Thesis MSC
Thesis MSC
A Thesis in
The Department of Mathematics
and Statistics
August 2008
Alexandra
c Lemus Rodrı́guez 2008
Abstract
Spectral Estimates for Schrödinger Operators.
Alexandra Lemus Rodrı́guez.
In quantum mechanics, one of the most studied problems is that of solving the
Schrödinger equation to find its discrete spectrum. This problem cannot always be
solved in an exact form, and so comes the need of approximations. This thesis is
based on the theory of the Schrödinger operators and Sturm-Liouville problems. We
use the Rayleigh-Ritz variational method (mix-max theory) to find eigenvalues for
these operators.
The variational analysis we present in this thesis relies on the sine-basis, which
we obtain from the solutions of the particle-in-a-box problem. Using this basis we
approximate the eigenvalues of a variety of potentials using computational implemen-
tations. The potentials studied here include problems such as the harmonic oscillator
in d dimensions, the quartic anharmonic oscillator, the hydrogen atom, a confined
hydrogenic system, and a highly singular potential. When possible the results are
compared either with those obtained in exact form or results from the literature.
iii
Acknowledgements
iv
To my mother and father,
v
Contents
List of Tables ix
Introduction 1
1 Schrödinger Operators 3
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Operator theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Variational characterization of the spectrum of an
operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Schrödinger operators . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 The dimension d . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2 Sturm-Liouville Problems 17
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Sturm-Liouville theory . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.1 Operator form of SLP . . . . . . . . . . . . . . . . . . . . . . 20
2.2.2 Eigenvalues and Eigenfunctions of the SLP . . . . . . . . . . . 21
2.3 Schrödinger normal form . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.1 The Liouville transformation . . . . . . . . . . . . . . . . . . . 23
vi
2.3.2 A numerical approach to solving the SLP . . . . . . . . . . . . 25
2.4 The particle-in-a-box . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4 Results 38
4.1 Implementation of the method . . . . . . . . . . . . . . . . . . . . . . 38
4.2 Dimension d = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.2.1 The harmonic oscillator . . . . . . . . . . . . . . . . . . . . . 41
4.2.2 The potential V (x) = |x| . . . . . . . . . . . . . . . . . . . . . 44
4.2.3 Polynomial potential: V (x) = x4 . . . . . . . . . . . . . . . . 46
4.2.4 Polynomial potential: V (x) = x2 + x4 . . . . . . . . . . . . . . 47
4.3 Higher dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.3.1 The harmonic oscillator . . . . . . . . . . . . . . . . . . . . . 49
4.3.2 The hydrogen atom . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3.3 Confined hydrogenic atoms . . . . . . . . . . . . . . . . . . . . 53
4.3.4 Highly-singular potentials . . . . . . . . . . . . . . . . . . . . 55
Conclusions 57
B Shooting Method 66
vii
List of Figures
viii
List of Tables
ix
Introduction
∂ ~2
i~ Ψ(x, t) = − ∆x Ψ(x, t) + V (x)Ψ(x, t). (1)
∂t 2m
In this equation, ~ ≈ 6.6255 × 10−27 erg sec is known as Planck’s constant, V (x) is
a potential independent of time, and the operator ∆x is the Laplacian corresponding
to the kinetic energy.
One of the main problems in Quantum Mechanics is to determine the solutions of
this equation, in particular, finding the energy levels of a particle. There are many
different approaches to solve it, one is to find exact solutions and the other is to use
methods that will give estimates of the solutions that cannot be found in an exact
manner. The examples in physics that have an exact solution are very few compared
to the ones that do not. This motivates an approach that allows us to approximate
solutions. In this thesis we describe an approach using variational analysis.
In Chapter 1 we review the theory concerning Schrödinger operators. In particular,
its relation with functional analysis, and we present the variational characterization
of the discrete spectrum of a Schrödinger operator, related to the min-max principle.
In Chapter 2 we review the theory of Sturm-Liouville problems. This is of im-
portance because we can transform a number of problems in the Schrödinger normal
form.
1
In Chapter 3 we describe the variational analysis of quantum mechanical problems
and introduce a special variational basis related to the solutions of the particle-in-a-
box problem.
In Chapter 4 we describe the implementation of a program to approximate the
eigenvalues and the eigenfunctions of a problem. This chapter is also devoted to the
results obtained in the application of the variational method to a variety of explicit
problems.
2
Chapter 1
Schrödinger Operators
1.1 Introduction
For the problem of finding the solutions to the Schrödinger equation given in (1), we
can rescale the equation, and given the assumption that Ψ(x, t) = ψ(x)Θ(t), derive a
time independent Schrödinger equation which is represented as follows
In this chapter we will relate this equation to the famous Schrödinger operator.
The results from this part are mainly studied by Griffiths [9], Gustafson and Sigal
[10], Hall et al. [13], Hannabus [14], Kryezig [16], and Reed and Simon [19].
3
is the L2 -space defined as follows
Z
2 d d 2
L (R ) = ψ : R → C | |ψ| < ∞ .
Rd
Z
(ψ, φ) = ψφ dx, with x ∈ Rd , (1.2)
Rd
where ψ represents the complex conjugate of ψ. The above inner product defines the
norm
1
||ψ|| = (ψ, ψ) 2 .
||Aψ||
||A|| = sup = sup ||Aψ|| < ∞.
{ψ∈D(A)} ||ψ|| {ψ∈D(A) | ||ψ||=1}
This result is important because the domain in which an operator is defined can
4
be a very complicated set, with this lemma we can find operators that are extended
to bounded operators and not worry about the domain itself.
BA = 1|Ran(A) , AB = 1|Ran(B) ,
We note that finding the inverse of an operator is equivalent to solving the equation
Aψ = f for all f ∈ Ran(A).
5
for all ψ, φ ∈ D(A).
Theorem (1.1) and the above lemma (1.2) suggest that the class of operators that
we can use is sufficiently wide.
for all ψ ∈ D(A), ψ 6= 0. Similarly we may define non-negative, negative and non-
positive operators.
Theorem 1.2. [10] If A is a self-adjoint operator, then for any z ∈ C with Im(z) 6= 0,
the operator A − zI has a bounded inverse, and this inverse satisfies
The above theorems are important concerning the domains of linear operators.
6
1.2.1 Variational characterization of the spectrum of an
operator
ρ(A) = C \ σ(A)
We note that for λ ∈ ρ(A), the operator (A − λ)−1 , called the resolvent of A, is
well-defined.
isolated meaning that some neighborhood of λ is disjoint from the rest of σ(A).
7
where ψ ∈ D(A) is called the Rayleigh quotient.
Let us note that if ||ψ|| = 1, the Rayleigh quotient becomes (ψ, Aψ).
Let T be a self-adjoint operator on H as previously defined. We can apply vari-
ational techniques to derive a characterization of the spectrum of T , that is, to find
its eigenvalues.
Theorem 1.4. [10] Given T as above with spectrum σ(T ) ⊂ [a, ∞), then
(ψ, T ψ)
≥a
(ψ, ψ)
The above theorem states that the Rayleigh quotient is bounded below.
Theorem 1.5. [10] Let S(ψ) = (ψ, T ψ) for ψ ∈ D(T ) with ||ψ|| = 1. Then
inf σ(T ) = inf S. Moreover, λ = inf σ(T ) is an eigenvalue of T if and only if there is
a minimizer ψ ∈ D(T ) for S(ψ) such that ||ψ|| = 1.
The above result leads to the Ritz variational principle, given that for any ψ ∈
D(T ),
(ψ, T ψ) ≥ λ = inf σ(T )
and equality holds if and only if T ψ = λψ, this particular ψ is the corresponding
eigenfunction, for all other ψ ∈ H, all values of (ψ, T ψ) are upper bounds to the
operator’s eigenvalues.
Theorem 1.6. (Min-max principle) [10] The operator T has at least n eigenvalues
(counting multiplicities) less than inf σess (T ), if and only if λn < inf σess (T ) where
8
Theorem 1.7. (The Rayleigh-Ritz theorem) [19] Let T be a semibounded
self-adjoint operator. let Dn ⊂ D(T ) be an n-dimensional subspace, and let P be the
orthogonal projection onto Dn . Let TDn = P T P . Let λ
b1 , λ
b2 , . . . , λ
bn be the eigenvalues
b1 ≤ λ
of TDn Dn , ordered by λ b2 ≤ . . . ≤ λ
bn . Then
λi (T ) ≤ λ
bi , i = 1, . . . , n.
λi ≤ λ
bi , i = 1, . . . , min(k, n).
It is important to point out that the min-max principle can be used to find up-
per estimates for the eigenvalues of operators, thus having both quantitative and
qualitative consequences.
In practice, we simply analyze T in an N -dimensional linear space
DN = span{φ1 , . . . , φN }, where {φi }N
i is an orthonormal basis for DN . The eigen-
values of the N × N matrix [(φi , T φj )] then provide upper bounds to the first N
eigenvalues of T . Very often, as N is increased, these upper approximations steadily
improve. We shall use this technique in later chapters.
Let H = L2 (Rd ) with inner product (ψ, φ) as defined in (1.2), at the beginning of this
chapter. We can write equation (1.1) as a linear operator.
H = −∆ + V (1.4)
9
is called a Schrödinger operator; where −∆ is the Laplacian in d dimensions, and
V : Rd → R a given potential corresponding to a quantum mechanical problem.
Theorem 1.8. [10] Let V (x) be a continuous function on Rd satisfying V (x) ≥ 0 and
V (x) → ∞ as |x| → ∞, then
2. σess (H) = [0, ∞), which means that H can have only negative isolated eigenval-
ues, possibly accumulating at 0.
The Dirichlet boundary conditions are given by ψ|∂Λ = 0, which means that
ψ = 0 outside Λ. We shall refer to such a Sturm-Liouville problem by the term
“particle-in-a-box”.
10
The previous results give us very valuable information in the study of the solutions
of different problems of quantum mechanics corresponding to various potentials. The
eigenvectors associated with the discrete spectrum of H are called bound states, while
the points in the discrete spectrum are called bound state energies, or energy levels.
We can use the variational characterization of the spectrum described in the pre-
vious section to analyze a series of problems.
The most common quantum mechanical problems are usually posed in dimension
d = 1, 2, 3.
If the dimension is d = 1 the two parts of the Schrödinger operator become
∂2
∆= ∂x2
, and V : R → R.
When d > 1 the structure of the Laplacian becomes more complicated. In order
to work in higher dimensions, we transform the problem from cartesian coordinates
into a more appropriate system. Let x ∈ Rd be x = (x1 , . . . , xd ), we can transform
this vector into another vector ρ = (r, θ1 , . . . , θd−1 ), where r = ||x||. Then the we
assume that the wave function is now given by
with ψ(r) being the spherically symmetric factor, and Y` the spherical harmonic
factor, where ` = 0, 1, 2, . . .. The derivation of this change of variables can be found
in Sommerfeld [22].
The above analysis is particularly suitable for central potentials given as functions
of r = ||x|| rather than as functions of the individual components of x ∈ Rd . In these
cases we denote the potential function as V (r).
Given a spherically symmetric potential V (r) in a d-dimensional space, if we
11
change the system of coordinates described above, and remove the spherical harmonic
factor, as shown in Hall, et al. [13], we get the following radial Schrödinger equation
d2 ψ d − 1 dψ `(` + d − 2)
− − + ψ + V (r)ψ = Eψ. (1.6)
dr2 r dr r2
A correspondence can now be made with a problem on the half line in one dimen-
sion, we define the radial wave function
d2 R
− + U R = ER, (1.7)
dr2
(2` + d − 1)(2` + d − 3)
U (r) = V (r) + . (1.8)
4r2
Given the above analysis is now clear how one can work with central potentials in
dimensions d > 1. We also note that the above equation (1.7) is in the form of a
Sturm-Liouville problem, but on the semi-infinite interval [0, ∞). We shall return to
this in chapter 2.
1.3.2 Examples
Clearly, the family of potentials which have exact analytical solutions is very small
compared to the family of those that do not. We now exhibit a series of examples that
have exact solutions. These examples are of utmost importance because they provide
us with exactly soluble test problems for our variational methods: if the method gives
12
good results for these, we can apply it to others with some degree of confidence.
d2
− ψ + (ωχ)2 ψ = Eχ ψ,
dχ2
d2
1
− 2 ψ + ωs4 x2 ψ = Eχ ψ,
s2 dx
1
ω 4 Ex ψ = Eχ ψ
Therefore we need only to solve the following differential equation defined by the
Schrödinger operator
d2
− ψ + x2 ψ = Eψ (1.9)
dx2
with solutions
x2
ψn (x) = Hn (x)e− 2 , (1.10)
where Hn is the Hermite polynomial of order n. The energy levels are given by
En = 2n + 1, n = 0, 1, 2, . . . (1.11)
13
Figure 1.1: Harmonic oscillator
We can also find the solutions for this problem in dimension d = 3, studied in
Griffiths [9] and Greiner [8], where the symmetric potential is V (r) = r2 , and the
equation needed to solve is the following
d2 R
− + U R = ER, (1.12)
dr2
where ` = 0, 1, 2, . . ., and n = 1, 2, 3, . . ..
Using algebra we can extend to the problem for dimensions d ≥ 2. This is done
comparing the effective potentials, the one for dimension d = 3 given by (1.13) and
(2`0 + d − 1)(2`0 + d − 3)
U (r) = r2 + . (1.15)
4r2
for dimension d.
14
We set the equation
0 d−1 0 d−3
`(` + 1) = ` + ` +
2 2
d 3
` = `0 + − ,
2 2
and substituting back into the values of the energy (1.14), thus obtaining
Hydrogen-like atoms. The electron in the hydrogen atom is bound by the Coulumb
potential. The hydrogen atom is modeled as an infinitely heavy and fixed proton with
an electron revolving around it. The potential that defines this problem is given by
e 2
V (x) = − |r| with r = ||x|| for x ∈ Rd .
If d = 1 the problem of the hydrogen atom is as interesting as it is troublesome,
given its nature, the singularity splits the space in two pieces acting as a barrier.
The solutions to this problem are not trivial and require a thorough analysis of the
geometry of the problem as studied by [2].
Thus, the hydrogen atom is usually studied in 3 dimensions, [9], [8]. The discrete
spectrum of this problem in d = 3 is given by the energy levels
e2
En` = − , (1.17)
4(n + `)2
where ` = 0, 1, 2, . . ., and n = 1, 2, 3, . . ..
We can also scale and extend this problem to the d-dimensional case using the
15
Figure 1.2: Hydrogen atom
arguments and algebraic analysis as in the above example. In this case we get the
following energy levels
e2
En`d = − , (1.18)
d 3 2
4 n+`+ 2
− 2
The above examples are discussed as they will serve as test problems in the following
chapters. For more insight on the solutions for this problems see literature from
Greiner [8], Griffiths [9], Gustafson and Sigal [10], and Hannabus [14].
16
Chapter 2
Sturm-Liouville Problems
2.1 Introduction
Sturm-Liouville systems arise in a large number of physical problems. They are one-
dimensional models of oscillating systems, and therefore we can also relate the notion
of energy to these particular problems.
The general form of a classical Sturm-Liouville problem is a second-order real
differential equation given by
d du
− p(x) + q(x)u = λw(x)u (2.1)
dx dx
d2 u
− + Q(t)u = λu, (2.2)
dt2
17
where Q(t) = q(x)/w. In fact, equation (2.2) is known as the Liouville or Schrödinger
normal form. Even more, by means of the Liouville transformation that we will de-
scribe in the section 2.3, the general Sturm-Liouville problem (2.1) can be transformed
into this normal form.
From now on we will refer to these kind of problems as Sturm-Liouville problems
or SLP.
2. p, q and w are defined on the closed interval [a, b] and are continuous (maybe
except for a finite number of discontinuities), with p and w strictly positive.
are imposed at the end points of the interval, where a1 , a2 , b1 and b2 are real
numbers different from zero.
18
2.2 Sturm-Liouville theory
Let the functions g(t) and h(t) be continuous functions in the open interval α < t < β.
Then, there exists one and only one function y(t) satisfying the differential equation
(2.5) and the prescribed initial conditions (2.6) on the entire interval α < t < β. In
particular, any solution y = y(t) of (2.5) which satisfies y(t0 ) = 0 and y 0 (t0 ) = 0 at
the time t = t0 must be identically zero.
Although this theorem refers to the homogeneous case only, it can be extended to
the non-homogeneous case, [3] and [5].
From the above theorem if we let p, q and w be as we defined in the previous
section on the interval [a, b], the Sturm-Liouville equation (2.4) has unique solutions
satisfying the initial conditions
19
for c ∈ [a, b].
In the case of the Sturm-Liouville problems local solutions will not suffice: we need
global solutions that satisfy boundary conditions, not only initial solutions. This is
not a trivial problem, and further analysis is required as the solutions are related to
the eigenvalues and the eigenfunctions.
We can relate the Sturm-Liouville theory to the operator theory from chapter 1.
Because (2.1) is a linear equation, we can define the differential operator
1 d d
L= − p(x) + q(x) (2.7)
w(x) dx dx
on the interval a < x < b. The domain of this operator is contained in the set of
admissible solutions for the Sturm-Liouville problems, which are the square integrable
functions with respect to the weight function w
Z b
2 2
L ([a, b]) = u : [a, b] → C | |u(x)| w(x)dx < ∞ .
a
Z b
(u, v) = uvwdx, (2.8)
a
20
Another fundamental property of this operator is that it is self-adjoint and sym-
metric with respect to the weight function w(x), implying that with the given inner
product (2.8) the following equality holds true
Lemma 2.1. (Green’s Identity) [18] Let u, v ∈ D(L) be twice differentiable func-
tions defined on [a, b], then
Z b
b
(Luv − uLv) wdx = [puv 0 − pu0 v]a .
a
The above lemma is useful to solve this boundary condition problems. In partic-
ular, Green’s identity relies on the fact that L is self-adjoint, and if u and v satisfy
the boundary conditions in (2.3), we have that [puv 0 − pu0 v]ba = 0.
There exists a different and equivalent approach to the operator theory of the
Sturm-Liouville problems using the operator form
d d
L=− p(x) + q(x), (2.9)
dx dx
Lu = λu (2.10)
21
Property 2.1. [18]
Let u be as above and the functions p, q and w that define the operator L in (2.7) be
piecewise continuous on [a, b], with p, w > 0, then:
1. The eigenvalues λk are simple, this is, there do not exist two linearly independent
eigenfunctions with the same value.
(f,uk )
where ck = (uk ,uk )
.
If we take p = w = 1 all the above theory is equivalent to the theory for the
Schrödinger operator in dimension d = 1. We note that the normal Liouville form
22
(2.2) is equivalent to the one-dimensional Schrödinger equation that we studied in
chapter 1.
For dimension d = 1 we can rewrite equation (1.1) as
d2
− ψ + V (x)ψ = Eψ (2.11)
dx2
where ψ, V (x) and E represent the wave function, potential, and level of energy of
the system respectively.
As mentioned before, there are a number general Sturm-Liouville problems that
can be transformed into the Liouville normal form; we demonstrate this in the fol-
lowing subsection.
Let x = x(t), this makes (2.1) transform into a new Sturm-Liouville differential
equation that is of the form
d p(x(t)) du
− + q(x(t))x0 u = λw(x(t))x0 u, (2.12)
dt x0 dt
where we assume that the mapping x(t) : (α, β) → (a, b) is onto and that x0 has the
same sign over the interval α < t < β.
23
Dependent variable transformation.
d d
− p mv + qmv = λwmv,
dx dx
which is not in self-adjoint form, though we can multiply by m both sides of the above
expression to obtain
pm2 v 0
0
+ − (pm0 ) 0 m + qm2 v = λwm2 v,
(2.13)
Liouville’s transformation.
d dv
− P + Qv = λW v, (2.14)
dt dt
with
pm2
P = 0 ,
x
Q = (−(pm0 )0 + qm)mx0 ,
W = wm2 x0 .
A1 v(α) = A2 (P v 0 )(α),
(2.15)
0
B1 v(β) = B2 (P v )(β).
24
where
Furthermore, if p, q and w are such that w/p and q/w are defined, then the general
SLP in (2.1) is finally converted to the Liouville normal form or Schrödinger form
d2 v
− + Iv = λv, (2.16)
dt2
Z r
w
t= dx,
p
(2.17)
− 14
m = (pw) .
For some symmetric problems that can be transformed to the Sturm-Liouville form,
it is possible to develop what is called a shooting method to calculate its eigenvalues
and eigenfunctions.
Given the problem (2.16) we can solve it numerically defining it as an initial value
problem satisfying the left hand boundary conditions over a specific range, depending
on the nature of the problem. The differential equation is then solved for a sequence of
trial values of the eigenvalue λ. The eigenvalue is adjusted until the required number
of zeros is obtained and the right hand boundary conditions are met. We refer to
Theorem (2.2).
25
Though the process is quite straightforward it is not always easy to implement such
a numerical program. But it opens the door to have another option to approximate
the eigenvalues which we can later compare to the original method studied in this
work, that is the variational technique. For an example see Appendix B.
The infinite square well potential, also known as the one-dimensional particle-in-a-
box problem, is an important in quantum mechanics. The potential that defines the
Schrödinger operator is the following:
0, if 0 ≤ x ≤ b;
V (x) =
∞, otherwise.
The Schrödinger operator H defines the following equation which is also a regular
Sturm-Liouville problem, with ψ defined on the interval 0 ≤ x ≤ b and boundary
conditions ψ(0) = ψ(b) = 0, and with
d2
− ψ = Eψ. (2.18)
dx2
26
Solving this equation we find that for n = 1, 2, . . . we have the following energy levels
or eigenvalues
nπ 2
En = , (2.19)
b
r
2 nπx
ψn (x) = sin . (2.20)
b b
We can verify the importance of theorem (2.2) with this example. The energy
levels form an increasing sequence, and as these levels go up, each successive wave
function or state has one more node. The eigenfunctions form a complete orthonor-
mal set: from Fourier analysis we know that each ψ ∈ L2 ([0, b]) can be written
ψ = ∞ 2 2
P
n=1 cn ψn . We also note that L ([0, b]) ⊂ L (R), this basis might be useful
27
Chapter 3
3.1 Introduction
E0 ≤ (ψ, Hψ) ,
28
where E0 is the ground-state energy. This means that by finding a suitable trial
function we can obtain an upper bound of the ground state energy [9].
Even more, if we choose ψ as a function of x and other parameters, we can
manipulate these parameters in order to obtain a better approximation, that is to
say, a lower upper bound. If we know the exact solution of a particular problem we
can compare results and study the accuracy of the variational estimate.
Based on the Rayleigh-Ritz principle and theorem (1.6), we can develop a variational
method to approximate the energy levels of a problem.
Given a specific problem with a potential V (x), where the corresponding Schrödinger
operator H = −∆ + V is self-adjoint and bounded below, and its domain
D(H) ⊂ H = L2 (Rd ), we consider of the following steps leading to approximate
energy levels.
N
X
ψ= ci φi (3.1)
i=1
where ci = (ψ, φi ) with i = 1, 2, . . . are the Fourier coefficients of the series expansion
of ψ. If the variational basis is infinite we can substitute ∞ for N .
29
2. Optimization. We suppose that ψ is the trial wave function solution to the
problem posed by H, we can then write ψ as a series of the variational basis B as in
(3.1).
The variational problem now becomes that of minimizing the energies or eigen-
values of H with respect to the free parameter or parameters of the above functions.
This is equivalent to calculating the eigenvalues of the Hamiltonian matrix H and
minimizing them over the same parameter set. The Hamiltonian matrix is defined as
(φ1 , Hφ1 ) (φ1 , Hφ2 ) . . . (φ1 , HφN )
(φ2 , Hφ1 ) (φ2 , Hφ2 ) . . . (φ2 , HφN )
H= , (3.2)
.. .. ... .
..
. .
(φN , Hφ1 ) (φN , Hφ2 ) . . . (φN , HφN )
and, as a consequence of the linearity of H, we can separate this matrix into the sum
of two parts, corresponding to the kinetic and potential energies. Thus
H = K + P, (3.3)
with
(φ1 , −∆φ1 ) (φ1 , −∆φ2 ) . . . (φ1 , −∆φN )
(φ2 , −∆φ1 ) (φ2 , −∆φ2 ) . . . (φ2 , −∆φN )
K= ,
.. .. ... .
..
. .
(φN , −∆φ1 ) (φN , −∆φ2 ) . . . (φN , −∆φN )
(3.4)
(φ1 , V φ1 ) (φ1 , V φ2 ) . . . (φ1 , V φN )
(φ2 , V φ1 ) (φ2 , V φ2 ) . . . (φ2 , V φN )
P= .
.. .. ... ..
. . .
(φN , V φ1 ) (φN , V φ2 ) . . . (φN , V φN )
30
We note that the kinetic component of the matrix H is the same regardless of the
potential for any given problem, this is helpful in a numerical application of the
method.
By the variational principle, we know that the eigenvalues of H will be upper
bounds to the eigenvalues of the operator H. That is, if the eigenvalues of H are
given by ε1 ≤ ε2 ≤ . . ., and the energy levels or eigenvalues of H are given by
E1 ≤ E2 ≤ . . ., then the following hold true
E1 ≤ ε1 ,
E2 ≤ ε2 ,
..
.
N
X
ψk = cki φi . (3.5)
i=1
This means that the components of the eigenvector are the Fourier coefficients of the
series expansion, [11], [15].
31
basis. Then the dimension of the matrix H is truncated and denoted by the letter
N . The numerical accuracy of the method depends greatly on how large this value
is. Secondly, we depend on a computer program to solve the eigenvalue problem for
H. And last, the optimization over the parameters can be very complicated as well.
The implementation of this method will be discussed in chapter 4 through an
example.
√
ψn (χ) = 2 sin (nπχ) ,
√
x 1
ψn (x) = 2 sin nπ + , (3.6)
2L 2
32
Figure 3.1: Shifted particle-in-a-box
Let
1 x 1
φn (x) = √ sin (n + 1)π + (3.7)
L 2L 2
be the functions that define the variational basis B = {φ1 , φ2 , . . .}. This basis has
the following properties that we require:
1. B is a complete set, in the sense that a function in the same Hilbert space
L2 ([−L, L]) ⊂ L2 Rd can be expressed as a linear combination of B.
3. The functions are even or odd with respect to the origin depending the index.
φi is even if i = 0, 2, . . ., or it is odd if i = 1, 3, . . ..
Then, with the above basis we can implement a program to calculate the eigen-
values of the matrix H as in (3.2). In this case, the calculation of the matrices H, K,
and P depends on the members of B.
The implementation of such a program, however, can be simpler if we analyze the
properties of V (x) first. We can have three different choices to build up the matrix
33
H thanks to the above basis, for any suitable potential, for even potentials, and for
polynomial potentials. The cases where the potentials are either even functions or
polynomials present interesting opportunities when it comes to the implementation
of a computer algorithm to attack the problem.
If the potential is even, in order to construct the H matrix, we can split it up in
two different matrices, the even part and the odd part, this is thanks to the property
of the functions in the basis B of being alternately even and odd: when we multiply
by the potential the functions in the basis will retain their even or odd property.
If the potential is a polynomial function, we have a very interesting result.
Property 3.1. Let Px and Px2 be the matrices corresponding to the potentials
V1 (x) = x and V2 (x) = x2 as defined in (3.4) respectively, then
and that
2 2
(φ1 , x φ1 ) (φ1 , x φ2 ) . . .
P x2 = (φ
2 , x 2
φ1 ) (φ2 , x 2
φ2 ) . . . ,
.. .. ...
. .
P∞
we know that (P2x )ij = (φi , xφr ) (φr , xφj ), and that (Px2 )ij = (φi , x2 φj )
r=1
As B is a complete basis for L2 ([−L, L]) we can write the function xφj as the following
34
series
∞
X
xφj = φr (φr , xφj ) (3.9)
r=1
for j = 1, 2, . . ..
If we multiply (3.9) by x we obtain
∞
!
X
x2 φj = x(xφj ) = x φr (φr , xφj )
r=1
∞
X
= xφr (φr , xφj ) .
r=1
∞
!
X
φi , x2 φj =
φi , xφr (φr , xφj ) ,
r=1
∞
X
φi , x2 φj = (φi , xφr (φr , xφj ))
r=1
∞
X
= (φi , xφr ) (φr , xφj ) .
r=1
35
3.4 The basis for the analysis in dimension d ≥ 2
We can use a similar analysis from the previous section in the case of higher di-
mensions, as we explained in chapter 1, in d ≥ 2, the introduction of a set of d − 1
coordinate angles, and the used of a suitable radial transformation, allows us to recast
the residual problem in 1 dimension, namely (1.7) is this one-dimensional problem
for which we shall use the basis functions that are described below.
For d ≥ 2, we shall consider central potentials, and for the analysis of this problem,
we need to work with the effective potential given by (1.8). In fact, the matrix H has
one more component which corresponds to the effective potential.
Let Q(`, d) = 14 (2` + d − 2) (2` + d − 3), and the matrix U defined as follows:
φ1 , r12 φ1 φ1 , r12 φ2
...
1
φ2 , r12 φ2
U=
φ2 , r2 φ1 ...
.. .. ..
. . .
Then
H = K + P + Q(`, d)U. (3.10)
36
B to be exactly the set of solutions from the particle-in-a-box given by
r
2 nπr
φn (r) = sin , (3.11)
b b
with r ∈ [0, b], and n = 1, 2, . . .. That is to say, we use variational trial functions in
L2 ([0, b]) ⊂ L2 ([0, ∞)).
This basis satisfies again the properties mentioned before such as orthonormality
and completeness for L2 ([0, b]). Here the box size b is a variational parameter.
Highly singular potentials. These potentials are functions that have at least one
term of the form 1/rs with s > 2. In these cases we cannot use the above basis
because the inner product (φi , Hφj ) is not defined on [0, b], therefore we will use, in
addition to b, a parameter a > 0 so that the variational basis is in L2 ([a, b]).
As in the case of dimension d = 1 we apply a transformation that shifts the
interval [0, b] to a new interval [a, b] where a, b > 0 are the variational parameters.
r−a
The transformation is given by χ = b−a
, and we obtain the shifted functions
√
r−a
ψn (r) = 2 sin nπ , (3.12)
b−a
r
2 r−a
ψn (r) = sin nπ , (3.13)
b−a b−a
with n = 1, 2, . . ..
37
Chapter 4
Results
In order to use the variational method to approximate the discrete spectrum of the
Schrödinger operators studied in chapter 3, we implemented the following analysis,
using the mathematical software Maple 11.
We will explain the main idea of the implementation by using the example of the
harmonic oscillator in dimension d = 1. The implementation for other problems is
very similar to this, and we shall discuss it at the end of this section.
Given the basis B = {φn }n , with φn defined as in (3.7), we need to calculate the
matrices K and P, corresponding to the problem of the harmonic oscillator with the
potential V (x) = x2 . The components of the matrices are given analytically by the
expressions:
i2 π 2
d2 , if i = j;
4L2
Kij = φi , − 2 φj =
dx 0,
otherwise,
and
L2 (i2 π 2 −6)
, if i = j;
3i2 π 2
Pij = (φi , V (x)φj ) =
4L2 (
4(−1)i+j ij+4ij )
π 2 (i2 −j 2 )2
, otherwise.
38
The Hamiltonian matrix is then given by the sum of the above two matrices, as in
(3.3).
For a numerical analysis the above matrices need to be finite, so the first thing to
do is to choose the dimension N , and then i, j = 1, 2, . . . , N . This means that we can
build numerical matrices of dimension N which depend on a fixed box size L, which
becomes a variational parameter. For each value of L we can obtain the eigenvalues
using the built in functions of the mathematical software, in our case Maple 11. The
problem then arises when we need to find which is the specific L that will give us an
acceptable approximation of the energy levels, in the sense that each eigenvalue we
obtain from the variational analysis is the lowest upper bound to the exact solutions.
In the case of the harmonic oscillator we can compare the results we obtain from this
numerical program with the exact solutions given by (1.11).
For the optimization process over L, we take a simple approach, justified by the
very flat region of a graph of the energy levels against the values of the variational
parameter L, near the minimum, namely, we calculate the eigenvalues of H for a
sequence of values of L. We chose an initial value of L = L0 and a step size s, the
next value of L would be equal to L0 +s, and so on, for a certain number of iterations.
The matrix would depend on the parameter Lk = L0 + ks for k = 0, 1, 2, . . . , M ,
where M is the final number of iterations. We can make an educated guess to choose
a suitable L0 depending on each problem, or choose a large value of s first, to get a
rough idea of an L that would be close to the minimum.
As we already know the exact solutions of the harmonic oscillator, we know that
its wave functions decay swiftly to 0 as x grows. Thus, we assume that the value of
an optimal L will not be too large, but this might not be the case for other problems.
The value of L might also differ from one state to another, that is, the value needed
to optimize the energy of the ground state might be different from the following state,
and so on.
39
When we find an acceptable value of L, the one that minimizes the given eigen-
value, we can also obtain the corresponding eigenvectors of the matrix, and use this
to graph the approximate eigenfunction as a linear combination of the basis, as stated
in (3.5).
The implementation details for different problems depend on the potential and
on the dimension. For an example concerning potentials, the harmonic oscillator
potential is a simple function, and moreover, an even function: in this case the inner
product (φi , V (x)φj ) has an analytical expression that can be exactly calculated by
a program like Maple 11. If the potential is such that we need to use numerical
integration to calculate the inner product, the calculation of the matrix P becomes
slower and possibly less accurate.
On the other hand, we do have the approximation given in (3.8) by the relation
Pxn ≈ Pnx , which allows us to calculate the Hamiltonian matrix of a polynomial
potential such as V (x) = m k
P
k=0 ak x by the computation of only two matrices, K and
Px . Thus we have
m
X
H=K+ ak Pkx . (4.1)
k=0
40
4.2 Dimension d = 1
Table 4.1: Approximation of the energy levels of the harmonic oscillator in dimension d = 1.
n E EV LV EP LP
0 1 1.000000001 6.1 0.999999998 7.3
1 3 3.000000001 5.6 2.999999998 6.5
2 5 5.000000001 5.8 4.999999997 6.6
3 7 7.000000002 5.9 6.999999996 6.6
4 9 9.000000003 6.3 6.999999996 6.6
5 11 11.00000000 6.3 10.99999999 8.3
22 45 45.00000006 8.9 45.00000005 8.9
23 47 47.00000010 9.0 46.32075646 11.7
24 49 49.00000120 8.9 49.00000117 8.9
25 51 51.00000185 9.0 50.32353151 11.2
26 53 53.00001881 8.9 53.00001868 8.9
32 65 65.01513530 8.9 62.32028554 10.1
34 69 69.07941805 8.9 69.07905146 8.9
37 75 75.36273369 9.0 74.35832785 9.3
42 85 89.861473970 8.9 98.14418440 8.9
41
For this case we have also plotted the energy levels E versus the parameter L,
as shown in figure (4.1): we observe that the energy levels as a function of L are
U-shaped and flat near the minima. Also, the values of L we get as minimizers for
each of the states are not that different from each other. From the graph we can
think of choosing only one value of L to obtain a “good” approximation for all the
eigenvalues considered. If N is large enough, say, bigger than 20, any value of L in
the range [6, 10] gives good approximations to the energy levels.
Figure 4.1: Energy vs. Parameter L for the harmonic oscillator in dimension d = 1.
Once the optimal L is determined, the eigenvectors of the first 3 states also ob-
tained and graphed with their corresponding eigenfunctions as defined in (3.5). Fig-
ures (4.2), (4.3), and (4.4), represent the trial wave functions obtained, each one
calculated with the values of L stated in table (4.1).
We note that the approximations obtained by both methods are good. After ex-
perimenting with different values of N we can confirm that the larger this value is the
better the approximations become, that is to say, closer to the exact solution. Also,
by this same observations we can say that up to N/2 states are good approximations
42
Figure 4.2: Trial wave function for the ground state ψ0 of the harmonic oscillator in dimension d = 1 with minimizer
L = 6.1.
Figure 4.3: Trial wave function for the first excited state ψ1 of the harmonic oscillator in dimension d = 1 with
minimizer L = 5.6.
in both approaches, the direct and the polynomial one. However, it is interesting to
note that the polynomial approach, as it involves yet another approximation, from
proposition (3.8), does not always give upper bounds. This is clear for some of the
results showed in table (4.1), for example states n = 25, 32, 37. In contrast we observe
that the variational approach always gives upper bounds even though N is not very
large. To overcome the problem of inaccuracy of the polynomial approach we need
only to increase the value of N to be greater than 100, this will give better results.
43
Figure 4.4: Trial wave function for the second excited state ψ2 of the harmonic oscillator in dimension d = 1 with
minimizer L = 5.8.
The function V (x) = |x| is symmetric and non-differentiable at the origin. This is a
linear potential that represents the attraction of a small particle to a large plate with
a hole in it under Newtonian gravity. The exact analytical solutions to this problem
are given in terms of the zeros of Airy functions and its first derivatives, studied in
Flügge [7] and shown in Abramowitz et al. [1].
44
In table (4.2) we show the results for this problem found by applying the varia-
tional method with a matrix of dimension N = 50, and a step size s = 0.25 for the
optimization process. Again, n represents the state. As there exist suitable expres-
sions for the exact solutions to this problem, we compare the results obtained in the
variational method to these. We can also use a shooting method which is known to
give good approximations for the energy levels. E and ES denote the exact energy lev-
els and those obtained by this shooting method respectively, while EV represents the
approximation to the eigenvalue using an implementation of the variational method,
and L is the minimum value of the variational parameter.
Table 4.2: Approximation of the energy levels of the potential V (x) = |x|.
n E ES EV L
0 1.01879297 1.01879305 1.01879344 6.25
1 2.33810741 2.33810752 2.33810745 8.00
2 3.24819758 3.24819770 3.24819827 8.25
3 4.08794944 4.08794959 4.08794960 9.25
4 4.82009921 4.82009938 4.82010046 9.75
5 5.52055983 5.52056000 5.52056026 10.50
6 6.16330736 6.16330750 6.16330960 10.75
7 6.78670809 6.78670828 6.78670915 11.50
8 7.37217726 7.37277509 7.37218128 12.00
9 7.94413359 7.94413380 7.94413594 12.50
10 8.48848673 8.48848690 8.48849407 12.75
11 9.02265085 9.02265108 9.02265603 13.50
12 9.53544905 9.53544923 9.53546280 13.75
Following the theory from chapter 3, and the observations made in the harmonic
oscillator the energy levels obtained by the variational methods are optimized upper
bounds to the energy levels. In this case, when we optimize over the variational pa-
rameter L, we obtained lower upper bounds, in fact, the goal was to obtain the lowest
upper bound. Here we show the best lower bound calculated using the optimization
process.
For the shooting method, as described in Appendix B, we need to solve an ordinary
differential equation numerically, which depends strongly on the software we are using
45
where we cannot control the accuracy of the results, therefore, the eigenvalues we
obtained by the shooting process are just approximations. We cannot claim how
accurate they are, but this is a matter for further numerical analysis which is not
the main point in this thesis. Though the approximations obtained by the shooting
method are very close to the exact solutions, if we face a problem that has no exact
solutions we can only use this approximations to get an idea about the eigenvalues,
not as a point of comparison to determine how accurate the results obtained from the
variational method are.
We considered the potential V (x) = x4 to test the accuracy of the method using
the polynomial approach, and compared it to the direct approach of the variational
method and the results obtained from the shooting method as well.
We used a matrix of dimension N = 50, and a step size of s = 0.5. Again
the subindexes S, V , and P represent the results obtained with the shooting, the
variational, and the polynomial methods respectively.
From this results we note that in some cases, the eigenvalues obtained from the
variational method are smaller than the ones obtained from the shooting method.
46
This reinforces the notion that the approximations obtained by the shooting method
cannot be compared to the ones obtained by the variational one.
We do observe that for the first states, the approximations obtained from the
variational and polynomial approaches are close. The polynomial approach has the
advantage that it is fast for an optimization process compared to the variational
method, therefore if we choose a large N we can get good approximations using it.
Table (4.4) reflects the results after applying the method to a matrix of dimension
N = 50, and using a step size s = 1 for the optimization process. Again, n represents
the state, and because this kind of potentials do not have exact solutions, yet again
we need to make use of a shooting method to approximate the solutions given by
ES , EV , as usual, represents the approximation to the eigenvalue using the direct
47
approach, with the respect value LV of the optimized variational parameter, and
EP represents the approximation using the polynomial approach, with corresponding
optimized variational parameter LP .
The results in this table confirm what we mentioned for the problem in the previ-
ous section, regarding the accuracy of the three methods used. However, calculating
these energies using the variational approach is computationally expensive, because
in order to construct the matrix H of dimension N we need to integrate numerically
for each component of P for each value of L. This makes us think that the polynomial
approach might be more useful, as it is faster, we need only to calculate the matrix
Px which has an analytical form depending on L, and multiply it by itself according
to the relation (4.1).
One of the problems is that after experimenting with L using this approach,
sometimes we get values for the higher excited states that are lower than the energy
levels of the problem, we show in figure (4.7) the behaviour of the eigenvalues in
function of L.
We can see from this that after L = 10, using the polynomial approach, the
numerical results might be not so reliable, this can be the explanation that this
approach has sometimes good results, but some other times goes even lower than
the shooting or exact results, as seen in previous examples. We also observe that
48
Figure 4.7: Energy vs. L of the quartic anharmonic oscillator in dimension d = 1.
before this specific value of the variational parameter, the eigenvalues have a smooth
behaviour, if we choose an L before this critical value and take a large value of N ,
we then again obtain good results.
The energy values for the harmonic oscillator in dimensions d = 2, 3, 4, 5 and quantum
number ` = 0, 1, 2, 3 are calculated. We used a matrix of dimension N = 40 and a step
size of s = 0.25. This problem has known exact solutions (1.16), which can be used
49
for comparisons. Since the expression of each element of the matrices corresponding
to the kinetic and potential energies is given by an exact analytical function of L, the
calculations of these eigenvalues are quite fast.
We present two different tables of results. For all of them ` represents the quantum
number related to the angular momentum, n is the state (1 plus the number of nodes
in the radial function for a given `), E is the exact value of the energy levels, and EV
the variational approximation with minimal parameter b.
Table 4.5: Approximation of the energy levels of the harmonic oscillator in dimension d = 2.
` n E EV b
0 1 2 2.28622395 4.00
2 6 6.30816202 4.00
3 10 10.32728254 4.50
4 14 14.34185985 5.25
1 1 4 4.00073469 4.25
2 8 8.00185191 4.75
3 12 12.00337390 5.25
4 16 16.00524280 5.50
2 1 6 6.00000262 5.00
2 10 10.00001068 5.50
3 14 14.00002888 6.00
4 18 18.00005965 6.25
1
U (r) = r2 − ,
4r2
and we can see in Figure (4.8). For this case the singular term makes the potential
tend to −∞ when r is close to 0. This shows that the original oscillating behaviour
might be lost and the matrix we get from the variational analysis is not as quite
accurate. This is a hint that we might have problems with singular problems, even
when they are weakly-singular. Here the difficulty is removed for ` > 0 (or d > 2)
50
since the effective potential U (r), generally given by (1.8), has a positive pole at r = 0
and the particle is confined in effectively a “soft box” between this pole and the rising
potential at large r. This point will be discussed further in the conclusion.
Table (4.6) shows a small sample of the results for dimensions d = 3, 4, 5 and
quantum numbers ` = 0, 1, 2, 4. Most of the results are very accurate, except for
some values of `.
51
Table 4.6: Approximation of the energy levels of the harmonic oscillator in dimension d = 3, 4, 5.
d ` n E EV b
3 0 1 3 3.00000000 6.00
5 19 19.00000001 7.00
10 39 39.00000001 8.75
1 1 5 5.00007348 4.50
5 21 21.00167944 6.25
10 41 41.00907276 7.75
2 1 7 7.00000000 6.00
5 23 23.00000001 7.75
10 43 43.00000001 9.25
3 1 9 9.00000001 6.00
5 25 25.00000076 7.25
10 45 45.00002070 8.50
4 0 1 4 4.00073469 4.25
5 20 20.00745550 6.00
10 40 40.02454449 7.50
1 1 6 6.00000262 5.00
5 22 22.00011370 6.50
10 42 42.00094014 8.00
2 1 8 8.00000002 6.00
5 24 24.00000248 7.25
10 44 44.00004592 8.50
3 1 10 10.00000000 6.00
5 26 26.00000008 7.50
10 46 46.00000274 8.75
5 0 1 5 5.00007348 4.50
5 21 21.00167944 6.25
10 41 41.00907276 7.75
1 1 7 7.00000000 6.00
5 23 23.00000001 7.75
10 43 43.00000001 9.25
2 1 9 9.00000001 6.00
5 25 25.00000076 7.25
10 45 45.00002070 8.50
3 1 11 11.00000001 6.00
5 27 27.00000000 8.00
10 47 47.00000001 10.00
The hydrogen atom is studied in section 1.3.2. We now show results obtained with
our method. Here the size of the matrix is N = 25, and the step size used in the
52
optimization process is s = 1, the step size is chosen so large because this weakly
bound system is all together large. The variational method is slow in this case, given
the singular nature of the problem and because for the matrix P (corresponding to
the potential energy) we need to integrate numerically for each term.
Table 4.7: Approximation of the energy levels of the hydrogen atom in dimension d = 3.
` n E EV b
0 1 -0.2500000000 -0.2494292776 13
2 -0.06250000000 -0.06173021569 32
3 -0.02777777778 -0.02682855454 57
4 -0.01562500000 -0.01457726183 90
1 1 -0.06250000000 -0.06231120892 33
2 -0.02777777778 -0.02747649731 60
3 -0.01562500000 -0.01526320869 94
4 -0.01000000000 -0.009656788911 143
2 1 -0.02777777778 -0.02777640178 75
2 -0.01562500000 -0.01561644406 108
3 -0.01000000000 -0.009970374676 146
4 -0.006944444444 -0.006872824074 189
The results we get are not as accurate as the ones obtained for the harmonic
oscillator. The hydrogen atom is a complicated problem; its energy levels are all
negative and they they get closer and closer to each other as n grows. The most
serious difficulty is posed by the weak binding leading to a spread-out wave function,
quite unlike a particle in a box. We can see that in the values of the variational
parameter, each following state needs a larger b, and similarly with each next quantum
number `.
With our variational approach we can think that we are confining the potential we
want to study to a box of specific size, and we want to find the optimal size that
will give us the best approximations to the energy levels. In the case of the hydrogen
atom mentioned in the previous section, we find that we need bigger boxers for each
53
following state, this is because, as already mentioned, the problem is spread out.
But this opens up the possibility to study the behaviour of a system that is already
confined.
The example of an hydrogen atom confined to a spherical box has been studied by
Varshni [23] and by Ciftci, Hall, and Saad [4]. For this sort of problem the Schrödinger
equation is given by:
d2
` (` + 1) A
− 2 ψ(r) + − ψ(r) = Eψ(r), (4.2)
dr r2 r
with boundary conditions ψ(0) = ψ(b) = 0, and A > 0. In the latter study [4], exact
solutions are given for specific radii. We then applied the variational method to this
problem using the radii suggested by the exact solutions and obtained the following
results for A = 1.
Table 4.8: Approximation of the energy levels of a confined hydrogenic atom in dimension d = 3.
` n b E EV N
0 1 4 -0.06250000000 -0.06249993558 200
1 1 12 -0.02777777778 -0.02777772347 150
2 1 24 -0.01562500000 -0.01562500004 150
3 1 40 -0.01000000000 -0.01000000002 100
0 1 3.803847576 -0.02777777778 -0.02777730093 100
2 14.19615242 -0.02777777778 -0.02777473502 100
1 1 11.05572809 -0.01562500000 -0.01562458222 100
2 28.94427191 -0.01562500000 -0.01562321913 100
2 1 21.77124344 -0.01000000000 -0.009999999956 150
2 48.22875656 -0.01000000000 -0.009999999937 150
3 1 36 -0.006944444444 -0.006944444423 150
2 72 -0.006944444444 -0.006944444436 150
Here ` is the quantum number, n is 1 plus the number of nodes of the wave
function, E the exact solution for the energy, and EV the variational approximation
using a matrix of dimension N . In contrast to what happened with the hydrogen
atom, we find that for the confined atom the approximations are comparable. We
also show the graphs for the approximations of the confined wave functions for radii
54
b = 4 and b = 72 in figures (4.9) and (4.10) respectively.
Highly-singular potentials are problems that have been widely studied as it is difficult
to find exact solutions. There are a special kind of potentials that are called quasi
exactly solvable, this means that we can find a part of the energy spectrum exactly
when the potential parameters satisfy specific conditions. For example, Dong et al.
55
[6] studied the potential
V (r) = ar2 + br−4 + cr−6
by assuming that the radial equations for the ground state and first excited state were
given by:
√ √
R` (r) = rκ α + βr2 + γr−2 exp − ar2 + cr−2 /2,
(4.3)
with β = γ = 0, and κ = κ0 for the ground state. Although there are more
studies on this matter, where different wave functions are used to calculate the exact
solutions, and the parameters a, b, and c have to satisfy certain constraints.
We find the the exact solutions for the ground state energy in the cases where
a = b = c = 1 and a = 1, b = c = 9 mentioned in Hall et al. [13].
For a = b = c = 1 we have the potential
56
Conclusions
Throughout the history of physics, scientists have studied how to find solutions to all
sort of mathematical problems posed by physical theories. In this way the explicit
implications of these theories can be revealed, and verification and falsification of the
theories become possible. In this thesis we have studied a small portion of this wide
world of physical problems, namely that related to Schrödinger operators and their
discrete spectra.
The variational method we have adopted relies strongly in the use of the sine-basis
obtained from the solutions of the particle-in-a-box problem, which we have presented
in section 2.4. One of the questions we wanted to answer was, can this basis be applied
effectively to construct a variational analysis for unconfined problems?
From the results shown in chapter 4. we conclude that this basis is appropriate
for the study of problems that have some similar characteristics to the particle-in-a-
box problem. That is why a potential such as the harmonic oscillator, whose wave
functions vanish very quickly with |x|, as though the particle were already in a box,
leads to good results. This points out that in theory, even if we are working with
potentials that are highly singular but strongly U-shaped, then the problem would
be suitable for a variational analysis in the sine-basis.
Other kinds of singularities that do not sufficiently confine the particle might
represent a problem, such as the harmonic oscillator in dimension d = 2 and with
quantum number ` = 0. An effective use of the sine-basis for this case is still an open
57
issue.
We also conclude that the sine-basis does not work too well with loosely-bound
systems such as the hydrogen atom in dimension d = 3. The explanation for this is
that such problems are not effectively confined, and require large boxes to approximate
their energy levels; and the resulting computations proved to be slow.
However, as we found for the hydrogen atom, if the quantum system studied is
already confined in a box, the approximations for the eigenvalues are highly accurate.
The study of confined systems has been of great interest in recent years, since the early
work of Michels [17]. These physical systems are important because of the necessity
of finding information about atoms or molecules trapped in microscopic cavities, or
places in high pressure environments. It appears that this sine-basis gives excellent
results when applied to estimate the spectra of this class of physical problem.
58
Bibliography
[3] M. Braun. Differential Equations and Their Applications. Springer, New York,
1993.
[4] H. Ciftci, R.L. Hall, and N. Saad. Study of a confined Hydrogen-like atom by the
Asymptotic Iteration Method. arXiv:0807.4135v1 [math-ph] 25 July 2008, (IJQC,
In Press).
[6] S.H. Dong, X.W. Hou, and Z.Q. Ma. Schrödinger Equation with the Potential
V (r) = ar2 + br−4 + cr−6 . arXiv:quant-ph/9808037v1 21 Aug 1998.
59
[10] S.J. Gustafson and I.M. Sigal. Mathematical concepts of quantum mechanics.
Springer, New York, 2006.
[11] R. Hall and N. Saad. Variational Analysis for a generalized spiked harmonic
oscillator. J. Phys. A 33 569-578 (2000).
[12] R.L. Hall. A simple eigenvalue formula for the quartic anharmonic oscillator.
Can. J. Phys. A 63, 311 (1985).
[13] R.L. Hall, Q.D. Katatbeh, and N. Saad. A basis for variational calculations in
d dimensions. J. Phys. A Math.Gen. 37 11620 (2004).
[15] P. Kościk and A. Okopińska. The optimized Rayleigh-Ritz scheme for determining
the quantum-mechanical spectrum. J. Phys. A Math.Theor. 40 10851 (2007).
[16] E. Kreyszig. Introductory functional analysis with applications. Wiley, New York,
1989.
60
[22] A. Sommerfeld. Partial differential equations in physics, volume 1. Academic
Press Inc., 1941.
[23] Y.P. Varshni. Critical cage radii for a confined hydrogen atom. J. Phys. B. At.
Mol. Opt. Phys. 31 2849 (1998).
61
Appendix A
62
> Po:= proc (L, n)
> local i, j, k, m, p;
> p := Matrix(n,n);
> for i to n do
> p[i,i]:= evalf(L^2*(i^2*Pi^2-6)/i^2/(3*Pi^2));
> for j from i+1 to n do
> p[i,j] := evalf(4*L^2/Pi^2/(i^4-2*i^2*j^2+j^4)*
(4*(-1)^(i+j)*i*j+4*i*j));
> p[j,i] := p[i,j];
> end do
> end do;
> p;
> end proc;
The Hamiltonian matrix.
> HM:= proc (L, n)
> local HM;
> HM := Ki(L,n)+Po(L,n);
> end proc;
This is the optimization process.
n : = dimension of the Hamiltonian matrix.
L : = initial value of the variational parameter
s : = stepsize so the next variational parameter used in the
process will be L+s
it : = number of iterations
dig : = number of digits in the approximation
flag : = 1 or 2
If flag : = 1 OPT will give a matrix with the first row being
63
that containing the values of the parameters L, and the next
rows will be the corresponding eigenvalues of the Hamiltonian
matrix.
If flag : = 2 OPT will give a matrix which in the first column
will obtain the ordered eigenvalues of the Hamiltonian matrix,
and in the second column we will have their corresponding values
of the optimum parameter.
> OPT:= proc (n, L, s, it,dig,flag)
> local M, P, T, i, j, l,E;
> Digits:=dig;
> l := L;
> #for i from 1 to n+1 do;
> # T[i,1]:=i-2;
> #end do;
> for j to it do
> T[1,j]:=l;
> M := HM(l,n);
> P := evalf(simplify(LinearAlgebra:-Eigenvalues(M)));
> for i to n do
> P[i] := Re(P[i])
> end do;
> P := sort(P);
> for i to n do
> T[i+1,j] := P[i];
> # print(j,l,T[i,j],i-1)
> end do;
> l := l+s;
64
> end do;
> T:=convert(T,array);
> for i from 2 to n+1 do
> E[i-1,1]:=10^10;
> for j from 1 to it do
> if T[i,j] < E[i-1,1] then
> E[i-1,1]:=T[i,j];
> E[i-1,2]:=T[1,j];
> end if;
> end do;
> end do;
> E:=convert(E,array);
> if flag=1 then
> T;
> else
> E;
> end if;
> end proc;
65
Appendix B
Shooting Method
66
that the function corresponding to the particular state to be ng. We then choose
upper and lower values λu and λl respectively,
1. Let λ be the average (λu + λl ) /2. This value will be our trial eigenvalue.
2. Solve the equation
d2 u
− + Q(x)u = λu, (B.1)
dx2
67