Examples of Complete Solvability of 2D Classical Superintegrable Systems
Classical (maximal) superintegrable systems in n dimensions are Hamiltonian systems with 2n−1 independent constants of the motion, globally defined, the maximum number possible. They are very special because they can be solved algebraically. In this paper we show explicitly, mostly through examples...
Збережено в:
Дата: | 2015 |
---|---|
Автори: | , , , |
Формат: | Стаття |
Мова: | English |
Опубліковано: |
Інститут математики НАН України
2015
|
Назва видання: | Symmetry, Integrability and Geometry: Methods and Applications |
Онлайн доступ: | http://dspace.nbuv.gov.ua/handle/123456789/147159 |
Теги: |
Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
|
Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Цитувати: | Examples of Complete Solvability of 2D Classical Superintegrable Systems / Y. Chen, E.G. Kalnins, Q. Li, W. Miller Jr // Symmetry, Integrability and Geometry: Methods and Applications. — 2015. — Т. 11. — Бібліогр.: 42 назв. — англ. |
Репозитарії
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-147159 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1471592019-02-14T01:25:58Z Examples of Complete Solvability of 2D Classical Superintegrable Systems Chen, Y. Kalnins, E.G. Li, Q. Miller Jr., W. Classical (maximal) superintegrable systems in n dimensions are Hamiltonian systems with 2n−1 independent constants of the motion, globally defined, the maximum number possible. They are very special because they can be solved algebraically. In this paper we show explicitly, mostly through examples of 2nd order superintegrable systems in 2 dimensions, how the trajectories can be determined in detail using rather elementary algebraic, geometric and analytic methods applied to the closed quadratic algebra of symmetries of the system, without resorting to separation of variables techniques or trying to integrate Hamilton's equations. We treat a family of 2nd order degenerate systems: oscillator analogies on Darboux, nonzero constant curvature, and flat spaces, related to one another via contractions, and obeying Kepler's laws. Then we treat two 2nd order nondegenerate systems, an analogy of a caged Coulomb problem on the 2-sphere and its contraction to a Euclidean space caged Coulomb problem. In all cases the symmetry algebra structure provides detailed information about the trajectories, some of which are rather complicated. An interesting example is the occurrence of ''metronome orbits'', trajectories confined to an arc rather than a loop, which are indicated clearly from the structure equations but might be overlooked using more traditional methods. We also treat the Post-Winternitz system, an example of a classical 4th order superintegrable system that cannot be solved using separation of variables. Finally we treat a superintegrable system, related to the addition theorem for elliptic functions, whose constants of the motion are only rational in the momenta. It is a system of special interest because its constants of the motion generate a closed polynomial algebra. This paper contains many new results but we have tried to present most of the materials in a fashion that is easily accessible to nonexperts, in order to provide entrée to superintegrablity theory. 2015 Article Examples of Complete Solvability of 2D Classical Superintegrable Systems / Y. Chen, E.G. Kalnins, Q. Li, W. Miller Jr // Symmetry, Integrability and Geometry: Methods and Applications. — 2015. — Т. 11. — Бібліогр.: 42 назв. — англ. 1815-0659 2010 Mathematics Subject Classification: 20C99; 20C35; 22E70 DOI:10.3842/SIGMA.2015.088 http://dspace.nbuv.gov.ua/handle/123456789/147159 en Symmetry, Integrability and Geometry: Methods and Applications Інститут математики НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
description |
Classical (maximal) superintegrable systems in n dimensions are Hamiltonian systems with 2n−1 independent constants of the motion, globally defined, the maximum number possible. They are very special because they can be solved algebraically. In this paper we show explicitly, mostly through examples of 2nd order superintegrable systems in 2 dimensions, how the trajectories can be determined in detail using rather elementary algebraic, geometric and analytic methods applied to the closed quadratic algebra of symmetries of the system, without resorting to separation of variables techniques or trying to integrate Hamilton's equations. We treat a family of 2nd order degenerate systems: oscillator analogies on Darboux, nonzero constant curvature, and flat spaces, related to one another via contractions, and obeying Kepler's laws. Then we treat two 2nd order nondegenerate systems, an analogy of a caged Coulomb problem on the 2-sphere and its contraction to a Euclidean space caged Coulomb problem. In all cases the symmetry algebra structure provides detailed information about the trajectories, some of which are rather complicated. An interesting example is the occurrence of ''metronome orbits'', trajectories confined to an arc rather than a loop, which are indicated clearly from the structure equations but might be overlooked using more traditional methods. We also treat the Post-Winternitz system, an example of a classical 4th order superintegrable system that cannot be solved using separation of variables. Finally we treat a superintegrable system, related to the addition theorem for elliptic functions, whose constants of the motion are only rational in the momenta. It is a system of special interest because its constants of the motion generate a closed polynomial algebra. This paper contains many new results but we have tried to present most of the materials in a fashion that is easily accessible to nonexperts, in order to provide entrée to superintegrablity theory. |
format |
Article |
author |
Chen, Y. Kalnins, E.G. Li, Q. Miller Jr., W. |
spellingShingle |
Chen, Y. Kalnins, E.G. Li, Q. Miller Jr., W. Examples of Complete Solvability of 2D Classical Superintegrable Systems Symmetry, Integrability and Geometry: Methods and Applications |
author_facet |
Chen, Y. Kalnins, E.G. Li, Q. Miller Jr., W. |
author_sort |
Chen, Y. |
title |
Examples of Complete Solvability of 2D Classical Superintegrable Systems |
title_short |
Examples of Complete Solvability of 2D Classical Superintegrable Systems |
title_full |
Examples of Complete Solvability of 2D Classical Superintegrable Systems |
title_fullStr |
Examples of Complete Solvability of 2D Classical Superintegrable Systems |
title_full_unstemmed |
Examples of Complete Solvability of 2D Classical Superintegrable Systems |
title_sort |
examples of complete solvability of 2d classical superintegrable systems |
publisher |
Інститут математики НАН України |
publishDate |
2015 |
url |
http://dspace.nbuv.gov.ua/handle/123456789/147159 |
citation_txt |
Examples of Complete Solvability of 2D Classical Superintegrable Systems / Y. Chen, E.G. Kalnins, Q. Li, W. Miller Jr // Symmetry, Integrability and Geometry: Methods and Applications. — 2015. — Т. 11. — Бібліогр.: 42 назв. — англ. |
series |
Symmetry, Integrability and Geometry: Methods and Applications |
work_keys_str_mv |
AT cheny examplesofcompletesolvabilityof2dclassicalsuperintegrablesystems AT kalninseg examplesofcompletesolvabilityof2dclassicalsuperintegrablesystems AT liq examplesofcompletesolvabilityof2dclassicalsuperintegrablesystems AT millerjrw examplesofcompletesolvabilityof2dclassicalsuperintegrablesystems |
first_indexed |
2025-07-11T01:29:47Z |
last_indexed |
2025-07-11T01:29:47Z |
_version_ |
1837312143321464832 |
fulltext |
Symmetry, Integrability and Geometry: Methods and Applications SIGMA 11 (2015), 088, 51 pages
Examples of Complete Solvability
of 2D Classical Superintegrable Systems?
Yuxuan CHEN †, Ernie G. KALNINS ‡, Qiushi LI † and Willard MILLER Jr. †
† School of Mathematics, University of Minnesota, Minneapolis, Minnesota, 55455, USA
E-mail: yc397@cam.ac.uk, lixx0939@umn.edu, miller@ima.umn.edu
‡ Department of Mathematics, University of Waikato, Hamilton, New Zealand
E-mail: math0236@waikato.ac.nz
Received May 05, 2015, in final form October 27, 2015; Published online November 03, 2015
http://dx.doi.org/10.3842/SIGMA.2015.088
Abstract. Classical (maximal) superintegrable systems in n dimensions are Hamiltonian
systems with 2n − 1 independent constants of the motion, globally defined, the maximum
number possible. They are very special because they can be solved algebraically. In this
paper we show explicitly, mostly through examples of 2nd order superintegrable systems
in 2 dimensions, how the trajectories can be determined in detail using rather elementary
algebraic, geometric and analytic methods applied to the closed quadratic algebra of sym-
metries of the system, without resorting to separation of variables techniques or trying to
integrate Hamilton’s equations. We treat a family of 2nd order degenerate systems: os-
cillator analogies on Darboux, nonzero constant curvature, and flat spaces, related to one
another via contractions, and obeying Kepler’s laws. Then we treat two 2nd order nondege-
nerate systems, an analogy of a caged Coulomb problem on the 2-sphere and its contraction
to a Euclidean space caged Coulomb problem. In all cases the symmetry algebra structure
provides detailed information about the trajectories, some of which are rather complicated.
An interesting example is the occurrence of “metronome orbits”, trajectories confined to an
arc rather than a loop, which are indicated clearly from the structure equations but might be
overlooked using more traditional methods. We also treat the Post–Winternitz system, an
example of a classical 4th order superintegrable system that cannot be solved using separa-
tion of variables. Finally we treat a superintegrable system, related to the addition theorem
for elliptic functions, whose constants of the motion are only rational in the momenta. It is
a system of special interest because its constants of the motion generate a closed polynomial
algebra. This paper contains many new results but we have tried to present most of the
materials in a fashion that is easily accessible to nonexperts, in order to provide entrée to
superintegrablity theory.
Key words: superintegrable systems; classical trajectories
2010 Mathematics Subject Classification: 20C99; 20C35; 22E70
Contents
1 Introduction 2
2 Examples of 2D 2nd degree degenerate systems 4
2.1 The D4(b) oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Contraction to the Poincaré upper half plane . . . . . . . . . . . . . . . . . . . . 9
2.3 Contraction to the D3 oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3.1 Embedding D3 in 3D Minkowski space . . . . . . . . . . . . . . . . . . . . 13
?This paper is a contribution to the Special Issue on Analytical Mechanics and Differential Geometry in honour
of Sergio Benenti. The full collection is available at http://www.emis.de/journals/SIGMA/Benenti.html
mailto:yc397@cam.ac.uk
mailto:lixx0939@umn.edu
mailto:miller@ima.umn.edu
mailto:math0236@waikato.ac.nz
http://dx.doi.org/10.3842/SIGMA.2015.088
http://www.emis.de/journals/SIGMA/Benenti.html
2 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
2.3.2 Contraction of the D3 oscillator to the isotropic oscillator . . . . . . . . . 14
2.4 The Higgs oscillator S3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4.1 Contraction of D4(b) to the Higgs oscillator on the sphere . . . . . . . . . 21
2.4.2 Contraction of the Higgs to the isotropic oscillator . . . . . . . . . . . . . 22
3 Examples of 2D 2nd degree nondegenerate systems 23
3.1 The system S7 on the sphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.1.1 S7 in polar coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1.2 Trajectories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2 System E16 in Euclidean space as a contraction of S7 . . . . . . . . . . . . . . . 28
4 The Post–Winternitz superintegrable systems 36
4.1 The classical system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5 Classical elliptic superintegrable systems 37
5.1 Elliptic superintegrability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.2 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.3 Another version of the example . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
A Review of some basic concepts in Hamiltonian mechanics 46
A.1 Classical integrability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
A.2 Classical superintegrability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
1 Introduction
Classical and quantum mechanical Hamiltonian systems that can be solved explicitly, both
algebraically and analytically, and with adjustable parameters, are relatively rare and highly
prized. Famous classical examples are the anharmonic oscillator (Lissajous patterns) and Kepler
systems (planetary orbits) and the Hohmann transfer for orbital navigation [7]. Famous quantum
examples are the Coulomb system (energy levels of the hydrogen atom, leading to the periodic
table of the elements), and the quantum isotropic oscillator. The solvability of these systems
is related to their symmetry, not necessarily group symmetry. This higher order symmetry is
captured by the concept of superintegrability. A natural classical Hamiltonian system (with the
Hamiltonian as kinetic energy plus potential energy) on an n-dimensional Riemannian space is
said to be (maximally) superintegrable if it admits the maximally possible 2n − 1 functionally
independent constants of the motion globally defined (usually required to be polynomial or at
least rational in the momenta). Similarly a quantum Hamiltonian system H = ∆n+V (where ∆n
is a Laplace–Beltrami operator on an n-dimensional Riemannian manifold and V is a potential
function) is (maximally) superintegrable if it admits 2n − 1 algebraically independent partial
differential operators commuting with H. There is a rapidly growing literature concerning these
systems, e.g., [2, 3, 4, 6, 8, 9, 10, 11, 13, 20, 28, 30, 31, 32, 33, 38, 39, 40, 42]. In this paper we
consider only classical systems and n = 2 so all of our systems admit 3 independent constants
of the motion. By taking Poisson brackets of the classical constants of the motion we generate
a symmetry algebra, not necessarily a Lie algebra, which is never abelian. (This contrasts with
integrable Hamiltonian systems which admit n constants of the motion in involution, so that the
symmetry algebra is always abelian.) It is this nonabelian symmetry algebra and its structure
that is responsible for the solvability of superintegrable systems.
To be more explicit, along a specific classical trajectory each constant of the motion Lj takes
a fixed value `j , so the trajectory can be characterized as the intersection of 2n−1 hypersurfaces
Lj = `j in the 2n-dimensional phase space. Thus in principle the path of the trajectory can be
Examples of Complete Solvability of 2D Classical Superintegrable Systems 3
determined algebraically, though not how it is traced out in time. Since the Poisson bracket of
two constants of the motion is again a constant of the motion, the nonabelian symmetry algebra
gives us relationships between the symmetries. In the quantum case each bound state eigenspace
of the Hamiltonian is invariant under the action of the symmetry algebra, so that a knowledge
of the irreducible representations of the symmetry algebra gives useful information about the
dimensions of the eigenspaces and of the eigenvalues themselves.
In this paper we illustrate the value of superintegrabilty by studying and solving several
families of classical superintegrable systems. Second degree superintegrable systems are those
whose generating symmetries are all polynomials in the momenta of degree ≤ 2. All of these
systems have been classified for 2 dimensions (as have the systems with nondegenerate potentials
in 3 dimensions) [5, 19, 23]. They occur on constant curvature spaces (with 3 Killing vectors,
admitting the most symmetries), on the 4 Darboux spaces (admitting 1 Killing vector) and
6 families of Koenigs spaces (admitting no Killing vectors) [25, 33]. Thus, after the constant
curvature spaces, the Darboux spaces admit the most symmetries. The Darboux metrics an be
written as
D1: ds2 = 4x
(
dx2 + dy2
)
, D2: ds2 =
x2 + 1
x2
(
dx2 + dy2
)
,
D3: ds2 =
ex + 1
e2x
(
dx2 + dy2
)
, D4(b) : ds2 =
2 cos 2x+ b
sin2 2x
(
dx2 + dy2
)
,
An example of a Koenigs space is
ds2 =
(
c1
x2 + y2
+
c2
x2
+
c3
y2
+ c4
)(
dx2 + dy2
)
.
We first treat some analogs of the harmonic oscillator. (A treatment of the Kepler system and
its analog on the 2-sphere from this point of view is contained in Chapter 3 of the article [33].)
We start by studying a system on the Darboux space D4(b) with a 1-parameter potential.
Then by taking limits (contractions [15, 24]) of this system we obtain systems on the Darboux
space D3, on the Poincaré upper half plane, the 2-sphere (the Higgs oscillator [14]), and finally
the isotropic oscillator on Euclidean space.
The second class of examples are nondegenerate (3-parameter potential) systems. The first,
S7 in our listing [23], can be regarded as a caged version of an analog to the Coulomb potential
on the 2-sphere. It contracts to the caged Coulomb system E16 in Euclidean space.
The preceding examples can also be studied analytically via separation of variables. However,
the Post–Winternitz system [36] cannot, see also [29]. It is 4th degree superintegrable but non-
separable. However, we show that the classical orbits can be found exactly via superintegrabilty.
The last examples are different. They are separable in elliptic coordinates and can be derived
via an action-angle construction. The usual action-angle construction of a superintegrable and
separable system requires the addition theorem for trigonometric or hyperbolic functions and
leads to polynomial superintegrability, e.g., [20]. The construction here uses the addition formula
for elliptic functions [35]. It leads to classical systems where some of the constants are nontrivially
rational in the canonical momenta. We present two examples where, however, the systems have
polynomial symmetry algebras, so we consider them as superintegrable and worthy of study.
The new feature here is that the symmetry algebra closes polynomially, even though the system
is only rationally superintegrable. These are the first examples of such behavior known to us.
Again we can determine the trajectories exactly.
With the exception of the elliptic systems, all of the superintegrable systems in this paper
have been derived and classified before; we take the structure equations as given and show
how superintegrability alone leads to formulas for the trajectories. These formulas and their
analysis are new. The elliptic systems have not been found elsewhere to our knowledge, so we
demonstrate the procedure to derive them.
4 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Darboux 4 (D4(b)) oscillator
Higgs oscillatorD3 oscillator Poincaré upper half plane
isotropic oscillator E5 (linear potential)
Figure 1. A diagram showing the contractions of oscillators.
This paper is partly pedagogical (the Introduction and the Appendix) but mostly new re-
search (Sections 2–5). In the Appendix we give a brief review of fundamental definitions and
results from classical Hamiltonian mechanics, adapted from [33], needed as background for our
computations. The point here is to illustrate how, using the structure equations of super-
integrable systems alone, we can derive and classify the trajectories via algebra. Hamilton’s
equations are used only to determine the periods of orbits for systems with degenerate poten-
tial. Separation of variable techniques are not used, except in the last section; our approach
is easy to understand geometrically. Analogously the spectra of the corresponding quantum
systems can be obtained algebraically from the structure relations, though we do not treat this
here. Some recent papers, e.g., [26, 27], adopt a related but different approach by using ladder
operators constructed from the structure algebra to compute trajectories and spectra for sys-
tems with degenerate potential. We know of no prior treatments of the nondegenerate caged
systems S7 and E16, or of the use of the structure algebra to call attention to special orbits, such
as those for which R2 = 0. We point out the contractions that relate our various superintegrable
systems.
2 Examples of 2D 2nd degree degenerate systems
Our first examples are degenerate systems. These have 1-parameter potentials and always admit
a symmetry that is a 1st degree polynomial in the momenta, hence a group symmetry that can
be interpreted as invariance with respect to rotation or translation corresponds to a constant of
the motion which leads to an analog of Kepler’s 2nd Law. The only possibilities in 2 dimensions
are constant curvature and Darboux spaces [22]. In [33] there is an example of an analog of the
Kepler problem on the 2-sphere that satisfies Kepler’s three laws and then by a limiting process
(a contraction) goes to the Kepler system in Euclidean space. Here we will treat an analog of
the harmonic oscillator on the Darboux space D4(b) and show that it obeys analogs of Kepler’s
laws. Then by taking contractions to superintegrable systems on the Darboux space D3, on
the Poincaré upper half plane (equivalent to a system on a hyperboloid in Euclidean space)
and, finally, to the isotropic oscillator system in Euclidean space, we will see that using ideas
from superintegrability theory alone we can understand the basic properties of these systems:
conservations laws, explicit trajectories, etc., and how they are related. Fig. 1 describes the
contraction relationships that we will exploit. An important feature of degenerate systems is
that they always admit 4 linearly independent symmetries that are 2nd degree in the momenta,
whereas only 3 can be functionally independent. Thus there must be a relation between these
symmetries. This relation emerges from the structure algebra obeyed by the symmetries.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 5
2.1 The D4(b) oscillator
We consider the superintegrable system
H =
sinh2(2x)
2 cosh(2x) + b
(
p2
x + p2
y
)
+
α
2 cosh(2x) + b
, (2.1)
with a basis of constants of the motion, H, J = py, and
Y1 = − cos(2y)
(
sinh2(2x)
2 cosh(2x) + b
(
p2
x + p2
y
)
− cosh(2x)p2
y
)
− sin(2y) sinh(2x)pxpy −
α cos(2y)
4 cosh2(x)− 2 + b
, (2.2)
Y2 = − sin(2y)
(
sinh2(2x)
2 cosh(2x) + b
(
p2
x + p2
y
)
− cosh(2x)p2
y
)
+ cos(2y) sinh(2x)pxpy −
α sin(2y)
4 cosh2(x)− 2 + b
, (2.3)
Here the spaces D4 are indexed by a parameter b > −2. (In the limit as b → −2 the space
becomes a 2-sphere and this system becomes the Higgs oscillator [14].) The variable y can
be interpreted as an angle, and the space and potential are periodic in y with period π. The
constants of the motion generate an algebra under the Poisson bracket obeying the structure
equations
{J ,H} = {Y1,H} = {Y2,H} = 0, (2.4)
{J ,Y1} = −2Y2, {J ,Y2} = 2Y1, {Y1,Y2} = 4J 3 + 2bJH − αJ , (2.5)
with Casimir
Y2
1 + Y2
2 = J 4 +H2 + bJ 2H− αJ 2. (2.6)
Note that there are 4 linearly independent symmetries J 2, H, Y1, Y2 as polynomials in the
momenta, but there can be only 3 functionally independent generators. The dependence relation
is given by (2.6).
From the first two equations (2.5) we see that (Y1,Y2) transforms like a 2-vector under
rotations about the 3-axis. The sum Y2
1 + Y2
2 is expressed in terms of constants of motion so
the sum is constant. We set Y2
1 + Y2
2 ≡ κ2 where κ is the length of the 2-vector. Thus we can
choose a preferred coordinate system such that Y1 = κ and Y2 = 0.
To determine a trajectory, we need to express x and y in terms of the constants of the motion
along the trajectory. We do this by eliminating the momenta from equations (2.1)–(2.6). We
see that Y1 and Y2 can be expressed in an alternate form:
Y1 = − cos(2y)H+ cos(2y) cosh(2x)p2
y − sin(2y) sinh(2x)pxpy, (2.7)
Y2 = − sin(2y)H+ sin(2y) cosh(2x)p2
y + cos(2y) sinh(2x)pxpy. (2.8)
To eliminate px, we multiply equation (2.7) by cos(2y), equation (2.8) by sin(2y) respectively,
and add them together, which gives us Y1 cos(2y) + Y2 sin(2y) = −H + cosh(2x)p2
y. Using the
facts that Y1 = κ, Y2 = 0 and J = py, we get the orbit equation,
cosh(2x)J 2 − cos(2y)κ = H. (2.9)
Since cosh(2x) is an even function in x, there will be two orbits for positive x and negative x
that are symmetric with respect to x-axis. Therefore, without loss of generality, we restrict
ourselves to positive x.
6 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Analog of Kepler’s second law of planetary motion. We exploit the fact that there is
a 1st degree constant of the motion. We consider an orbital motion as taking place in the s1−s2
plane with polar coordinates s1 = r cos θ, s2 = r sin θ where r = (2 cosh(2x) + b)/(2 sinh 2x) and
θ = 2y. In these coordinates the metric is
ds2 = 2
1 + 2br2 +
√
16r4 + 4br2 + 1
(1 + 2br2)2
dr2 + r2dθ2.
If α > 0 the potential is attractive to the origin in the plane; if α < 0 the potential is repulsive
and the trajectories are unbounded.
Theorem 2.1. Trajectories of the D4(b) oscillator sweep out equal areas in equal times with
respect to the origin in the s1 − s2 plane.
Proof. In the interval from some initial time 0 to time t the area swept out by the segment of
the straight line connecting the origin and the object is A(t) = 1
2
∫ θ(t)
θ(0) r
2(θ)dθ. Thus the rate at
which the area is swept out is
dA
dt
=
dA
dθ
dθ
dt
=
1
2
r2(θ(t))
dθ
dt
= r2dy
dt
, (2.10)
since dy/dt = (dy/dθ) · (dθ/dt) = (1/2) · (dθ/dt). To get (dy/dt), we take the Poisson bracket
of H and y, which is
{H, y} =
dy
dt
=
2py sinh2(2x)
2 cosh(2x) + b
=
2J sinh2(2x)
2 cosh(2x) + b
,
then plug in this expression for (dy/dt) into equation (2.10), to get, (dA/dt) = J /2 which is
a constant. �
Analog of Kepler’s third law of planetary motion.
Theorem 2.2. For α > 0 the period T of an orbit is
T =
1
2
π
(
2 + b√
(−2H− bH+ α)
+
2− b√
(2H− bH+ α)
)
. (2.11)
Proof. The total area swept out as the trajectory goes through one period is
A(T ) =
1
2
∫ 2π
0
r2(θ)dθ (2.12)
=
1
16
J 2π
(2 + b)
√
−κ2 +H2 + 2J 2H+ J 4 + (2− b)
√
−κ2 +H2 − 2J 2H+ J 4√
(−κ2 +H2 − 2J 2H+ J 4)(−κ2 +H2 + 2J 2H+ J 4)
.
However, from the second law we see that A(T ) = (T/2) · J and, using the fact that κ2 =
J 4 +H2 + bJ 2H− αJ 2, the period can be expressed in the form (2.12). �
Now we begin an analysis of the trajectories. We first assume that α > 0 so that the potential
is attractive.
Restriction on H. Recall there is a restriction when we try to express r in terms of θ and
other constants of the motion. That is, cosh(2x) = (H + cos θκ)/(J 2) should always be larger
than 1 for any θ. Then since κ > 0, we have the restriction H − κ > J 2 implying that for
a closed trajectory, H should always be positive. Here κ is the length of the 2-vector (Y1,Y2).
To be explicit, κ2 = J 4 + H2 + bJ 2H − αJ 2. Then by squaring both sides of an alternate
Examples of Complete Solvability of 2D Classical Superintegrable Systems 7
-2 -1.5 -1 -0.5 0 0.5 1
-1.5
-1
-0.5
0
0.5
1
1.5
S
2
vs S
1
Figure 2. Orbit plot for J = 10, H = 130, κ = 20, b = 0 (equal axes).
form of the restriction equation, H − J 2 > κ, and plugging in the expression for κ2, we get
(α − (b + 2)H)J 2 > 0. Noticing the fact that b + 2, H, and J 2 are all larger than 0, we reach
the conclusion, H < α/(b + 2) for a bounded orbit. For larger values of H the trajectory is
unbounded.
Case of bounded trajectory. To plot the trajectories on the s1− s2 plane it is convenient
to write s1 and s2 both in terms of θ, which is
s1 =
1
2
√
J 2(2H+ 2 cos θ · κ+ bJ 2)
(H+ cos θ · κ)2 − J 4
cos θ,
s2 =
1
2
√
J 2(2H+ 2 cos θ · κ+ bJ 2)
(H+ cos θ · κ)2 − J 4
sin θ. (2.13)
(Note: it is useful to divide the r part of s1 and s2 by J 4 and introduce new constantsH′ = H/J 2
and κ′ = κ/J 2 so we can eliminate one constant.) Since in s1 and s2, J always appears in the
form of J 2, so for every positive J , there will always be a duplicate case for the correspon-
ding −J , and without loss of generality, we can restrict our discussion to J > 0.
A typical plot of the trajectory on the s1-s2 plane is Fig. 2, and as one or more of H, J and b
becomes larger and larger, they will dominate the r term and make r less susceptible to the
change of cos θ. Thus the plots will look more and more circular. Furthermore, if κ becomes very
large, the plots will tend to move towards the negative s1 direction and appear in an elongated
form as in Fig. 3. The plot will always be symmetric with respect to s1-axis, because s1(θ) is
even and s2(θ) is odd in θ. Also, since the plot will always lean towards the negative s1 side,
larger r always occurs at smaller s1 value.
To trace out an orbit in time, we would need a starting point to integrate Hamilton’s equa-
tions. Conventionally, we choose a point that is closest to the origin, which we call perihelion.
From the plots of the orbits and an easy analysis of equations (2.13), it is obvious that this
point is the intersection of the trajectory with the positive s1 axis. The point on the x-y plane
that corresponds to this perihelion is (x0, 0) where x0 = 1
2 arcCosh(κ+H
J 2 ). Plugging y = 0 into
equation (2.8) gives pxpy = 0. Realizing py = J 6= 0, we know px must be equal to 0. We
see that the perihelion points on the phase space trajectory are uniquely determined by the
constants of the motion as follows:
cosh(2x) =
H+ κ
J 2
, y = 0, px = 0, py = J , (2.14)
8 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
-0.04
-0.03
-0.02
-0.01
0
0.01
0.02
0.03
0.04
S
2
vs S
1
Figure 3. Orbit plot for J = 1, H = 1002, κ = 1000, b = 0 (axes are not set equal).
s
1
-50 -25 0 5
s
2
-2
0
2
S
2
vs S
1
Figure 4. s2 vs s1 for J = 10, H = 120, κ = 20, b = 0.
and we can see that if we know the perihelion point on the phase space as in equation (2.14),
a set of constants of the motion can be uniquely determined.
Case of escape velocity. This is the case when H = J 2 + κ, a bifurcation point on
the momentum map. In polar coordinates, as θ → π we have r → ∞, whereas for s1, s2 in
equations (2.13), s1 → −∞ and s2 → 0. A plot for s2 vs. s1 is given in Fig. 4.
In Fig. 4, the two tails will extend to infinity in the direction of negative s1-axis, and they
will get closer and closer to the s1-axis for smaller and smaller s1 value but they will never touch
the s1-axis.
Case of unbounded trajectory. Here, there is more than one value of y, i.e., θ/2 that has
no corresponding real value of x in equation (2.9).
Example 2.3 (H ≤ J 2). In this case, since H ≤ J 2, r will blow up before s1 goes becomes
negative, so the entire trajectory will be bounded on the positive-s1 side of the plane. A boundary
case (H = J 2) is plotted in Fig. 5. For Fig. 5 the constants are J = 10, H = 100, κ = 20,
b = 0, so for 0 ≤ θ < π/2 and 3π/2 < θ ≤ 2π, corresponding real values of x exist and there is
a trajectory. For θ = π/2 and 3π/2, r will goes to infinity and at the same time, s2 will go to
infinity and s1 to 0 which is represented by the two tails in Fig. 5.
Example 2.4 (J 2 ≤ H < J 2 + κ). In this case, the trajectory will extend to the negative-s1
side of the plane. A typical plot for an orbit is given in Fig. 6.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 9
0 0.2 0.4 0.6 0.8 1 1.2
-50
-40
-30
-20
-10
0
10
20
30
40
50
S
2
vs S
1
Figure 5. s2 vs s1 for J = 10, H = 100, κ = 20, b = 0.
-30 -25 -20 -15 -10 -5 0 5
-50
-40
-30
-20
-10
0
10
20
30
40
50
-30 -25 -20 -15 -10 -5 0 5
-50
-40
-30
-20
-10
0
10
20
30
40
50
S
2
vs S
1
Figure 6. s2 vs s1 for J = 10, H = 110, κ = 20, b = 0.
2.2 Contraction to the Poincaré upper half plane
Using the Hamiltonian H, (2.1) modified by a constant as H′ = H− [α/(2 + b)], we let x = εY ,
y = εX, α = −[(2 + b)β]/[2ε2], J = K/ε, Y1 − (2J 2)/(ε2) +H ≈ 2X1, Y2 ≈ (2X2)/(ε), and go
to the limit as ε→ 0. Then
H′ = 4Y 2
b+ 2
(
p2
X + p2
Y
)
+ βY 2, K = pX ,
X1 =
(
X2 − Y 2
)
p2
X − 2XY pXpY −
βX2
4
(b+ 2), X2 = Xp2
X + Y pXpY +
β(b+ 2)
4
X.
The structure equations become
{K,X1} = −2X2, {K,X2} = K2 +
β
4
(b+ 2), {X1,X2} = −2KX1 +
b+ 2
2
KH′,
and the Casimir is
X2 + (
β(b+ 2)
4
+K2)X1 −
b+ 2
4
K2H′ = 0. (2.15)
10 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Circle
X
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Y
0.2
0.4
0.6
0.8
1
Origin
Figure 7. Circular trajectory on Poincaré upper half plane with β = 0, b = 0, H′ = 2, K = 1, X2 = 1.
If the potential is turned off this is essentially the Poincaré upper half plane model of hyperbolic
geometry; for b = 4 it is exactly that. The space here consists of all real points (X,Y ) with
Y > 0. In this limit there is no longer any periodicity. Using these identities to eliminate the
momenta we arrive at the equation for the trajectories:(
4K2 + β(b+ 2)
2
X − 2X2
)2
+K2
(
4K2 + β(b+ 2)
)
Y 2 −H′K2(b+ 2) = 0. (2.16)
Notice that the orbit equation is even in Y , so the trajectory will always be symmetric with
respect to the X-axis. When restricting to the Poincaré upper half plane with Y > 0, the
trajectory will be the upper half of the trajectory traced out by the orbit equation. If β > 0, so
that the potential is attractive to the boundary Y = 0, this describes the portion of the ellipse
(x2/A2) + (y2/B2) = 1 in the upper half plane, where
A2 =
4H′K2(b+ 2)
(4K2 + β(b+ 2))2
, B2 =
H′(b+ 2)
4K2 + β(b+ 2)
,
x = X − 4X2
4K2 + β(b+ 2)
, y = Y.
If β ≤ 0 the potential is repulsive. If however, 4K2 + (b + 2)β > 0 the trajectory (2.16) is
again a portion of an ellipse. There is a special case that when β = 0, we will have A2 = B2
such that the trajectory will be upper half of a circle. A plot of a trajectory when β = 0 is given
in Fig. 7. And an example of elliptic trajectory is given in Fig. 8.
If 4K2 + (b + 2)β < 0 the trajectory is the upper sheet y > 0 of the hyperbola (x2/A2) −
(y2/B′2) = 1 where
A2 =
4H′K2(b+ 2)
(4K2 + β(b+ 2))2
, B′
2
= − H′(b+ 2)
4K2 + (b+ 2)β
,
x = X − 4X2
4K2 + β(b+ 2)
, y = Y.
If 4K2 + (b + 2)β = 0 equation (2.16) becomes degenerate. From (2.15) we now find
X1 = −Y 2K2−2XY pYK and haveH′ = [4Y 2/(b+2)]·p2
Y ≥ 0. Eliminating pY from these two ex-
pressions we find the equation (K2Y 2+X1)2−H′K2(b+2)X2 = 0 for the unbounded trajectories.
These are upper halves of parabolas with orbit equation Y 2 = −(X1/K) ±
√
H′(b+ 2)X, and
we notice that a certain subset of H′, X1, K and b would correspond to two half-parabolas which
are symmetric with respect to y-axis. Interestingly, when X1/K ≤ 0, the two half-parabolas will
Examples of Complete Solvability of 2D Classical Superintegrable Systems 11
Ellipse
X
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Y
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Origin
Figure 8. Elliptic trajectory on Poincaré upper half plane with β = 2, b = 0, H′ = 2, K = 1, X2 = 1.
Hyperbola
X
-10 -8 -6 -4 -2 Origin 2 4 6
Y
0
1
2
3
4
5
6
Figure 9. Hyperbolic trajectory on Poincaré upper half plane with β = −3, b = 0, H′ = 2, K = 1,
X2 = 1.
A Parabolic Trajectory
X
-15 -10 -5 Origin 5 10 15
Y
0
2
4
6
Figure 10. Parabolic trajectory on Poincaré upper half plane with β = −2, b = 0, H′ = 2, K = 1,
X1 = 1.
A Intersecting Parabolic Trajectory
X
-15 -10 -5 Origin 5 10 15
Y
0
1
2
3
4
5
6
7
Figure 11. Parabolic trajectory on Poincaré upper half plane with β = −2, b = 0, H′ = 2, K = −1,
X1 = 1.
cross as in Fig. 11. A special case occurs when H′ = 0 so that the trajectory becomes a straight
line. Since H′ = (4Y 2 · p2
Y )/(b + 2), we see this implies that pY = 0 so the trajectory must be
parallel to the x-axis. Furthermore, if H′ = 0 and X1/K = 0, the trajectory will be exactly the
y-axis. Moreover, for H′ = 0 and X1/K < 0, there will be no trajectory.
12 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
We mention that this Poincaré upper half plane model can be mapped to the unit disk, the
Poincaré disk, as well as to the upper sheet of the 2-sheet hyperboloid is 3-space. Further, this
model can be contracted to the flat space system H = p2
x + p2
y + αx, E5 in our listing [23].
Indeed, if we set
X = y, Y = x+
√
b+ 2
2ε
, β =
α
√
b+ 2
ε
, H(ε) = ε2
(
H′ − α(b+ 2)3/2
4ε
)
,
then lim
ε→0
H(ε) = p2
x+p2
y+αx. We don’t go into detail here but we give another detailed example
in the next section.
2.3 Contraction to the D3 oscillator
The oscillator in the space Darboux 3 (D3) has Hamiltonian
H =
1
2
e2x
ex + 1
(
p2
x + p2
y
)
+
β
ex + 1
(2.17)
with a single Killing vector J = py and a symmetry algebra basis, {H,J 2,X1,X2}. Here,
X1 =
1
2
ex sin ypxpy +
1
4
e2x
ex + 1
cos yp2
x −
1
4
ex(ex + 2)
ex + 1
cos yp2
y +
1
2
β cos y
ex + 1
,
X2 = −1
2
ex cos ypxpy +
1
4
e2x
ex + 1
sin yp2
x −
1
4
ex(ex + 2)
ex + 1
sin yp2
y +
1
2
β sin y
ex + 1
.
The structure relations are
{J ,X1} = −X2, {J ,X2} = X1, {X1,X2} =
1
2
JH − β
2
J ,
and there is the functional relation 4X 2
1 + X 2
2 − 1
4H
2 − 1
2J
2H+ β
2J
2 = 0.
Theorem 2.5. The D3 oscillator system is a contraction of the D4 oscillator system.
Proof. We can get (2.17) as a limiting case of (2.1) by taking
x =
x′
2
− 1
2
ln(ε), y =
y′
2
, b =
1
ε
, H′ = 8εH, α =
β
8ε2
. (2.18)
Then we have H′ = 1
2
e2x
′
ex′+1
(p2
x′ + p2
y′), J ′ = 2J = py′ , and Y ′1 ≈ −4εY1, Y ′2 ≈ −4εY2.
First, it is obvious that J ′ = 2J = py′ . Then as ε→ 0, since
cosh(2x) =
ex
′−ln(ε) + e−x
′+ln(ε)
2
=
ex
′
+ ε2e−x
′
2ε
, sinh(2x) =
ex
′ − ε2e−x′
2ε
,
plugging these two identities into the Hamiltonian equation (2.1) and also using equation (2.18),
we have
H′ = 8ε
e2x
′−2ε2+ε4e−2x′
4ε2
ex′+ε2e−x′
ε + 1
ε
(
1
4
p2
x′ +
1
4
p2
y′) + 8ε
(
β
8ε2
ex′+ε2e−x′
ε + 1
ε
)
=
1
2
e2x′ − 2ε2 + ε4e−2x′
ex′ + ε2e−x′ + 1
(
p2
x′ + p2
y′
)
+
β
ex′ + ε2e−x′ + 1
=
1
2
e2x′
ex′ + 1
(
p2
x′ + p2
y′
)
+
β
ex + 1
,
Examples of Complete Solvability of 2D Classical Superintegrable Systems 13
because the terms with ε in front of them will all go to 0 as ε → 0. This H′ is exactly the
Hamiltonian of the Darboux-3 oscillator.
Similarly, by substituting the identities in equation (2.18) into equations (2.7) and (2.8), we
have,
Y ′1 = −4ε(− cos y′)
e2x′ − 2ε2 + ε4e−2x′
4ε(ex′ + ε2e−x′ + 1)
(
1
4
p2
x′ +
1
4
p2
y′
)
− cos y′
β
8ε(ex′ + ε2e−x′ + 1)
+ cos y′
(
ex
′
+ ε2e−x
′
2ε
)
1
4
p2
y′ − sin y′
(
ex
′ − ε2e−x′
2ε
)
1
4
px′py′ =
1
4
e2x′
ex′ + 1
cos(y′)p2
x′
− 1
4
(
2ex
′
(ex
′
+ 1)
ex′ + 1
− e2x′
ex′ + 1
)
cos(y′)p2
y′ +
1
2
ex
′
sin(y′)px′py′ +
1
2
β cos y′
ex′ + 1
=
1
4
e2x′
ex′ + 1
cos(y′)p2
x′ −
1
4
ex
′
(ex
′
+ 2)
ex′ + 1
cos(y′) p2
y′ +
1
2
ex
′
sin(y′)px′py′ +
1
2
β cos y′
ex′ + 1
,
as ε→ 0. In the same sense,
Y ′2 =
1
4
e2x′
ex′ + 1
sin(y′)p2
x′ −
1
4
ex
′
(ex
′
+ 2)
ex′ + 1
sin(y′) p2
y′ −
1
2
ex
′
cos(y′)px′py′ +
1
2
β sin y′
ex′ + 1
.
We notice they are exactly the same vectors as X1 and X2. So we have proved that in this
limiting case, Darboux-4 space contracts to Darboux-3. �
Using this transformation, we can get the orbit equation for the D3 oscillator from the orbit
equation for the D4(b) oscillator. First, since κ′2 = Y ′21 + Y ′22 = 16ε2(Y2
1 + Y2
2 ) = 16ε2κ2 and
assuming ε > 0, we have κ′ = 4εκ. Then putting κ = 1
4εκ
′ and the identities (2.18) into (2.9),
we derive
(
ex
′
+ε2e−x
′
2ε
)(J ′
2
)2 − cos(y′)(κ
′
4ε) = H′
8ε . Thus, ex
′J ′2 − 2 cos(y′)κ′ = H′ as ε→ 0.
2.3.1 Embedding D3 in 3D Minkowski space
We can embed D3 as a surface in 3D Minkowski space with coordinates X, Y , Z in such a way
as to preserve rotational symmetry. For example, let
X = 2
√
2e−
x
2
√
1 + e−x cos
y
2
, Y = 2
√
2e−
x
2
√
1 + e−x sin
y
2
,
Z =
√
6
12
ln
( √
3
6 (6 + 5ex)
√
3 + 2e2x + 5ex + 1
√
3
6 (6 + 5ex)
√
3 + 2e2x + 5ex − 1
)
− e−x
√
2
√
3 + 2e2x + 5ex.
Then dX2 + dY 2 − dZ2 = 2(ex+1)
e2x
(dx2 + dy2). Such embeddings are not unique.
Discussion of trajectories. To see how the trajectory behaves on this surface, we impose
the orbit equation ex
′J ′2 − 2 cos(y′)κ′ = H′ to X, Y and Z. From the orbit equation, we have
ex =
H+ 2 cos y · κ
J 2
=
H
J 2
+
2κ
J 2
cos y = a+ b cos y, (2.19)
where we ignore all the primes on the letters for convenience. Also, we treat a = H
J 2 and b = 2κ
J 2
as two new constants, so the number of constants is reduced by 1. Then we can express X, Y
and Z all in terms of y. We notice that since ex > 0, the trajectory will be closed when a > b
and unbounded when a < b.
Example 2.6. Plots of an“elliptical” shape trajectory and its overhead 2-D view are given in
Figs. 12 and 13. (The size of the “cone” that appears on the graphics is determined by the
minimum value we choose for ex in MATLAB. The smaller the minimum value, the larger the
“cone”. For these two plots, the minimum value for ex we choose is 1 and we adjust to different
sizes of “cones” for other examples to make the plots clearer.)
14 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Figure 12. 3-D View of trajectory with a = 2, b = 1.
Figure 13. Overhead view of trajectory with a = 2, b = 1.
Example 2.7. As a becomes smaller, the “ellipse” is prolonged, see Figs. 14 and 15. Seen from
above the shape of the trajectory two merging ellipses. (The minimum value for ex we choose
is 0.1 for these two plots.)
Example 2.8 (escape velocity). For a = b, we encounter the boundary case between bounded
trajectory and unbounded trajectory. A plot of the case is given in Fig. 16. The shape of the
trajectory when looking from above is like two very “thin” parabolas meeting at the top of the
“cone”. (The minimum value for ex we choose is 0.001 for these two plots.)
Example 2.9 (unbounded trajectory). When a < b, there are no real x values corresponding
to a range of values of y, clearly seen from equation (2.19), so the trajectory is unbounded.
Examples are given in Figs. 17 and 18 which are a standard 3-D view and a closeup overhead view.
2.3.2 Contraction of the D3 oscillator to the isotropic oscillator
The isotropic oscillator in Euclidean space has the Hamiltonian
H = p2
x + p2
y + ω2
(
x2 + y2
)
(2.20)
Examples of Complete Solvability of 2D Classical Superintegrable Systems 15
Figure 14. 3-D View of trajectory with a = 1.1, b = 1.
Figure 15. Overhead view of trajectory with a = 1.1, b = 1.
Figure 16. 3-D View of escape velocity trajectory with a = b = 1.
and basis symmetries
K = xpy − ypx, L1 =
1
2
(
p2
x − p2
y
)
+
1
2
ω2
(
x2 − y2
)
, L2 = pxpy + ω2xy, K2.
The structure equations are
{L1,K} = 2L2, {L2,K} = −2L1, {L1,L2} = −2ω2K.
16 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Figure 17. 3-D View of trajectory with a = 0.7, b = 1.
Figure 18. Closeup overhead view of trajectory with a = 0.7, b = 1.
The functional relation is L2
1 + L2
2 = 1
4H
2 − ω2K2. We can obtain this system as a limit of the
D3 oscillator as follows: In (2.17) we set x = x′ + ln(1
ε ), y = y′, H′ = 2εH, β = (2ω2)/ε2. Then
as ε→ 0 we have K = py′ and
H′ = ex
′
(p2
x′ + p2
y′) +
4ω2
ex′
,
X ′1 = εX1 =
ex
′
2
(
sin y′px′py′ +
1
2
cos y′
(
p2
x′ − p2
y′
))
+
ω2 cos y′
ex′
,
X ′2 = εX2 =
ex
′
2
(
− cos y′px′py′ +
1
2
sin y′
(
p2
x′ − p2
y′
))
+
ω2 sin y′
ex′
.
In terms of flat space Cartesian coordinates X = r cos θ, Y = r sin θ we have
ex
′
=
4
r2
, y′ = 2θ, H′ = p2
X + p2
Y + ω2
(
X2 + Y 2
)
,
X ′1 =
1
4
(
p2
X − p2
Y
)
+
1
4
ω2
(
X2 − Y 2
)
, X ′2 =
1
2
pXpY +
1
2
ω2XY,
Examples of Complete Solvability of 2D Classical Superintegrable Systems 17
Figure 19. Elliptical trajectories.
with K = 1
2(XpY − Y pX). From the expressions of H′ and X ′1, we get
p2
x =
H′ + 4X ′1 − 2ω2X2
2
, p2
y =
H′ − 4X ′1 − 2ω2Y 2
2
. (2.21)
Then from the expression of X ′2, we have pxpy = 2X ′2 − ω2XY . Squaring this equation and
equating it to the product of the two equations (2.21) we obtain the orbit equation,
2ω2
(
X2 + Y 2
)
(H′ − 4X ′1)− 16ω2X ′2XY = H′ − 16X ′21 − 16X ′22 .
The shape of the orbit depends only on 8X ′2/(H′ − 4X ′1), so we are really investigating the
equation X2 + Y 2 − aXY = 1 where a = 8X ′2/(H′ − 4X ′1). We take the right side to be 1 since
we now only care about the shape, not the scale.
When a = 0, i.e., X ′2 = 0, the trajectory is just a circle. When 0 < a < 2, i.e., 0 <
8X ′2/(H′ − 4X ′1) < 2, the trajectory is a tilted ellipse with y = x as the major axis. In Fig. 19,
we show three trajectories for different values of a, and also the major axis. As a increases, the
ellipse becomes more elongated in the major axis direction.
When a = 2, i.e., H′ − 4X ′1 = 4X ′2, there is a bifurcation point on the momentum map.
The trajectory splits into two parallel straight lines. For a > 2, i.e., 8X ′2/(H′ − 4X ′1) > 2,
the trajectories are hyperbolas with y = x as the symmetry axis. In Fig. 20, we present
three hyperbolic trajectories, shown in darker colors as a increases. It is also possible that
8X ′2/(H′ − 4X ′1) < 0, in which case the trajectories will be symmetric about the line y = −x.
2.4 The Higgs oscillator S3
The classical system S3 on the 2-sphere is determined by Hamiltonian [14]
H = J 2
1 + J 2
2 + J 2
3 +
α
(
s2
1 + s2
2 + s2
3
)
s2
3
, (2.22)
where J1 = s2p3−s3p2 and J2, J3 are cyclic permutations of this expression. For computational
convenience we have embedded the 2-sphere in Euclidean 3-space. Thus we can write
H′ = p2
1 + p2
2 + p2
3 +
α
s2
3
=
H+ (s1p1 + s2p2 + s3p3)2
s2
1 + s2
2 + s2
3
18 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Figure 20. Hyperbolic trajectories.
and use the Euclidean space Poisson bracket {F ,G} =
3∑
i=1
(−∂siF∂piG + ∂piF∂siG) for our
computations, but at the end we restrict to the unit sphere: s2
1 + s2
2 + s2
3 = 1 and s1p1 + s2p2 +
s3p3 = 0. The Hamilton equations for the trajectories sj(t), pj(t) in phase space are
dsj
dt
= {H, sj},
dpj
dt
= {H, pj}, j = 1, 2, 3.
The classical basis for the constants of the motion is
L1 = J 2
1 + α
s2
2
s2
3
, L2 = J1J2 − α
s1s2
s2
3
, X = J3. (2.23)
The structure relations are
{X ,L1} = −2L2, {X ,L2} = 2L1 −H+ X 2 + α, {L1,L2} = −2(L1 + α)X , (2.24)
and the Casimir relation is
L2
1 + L2
2 − L1H+ L1X 2 + αX 2 + αL1 = 0.
To analyze the classical trajectories it is convenient to replace the basis elements L1, L2 in the
algebra with the new basis set
S1 =
1
2
(
J 2
1 − J 2
2 −
α
(
s2
1 − s2
2
)
s2
3
)
, S2 =
1
2
(
2J1J2 −
2αs1s2
s2
3
)
. (2.25)
Thus S1 = L1 + 1
2(X 2−H+α), S2 = L2. Now the first two equations (2.24) become {J3,S1} =
−2S2, {J3,S2} = 2S1, so (S1,S2) transforms as a 2-vector with respect to rotations about the
3-axis. Indeed, a rotation through the angle β about the 3-axis rotates the vector by 2β. The
remaining relations now become
{S1,S2} = −X
(
H−X 2 + α
)
, S2
1 + S2
2 =
1
4
(
X 2 −H+ α
)2 − αX 2 ≡ κ2.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 19
This verifies that the length κ of the 2-vector is unchanged under a rotation and shows that
(S1,S2) is similar to the Laplace–Runge–Lenz vector for the Kepler problem in Euclidean space.
(However, a better superintegrable 2-sphere analog of the Kepler problem is the potential
αs3/
√
s2
1 + s2
2, called S6 in our listing [23].) In analogy with the choice of periaptic coordi-
nates to simplify the Kepler problem, we choose the preferred coordinate system such that the
vector (S1,S2) points along the 1-axis, i.e., S2 = 0, S1 = κ ≥ 0. Then equations (2.22) and (2.23)
become
J1J2 =
αs1s2
s2
3
, J 2
1 − J 2
2 = 2κ+
α
(
s2
1 − s2
2
)
s2
3
,
J 2
3 = X 2, J 2
1 + J 2
2 + J 2
3 = H− α
s2
3
. (2.26)
The last 3 equations can be solved to give
J 2
1 = κ+
1
2
(
H−X 2 − α
)
− αs2
2
s2
3
,
J 2
2 = −κ+
1
2
(
H−X 2 − α
)
− αs2
1
s2
3
, J 2
3 = X 2.
Substituting these results into the square of the first equation (2.26) and simplifying, we get the
result (assuming α 6= 0)[
κ+
H−X 2 − α
2
]
s2
1 +
[
−κ+
H−X 2 − α
2
]
s2
2 −X 2s2
3 = 0. (2.27)
This is the equation of a cone As2
1 + Bs2
2 + Cs2
3 = 0. The trajectories lie on the intersection
of this cone and the unit sphere s2
1 + s2
2 + s2
3 = 1. Thus we get conic sections again, as in the
Euclidean Kepler problem, but this time the sections are intersections with the unit sphere,
rather than planes. The possible types of trajectory will depend on the signs of A, B, C.
The projection on the s1 − s2 plane is the curve(
H+ X 2 − α
2
+ κ
)
s2
1 +
(
H+ X 2 − α
2
− κ
)
s2
2 = X 2. (2.28)
The identities[
κ+
1
2
(H−X 2 − α)
] [
−κ+
1
2
(H−X 2 − α)
]
= αX 2, (2.29)[
κ+
1
2
(H+ X 2 − α)
] [
−κ+
1
2
(H+ X 2 − α)
]
= HX 2, (2.30)[
κ+
1
2
(H−X 2 + α)
] [
−κ+
1
2
(H−X 2 + α)
]
= αH (2.31)
will be important in the analysis of trajectories to follow.
Analog of Kepler’s second law of planetary motion. We see from equations (2.28)
and (2.30) that for the nonzero angular momentum X the projection of the motion on the unit
circle in the s1 − s2 plane is a segment of an ellipse, hyperbola or straight line and that none of
these curves pass through the center (s1, s2) = (0, 0) of the circle. Thus as the particle moves
along its trajectory (s1(t), s2(t), s3(t)), the line segment connecting the projection (s1(t), s2(t)) to
the center of the circle sweeps out an area. Introducing polar coordinates s1(t) = r(φ(t)) cosφ(t),
s2(t) = r(φ(t)) sinφ(t) we see that in the interval from some initial time 0 to time t the area
20 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
swept out is A(t) = 1
2
∫ φ(t)
φ(0) r
2(φ)dφ. Thus the rate at which the area is swept out is dA
dt =
dA
dφ
dφ
dt = 1
2r
2(φ(t))dφdt . Now note from Hamilton’s equations that along the trajectory
X = J3 = s1(t)p2(t)− s2(t)p1(t) = s1
ds2
dt
− s2
ds1
dt
= r2dφ
dt
.
Thus dA
dt = X
2 , a constant.
Theorem 2.10. If the angular momentum is nonzero, the projection of the trajectory on the
unit circle in the s1 − s2 plane sweeps out equal areas in equal times.
Now we begin an analysis of the trajectories. There are two cases, depending on whether the
potential is repulsive (α > 0) or attractive (α < 0).
Case 1: α > 0. This is the case of a repulsive potential. The equator of the sphere repels
the particle. Thus motion is confined to a hemisphere. We start an analysis of the types of
trajectories.
If X 6= 0 and using the fact that all trajectories will be periodic for a repulsive force, we see
that there will necessarily be at least one point on the phase space trajectory for which p2 = 0.
Substituting into the phase space conditions (2.26) it is easy to see that this is possible only
for s1 = 0. Solving all the equations completely we find that these points on the phase space
trajectory are uniquely determined by the constants of the motion as follows:
s1 = 0, s2
2 =
κ+ 1
2
(
H+ X 2 − α
)
H
, p2
1 = −κ+
1
2
(
H+ X 2 − α
)
,
p2 = 0, p3 = 0. (2.32)
This prescription gives us a point on each trajectory, comparable to the aphelion for the Kepler
system, where we can start to trace out the orbit. It remains to analyze the possible orbits. To
see what is the available parameter space for the case X 6= 0, note that first we must require
κ ≥ 0. Noting the identity (2.29) we see that the quantities in brackets must have the same
sign. However, if that sign is negative then the cone (2.27) will degenerate to a point and not
intersect the sphere. Thus to obtain trajectories it is necessary that the constants of the motion
satisfy −κ+ 1
2(H−X 2−α) > 0, κ ≥ 0. On the other hand, if these conditions are satisfied, we
see from equations (2.32) that there exist trajectories for which the corresponding constants of
the motion are assumed. Thus the conditions are necessary and sufficient.
An Analog of Kepler‘s 3rd law of planetary motion. For nonzero angular momen-
tum X the projection of the motion on the unit circle in the s1 − s2 plane is an ellipse (2.28)
enclosing the center of the circle. The area of this ellipse is easily seen to be πX/
√
H. Let T be
the period of the trajectory. Thus T is the length of time for the projection of the trajectory to
trace out the complete ellipse and A(T ) = πX/
√
H. Since the area of the ellipse is swept out at
the constant rate dA
dt = X
2 we have A(T ) = dA
dt T = XT
2 . Equating these two expressions for A(t)
and solving for T we find T = 2π/
√
H.
Theorem 2.11. For X 6= 0 the period of an orbit is T = 2π/
√
H.
If X = 0 then from (2.27) and (2.28) we see that the projection of the motion in the s1 − s2
plane is the straight line segment s1 = 0. Thus this motion takes place in a plane and the
trajectory is a portion of a great circle passing through the poles of the sphere. Here, κ =
1
2(H−α), H ≥ α > 0. The motion is periodic and the points of closest approach to the equator
are those such that s2
2 = (H− α)/H. The period is again T = 2π/
√
H. There is a special case
of equilibrium at a pole when H = α.
Case 2: α < 0. This is the case of an attractive potential, where the equator of the sphere
attracts the particle. Again, motion is confined to a hemisphere. We first consider the case
Examples of Complete Solvability of 2D Classical Superintegrable Systems 21
where X 6= 0. Without loss of generality we can assume X > 0. Then the projection of the
motion is counter-clockwise. Now κ > 0 and from the identity (2.30) we see that there will be
3 classes of trajectories, depending on the value of H.
If H > 0 then the coefficients of s2
1 and s2
2 in the projection formula (2.28) must have the
same sign, necessarily positive for a real trajectory. Thus
H > 0: H+ X 2 − α > 2κ > H−X 2 − α, 2κ+H−X 2 − α > 0,
and the projection will be a segment of an ellipse. It is straight-forward to check that the ellipse
always intersects the circle at 4 points. All of these trajectories lead to annihilation at the
equator. There will necessarily be a “perihelion” point on each trajectory: s2 = 0, p1 = p3 = 0,
s2
1 = X 2/
[
κ+ 1
2(H+ X 2 − α)
]
, p2
2 = κ+ 1
2(H+ X 2 − α).
If H < 0 then the coefficient of s2
1 is positive and the coefficient of s2
2 is negative in the
projection formula (2.28). Thus
H < 0: H+ X 2 − α− 2κ < 0, H+ X 2 − α+ 2κ > 0,
and the projections will be segments of hyperbolas. Each branch of the hyperbola intersects the
circle at exactly 2 points. Again, these trajectories lead to annihilation at the equator. Again
there is a “perihelion” point on each trajectory: s2 = 0, p1 = p3 = 0, s2
1 = X 2/
[
κ+1
2(H+X 2−α)
]
,
p2
2 = κ+ 1
2(H+ X 2 − α).
If H = 0 then κ = 1
2(X 2 − α) so the coefficient of s2
1 is positive and the coefficient of s2
2 is 0
in the projection formula (2.28). Thus
H = 0: X 2 − α = 2κ
and the projections will be segments of straight lines. s2
1 = X 2/(X 2 − α). Here p1 = 0 and
p2
2 = X 2−α, a constant in accordance with the analog of Kepler’s Second Law. These trajectories
lead to annihilation at the equator. There is a “perihelion” point on each trajectory: s2 = 0,
p1 = p3 = 0, s2
1 = X 2/(X 2 − α), p2
2 = X 2 − α.
Finally, we suppose X = 0, so that the motion takes place in a plane through the poles.
If H − α > 0 then the motion takes place in the plane s1 = 0. All trajectories annihilate at
the equator. The projected trajectories each pass through the center of the circle, and we have
p2
2 = H−α−Hs2
2. If H−α < 0 then the motion takes place in the plane s2 = 0. All trajectories
annihilate at the equator. The projected trajectories do not pass through the center of the circle,
and we have p2
1 = H − α − Hs2
1. If H = α then there are possible trajectories on any plane
passing through the poles. If we go to new rotated s1, s2 coordinates such that the motion takes
place in the plane s1 = 0, then p1 = 0 and p2
2 = −αs2
2. All trajectories annihilate at the equator,
except for unstable equilibria at the poles.
2.4.1 Contraction of D4(b) to the Higgs oscillator on the sphere
By letting ε → 0 in b = −2 + ε2 and defining new constant α = 4β, the Hamiltonian equa-
tion (2.1) becomes
H =
sinh2(2x)
2 cosh(2x)− 2
(
p2
x + p2
y
)
+
4β
2 cosh(2x)− 2
=
(2 sinhx coshx)2
(2 sinhx)2
(
p2
x + p2
y
)
+
4β
(2 sinhx)2
= cosh2 x
(
p2
x + p2
y
)
+
β(cosh2 x− sinh2 x)
sinh2 x
= cosh2 x
(
p2
x + p2
y
)
+
β cosh2 x
sinh2 x
− β.
We can ignore the constant term −β and write
H′ = cosh2 x
(
p2
x + p2
y
)
+
β cosh2 x
sinh2 x
,
such that H′ = H+ β. This is the Higgs oscillator on the 2-sphere.
22 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
In terms of the Euclidean embedding of the sphere we have s1 =cos y/coshx, s2 =sin y/coshx,
s3 = sinhx/coshx, so s2
1+s2
2+s2
3 = 1. If we set s1 = r′ cos θ′, s2 = r′ sin θ′ as in polar coordinates,
then r′ = 1/coshx and θ′ = y. However, we notice that in the analog of Kepler’s 2nd law part
of the Section 2.1, r =
√
2 cosh(2x) + b/(2 sinh 2x) and θ = 2y. As ε → 0, r = 1/(2 coshx).
Thus r′ = 2r and θ′ = θ/2, so this contraction space is a double covering of the original D4(b)
space. We should be careful of this fact in the further computations. Expressed in the phase
space (s1, s2, s3, p1, p2, p3), we have H′ = J 2
1 +J 2
2 +J 2
3 + β
s23
, where J1 = s2p3− s3p2 and J2, J3
are cyclic permutations of this expression. Considering the potential part of the Hamiltonian,
we have a new basis set, Y ′1 = J 2
1 − J 2
2 − β(s2
1 − s2
2)/s2
3, Y ′2 = 2J1J2 − (2βs1s2/s
2
3), J ′ = J3.
We will show that in this limit, the orbit equation (2.9) will yield the orbit equation (2.28).
First, we transform cosh(2x) and cos(2y) in equation (2.9) to express them in terms of s1 and s2.
(s3 can be expressed in terms of s1 and s2.) From the fact that cosh2 x = 1/(s2
1 + s2
2), we have
cosh(2x) = 2 cosh2 x− 1 =
2
s2
1 + s2
2
− 1, cos(2y) = cos2 y − sin2 y =
s2
1 − s2
2
s2
1 + s2
2
.
Then equation (2.9) becomes [2/(s2
1 + s2
2)− 1]J ′2 − [(s2
1 − s2
2)/(s2
1 + s2
2)]κ′ = H, where κ′ is the
length of the two vector (Y ′1,Y ′2) and we set Y ′2 = 0 such that κ′ = Y ′1. Multiplying the
equation by (s2
1 + s2
2), plugging in the fact that H = H′− β and doing some rearrangements, we
obtain(
H′ + J ′2 − β + κ′
2
)
s2
1 +
(
H′ + J ′2 − β − κ′
2
)
s2
2 = J ′2, (2.33)
where we realize that κ′ = Y ′1 which is 2 times the basis S1 in equation (2.25), so κ′ = 2κ
where κ is the one in equation (2.28). Therefore, we can see that (2.33) is the same orbit
equation as (2.28).
Analog of Kepler’s second law of planetary motion. Following the same procedure
as in Section 2.1, and noticing that r′ = 2r and θ′ = θ/2, we have, dA′
dt = 1
2
(
4r2(θ(t))
)
dθ
2dt =
r2 dθ
dt = 2dAdt . Therefore, dA′
dt = J ′, which matches the expression for the Higgs oscillator.
Period of the orbit. With the same procedure as in Section 2.1 and r′ = 2r and θ′ = θ/2,
we find area swept out as a trajectory goes through one period T ′ is A′ = 1
2
∫ 2π
0 r′2(θ′)dθ′ =
2
∫ 2π
0 r2(θ)dθ = 4A where A is the area in equation (2.12). Also, from the second law we see
that A′ = 1
2J
′T ′. Hence,
T ′ =
A′
J ′
=
4A
J ′
= 2T = π
(
2 + b√
(−2H− bH+ α)
+
2− b√
(2H− bH+ α)
)
,
where T is the same as equation (2.11).
In the limit as ε→ 0, we take b = −2 into the above equation of T ′ and get ,
T ′ = π
4√
4H+ α
=
2π√
4(H′ − β) + 4β
=
π√
H′
,
which differs from the period in Theorem 2.11 by a factor of 1
2 , due to the fact that the con-
traction is a double covering of the original D4 space.
2.4.2 Contraction of the Higgs to the isotropic oscillator
This contraction has a simple geometric interpretation, the contraction of a 2-sphere to a plane.
We can consider the Higgs oscillator as living in a 2-dimensional bounded “universe” of radius 1
Examples of Complete Solvability of 2D Classical Superintegrable Systems 23
in some set of units. Suppose an observer is situated in this universe “near” the attractive north
pole. Our observer uses a system of units with unit length ε where 0 < ε � 1 and we suppose
that using these units the universe appears flat to the observer. Thus in the observer’s units
we have s1 = εX, s2 = εY , s3 =
√
1− ε2(X2 + Y 2) = 1 − (X2 + Y 2)ε2/2 + O(ε4). Here, ε2 is
so small that to the observer it appears that s3 = 1. Thus, to the observer, it appears that the
universe is the plane s3 = 1 with local Cartesian coordinates (X,Y ). We compare the actual
system on the 2-sphere with the system as it appears to the observer and we assume that ε2 is
so small that it can be neglected, unless we are dividing by it. Now we have
s1 = εX, s2 = εY, s3 ≈ 1, p1 =
pX
ε
, p2 =
pY
ε
, p3 ≈ −(XpX + Y pY ).
We define new constants ω2, h by α = ω2/ε4, E − X 2 − ω2/ε4 = h/ε2, where J3 = X . Then
substituting these results into the Higgs Hamiltonian equation H ≡ J 2
1 +J 2
2 +J 2
3 + α/s2
3 = E,
we find (p2
X + p2
Y )/ε2 +ω2/
(
ε2
√
X2 + Y 2
)
= h/ε2. Multiplying both sides of this equation by ε2
we obtain the Hamiltonian equation for the Euclidean space isotropic oscillator
H̃ ≡ p2
X + p2
Y + ω2
(
X2 + Y 2
)
= h,
in agreement with (2.20). Using the same procedure we find that the constants of the motion
become K = X = xpy − ypx, and
L̃1 = lim
ε→0
ε2L1 = p2
Y + ω2Y 2, L̃2 = lim
ε→0
ε2L2 = −pXpY − ω2XY,
an alternate basis for the symmetries of the isotropic oscillator. Similarly, the structure equations
for the Higgs oscillator go in the limit to the structure equations of the isotropic oscillator.
3 Examples of 2D 2nd degree nondegenerate systems
3.1 The system S7 on the sphere
The classical system S7 on the 2-sphere [23] is determined by the Hamiltonian
H = J 2
1 + J 2
2 + J 2
3 +
a1s1
s2
2
√
s2
1 + s2
2
+
a2
s2
2
+
a3s3√
s2
1 + s2
2
,
where J1 = s2p3−s3p2 and J2, J3 are cyclic permutations of this expression. We have embedded
the 2-sphere in Euclidean 3-space, so
p2
1 + p2
2 + p2
3 = J 2
1 + J 2
2 + J 2
3 +
(s1p1 + s2p2 + s3p3)2
s2
1 + s2
2 + s2
3
and we can use the Poisson bracket {F ,G} =
3∑
i=1
(−∂siF∂piG + ∂piF∂siG) for our computations,
but at the end we restrict to the unit sphere: s2
1 + s2
2 + s2
3 = 1 and s1p1 + s2p2 + s3p3 = 0. The
Hamilton equations for the trajectories sj(t), pj(t) in phase space are
dsj
dt
= {H, sj},
dpj
dt
= {H, pj}, j = 1, 2, 3.
The classical basis for the constants of the motion is
L1 = J 2
3 −
a1
√
s2
1 + s2
2s1
s2
2
+
a2
(
s2
1 + s2
2
)
s2
2
,
24 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
L2 = −J1J3 +
a1s3
(
s2
2 + 2s2
1
)
2s2
2
√
s2
1 + s2
2
+
a2s1s3
s2
2
+
a3s1
2
√
s2
1 + s2
2
,
The structure relations are R = {L1,L2} and
{R,L1} = a1a3 + 4L1L2,
{R,L2} = −6L2
1 − 2L2
2 + 4HL1 − 2a2H+ 4a2L1 +
a2
3 − a2
1
2
,
R2 + 4L3
1 − 4L2
1H+ 4L2
2L1 − 4a2L2
1 + 4a2HL1 − a2
1H
−
(
a2
3 − a2
1
)
L1 + 2a1a3L2 + a2a
2
3 = 0.
3.1.1 S7 in polar coordinates
In terms of polar coordinates r, θ where s1 = r cos θ, s2 = r sin θ, s3 = ±
√
1− r2, and 0 ≤ r ≤ 1,
0 ≤ θ < π, we have
H =
p2
θ
r2
+
(
1− r2
)
p2
r +
a1 cos θ
r2 sin2 θ
+
a2
r2 sin2 θ
± a3
√
1− r2
r
,
L1 = p2
θ +
2a1 cos θ + 2a2
1− cos(2θ)
, (3.1)
L2 = ±
√
1− r2
(
cos θ pθ
r
+ sin θpr
)
pθ ±
√
1− r2
(a1 + a1 cos2 θ + 2a2 cos θ)
2r(1− cos2 θ)
+
1
2
a3 cos θ.
Here, ± is interpreted as + in the northern hemisphere and − in the southern hemisphere.
Hamilton’s equations give
ṙ = 2
(
1− r2
)
pr, θ̇ =
2pθ
r2
, ṗθ =
a1 + 2a2 cos θ + a1 cos2 θ
r2 sin3 θ
,
ṗr =
2p2
θ
r3
+ 2rp2
r +
2a1 cos θ
r3 sin2 θ
+
2a2
r3 sin2 θ
± a3√
1− r2
± a3
√
1− r2
r2
.
Assumptions. We require a3 < 0, and a1, a2 > 0. and restrict our attention to trajectories
caged in one hemisphere (eastern or western):
−1 < s3 < 1, 0 < s2 < 1, −1 < s1 < 1, or 1 ≥ r > 0, π > θ > 0.
Note that the north pole is attractive and the south pole is repulsive.
3.1.2 Trajectories
Eliminating pθ, pr in the expressions for H, L1, L2 we find the implicit equation for the trajec-
tories:
0 = cos2 θa2
3r
2 − 4L2
1r
2 cos2 θ + 4Hr2L1 cos2 θ − 4 cos θa3r
2L2 − 8 cos θL1L2r
√
1− r2
− 4L1a1 cos θr2 − 2a1 cos θa3r
√
1− r2 + 4Hr2a1 cos θ − r2a2
1 + 4L2
1 − 4L1a2
− 4a1L2r
√
1− r2 + a2
1 + 4L2
2r
2 − 4a2a3
√
1− r2r + 4Hr2a2 − 4Hr2L1
+ 4a3
√
1− r2rL1.
This is a quadratic equation for cos θ as a function of r, with solutions
cos θ =
N(r)± 2
√
S(r)
D(r)
, (3.2)
Examples of Complete Solvability of 2D Classical Superintegrable Systems 25
in the upper hemisphere, where
N = a1a3
√
1− r2 + 2a3rL2 − 2Hra1 + 4L1L2
√
1− r2 + 2rL1a1,
S =
(
Hr2 −
√
1− r2ra3 − L1
)
R2, D = r
(
4HL1 − 4L2
1 + a2
3
)
.
We note that
N2 − 4S =
(
a2
3 + 4HL1 − 4L2
1
)(
4L2
2r
2 − 4a3
√
1− r2ra2 + 4L1a3
√
1− r2r + 4Hr2a2
− 4Hr2L1 − 4L1a2 − r2a2
1 − 4a1L2
√
1− r2r + 4L2
1 + a2
1
)
.
In the lower hemisphere the trajectories are given by
cos θ =
Ñ(r)± 2
√
S̃(r)
D̃(r)
,
Ñ = −a1a3
√
1− r2 + 2a3rL2 − 2Hra1 − 4L1L2
√
1− r2 + 2rL1a1, (3.3)
S̃ =
(
Hr2 +
√
1− r2ra3 − L1
)
R2, D̃ = r
(
4HL1 − 4L2
1 + a2
3
)
.
Since R2 ≥ 0 we see that if H ≤ L1 then S̃ ≤ 0. Thus for this case, the trajectory is never in the
lower hemisphere. If H ≥ L1 an elementary analysis shows that S̃(r) > 0 exactly in an interval
0 < α ≤ r ≤ 1 and is nondecreasing on that interval.
Case 1: a2 > a1 > 0. All trajectories are closed and periodic. We must have L1 > 0 for
trajectories and we initially assume R2 > 0. The equations for perigee and apogee (distance
from the projection of the trajectory in the equatorial plane to the origin) are
r2 =
a2
3 + 2HL1 ± a3
√
a2
3 + 4HL1 − 4L2
1
2(a2
3 +H2)
.
Thus, in order to have physical trajectories we must have
a2
3 + 4HL1 − 4L2
1 ≥ 0 and a2
3 + 2HL1 ≥ 0.
Suppose the trajectory touches the equatorial plane at r = 1, θ = θ0. Then taking a limit in (3.1)
as the trajectory goes to the boundary we find that if the trajectory touches the equatorial plane
it does so at an angle cos θ which is a solution of the quadratic equation(
1
4
a2
3 + (H−L1)L1
)
cos2 θ − (a3L2 − a1(H−L1)) cos θ + L2
2 − (H−L1)(L1 − a2) = 0.
Indeed, cos θ0 =
[
2(a3L2 + a1(L1 −H))± 2
√
(H−L1)R2
]
/[a2
3 + 4L1(H−L1)]. Thus, a neces-
sary condition for the trajectory to reach the equatorial plane is that H ≥ L1.
If the trajectory just touches the equatorial plane but doesn’t go into the lower hemisphere
then we must have H = L1 and −a3
2 > L2 > a3
2 . An example of touching the equatorial
plane is Fig. 21. The different colors in the accompanying graphs correspond to the different
curves (3.2), (3.3) that make up the trajectories. If the trajectory passes through the equatorial
plane then we must have H > L1 and 4L2
1− 4a− 2L1 + a2
1 ≥ 0. The angle of crossing lies in the
interval
− a1
2L1
− 1
2L1
√
4L2
1 − 4a2L1 + a2
1 ≤ cos θ0 ≤ −
a1
2L1
+
1
2L1
√
4L2
1 − 4a2L1 + a2
1,
and for each θ0 in that interval the possible values of L2 are
L2 =
a3 cos θ0
2
±
√
(L1 −H)(cos2 θ0 L1 − L1 + a1 cos θ0 + a2).
26 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
−1−0.8−0.6−0.4−0.200.20.40.60.81
−1
−0.5
0
0.5
1
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
s2
s1
s
3
Figure 21. Touching the equatorial plane: a1 = 3, a2 = 5, a3 = −6, L1 = 6, H = 6, L2 = 1.
−1−0.8−0.6−0.4−0.200.20.40.60.81−1−0.500.51
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
s1
s2
s
3
Figure 22. Case 1: a1 = 3, a2 = 4, a3 = −5, L1 = 4, L2 = 0, H = 6.
−1−0.500.51 −1−0.500.51
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
s2
s1
s
3
Figure 23. Case 1: a1 = 2, a2 = 3, a3 = −4, L1 = 3, L2 = 0.5, H = 5.
Note that for each choice of the constants of the motion, there are always two crossing angles.
Thus if a trajectory crosses into the lower hemisphere, it must return to the upper hemisphere:
No trajectory remains confined to the lower hemisphere. Examples are Figs. 22 and 23. If the
angle is fixed, we can use the formula to get L2: Examples are Figs. 24 and 25. If R2 = 0 then
S(r) ≡ 0. The previous analysis is correct, except the trajectory is a single arc, rather than
a loop. We call this a metronome orbit. The particle moves back and forth along the arc with
a fixed period. We give no more details here because we will study the analogous systems in
our treatment of E16.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 27
−1−0.8−0.6−0.4−0.200.20.40.60.81
−1
0
1
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
s2
s1
s
3
Figure 24. a1 = 8, a2 = 10, a3 = −10, L1 = 40, H = 50, L2 = −15.1491.
−1−0.8−0.6−0.4−0.200.20.40.60.81 −101
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
s2s1
s
3
Figure 25. a1 = 8, a2 = 10, a3 = −10, L1 = 40, H = 50, L2 = −12.9919.
−1
−0.5
0
0.5
1
−1
−0.5
0
0.5
1
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
s2
s1
s
3
Figure 26. a1 = 2, a2 = 1, a3 = −4, L1 = 5, H = 8, L2 = 2.
Case 2: a1 > a2. Then the portion −1 < s1 < −a2/a1 of the s1-axis becomes attractive, the
trajectories are not closed; each end of the trajectory impacts this interval, assuming R2 > 0.
If H > L1 then one impact occurs in the northern hemisphere and one impact in the southern
hemisphere. If H ≤ L1 both impacts occur in the northern hemisphere.
An example for H > L1 is Fig. 26, and for case H < L1 is Fig. 27. If R2 = 0 then the
trajectory is a single arc that impacts the interval −1 < s1 < −a2/a1 of the s1-axis in the
northern hemisphere.
28 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Figure 27. a1 = 4, a2 = 2, a3 = −10, L1 = 5, H = 2, L2 = 1.
3.2 System E16 in Euclidean space as a contraction of S7
E16 can be described as a Kepler–Coulomb system with barrier in two dimensions [23, 30]. The
system has Hamiltonian,
H = p2
x + p2
y +
a1√
x2 + y2
+
a2y
x2
√
x2 + y2
+
a3
x2
.
In terms of the rotation generator J = xpy − ypx the defining constants of the motion are H
and
L1 = J 2 +
a2y
√
x2 + y2
x2
+ a3
y2
x2
,
L2 = pxJ −
a1
2
y√
x2 + y2
− a2
(
y2
x2
√
x2 + y2
+
1
2
√
x2 + y2
)
− a3
y
x2
.
The time derivative of a function F of the phase space parameters along a trajectory is given
by Ḟ ≡ ∂F
∂t = {H, F}, so in Cartesian coordinates we have the equations of motion
ẋ = 2px, ṗx =
1
x3(x2 + y2)3/2
(
a1x
4 + a2y
(
3x2 + 2y2
)
+ 2a3
(
x2 + y2
)3/2)
,
ẏ = 2py, ṗy =
1
(x2 + y2)3/2
(a1y − a2).
In polar coordinates x = r cos θ, y = r sin θ we have
pθ = xpy − ypx, rpr = xpx + ypy,
px = −sin θ
r
pθ + cos θpr, py =
cos θ
r
pθ + sin θpr,
H = p2
r +
p2
θ
r2
+
a1
r
+
a2 sin θ
r2 cos2 θ
+
a3
r2 cos2 θ
, L1 = p2
θ +
a2 sin θ
cos2 θ
+
a3 sin2 θ
cos2 θ
,
L2 =
(
cos θpr −
sin θ
r
pθ
)
pθ −
a1 sin θ
2
− a2
r
(
sin2 θ
cos2 θ
+
1
2
)
− a3 sin θ
r cos2 θ
.
Setting R = {L1,L2} we can verify that
R2 = 4(L1 + a3)
(
L1H−L2
2
)
+ a2
1L1 + 2a1a2L2 + a2
2H. (3.4)
Note also that H = p2
r + a1
r + L1+a3
r2
. In polar coordinates the equations of motion are
θ̇ = 2
pθ
r2
, ṙ = 2pr, ṗr = 2
L1 + a3
r3
+
a1
r2
,
Examples of Complete Solvability of 2D Classical Superintegrable Systems 29
ṗθ =
1
r3 cos3 θ
(
a2
(
cos2 θ − 2
)
− 2a3 sin θ
)
.
From these equations we see that points on a trajectory such that θ̇ = 0 are characterized by
pθ = 0. If such a point has coordinates (θ0, r0, pr0 , 0) we find that
sin θ0 =
−a2 ±
√
a2
2 + 4L1(L1 + a3)
2(L1 + a3)
, pθ0 = 0,
r0 =
a2(1 + sin2 θ0) + 2a3 sin θ0
(1− sin2 θ0)(a1 sin θ0 − 2L2)
, p2
r0 = H− a1
r0
− L1 + a3
r2
0
.
Similarly, if ṙ = 0 then pr = 0 and if such a point has coordinates (θ1, r1, 0, pθ1) we find in
particular that r1 = (a1 ±
√
a2
1 + 4H(L1 + a3))/(2H).
Assumption 1 : a1 < 0, a2 > 0, a3 > 0.
Possible values taken by the constants of the motion depend on the relative sizes of a2 and a3.
Case 1: a3 > a2. Then L1 + a3 > 0 and we must have a2
1 + 4H(L1 + a3) ≥ 0 for trajectories.
Unbounded trajectories correspond to H ≥ 0, and for H > 0 perigee occurs at
pr1 = 0, r1 =
a1 +
√
a2
1 + 4H(L1 + a3)
2H
,
p2
θ1 =
L1 − a2 sin θ1 − (L1 + a3) sin2 θ1
1− sin2 θ1
, sin θ1 = − 2r1L2 + a2
2L1 + 2a3 + a1r1
.
For H = 0 perigee occurs at
pr1 = 0, r1 = −L1 + a3
a1
,
p2
θ1 =
L1 − a2 sin θ1 − (L1 + a3) sin2 θ1
1− sin2 θ1
, sin θ1 = − a2
L1 + a3
+ 2
L2
a1
.
Bounded orbits correspond to H < 0, in which case perigee occurs at
pr1 = 0, r1 =
a1 +
√
a2
1 + 4H(L1 + a3)
2H
,
p2
θ1 =
L1 − a2 sin θ1 − (L1 + a3) sin2 θ1
1− sin2 θ1
, sin θ1 =
2r1L2 + a2
2L1 + 2a3 + a1r1
.
and apogee at
pr2 = 0, r2 =
a1 −
√
a2
1 + 4H(L1 + a3)
2H
,
p2
θ2 =
L1 − a2 sin θ2 − (L1 + a3) sin2 θ2
1− sin2 θ2
, sin θ2 =
2r2L2 + a2
2L1 + 2a3 + a1r2
.
We must have a2
2 + 4L1(L1 + a3) ≥ 0 for trajectories. There is an equilibrium point at
pr0 = 0, pθ0 = 0, r0 = −2
L1 + a3
a1
, sin θ0 =
a3 −
√
a2
3 − a2
2
a2
.
Period of bounded orbit. Since dr
dt = 2
√
H− a1
r −
L1+a3
r2
, we have r dr
2
√
Hr2−a1r−(L1+a3)
=
dt. Thus the time needed to go from perigee r1 to apogee r2 is
lim
R1→r1+
lim
R2→r2−
∫ R2
R1
r dr
2
√
Hr2 − a1r − (L1 + a3)
=
T
2
.
30 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Clearly, this is the same time as needed to go from apogee to perigee. Hence the period T is
T = − a1
2(−H)3/2
[
lim
α→∞
arctanα− lim
α→−∞
arctanβ
]
= − a1π
2(−H)3/2
.
Thus the period depends only on the energy.
The trajectories. Eliminating pθ, pr in the expressions for H, L1, L2 we find the orbit
equations
r =
N(sin θ)± 2
√
S(sin θ)
D(sin θ)
, N = (a1a2 − 4a3L2 − 4L1L2) sin θ − 2a1L1 − 2a2L2,
S =
(
−(L1 + a3) sin2 θ + L1 − a2 sin θ
)(
4(L1 + a3)(HL1 − L2
2) + a2
1L1 + a2
2H+ 2a1a2L2
)
,
D =
(
4H(L1 + a3
)
+ a2
1) sin2 θ + (4a1L2 + 4a2H) sin θ − 4HL1 + 4L2
2. (3.5)
From structure equation (3.4) and the fact that R2 > 0 we see that the 2nd factor in S is
a nonnegative constant. Further, from (3.4) we see that the denominator does not vanish for
H < 0, that it has a single root sin θ = −2L2/a1 for H = 0, and a double root for H > 0.
For the single root case with H = 0, assuming the plus sign in (3.5), we set z = sin θ+2L2/a1
in (3.5) and expand r in a Laurent series series to find the asymptotic expression
r = −4
R2
a3
1z
2
+
(2a1a2 − 8L1L2 − 8a3L2)
a2
1z
+
1
4
(
4R2(L1 + a3) + 16L2
2(L1 + a3)2 − 8a1a2L2(L1 + a3) + a2
1a
2
2
)
a1R2
+O(z).
Note that here R2 = −4(L1 + a3)L2
2 + a2
1L1 + 2a1a2L2. If the minus sign is assumed in (3.5),
then we have the finite limit lim
z→0
r = − a1
4R2 [4L1(L1 + a3) + a2
2].
Parametrization of case 1 trajectories. Recall that a1, a2, a3 are fixed constants with
a1 < 0, a3 > a2 > 0. To parametrize the trajectories we use perigee coordinates `1, tp, rp where
`1 = L1 + a3 > 0, −1 < tp < 1, rp > 0. Then
H =
`1
r2
p
+
a1
rp
, L2 = − tp
rp
`1 −
1
2
a1tp −
a2
2rp
,
H > 0↔ `1
rp
> −a1, H = 0↔ `1
rp
= −a1, H < 0↔ `1
rp
< −a1.
Note that(
N + 2
√
S
)(
N − 2
√
S
)
= (a2
2 + 4(`1 − a3)`1)D. (3.6)
For H < 0, apogee occurs for radial distance ra =
a1r2p−rp|2`1+a1rp|
2(`1+a1rp) . For wedge boundaries S1, S2
(radial lines containing a point on the trajectory where S vanishes) to exist at all, we must
have a2
2 + 4(`1 − a3)`1 ≥ 0. Then the righthand boundary will always exist, but the left hand
boundary will exist only if 2`1 ≥ a2. (However, from the R2 > 0 inequality below, the left hand
boundary must always exist.) The location of the wedge boundaries is defined by t = sin θ,
where
S1 : t1 =
−a2 −
√
a2
2 + 4`1(`1 − a3)
2`1
, S2 : t2 =
−a2 +
√
a2
2 + 4`1(`1 − a3)
2`1
,
R2 =
(
2`1 + a1rp
rp
)2 (
`1 − a3 − a2tp − `1t2p
)
.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 31
Figure 28. Case 1: positive energy.
In order for the system to correspond to a trajectory, we must have R2 ≥ 0. We will first require
that R > 0, the generic case. We see that this is violated if either 1) 2`1 + a1rp = 0, or 2) tp lies
outside the open interval (t1, t2). Thus we have the conditions
−
√
a2
2 + 4`1(`1 − a3)
2`1
< tp +
a2
2`1
<
√
a2
2 + 4`1(`1 − a3)
2`1
, 2`1 + a1rp 6= 0.
For H > 0 the zeros of the denominator occur at the values t± of sin θ such that
t± =
a1rptp − a2 ± 2
√
(`1 + a1rp)(`1 − a3 − a2tp − `1t2p)
2`1 + a1rp
.
For H > 0, if the numerator of the trajectory equation (3.5) vanishes at sin(t) = tn then tn = t+
or t− and t1 < tn < t2. Since every zero of the denominator is a zero of one of the numerators,
we must have t1 < t− < t+ < t2. Further, if 2`1 < a2 then D has at most one zero. Thus, if this
case occurs then there is a single asymptote and a single wedge boundary.
Summary of necessary Case 1 conditions for trajectories with R2 > 0.
1) `1 = L1 + a3 > 0, −1 < tp < 1, rp > 0,
2) a2
2 + 4(`1 − a3)`1 ≥ 0,
3) 2`1 + a1rp 6= 0,
4) − a2 +
√
a2
2 + 4`1(`1 − a3)
2`1
< tp <
−a2 +
√
a2
2 + 4`1(`1 − a3)
2`1
.
For examples, see Figs. 28, 29, 30, 31.
Case 1a: R2 = 0. In this case equation (3.6) for the trajectories simplifies since S ≡ 0. We
have N2 = (a2
2 + 4(`1 − a3)`1)D so
r(θ) =
a2
2 + 4(`1 − a3)`1
N
.
Because N is linear in t = sin θ, it is easily seen that the trajectories are segments of ordinary
conic sections: lines, ellipses, circles, hyperbolas and parabolas. Now R2 = (
2`1+a1rp
rp
)2(`1−a3−
a2tp− `1t2p) = 0, so there are two possibilities: I) 2`1 +a1rp = 0, and II) `1−a3−a2tp− `1t2p = 0.
32 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Figure 29. Case 1: negative energy.
Figure 30. Case 1: zero energy.
Figure 31. Case 1: zero energy.
Case 1aI): Circular arc. 2`1 + a1rp = 0. We can express the constants of the motion in
terms of rp, tp alone:
`1 = −a1rp
2
, H =
a1
2rp
< 0, L2 = − a2
2rp
.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 33
The equation for the trajectories simplifies to r(θ) =
a22+4(`1−a3)`1
N = rp, so the trajectory is
a segment of a circle centered at the origin with radius rp. The wedge boundaries are t1 =
−a2−
√
a22+a21r
2
p+2a1a3rp
a1rp
, t2 =
−a2+
√
a22+a21r
2
p+2a1a3rp
a1rp
. Thus for this case to occur, rp must satisfy
a2
2 + a2
1r
2
p + 2a1a3rp ≥ 0. We conclude that this case always occurs for rp sufficiently small or
sufficiently large, but that there is a forbidden intermediate band for which no circular trajectory
exists.
Case 1aII): Conic section arc. `1−a3−a2tp− `1t2p = 0. Again we can express the constants
of the motion in terms of rp, tp alone:
`1 =
a3 + a2tp
1− t2p
, H =
a3 + a2tp
r2
p(1− t2p)
+
a1
rp
, L2 = − tp(a3 + a2tp)
rp(1− t2p)
− a1tp
2
− a2
2rp
.
The equation for the trajectories becomes
r(θ) = ±
rp(a2 + 2tpa3 + a2t
2
p)
(2a2tp − a1t2prp + a1rp + 2a3) sin θ + (1− t2p)(a2 − a1tprp)
= ± 1
α sin θ + β
,
where
α =
(2a2tp − a1t
2
prp + a1rp + 2a3)
rp(a2 + 2tpa3 + a2t2p)
, β =
(1− t2p)(a2 − a1tprp)
rp(a2 + 2tpa3 + a2t2p)
,
and the sign is chosen so that r > 0. Setting y = r sin θ, x = r cos θ we see that the trajectory
is a segment of the curve αy + βr = ±1 or βr = −αy ± 1, so β2(x2 + y2) = α2y2 ± 2αy + 1,
α2 + β2 > 0. We have the following possibilities for the curve segments:
β2 > α2 > 0: ellipse, 0 < β2 < α2 : hyperbola,
β2 = α2 : parabola, α = 0: circle,
β = 0: horizontal line.
The wedge boundaries are tp, −a2+a3tp
a3+a2tp
. For the circular arc we have rp = −2
a2tp+a3
a1(1−t2p)
. (This
coincides with Case 1aI.) For the horizontal line we have tp =
a2rp
a1
< 0, so trajectories are possible
only for rp < −a1/a2. Only strictly positive energy solutions are possible for the horizontal line.
For the parabolic arc we have rp = −2a3 − a2 − 2a2tp + a2t
2
p/a1(1 + tp)(1− tp)2. The energy is
always negative. Only strictly negative energies are possible for the elliptical arc.
In general, orbits are possible for H < 0 if and only if the constants of the motion satisfy the
inequality a3 + a2tp/(1− t2p)(−a1) < rp. Trajectories are possible in the positive energy case if
and only if rp < a3 + a2tp/(1− t2p)(−a1).
Case 2: a3 < a2. Now there are no restrictions on the sign of `1 = L1 + a3. The force in the
angular direction is always negative. In the first quadrant the force in the y direction is always
negative. It follows from this that there are no periodic orbits contained entirely in the first
quadrant. The positive y-axis is repelling, but the negative y-axis is attracting.
We assume, initially, that R2 > 0. Equations (3.5) for the trajectories still hold. The
condition a2
2 + 4`1(`1 − a3) > 0 is now satisfied automatically. There is at most a single wedge
boundary
t0 = sin θ0 =
−a2 +
√
a2
2 + 4`1(`1 − a3)
2`1
< 0.
The following general types of behavior occur:
1. If `1 < 0 the force in the radial direction is always negative. The trajectories all impact
the lower y-axis and are perpendicular to the axis at impact. Since ṗr is always strictly negative
34 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
along a trajectory, perigee must occur at yp = −rp on the negative y-axis. Note that R2 is
linear in H, always with positive linear coefficient. The critical value of H such that R2 = 0
is E0 = [4a3L2
2 + 4L1L2
2 − 2a1a2L2 − a2
1L1]/[4L2
1 + a2
2 + 4a3L1]. The qualitative behavior of the
trajectories divides into two basic classes, depending on the sign of N(sin θ) for θ = −π
2 . Note
that
N(−1) = (2L2 − a1)(2`1 − a2) + 2a1(a3 − a2), S(−1) = (a2 − a3)R2.
Case 2a: N(−1) > 0. Then if we increase H from E0 to E0 + ε for arbitrarily small ε > 0
there will be two points of intersection of the trajectory with the negative y-axis, one on the
curve r = (N + 2
√
S)/D and one on the curve r = (N − 2
√
S)/D. The trajectory will remain
bounded and there will be a wedge boundary. If we further increaseH, leaving L1,L2 unchanged,
this behavior will persist until the critical value E1 = 1
4 [(2L2 − a1)2]/[(a2 − a3)] ≥ E0, where
D(−1) = 0. Here the intersection point with the negative y-axis and the curve r = (N−2
√
S)/D
has moved to y = −∞ whereas the intersection point with the curve r = (N + 2
√
S)/D remains
finite, As H is further increased the behavior of the trajectory is that it is unbounded and
asymptotic to a radial line in the 2nd quadrant, that there is a wedge boundary and a single
intersection point with the negative y-axis.
Case 2b: N(−1) < 0. Again, we increase H from E0 to E0 + ε for arbitrarily small ε > 0.
Now there are no points of intersection with the negative y-axis and, in fact, the trajectory
is non-physical until H is increased to ε beyond the critical value E1. The curve is entirely
described by the function r = (N + 2
√
S)/D. The trajectory is unbounded, asymptotic to
a radial line in the 2nd quadrant and intersects the negative y-axis a single time. There is no
wedge boundary. Note that for this case, the region {N(−1) < 0,H ≤ E1} is nonphysical. (We
do not have a complete proof of this but strong numerical evidence.)
2. If `1 > 0 the behavior is different. The trajectories do not necessarily intersect the negative
y-axis. Some trajectories are disjoint. The qualitative behavior of the trajectories divides into
two basic classes, depending on the sign of N(sin θ) for θ = −π
2 . Note again that
N(−1) = (2L2 − a1)(2`1 − a2) + 2a1(a3 − a2), S(−1) = (a2 − a3)R2.
Case 2c: N(−1) > 0. If we increase H from E0 to E0 + ε for arbitrarily small ε > 0 there
will be two points of intersection of the trajectory with the negative y-axis, one on the curve
r = (N+2
√
S)/D and one on the curve r = (N−2
√
S)/D. The trajectory will remain bounded
if H < 0 (with a wedge boundary) and be unbounded if H > 0. The unbounded trajectory
will divide into two disconnected parts. If we further increase H, leaving L1,L2 unchanged,
this behavior will persist until the critical value E1 = 1
4 [(2L2 − a1)2][(a2 − a3)] ≥ E0, where
D(−1) = 0. Here the intersection point with the negative y-axis and the curve r = (N−2
√
S)/D
has moved to y = −∞ whereas the intersection point with the curve r = (N + 2
√
S)/D remains
finite, As H is further increased the behavior of the trajectory is that it is unbounded and
asymptotic to a radial line in the 2nd quadrant, and a single intersection point with the negative
y-axis. There is a wedge boundary.
Case 2d: N(−1) < 0. Again, we increase H from E0 to E0 + ε for arbitrarily small ε > 0.
There are no points of intersection with the negative y-axis and, the trajectory is non-physical
for H < 0, but physical for the H > 0 interval until H is increased to ε beyond the critical va-
lue E1. Then the trajectory is unbounded and asymptotic to a radial line in the 2nd quadrant.
Physical trajectories have a wedge boundary. See Fig. 32.
3. R2 = 0. In this case equation (3.6) for the trajectories simplifies since S ≡ 0. We have
N2 = (a2
2 + 4(`1 − a3)`1)D so
r(θ) = ±a
2
2 + 4(`1 − a3)`1
N
= ± a2
2 + 4(`1 − a3)`1
(a1a2 − 4L2`1)t+ 2(a1a3 − a2L2 − a1`1)
= ± 1
α sin θ + β
,
Examples of Complete Solvability of 2D Classical Superintegrable Systems 35
Figure 32. Case 2d: N(−1) < 0.
where α = a1a2−4L2`1
a22+4(`1−a3)`1
, β = 2(a1a3−a2L2−a1`1)
a22+4(`1−a3)`1
, and the sign is chosen so that r > 0. Setting
y = r sin θ, x = r cos θ we see that the trajectory is a segment of the curve αy + βr = ±1 or
βr = −αy± 1, so β2(x2 + y2) = α2y2± 2αy+ 1,α2 + β2 > 0. We have the following possibilities
for the curve segments:
β2 > α2 > 0: ellipse, 0 < β2 < α2 : hyperbola,
β2 = α2 : parabola, α = 0: circle,
β = 0: horizontal line.
Since R2 = 0 we can express the constants of the motion in terms of α, β alone:
H =
a2α
3 − a1α
2 − 2a3α
2β + a2αβ
2 + a1β
2
2β
, `1 =
a2α− a1
2β
,
L2 = −a2α
2 − a1α− 2a3αβ + a2β
2
2β
.
Case 2e: circular arc, a1a2 − 4L2`1 = 0. Then H = − a21
4`1
. The radius is r0 = ±2`1
a1
. There
is a wedge boundary at t0 = sin θ0 =
−a2+
√
a22+4`1(`1−a3)
2`1
. The trajectory impacts the negative
y-axis. See Fig. 33.
Case 2f: horizontal line segment, a2L2 + a1`1 = a1a3, y0 = a2
a1
. The energy is H =
a21(`1−a3)
a22
.
If the energy is negative there is a wedge boundary at t0 = sin θ0 =
−a2+
√
a22+4`1(`1−a3)
2`1
< 0, and
the trajectory impacts the negative y-axis. If the energy is positive, the trajectory is unbounded
and there is no wedge boundary, but the trajectory still impacts the negative y-axis.
Case 2g: parabolic arc. If α = β we have
L2 =
a1(a2 − 2a3 + 2`1)
2(2`1 − a2)
, H =
(a2 − a3)a2
1
(2`1 − a2)2
and the energy is positive. If α = −β we have
L2 =
a1(a2 + 2a3 − 2`1)
2(2`1 + a2)
, H = −(a2 + a3)a2
1
(2`1 + a2)2
and the energy is negative. There is a wedge boundary at t0 = sin θ0 =
−a2+
√
a22+4`1(`1−a3)
2`1
, and
the trajectory impacts the negative y-axis.
36 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Figure 33. Case 2e: circular arc.
Case 2h: elliptic arc.
H =
4`1L2
2 − 2a1a2L2 − a2
1(`1 − a3)
a2
2 + 4`1(`1 − a3)
,
the general expression. There is a wedge boundary at t0 = sin θ0 =
−a2+
√
a22+4`1(`1−a3)
2`1
, and the
trajectory impacts the negative y-axis.
Case 2i: hyperbolic arc. Here again,
H =
4`1L2
2 − 2a1a2L2 − a2
1(`1 − a3)
a2
2 + 4`1(`1 − a3)
,
the general expression. There is a wedge boundary at t0 = sin θ0 =
−a2+
√
a22+4`1(`1−a3)
2`1
, and the
trajectory impacts the negative y-axis.
There are no radial line trajectories in this case.
4 The Post–Winternitz superintegrable systems
These systems are especially interesting as a pair of classical and quantum superintegrable
systems that do not admit separation of variables, so they cannot be solved by traditional
methods. We treat only the classical case here and show that its superintegrability allows us to
determine the trajectories exactly. The quantum case can also be solved.
4.1 The classical system
The classical Post–Winternitz system is defined by the Hamiltonian [36]
H =
1
2
(
p2
x + p2
y
)
+
cy
x2/3
.
The generating constants of the motion are of 3rd and 4th degree:
L1 = 3p2
xpy + 2p3
y + 9cx1/3px + 6
cypy
x2/3
,
L2 = p4
x + 4cp2
x
y
x2/3
− 12cx1/3pxpy − 2c2 (9x2 − 2y2)
x4/3
.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 37
−0.06 −0.05 −0.04 −0.03 −0.02 −0.01 0
−0.3
−0.25
−0.2
−0.15
−0.1
−0.05
0
0.05
x
y
Figure 34. c = 5, E = 2, L1 = 1, L2 = 4.
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035
−0.02
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
x
y
Figure 35. c = −12, E = 4, L1 = 3, L2 = 6.
Here {L1, L2} = −188c3, so the constants of the motion generate a Heisenberg algebra. Since
there is no 2nd degree constant other than H, separation of the Hamilton–Jacobi equations is
not possible in any coordinate system, so separation of variables methods cannot be used to
compute the trajectories. However, we can find a parametric description of the trajectories.
Choosing constants H = E, L1 = `1 and L2 = `2, we can verify the two identities
−p3
y + 9cx1/3px − `1 + 6Epy = 0,(
−1
3
p4
y + 4py2E + 4E2 − `2 −
4
3
`1py
)
p2
x −
2
9
(
−p3
y − `1 + 6Epy
)2
= 0.
From these results we can solve for px and x, y as functions of the parameter py:
px =
2
(
−p3
y − `1 + 6Epy
)√
−6p4
y + 72Ep2
y − 24`1py + 72E2 − 18`2
,
x = − 1
5832c3
(
−6p4
y + 72Ep2
y − 24`1py + 72E2 − 18`2
)3/2
,
y =
1
324c3
(
−2`21 − 36E2p2
y + p6
y + 8`1p
3
y − 18Ep4
y + 9`2p
2
y + 72E3 − 18`2E
)
, (4.1)
and equations (4.1) enable us to plot the trajectories. See Figs. 34, 35 and 36. There are no
closed orbits. The trajectories always impact the y-axis from the left or right, depending on the
signs of the parameters.
5 Classical elliptic superintegrable systems
5.1 Elliptic superintegrability
Here we extend the ideas in [40, 41] to a system that separates in Jacobi elliptic coordina-
tes [35]. We start with a simple example, the Higgs oscillator S3 (2.22) [23], expressed in elliptic
38 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
0 0.05 0.1 0.15 0.2 0.25
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
x
y
Figure 36. c = −1, E = 0, L1 = 5, L2 = 5.
coordinates.
H =
1
sn2(α, k)− sn2(β, k)
(
p2
α − p2
β +
A
sn2(α, k)
− A
sn2(β, k)
)
.
The Hamilton–Jacobi equation can be taken as H = E for pα = ∂W
∂α , pβ = ∂W
∂β . The separation
equations in the coordinates α, β are
p2
α − E sn2(α, k) +
A
sn2(α, k)
+ Λ1 = 0, p2
β − E sn2(β, k) +
A
sn2(β, k)
+ Λ1 = 0. (5.1)
Here, H and L1 = −p2
α + sn2(α, k)H − A
sn2(α,k)
are constants of the motion, and L1 = Λ1 on
a coordinate hypersurface. To find another constant of the motion via the action-angle variable
construction we introduce M and N which are computed according to
M =
1
2
∫
dα
pα
, N =
1
2
∫
dβ
pβ
,
where pα, pβ are expressed in terms of coordinates α, β by (5.1). Applying action-angle theory,
see, e.g., [20, 21], we see that M −N is a constant of the motion. The integral M can be written
as
M =
1
2
∫
sn(α, k)dα√
E sn4(α, k)− Λ1 sn2(α, k)−A
,
with a similar integral for N . This can be calculated as an elliptic integral by choosing the new
variable ρ = sn2(α, k). We then have the relations
dρ
dα
= 2sn(α, k)cn(α, k)dn(α, k), cn(α, k) =
√
1− ρ, dn(α, k) =
√
1− k2ρ.
The change of variables is then tantamount to the computation of
M =
1
2
∫
cn(α, k)dn(α, k)sn(α, k) dα
cn(α, k)dn(α, k)
√
E sn4(α, k)− Λ1 sn2(α, k)−A
=
1
4k
√
E
∫
dρ√
(ρ− a)(ρ− b)(ρ− c)(ρ− d)
,
where a = 1, b = 1
k2
and c, d are roots of ρ2 − Λ1
E ρ −
A
E . A similar calculation applies to the
substitution µ = sn2(β, k) with exactly the same a, b, c and d. A convenient choice of integral is∫
dt√
(t− a)(t− b)(t− c)(t− d)
Examples of Complete Solvability of 2D Classical Superintegrable Systems 39
=
2√
(a− c)(b− d)
sn−1
(√
(b− d)(t− a)
(a− d)(t− b)
,
√
(b− c)(a− d)
(a− c)(b− d)
)
.
This integral can be used for t = ρ or µ. If we form sn(2k
√
E(a− c)(b− d)(M − N), κ) for
κ =
√
(b− c)(a− d)/(a− c)/(b− d) and use the addition formula [35]
sn(u+ v, κ) =
sn(u, κ)cn(v, κ)dn(v, κ) + sn(v, κ)cn(u, κ)dn(u, κ)
1− κ2sn2(u, κ)sn2(v, κ)
,
we obtain a constant of the motion. In terms of the coordinates it is helpful to write
pα =
√
E
√
(ρ− c)(ρ− d)
ρ
, pβ =
√
E
√
(µ− c)(µ− d)
µ
.
Then for Z =
√
b−d
a−c
1
a−d we have
sn(u, κ)cn(v, κ)dn(v, κ) = Z
√
ρ− a
ρ− b
b− a
µ− b
√
(µ− c)(µ− d)
with a similar expression for sn(v, κ)cn(u, κ)dn(u, κ). We also note that
κ2sn(ρ, κ)2sn(µ, κ)2 =
(ρ− a)(µ− a)(b− c)(b− d)
(ρ− b)(µ− b)(a− c)(a− d)
,
as well as√
(µ− c)(µ− d) = 2µ
√
(1− µ)(1− k2µ)
pµ√
E
,√
(ρ− c)(ρ− d) = 2ρ
√
(1− ρ)(1− k2ρ)
pρ√
E
,
sn(u, κ) =
√
(ρ− 1)(d k2 − 1)
(d− 1)(ρ k2 − 1)
, cn(u, κ) =
√
(ρ− d)(k2 − 1)
(d− 1)(1− ρ k2)
,
dn(u, κ) =
√
(ρ− c)(k2 − 1)
(c− 1)(1− ρ k2)
, κ =
√
(d− 1)(c k2 − 1)
(c− 1)(d k2 − 1)
,
with similar expressions for sn(v, κ), cn(v, κ) and dn(v, κ) with ρ replaced by µ. From this we
deduce that
√
E
Z sn(2k
√
E(a− c)(b− d)(M −N), κ) = 1
X where X is a constant of the motion,
linear in the momenta, viz X =
√
(1−k2ρ)(1−k2µ)(1−ρ)(1−µ)
(ρ−µ) (ρpρ − µpµ), in involution with the
classical Hamiltonian
H =
1
ρ− µ
(
ρ(1− ρ)
(
1− k2ρ
)
p2
ρ − µ(1− µ)
(
1− k2µ
)
p2
µ
)
− A
ρµ
.
This is just the first degree symmetry that we expect for this Hamiltonian. In this case we
choose the variables ρ and µ to obtain the constant. In particular, cd = −A/E, c+d = −Λ1/E.
This seems to be just a complicated way of deriving a straightforward result. However, we
can use the TTW trick [40, 41] by looking at the Hamiltonian
H =
1
sn2(pα, k)− sn2(qβ, k)
(
p2
α − p2
β +
A
sn2(pα, k)
− A
sn2(qβ, k)
)
,
40 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
where p and q are arbitrary nonzero numbers. The separation equations are
p2
α − E sn2(pα, k) +
A
sn2(pα, k)
+ Λ1 = 0,
p2
β − E sn2(qβ, k) +
A
sn2(qβ, k)
+ Λ1 = 0. (5.2)
Here, H and L1 = p2
α − sn2(pα, k)H + A
sn2(pα,k)
are constants of the motion. To find another
constant of the motion via the action-angle variable construction we introduce M and N which
are computed according to M = 1
2
∫
dα
pα
, N = 1
2
∫ dβ
pβ
, where pα, pβ are expressed in terms of
coordinates α, β by (5.2). Again M − N is a constant of the motion. The integral M can be
written as
M =
1
2
∫
sn(pα, k) dα√
E sn4(pα, k)− Λ1 sn2(pα, k)−A
,
with a similar integral for N .
This can be calculated as an elliptic integral by choosing the new variable ρ = sn2(pα, k).
We then have the relations
dρ
dα
= 2p sn(pα, k)cn(pα, k)dn(pα, k),
cn(pα, k) =
√
1− ρ, dn(pα, k) =
√
1− k2ρ.
The change of variables is then tantamount to the computation of
M =
1
2p
∫
cn(α̂, k)dn(α̂, k)sn(α̂, k) dα̂
cn(α̂, k)dn(α̂, k)
√
E sn4(α̂, k)− Λ1 sn2(α̂, k)−A
=
1
4k
√
Ep
∫
dρ√
(ρ− a)(ρ− b)(ρ− c)(ρ− d)
,
where α̂ = pα, a = 1, b = 1
k2
and c, d are roots of ρ2 − Λ1
E ρ −
A
E . A similar calculation applies
to the substitution µ = sn2(qβ, k) with exactly the same a, b, c and d. A convenient choice of
integral is∫
dt√
(t− a)(t− b)(t− c)(t− d)
=
2√
(a− c)(b− d)
sn−1
(√
(b− d)(t− a)
(a− d)(t− b)
,
√
(b− c)(a− d)
(a− c)(b− d)
)
.
This integral can be used for t = ρ or µ. If we now form sn(2k
√
E(a− c)(b− d)pq(M − N))
and use the addition formula
sn(qu+ pv, κ) =
sn(qu, κ)cn(pv, κ)dn(pv, κ) + sn(pv, κ)cn(qu, κ)dn(qu, κ)
1− κ2sn2(qu, κ)sn2(pv, κ)
we obtain a constant of the motion. Thus computation yields a rational constant of the motion
whenever p/q is a rational number.
Let us restrict ourselves to p = 1, q = 2. Then u = α/2 and v = β where
pα =
√
E
√
(ρ− c)(ρ− d)
ρ
, pβ =
√
E
√
(µ− c)(µ− d)
µ
, Y =
√
b− d
a− d
.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 41
We form
T1 = sn(2u, κ)cn(v, κ)dn(v, κ)
= −2Y
√
(ρ− a)(ρ− b)(ρ− c)(ρ− d)(µ− c)(µ− d)
(µ− b)(ρ2(a+ b− c− d) + 2ρ(cd− ab)− cd(a+ b) + ab(c+ d))
,
T2 = sn(u, κ)cn(2v, κ)dn(2v, κ)
= −Y
√
µ− a
µ− b
(
(c+ b− a− d)ρ2 + 2(ad− bc)ρ+ cb(a+ d)− ad(b+ c)
)
×
(
(d+ b− a− c)ρ2 + 2(ac− bd)ρ+ cb(d− a) + ad(b− c)
)
×
(
(a+ b− c− d)ρ2 + 2(cd− ab)ρ+ ab(c+ d)− cd(a+ b)
)−2
.
We also note that
T3 = κ2sn(2u, κ)2sn(µ, κ)2
= 4
(µ− a)(ρ− a)(ρ− b)(ρ− c)(ρ− d)(b− d)(b− c)
(µ− b) ((a+ b− c− d)ρ2 + 2(cd− ab)ρ+ ab(c+ d)− cd(a+ b))2 ,√
(µ− c)(µ− d) = 4µ
√
(1− µ)(1− k2µ)
pµ√
E
,√
(ρ− c)(ρ− d) = 2ρ
√
(1− ρ)(1− k2ρ)
pρ√
E
,
and ρ = sn2(α, k) and µ = sn2(2β, k). From this we deduce that
√
E
Y sn(4k
√
E(a− c)(b− d)(M−
N), κ) = P , where P is a constant of the motion rational in the momenta. We can calculate the
components of this constant as follows: T1 = n1/d1, where
n1 = −16(1− ρ)
(
1− k2ρ
)
ρµk′
2√
1− µpρpµ,
d1 = (k2µ− 1)
((
k2Λ1 − E(1 + k2)
)
ρ2 + 2
(
Ak2 + E
)
ρ− Λ1 −A(k2 + 1)
)
,
T2 = n2/d2, where
n2 =
√
1− µ
[
−
(
E2k′
4
+ k4
(
Λ2
1 + 4AE
))
ρ4 +
(
−k4
(
− 8AE + 2Λ1E − 2Λ2
1
)
+ k2
(
−8AE − 4Λ1E − Λ2
1
)
+ 2Λ1E
)
ρ3 + 6
(
2k2AE +AE + k2(Λ2
1 +AE)
)
ρ2
+
(
−2k4Λ1A+ 2k2
(
−Λ2
1 + 2Λ1A− 4AE
)
− 2Λ2
1 − 2Λ1A− 8AE
)
ρ
+
(
−k4A2 + 4AE +
(
2k2 − 1
)
A2 − Λ2
1
)]
,
d2 =
√
k2µ− 1
[(
k2Λ1 − E
(
1 + k2
))
ρ2 + 2
(
Ak2 + E
)
ρ− Λ1 −A
(
k2 + 1
)]2
,
T3 = n3/d3, where
n3 = 4(ρ− 1)
(
ρk2 − 1
)
(µ− 1)
(
−Eρ2 + Λ1ρ+A
)(
−E +Ak4 + Λ1k
2
)
,
d3 = (k2µ− 1)
[(
k2Λ1 − E
(
1 + k2
))
ρ2 + 2
(
Ak2 + E
)
ρ− Λ1 −A
(
k2 + 1
)]2
.
We could, of course, use the expressions above given in terms of c+ d and cd.
If we set x = sn2(α, k) = ρ and y = sn2(2β, k) = µ then H and L1 become
H =
1
x− y
(
x(1− x)
(
1− k2x
)
p2
x − 4y(1− y)
(
1− k2y
)
p2
y
)
− A
xy
, (5.3)
L1 =
xy
x− y
(
(1− x)
(
1− k2x
)
p2
x − 4(1− y)
(
1− yk2
)
p2
y
)
−A
(
1
x
+
1
y
)
.
42 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
For A = 0 the extra constant of the motion is G = N/D where
N =
√
y
((
−yxk2 − x2k − x2 + 2x− y
)
p2
x − 4y(y − 1)
(
−1 + k2y
)
p2
y
+ 4x(y − 1)
(
−1 + k2y
)
pxpy
)
,
D =
√
y − 1
((
x2k2 + x2yk2 − x2 − 2xyk2 + y
)
p2
x + 4y(y − 1)
(
−1 + k2y
)
p2
y
− 4y
(
−1 + k2y
)
(x− 1)pxpy
)
.
Setting R = {L1, G} we obtain the polynomial structure relations
R2 − 4
(
1− k′2G2
)(
L1 +G2H − L1G
2
)
= 0,
{G,R} = 2
(
G2 − 1
)(
G2k′
2 − 1
)
,
{L1, R} = −4G
((
2− k2
)
L1 + 2k′
2
HG2 − 2k′
2
L1G
2 −H
)
.
Turning to the case when A 6= 0 we have G = N/D, where
N =
√
y − 1
[
x2y
(
x2yk2 + x2k2 − x2 − 2xyk2 + y
)
p2
x + 4x2y2(y − 1)
(
−1 + k2y
)
p2
y
− 4x2y2
(
−1 + k2y
)
pxpy +A(x− y)2
]
,
D =
√
−1 + k2y
[
x2y
(
−2xy + y − k2x2 + x2yk2 + x2
)
p2
x + 4x2y2(y − 1)
(
−1 + k2y
)
p2
y
− 4x2y2(y − 1)
(
−1 + k2x
)
pxpy +A(x− y)2
]
.
The structure relations are
{G,R} = −2
(
G2 − 1
)(
k′
2
G2 − 1
)
,
{L1, R} = −4G
[
2L1k
2G2 + 2Ak4G2 − 2HG2 − L1 − L1k
2 − 2Ak2 + 2H
]
, (5.4)
R2 + 4
[
Ak4G2 − L1k
2G2 + L1G
4 − 2Ak2G+A− L1G
2 + 2HG2 + L1 −H −H4
]
= 0.
All of the examples that we have constructed using the addition theorem for elliptic functions
obey polynomial structure equations, but we have no proof that this is true in general.
5.2 Example
H =
1
y2 − x2
((
1− x2
)
p2
x − 4
(
1− y2
)
p2
y
)
− A
(1− x2)(1− y2)
,
L1 =
(1− x2)(1− y2)
(y2 − x2)
(
p2
x − 4p2
y
)
+
A(x2 + y2 − 2)
(1− x2)(1− y2)
,
L2 =
(yx4 − 2x2y + y3)p2
x + (−4x2y + 4x2y3)p2
y + (−4x3y2 + 4x3)pxpy −A x2y(x2−y2)2
(1−y2)(1−x2)2
(x4 − 2y2x2 + y2)p2
x + (4x2 − 4x2y2)p2
y + (−4xy + 4xy3)pxpy +A x2(x2−y2)2
(1−y2)(1−x2)2
.
This is the special case of (5.3), (5.4) where we have replaced x, y by 1−x2, 1− y2, respectively,
set k = 0 and rescaled. Note that H = 4p2
y + L1
(1−y2)
+ A
(1−y2)2
, and (1−x2)
(y2−x2)
(
p2
x − 4p2
y
)
=
L1
(1−y2)
− A(x2+y2−2)
(1−x2)(1−y2)2
. We have {H,L1} = {H,L2} = 0, and with R = {L1, L2}, the polynomial
structure relations are
{L1, R} − 16L2L1 + 32HL2 − 32L3
2H = 0, {L2, R} − 8
(
1− L2
2
)
= 0,
R2 + 16L1 − 16L2
2L1 + 32HL2
2 − 16L4
2H − 16H + 16A = 0.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 43
Using the three constants of the motion to eliminate the momenta terms we find
p2
x =
H(x2 − 1)2 + L1(x2 − 1)−A
(1− x2)2
, p2
y =
H(y2 − 1)2 + L1(y2 − 1)−A
4(1− y2)2
,
and the trajectories must satisfy the implicit equation
0 =
(
Hx2(A+ L1 − L1x
2 −Hx4 + 2Hx2 −H)y2 +
1
4
(
−H + L1 +Hx4 +A
)2)
L2
2
+
(
3Hx4L1 −
1
2
H2x8 − 1
2
A2 +AH − 3H2x4 + L1x
2A+ 2H2x6 + 3Hx4A
−AL1 − L1Hx
6 + L1H + 2H2x2 − 1
2
L2
1 + L2
1x
2 − 1
2
H2 − 3L1x
2H − 2Hx2A
)
yL2
+
1
4
(
−H + L1 +Hx4 +A
)2
y2
− x2(A+ L1 −H)
(
A+ L1 − L1x
2 −Hx4 + 2Hx2 −H
)
.
The left-hand side of this equation is quadratic in y, so we can solve for y as a function of x.
We obtain the solutions y = N±2
√
S
D , where
N = L2(−1 + x)4(x+ 1)4H2 − 2L2
(
−2Ax2 − 3L1x
2 − L1x
6 + 3L1x
4 + 3Ax4 +A+ L1
)
H
+ L2(A+ L1)
(
A+ L1 − 2L1x
2
)
,
S = −x2
(
A+ L1 − L1x
2 −Hx4 + 2Hx2 −H
)(
−H + L1 +Hx4 +A
)2R2
16
,
D = −(−1 + x)2(x+ 1)2
(
x2 + 2L2x+ 1
)(
−x2 + 2L2x− 1
)
H2
+
(
4L2
2Ax
2 + 2L1x
4 − 2L1 + 4L2
2L1x
2 − 2A− 4L2
2L1x
4 + 2Ax4
)
H + (A+ L1)2.
Let (x0, y0) be a point on the trajectory such that px0 = 0. A straightforward calculation gives
x2
0 =
2H − L1 ±
√
L2
1 + 4AH
2H
, y0 = −L2,
p2
y0 = −A−H −HL
4
2 + 2HL2
2 + L1 − L1L
2
2
4(L2 − 1)2(L2 + 1)2
.
Assumption: A > 0. Note that
x2
0 = 1 +
(
−L1 ±
√
L2
1 + 4AH
2H
)
. (5.5)
It follows that if we are in the region x > 1 and H > 0, then since
√
L2
1 + 4AH > |L1|, only
the plus sign is possible in (5.5). We conclude that the trajectory must be unbounded to the
right. Examples are Figs. 37 and 38. (The different colors in the figures correspond to different
functions describing the trajectories.)
We investigate what happens if the trajectory intersects the lines y = ±x. If the intersection
occurs for y = x then we find that x = −L2 or px = 2py, or both. If the intersection occurs for
y = −x then we find that x = L2, or px = −2py, or both. In both cases, we have p2
x = 4p2
y =
H(L2
2−1)2+L1(L2
2−1)−A
(L2
2−1)2
. Thus, in both cases if |L2| 6= 1, the trajectory intersects the lines y = ±x
at finite momentum. However, the velocity is infinite. Indeed ẋ = {H,x} = 2px(1−x2)/(y2−x2),
which blows up unless px = 0. This suggests that there are no bounded orbits in the square
−1 < x, y < 1. However we find some bounded orbits in the region x > 1 for negative energies.
Examples of these bounded orbits are Figs. 39 and 40.
44 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
−0.4
−0.3
−0.2
−0.1
0
0.1
0.2
0.3
0.4
0.5
x
y
y+
y−
Figure 37. Case: A = 0.5, L1 = 0.1, L2 = −0.4, H = 1.
1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
−0.4
−0.3
−0.2
−0.1
0
0.1
0.2
0.3
0.4
0.5
x
y
y+
y−
Figure 38. Case: A = 1, L1 = 0.2, L2 = 0, H = 2.
1 1.5 2 2.5 3 3.5
−3.5
−3
−2.5
−2
−1.5
−1
x
y
y+
y−
Figure 39. Case: A = 5, L1 = 10, L2 = 3, H = −1.
1 1.5 2 2.5
1
1.5
2
2.5
x
y
y+
y−
Figure 40. Case: A = 1, L1 = 10, L2 = −2, H = −2.
5.3 Another version of the example
We set y = iz where z is real, multiply H by −1 and L2 by −i to get the system
H =
(1− x2)p2
x + 4(1 + z2)p2
z
x2 + z2
+
A
(1− x2)(1 + z2)
,
Examples of Complete Solvability of 2D Classical Superintegrable Systems 45
L1 = −(1 + z2)(1− x2)(p2
x + 4p2
z)
x2 + z2
+
A(x2 − z2 − 2)
(1 + z2)(1− x2)
,
L2 =
zx4p2
x − 4x3pzpxz
2 − 4x3pzpx − 2x2zp2
x + 4x2p2
zz + 4x2p2
zz
3 − p2
xz
3 − Ax2z(z2+x2)2
(1+z2)(1−x2)2
x4p2
x − 4p2
zx
2 − 4pxzpzx− 4p2
zx
2z2 + 2p2
xz
2x2 − p2
xz
2 − 4xz3pzpx + Ax2(x2+z2)2
(1+z2)(1−x2)2)
.
We will consider this system for A > 0 and resticted to the vertical strip of points (x, z) where
−1 < x < 1. Then H must be > 0. The motion is unbounded, there are no bounded orbits, and
the origin (0, 0) and the boundary lines x = ±1 are repulsive.
The polynomial structure relations are R = {L1, L2} and
{L2, R}+ 8
(
1 + L2
2
)
= 0, {L1, R} − 16L2L1 − 32HL2 − 32L3
2H = 0,
−R2 + 16A+ 16L1 + 16L2
2L1 + 32HL2
2 + 16L4
2H + 16H = 0.
Expressing the momenta in terms of the coordinates we have
p2
x = −(A+H + x4H + L1 − x2L1 − 2Hx2
(1− x2)2
,
p2
z =
1
4
(2Hz2 +Hz4 + L1z
2 +A+H + L1)
(1 + z2)2
.
The equations for the trajectories turn out to be y = N±2
√
S
D , where
N = L2
(
1− x2
)4
H2 + 2L2
(
L1 + 3Ax4 +A− L1x
6 + 3L1x
4 − 2Ax2 − 3L1x
2
)
H
+ L2(L1 +A)
(
L1 +A− 2L1x
2
)
,
S = −x
2
16
(
A+H +Hx4 + L1 − L1x
2 − 2Hx2
)(
A+H + L1 −Hx4
)2
R2,
D =
(
1− x2
)2(
x4 + 4L2
2x
2 + 2x2 + 1
)
H2
+
(
−2Ax4 + 2L1 − 4L2
2L1x
4 + 2A− 2L1x
4 + 4L2
2L1x
2 + 4L2
2Ax
2
)
H + (L1 +A)2.
The points (x0, z0, px0 , pz0) on the trajectory where px0 = 0 satisfy
z0 = −L2, p2
z0 =
1
4
(A+H(1 + L2
2)2 + L1 + L1L
2
2)
(1 + L2
2)2
,
x2
0 = 1 +
L1 ±
√
L2
1 − 4AH
2H
.
Requiring first thatR2 > 0, we see that L1 < 0 is a necessary condition for trajectories (otherwise
S < 0 in the strip −1 < x < 1. If |L1|/2H > 1 then a further necessary condition for S > 0
in a strip is L1 + H + A < 0. If |L1|/2H ≤ 1 then a necessary condition is L2
1 − 4AH > 0.
Examples here are Figs. 41 and 42. Note that here there are 4 different trajectories corresponding
to a single choice of constants of the motion.
Case when R2 = 0. When L1 = −9, L2 = 0, A = 1, H = 8, we have R2 = 0 such that S = 0.
The trajectory in this case is a straight line on the x-axis. The two boundary points where px = 0
for x > 0 are 0 and 0.9354. The velocity of the particle is ẋ = {H,x} = px(1−x2)
x2+z2
. Substituting
z = 0 and the expression for px in terms of x into the equation gives,ẋ =
√
L1+2H
x2
−H Thus
lim
x→0
ẋ =∞, so the particle will go through the origin instead of bouncing back.
46 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
Figure 41. Case: A = 1, L1 = −5, L2 = 3, H = 2.
Figure 42. Case: A = 1, L1 = −10, L2 = −3, H = 10.
A Review of some basic concepts in Hamiltonian mechanics
Hamiltonian mechanics is a reformulation of classical Newtonian mechanics alternate to La-
grangian mechanics. In the Hamiltonian formalism a physical system describing the motion of
a particle at time t involves n generalized coordinates qj(t), and n generalized momenta pj(t).
The phase space of the system is described by points (pj , qj) ∈ R2n. The generalized coordina-
tes q′js and generalized momenta p′js are derived from the Lagrangian formulation. The dynamics
of the system are given by Hamilton’s equations [1, 12]
dqj
dt
= +
∂H
∂pj
,
dpj
dt
= −∂H
∂qj
, (A.1)
where H = H(q1, . . . , qn, p1, . . . , pn, t) is the Hamiltonian, which in this paper corresponds to
the total (time-independent) energy of the system: H = T + V , where T and V are kinetic and
potential energy, respectively. Solutions of these equations give the trajectories of the system.
Examples of Complete Solvability of 2D Classical Superintegrable Systems 47
Here T is a function of q alone while V is a function of q alone. Explicitly,
H =
1
2m
∑
j,k
gjk(q)pjpk + V (q), (A.2)
where gjk is a contravariant metric tensor on some real or complex Riemannian manifold. That
is g−1 = det(gjk) 6= 0, gjk = gkj and the metric on the manifold is given by ds2 =
n∑
j,k=1
gjkdq
jdqk,
where (gjk) is the covariant metric tensor, the matrix inverse to (gjk). Under a local transfor-
mation q′j = fj(q) the contravariant tensor and momenta transform according to
(g′)`h =
∑
j,k
∂q′`
∂qj
∂q′h
∂qk
gjk, p′` =
n∑
j=1
∂qj
∂q′`
pj ,
so H is coordinate independent. Here m is a scaling parameter that can often be interpreted
as the mass of the particle. For the Hamiltonian (A.2) the relation between momenta and
the velocities is pj = m
n∑̀
=1
gj`q̇`, so that T = 1
2m
∑
j,k
gjk(q)pjpk = m
2
n∑
`,h=1
g`h(q)q̇`q̇h. Once the
velocities are given, the momenta are scaled linearly in m. In mechanics the exact value of m may
be important, but for our purposes can be scaled to any nonzero value using the above formulas.
To make direct contact with mechanics we will often set m = 1; for structure calculations we
will usually set m = 1/2.
The Poisson bracket of two arbitrary functions A(p,q), B(p,q) on the phase space is the
function
{A,B}(p,q) =
n∑
j=1
(
∂A
∂pj
∂B
∂qj
− ∂A
∂qj
∂B
∂pj
)
.
The Poisson bracket obeys the following properties, for A, B, C functions on the phase space
and a, b constants,
anti-symmetry: {A,B} = −{B,A},
Bilinearity: {A,mB + nC} = m{A,B}+ n{A, C},
Jacobi identity: {A, {B, C}}+ {B, {C,A}}+ {C, {A,B}} = 0,
Leibniz rule: {A,BC} = {A,B}C + B{A, C},
chain rule: {f(A),B} = f ′(A){A,B}.
In terms of the Kronecker delta δjk, coordinates (q,p) satisfy canonical relations {pj , pk} =
{qj , qk} = 0, and {pj , qk} = δjk. Using the Poisson bracket, we can rewrite Hamilton’s equa-
tions (A.1) as
{H, qj} =
dqj
dt
=
∂H
∂pj
, {H, pj} =
dpj
dt
=
∂H
∂qj
.
For any function F(q,p), its dynamics along a trajectory q(t), p(t) is dF
dt = {H,F}. Thus
F(q,p) will be constant along a trajectory if and only {H,F} = 0. If {H,F} = 0, we say
that F is a constant of the motion.
A.1 Classical integrability
A system with Hamiltonian H is integrable if it admits n constants of the motion L1 = H,
L2, . . . ,Ln that are in involution:
{Lj ,Lk} = 0, 1 ≤ j, k ≤ n, (A.3)
48 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
and are functionally independent in the sense that det
(∂Lj
∂pk
)
6= 0. Suppose H is integrable
with associated constants of the motion Lj . Then by the inverse function theorem we can
solve the n equations Lj(q,p) = cj for the momenta to obtain pk = pk(q, c), k = 1, . . . , n,
where c = (c1, . . . , cn) is a vector of constants. For an integrable system, if a particle with
position q lies on the common intersection of the hypersurfaces Lj = cj for constants cj , then its
momentum p is completely determined. Also, if a particle following a trajectory of an integrable
system lies on the common intersection of the hypersurfaces Lj = cj at time t0, where the Lj
are constants of the motion, then it lies on the same common intersection for all t near t0.
Considering pj(q, c) and using the conditions (A.3) and the chain rule it is straightforward to
verify
∂pj
∂qk
= ∂pk
∂qj
. Therefore, there exists a function u(q, c) such that p` = ∂u
∂q`
, ` = 1, . . . , n. Note
that Pj(q, ∂u∂q) = cj , j = 1, . . . , n, and, in particular, u satisfies the Hamilton–Jacobi equation
H
(
q,
∂u
∂q
)
= E, (A.4)
where E = c1. By construction det( ∂u
∂qj∂ck
) 6= 0, and such a solution of the Hamilton–Jacobi
equation depending nontrivially on n parameters c is called a complete integral. This argument
is reversible: a complete integral of (A.4) determines n constants of the motion in involution,
P1, . . . ,Pn.
Theorem A.1. A system is integrable if and only if (A.4) admits a complete integral.
A powerful method for demonstrating that a system is integrable is to exhibit a complete
integral by using additive separation of variables.
It is a standard result in classical mechanics that for an integrable system one can integrate
Hamilton’s equations and obtain the trajectories [1, 12]. The Hamiltonian formalism is is well
suited to exploiting symmetries of the system and an important tool in laying the framework
for quantum mechanics.
A.2 Classical superintegrability
Let F = (f1(q,p), . . . , fN (q,p)) be a set of N functions defined and locally analytic in some
region of a 2n-dimensional phase space. We say F is functionally independent if the N × 2n
matrix
(
∂f`
∂qj
, ∂f`∂pk
)
has rankN throughout the region. NecessarilyN ≤ 2n. The set is functionally
dependent if the rank is strictly less than N on the region. In this case there is a nonzero analytic
function F of N variables such that F (f1, . . . , fn) = 0 identically on the region. Conversely,
if F exists then the rank of the matrix is < N . We say that the Hamiltonian system H is
(polynomially) integrable if there exists a set of N = n constants of the motion L1 = H, . . . ,Ln,
each polynomial in the momenta globally defined (except possibly for isolated singularities),
that is functionally independent and in involution: {Lj ,Lk}=0 for 1 ≤ j, k ≤ n. A classical
Hamiltonian system in n dimensions is maximally (polynomially) superintegrable if it admits
2n − 1 functionally independent, globally defined constants, polynomial in the momenta (the
maximum number possible [33]). At most n functionally independent constants of the motion
can be in mutual involution [1]. However, several distinct n-subsets of the 2n − 1 polynomial
constants of the motion for a superintegrable system could be in involution. In that case the
system is multi-integrable. An important feature of superintegrable systems is that the orbits
traced out by the trajectories can be determined algebraically, without the need for integration.
Along any trajectory each of the symmetries is constant: Ls = cs, s = 1, . . . , 2n − 1. Each
equation Ls(q,p) = cs determines a (2n − 1)-dimensional hypersurface in the 2n-dimensional
phase space, and the trajectory must lie in that hypersurface. Thus the trajectory lies in the
Examples of Complete Solvability of 2D Classical Superintegrable Systems 49
common intersection of 2n − 1 independent hypersurfaces; hence it must be a curve. Another
important feature of the trajectories is that all bounded orbits are periodic [34].
The polynomial constants of the motion for a system with Hamiltonian H form the (poly-
nomial) symmetry algebra SH of the system, closed under scalar multiplication and addition,
multiplication and the Poisson bracket: Indeed if H is a Hamiltonian with constants of the
motion L, K. Then αL+ βK, LK and {L,K} are also constants of the motion. The n defining
constants of the motion of a polynomially integrable system do not generate a very interesting
symmetry algebra, because all Poisson brackets vanish. However, for the 2n − 1 generators of
a polynomial superintegrable system the brackets cannot all vanish and the symmetry algebra
has nontrivial structure. The degree (or order) O(L) of a polynomial constant of the motion L
is its degree as a polynomial in the momenta. Here H has degree 2. The degree O(Fk) of a set
of generators Fk = {L1, . . . ,Lk}, is the maximum degree of the generators. Let S = SFk be
a symmetry algebra of a Hamiltonian system generated by the set Fk. Clearly, many different
sets F ′k′ can generate the same symmetry algebra. Among all these there will be a set of genera-
tors F 0
k0
for which ` = O(F 0
k0
) is a minimum. Here ` is unique, although F 0
k0
is not. We define
the degree (or order) of S to be `. The theory of 2nd degree superintegrable systems has been
developed in papers such as [8, 16, 17, 18, 19, 22, 28, 37]; all such systems are known as are the
structures of the symmetry algebras, all of which close at degree 6.
Acknowledgement
This work was partially supported by a grant from the Simons Foundation (# 208754 to Willard
Miller, Jr.). We thank Galliano Valent for correcting an error in an earlier draft.
References
[1] Arnold V.I., Mathematical methods of classical mechanics, Graduate Texts in Mathematics, Vol. 60,
Springer-Verlag, New York – Heidelberg, 1978.
[2] Ballesteros Á., Enciso A., Herranz F.J., Ragnisco O., Riglioni D., Quantum mechanics on spaces of non-
constant curvature: the oscillator problem and superintegrability, Ann. Physics 326 (2011), 2053–2073,
arXiv:1102.5494.
[3] Ballesteros Á., Enciso A., Herranz F.J., Ragnisco O., Riglioni D., Superintegrable oscillator and Kepler
systems on spaces of nonconstant curvature via the Stäckel transform, SIGMA 7 (2011), 048, 15 pages,
arXiv:1103.4554.
[4] Ballesteros Á., Herranz F.J., Musso F., The anisotropic oscillator on the 2D sphere and the hyperbolic plane,
Nonlinearity 26 (2013), 971–990, arXiv:1207.0071.
[5] Capel J.J., Kress J.M., Invariant classification of second-order conformally flat superintegrable systems,
J. Phys. A: Math. Theor. 47 (2014), 495202, 33 pages, arXiv:1406.3136.
[6] Chanu C., Degiovanni L., Rastelli G., Superintegrable three-body systems on the line, J. Math. Phys. 49
(2008), 112901, 10 pages, arXiv:0802.1353.
[7] Curtis H.D., Orbital mechanics for engineering students, 3rd ed., Elsevier, 2013.
[8] Daskaloyannis C., Quadratic Poisson algebras of two-dimensional classical superintegrable systems and
quadratic associative algebras of quantum superintegrable systems, J. Math. Phys. 42 (2001), 1100–1119,
math-ph/0003017.
[9] Evans N.W., Superintegrability in classical mechanics, Phys. Rev. A 41 (1990), 5666–5676.
[10] Evans N.W., Verrier P.E., Superintegrability of the caged anisotropic oscillator, J. Math. Phys. 49 (2008),
092902, 10 pages, arXiv:0808.2146.
[11] Fordy A.P., Quantum super-integrable systems as exactly solvable models, SIGMA 3 (2007), 025, 10 pages,
math-ph/0702048.
[12] Goldstein H., Poole C.P., Safko J.L., Classical mechanics, Addison-Wesley Press, Inc., Boston, 2001.
http://dx.doi.org/10.1007/978-1-4757-2063-1
http://dx.doi.org/10.1016/j.aop.2011.03.002
http://arxiv.org/abs/1102.5494
http://dx.doi.org/10.3842/SIGMA.2011.048
http://arxiv.org/abs/1103.4554
http://dx.doi.org/10.1088/0951-7715/26/4/971
http://arxiv.org/abs/1207.0071
http://dx.doi.org/10.1088/1751-8113/47/49/495202
http://arxiv.org/abs/1406.3136
http://dx.doi.org/10.1063/1.3009575
http://arxiv.org/abs/0802.1353
http://dx.doi.org/10.1063/1.1348026
http://arxiv.org/abs/math-ph/0003017
http://dx.doi.org/10.1103/PhysRevA.41.5666
http://dx.doi.org/10.1063/1.2988133
http://arxiv.org/abs/0808.2146
http://dx.doi.org/10.3842/SIGMA.2007.025
http://arxiv.org/abs/math-ph/0702048
50 Y. Chen, E.G. Kalnins, Q. Li and W. Miller Jr.
[13] Granovskii Y.I., Zhedanov A.S., Lutsenko I.M., Quadratic algebras and dynamics in curved space. II. The
Kepler problem, Theoret. and Math. Phys. 91 (1992), 604–612.
[14] Higgs P.W., Dynamical symmetries in a spherical geometry. I, J. Phys. A: Math. Gen. 12 (1979), 309–323.
[15] Inönü E., Wigner E.P., On the contraction of groups and their representations, Proc. Nat. Acad. Sci. USA
39 (1953), 510–524.
[16] Kalnins E.G., Kress J.M., Miller Jr. W., Second order superintegrable systems in conformally flat spaces.
I. Two-dimensional classical structure theory, J. Math. Phys. 46 (2005), 053509, 28 pages.
[17] Kalnins E.G., Kress J.M., Miller Jr. W., Second order superintegrable systems in conformally flat spaces.
II. The classical two-dimensional Stäckel transform, J. Math. Phys. 46 (2005), 053510, 15 pages.
[18] Kalnins E.G., Kress J.M., Miller Jr. W., Second order superintegrable systems in conformally flat spaces.
III. Three-dimensional classical structure theory, J. Math. Phys. 46 (2005), 103507, 28 pages.
[19] Kalnins E.G., Kress J.M., Miller Jr. W., Second-order superintegrable systems in conformally flat spaces.
V. Two- and three-dimensional quantum systems, J. Math. Phys. 47 (2006), 093501, 25 pages.
[20] Kalnins E.G., Kress J.M., Miller Jr. W., Families of classical subgroup separable superintegrable systems,
J. Phys. A: Math. Theor. 43 (2010), 092001, 8 pages, arXiv:0912.3158.
[21] Kalnins E.G., Kress J.M., Miller Jr. W., Tools for verifying classical and quantum superintegrability, SIGMA
6 (2010), 066, 23 pages, arXiv:1006.0864.
[22] Kalnins E.G., Kress J.M., Miller Jr. W., Winternitz P., Superintegrable systems in Darboux spaces, J. Math.
Phys. 44 (2003), 5811–5848, math-ph/0307039.
[23] Kalnins E.G., Kress J.M., Pogosyan G.S., Miller Jr. W., Completeness of superintegrability in two-
dimensional constant-curvature spaces, J. Phys. A: Math. Gen. 34 (2001), 4705–4720, math-ph/0102006.
[24] Kalnins E.G., Miller Jr. W., Post S., Contractions of 2D 2nd order quantum superintegrable systems and the
Askey scheme for hypergeometric orthogonal polynomials, SIGMA 9 (2013), 057, 28 pages, arXiv:1212.4766.
[25] Koenigs G., Sur les géodésiques à intégrales quadratiques, in Darboux G., Lecons sur la théorie générale des
surfaces et les applications geométriques du calcul infinitesimal, Vol. 4, Chelsea, New York, 1972, 368–404.
[26] Kuru S., Negro J., ‘Spectrum generating algebras’ of classical systems: the Kepler–Coulomb potential,
J. Phys. Conf. Ser. 343 (2012), 012063, 5 pages.
[27] Latini D., Ragnisco O., The classical Taub–Nut system: factorization, spectrum generating algebra and
solution to the equations of motion, J. Phys. A: Math. Theor. 48 (2015), 175201, 13 pages, arXiv:1411.3571.
[28] Létourneau P., Vinet L., Superintegrable systems: polynomial algebras and quasi-exactly solvable Hamilto-
nians, Ann. Physics 243 (1995), 144–168.
[29] Maciejewski A.J., Przybylska M., Yoshida H., Necessary conditions for classical super-integrability of a cer-
tain family of potentials in constant curvature spaces, J. Phys. A: Math. Theor. 43 (2010), 382001, 15 pages,
arXiv:1004.3854.
[30] Makarov A.A., Smorodinsky J.A., Valiev K., Winternitz P., A systematic search for nonrelativistic systems
with dynamical symmetries. Part I: The integrals of motion, Nuovo Cimento 52 (1967), 1061–1084.
[31] Marquette I., Classical ladder operators, polynomial Poisson algebras, and classification of superintegrable
systems, J. Math. Phys. 53 (2012), 012901, 12 pages, arXiv:1109.4471.
[32] Mayrand M., Vinet L., Hidden symmetries of two-dimensional systems with local degeneracies, J. Math.
Phys. 33 (1992), 203–212.
[33] Miller Jr. W., Post S., Winternitz P., Classical and quantum superintegrability with applications, J. Phys. A:
Math. Theor. 46 (2013), 423001, 97 pages, arXiv:1309.2694.
[34] Nehorošev N.N., Action-angle variables, and their generalizations, Trans. Moscow Math. Soc. 26 (1972),
180–198.
[35] Olver F.W.J., Lozier D.W., Boisvert R.F., Clark C.W. (Editors), NIST handbook of mathematical func-
tions, U.S. Department of Commerce, National Institute of Standards and Technology, Washington, DC,
Cambridge University Press, Cambridge, 2010.
[36] Post S., Winternitz P., A nonseparable quantum superintegrable system in 2D real Euclidean space,
J. Phys. A: Math. Theor. 44 (2011), 162001, 8 pages, arXiv:1101.5405.
[37] Rodŕıguez M.A., Tempesta P., Winternitz P., Symmetry reduction and superintegrable Hamiltonian systems,
J. Phys. Conf. Ser. 175 (2009), 012013, 8 pages.
http://dx.doi.org/10.1007/BF01017335
http://dx.doi.org/10.1088/0305-4470/12/3/006
http://dx.doi.org/10.1063/1.1897183
http://dx.doi.org/10.1063/1.1894985
http://dx.doi.org/10.1063/1.2037567
http://dx.doi.org/10.1063/1.2337849
http://dx.doi.org/10.1088/1751-8113/43/9/092001
http://arxiv.org/abs/0912.3158
http://dx.doi.org/10.3842/SIGMA.2010.066
http://arxiv.org/abs/1006.0864
http://dx.doi.org/10.1063/1.1619580
http://dx.doi.org/10.1063/1.1619580
http://arxiv.org/abs/math-ph/0307039
http://dx.doi.org/10.1088/0305-4470/34/22/311
http://arxiv.org/abs/math-ph/0102006
http://dx.doi.org/10.3842/SIGMA.2013.057
http://arxiv.org/abs/1212.4766
http://dx.doi.org/10.1088/1742-6596/343/1/012063
http://dx.doi.org/10.1088/1751-8113/48/17/175201
http://arxiv.org/abs/1411.3571
http://dx.doi.org/10.1006/aphy.1995.1094
http://dx.doi.org/10.1088/1751-8113/43/38/382001
http://arxiv.org/abs/1004.3854
http://dx.doi.org/10.1007/BF02755212
http://dx.doi.org/10.1063/1.3676075
http://arxiv.org/abs/1109.4471
http://dx.doi.org/10.1063/1.529946
http://dx.doi.org/10.1063/1.529946
http://dx.doi.org/10.1088/1751-8113/46/42/423001
http://dx.doi.org/10.1088/1751-8113/46/42/423001
http://arxiv.org/abs/1309.2694
http://dx.doi.org/10.1088/1751-8113/44/16/162001
http://arxiv.org/abs/1101.5405
http://dx.doi.org/10.1088/1742-6596/175/1/012013
Examples of Complete Solvability of 2D Classical Superintegrable Systems 51
[38] Tempesta P., Turbiner A.V., Winternitz P., Exact solvability of superintegrable systems, J. Math. Phys. 42
(2001), 4248–4257, hep-th/0011209.
[39] Tempesta P., Winternitz P., Harnad J., Miller W., Pogosyan G., Rodriguez M. (Editors), Superintegra-
bility in classical and quantum systems, CRM Proceedings and Lecture Notes, Vol. 37, Amer. Math. Soc.,
Providence, RI, 2004.
[40] Tremblay F., Turbiner A.V., Winternitz P., An infinite family of solvable and integrable quantum systems
on a plane, J. Phys. A: Math. Theor. 42 (2009), 242001, 10 pages, arXiv:0904.0738.
[41] Tremblay F., Turbiner A.V., Winternitz P., Periodic orbits for an infinite family of classical superintegrable
systems, J. Phys. A: Math. Theor. 43 (2010), 015202, 14 pages, arXiv:0910.0299.
[42] Tsiganov A.V., On maximally superintegrable systems, Regul. Chaotic Dyn. 13 (2008), 178–190,
arXiv:0711.2225.
http://dx.doi.org/10.1063/1.1386927
http://arxiv.org/abs/hep-th/0011209
http://dx.doi.org/10.1088/1751-8113/42/24/242001
http://arxiv.org/abs/0904.0738
http://dx.doi.org/10.1088/1751-8113/43/1/015202
http://arxiv.org/abs/0910.0299
http://dx.doi.org/10.1134/S1560354708030040
http://arxiv.org/abs/0711.2225
1 Introduction
2 Examples of 2D 2nd degree degenerate systems
2.1 The D4(b) oscillator
2.2 Contraction to the Poincaré upper half plane
2.3 Contraction to the D3 oscillator
2.3.1 Embedding D3 in 3D Minkowski space
2.3.2 Contraction of the D3 oscillator to the isotropic oscillator
2.4 The Higgs oscillator S3
2.4.1 Contraction of D4(b) to the Higgs oscillator on the sphere
2.4.2 Contraction of the Higgs to the isotropic oscillator
3 Examples of 2D 2nd degree nondegenerate systems
3.1 The system S7 on the sphere
3.1.1 S7 in polar coordinates
3.1.2 Trajectories
3.2 System E16 in Euclidean space as a contraction of S7
4 The Post–Winternitz superintegrable systems
4.1 The classical system
5 Classical elliptic superintegrable systems
5.1 Elliptic superintegrability
5.2 Example
5.3 Another version of the example
A Review of some basic concepts in Hamiltonian mechanics
A.1 Classical integrability
A.2 Classical superintegrability
References
|