Complete sets of invariants for dynamical systems that admit a separation of variables

Similar documents
Nondegenerate 2D complex Euclidean superintegrable systems and algebraic varieties

Variable separation and second order superintegrability

Equivalence of superintegrable systems in two dimensions

Nondegenerate three-dimensional complex Euclidean superintegrable systems and algebraic varieties

Superintegrability in a non-conformally-at space

Superintegrable 3D systems in a magnetic field and Cartesian separation of variables

Superintegrability and exactly solvable problems in classical and quantum mechanics

COULOMB SYSTEMS WITH CALOGERO INTERACTION

Models of quadratic quantum algebras and their relation to classical superintegrable systems

Nondegenerate 3D complex Euclidean superintegrable systems and algebraic varieties

Models for the 3D singular isotropic oscillator quadratic algebra

arxiv: v1 [math-ph] 31 Jan 2015

Calogero model and sl(2,r) algebra

arxiv: v1 [math-ph] 8 May 2016

Remarks on Quadratic Hamiltonians in Spaceflight Mechanics

= 0. = q i., q i = E

General N-th Order Integrals of Motion

Warped product of Hamiltonians and extensions of Hamiltonian systems

Liouville integrability of Hamiltonian systems and spacetime symmetry

arxiv: v1 [math-ph] 22 Apr 2018

Complete integrability of geodesic motion in Sasaki-Einstein toric spaces

GEOMETRIC QUANTIZATION

arxiv: v1 [math-ph] 15 Sep 2009

Quantum Mechanics- I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras

Review and Notation (Special relativity)

Modern Geometric Structures and Fields

Theorem 5.3. Let E/F, E = F (u), be a simple field extension. Then u is algebraic if and only if E/F is finite. In this case, [E : F ] = deg f u.

Structure relations for the symmetry algebras of classical and quantum superintegrable systems

Integrability approaches to differential equations

Symmetries in Quantum Physics

The Toda Lattice. Chris Elliott. April 9 th, 2014

arxiv:gr-qc/ v1 7 Nov 2000

REVIEW. Hamilton s principle. based on FW-18. Variational statement of mechanics: (for conservative forces) action Equivalent to Newton s laws!

arxiv: v2 [nlin.si] 27 Jun 2015

The Helically Reduced Wave Equation as a Symmetric Positive System

SEPARABLE COORDINATES FOR THREE-DIMENSIONAL COMPLEX RIEMANNIAN SPACES

Left-invariant Einstein metrics

Symmetries, Groups Theory and Lie Algebras in Physics

arxiv:math-ph/ v1 25 Feb 2002

arxiv:math/ v1 [math.ra] 13 Feb 2004

Two-variable Wilson polynomials and the generic superintegrable system on the 3-sphere

How curvature shapes space

Illustrating Dynamical Symmetries in Classical Mechanics: The Laplace-Runge-Lenz Vector Revisited

General tensors. Three definitions of the term V V. q times. A j 1...j p. k 1...k q

Lagrangian Description for Particle Interpretations of Quantum Mechanics Single-Particle Case

2 Canonical quantization

Normal form for the non linear Schrödinger equation

arxiv: v4 [math-ph] 3 Nov 2015

Physics 106b: Lecture 7 25 January, 2018

LECTURE 1: LINEAR SYMPLECTIC GEOMETRY

A MARSDEN WEINSTEIN REDUCTION THEOREM FOR PRESYMPLECTIC MANIFOLDS

The 3 dimensional Schrödinger Equation

Symmetry Preserving Numerical Methods via Moving Frames

SEMISIMPLE LIE GROUPS

The Spinor Representation

Proper curvature collineations in Bianchi type-ii space-times( )

Supersymmetric Quantum Mechanics and Geometry by Nicholas Mee of Trinity College, Cambridge. October 1989

The Klein-Gordon equation

Summary of Prof. Yau s lecture, Monday, April 2 [with additional references and remarks] (for people who missed the lecture)

Canonical Forms for BiHamiltonian Systems

CONSIDERATION OF COMPACT MINIMAL SURFACES IN 4-DIMENSIONAL FLAT TORI IN TERMS OF DEGENERATE GAUSS MAP

Generalized Stäckel Transform and Reciprocal Transformations for Finite-Dimensional Integrable Systems

TRANSITIVE HOLONOMY GROUP AND RIGIDITY IN NONNEGATIVE CURVATURE. Luis Guijarro and Gerard Walschap

Stress-energy tensor is the most important object in a field theory and have been studied

OBSTRUCTION TO POSITIVE CURVATURE ON HOMOGENEOUS BUNDLES

The harmonic map flow

SELF-SIMILAR PERFECT FLUIDS

Analytical Mechanics for Relativity and Quantum Mechanics

As always, the story begins with Riemann surfaces or just (real) surfaces. (As we have already noted, these are nearly the same thing).

Page 684. Lecture 40: Coordinate Transformations: Time Transformations Date Revised: 2009/02/02 Date Given: 2009/02/02

M3-4-5 A16 Notes for Geometric Mechanics: Oct Nov 2011

The two body problem involves a pair of particles with masses m 1 and m 2 described by a Lagrangian of the form:

Quaternionic Complexes

Special Conformal Invariance

Lecture 10: A (Brief) Introduction to Group Theory (See Chapter 3.13 in Boas, 3rd Edition)

SYMPLECTIC GEOMETRY: LECTURE 5

QUALIFYING EXAMINATION Harvard University Department of Mathematics Tuesday 10 February 2004 (Day 1)

Tangent bundles, vector fields

14 Ordinary and supersingular elliptic curves

Statistical Interpretation

Generalized MICZ-Kepler system, duality, polynomial and deformed oscillator algebras arxiv: v1 [math-ph] 26 Apr 2010

String Theory and The Velo-Zwanziger Problem

Conformal geometry and twistor theory

The Symplectic Camel and Quantum Universal Invariants: the Angel of Geometry versus the Demon of Algebra

LECTURE 16: LIE GROUPS AND THEIR LIE ALGEBRAS. 1. Lie groups

arxiv:hep-th/ v1 2 Feb 1996

Null-curves in R 2,n as flat dynamical systems

An Exactly Solvable 3 Body Problem

2. Lie groups as manifolds. SU(2) and the three-sphere. * version 1.4 *

GLASGOW Paolo Lorenzoni

The Commutation Property of a Stationary, Axisymmetric System

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R

Self-intersections of Closed Parametrized Minimal Surfaces in Generic Riemannian Manifolds

Part II. Geometry and Groups. Year

Variable-separation theory for the null Hamilton Jacobi equation

Poisson Manifolds Bihamiltonian Manifolds Bihamiltonian systems as Integrable systems Bihamiltonian structure as tool to find solutions

February 1, 2005 INTRODUCTION TO p-adic NUMBERS. 1. p-adic Expansions

A Note on Poisson Symmetric Spaces

Transcription:

Complete sets of invariants for dynamical systems that admit a separation of variables. G. Kalnins and J.. Kress Department of athematics, University of Waikato, Hamilton, New Zealand, e.kalnins@waikato.ac.nz and jonathan@math.waikato.ac.nz G. Pogosyan Laboratory of Theoretical Physics, Joint Institute for Nuclear Research, Dubna, oscow Region, 14980, Russia, pogosyan@thsun1.jinr.dubna.su and W. iller, Jr. School of athematics, University of innesota, inneapolis, innesota, 55455, U.S.A., miller@ima.umn.edu February 28, 2002 Consider a classical Hamiltonian Abstract in dimensions consisting of a kinetic energy term plus a potential. If the associated Hamilton-Jacobi equation admits an orthogonal separation of variables, then it is possible to generate algorithmically a canonical basis where, are the other 2nd-order constants of the motion associated with the separable coordinates, and!"#$&, ()+*,-. The.0/21 functions 34563 form a basis for the invariants. We show how to determine for exactly which spaces and potentials the invariant is a polynomial in the original momenta. We shed light on the general question of exactly when the Hamiltonian admits a constant of the motion that is polynomial in the momenta. For 72. we go further and consider all cases 1

" where the Hamilton-Jacobi equation admits a 2nd-order constant of the motion, not necessarily associated with orthogonal separable coordinates, or even separable coordinates at all. In each of these cases we construct an additional constant of the motion. 1 Introduction The quest for integrable systems has a long history. Basically, the question is, given a classical Hamiltonian where, how can one find all the solutions to the Poisson bracket condition! $# & ( )+* (1) where,-./0/,[1]. There is no known comprehensive solution to this problem. However, if the associated Hamilton-Jacobi equation1)3254 89 is additively separable in the orthogonal variables then a complete integral 276 of the equation can be constructed by quadratures and one can find a basis of <; function- ally independent solutions to equation (1). Indeed there is an explicit canonical change of coordinates from the variables /0 with!+> to variables?@0a where B +, B CDCDC&B are the other 2nd-order constants of the motion associated with the orthogonal separable -coordinates, and F F G B B H-*, 0B @I> -. Thus the <; functions CDCDC& B CDCDCKB form a basis for the invariants. ach invariant J can be expressed as a sum of the form " L # 0AN5 (2) see [1]. Numerous examples have been found through this approach, but important problems remain. any of the known interesting dynamical systems have extra constants of the motion which are polynomial in the canonical momenta 0OP CDCDC&;. This often enables global statements to be made about the system in question, e.g., the existence of closed orbits. However, though many interesting results have been obtained, e.g. [2], an algorithmic way of generating all polynomial solutions to (1) is not known. In particular from the -based integrals in (2) it is difficult to tell if is a polynomial in the momenta. In this article we adopt a -based approach to the calculation of the invariants in which the term L take the form L $ L AQ, and we can say in advance for exactly which separable metrics and potentials is a polynomial in the momenta. We give, in principle, a complete solution to this problem. oreover, we show how 2

to characterize each term L in (2) by the Poisson brackets L 0B. [Note Although the term L L AQ always exists, there are cases where it cannot be expressed as L L AQ, i.e., as a function of L alone. These are exactly the cases where L is an ignorable variable, i.e., where the components of the metric tensor in the -coordinates do not depend on L and where, also, the potential doesn t depend on L. However, these special cases where L, and the invariant of which it is a component term always have polynomial dependence (after multiplication by a linear combination of second-order invariants) can be handled separately or by requiring that L depends on a variable with some dependence, such as L L L 0AN treated below.] Of course, the system could admit a polynomial invariant such that 0A HA CDCDC5 are not polynomials. It is a much more difficult problem to classify all such possibilities for polynomial as functions of possibly nonpolynomial. We make some progress is functionally independent, even if KCDCDC& toward the solution of this problem, through the consideration of important examples. These questions of when a system with n second-order constants of the motion (generated by an orthogonal separation of variables) admits additional poynomial constants of the motion are closely related to the concept of superintegrability, [3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16]. For dimension ; in this paper, we go beyond the formulation discussed above and consider all cases where the Hamilton-Jacobi equation admits a 2ndorder constant of the motion, not necessarily associated with orthogonal separable coordinates, or even separable coordinates at all. In each of these cases we construct an additional constant of the motion. 2 Cartesian systems in two dimensions Let us first consider two dimensional uclidean space. In Cartesian coordinates the Hamiltonian has the form 5 If we have separation of variables in Cartesian coordinates the potential must take the form / 5 (3) We immediately observe that there are already two invariants arising from the separation, namely and. Our problem is to calculate a third invariant and determine when it can be chosen to be a polynomial 3

in the canonical momenta. To do this we compute two functions that satisfy the conditions! and (4) These equations can be solved in principle if we know the original functions and. Indeed if we write out the first of these conditions we obtain 7 This equation can be readily solved to give where and. (We consider to be a function of to compute the integral. An arbitrary function / can be added to the integral, but this makes no difference since are invariants.) Once and have been determined, we see that must be an invariant. It is immediately clear that if, where is an integer then is a polynomial in. As examples of this consider 1.. 2.. < It follows from these two examples that the Hamiltonian has in addition to the obvious invariants the additional invariant!("< # $ (5) 4

" From this observation we conclude that all potentials of the form P (6) have the superintegrability property with three functionally independent invariants which are polynomial in and. This includes the known examples corresponding to. If is determined by a polynomial relation of the form # we can go even further. Then the function is always a polynomial in the canonical momentum. As an example consider + (7) The inverse function is and the corresponding function is given by 8, It is clear that all that we have done applies also to potentials that separate in ; dimensions, in Cartesian coordinates. There is only one further Cartesian case for which polynomial invariants can be generated. Let us consider the case when. The corresponding function is given by If ) has the constant of motion this establishes that the Hamiltonian 7 (8) 5 (9) in addition to the constants and. In general this invariant is not polynomial in the canonical momenta. However, if is a fraction for integers then "!! and!!7 will be a rational invariant whose common denominator is a product of powers of 5

and. The numerator is then an additional polynomial invariant, e.g., consider 3 Then F!5) which indicates that, is an additional invariant. In general,! will be a polynomial invariant, functionally independent of and. 3 General two-dimensional separable systems If we extend this problem to the case of orthogonal separable coordinates in a general Riemannian space, we know that the Hamiltonian in a given set of coordinates with a separable potential has the form J (10) and, due to the separability, there is the invariant [17, 18] &$ &$ We can implement the same ansatz as we have done previously by looking for a function which satisfies The condition has the form Assuming that where 5 (11) (12) *, we see that this equation has the solution (We consider to be a function of. An arbitrary function / can be added to the integral, but this makes no difference since 6

and are invariants.) There is a similar condition for the function. The new invariant is. It is straightforward to verify the condition!k (13) Indeed,, @. This implies that the set! is functionally independent. Similarly we can construct functions 5 that satisfy Assuming that solutions ) * for O where Setting (14), we see that these equations have the, we see that, not an invariant, satisfies Let us illustrate what can happen with some examples. 1. We choose parabolic coordinates in uclidean space 0 +* (15),. First consider the parabolic-separable Hamiltonian + (16) We can immediately associate with this the extra invariant If we look for our functions ) 8 and 7, as before we obtain

O O O If we now consider the constant be written in the form where! $ 8 5, we find that it can (17) is an additional invariant quadratic in the canonical momenta [19]. 2. Consider the Hamiltonian in Cartesian coordinates (18) In parabolic coordinates this Hamiltonian has the form The second order invariant associated with this separation is The additional invariant calculated by our method is given by arccosh arccosh! which is clearly transcendental. (19) 3. If we consider the Hamiltonian OF O 7 (20) then using the semihyperbolic coordinates O HJO0 7 and applying our construction, we find O G O O O K thus giving rise to the additional constant O. 8

4. Let s now look at an example of a potential where our construction yields elliptic integrals. We consider the potential IF. If we carry out the construction using parabolic coordinates then the functions and are given by the integrals &, where is the quadratic constant associated with the separation of variables in parabolic coordinates. If we change variables according to, then both and are given by integrals of the form where and & & F There are a variety of ways of evaluating elliptic integrals of this type. We recall that all our considerations are in the complex domain. As an example, we can choose to use the complex equivalent of the integral valid for & K and for which Then if we calculate sn for elliptic functions we obtain sn < snn5 7 5 5 using the addition formulas where is the second quadratic constant associated with this superintegrable system. Because of the various ways of evaluating elliptic integrals there are a number of ways of uncovering the presence of. In analogy with the constructions (5)-(7), we can find Riemannian spaces and potentials with polynomial invariants of arbitrarily high order. Set +B 9 J> (21)

> where B is a polynomial of order ; and >D are constants. Then there exists a function, inverse to B, i.e., B ), such that and >,, where is a polynomial in the momenta. The cartesian coordinate constructions (5)-(7) correspond to the special case +*. The solution of the equation (11) can be understood in a more general context. We have the dual relations @ 5 * (22) (Since and occur only as we will, without loss of generality, set -* in the theoretical developments to follow, and then replace by in the examples.) Thus we have is linear in, i.e., +* The condition that *, leads to the following necessary and sufficient conditions that the function correspond to an invariant on a Riemannian manifold with potential J* * (23) This equation admits an infinite dimensional conformal symmetry group. Indeed if is a solution then N is also a solution, for any nonconstant function. Also, this group contains the subgroup of inhomogeneous affine symmetries if is a solution then so is N,!7 where - are constants, - +* and J Note that the function > satisfies (23), so any function of must also satisfy the requirement. This puts (21) in the proper context. A more general solution is >, where again any function of also satisfies the requirement. quation (23) also occurs in the theory of level sets, used in computational geometry and computer vision, [20], since it describes the family of functions whose level sets are always straight lines in the plane. We have seen that the construction (21) always leads to a polynomial invariant, up to multiplication by a polynomial in and. In fact these are the only polynomial invariants that can be constructed directly from the integration. This follows from 10

> Theorem 1 The function polynomial dependence on with +* is a solution of equation (21) with if and only if it is of the form +B where B is a (nonconstant) polynomial and *. PROOF Let > CDCDC > are constants with be a solution of (21) with and J*. Substituting this expression into (21) and equating the coefficient of on both sides of the resulting expression, we find the condition, so ) >. Now we make the change of variables, where > *. It follows that CDCDC in the new coordinates, and is a solution of +* (24) Substituting the polynomial into (24) and equating coefficients of, we find * or. Using this information, we return to our original expression for the polynomial and make a new change of variables of the form > > (25) where > -*, and8 are chosen such that the transformed coefficient of vanishes. In these variables CDCDC 5 We substitute this expression into (24), and equating coefficients of we find *, so is a polynomial in of order. Proceeding in this fashion to equate coefficients of for!1 CDCDC in order, we find that the first occurence of L in this sequence of equations takes the form L L CDCDC5 L where L is a polynomial of order 3 at most. It follows by induction on that each L is a polynomial in. At his point we have shown that is a polynomial in both and in. Let be the maximal power of that occurs in. If * we are done. Assume. If we use the argument of the first paragraph of this proof with 11

and interchanged, we see that the coefficient of in must take the form with J*. Since is a polynomial in we must have +*. Thus CDCDC Now substitute this expression into (24) and equate coefficients of where ; is maximal. Suppose. The highest power term in is. The highest power term in is, but this is of lower order. The highest power term in is where is the coefficient of in. Here. If then the highest power term is the coefficient of, so. If then +*, so +*. Thus, the only possiblity is, so CDCDC Substituting this expression into the differential equation we see that +*, or. But this is impossible since and the coefficient of must be *. Hence depends only on. There is a similar argument for the case. QD If we limit our search for potentials to a space in which is prescribed, then the general conditions (23) are replaced by +* J* (26) quation (26) admits the complete integral 8 where is the function inverse to. From this one can use standard techniques (method of characteristics, envelopes of solutions) from the theory of quasilinear first order partial differential equations to construct solutions of (26) that satisfy particular initial conditions or that depend on arbitrary functions, [21] (chapter II) or [22] (section 88). Note Standard Hamilton-Jacobi theory gives essentially these same constants of the motion, but from a different viewpoint, [1]. Our expression for, for example, is!( where!, etc. Standard Hamilton-Jacobi theory gives 12

!, etc., whereas in our approach, etc. In both cases the condition (12) is satisfied. Our approach makes it easier in some cases to determine if polynomial invariants exist. It also points out the bracket relations between and the operators defining the separation, e.g., (11). In the standard theory xamples abound of spaces for which these constructions apply. We illustrate this with a family of surfaces in inkowski space!. The surfaces involve a horispherical coordinate and take the form / ) (27) The metric on the surface is where, and we can construct a polynomial invariant for the surface (and for an appropriate added potential) provided that the function has a polynomial inverse function, i.e., where is a polynomial. Clearly and any polynomial will determine a surface with a polynomial invariant. For example, choose. Then we can take ( and. The resulting will be third-order polynomial in and. Similarly, we can determine a potential term with +* such that N is a polynomial in and. Rather than make either of the choices or for the independent variable in (12) we could choose some other function, adapted to the specific problem at hand. For example, let us take @ for some given function, and require. Solving (12) in these variables we find where (28) 5 ) 5 This approach will work even if and are constants; it is guaranteed to yield a polynomial invariant if we require B (29) where B is a polynomial of order ; and the are constants. Then there exists a function, inverse to B, such that 5) 13 5

" " quating coefficients of we find the condition can solve for by quadratures. quating coefficients of term, we obtain expressions for and It follows that and we and the constant 7 is a polynomial in the momenta. 4 Lie form and nonorthogonal separation in two dimensions We know that if a Hamiltonian # - admits a constant of the motion that is quadratic in the momenta # -!+* (30) and if the roots of the determinant - - are distinct, then the eigenforms define new (separable) variables and the Hamiltonian can be written in Liouville form However, it may be that the roots of this determinant are equal. In this case cannot be put into Liouville form, but rather Lie form, which for a suitable choice of variables (non-separable) is The associated quadratic constant of the motion is (31) (32) We now ask the question When the roots of are equal, how can we calculate the third invariant? We are interested in the the same question when a potential is added to the Hamiltonian. These questions can readily be answered. Indeed if 14

B we look for a function equation that is in involution with, we obtain the J* (33) If we solve (31), (32) for and in terms of the variables obtain @ 5 The equation (33) for then has the form and, we J* From this condition a second invariant can be readily obtained in the form (34) We now extend these considerations by considering the possibility of adding a potential. If we do this and have an extra quadratic constant then and have the forms Solving (35) for and gives Then the equation for @ has the form 5 1 H 5 (35) 5 (36) J* This equation can, in principle, be solved directly. In fact for suitable redefinition of the variables, B equation (36) can be put in the form B! that can be solved by the further transformation +B 5 15 J* (37)

Then, provided that (37) reduces to B! )+* +* From this we immediately deduce an extra constant of the motion of the form The equation for has the solution B (38) There is one remaining possibility for a quadratic constant of the motion (30) in two dimensions the constant may be associated with nonorthogonal separation of variables. In two dimensions there is only one case separation in light cone (null) coordinates, []. For this case the Hamiltonian takes the form! and there is a Killing vector, so is a second-order constant of the motion. In addition there is a quadratic constant Thus we have answered the following questions. O 1. If a Hamiltonian with potential admits a quadratic constant of the motion in two dimensions how does one calculate the third constant? 2. A subset of problem 1 is when we require separation only and ask to calculate the third constant. 5 Systems in three dimensions Let us now look at how the orthogonal separation of variable considerations extend to three dimensions. If we have a general separable coordinate system in three dimensions we could take the Hamiltonian to be [18, 23] $5 (39) 16

where 7 / and is the determinant of the Stäckel matrix This automatically gives us two more invariants (40) $ $5 5 (41)! $5 5 (42) We need to find an additional two invariants, such that the five form a functionally independent set. If we look for a function such that then this function satisfies the equation (43) 5 (44) which looks like the form we have been using in two dimensions. There are similar equations for the corresponding functions for O +. For 0 0 K with this has the solution where and (. (Here, we consider to be a function of to compute the integral. We also assume that *.) The corresponding invariant that we can calculate from these three functions is. This is based on the obvious identity 7 )+* 17

Note As in the two-dimensional case, the solution of the equation (44) can be understood in a more general context. We have the dual relations!d 0 7 0!D (45) where *. (Since and occur only as generality, set (+* in the equations immediately following, and then replace we can, without loss of by! in the examples.) Thus we have +* +* The condition that 0 is linear in and, i.e., +*, leads to the following necessary and sufficient conditions that the function correspond to an invariant on a Riemannian manifold with potential * +* * (46) * These equations admit an infinite dimensional conformal symmetry group. Indeed if is a solution then is also a solution, for any nonconstant function. Also, this group contains the subgroup of inhomogeneous affine symmetries if 0 is a solution then so is 7, 7 N 7! where are constants, - +* and J 7 As in the two-dimensional case, the only polynomial functions a very special form. Theorem 2 The function with polynomial dependence on where B +B 0 with is a (nonconstant) polynomial and > *. of are of +* is a solution of equations (46) if and only if it is of the form > > are constants with PROOF The proof is similar to that of Theorem 1. It follows from this theorem and the first two equations (46) that B +B 18

where, the B are polynomials of strict order > in their first arguments and > Furthermore the coefficients of the -st power of their first arguments can be asumed to be zero. Comparing the coefficients of the highest power of in, we see that this coefficient must be of the form >< where now the > are constants. Substituting this into the third equation (46) and equating coefficients of, we see that J*. quating the coefficients of in the B we see that > where the coefficients are constants. Then, substituting this result into the 3rd equation again and comparing coefficients of we see that I*. At this point we have shown that B where B is a polynomial of order exactly in its first argument. The proof that B is independent of its second and third arguments follows exactly as in the last part of the proof of Theorem 1. QD If we limit our search for potentials to a space in which 7 are prescribed, then the general conditions (46) are replaced by +* +* J* (47) From this one can use standard techniques (method of characteristics, envelopes of solutions) from the theory of systems of quasilinear first order partial differential equations to construct solutions of (47) that satisfy particular initial conditions or that depend on arbitrary functions. The invariant also commutes with. Indeed, from the fact that we can verify that (44) implies (48) The corresponding conditions are satisfied by and. Then the fact that 0 +* is implied by the obvious identity 7 19 J*

Finally, from the fact that! we can verify that (44) implies!d (49) The corresponding conditions are satisfied by and. Then the fact that D0 is implied by the identity 7 ) (50) Similarly, we can define a new invariant by requiring that a new function satisfy 7! (51) with analogous conditions for and. For! this has the solution where with. [Note that for to be a polynomial in we must have a. If * this reduces to requiring to be a polynomial in. If * we can replace the variable by 0. Then and our original analysis goes polynomial in with through with replaced by. It is guaranteed to yield a polynomial invariant if we require!jb 7! (52) where B is a polynomial of order ; and the are constants. Then there exists a function, inverse to B, such that! D quating coefficients of we find the condition and we can solve for by quadratures. quating coefficients of, and the constant term, we find ( and 20 7

" " " " It follows that is a polynomial in the momenta.] The corresponding invariant that we can calculate from these three functions. This is based on the obvious identity is Then it follows that 5 D!D J* with analogous results for,. Thus, from the definition of we see that 0. Finally we define a function by requiring with similar conditions for and. For! this has the solution Then it follows that!d 5 (53) with with analogous relations for and. In summary, all brackets between the six functions are zero except that!d!! (54) Thus the mapping $! is canonical. Note Standard Hamilton-Jacobi theory gives exactly these same constants of the motion, from a different viewpoint, [1]. Our expression for, for example, is where theory gives! and 21. Standard Hamilton-Jacobi

!D, whereas in our approach!. In both cases the condition (44) is satisfied. Our approach makes it straightforward to determine exactly when the are polynomials in the momenta. It also points out the bracket relations between the and the operators defining the separation, e.g., (43, 48, 49,51, 53). The generalization to ; dimensions is straightforward. In the standard theory References [1] V.I. Arnold. athematical ethods of Classical echanics. (translated by K. Vogtmann and A. Weinstein) Graduate Texts in athematics, 60, Springer-Verlag, New York, 1978. [2] ax Karlovini and Kjell Rosquist. Third rank Killing tensors in general relativity. The (1+1)-dimensional case; Gen.Rel.Grav. 31, 1271-1294, (1999). [3] N.W.vans. Superintegrability in Classical echanics; Phys.Rev. A 41 (1990) 5666; [4] N.W.vans. Group Theory of the Smorodinsky-Winternitz System; J.ath.Phys. 32, 3369 (1991). [5] N.W.vans. Super-Integrability of the Winternitz System; Phys.Lett. A 147, 483 (1990). [6] S.Wojciechowski. Superintegrability of the Calogero-oser System. Phys. Lett. A 95, 279 (1983); [7] L.P.isenhart. numeration of Potentials for Which One-Particle Schrödinger quations Are Separable; Phys.Rev. 74, 87 (1948). [8] J.Friš, V.androsov, Ya.A.Smorodinsky,.Uhlir and P.Winternitz. On Higher Symmetries in Quantum echanics; Phys.Lett. 16, 354 (1965). [9] J.Friš, Ya.A.Smorodinskii,.Uhlír and P.Winternitz. Symmetry Groups in Classical and Quantum echanics; Sov.J.Nucl.Phys. 4, 444 (1967). [10] A.A.akarov, Ya.A.Smorodinsky, Kh.Valiev and P.Winternitz. A Systematic Search for Nonrelativistic Systems with Dynamical Symmetries; Nuovo Cimento A 52, 1061 (1967). [11] D.Bonatos, C.Daskaloyannis and K.Kokkotas. Deformed Oscillator Algebras for Two-Dimensional Quantum Superintegrable Systems; Phys. Rev. A 50, 3700 (1994). 22

[12] F.Calogero. Solution of a Three-Body Problem in One Dimension; J.ath.Phys. 10, 2191 (1969). [13] A.Cisneros and H.V.cIntosh. Symmetry of the Two-Dimensional Hydrogen Atom; J.ath.Phys. 10, 277 (1969). [14] L.G.ardoyan, G.S.Pogosyan, A.N.Sissakian and V..Ter-Antonyan. lliptic Basis for a Circular Oscillator. Nuovo Cimento, B 88, 43 (1985), Two- Dimensional Hydrogen Atom I. lliptic Bases; Theor.ath.Phys. 61, 1021 (1984); Hidden symmetry, Separation of Variables and Interbasis xpansions in the Two-Dimensional Hydrogen Atom. J.Phys., A 18, 455 (1985). [15] B.Zaslow and..zandler. Two-Dimensional Analog of the Hydrogen Atom. Amer. J.Phys. 35, 1118 (1967). [16] J.Hietarinta. Direct methods for the search of the second invariant. Physics Report, 147, 87-154, (1987). [17] P. Stäckel. Habilitationsschrift, Halle, 1891 [18] L.P. isenhart. Separable Systems of Stäckel; Annals of ath. (2) 35, 284-305 (1934). [19].G.Kalnins, W.iller Jr. and G.S.Pogosyan. Superintegrability and associated polynomial solutions. uclidean space and the sphere in two dimensions; J.ath.Phys. 37, 6439, (1996). [20] J.A. Sethian. Level Set ethods and Fast arching ethods volving Interfaces in Computational Geometry, Fluid echanics, Computer Vision, and aterials Science; (Chapter 1) Cambridge onograph on Applied and Computational athematics, Cambridge University Press, 1999. [21] R. Courant and D. Hilbert. ethods of athematical Physics, Volume II. Interscience Publishers, New York, 1962. [22] D. Zwillinger. Handbook of Differential quations. Academic Press, San Diego, CA, 1989. [23].G. Kalnins and W. iller. Separable coordinates for three-dimensional complex Riemannian spaces; J. Diff. Geometry 14, 221-236 (1979). [24].G. Kalnins and W. iller. Killing tensors and variable separation for Hamilton-Jacobi and Helmholtz equations; SIA J. ath. Anal. 11, 1011-1026 (1980). 23

[25] W.iller, Jr. Symmetry and Separation of Variables. Addison-Wesley Publishing Company, Providence, Rhode Island, 1977. [26].G. Kalnins, Separation of Variables for Riemannian Spaces of Constant Curvature, Pitman, onographs and Surveys in Pure and Applied athematics 28, Longman, ssex, ngland, 1986. [27].G.Kalnins, J. Kress, W.iller Jr. and G.S.Pogosyan. Completeness of superintegrability in two-dimensional constant curvature spaces; J. Phys. A ath Gen. 34, 4705 4720, (2001). 24