is
Yi , i = 1, . . . , n. By Theorem 2.3 the solution Y is nonsingular at each point of J.
Choose
P = Y 0 Y −1 .
Let P ∈ Mn (Lloc (J)). From Theorem 2.2 we know that for each point u of J there
is exactly one matrix solution X of (2.13) satisfying X(u) = In where In denotes
the n × n identity matrix.
DEFINITION 2.5 (The Fundamental Matrix Φ). For each fixed u ∈ J let Φ(·, u)
be the fundamental matrix of (2.13) satisfying
Φ(u, u) = In .
Note that for each fixed u in J, Φ(·, u) belongs to Mn (ACloc (J)). Furthermore, if J
is compact and P ∈ Mn (L(J)), then u can be an endpoint of J and Φ(·, u) belongs
to Mn (AC(J)). By Theorem 2.3, Φ(t, u) is invertible for each t, u ∈ J and we note
that
We also write
9
Z t
(2.18) Y (t) = Φ(t, u, P ) C + Φ(t, s, P ) F (s) ds, t∈J
u
PROOF: Clearly Y (u) = C. Differentiate (2.18) and substitute into the equation
(2.3). 2
Since we need a Gronwall inequality which is more general than the one usually
found in the literature we state and proof it here.
Z t
(2.19) y(t) ≤ f (t) + g(s) y(s)ds , a ≤ t ≤ b,
a
then
Z t Z t
(2.20) y(t) ≤ f (t) + f (s) g(s) exp( g(u) du ) ds , a ≤ t ≤ b.
a s
Z t
(2.21) y(t) ≤ c exp g(s)ds , t ∈ J.
a
(ii) (The “left” Gronwall inequality) Let J = [a, b]. Assume g in L(J) with
g ≥ 0 a.e. f real valued and continuous on J. If y is continuous, real valued, and
satisfies
Z b
(2.23) y(t) ≤ f (t) + g(s) y(s)ds , a ≤ t ≤ b,
t
then
!
Z b Z s
(2.24) y(t) ≤ f (t) + f (s) g(s) exp( g(u) du ) ds , a ≤ t ≤ b.
t t
!
Z b
(2.25) y(t) ≤ c exp g(s)ds , a ≤ t ≤ b.
t
!
Z b
(2.26) y(t) ≤ f (t) exp g(s)ds , a ≤ t ≤ b.
t
Rt
PROOF: For part (i) let z(t) = a
g y, t ∈ J and note that
z 0 = gy ≤ gf + gz a.e.
Rt
Hence with G = exp( a g) we have
(z exp(−G))0 ≤ g f exp(−G)
and
11
Z t
y(t) ≤ f (t) + z(t) ≤ c exp(G) + exp(G) (g f G)
a
Rb
from which the conclusion follows. For part (ii) let z(t) = t
g y and note that
z 0 + gy ≥ −g f,
(2.27) Y 0 = P Y + F on J, Y (u) = C.
Then
! !
Z b Z b
(2.28) |Y (t)| ≤ |C| + |F | exp |P | , a < t < b.
a a
Z t
(2.29) Y (t) = C + (P (s) Y (s) + F (s) ) ds, a < t < b.
u
Z t Z t
|Y (t)| ≤ |C| + | (P Y + F ) | ≤ |C| + (|P | |Y | + |F | )
u u
Z b ! Z
t
≤ |C| + |F | + (|P ||Y |) , u ≤ t < b.
u u
! !
Z b Rt
Z b Rb
|Y (t)| ≤ |C| + |F | e( u
|P |)
≤ |C| + |F | e( u
|P |)
, u ≤ t < b.
u u
12
Z t Z u
|Y (t)| ≤ |C| + (P Y + F ) ≤ |C| + (|P | |Y | + |F |)
u t
Z u Z u
≤ |C| + |F | + (|P ||Y |) , a ≤ t < u.
a t
Z u R Z u R
u u
|Y (t)| ≤ |C| + |F | e( t
|P |)
≤ |C| + |F | e( a |P |) , a < t ≤ u.
a a
Below we will show that, under the conditions of Theorem 2.8, a < t < b can
be replaced with a ≤ t ≤ b in (2.28). For this Y (a) and Y (b) are defined as limits.
This holds for both finite and infinite endpoints a, b.
for some c ∈ (a, b). For some u ∈ J and C ∈ Mn,m (C), let Y be the solution of the
IVP (2.3), (2.4) on J. Then
for some c ∈ (a, b). For some u ∈ J and C ∈ Mn,m (C), let Y be the solution of the
IVP (2.3), (2.4) on J. Then
Z bj Z bj
|Y (bj ) − Y (bi )| = PY ≤ B |P |.
bi bi
¿From this and the absolute continuity of the Lebesgue integral it follows that
{Y (bi ) : i ∈ N } is a Cauchy sequence and hence converges to a finite limit. 2
for some c ∈ (a, b). Let, for some u ∈ J and C ∈ Mn,m (C), Y be the solution of
the IVP (2.3), (2.4) with F = 0 on J. Then
where Y (a) is given by (2.32). Moreover, given any C ∈ Mn,m (C) there exists a
unique solution Y of the “endpoint” value problem:
(2.38) Y 0 = P Y, Y (a) = C.
for some c ∈ (a, b). Let, for some u ∈ J and C ∈ Mn,m (C), Y be the solution of
the IVP (2.3), (2.4) with F = 0 on J. Then
where Y (b) is given by (2.34). Moreover, given any C ∈ Mn,m (C) there exists a
unique solution Y of the “endpoint” value problem:
14
(2.41) Y 0 = P Y, Y (b) = C.
Note that the endpoints a and b in Theorem 2.10 may be finite or infinite.
PROOF: Note that (2.37) or (2.40) do not follow directly from (2.14) and (2.32)
or (2.34) since the rank of a matrix is not a continuous function of the matrix. We
argue as follows: Let Y (u) = C, rank C = r. If r = 0, then Y (t) = 0 for all t ∈ J
and Y (b) = 0 by (2.34). If r > 0, let C1 , . . . , Cr be linearly independent columns
of Y (u) and construct a nonsingular n × n matrix D by adding n − r appropriate
columns to Cj , j = 1, ..., r. Let Z denote the solution of (2.13) determined by the
initial condition Z(u) = D. It follows from (2.15) that Z(t) is nonsingular for each
t ∈ J and hence Z(b) is nonsingular by (2.34) and (2.15). (Note that it does
not follow directly from (2.34) alone that Y (b) is nonsingular since the rank of a
matrix is not a continuous function of its coefficients.) Therefore Z1 (b), ..., Zr (b)
are linearly independent. From the uniqueness part of the existence-uniqueness
theorem - Theorem 2.2 - Y (t) = Z(t) for t ∈ J and hence also for t = b by (2.34).
The proof for the endpoint a is similar. This establishes (2.37) and (2.40). To
prove the moreover parts of the Theorem consider the fundamental matrix Φ , see
Definition 2.5, choose u ∈ J and determine the solution Y of (2.13) by the initial
condition Y (u) = Φ(b, u) C then Y (b) = C. Note that Φ(b, u) exists by (2.34). The
proof of (2.38) is similar. 2
Then
Rb
(2.43) |Y (t) − Z(t)| ≤ K e( a
|Q|)
, a ≤ t ≤ b,
where
Z v Z v Z b Z b
(2.44) K = |C − D| + |F | + M |P | + |F − G| + M |P − Q|,
u u a a
and
! !
Z b Z b
(2.45) M= |C| + |F | exp |P | .
a a
15
PROOF: For a < t < b this follows from Theorem 2.8 and the Gronwall inequality.
The case t = a and t = b then follows from Theorem 2.9. 2
Z b Z b
(2.46) |u − v| + |C − D| + |P − Q| + |F − G| < δ,
a a
then
¯
Note that Y (t, u, C, P, F ) is jointly continuous in u, C, P, F, uniformly for t in J.
PROOF: The absolute continuity of the Lebesgue integral and (2.46) imply that
the constant K in (2.44) can be made arbitrarily small. The conclusion then follows
from Theorem 2.11. 2
THEOREM 2.13. Let J = (a, b), −∞ ≤ a < b ≤ ∞, let Pk ∈ Mn,n (Lloc (J)),
Fk ∈ Mn,m (Lloc (J)), Ck ∈ Mn,m , uk ∈ J, k ∈ N0 = {0, 1, 2, . . . }. Assume
(i) Pk → P0 as k → ∞
locally in Lloc (J) in the sense that for each compact subinterval K of J we
have
Z
|Pk − P0 | → 0 as k → ∞;
K
(ii) Fk → F0 as k → ∞
locally in Lloc (J) in the sense that for each compact subinterval K of J we
have
Z
|Fk − F0 | → 0 as k → ∞;
K
(iii) Ck → C0 ∈ C as k → ∞;
(iv) uk → u0 ∈ J as k → ∞.
16
Then
Y (t, uk , Ck , Pk , Fk ) → Y (t, u0 , C0 , P0 , F0 ) as k → ∞
Y (t, uk , Ck , Pk , Fk ) → Y (t, u0 , C0 , P0 , F0 ) as k → ∞
Theorem 2.12 shows that the solution of the initial value problem (2.3), (2.4) with
P ∈ Mn,n (Lloc (J)) , F ∈ Mn,m (Lloc (J)) , C ∈ Mn,m (C) depends continuously on
all the given data. In this section we show that this dependence is differentiable.
If such a map T 0 (x) exists, it is unique and is called the (Frechet) derivative of T
at x. A map T is differentiable on a set S ⊂ X if it is differentiable at each point of
S. In this case the derivative is a map : x → T 0 (x) from S into the Banach space
L(X, Z) of all bounded linear operators from X into Z denoted by T 0 . To say that
T 0 is continuously differentiable on S or T is C 1 on S means that the map T 0 is
continuous in the operator topology of the Banach space L(X, Z).
The differentiability of the solution
Y = Y (t, u, C, P, F )
17
Z t
Y (t, u) = Φ(t, u) C + Φ(t, s) F (s) ds.
u
∂Y
(2.48) Y 0 (C) = (t, u, C, P, F ) = Φ(t, u, P ).
∂C
Thus we have
(2.50)
Z t
∂Y
Y 0 (F )(H) = (t, u, C, P, F )(H) = Φ(t, s)H(s) ds, H ∈ Mn,m (L(J)).
∂F u
Here the right side of equation (2.50) defines a bounded linear operator on the space
Mn,m (L(J)). The derivative Y 0 (F ) is constant in F .
18
Z t
Y (t, u, C, P, F + H) − Y (t, u, C, P, F ) = Φ(t, s, P )H(s) ds.
u
The conclusion follows from this equation and the definition of derivative. 2
Before stating the next Theorem we give two lemmas. These may be of inde-
pendent interest.
LEMMA 2.18. Let J = (a, b), −∞ ≤ a < b ≤ ∞, let P ∈ Mn (Lloc (J)), let
u ∈ J. Then for any t ∈ J we have
Z t Z t Z r
Φ(t, u, P ) = I+ P+ P (r) P (s) drds
u u u
Z t Z r Z s
(2.51) + P (r) P (s) P (x) dxdsdr + · · · .
u u u
PROOF: This follows directly from the successive approximations proof of the
existence-uniqueness
Rt Theorem : Start with the first approximation Φ0 = I; then
Φ1 = I + u P, etc. 2
where
PROOF: The proof consists in showing that both sides satisfy the same initial value
problem and then using the existence-uniqueness Theorem. 2
Z s Z s
(2.54) P (t) ( H) = ( H) P (t), s, t, u ∈ J,
u u
PROOF: It follows from Lemmas 2.18, 2.19 and hypothesis (2.54) that Φ(·, u, P ) H =
H Φ(·, u, P ) and hence S = H in (2.53). 2
THEOREM 2.21. Let J = [a, b]. Fix t, u ∈ J, C ∈ Mn,m (C), F ∈ Mn,m (L(J)).
For P ∈ Mn (L(J)) let Y = Y (t, u, C, P, F ) be the unique solution of (2.3), (2.4).
Then the map P → Y (t, u, C, P, F ) from the Banach space Mn (L(J)) to Mn,m (C)
is differentiable and its derivative
20
∂Y
(2.56) Y 0 (P ) = (t, u, C, P, F )
∂P
is the bounded linear transformation from the Banach space Mn (L(J)) to the Ba-
nach space Mn,m (C) given by
Z t
0 −1
Y (P ) H = Φ(t, u, P ) Φ (r, u, P )H(r)Φ(r, u, P ) dr C
u
(2.57)
Z t Z t
−1
+ Φ(t, r, P ) Φ (s, u, P )H(s)Φ(s, u, P ) ds F (r) dr, H ∈ Mn (L(J)).
u r
Y (t, P + H) − Y (t, P )
Z t
= Φ(t, P + H) C + Φ(t, s, P + H) F (s) ds − Φ(t, P ) C
u
Z t
− Φ(t, s, P ) F (s) ds
u
Z t
= Φ(t, u, P ) [Φ(t, u, S) − I] C + Φ(t, s, P )[Φ(t, r, S) − I]F (r)dr
u
Z t Z t Z x Z t
= Φ(t, u, P ) S+ S(x) S(y)dydx + . . . C + Φ(t, r, P )
u u u u
Z t Z t Z x
S+ S(x) S(y)dydx + . . . F (r)dr
r r r
Hence
Z t
Y (t, u, P + H) − Y (t, u, P ) − Φ(t, u, P ) S(r)dr C
u
Z t Z t
− Φ(t, r, P ) S(x)dx F (r)dr
u r
Z t Z x
= Φ(t, u, P ) S(x) S(y)dydx + . . . C
u u
Z t Z t Z x
+ Φ(t, r, P ) S(x) S(y)dydx + . . . F (r)dr
u r r
= E(H).
21
Noting that |S|(b−a)| ≤ |k H| for some k ∈ R, that |Φ(t, u, P )| and |Φ−1 (t, u, P )|
are bounded on J, there exists an M > 0 such that
|E(H)| ≤ M |C| [|S(b − a)|2 + |S(b − a)|3 . . . ] + M |F | [|S(b − a)|2 + |S(b − a)|3 . . . ]
≤ M |C| |kH| [|kH| + |kH|2 + . . . ] + M |F | |kH| [|kH| + |kH|2 + . . . ]
|E(H)|
→ 0 as |H| → 0 in Mn (L(J)).
|H|
THEOREM 2.22. Let the hypotheses and notations of Theorem 2.21 hold and
assume, in addition, that the commutativity hypothesis (2.54) is satisfied. Then
1.
(2.58) H(t) Φ(t, u, P ) = Φ(t, u, P ) H(t), t, u ∈ J.
Z t
Y 0 (P )(H) = Φ(t, u, P ) H(s) ds C +
u
Z t Z t
(2.60) Φ(t, r, P ) H(s) ds F (r) dr, t, u ∈ J.
u r
Note however that Y 0 (P ) is not the operator defined by the right hand side of
(2.60) since H cannot be restricted to satisfy the commutativity hypothesis (2.54)
in the definition of the derivative Y 0 (P ).
PROOF: This follows from Theorem 2.21 and Lemma 2.20. 2
REMARK 1. In the special case when P and H are constant matrices we have
Z t
Y 0 (P )(H) = exp((t − u)P ) exp((u − r)P )H exp((r − u)P ) dr C
u
22
Z t Z t
(2.61) + exp((t − r)P ) exp((u − s)P )H exp((s − u)P )ds F (r) dr.
u r
Note that if P and H are constant and commute, then (2.61) reduces to
Z t
0
Y (P )(H) = (t − u) exp ((t − u)P ) HC + (t − r) exp((t − r)P )HF (r) dr.
u
But this reduction does not hold, in general, for constant matrices which do not
commute.
The Frechet derivative of E is the bounded linear operator from Mn (C) into
Mn (C) given by
Z 1
0
(2.62) E (A) H = e A
e− r A H er A dr, H ∈ Mn (C).
0
Note that (2.62) reduces to the more familiar formula E 0 (A) = E(A) for all
A ∈ Mn (C) only in the one dimentional case n = 1. When n > 1 (2.62) reduces to
E 0 (A) = E(A) only for constant multiples A = cIn , c ∈ C , of the identity since
only multiples of the identity satisfy the commutativity condition with respect to
all matrices in Mn (C).
∂Y
Y 0 (z) = ?
∂z
Z t
Y 0 (z) = Φ(t, u, P + zW ) Φ−1 (r, u, P + zW ) W (r) Φ(r, u, P + zW ) dr C
u
Z t Z t
+ Φ(t, r, P + zW ) Φ−1 (s, u, P + zW ) W (s) Φ(s, u, P + zW ) ds F (r)dr
u r
PROOF:
[Y (t, u, C, P + (z + h)W, F ) − Y (t, u, C, P + zW, F )]
= [Φ(t, u, P + (z + h)W ) − Φ(t, u, P + zW )] C
Z t
(2.63) + [Φ(t, r, P + (z + h)W ) − Φ(t, r, P + zW )] F (r) dr.
u
Let
S(z) = Φ−1 (·, u, P + zW )W (·) Φ(·, u, P + zW ).
Z t
Y 0 (W ) H = z Φ(t, u, P + zW ) Φ−1 (r, u, P + zW )H(r) Φ(r, u, P + zW ) dr C
u
Z t Z t
−1
+z Φ(t, r, P + zW ) Φ (s, u, P + zW )H(s) Φ(s, u, P + zW ) ds F (r) dr,
u r
for H ∈ L(J).
PROOF: The proof is similar to that of Theorem 2.21 and hence omitted. 2
(2.66) (Z ∗ CY )0 = Z ∗ CF + G∗ CY.
The fundamental matrices of adjoint systems are closely related to each other.
The next result gives this relationship. It plays an important role in the theory of
adjoint and, in particular, self-adjoint boundary value problems.
(2.67) E −1 E ∗ = I or E −1 E ∗ = −I
and define
(2.68) P + = −E −1 P ∗ E.
25
Then
using (2.67) and (2.68). Hence Z(t) = I, for t ∈ J. That this is equivalent to
(2.69) follows from the representation Φ(t, s, P + ) = Y (t) Y −1 (s), s, t ∈ J, for any
fundamental matrix Y of Y 0 = P + Y. 2
(2.70) Y = [Y1 , Y2 , . . . , Yd ].
(2.71) Y 0 = PY .
26
If
(2.72) rank [Y1 , Y2 , . . . , Yd ](t) = d
M = [Y1 , Y2 , . . . , Yd , Yd+1, . . . , Yn ]
2.10 Comments.
Most of Section 2 is based on the paper [59] by Kong and Zettl. Below we comment
on each subsection separately.
1. The notation for matrix functions such as Mn (L(J)) is taken from [70].
2. The sufficiency of the local integrability conditions of Theorem 2.2 are well
known - see [72] or [81]; the necessity given by Theorem 2.4 is due to Everitt
and Race - see [25]. Except for the use of the Bielecki norm the first proof
of Theorem 2.2 is the standard successive approximations argument although
it is dressed in the clothes of the Contraction Mapping Theorem in Banach
space here. The advantage of the Bielecki norm is that it yields a global proof;
the sup norm would only give a local proof and then one has to patch together
the intervals of existence. The second proof is a minor variant of the usual
successive approximations argument.
The constancy of the rank of solutions given by Theorem 2.3 is known -
see [42] or [75] but we haven’t seen it stated under these general conditions.
It is surprising how many authors, including the two just mentioned assume
continuity of the coefficients when only Lebesgue integrability is needed. This
is of some consequence both theoretically and numerically when coefficients
are approximated by piece-wise constants, piece-wise linear functions, etc.
3. The variation of parameters formula given by Theorem 2.6 is standard, but
our notation is not. We use a notation which shows the dependence of the
fundamental matrix on the coefficient matrix P . This is handy for the differ-
entiation results that follow.
4. A detailed discussion of the Gronwall inequality is given here because it is a
very useful tool and we do not want any continuity assumptions on f and g.
The Gronwall inequality has many extensions: see - [11],
5. Theorem 2.8 is elementary but we have not seen it stated in this generality.
The continuous extensions of solutions given by Theorem 2.9 are a special
case of much more powerful results e.g. Levinsons asymptotic theorem -
[14]. Often the existence of limits of solutions are stated only for infinite
endpoints. We want to emphasize here that the relevant consideration is not
whether the endpoint is finite or infinite but whether the coefficient matrix
P and the inhomogeneous term F are integrable or not all the way to the
endpoint. Theorem 2.10 may be new in [59].
6. Theorems 2.11, 2.12, and 2.13 illustrate clearly that the natural space in which
to study solutions of linear ode’s is L1loc (J) in the singular case and L1 (J) in
the regular case.
7. Sections 7 and 9 were motivated to some extend by the elegant treatment of
the inverse spectral theory for regular Sturm-Liouville problems by Poeschel
and Trubowitz in [76]. Theorem 2.14 is standard - see [81] or [3]. Theorems
2.14 and 2.15 are trivial consequences of the variation of parameters formula.
Theorems 2.21, 2.24, 2.25, are not so trivial consequences of the Variation of
Parameters Formula; and may be new in [59].
8. Adjoint systems of this type were used by Atkinson [3] . They will be used
in the next chapter to provide an elegant proof of a very general Lagrange
identity due to Everitt and Neumann, see [24]. Theorem 2.27, the Adjointness
Lemma, is due to Zettl, see [82], [83], [87].
28
9. These kinds of inverse problems are discussed by Hartman [42] and Petrovski
[75] but not in this generality.
(3.1) −(py 0 )0 + qy = f on J
where
THEOREM 3.2. Every initial value problem (IVP) (3.1), (3.2), (3.3) has a solu-
tion defined on J and this solution is unique if and only if
0 1/p 0 y
(3.5) P = , F = , Y =
q 0 f py 0
29
(3.6) Y 0 = P Y + F on J,
in the sense that, given any scalar solution y of (3.1) the vector Y defined by
(3.5) is a solution of the system (3.6) and conversely, given any vector solution Y
of system (3.6) its top component y is a solution of (3.1). Theorem 3.2 follows from
this system representation and Theorems 2.2 and 2.4. 2
Notation. Given (3.3) and (3.4), the unique solution y of (3.1), (3.2) and its
quasi-derivative py 0 are denoted by
(3.8) −(py 0 )0 + qy = λ w y on J, λ ∈ C,
where
Theorem 3.2 and its proof readily extend to the case when q is replaced by q − λw.
Below, when we study the dependence of y and py 0 on one of these quantities
with all the others fixed we further abbreviate this notation by simply omitting all
the fixed variables. Thus we write y = y(·, c) when we wish to study the unique
solution as a function of c, y = y(·, q) to study the dependence of y on q, etc.
DEFINITION 3.3 (Regular and singular endpoints). Let J = (a, b), −∞ ≤ a <
b ≤ ∞, and consider the equation (3.8) with conditions (3.9).
The end-point a is said to be regular if
30
both exist and are finite for the solution y of every initial value problem (3.1), (3.2),
(3.3), (3.4) if and only if
both exist and are finite for the solution y of every initial value problem (3.1), (3.2),
(3.3), (3.4) if and only if
has a unique solution on J; this solution is real if all the data is real.
• Assume (3.16) and define y(b) and (py 0 )(b) by (3.15). Then each “initial value
problem ” consisting of equation (3.1) and the terminal value conditions
has a unique solution on J; this solution is real if all the data is real.
PROOF: This follows from the moreover part of Theorem 2.10. 2
THEOREM 3.6. Let (3.1), (3.2), (3.3) and (3.4) hold. Using the notation (3.7)
each solution y of (3.1), (3.2) and its quasi-derivative py 0 is a jointly continuous
function of all its variables, uniformly on compact subintervals of J = (a, b). More
precisely, given cj ∈ J, hj , kj ∈ C, 1/pj , qj , fj ∈ Lloc (J), j = 1, 2, and given > 0
and a compact subinterval K = [a1 , b1 ] of J containing c1 and c2 , there exists a
δ > 0 such that if
(3.19)
Z
|c1 − c2 | + |h1 − h2 | + |k1 − k2 | + (|1/p1 − 1/p2 | + |q1 − q2 | + |f1 − f2 |) < δ,
K
then
and
y(t, c, h, k, 1/p, q, w, λ)
of (3.8), (3.9) and its quasi-derivative (py 0 ) with respect to t follows from the
definition of solution; the differentiability of y and (py 0 ) with respect to c is a
consequence Lemma 2.16. The differentiability of y and of (py 0 ) with respect to the
other variables is studied in this subsection.
THEOREM 3.7. Let (3.3), (3.8), (3.9) hold. Let u, v be solutions of (3.8) deter-
mined by the initial conditions
Using the notation (3.10) with the associated convention mentioned in the para-
graph below (3.10), we have that each of the following maps from C to C :
y 0 (h) = v(t) h, h ∈ C
0
(py 0 ) (h) = (pv 0 )(t) h, h ∈ C
y 0 (k) = u(t) k, k ∈ C
0
(py 0 ) (k) = (pu0 )(t) k, k ∈ C,
respectively. Note that here y 0 (h) denotes the derivative of y with respect to h, and
in (py 0 )0 (h) the outside prime denotes the derivative of the quasi-derivative (py 0 )
with respect to h. Thus the two primes in (py 0 )0 have different meanings in this
formula - the outside one is for differentiation with respect to h , the inside one
for diffentiation with respect to t - but since t is fixed here this should not cause
33
confusion. Similar remarks apply to the formulas for differentiation with respect to
k.
Let K = [a1 , b1 ] be a compact subinterval of J. Each of the following maps from
C to the Banach space C(K):
y 0 (k)(g) = u g, g ∈ C(K)
respectively.
PROOF: This is a straightforward consequence of the variation of parameters for-
mula and the definition of the Frechet derivative. See the above remarks about
notation. 2
0 1/p 0 0
P = , W =
q 0 w 0
THEOREM 3.8. Let (3.3), (3.8), (3.9) hold, and let K be a compact subinterval
of J. Fix t, c ∈ K, h, k, λ ∈ C, 1/p, w ∈ L(K); then the maps q → y(t, q), w →
y(t, w), 1/p → y(t, 1/p) from L(K) to C as well as the map λ → y(t, λ) from C → C
are differentiable and their derivatives are given by
Z t
y 0 (t, q)(r) = − φ1,2 (t, s) y(t, s) r(s) ds, r ∈ L(K),
c
Z t
0
(py )(t, q)(r) = − φ2,2 (t, s) y(t, s) r(s) ds, r ∈ L(K),
c
34
Z t Z s
0 0
y (t, 1/p)(r) = φ1,2 (t, s) q(s) (py )(x) r(x) dx ds
c c
Z t
+ (py 0 )(x) r(x) dx, r ∈ L(K),
c
Z t Z s
0 0
(py )(t, 1/p)(r) = φ2,2 (t, s) q(s) (py )(x) r(x) dx ds, r ∈ L(K),
c c
Z t
y 0 (t, w)(r) = λ φ1,2 (t, s, w) y(s, w) r(s) ds, r ∈ L(K);
c
Z t
(py 0 )(t, w)(r) = λ φ2,2 (t, s, w) y(s, w) r(s) ds, r ∈ L(K);
c
Z t
y 0 (t, λ) = φ1,2 (t, s, λ) w(s) y(s, λ) ds, λ ∈ C.
c
PROOF: We prove some of these, the proofs of the others are similar. Let
Let x = z − y. Then
Z t
x(t) = φ1,2 (t, s)(−r(s)) z(s)ds.
c
Letting z = y + (z − y) we get
Z t Z t
z(t) − y(t) + φ1,2 (t, s)(r(s) y(s)ds = − φ1,2 (t, s)[ z(s) − y(s)] r(s) ds
c c
= o(r) as r → 0 in L(J).
The last equality follows from the fact that φ1,2 is bounded on K × K and z → y
uniformly on K by the furthermore part of Theorem 3.6. Similarly we get
Z t
0 0 0
(pz − py )(t) = (px )(t) = φ2,2 (t, s)(−r(s) z(s)ds
c
35
and from this, proceeding as above, we obtain the formulas for y(t, q)(r) and for
(py 0 )(t, q)(r).
To derive the formulas for the derivatives with respect to 1/p we proceed as
follows. Let
1 1
= + r, r ∈ L(J)
pr p
Z t
1
x(t) = (py 0 − pr z 0 ).
c p
Then
Z t
−(px0 )0 + qx = f, x(c) = 0, (px0 )(c) = 0, f (t) = −q(t) (pr z 0 ) r.
c
Z t Z s
0
x(t) = − φ1,2 (t, s) q(s) (pr z )(u) r(u)du ds,
c c
Z t Z s
(px0 )(t) = − φ2,2 (t, s) q(s) (pr z 0 )(u) r(u)du ds.
c c
Z t
1 1
z(t) − y(t) = (pr z 0 ) − (py 0 )]
[
c pr p
Z t Z t
1 1
= [ (pr z 0 ) − (py 0 )] + (pr z 0 ) r
c p p c
Z t
= −x(t) + (pr z 0 ) r
c
Z t Z s Z t
= φ1,2 (t, s) q(s) (pr z 0 )(u) r(u)du ds + (pr z 0 ) r.
c c c
Setting
pr z 0 = py 0 + [pr z 0 − py 0 ]
we obtain
36
Z t Z s Z t
z(t) − y(t) − φ1,2 (t, s) q(s) (py 0 )(u) r(u)du ds + (py 0 ) r
c c c
Z t Z s Z t
= φ1,2 (t, s) q(s) [(pr z 0 ) − py 0 ](u) r(u)du ds + [(pr z 0 ) − py 0 ] r
c c c
= o(r) as r → 0 in L(J).
The last equality follows from the boundedness of φ1,2 on K ×K, from q ∈ L(K),
and from the fact that, by Theorem 3.6, (pr z 0 ) → py 0 uniformly on K as r → 0 in
L(J). 2
There is an interesting and subtle point involved in the proof of Theorem 3.9:
1
p + r may be identically zero on a subinterval of J. Note that the solutions y
depend on p1 , not on p. Therefore p1 may be identically zero on a subinterval of J or
even on all of J. This is allowed by the existence-uniqueness Theorem 2.2 and the
subsequent theorems. However, the equation (3.8) has to be interpreted properly
in this case as Atkinson [3] has pointed out. In fact Atkinson uses the notation
1
−( y 0 )0 + qy = λ w y
p
for equation (3.8) but this notation has not been widely accepted.
For regular equations each solution y and its quasi-derivative py 0 are not only
entire functions of λ but have order at most 1/2.
Then every nontrivial solution y of (3.8) and its quasi-derivative py 0 are entire
functions of λ of order at most 1/2. More precisely, there exist positive constants
M, B, δ such that
√
|λ|
|y(t, λ)| ≤ B eM , a ≤ t ≤ b, |λ| ≥ δ
√
|(py 0 )(t, λ)| ≤ B eM |λ|
, a ≤ t ≤ b, |λ| ≥ δ
PROOF: Let v = py 0 then v 0 = (q − λw) y. Fix λ and let prime “0 ” denote differen-
tiation with respect to t. Then
we get
and hence
p 1 1 p
[log (|λ| |y|2 + |v|2 )]0 ≤ |λ| + p |q| + |λ| |w|.
|p| |λ|
An integration yields
√ Rt 1 Rt
|λ| a ( |p| +|w|)+ √1 |q|
|λ| |y(t, λ)|2 + |v(t, λ)|2 ≤ C e |λ\ a
√
Z b √1
Rb
M |λ| 1 a
|q|
≤Be , 0<M = ( + |w|) < ∞, e |λ|
< B < ∞.
a |p|
(3.23) −(py 0 )0 + qy = λ w y, λ ∈ C, on J,
with
1/p, q, w ∈ L(a, d)
38
holds for some (and hence any) d ∈ J; is limit-circle if all solutions of the equation
Rd
(3.23) are in L2w (a, d) = {f : (a, d) → C, a |f |2 w < ∞} for some (and hence any)
d ∈ (a, b); is LP if it is not LC; is O if there is a nontrivial solution with an infinite
number of zeros in any right neighborhood of a ; is NO if it is not O; is LCO if it
is both LC and O; and is LCNO if it is both LC and NO. Similar definitions are
made at b. An endpoint is called singular if it is not regular.
It is well known [81] that the LC, LP, classifications are independent of λ ∈ C
and that the LCO and LCNO classifications are independent of λ ∈ R. At an LP
endpoint the O classification, in general, depends on λ.
p, q, w : J → R, p > 0, a.e., λ ∈ R.
Then the zeros of every nontrivial solution y of (3.23) are isolated in the interior of
J and also at regular endpoints of J i.e. if a nontrivial solution y has a zero at a
regular endpoint of J then there is an appropriate one sided neighborhood of this
endpoint in which y has no other zero. Thus only the singular endpoints of J can
be accumulation points of zeros of y.
PROOF: Let y be a nontrivial solution of (3.23). First we show that if y has
consecutive zeros at c, d ∈ (a, b), c < d, then (py 0 )(h) = 0 for some h ∈ (c, d). We
have
Z d Z d Z d
0 1 1
0 = y(d) − y(c) = y = (py 0 ) = (py 0 )(h)
c c p c p
by the Mean Value Theorem for the Lebesgue integral. (Recall that (py 0 ) is con-
Rd
tinuous on J.) Hence either (py 0 )(h) = 0 or c p1 = 0, but the latter would imply
that p1 = 0 a.e. in (c, d) in contradiction to the hypothesis that p > 0 a.e. in (a, b).
Now to prove the main Proposition suppose there exists a sequence {tn ∈ (a, b) :
n ∈ N0 } such that tn → t0 and y(tn ) = 0, n ∈ N0 . Then y(t0 ) = 0 and from the
first part of the proof we get a sequence {sn : n ∈ N } with sn → t0 such that
(py 0 )(sn ) = 0. Since (py 0 ) is continuous in (a, b) it follows that (py 0 )(t0 ) = 0. But
y(t0 ) = 0 and (py 0 )(t0 ) = 0 implies that y is identically zero on J by the uniqueness
of initial value problems. 2
We end this subsection with an example to show that when p changes sign the
behavior of the classical and quasi-derivatives of a solution can be quite different.
1
p(t) = , 0 < t ≤ 1.
cos(log(t))
39
−(py 0 )0 = 0 on (0, 1)
has
Z
y= cos(log(t)) dt, v(t) = 1
tk = e−kπ/2 → 0, as k → ∞,
y v
W (y, v)(t) = (t) = −1, 0 ≤ t ≤ 1,
py 0 pv 0
y v
(t) = −cos(log(t)), 0 < t ≤ 1
y0 v0
M y = [−(py 0 )0 + qy]
THEOREM 3.12. Let (3.3), (3.8), and (3.9) hold; and suppose that p, q, w are
real-valued and p > 0, w > 0 a.e. on J. Assume each endpoint is either regular
or LCNO. Let ∆ and M be defined as in subsection 6. Then there exist functions
u, v ∈ ∆ satisfying the following conditions:
1. They are real valued.
2. For some real λ = λa , u is a principal solution at a and v is a nonprincipal
solution at a.
3. For some real λ = λb , u is a principal solution at b and v is a nonprincipal
solution at b.
4. These functions u, v need not be solutions through the interior of (a, b), and,
in case λa = λb , they need not be the same solution near a and near b.
5. [u, v](a) = limt→a+ [u, v](t) = 1,
6. [u, v](b) = limt→b− [u, v](t) = 1,
7. v > 0 on J = (a, b).
PROOF: See Subsection 5.2 of Section 5 below for the definition of principal and
non-principal solution; and see Lemma 7 in Niessen and Zettl [74] for a proof. 2
THEOREM 3.13. Let (3.3), (3.8), (3.9) hold. Assume p, q, w are real valued,
w > 0 a.e. and the left endpoint a is R or LC. Suppose u, v are real valued linearly
independent solutions on some interval (a, d] for some fixed real λ0 . Given any λ ∈ R
and any h, k ∈ R the singular initial value problem consisting of the equation
−(py 0 )0 + qy = λ w y on J
PROOF: Since u, v are linearly independent solutions near a we have [u, v](a) 6= 0
and we can assume that [u, v](a) = 1. Let
u v
U= , Z = U −1 Y, Y 0 = (P − λW )Y, U 0 = (P − λ0 W )U,
pu0 pv 0
0 0
and, using the notation from (3.5) for P, Y, let with W = . Note that
w 0
U is a fundamental matrix solution for a fixed λ0 but Y is a vector solution for an
arbitrary λ ∈ R . A direct computation reveals that
where
−v 2 w
−1 −uvw
G=U WU = ∈ L(a, d).
u2 w uvw
Note that G ∈ L(a, d) follows from the Cauchy-Schwarz inequality coupled with
the assumption that a is in the LC case. Hence by Theorem 2.10 all initial value
problems
z1
have a unique solution. From Y = U Z, Z = we get
z2
y(t) v(t)
z1 (t) = = [y, v](t), a < t < d
(py 0 )(t) (pv 0 )(t)
u(t) y(t)
z2 (t) = = −[y, u](t), a < t < d.
(pu0 )(t) (py 0 )(t)
By letting t → a we get z1 (a) = k, z2 (a) = −h. Since this holds for arbitrary
h, k the proof is complete. 2
REMARK 3. Note that the assumption p > 0 is not needed in Theorem 3.13,
nor have we assumed that a is LCNO but only that a is LC. The transformation
Z = U −1 Y transforms the singular scalar equation (3.8) into a first order regular
system.
42
A singular equation with an LCNO endpoint can be “regularized” using the function
v from a regularizing pair u, v of functions as defined in subsection 8.
THEOREM 3.14. Consider the equation (3.8) with J given by (3.3) and with
Assume that the left endpoint a is LCNO and let u, v be a pair of regularizing
functions at a i.e. on (a, d) for some d ∈ (a, b) as defined in section 7. Define
P = v 2 p, W = v 2 w, Q = w v M v on J,
(3.25) −(P z 0 )0 + Qz = λ W z on J.
Then we have
exist and are finite. Thus the solution z and its quasi-derivative (P z 0 ) can be
continuously extended to the (finite or infinite) endpoint a.
3. Note that v is independent of λ ∈ R but does depend on (M, w) i.e. on
1/p, q, w and on the endpoint a i.e. on some neighborhood (a, d) for d ∈ (a, b).
4. The one-to-one mappings y(t, λ) = v(t) z(t, λ) and (py 0 )(t, λ) = v(t) (P z 0 )(t, λ)
can be given more explicitly, using the notation (3.10), by
for c ∈ J, and h, k ∈ C.
PROOF: See Niessen and Zettl [74]. Although the explicit formulas for the 1-1 map
y → vz are not given by these authors it can easily be obtained from there. 2
y
z(a) = (a); (P z 0 )(a) = (vpy 0 − ypv 0 )(a) = [v, y](a)
v
but neither the numerator y nor the denominator v , nor the individual terms in
(P z 0 )(a), can be evaluated separately at a. Of course there is an entirely analogous
theorem and remark for the endpoint at b. If each endpoint is either R or LCNO
then Theorem 3.14 holds on the entire interval J.
THEOREM 3.15. Let the notation of Theorem 3.14 hold. If each endpoint of
the interval J is either R or LCNO, then there exist a pair of regularizing functions
u, v defined on all of J such that the conclusions of Theorem 3.14 hold on the whole
interval J.
PROOF: See Niessen and Zettl [74]. 2
we have
Note that v > 0 near +1 and near -1. Every solution y can be factored as follows:
where z is continuous on the closed interval [−1, 1]. This shows, in particular, that
the asymptotic behavior of solutions of the Legendre equation satisfying a fixed
initial condition is independent of λ ∈ C .
44
3.10 Comments.
1. In Theorem 3.2 the sufficiency of condition (3.4) is well known, the necessity
is due to Everitt and Race [25].
2. The results of Theorems 3.4, 3.5 and 3.6 are surely not new, but we don’t
know of a reference where they can be found in this generality.
3. Again, other than [59], we don’t know of a reference where the continuous
dependence of solutions of initial value problems on 1/p, q, w in the L1 norm
is established. The continuous dependence of solutions on initial conditions
i.e. on c, h, k is discussed by Hille [43].
4. The differentiable dependence of solutions and their quasi-derivatives on each
parameter as well as the formulas for the derivatives seems to be new. The
proof of the differentiable dependence of y and of py 0 on 1/p is due to Qingkai
Kong and published here for the first time with his permission.
Is y jointly differentiable in all its variables: c, h, k, 1/p, q, w ? Ditto for
py 0 . What are the derivatives ?
Theorem 3.9 is adapted from Atkinson [3]. It follows from the asymptotic
form of the eigenvalues of regular self-adjoint SLP that for p, q, w real-valued
p ≥ 0, w > 0 that the non-trivial solutions - as functions of λ - are of order
exactly 1/2.
Under what more general conditions on p, q, w are the non-trivial solutions
of exact order 1/2 ?
Under what general conditions (excluding trivial cases such as p = q =
w = 0) are the solutions of the SL equation of order zero as functions of λ
? Are there general classes of equations for which the solutions have order r,
for 0 < r < 1/2 as functions of λ ?
5. The definitions of R, LC, LP, O, NO, LCNO, LCO are standard except, as
pointed out in Remark 2, we classify an infinite endpoint as regular if 1/p, q, w
are integrable in a neighborhood of this point. This contrasts with the usual
practice. For a definition of oscillation of difference equations see [58].
The standard proof of the invariance of the L2w (J) solutions with respect
to λ is based on the variation of parameter formula [14].
The invariance of the LCO case with respect to real λ follows from the
spectral theory of ordinary differential operators, see [81]. In general for the
symmetric case i.e. p, q, w real and 1/p, q, w ∈ Lloc (a, d), a < d < b, p ≥
0, w > 0 there exists a σ0 , −∞ ≤ σ0 ≤ ∞, such that the equation (3.8) is O
at a for λ > σ0 and is NO for λ < σ0 . Examples show that for λ = σ0 the
equation can be O or NO. This “oscillation number” σ0 is also the starting
point of the essential (continuous) spectrum of every self-adjoint realization of
the equation. The case σ0 = −∞ is interpreted as meaning that the essential
spectrum is not bounded below; the case σ0 = ∞ means that the essential
spectrum is empty. The latter holds for all SLP for which each endpoint
is either R or LC since in this case the spectrum is discrete. The essential
but not discrete spectrum is the same for all self-adjoint realizations of the
equation (3.8). For proofs of these statements as well as further information
the reader is referred to [81].
45
6. The definition of the maximal domain and of the Lagrange sesquilinear form
is standard. These play an important role in the theory of boundary value
problems.
7. This section is based on Niessen and Zettl [74].
8. The “system regularization” of LC endpoints based on the fundamental ma-
trix U is not new. It has been used by Fulton and by Fulton and Krall. It
has other applications besides the one given here to singular IVP:
(a) It can be used to prove the LC invariance with repect to λ : ¿From
(λ0 = 0 in this case) maps solutions y of the Legendre equation into solutions
z of the regular equation in a 1-1 onto manner. All solutions z of the regular
equation are continuous on the closed interval [−1, 1]. So the singular behavior
is contained in the transformation function v. Since v is independent of λ this
shows that the singular behavior of the Legendre equation is independent of
λ ∈ C . The invariance of the LC classification with respect to λ ∈ C and
the invariance of the LCNO classifications with respect to λ ∈ R are merely
specific instances of this general invariance property.
Of course, these remarks apply to all other equations where each endpoint
is either regular or LCNO.
4.1 Introduction.
where
y
(4.3) AY (a) + BY (b) = 0, Y = , A, B ∈ M2 (C).
py 0
By Theorem 3.4, Y (a), Y (b) exist and are finite so that (4.3) is well defined. Let
0 1/p 0 0
(4.4) P = , W = .
q 0 w 0
Then the scalar equation (4.1) is equivalent with the first order system
47
0 1/p y
(4.5) Y 0 = (P − λW )Y = Y, Y = .
q − λw 0 py 0
(4.6) Φ0 = (P − λW ) Φ, Φ(u) = I, u ∈ J, λ ∈ C,
Although the notation ∆ was used in section 3 to denote the maximal domain
there should be no confusion with its use here as the characteristic function.
LEMMA 4.1. Let (4.1) to (4.7) hold. Then the characteristic function ∆ is de-
fined and continuous at a and b for fixed P, w, λ and is an entire function of λ for
fixed a, b, P, w.
PROOF: It follows from Theorems 2.9 and 2.10 that, for fixed P, w, λ, ∆(a, b) exists
and is continuous at a and b. The entire dependence on λ follows from the second
proof using successive approximations of Theorem 2.2, the existence-uniqueness
theorem. Each successive approximation is a polynomial in λ. Since these converge
uniformly on each compact subset K of the complex plane to the solution, this
solution is analytic on K. Thus the solution is entire in λ since this holds for each
such K. 2
PROOF: Suppose ∆(λ) = 0. Then (4.8) has a nontrivial vector solution for C. Let
Y (a) = C and solve the IVP
Y 0 = (P − λW )Y, Y (a) = C, on J.
Then
From this it follows that the top component of Y, say, y is an eigenvector of the
BVP (4.1), (4.2), (4.3); that means λ is an eigenvalue ofthis BVP. Conversely, if
y
λ is an eigenvalue and y an eigenvector of λ, then Y = satisfies Y (b) =
py 0
Φ(b, a, λ) Y (a) and consequently [A + B Φ(b, a, λ)] Y (a) = 0. Since Y (a) = 0 would
imply that y is the trivial solution in contradiction to it being an eigenvector, we
have that det[A+B Φ(b, a, λ)] = 0. If (4.8) has two linearly independent solutions for
C, say C1 , C2 , then solve the IVP with the initial conditions Y (a) = C1 , Y (a) = C2
to obtain solutions Y1 , Y2 . Then Y1 , Y2 are linearly independent vector solutions of
(4.5) and their top components y1 , y2 are linearly independent solutions of (4.1).
Conversely, if y1, y2 are linearly dependent solutions of (4.1) we can reverse the
steps above to obtain two linearly independent solutions of the algebraic system
(4.8).
LEMMA 4.3. For the BVP (4.1), (4.2), (4.3) exactly one of the following four
cases holds:
1. There are no eigenvalues in C .
2. Every complex number is an eigenvalue.
3. There are a nonzero finite number of eigenvalues in C .
4. There are an infinite but countable number of eigenvalues in C and these have
no finite accumulation point in C.
PROOF: This follows directly from Lemmas 4.1 and 4.2 and the fact that the zeros
of an entire function are isolated and have no accumulation point in the finite plane
C. 2
It is convenient to separate the boundary conditions (BC) (4.3) into two mutually
exclusive classes: separated and coupled. Note that, since the BC are homogeneous,
multiplication by a nonzero constant or a nonsingular matrix leads to equivalent
boundary conditions.
LEMMA 4.4 (Separated BC). Let (4.1) to (4.7) hold. Fix P, W, J and assume
A1 A2 0 0
(4.9) A= , B= .
0 0 B1 B2
Then
for λ ∈ C.
PROOF: This follows directly from the definition of ∆. 2
49
LEMMA 4.5 (Coupled self-adjoint BC). Let (4.1) to (4.7) hold. Fix P, W, J and
assume that
−1 0
(4.10) B= , A = eiα K, −π ≤ α ≤ π, K ∈ SL2 (R),
0 −1
(4.11)
D(λ, K) = k11 φ11 (b, a, λ) + k12 φ21 (b, a, λ) + k21 φ12 (b, a, λ) + k22 φ22 (b, a, λ)
Eberhard and Freiling [20], Mennicken and Möller [67], have established the
existence of infinitely many eigenvalues for the symmetric equation (4.1) but with
non-self-adjoint BC. We have
50
Then
1. the BVP (4.1), (4.9) with (A1 , A2 ) 6= (0, 0) 6= (B1 , B2 ) has an infinite number
of eigenvalues;
2. the BVP (4.1) together with the coupled BC
Y (b) = A Y (a)
We now pause to consider the simplest SLP for at least two reasons: (i) to illustrate
the results of the previous section and (ii) to indicate some of the coming attractions
of section 5. It is remarkable how many properties of SLP for the simplest SL
equation
−y 00 = λ y
hold for the general case. This is a third reason for discussing this equation here.
Consider the equation
We include the case when one or both endpoints are infinite here, even though this
is a singular problem then (p = 1 = w are not in L(J) if J is unbounded) and
singular problems are not discussed, in general, until the next section, to highlight
the interplay between regular and singular problems.
We content that to fully understand regular SLP requires a perspective which
includes the singular case.
Each infinite endpoint is in the LP case since the constant 1 is a non L2 solution
for λ = 0. Thus there is one and only one self-adjoint realization, say S, of the
equation (4.13) in the space L2 (−∞, ∞). The spectrum of S, σ(S), contains no
51
eigenvalues and thus coincides with the essential (continuous) spectrum σe (S); we
have
In this case
0, 1 0, 0
P = , W = .
0, 0 1, 0
Since these are fixed for this example we will omit them in the notation for Φ. This
fundamental matrix Φ = Φ(t, u, P, W, λ) = Φ(t, u, λ) is determined as the unique
solution of the initial value problem
Φ0 = (P − λW ) Φ, Φ(u) = I, t, u ∈ R, λ ∈ C.
√
To compute Φ we choose an analytic branch of the square root function z as
follows :
√
µ= λ = s + it, s > 0, t > 0, f or λ 6= 0,
and obtain
1
cosh(iµ(t − u)), iµ sinh(iµ(t − u))
(4.14) Φ(t, u, λ) = ,
iµ sinh(iµ(t − u)), cosh(iµ(t − u))
√
t, u ∈ R, λ ∈ C, λ 6= 0, µ = λ,
and
1, t − u
(4.15) Φ(t, u, 0) = , t, u ∈ R.
0, 1
Note that for fixed t, u ∈ R the fundamental matrix Φ(t, u, λ) is analytic at λ for
each λ ∈ C including λ = 0. (This can be confirmed from the series expansions of
the hyperbolic sinh and cosh functions.)
For the convenience of the reader we now recall some definitions and properties
of hyperbolic functions which will be used below.
1. 2 sinh z = ez − e−z z ∈ C, 2 cosh z = ez + e−z , z ∈ C, tanh z = sinh z/ cosh z,
2. sinh z = −i sin iz, cosh z = cos iz, tanh z = −i tan iz
3. sinh(z + 2kπi) = sinh z, cosh(z + 2kπi) = cosh z
4. (sinh z)0 = cosh z, (cosh z)0 = sinh z
52
1, 0 c, 0
(4.17) A= , B= .
0, 1 0, d
√
(4.18) ∆(λ) = 1 + cd + (c + d) cosh(iµ(b − a)), µ = λ 6= 0.
1 + cd
cosh((iµ(b − a)) = cos(−µ(b − a)) = − = r.
c+d
and both integrals must be taken along a path which does not cross the
real axis.
When r is real and −1 ≤ r ≤ 1 then the roots for µ(b − a) are real and
we get
(π/2 + 2nπ)2
λn = , n ∈ N0 .
(b − a)2
(2nπ)2
λP
n = , n ∈ N0 .
(b − a)2
((2n + 1)π)2
λSn = , n ∈ N0 .
(b − a)2
54
1 + e2iα
r=− = − cos α, t0 (α) = arccos(− cos α) = π − α ∈ (0, π).
2eiα
So
(b − a)µ = π − α + 2kπ, k ∈ N0 ,
and therefore
(π − α + 2nπ)2
λn (α) = , n ∈ N0 .
(b − a)2
A1 , A2 0, 0
A= , B= , Aj , Bj ∈ C, j = 1, 2.
0, 0 B 1 , B2
1
∆(λ) = ( A1 B1 − iµA2 B2 ) sinh((iµ(b − a)).
µ
Since µ = s > 0 we get the following eigenvalues from the periodic factor:
(nπ)2
λn = , n ∈ N.
(b − a)2
(nπ)2
λn = , n ∈ N.
(b − a)2
µ(b − a) = kπ, k ∈ Z.
Since λ = 0 is also an eigenvalue in this case and the first factor does not
produce an eigenvalue we get that
(nπ)2
λN
n = , n ∈ N0 .
(b − a)2
56
4. A1 = 1 = B2 , A2 = 0 = B1 .
(π/2 + 2nπ)2
λDN
n = , n ∈ N0 .
(b − a)2
5. A1 = 0 = B2 , A2 = 1 = B1 .
Here we have the same roots and hence the same eigenvalues as in the
previous case
(π/2 + 2nπ)2
λN
n
D
= , n ∈ N0 .
(b − a)2
6. A1 B2 − A2 B1 6= 0.
( µ1 A1 B1 − iµA2 B2 )
coth(iµ(b − a)) = .
A1 B2 − A2 B1
The roots of this equation are not so easy to find explicitly since the unknown µ
appears on both sides. However, numerical approximations can be obtained from
a root finder code.
Some observations
1. Let {ak : k ∈ N } be a decreasing sequence to −∞; {bk : k ∈ N }an increasing
sequence to +∞ and let
E = {λD
n (ak , bk ) : n ∈ N0 , k ∈ N.
Then E is dense in [0, ∞). Thus every point of the essential spectrum of the
self-adjoint realization S of the Fourier equation on (−∞, ∞) is the limit of
a sequence of eigenvalues of regular problems on the intervals (ak , bk ), k ∈
N. This illustrates a general result, to be discussed in section 5, about the
approximation of the spectrum of singular SLP by eigenvalues of a sequence
of regular SLP.
2. For any n ∈ N0 the Dirichlet eigenvalue λD n is greater than or equal to λn for
any other self-adjoint boundary condition.
3. λD
n (a, b) → 0 as (b − a) → ∞ for each n ∈ N0 .
4. λD
n (a, b) → ∞ as (b − a) → 0 for each n ∈ N0 .
57
We want to show that if two SLP are “close” to each other then their eigenvalues
and eigenfunctions are also “close” to each other. To study the “closeness” of two
BVP we introduce a “boundary value problem space” with a distance function. Let
0 0 0 0
J = (a , b ), −∞ ≤ a < b ≤ ∞,
such that
0 0
−∞ ≤ a < a < b < b ≤ ∞, A, B ∈ M2 (C), p, q, w : J → C, 1/p, q, w ∈ Lloc (J).
where
q on [a, b]
(4.21) q̃ =
0 otherwise
and 1/p,
g w̃ are defined similarly. Now we introduce the Banach space
58
Z b0
(4.23) ||ω|| = ||ω̃|| = |a| + |b| + ||A|| + ||B|| + |1/p|
g + |q̃| + |w̃|
a0
Are the eigenvalues of a regular SLP continuous functions of the problem? The
answer is YES and NO.
NO, because for a fixed index n, the n-th eigenvalue λn of a self-adjoint SLP is,
in general, not a continuous function of the problem i.e. of ω.
YES, because every isolated eigenvalue can be embedded in a “continuous eigen-
value branch”.
In this section these statements will be made precise. Also the continuous de-
pendence of the eigenfunctions on the problem is studied here.
Consider the SLP consisting of the equation
(4.27) p, q, w : (a, b) → R, 1/p, q, w ∈ L(a, b), p > 0, w > 0, a.e. on (a, b).
59
It is well known that this problem is self-adjoint and has an infinite but countable
number of eigenvalues {λn : n ∈ N0 }, these are all real, simple, bounded below and
can be indexed to satisfy
REMARK 8. Kong and Zettl [57] have shown that each continuous eigenvalue
branch is in fact differentiable everywhere including the point A0 (or B0 ) where the
index jumps. This also follows from Möller and Zettl [69].
EXAMPLE 4.8. Consider the BVP with Dirichlet BC and the equation
−1, if 0 ≤ t ≤ ε,
pε (t) =
1, ε < t ≤ 1.
Then for ε = 0 the spectrum is bounded below but for each ε > 0 the spectrum is
unbounded below. Note that 1/pε → 1/p0 in L(0, 1).
Furthermore, if λ(ω0 ) is simple then there is exactly one λ(ω) satisfying (4.29); if
λ(ω0 ) is a double eigenvalue, then either (4.29) holds for a double eigenvalue of ω
or for exactly two simple ones.
PROOF. See Kong and Zettl [57] Theorem 3.1. 2
Theorem 4.10 that every SLP ω which is sufficiently close to ω0 , whether it is self-
adjoint or not, must have n eigenvalues close to n eigenvalues of λ(ω0 ), for any
positive integer n. Note that it does not follow directly from Theorem 4.10 that ω
must have an infinite number of eigenvalues close to those of ω0 .
Z b
(4.30) |u|2 w = 1.
a
Next we state a result for normalized eigenfunctions. Note that these are not
uniquely determined. In the case of a simple eigenvalue they are unique up to
sign, but for a double eigenvalue there are pairs of linearly independent normalized
eigenfunctions.
Below when considering SLP on two subintervals of J we extend the solutions
and their quasi-derivatives continuously to the whole of the open interval J. This
is done to facilitate comparisons between solutions and their quasi-derivatives on
different intervals.
THEOREM 4.11. Let the notation and hypotheses of Theorem 4.10 hold.
(i) Assume the eigenvalue λ(ω0 ) is simple for some ω0 ∈ Ω and let u(·, ω0 ) denote
a normalized eigenfunction of λ(ω0 ). Then there is a neighborhood M of ω0 in Ω
such that λ(ω) is simple for every ω ∈ M and there exist normalized eigenfunctions
u(·, ω) of λ(ω) for ω ∈ M such that
both uniformly on any compact subinterval K of (a0 , b0 ). Note that in this case,
given two linearly independent normalized eigenfunctions uj of λ(ω0 ) there ex-
ist a pair uj of linearly independent normalized eigenfunctions of λ(ω) such that
uj (·, ω) → uj (·, ω0 ) as ω → ω0 in Ω for j = 1, 2.
62
0 0
−∞ ≤ a < a < b < b ≤ ∞, A, B ∈ M2 (C),
0, −1
A E A∗ ∗
= B E B , rank (A|B) = 2, E = ,
1, 0
p, q, w : J → R, 1/p, q, w ∈ Lloc (J), w > 0, J = (a0 , b0 ).
Note the different normalization in (4.36) for β than that used for α in (4.35).
This is for convenience in stating some of the results below.
3. All real coupled self-adjoint BC. These can be formulated as follows:
y(b) y(a)
(4.37) =K
(py 0 )(b) (py 0 )(a)
k11 k12
(4.38) K= , kij ∈ R, det K = 1.
k21 k22
y(b) y(a)
(4.39) = exp(i α) K
(py 0 )(b) (py 0 )(a)
Most of the following results are well-known. See [81] for some proofs with only
integrable coefficients; see [68] for the case when p changes sign, and see [8], [18]
for the case of complex couple BC.
Then for ω ∈ Ωs the BVP ω has only real and simple eigenvalues; there are an
infinite but countable number of them; they are bounded below and can be ordered
to satisfy
Notation. Let
For ω ∈ Ωrc the BVP ω has only real eigenvalues; each of these may be simple or
double; there are an infinite but countable number of them and they can be ordered
to satisfy
(4.45) −∞ < λ0 ≤ λ1 ≤ λ2 ≤ . . . ; λn → +∞ , as n → ∞.
Notation. Let
Note that there is some arbitrariness in the indexing of the eigenfunctions corre-
sponding to a double eigenvalue.
For ω ∈ Ωcc the BVP ω has only real and simple eigenvalues; there are an infinite
but countable number of them and they can be ordered to satisfy
Z b r !−2
λn w
(4.49) → c = π2 , as n → ∞.
n2 a p
If we fix all variables except α and shorten the notation to λn = λn (α), then we
have λn (−α) = λn (α),and the complex conjugate of an eigenfunction of λn (α) is
an eigenfunction of λn (−α).
(b) Assume that p changes sign in the interval [a,b], i.e. p is positive on
a subset of [a, b] of positive Lebesgue measure and p is negative on a subset of the
interval [a, b] of positive Lebesgue measure. Then
65
Each BVP ω ∈ Ωs has only real and simple eigenvalues; there are an infinite
but countable number of them; they are unbounded below and above and can be
ordered to satisfy
Each BVP ω ∈ Ωrc has only real eigenvalues; each of these may be simple or
double; there are an infinite but countable number of them; they are unbounded
below and above and can be ordered to satisfy
The notations for eigenvalues λn and eigenfunctions un , n ∈ Z, for part (b) are the
same as those introduced in part (a) for n ∈ N0 .
PROOF. The fact that the eigenvalues are unbounded below when p changes sign
was estalished by M. Möller. This holds even if there is no subinterval on which
p is negative. The fact that the eigenvalues for any ω ∈ Ωcc are all simple follows
from Weidmann [81], although it doesn’t appear to be stated explicitly there for
the general case. See also [8]. The other results are standard. 2
Also assume that p > 0 a.e. on J and denote the eigenvalues for this boundary
condition by λn (α, K), abbreviated to λn (K) when α = 0, for n ∈ N0 .
Suppose that either k12 < 0 or k12 = 0 and k11 + k22 > 0. Then
1. λ0 (K) is simple;
2. λ0 (K) < λ0 (−K) and
3. the following inequalities hold for −π < α < 0 and 0 < α < π :
−∞ < λ0 (K) < λ0 (α, K) < λ0 (−K) ≤ λ1 (−K) < λ1 (α, K) < λ1 (K)
≤ λ2 (K) < λ2 (α, K) < λ2 (−K) ≤ λ3 (−K) < . . .
66
λ0 (α, K) < λ0 (β, K) < λ1 (β, K) < λ1 (α, K) < λ2 (α, K) < λ2 (β, K)
< λ3 (β, K) < λ3 (α, K) < . . .
Suppose that either k12 > 0 or k12 = 0 and k11 + k22 < 0. Then
1. λ0 (−K) is simple;
2. λ0 (−K) < λ0 (K) and
3. the following inequalities hold for −π < α < 0 and 0 < α < π :
−∞ < λ0 (−K) < λ0 (α, K) < λ0 (K) ≤ λ1 (K) < λ1 (α, K) < λ1 (−K)
≤ λ2 (−K) < λ2 (α, K) < λ2 (K) ≤ λ3 (K) < . . .
λ0 (β, K) < λ0 (α, K) < λ1 (α, K) < λ1 (β, K) < λ2 (β, K) < λ2 (α, K)
< λ3 (α, K) < λ3 (β, K) < . . .
0 0
p ≥ 0 a.e. and q 2 /w ∈ Lloc (a , b ).
0
Then for any n ∈ N0 , λD
n (b) is strictly decreasing in (a, b ) and
λD +
n (b) → ∞ as b → a .
0
Q = q/w ∈ ACloc [a, b ), p(b) ≥ δ > 0 f or b ∈ (a, b0 ).
Then
67
1
(4.52) λ0 (a) = |pu0 |2 (a) − |u|2 (a)[q(a) − λ(a)w(a)] a.e. in U.
p(a)
1
(4.53) λ0 (b) = − |pu0 |2 (b) + |u|2 (b) [q(b) − λ(b)w(b)].
p(b)
REMARK 11. In his well known monograph on variational methods for eigen-
value problems Hans Weinberger states, without proof or reference to a proof, that
68
the Dirichlet eigenvalues are decreasing functions of the length of the interval but
that this is not true for the Neuman eigenvalues. Theorem 4.16 sheds a great deal
of light on this.
Assume that p > 0. For Dirichlet BC u(a) = u(a, a, 0, π) = 0 and u(b, b, 0, π) = 0
hence the second term in (4.52) and in (4.53) is zero; thus it is clear from these
formulas that the Dirichlet eigenvalues are increasing functions of the left endpoint
and decreasing functions of the right endpoint. It is also clear from (4.52), (4.53)
that this is not true, in general, for any other boundary conditions. However if q/w is
bounded above, say by C, then for any boundary conditions all eigenvalues greater
than C are increasing functions of the left endpoint a and decreasing functions
of the right endpoint b. Since for any fixed regular SLP the eigenvalues λn →
∞ asymptotically as n2 it is clear that if q/w is bounded above only the lower
eigenvalues may fail to be monotonic functions of the length of the interval.
1. Fix all components of ω except β and let λ = λ(β) and u = u(·, β). Then λ
is differentiable and
2. Fix all components of ω except α min(4.39) and let λ = λ(α) and u = u(·, α).
Then λ is differentiable at α for any α satisfying −π < α < 0 or 0 < α < π
and
u(b)
(4.57) λ0 (K) H = [pu0 (b), −u(b)] HK −1 , H ∈ M2,2 (C).
(pu0 )(b)
69
PROOF. See [57] for parts (1) and (2) and see [69] for part (3). 2
∗ ∗ 0 −1
AE A = B E B , rank (A|B) = 2, E = .
1 0
Z b
(4.58) λ0 (q) h = |u(·, q)|2 h, h ∈ L(a, b),
a
Z b
0
(4.59) λ (1/p) h = − |(pu0 )(·, 1/p)|2 h, h ∈ L(a, b),
a
Z b
(4.60) λ0 (w) h = −λ(w) |u(·, w)|2 h, (h ∈ L(a, b)),
a
Then λ(1/P ) ≥ λ(1/p). If 1/P < 1/p on a subset of [a, b] having positive
Lebesgue measure, the λ(1/P ) < λ(1/p).
3. Fix p, q. Suppose W ∈ L(a, b) and W ≥ w > 0 on [a, b]. Assume that λ(s(t))
is on the same continuous eigenvalue branch as λ(w) for all t ∈ [0, 1], where
Then λ(W ) ≥ λ(w) if λ(W ) < 0 and λ(w) < 0; but λ(W ) ≤ λ(w) if λ(W ) > 0
and λ(w) > 0. Furthermore, if strict inequality holds in the hypothesis on a set
of positive Lebesgue measure, then strict inequality holds in the conclusion.
PROOF. We give the proof for (1), the proofs of (2) and (3) are similar. Define
Z b
f 0 (t) = λ0 ((s(t)) s0 (t) = |u2 (r, s(t))| (Q(r) − q(r)) dr ≥ 0, t ∈ [0, 1].
a
Hence f (1) = λ(Q) ≥ λ(q) = f (0). The strict inequality part of the theorem also
follows from this. 2
Z b
(4.61) λ0 (ω) ρ = {−py 0 p̄z̄ 0 (1/r) + [g − λ(ρ)v]yz̄} + d∗ [CY (a) + DY (b)],
a
where ρ = (C, D, 1/r, g, v) ∈ Ω, y, z are biorthogonal solutions of the given and its
adjoint boundary value problems at λ(ω), i.e.
0 0 y
(4.62) −(py ) + qy = λ(ω) w y, AY (a) + BY (b) = 0, Y =
py 0
71
Z b
0 0 z
(4.63) −(p̄z ) + q̄z = λ̄(ω)w̄z, Z = , y z̄ w = 1,
p̄z 0 a
0 −1
(4.64) Z(a) = EA∗ d, Z(b) = −EB ∗ d, E = .
1 0
REMARK 12. Note that for Theorem 4.20 no self-adjointness hypothesis is need-
ed: the coefficients p, q and the weight function w may be complex-valued; the
boundary conditions need not be self-adjoint. The existence of infinitely many
eigenvalues for non-self-adjoint SLP is well known for so called Birkhoff regular and
Stone regular SLP, see [67], [20].
4.8 Comments.
5.1 Introduction.
In this section we discuss singular self-adjoint SLP. This is a field so vast that we
can only hope to give a brief introduction here. Following a review of the some
of the basic theory we will focus on two topics: (i) The behavior of eigenvalues of
regular SLP near a singular boundary and (ii) The approximation of the discrete
as well as continuous (essential) spectrum of a given singular problem with spectra
of regular problems. On these two topics we aim to bring the reader to the frontier.
To illustrate some of the basic behavior of the spectrum of SLP we discuss briefly
the 29 examples from the sleign2 package in Subsection 5.8.
73
Note that for this section we assume that p ≥ 0 and w > 0 unless explicitly stated
otherwise.
Recall that, according to Proposition 3.10 in section 3 no nontrivial solution of
(5.1) can have an accumulation point of zeros in the interior of J. The zeros, if any,
of any nontrivial solution of (5.1) inside the interval J are isolated. Thus only an
endpoint of J can be an accumulation point of zeros of a nontrivial solution of (5.1)
and that can happen only at a singular endpoint.
LEMMA 5.2. If (5.1) has a principal solution u at a, then every non-zero real
multiple of u is also a principle solution and no other solution is a principal solution
at a.
PROOF. This follows directly from the definition. 2
REMARK 14. If the equation in (5.1) is regular at a then for any solution y, y
and py 0 can be continuously extended to a and principal solutions u exist and satisfy
the initial conditions : u(a) = 0, (pu0 )(a) 6= 0. Any non-principal solution v at a
satisfies : v(a) 6= 0.
Z d
1
(5.3) = ∞;
a p u2
Z d
1
(5.4) < ∞;
a p v2
Z t
c
(5.5) u(t) = v(t) , a < t ≤ d,
a p v2
4. and
Boundary conditions of the form (4.3) do not make sense when one endpoint, say
a, is singular since Y (a) does not exist, in general. What takes their place ? This
depends on the endpoint classification e.g. LP or LC. Before going into the details
75
we need to set the stage first. Recall the definitions of the maximal domain ∆ and
the sesquilinear form [·, ·] from Subsection 3.6 of Section 3.
DEFINITION 5.5 (Maximal and Minimal Operators). Let (5.1) hold and let ∆
and M be defined as in
Subsection 3.6. Define
T1 f = M f, f or f ∈ ∆.
LEMMA 5.6. The maximal and minimal domains are dense in the Hilbert space
Z
H = L2 (J, w) = {f : J → C, |f |2 w < ∞},
J
T0 is a closed symmetric operator and T0∗ = T, T ∗ = T0 . Hence any self-adjoint
extension of T0 is also a self-adjoint restriction of T and conversely.
PROOF. See [72], [81]. 2
(5.7) T0 ⊂ S = S ∗ ⊂ T1 ,
and can be determined by two point boundary conditions. These, however, are
vacuous at an LP endpoint. To describe these conditions it is convenient to take
cases depending on the LP/LC classification of the endpoints. Here LC/LP will
mean that the left endpoint a is LC and the right endpoint b is LP, etc.
An operator S satisfying (5.7) is called a self-adjoint extension of T0 on J, or a
self-adjoint restriction of T1 on J, or simply a self-adjoint realization of the equation
(5.1) on J, or a self-adjoint realization of (M, w) on J.
[y, u]
(5.10) AY (a) + BY (b) = 0, Y = , A, B ∈ M2 (C),
[y, v]
0 −1
rank(A : B) = 2, AEA∗ = BEB ∗ , E = .
1 0
REMARK 15. Functions u, v needed in (5.10) can often, but not always, be
obtained by choosing linearly independent real solutions of (5.1) for some particular
real value of λ, e.g. λ = 0.
some of the literature this is also referred to as the continuous part of the spectrum.
Either one, but not both, of σd and σe may be empty.
Next we summarize some basic properties of the spectrum.
REMARK 16. We see from this Proposition that the essensial spectrum does
not depend on the boundary conditions; thus σe = σe (J, p, q, w). The eigenvalues
do depend on the boundary condition. To the left of the continuous spectrum
there may be no eigenvalues, a finite number of them, or an infinite number of
them. In the case of an infinite number of eigenvalues to the left of σ0 = inf σe
these can have no accumulation point other than, possibly, σ0 and −∞. There also
may be eigenvalues embedded in the continuous spectrum as well as in gaps of the
continuous spectrum. The essential spectrum may have no gaps, a finite number of
them, or an infinite number of them. A remarkable result of Hartmann [41] states
that any closed set of real numbers which is not bounded above is the spectrum of
a Sturm-Liouville operator ! See also Halvorsen [37]. In particular the spectrum of
an SLP can be a Cantor-like set which is not bounded above.
For specific examples of SLP which illustrate some of these features (but not
the Cantor-like behavior of the spectrum) the reader is referred to the examples
discussed in Subsection 5.8 which come with the code SLEIGN2; these, along with
the code, can be downloaded from the WWW using netscape or mosaic or lynx by
specifying the URL :
ftp://ftp.math.niu.edu./pub/papers/Zettl/Sleign2
or by accessing the web page
http://www.math.niu.edu./˜zettl/SL2/
If there is one proper self-adjoint extension of the minimal operator then there are
an infinite number of such extensions. Friedrichs singled out one of these which
he called “ausgezeichnete” and which has come to be known as “the Friedrichs
extension”. He singled one out by giving a construction which constructed such
an extension while preserving the lower bound. (However having the same lower
bound does not characterize the Friedrichs extension since there are, in general,
other self-adjoint extensions which also have the same lower bound.) To get the
Friedrichs extension you must use the Friedrichs construction or some equivalent
78
version of it. This construction works for any symmetric, densely defined, bounded
below, operator in a Hilbert space. When it is applied to T0 it makes no explicit
use of boundary conditions. Thus the question arises : What boundary conditions
determine the Friedrichs extension of the minimal operator T0 ?
SF y = S ∗ y, f or y ∈ D(SF ).
Then SF is called the Friedrichs extension of S. According to the well known result
of Friedrichs [28], SF is a self-adjoint operator with the same lower bound as S.
Friedrichs himself [29] addressed the question of which boundary condition de-
termines the Friedrichs extension for regular SLP ? For regular SLP the answer in
general is : the Dirichlet condition, see [73], [50], [10], [78]. We take up the singular
case next.
THEOREM 5.10. If the minimal operator T0 is bounded below with lower bound
c then the equation (5.1) is NO at a and at b for any λ < c. Conversely, if the
equation (5.1) is NO at a for some real λa and at b for some real λb , then the
minimal operator T0 is bounded below.
PROOF. See Niessen and Zettl [74] Corollary 2.1 and Theorem 4.2. 2
THEOREM 5.12. Let (5.1) hold. Let the maximal domain ∆, the Lagrange
form [·, ·] and the expression M be defined as in Subsection 3.6. Assume that each
endpoint is either regular or LCNO. Then the minimal operator is bounded below
and thus has a Friedrichs extension SF . Let ua be a principal solution at a for some
real λa and let ub be a principal solution at b for some real λb . Then the domain
D(SF ) of SF is given by
Note that (5.11) is independent of the principal solution chosen for λa and is
independent of λa ∈ R; similarly at b.
PROOF. See Niessen and Zettl [74], Theorems 4.2 and 4.3. 2
REMARK 17. At a regular endpoint, say a, (5.11) reduces to y(a) = 0. Thus the
BC (5.11) can be viewed as the singular analogues of the regular Dirichlet boundary
conditions. It is also shown in [74] that the conditions (5.11) are equivalent to
y(t) y(t)
lim+ = 0 = lim− ,
t→a va (t) t→b vb (t)
We now study the case when each endpoint is either regular or LCNO. The proper-
ties of the eigenvalues and eigenfunctions in this case are similar to the regular case.
In fact Niessen and Zettl [74] have shown that, given any SLP with endpoints which
are either regular or LCNO there exists a regular SLP which has exactly the same
spectrum as this singular problem and furthermore the eigenfunctions of the given
singular problem {yn : n ∈ N0 } are related to the eigenfunctions {zn : n ∈ N0 } of
the corresponding regular problem by the equation
for some function v in the maximal domain of the singular problem which satisfies
v(t) > 0 for t ∈ (a, b). Since each zn is a solution to a regular problem on (a, b) z
and its quasi-derivative can be continuously extended to the endpoints by Theorem
3.4. Hence the singular behavior at each endpoint is contained in v. In particular,
this shows that at each endpoint the singular (e.g. asymptotic) behavior of all
eigenfunctions (in fact of all solutions for all real λ) is the same.
THEOREM 5.13. Let (5.1) hold and assume that each endpoint is either R or
LCNO. Let S be a self-adjoint realization of (5.1).
1. Then the spectrum of S is discrete and bounded below. It consists of a
countably infinite sequence {λn : n ∈ N0 } of real eigenvalues tending to +∞
which can be ordered to satisfy
(5.13) −∞ < λ0 ≤ λ1 ≤ λ2 ≤ λ3 ≤ . . . → ∞.
Here the eigenvalues are counted according to their multiplicity. Each eigen-
value can have multiplicity one, in which case it is called simple, or two, in
80
which case it is called double. Therefore, in (5.13), equality cannot hold for
more than two consecutive terms.
2. Let yn be a real eigenfunction of λn . Then yn has at least n − 1 and at most
n + 1 zeros in (a, b). If [yn , ua ](a) = 0, where ua is a principal solution at a,
then yn has at most n zeros in (a, b).
3. Let N (λ) denote the number of eigenvalues of S in the interval (−∞, λ]. Then
s
Z b
N (λ) 1 w(t)
(5.14) √ → dt < ∞, as λ → ∞,
λ π a p(t)
4. and
λn 1
(5.15) → q 2 , as n → ∞.
n2 π 2 R b w(t)
a p(t) dt
The finiteness of the integral in (5.14) and (5.15) is a consequence of the assump-
tion that each endpoint is R or LCNO.
PROOF. See Theorem 5.2 in [74]. 2
THEOREM 5.14. Let (5.1) hold. Assume that each endpoint is regular or LCNO.
Let u, v be real maximal domain functions such that u, v are principal and non-
principal solutions at a for some real λa and at b for some real λb , respectively,
normalized to satisfy : [u, v](a) = 1 = [u, v](b). (Such u, v exist by Lemma 5.2 and
Theorem 5.3 ). Let S with spectrum σ(S) be the self-adjoint realization determined
by the normalized separated boundary conditions
Then
1. σ(S) = {λn : n ∈ N0 } and each eigenvalue is simple.These eigenvalues can be
ordered to satisfy
5. with the understanding that terms involving λ−1 , λ−2 are not present we have
the following inequalities for α ∈ [0, π), β ∈ (0, π], and n ∈ N0
λn−1 (0, β) λn (0, β)
(5.19) λn−2 (0, π) < < λn (α, β) ≤ ≤ λn (0, π).
λn−1 (α, π) λn (α, π)
PROOF. The regular case is known, see [81]. The singular case then follows from
the regular case and the transformation employed in [74] to “regularize” singular
LCNO endpoints. 2
THEOREM 5.15 (Canonical Coupled BC). Let (5.1) hold. Assume that each
endpoint is regular or LCNO. Then the canonical form of all coupled self-adjoint
boundary conditions is
where −π ≤ α ≤ π,
[y, θ]
Y = , θ, ϕ ∈ ∆, θ, ϕ real, K ∈ SL2 (R),
[y, ϕ]
(5.21) [θ, ϕ](a) = 1 = [θ, ϕ](b); Y (a) = lim+ Y (t), Y (b) = lim− Y (t).
t→a t→b
Fix p, q, w, a, b and let λn (α, K), n ∈ N0 denote the eigenvalues for BC (5.20);
when α = 0 this notation is abbreviated to λn (K). Suppose that
Then
1. λ0 (K) is simple;
2. λ0 (K) < λ0 (−K);
3. the following inequalities hold for -π < α < 0 and 0 < α < π :
−∞ < λ0 (K) < λ0 (α, K) < λ0 (−K) ≤ λ1 (−K) < λ1 (α, K) < λ1 (K)
(5.22) ≤ λ2 (K) < λ2 (α, K) < λ2 (−K) ≤ λ3 (−K) < . . .
λ0 (α, K) < λ0 (β, K) < λ1 (β, K) < λ1 (α, K) < λ2 (α, K) < λ2 (β, K)
(5.23) < λ3 (β, K) < λ3 (α, K) < . . .
82
Note that K ∈ SL2 (R) implies −K ∈ SL2 (R) and therefore if the hypothesis :
The next result characterizes the eigenvalues of singular SLP consisting of the
canonical form (5.20) of the coupled self-adjoint boundary conditions but for the
general equation (5.1) with real or complex valued coefficients.
where
with the boundary conditions (5.20), (5.21) and assume each endpoint is either R
or LC. For each λ ∈ C determine unique solutions u = u(·, λ), v = v(·, λ) by the
‘singular initial conditions’
Such solutions u, v exist by Theorem 3.13. Let K ∈ SL2 (R). Then for any α,
−π ≤ α ≤ π, a number λ ∈ C is an eigenvalue of the BVP (5.26), (5.25), (5.20),
(5.21), (5.24) if and only if
where for λ ∈ C
PROOF. See Theorem 3.1 in [8]. Although this Theorem is stated there only for
the case when p, q, w are real valued and w > 0 the proof given there holds with no
significant changes when p, q, w are complex valued. 2
For this subsection we change the notation for the interval J from
J = (a, b)
to
J = (a0 , b0 ).
The reason for this change of notation is that we wish to consider “approximations”
of a singular SLP on an interval (a0 , b0 ) by a sequence of regular SLP on truncated
intervals (ar , br ) where
where u, v are real valued maximal domain functions satisfying [u, v](a0 ) = 1.
The inherited BC on (ar , br ) are obtained by replacing a0 by ar in (5.30)
and by using the Dirichlet conditions y(br ) = 0 at br . Note that, although
u, v and their quasi-derivatives may not be defined at a0 they are well defined
at any point a0 < ar < br < b0 .
• Assume b0 is LC and a0 is LP. Then there is no singular boundary condition
at a0 and all singular self-adjoint BC at b0 have the form
where u, v are real valued maximal domain functions satisfying [u, v](b0 ) = 1.
The inherited BC on (ar , br ) are obtained by replacing b0 by br in these
conditions and by using the Dirichlet conditions y(ar ) = 0 at ar . Note that,
although u, v and their quasi-derivatives may not be defined at b0 they are
well defined at any point br , a0 < ar < br < b0 .
• Each endpoint is either regular or LC. In this case we have self-adjoint real-
izations determined by both separated and coupled BC. Let the BC on (a0 , b0 )
be determined by (5.10) but with our changed notation a0 for a and b0 for b.
To obtain the inherited BC just replace a, b in (5.10) by (ar , br ). Note that
although the same matrices A, B occur in the BC on (a0 , b0 ) and on (ar , br )
when these inherited boundary conditions are written in the usual form (4.3)
for regular BC the coefficient matrices, say A = A(ar ), B = B(br ) depend on
values of u, v and their quasi-derivatives at ar , br . In particular as the end-
point ar or br is changed, e.g. in the code sleign2, the inherited boundary
conditions change accordingly.
• The inherited operators and their spectral quantities are identified with the
superscript i : Sri , λin (ar , br ).
DEFINITION 5.20 (Start of the essential spectrum). For any operator S let
(5.31) σ0 = inf σe , −∞ ≤ σ0 ≤ ∞,
REMARK 18. We comment on the contrast between (5.32) and (5.29). This
markedly different behavior of the eigenvalues of regular problems shows the enor-
mous influence that the spectrum of a singular problem has on the regular problems
which are “close” to the singular one; in this case by virtue of the fact that the
endpoints of the regular problem are close to the endpoints of the singular one.
To understand the behavior of the eigenvalues of regular problems one needs a
perspective which includes the singular case. This is even more interesting when
viewed in the light of the asymptotic formula (5.15) for the eigenvalues on each
fixed interval (ar , br ).
PROOF. This is contained in Theorem 4.1 of [6]; see also Remark 1 (ii) on pages
15-16. 2
In this case, i.e. when the eigenvalues are unbounded below as well as above,
we follow the SLEIGN2 convention that λ0 denotes the smallest nonnegative eigen-
value. This makes the indexing scheme unique.
(5.36) λn (Sri ) → σ0 , r → ∞, n ∈ N0 .
4. etc
5. If S has an infinite number of eigenvalues {λn : n ∈ N0 } to the left of σ0 ,
then
(5.39) λn (Sri ) → λn , r → ∞, n ∈ N0 .
REMARK 19. Theorem 5.23 can be used to detect the number of eigenvalues to
the left of the essential spectrum. Using one of the numerical codes such as sleign2
or the Fulton and Pruess code sledge [30] one can ascertain which of (5.36), (5.37),
(5.38) or (5.39) holds.
REMARK 20. The so called Coffee-Evans equation has attracted a good deal of
attention in the literature because it has the interesting feature that, despite the
asymptotic behavior (5.15), the lower eigenvalues occur in clusters of three which are
close together. It is clear from Theorem 5.23 how to construct examples of regular
problems with clusters of three million or three billion eigenvalues as close together
as you please. This also shows that it easy to construct examples of regular SLP
which will defeat any numerical code for the computation of eigenvalues. See Zettl
[86] for some illustrations of this. (This paper can be downloaded from the author’s
web page following the instructions for downloading sleign2 given in Subsection 3
above.)
THEOREM 5.24. Let S be any self-adjoint realization of (M, w). The sequence
of inherited operators {Sri : r ∈ N} is spectral included for S i.e. given any λ ∈ σ(S)
there exists an n(r, λ) ∈ N0 for each r ∈ N such that
87
THEOREM 5.25. Assume that σ(S) is bounded below. Then the sequence of
inherited operators {Sri : r ∈ N} is spectral exact for S below σ0 (S), i.e. it is spectral
included and if the convergence (5.40) holds for some λ < σ0 , then λ ∈ σ(S).
PROOF. This is contained in [6]. 2
By Theorem 5.23 any point of the spectrum of a singular problem can, in princi-
ple, be approximated arbitrarily closely by eigenvalues from the inherited sequence
of regular problems. In practice this isn’t feasible since there are an uncountable
number of points in the spectrum of the singular problem and finding an index
sequence for each one is a hopeless task.
Nevertheless Theorems (5.23), (5.24), (5.25) together than be used to approx-
imate the spectrum of many singular SLP quite effectively. This is done by ap-
proximating, not the individual points of the spectrum, but the spectral bands and
gaps. It is remarkable that for so many singular problems the first few spectral
bands and gaps - or the absence of gaps - can be detected and approximated from
the distribution of a few thousand eigenvalues of the inherited problems which can
be computed with sleign2 or sledge. For illustrations of this scheme see Zettl [88].
This paper can be downloaded as part of the SLEIGN2 package of files from the
internet; see the instructions in Subsection 3 of this Section.
5.8 Examples.
Here we list a few examples to illustrate some of the concepts and results discussed
above. These are taken from the the SLEIGN2 package. We follow the notation
established above. We supplement the endpoint classifications given above with the
weakly regular (WR) classification used by sleign2 : The endpoint a is WR if
but at least one of 1/p, q, w is not bounded in (a, c) for any c, a < c < b, or
w(a) = 0.
1. The Legendre equation on (−1, 1).
1 1+t
u(t) = 1, v(t) = log( ), −1 < t < 1.
2 1−t
88
λn = n(n + 1), n ∈ N0 ;
respectively.
2. The Liouville form of the Bessel equation on (0, ∞).
ν 2 − 1/4
p = 1, q(t) = , w = 1.
t2
LP at ∞ for all ν.
At 0 :
• LCNO for −1 < ν < 1 but ν 2 6= 1/4
• R for ν 2 = 1/4
• LP for ν 2 > 1/4.
ν < 0 note that v is the principal solution and hence [y, v](0) = 0 is the
Friedrichs BC.
3. The Halvorsen equation on (0, ∞).
e−2/t
p(t) = 1, q(t) = , w(t) = 1;
t4
p = 1, q(t) = −1/t, w = 1
Solutions can be given in terms of Whittaker functions, see [7]; the given
u, v are maximal domain functions which are not solutions for any λ. This
equation on the interval (−1, 1) , hence with an interior singularity at 0, arose
in a model in connection with the study of eddies in the atmosphere, see Boyd
[13].
5. The regularized Boyd equation on (−∞, 0) and on (0, ∞).
where
LP at −∞ and ∞; W R at 0− and 0+ .
This is a WR form of example 4; the singularity at zero has been “regular-
ized” using quasi-derivatives. There is a one-to-one correspondence between
all self-adjoint BC of this example and example 4 considered as a “two inter-
val” problem. Each self-adjoint realization of the Boyd equation on (−∞, ∞)
is unitarily equivalent to a self-adjoint realization of the regularized Boyd
equation and conversely. Thus these problems have the same spectrum. Also
the eigenfunctions are closely related (but, of course not the same since one
is singular and the other regular). For details see [4], [7], [21], [74].
6. The Sears -Titchmarsh equation on (0, ∞).
LP at 0; LCO at ∞.
This equation was studied by Titchmarsh [79] and by the two named au-
thors .
For problems on [1, ∞) the spectrum is discrete but unbounded above and
below; for some numerical results see [7].
k k2
p(t) = 1/t, q(t) = 2
+ , w = 1, k ∈ R, k 6= 0.
t t
LCNO at 0 for all k; LP at ∞ for all k. This is only a very special case of
the named equation, see Homer [48] for the general equation and references
to the applied literature.
Even for this special case there are no representations of solutions in terms
of the well known special functions. Thus to determine boundary conditions
one must use maximal domain functions. Such functions are given by
p(t) = 1 − t7 , q = 0, w(t) = t7 ,
to determine the boundary condition vsctor Y (1). This example has a long
and celebrated history; see Fichera [27].
10. A weakly regular equation on (0, ∞).
√ 1
p(t) = t, q = 0, w(t) = √ .
t
LP at −∞ and ∞.
This is a form of the Mathieu equation. Since both endpoints are LP there
are no boundary conditions needed or allowed; there is a unique self-adjoint
realization of this equation which has no eigenvalues. It’s spectrum consists
of an infinite number of compact intervals, the spectral bands, separated by
gaps. The first few spectral band are rather thin, the first has a width on the
order of 10−4 .
The eigenvalues of this problem on the interval (−b, b), 0 < b < ∞, with
Dirichlet boundary conditions
y(−b) = 0 = y(b),
tend to “bunch up” in the spectral bands of the whole line problem, partic-
ularly the fist band. See the comments in the last paragraph of Subsection
5.7. From Theorem 5.23 we know that
λn (b) → σ0 = inf σe .
In particular, for any positive ε and any positive integer n the first n
eigenvalues differ from each other by less than ε if b is sufficiently large, e.g.
one can choose b large enough so that the first 7 million eigenvalues agree
to the first 1000 digits. Given any numerical code for the computation of
Sturm-Liouville eigenvalues, it can be defeated simply by choosing a large
enough b. All this for such a “simple” regular SLP. The singular problem on
(−∞, ∞) exerts a strong influence on the behavior of the eigenvalues of the
regular problems on (−b, b).
12. The Mathieu equation on (−∞, ∞) and on (0, ∞).
p = 1, q(t) = sin(t), w = 1.
LP at −∞ and at ∞; R at 0.
92
k h
p(t) = 1, q(t) = + 2 , w(t) = 1, k, h ∈ R.
t t
This is the two parameter version of the classical one-dimensional equation for
quantum theory modelling of the hydrogen atom; see [49], section 10 where
most of the results reported on here can be found. A few of these results can
be found in the commentary file xamples.tex of the package of files comprising
the SLEIGN2 package. For all h, k there are no positive eigenvalues, ∞ is LP
and the essential spectrum is [0, ∞). If k = 0 the equation reduces to Bessel,
see example #2 with h = ν 2 − 1/4.
LP at ∞ for all h, k.
At 0 :
• R for h = 0 = k,
• LCNO for h = 0 and all k 6= 0
• LCNO for −1/4 ≤ h < 3/4, but h 6= 0 and all k
• LCO for h < −1/4 and all k
• LP for h ≥ 3/4 and all k.
Let
p
ρ= h + 1/4, f or h ≥ −1/4.
: (a) For h ≥ 3/4 and k ≥ 0 there is at most one negative eigenvalue and
λ = 0 may be an eigenvalue; for h ≥ 3/4 and k < 0 there are infinitely
many negative eigenvalues given by
−k 2
λn = , n ∈ N0 ,
(2n + 2ρ + 1)2
93
k
u(t) = tρ+1/2 , v(t) = t1/2−ρ + t3/2−ρ .
1 − 2ρ
−k 2
λ0 =
(2ρ − 1)2
(c) if k < 0, 0 < ρ < 1/2 there are infinitely many negative eigenvalues
given by
−k 2
λn = , n ∈ N0
(2n − 2ρ + 1)2
(d) if k < 0, 1/2 < ρ < 1 there are infinitely many negative eigenvalues
given by
−k 2
λn = , n ∈ N0
(2n − 2ρ + 3)2
1/ρ
−A1 Γ(1 + ρ)
λ0 = −4
A2 Γ(1 − ρ)
√ √ √ √ √
u(t) = t + kt t, v(t) = 2 t + ( t + kt t) log(t).
94
is Eulers constant.
(g) h < −1/4, k ∈ R, the equation is LCO at 0. For k = 0 this equation
reduces to the Krall equation, see example 20. For k 6= 0 explicit
formulas for the eigenvalues are not available; some qualitative prop-
erties of the spectrum are :
for all k there are infinitely many negative eigenvalues going expo-
nentially to −∞
for k > 0 the point 0 is not an accumulation point of eigenvalues
for k ≤ 0 the eigenvalues also accumulate at 0.
14. The Marletta equation on (0, ∞).
3(t − 31)
p = 1, q(t) = , w = 1.
4(t + 1)(4 + t)2
p = 1, q(t) = t2 , w = 1
λn = 2n + 1, n ∈ N0 .
At -1 for all α :
• LP for β ≤ −1 and for β ≥ 1
• WR for −1 < β < 0
• LCNO for 0 ≤ β < 1
At +1 for all β:
95
λn = n(n + α + β + 1), n ∈ N0 .
It is interesting to observe that the required boundary condition for the Jacobi
polynomials is the Friedrichs condition in the LCNO case but not in the WR
case.
17. The rotation Morse oscillator on (0, ∞).
2
p = 1, w = 1, q(t) = − 2000(2E − E 2 ), E = e−1.7(t−1.3) .
t2
2α2 2β 2
p(t) = 1 − t2 , q(t) = + , w = 1, 0 ≤ α, 0 ≤ β.
1+t 1−t
At -1:
96
Note that u, v are maximal domain functions but not solutions. In [17] on
p. 1519 it is claimed that the boundary value problem determined by the
boundary conditions
λn = (n + α + β + 1) (n + α + β), n ∈ N0 .
19. The Donsch equation on (−1, 1). This is a modification of example 18 which
illustrates an LCNO/LCO mix. Replace α in 18 by iγ. This changes the
singularity at -1 from LCNO to LCO. For γ > 0 and 0 < β < 1/2 we have
k 2 + 1/4
p = 1, q(t) = 1 − , w = 1, k ∈ R, k 6= 0.
t2
LP at ∞ for all α.
At 0 :
• LP for α ≤ −1
• WR for −1 < α < 0
• LCNO for 0 ≤ α < 1
• LP for α > 1.
This is the classical form of the celebrated equation, which for parameter
values α > 1 produces the Laguerre polynomials as eigenfunctions; for the
appropriate boundary at 0, when needed, the eigenvalues are given by
λn = n, n ∈ N0 .
Remarkably, these are independent of α, see Abramovitz and Stegun [1], chap-
ter 22, section 22.6 for more details. See the file xamples.f (this is not a typo)
of the sleign2 package for details of the boundary condition functions u, v.
The code sleign2 has only very limited success with this problem; for nu-
merical computations the Laguerre/Liouville equation, which has the same
eigenvalues (for the appropriate corresponding boundary conditions) is more
convenient, see example 23 to follow.
23. The Laguerre/Liouville equation on (0, ∞).
α2 − 1/4 α + 1 t2
p = 1, w = 1, q(t) = − + , α ∈ R.
t2 2 16
LP at ∞ for all α
LCNO for −1 < α < 1 but α2 6= 1/4
R for α2 = 1/4
LP for α ≥ 1
See the xamples.f file of the sleign2 package for details of appropriate
boundary condition functions.
24. Jacobi/Liouville form of the Jacobi equation. See the files xamples.f and
xamples.tex of the sleign2 package for details.
25. The Meissner equation on (−∞, ∞).
98
1 f or t<0
p = 1, q = 0, w =
9 f or t≥0
LP at −∞ and ∞.
This equation is well known in the applied literature in connection with
the modelling of crystals in one dimension, see [19], [47].
Periodic boundary conditions on (−1/2, 1/2). We have λ0 = 0 and
Note that there are infinitely many simple and infinitely many double periodic
eigenvalues on the interval (−1/2, 1/2).
Semi-Periodic eigenvalues on (−1/2, 1/2).
λ4n = (2nπ + β)2 , λ4n+1 = (2nπ + γ)2 , λ4n+2 = (2(n + 1)π − γ)2 ,
√ √
2 −1 1 + 33 −1 1 − 33
λ4n+3 = (2(n + 1)π − β) , β = cos ( ), γ = cos ( ).
16 16
p = 1, w = 1, q(t) = 1000t
1 2t
p = 1, w = 1, q(t) = e − k et , k ∈ R.
4
LP at −∞ and at ∞.
The essential spectrum starts at 0; for k ≤ 1/2 there are no eigenvalues;
for
h < k − 1/2 ≤ h + 1, h = 0, 1, 2, 3, . . .
there are exactly h + 1 eigenvalues and these are all below the essential spec-
trum i.e. they are all negative. They can be given explicitly by
99
1 k2 − 1 1
p = 1, q(t) = + , w(t) = , k ∈ N.
4 4t2 t
LP at 0 and ∞.
This equation is studied in [49], part II, section 10. The spectrum is discrete
and can be given explicitly by
k+1
λn = n + , n ∈ N0 .
2
5.9 Comments.
Much of Section 5 is based on three papers : Bailey, Everitt, Weidmann and Zettl
[6], Niessen and Zettl [74] and the pre-print of Everitt and Zettl [26].
1. Comments are made separately for each subsection.
2. This treatment of principal and non-principal solutions is based on Niessen
and Zettl [74]. According to Hartman [42] these terms were coined by Leighton
in [64]. The principal solution is the “small” solution but using a term such
as small might lead to confusion since the same solution may be small at one
endpoint but not the other.
3. The structure of singular limit circle boundary conditions is well known, see
[81], [72], [2], [61] but it is not easy to find a clear and comprehensive treatment
of them in the literature. We hope that this paper makes a positive contribu-
tion in this area. The conditions (5.10) characterize all self-adjoint boundary
conditions for the case when each endpoint is either R or LC (either LCO or
LCNO). The canonical form (5.16) and (5.17) represents all separated self-
adjoint BC; (5.20) is the canonical form of all coupled self-adjoint BC, both
for the case R or LC/ R or LC. Clearly (5.16) is a canonical form of (5.8) and
(5.17) represents (5.9). For a different representation of singular self-adjoint
BC see the treatment of Dunford and Schwartz [17].
100
c
(−1)n y (2n) ± y = λ y on 0 < t < 1, c ∈ R, c 6= 0.
t
For the regular case very general results are known, see Niessen and Zettl [73],
Möller and Zettl [71].
5. Möller [68] has shown that, for regular as well as singular SLP, if either p
changes sign on the underlying interval then the spectrum is not bounded
below. This holds even if there is no subinterval of the underlying interval on
which p is negative. The corresponding result for very general higher order
problems was established by Möller and Zettl in [70].
Inequalities (5.19), (5.22), (5.23) can be found in Weidmann [81] for the
regular case with
c
K= 1 , c 6= 0, c ∈ R.
c
For earlier work see Jörgens [49], Rellich [77]. These were extended to the
singular case by Niessen and Zettl in [73], then further extended to more
general K by Bailey, Everitt and Zettl [8], and finally Eastham established
the general case [18].
In [8] there is also the characterization of the eigenvalues given by (5.27)
for the case when the expression M is symmetric and w > 0. It was only
during the writing of these notes that the author realized that the proof given
in [8] still holds, in the regular case, for complex valued p, q, w.
Fore an extension of the asymptotic formula (5.15) to the case when p or w
are allowed to change sign see Atkinson and Mingarelli [5]. Additional terms
in this formula can be obtained if the coefficients satisfy some smoothness
assumtions, see Harris [40], Hartman [42], and the references therein.
6. This Subsection is based on the unfinished paper of Everitt and Zettl [26]
which in turn is based, to a considerable extent, on Bailey, Everitt, Weidmann
and Zettl [6].
7. This Subsection is also based on [26].
8. These examples were taken from the sleign2 files: xamples.f, xamples.tex.
See Subsection 5.3 for instructions on how to download these files from the
internet.
101
1. The LP/LC dichotomy. See the monograph by Kauffman, Read and Zettl [53]
for a brief introduction to LP and LC criteria. There are many sufficient con-
ditions known for LC and also for LP and even some necessary and sufficient
conditions but no necessary and sufficient conditions which can be checked in
each case. Finding such conditions is still an open problem. One obstacle to
finding such conditions is that for the LP case to hold it is enough to give
conditions on a sequence of intervals, with almost no requirements on the co-
efficients outside these intervals, whereas essentially pointwise conditions are
required for the LC case. For an entirely different approach see Zettl [85] for
a construction of all LC expressions.
2. The O/NO alternative. See Kauffman, Read and Zettl [53] for a brief intro-
duction to O/NO criteria, also see the monograph of Havorsen and Mingarelli
[38] and its references. Many conditions, both necessary and sufficient, are
known but there are no known necessary and suffcient conditions which can
be checked in each case. As in the LC/LP situation the most general sufficient
conditions for O are of “interval” type, see e.g. Kwong and Zettl [62], but
such conditions are not appropriate for the NO case to hold. For a completely
different approach i.e. a constuction of all disconjugate (NO) equations see
Zettl [84].
3. Absolutely continuous spectrum. See the contributions to these Proceedings
from Last and Jitomirskaya and from Stolz. We have only considered the
simplest division of the spectrum into its discrete and essential (continuous)
parts. See the seminal paper of Gilbert and Pearson [36] for criteria involving
subordinate solutions - an extension of principal solutions - for the absolutely
continuous spectrum; see also Hinton and Shaw [46], [44], [45], Gesztesy, Gu-
rarie, Holden, Klaus, Sadun, Simon, and Vogl [33] and the references therein.
There is an extensive literature on spectral properties of Sturm-Liouville op-
erators by Simon and his 100 co-authors.
4. There is also a vast literature on so called left definite problems, when the
weight function w is allowed to change sign. See the monograph by Mingarelli,
and its references. Here the operator theory is studied in the setting of Krein
and Pontryagen spaces rather than Hilbert space. For oscillatory properties
of the eigenfunctions when the weight function changes sign or is identically
zero on subintervals, see Everitt, Kwong and Zettl, [22].
5. Another popular topic we have not discussed is the two, or multi, parameter
theory. See Volkmer [80] and the papers by Binding and Sleeman.
6. Multi interval theory. See the contribution of Everitt, Shubin, Stolz and Zettl
to these Proceedings and the references therein for an introduction to SLP
problems on infinitely many intervals. Also see Geszteszy and Kirsch [34],
[35], [32]. If an equation has an interior singularity then, in general, the
solutions cannot be continuously moved through this singularity. One can
study this problem on two separate intervals each of wich has this singular
point on the boundary. Take the direct sum of two self-adjoint operators from
the two intervals and you have a two-interval self-adjoint operator which is
not particularly interesting because it doesn’t “connect” the two intervals
together. More interesting operators are obtained by connecting solutions
102
through the singular point, even if they blow up there, in such a way as to get
a new “two interval” self-adjoint operator. Actually this construction is also
of interest when the interior point is not singular but regular, obtaining what
is known as “point interactions”. For an early contribution see Zettl [82].
7. Discreteness criteria. See the contribution of Read to these Proceedings and
the references therin. Also see [63].
8. Inverse spectral theory. See the landmark paper of Gelfand and Levitan
[31], the elegant exposition of Pöschel and Trubowitz [76] and the references
therein.
References
[1] A.B. Abramowitz and I.A. Stegun. Handbook of Mathematical Functions with Formulas and
Mathematical Tables. Dover Publications Inc., New York, 1970.
[2] N. I. Akhiezer and I. M. Glazman. Theory of linear operators in Hilbert space, Volumes I
and II. Pitman and Scottish Academic Press, London and Edinburgh, 1980.
[3] F. V. Atkinson. Discrete and continuous boundary value problems. Academic Press, New
York /London, 1964.
[4] F. V. Atkinson, W. N. Everitt, and A. Zettl. Regularisation of a Sturm-Liouville problem with
an interior singularity using quasi-derivatives. Differential and Integral Equations 1 (1988),
213-221.
[5] F. V. Atkinson and A. B. Mingarelli. Asymptotics of the number of zeros and of the eigen-
values of general weighted Sturm-Liouville problems. J. reine angew. Math. 375/376 (1987),
380-393.
[6] P. B. Bailey, W. N. Everitt, J. Weidmann, and A. Zettl. Regular approximations of singular
Sturm-L iouville problems. Results in Mathematics, v. 23 (1993), 3-22.
[7] P. B. Bailey, W. N. Everitt, and A. Zettl. Computing eigenvalues of singular Sturm-Liouville
problems. Results in Mathematics, v. 20 (1991), 391-423.
[8] P. B. Bailey, W. N. Everitt, and A. Zettl. Regular and singular Sturm-Liouville problems
with coupled boundary conditions. Proc. Royal Soc. of Edinburgh (A) 126 (1996), 505-514.
[9] P. B. Bailey, W. N. Everitt, and A. Zettl. Sleign2: An eigenfunction - eigenvalue code for
singular Sturm-L iouville problems. pre-print.
[10] J. V. Baxley. Eigenvalues of singular differential operators by finite difference methods I. J.
Math. Anal. Appl. 37 (1972), 244-254.
[11] P. R. Beesack. Gronwall inequalities. Carleton University, Dept. of Mathematics, Carleton
Mathematical Lecture Notes, no. 11, Ottawa, 1975.
[12] A. Bielecki. Une remarque sur la methode de Banach-Cacciopoli-Tikhonov dans la therie des
equations differentielles ordinaires. Bull. Acad. Polon. Sci. Cl. III 4(1956), 261-264.
[13] J. P. Boyd. Sturm-Liouville eigenvalue problems with an interior pole. J. Math. Physics 22
(1981), 1575-1590.
[14] E. A. Coddington and N. Levinson. Theory of ordinary differential equations. McGraw-Hill,
New York /London/Toronto, 1955.
[15] M. Dauge and B. Helffer. Eigenvalues V ariation. I. Neumann Problem for Sturm-Liouville
Operators. J. Diff. Equations, v. 104 (1993), 243-262.
[16] J. Dieudonne. Foundations of Modern Analysis. Academic Press, New York/London, 1969.
[17] N. Dunford and J. T. Schwartz. Linear operators , vol.II. Wiley, New York, 1963.
[18] M. S. P. Eastham. Eigenvalue inequalities for coupled Sturm-Liouville problems. private com-
munication.
[19] M. S. P. Eastham. Limit-circle differential expressions of the second-order with an oscillating
coefficient. Quart. J. Math. Oxford (2) 24 (1973).
[20] W. Eberhard and G. Freiling. Stone-regulare Eigenwertprobleme. Math. Z. 160(1978), 139-
161.
103
[21] W. N. Everitt, J. Gunson, and A. Zettl. Some comments on Sturm-Liouville eigenvalue prob-
lems with interior singularities. Zeitschrift fuer Angewandte Mathematik und Physik (ZAMP)
v. 38 (1987),813-838.
[22] W. N. Everitt, M. K. Kwong, and A. Zettl. Oscillation on eigenfunctions of weighted regular
Sturm-Liouville problems. J. London Math. Soc. 27(1983), 106-120.
[23] W. N. Everitt, M. Moeller, and A. Zettl. Discontinuous dependence of the n-th Sturm-
Liouville eigenvalue. to appear.
[24] W. N. Everitt and F. Neuman. A concept of adjointness and symmetry of ordinary differential
expressions based on the generalized Lagrange identity and Green’s formula. Lecture Note in
Math. 1032 , Springer Verlag, Berlin (1983), 161-169.
[25] W. N. Everitt and D. Race. On necessary and sufficient conditions for the existence of
Caratheodory solutions of ordinary differential equations. Quaestiones Math. 3, no.2, (1976),
507-512.
[26] W. N. Everitt and A. Zettl. Behavior of eigenvalues of regular Sturm-Liouville problems near
a singular boundary. pre-print.
[27] G. Fichera. Numerical and quantitative analysis. Pitman Press, London, 1978.
[28] K. O. Friedrichs. Spectraltheorie halbbeschraenkter Operatoren und Anwendungen auf die
Spektralzerlegung von Differentialoperatoren. I,II. Math. Ann.109 (1933/34), 465-487. Cor-
rections, ibid. 110 (1934/35), 777-779.
[29] K. O. Friedrichs. Ueber die ausgezeichnete Randbedingung in der Spektraltheorie der
halbbeschraenkten gewoehnlichen Differentialoperatoren zweiter Ordnung. Math. Ann. 112
(1935/36), 1-23.
[30] C.T. Fulton and S. Pruess. Mathematical software for Sturm-Liouville problems. ACM. Trans.
Math. Software 19 (1993), 360-376.
[31] I. M. Gelfand and B. M. Levitan. On the determination of a differential equation from its
spectral function. English transl., AMS transl. (2), 1 (1955), 253-304.
[32] F. Gesztesy. On the one-demensional Coulomb Hamiltonian. J. Physics A 13 (1980), 867-
875.
[33] F. Gesztesy, D. Gurarie, H. Holden, M. Klaus, L. Sadun, B. Simon, and P. Vogel. Trapping
and cascading of Eigenvalues in the Large Coupling Limit. CMP 118 (88), 597-634.
[34] F. Gesztesy and W. Kirsch. One-dimensional Schroedinger operators with interactions on a
discrete set. J. fuer Mathematik 362 (1985), 28-50.
[35] F. Gesztesy, C. Macedo, and L. Streit. An exactly solvable periodic Schroedinger operator.
J. Phys. A: Math. Gen. 18 (1985), 503-507.
[36] D. J. Gilbert and D. B. Pearson. On subordinacy and analysis of the spectrum of one-
dimensional Schroedinger operators. J. Math. Anal. Appl. 128 (1987), 130-56.
[37] S. G. Halvorsen. A function-theoretic property of solution of the equation x” + (l w-q)x=0.
Quart. J. Math. Oxford (2), 38 (1987), 73-76.
[38] S. G. Halvorsen and A. B. Mingarelli. Non-Oscillation Domains of Differential Equations
with Two Parameters. Lecture Notes in Mathematics no 1338, Springer-Verlag, Berlin, New
York, 1988.
[39] B. J. Harris. Eigenvalue problems for second-order differential equations. Proc. 1986-87 Fo-
cused Research Program on Spectral Theory and Boundry Value Problems, ANL 87-26.
[40] B. J. Harris. On the asymptotic properties of linear differential equations. Mathematika 34
(1987), 187-198.
[41] P. Hartman. On the number of L2-solutions of x” + q(t)x =l x. Amer. J. Math. 73 (1951),
635-645.
[42] P. Hartman. Ordinary differential equations. Wiley and Sons, Inc., New York /Lon-
don/Sydney, 1964.
[43] E. Hille. Lectures on ordinary differential equations. Addison-Wesley, London, 1969.
[44] D. B. Hinton and J. K. Shaw. Absolutely continuous spectra of perturbed periodic Hamil-
tonian systems. Rocky Mountain J. Math. 17, no.4, (1987), 727-748.
[45] D. B. Hinton and J. K. Shaw. Absolutely continuous spectra of second-order differential
operators with long and short range potentials. SIAM J. Math. Anal. 17 (1986), 182-196.
[46] D. B. Hinton and J. K. Shaw. On the absolutely continuous spectrum of the perturbed Hill’s
equations. Proc. London Math. Soc. (3) 50 (1985), 175-192.
[47] H. Hochstadt. A special Hill’s equation with discontinuous coefficients. Amer. Math. Monthly,
70 (1963), 18-26.
104
[48] M. S. Homer. Boundry value problems for the Laplace tidal wave equation. Proc. Royal Soc.
London Ser. A 428 (1990), 157-180.
[49] K. Joergens. Spectral theory of second order ordinary differential operators. Lectures delivered
at Aarhus Universitet 1962/63, Matimatisk Institut Aarhus 1964.
[50] H. Kalf. A characterization of the Friedrichs extension of Sturm-Liouville operators. J. Lon-
don Math. Soc. (2) 17 (1978), 511-521.
[51] H. G. Kaper, M. K.Kwong, and A. Zettl. Characterization of the Friedrichs extensions of
singular Sturm-Liouville expressions. SIAM J. Math. Anal., 7(1986), 772-777.
[52] H. G. Kaper and M. K. Kwong. Asymptotics for the Titchmarsh-Weyl m-coefficient for
integrable potentials. Proc. Royal Soc. Edinburgh 103A (1986), 347-358.
[53] R. M. Kauffman, T. T. Read, and A. Zettl. The deficiency index problem for powers of
ordinary differential expressions. Lecture Notes in Mathematics 621, Springer-Verlag.
[54] Q. Kong, H. Wu, and A. Zettl. Dependence of eigenvalues on the problem. Mathematische
Nachrichten, (to appear).
[55] Q. Kong and A. Zettl. Dependence of eigenvalues of Sturm-Liouville problems on the bound-
ary. J. Differential Equations, vol.126,no.2, (1996), 389-407.
[56] Q. Kong and A. Zettl. The derivative of the exponential map of matrices. pre-print.
[57] Q. Kong and A. Zettl. Eigenvalues of regular Sturm-Liouville problems. J. Differential Equa-
tions, vol. 131, no.1, (1996), 1-19.
[58] Q. Kong and A. Zettl. Interval oscillation conditions for difference equations. SIAM J. Math.
Analysis, 26, no.4 (1995), 1047-1060.
[59] Q. Kong and A. Zettl. Linear ordinary differential equations. WSSIAA v.3, (1994), (Special
volume dedicated to W.Walter), Inequalities and Applications, edited by R. P. Agarwal, 381-
397.
[60] A. M. Krall. Boundary value problems for an eigenvalue problem with a singular potential.
J. Diff. Equations, 45 (1982), 128-138.
[61] A. M. Krall and A. Zettl. Singular self-adjoint Sturm-Liouville problems. Differential and
Integral Equations v. 1 (1988), 423-432.
[62] M. K. Kwong and A. Zettl. Asymptotically constant functions and second-order linear oscil-
lation. J. of Math. Anal. and Applications, 93 (1983), 475-494.
[63] M. K. Kwong and A. Zettl. Discreteness conditions for the spectrum of ordinary differential
operators. J. Diff. Equations v. 40 n. 1 (1981), 53-70.
[64] W. Leighton. Principal quadratic functionals. Trans. Amer. Math. Soc. 67 (1949), 253-274.
[65] R. J. Lohner. Verified solution of eigenvalue problems in ordinary differential equations. pri-
vate communication.
[66] M. Marletta. Numerical tests of the SLEIGN software for Sturm-Liouville problems. ACM
TOMS, v17 (1991), 501-503.
[67] R. Mennicken and M. Moeller. Nonselfadjoint boundary eigenvalue problems. preprint, Lon-
don, 1997.
[68] M. Moeller. On the unboundedness below of the Sturm-Liouville operator. Differential and
Integral Equations, (to appear).
[69] M. Moeller and A. Zettl. Differentiable dependence of eigenvalues of operators in Banach
spaces. J. Operator Theory, (to appear).
[70] M. Moeller and A. Zettl. Semi-boundedness of ordinary differential operators. J. Differential
Equations, v.115, no.1, (1995), 24-49.
[71] M. Moeller and A. Zettl. Symmetric differential operators and their Friedrichs extension. J.
Differential Equations, v.115, no.1, (1995), 50-69.
[72] M. A. Naimark. Linear Differential Operators. Ungar, New York, 1968.
[73] H.-D. Niessen and A. Zettl. The Friedrichs extension of regular ordinary differential operators.
Proc. Roy. Soc. Edinburgh, 114A, (1990), 229-236.
[74] H.-D. Niessen and A. Zettl. Singular Sturm-Liouville problems: The Friedrichs extension and
comparison of eigenvalues. Proc. London Math. Soc. v.64, (1992), 545-578.
[75] I. G. Petrovski. Ordinary differential equations. Prentice Hall, Inc., London, 1966.
[76] J. Poeschel and E. Trubowitz. Inverse Spectral Theory. Academic Press, New York /Lon-
don/Sydney, 1987.
[77] F. Rellich. Halbbeschraenkte gewoehnliche Diffferentialoperatoren zweiter Ordnung. Math.
Ann. 122 (1950/51), 343-368.
105