The Radially Vibrating Spherical Quantum Billiard
The Radially Vibrating Spherical Quantum Billiard
The Radially Vibrating Spherical Quantum Billiard
r
X
i
v
:
n
l
i
n
/
0
1
0
7
0
5
5
v
1
[
n
l
i
n
.
C
D
]
2
4
J
u
l
2
0
0
1
The Radially Vibrating Spherical Quantum
Billiard
Mason A. Porter and Richard L. Libo
Center for Applied Mathematics
and
Schools of Electrical Engineering and Applied Physics
Cornell University
July, 2000
Abstract
We consider the radially vibrating spherical quantum billiard as a represen-
tative example of vibrating quantum billiards. We derive necessary conditions
for quantum chaos in d-term superposition states. These conditions are sym-
metry relations corresponding to the relative quantum numbers of eigenstates
considered pairwise. In this discussion, we give special attention to eigenstates
with null angular momentum (for which the aforementioned conditions are au-
tomatically satised). When these necessary conditions are met, we observe
numerically that there always exist parameter values for which the billiard
behaves chaotically. We focus our numerical studies on the ground and rst
excited states of the radially vibrating spherical quantum billiard with null an-
gular momentum eigenstates. We observe chaotic behavior in this conguration
and thereby dispel the common belief that one must pass to the semiclassical
( 0) or high quantum number limits in order to meaningfully discuss quan-
tum chaos. The results in the present paper are also of practical import, as the
radially vibrating spherical quantum billiard may be used as a model for the
quantum dot nanostructure, the Fermi accelerating sphere, and intra-nuclear
particle behavior.
1 Introduction
There has been considerable research in the last twenty years that seeks to marry
quantum mechanics and dynamical systems theory into a coherent whole.[4, 6] In
particular, the concept of quantum chaos extends the notions of classical Hamil-
1
tonian chaos to the quantum regime. There are three types of quantum chaotic
behavior: quantized chaos (quantum chaology), semiquantum chaos, and
genuine quantum chaos. Quantum chaology concerns the quantum structure
of classically chaotic systems, semiquantum chaos refers to the chaotic dynamics
of coupled classical and quantum systems, and genuine quantum chaos refers
to chaotic dynamics of fully quantum systems.[3] No example of the third type
of quantum chaos has been established, so the existence of such systems is an
open question.
In the present paper, we discuss semiquantum chaos in the context of vi-
brating billiard systems. We review the recent results of Libo and Porter[9]
and discuss them in further detail. We treat a two-term Galerkin projection
(superposition state) of the radially vibrating sphere and prove that only 2-
term superpositions whose normal modes have common rotational symmetry
behave chaotically. We extend this theorem to arbitrary superposition states
by applying it pairwise. In the proof of this theorem, we establish integrable
(non-chaotic) behavior by showing that the evolution equations reduce to a
two-dimensional autonomous dynamical system, whose non-chaotic properties
are known.[12] We then discuss two examples: one integrable superposition and
one chaotic one. We compute Poincare maps and thereby reveal chaotic char-
acteristics such as regions of ergodicity and KAM islands.
We also discuss the present results with respect to the phenomenology of
quantum chaos. The chaotic behavior in the radius a and conjugate momentum
P corresponds to classical Hamiltonian chaos. The normal modes
nlm
depend
on the radius, so they exhibit quantum-mechanical wave chaos. We also observe
chaos in the Bloch variables (x, y, z), which correspond to quantum-mechanical
probabilities. The dynamical equations describing the present system corre-
spond to a two degree-of-freedom (dof) Hamiltonian system, where one degree-
of-freedom is classical (corresponding to the so-called one degree-of-vibration
(dov) radial motion) and one is quantum-mechanical (corresponding to the cou-
pling coecient ). By coupling a single classical dof (which is necessarily
integrable) with a single quantum dof (which must also be integrable), we ob-
tain a genuinely chaotic system that provides an example of semiquantum chaos
since it consists of a classical system coupled to a quantum one. We remark that
we do not need to pass to the semiclassical ( 0) or high quantum num-
ber limits in order to observe chaos, as is commonly considered requisite for a
meaningful analysis of quantum chaos.[6]
The radially vibrating spherical quantum billiard has several practical appli-
cations that complement its theoretical import. The most important one is that
it may be used as a model for the quantum dot nanostructure.[10] At low tem-
peratures, this microdevice component experiences vibrations due to zero-point
motions, and at higher temperatures, it exhibits vibrations due to natural uc-
tuations. Another application is that the radially vibrating spherical quantum
billiard generalizes Fermis bouncing-ball model of cosmic ray acceleration.[1]
Additionally, the radially vibrating spherical quantum billiard models the in-
tradynamics of the nucleus, as the liquid drop and collective models of the
nucleus include boundary vibrations. Consequently, the importance of the ra-
2
dially vibrating spherical quantum billiard lies not only in its expansion of the
theory of quantum chaos but also in its applicability to problems in nuclear,
atomic, and mesoscopic physics.
2 Statement of the Problem
The spherical quantum billiard addresses the quantum dynamics of a particle
of mass m
0
conned to the interior of a spherical cavity of mass M m
0
with
smooth walls of radius a. The radius vibrates in an a priori unspecied manner,
so that a a(t). A two-component superposition state (Galerkin projection) of
this quantum billiard is given using Dirac notation by
[(r, , , t; a(t))` = A
1
(t)[nlm, t` + A
2
(t)[n
, t`, (1)
where A
1
(t) and A
2
(t) are complex amplitudes. The numbers n, l, m are, re-
spectively, the principal, orbital, and azimuthal quantum numbers. The eigen-
states of the present system are products of spherical Bessel functions and spher-
ical harmonics.[8] In coordinate representation,
'r[nlm, t` =
nlm
(r, , , t; a(t)) =
2
a(t)
3
1
j
l+1
(x
ln
)
j
l
rx
ln
a(t)
Y
lm
(, ),
(2)
where x
ln
is the nth zero of j
l
, the spherical Bessel function of order l.
For the system at hand, the time-dependent Schrodinger equation is given
by
i
t
= K, r a(t), (3)
where the quantum-mechanical Hamiltonian K, the kinetic energy of the parti-
cle, is
K =
2
2m
0
2
. (4)
The total Hamiltonian of the system is
H =
P
2
2M
+ V + K, (5)
where P is the momentum of the billiard boundary, and V V (a) is the
potential of the billiard surface. The potential energy V and kinetic energy
P
2
/2M of the billiard walls are classical quantities, and the conned particle
is quantum-mechanical. For this semiquantum system, we utilize the Born-
Oppenheimer approximation[2], so that only the particle kinetic energy K is
inserted into the Schrodinger equation (3). In this adiabatic approximation,
3
which is commonly used in mesoscopic physics, we are ignoring the eects of
Berry phase.[13]
Taking the expectation of (3) using the superposition state (1) gives
2
2m
0
=
1
a
2
1
[A
1
[
2
+
2
[A
2
[
2
K(A
1
, A
2
, a),
i
= i
A
1
A
1
+
A
2
A
2
+
11
[A
1
[
2
+
12
A
1
A
2
+
21
A
2
A
1
+
22
[A
2
[
2
,
(6)
where the energies of the two terms are given by
1
2
x
2
ln
2m
0
,
2
2
x
2
l
2m
0
. (7)
3 Integrable Conguration
Examining the superposition of [100` and [110` using (6) and orthogonality of
spherical harmonics shows that
11
=
12
=
21
=
22
= 0. (Note that
11
and
22
vanish no matter which eigenstates one considers.) Equating the inner
products (6) of both sides of the Schrodinger equation gives equations of motion
for the complex amplitudes:
i
A
1
=
1
a
2
1
A
1
, i
A
2
=
1
a
2
2
A
2
, (8)
which are integrated to yield
A
1
(t) = C
1
exp
i
1
a
2
(t)dt
, A
2
(t) = C
2
exp
i
1
a
2
(t)dt
.
(9)
From (9), one obtains a Hamiltonian in the radius a and conjugate momentum
P:
H =
P
2
2M
+ K(A
1
, A
2
, a) + V (a) =
P
2
2M
+
1
a
2
1
[C
1
[
2
+
2
[C
2
[
2
+ V (a). (10)
A Hamiltonian with no explicit time-dependence and one dof corresponds to a
two-dimensional autonomous system, which is known to be non-chaotic.[5, 12]
When there are no coupling terms, the degree-of-freedom of the resulting Hamil-
tonian corresponds to the degree-of-vibration of the quantum billiard, which is a
measure of the number of distance dimensions that undergo oscillations. When
a two-term superposition has a non-vanishing coupling coecient, the number
of degrees-of-freedom of the resulting Hamiltonian system is equal to the num-
ber of dov of the billiard plus one. In particular, this means that a superposition
state of a quantum billiard with more than one dov (such as the two dov rect-
angular quantum billiard) is expected to behave chaotically even if every one of
its coupling coecients vanishes.
4
Hamiltons equations for the present integrable conguration are
a =
P
M
H
P
,
P =
V
a
+
a
3
H
a
, (11)
where the energy parameter is given by
2
1
[C
1
[
2
+
1
[C
2
[
2
> 0. (12)
The bifurcation structure of (11) has been studied for quartic potentials V (a).[11]
4 Necessary Conditions for Chaos in k Coupled
States
Consider the superposition
= A
1
q1
+ A
2
q2
+ + A
k
q
k
, (13)
where q
i
(n
i
, l
i
, m
i
) is a vector of quantum numbers. If there does not exist
a pair of normal modes in the k-state superposition (13) with common angular
momentum quantum numbers (i.e., there is no pair i, i
such that l
i
= l
i
and m
i
= m
i
), then inserting (13) into the Schrodinger equation (3) returns a
diagonal quadratic form
A
1
A
1
+ +
A
k
A
k
=
11
[A
1
[
2
+ +
kk
[A
k
[
2
, (14)
as all the cross terms
ij
A
i
A
j
have vanishing coupling coecients
ij
by orthog-
onality of spherical harmonics with dierent angular momenta. The diagonal
terms in (14) stem from the Laplacian. As above, we obtain the Hamiltonian
H(a, P) =
P
2
2M
+
1
a
2
k
i=1
i
[C
i
[
2
+ V (a), (15)
where the C
i
are constants. The superposition (13) is non-chaotic, because the
Hamiltonian (15) is autonomous with one dof.
We thus conclude that a necessary condition for chaotic behavior in an ar-
bitrary nite superposition state of the radially vibrating spherical quantum
billiard is that at least one pair of normal modes in the eigenfunction expan-
sion have common angular momentum quantum numbers. In particular, by
considering small n
i
and n
i
, we obtain a chaotic superposition for eigenstates
with small energies. This even holds for some superpositions that include the
ground state! In most studies of quantum chaos, one must take the semiclas-
sical ( 0) or high quantum-number limits in order to meaningfully study
quantum chaos.[3, 6] In such studies, termed quantum chaology, one considers
the quantum signatures of classically chaotic systems in these regimes. In the
present system, on the other hand, we obtain genuinely chaotic behavior in a
semiquantum system. This phenomenon is often called semiquantum chaos.[3]
5
5 Chaotic Conguration
As an example of a chaotic conguration of the radially vibrating spherical
quantum billiard, consider the superposition state consisting of the ground and
rst excited states with null angular momentum
[(n, l, m)` = A
1
[100` + A
2
[200`, (16)
which gives the wavefunction
(r, t) = A
1
(t)
1
1
(r, t)e
i
E
1
t
+ A
2
(t)
2
2
(r, t)e
i
E
2
t
, (17)
where
n
(r, t) = j
0
nr
a(t)
, j
0
(x) =
sin(x)
x
,
n
=
2
a
3
2
j
1
(n)
. (18)
The superposition (16) has a coupling coecient
12
= 4/3.
Equating coecients in the quadratic form (6) gives the matrix equation
i
A
n
=
2
k=1
D
nk
A
k
, (19)
where D (D
ij
) is the Hermitian matrix
D =
1
a
2
i
a
a
i
a
a
2
a
2
, (20)
and the energy coecient
j
is given by
j
(j)
2
2m
0
, j 1, 2. (21)
Dening the density matrix[8] by
qn
= A
q
A
n
, introducing (dimensionless)
Bloch variables x
12
+
21
, y i(
21
12
), and z
22
11
, and using
(19), we obtain the following three dierential equations:
x =
0
y
a
2
2Pz
Ma
, y =
0
x
a
2
, z =
2Px
Ma
. (22)
In these equations,
0
(
2
1
)/. Rewriting the kinetic energy K(A
1
, A
2
, a)
in terms of the Bloch variable z gives
K(z, a) =
1
a
2
(
+
+ z
),
1
2
(
2
1
). (23)
Inserting K(z, a) into the Hamiltonian (5) gives Hamiltons equations:
a =
P
M
,
P =
V
a
+
2
a
3
[
+
+
(z x)] . (24)
6
Equations (22) and (24) constitute a set of ve coupled nonlinear ordinary
dierential equations, which can be shown to be equivalent to a two degree-of-
freedom Hamiltonian system. The constants of motion of the present system
are the radius of the Bloch sphere
x
2
+ y
2
+ z
2
[A
1
[
2
+[A
2
[
2
= 1 (25)
and the energy (total Hamiltonian)
H =
P
2
2M
+ V (a) + K(z, a). (26)
The equilibria of equations (22, 24) satisfy x = y = 0, z = 1, P = 0, and
a = a
, where a
), (27)
where the subscript of a
a
2
. (29)
With this choice of potential, equation (27) becomes
a a
0
=
a
2
0
V
0
a
3
. (30)
The solutions of (30) for the equilibrium radii a
correspond to the
values.
One computes that a
+
a
a
0
, from which it follows that a
a(0) a
+
,
so a(t) remains bounded in the interval [a
, a
+
].[9]
Now consider the superposition of the rst k null angular momentum eigen-
states,
(k)
(r) =
k
n=1
A
n
(t)
n
n
(r, t). (31)
In order to analyze this conguration, one rst examines the 2-term superposi-
tion
nq
= A
n
(t)
n
n
(r, t) + A
q
(t)
q
q
(r, t), n < q, (32)
7
and then superposes the couplings one obtains from each
nq
as n and q run
from 1 to k in order to obtain dynamical equations for the amplitudes A
i
. One
computes the coupling coecients
nq
to be
nq
= 2
qn
a(t)(n + q)(q n)
, n = q. (33)
The dynamical equations for A
i
are described by a k k matrix and are a
straightforward generalization of (19, 20).
5.1 Numerical simulations
The analysis for 2-term superpositions of null angular momentum eigenstates
follows that for the general case.[9] In the present case, the necessary conditions
for chaotic behavior are satised automatically, because the quantum numbers
m and l vanish for every normal mode under consideration. Consequently, any
k-term superposition (k 2) of null angular momentum eigenstates exhibits
chaotic behavior. We consider numerical simulations for the coupling of the
ground state and rst excited state of a billiard residing in a harmonic potential.
Vibrating Billiards
2.5 0.7
10.9763
-10.5015
a
P
Figure 1: Poincare Section (x = 0) in the (a, P)-plane illustrating that not all
invariant tori are destroyed in the present congruation.
Figure 1 shows a Poincare map in the (a, P)-plane corresponding to x = 0,
and Figure 2 shows a Poincare section projected onto the (x, y)-plane for P = 0.
For each of these two plots, we used the parameter values = 1, M = 10,
m = 1, V
0
/a
2
0
= 5, and a
0
= 1.25. The initial conditions for the two gures
are x(0) = sin(0.95) 0.156434, y(0) = 0, z(0) = cos(0.95) 0.987688,
a(0) 1.6, and P(0) 9.45.
The chaotic behavior of this conguration is evident in both plots, although
there is clearly still some non-chaotic structure present. In the language of
KAM theory, some of the nonresonant tori persist for the present choice of
initial conditions[5, 12]. One may also choose initial conditions corresponding
to a dierent level of persistence of the resonant tori. For example, Figure 3
shows an x = 0 Poincare map in the (a, P)-plane with the same initial conditions
and parameter values as above, except a(0) = 3 and P(0) = 10. Figure 4 shows
a P = 0 Poincare map in the (x, y)-plane for these conditions. There are fewer
invariant tori in these two gures than there are in Figures 12.
8
Vibrating Billiards
1 -1
1
-1
x
y
Figure 2: Poincare Section (P = 0) of the Bloch sphere projected onto the
(x, y)-plane. The structure in this diagram likewise illustrates the survival of
some invariant tori.
6 Phenomenology
In contrast to the present quantum-mechanical context, we note that for the clas-
sical radially vibrating spherical billiard, every orbit with null angular momen-
tum is well-dened and invariant under radial oscillations of the boundary.[7]
Due to conservation of angular momentum, one nds that for the stationary
spherical classical billiard, the enclosed particle sweeps out an annular domain
of constant inner radius.[9] Vibration of the wall of the sphere destroys this
constant, and chaotic motion is expected to develop. In the present quantum-
mechanical context, we note that null angular momentum wavefunctions are
composed only of spherical waves. The nodal surfaces of these wavefunctions
are likewise spherical. Accordingly, the chaotic signature of this conguration
in real space is the sequence of intersections with a xed radius that nodal sur-
faces make at any instant subsequent to a number of transversal times.[9] This
latter condition is consistent with the standard long-time behavior of chaotic
dynamical systems.[6]
In the language of Bl umel and Reinhardt[3], vibrating quantum billiards are
an example of semiquantum chaos. One has a classical system (the walls of the
billiard) coupled to a quantum-mechanical one (the enclosed particle). Consid-
ered individually, each of these subsystems is integrable, as each contributes a
single dof. When they are coupled, however, one observes chaotic behavior in
both of them. The classical variables (a, P) exhibit Hamiltonian chaos, whereas
the quantum subsystem (x, y, z) is truly quantum chaotic. Chaos on the Bloch
sphere is an example of quantum chaos because the Bloch variables (x, y, z) cor-
respond to the quantum probabilities of the wavefunction. Additionally, each
individual normal mode
n
depends on the radius a(t), so each eigenfunction is
an example of quantum-mechanical wave chaos for chaotic congurations of the
billiard. Moreover, because the evolution of the probabilities [A
i
[
2
is chaotic, the
wavefunction in the present conguration is a chaotic combination of chaotic
9
Vibrating Billiards
2.75672 0.6
13.7575
-13.7568
a
P
Figure 3: Poincare Section (x = 0) in the (a, P)-plane for slightly dierent
initial conditions in which fewer invariant tori persist. This is in accord with
KAM theory.
normal modes. Finally, we note that if we quantize the motion of the billiard
walls, we would obtain a higher-dimensional, fully-quantized system that ex-
hibits so-called quantized chaos.[3] In particular, the fully quantized version of
the present system would require passage to the semiclassical limit in order to
observe quantum signatures of classical chaos.
7 Conclusions
We considered the radially vibrating spherical quantum billiard as a represen-
tative example of vibrating quantum billiards. We derived necessary conditions
for quantum chaos in k-term superposition states. We gave special attention to
eigenstates with null angular momentum, for which these conditions are auto-
matically satised. We examined a numerical simulation of the superposition of
the ground and rst null angular momentum excited states. We observed chaotic
behavior in this conguration, thereby dispelling the common belief that one is
required to pass to the semiclassical ( 0) or high quantum number limits
in order to meaningfully study quantum chaos.
8 Acknowledgements
We express our gratitude to the conference organizers (especially Joshua Du)
for their excellent work. We also thank the speakers for the wonderful talks they
gave and the academic committee for their role in the organization process.
References
10
Vibrating Billiards
1 -1
1
-1
x
y
Figure 4: Poincare Section (P = 0) of the Bloch sphere projected onto the
(x, y)-plane. The initial conditions in this plot are the same as those in Figure
3.
[1] R. Badrinarayanan and J. V. Jose. Spectral properties of a Fermi acceler-
ating disk. Physica D, 83:129, 1995.
[2] R. Bl umel and B. Esser. Quantum chaos in the Born-Oppenheimer approx-
imation. Physical Review Letters, 72(23):36583661, 1994.
[3] R. Bl umel and W. P. Reinhardt. Chaos in Atomic Physics. Cambridge
University Press, Cambridge, England, 1997.
[4] G. Casati (Ed.). Chaotic Behavior in Quantum Systems. Plenum, New
York, NY, 1985.
[5] John Guckenheimer and Philip Holmes. Nonlinear Oscillations, Dynam-
ical Systems, and Bifurcations of Vector Fields. Number 42 in Applied
Mathematical Sciences. Springer-Verlag, New York, NY, 1983.
[6] Martin C. Gutzwiller. Chaos in Classical and Quantum Mechanics. Num-
ber 1 in Interdisciplinary Applied Mathematics. Springer-Verlag, New York,
NY, 1990.
[7] Richard L. Libo. Bohr correspondence principle for large quantum num-
bers. Foundations of Physics, 5(2):271293, 1975.
[8] Richard L. Libo. Introductory Quantum Mechanics. Addison-Wesley, San
Francisco, CA, 3rd edition, 1998.
[9] Richard L. Libo and Mason A. Porter. Quantum chaos for the radially
vibrating spherical billiard. Chaos, 10(2):366370, 2000.
[10] J Lucan. Quantum Dots. Springer, New York, NY, 1998.
11
[11] Mason A. Porter and Richard L. Libo. Bifurcations in one degree-of-
vibration quantum billiards. International Journal of Bifurcation and
Chaos, 11(4):903911, April 2001.
[12] Stephen Wiggins. Introduction to Applied Nonlinear Dynamical Systems
and Chaos. Number 2 in Texts in Applied Mathematics. Springer-Verlag,
New York, NY, 1990.
[13] Josef W. Zwanziger, M. Koenig, and A. Pines. Berrys phase. Annual
Reviews of Physical Chemistry, 41:601646, 1990.
12