On Superintegrable Systems with a Cubic Integral of Motion
Shang Lin1, 2, Huang Qing1, 2
School of Mathematics, Northwest University, Xi’an 710069, China
Center for Nonlinear Studies, Northwest University, Xi’an710069, China

 

† Corresponding author. E-mail:

Abstract
Abstract

We present a variety of superintegrable systems in polar coordinates possessing a cubic and a quadratic integral of motion, where Noether integrals of kinetic energy are used to build the integrals. In addition, the associated polynomial Poisson algebras with their algebraic dependence relations are exhibited.

PACS: ;02.30.Ik;
1 Introduction

In classical integrable systems, the term complete integrability is well defined. More precisely in n degrees of freedom, a Hamiltonian system can be integrated up to quadrature if there exist n involutive functions, which are functionally independent. When taking one of these functions as the Hamiltonian H and thinking of others as its first integrals, this Hamiltonian is said to be completely integrable in the Liouville sense. Even though the number of independent functions, which are in involution is n at most, maybe additional integrals of the Hamiltonian H exist and they will definitely generate a non-Abelian algebra of integrals of H. In general, these integrals of motion do not yield finite-dimensional Lie algebras, but more complicated algebraic structures, namely, Possion algebra. These systems, with more integrals, are called super-integrable. The maximal number of further independent integrals is . If there are exactly independent first integrals, we say that the Hamiltonian H is maximally superintegrable. The isotropic harmonic oscillator, the Kepler system, and the Calogero-Moser system are the well-known examples. Maximally superintegrable systems have some interesting features. For instance, they can be separable in more than one coordinate system and all bounded classical trajectories are closed.

There are numerous papers on superintegrable systems, both classical and quantum, see Refs. [15] and references therein. In most of these work, it was restricted to quadratic integrals of motion. The quadratic integrals have an intimate connection with the separation of variables in the Hamilton-Jacobi and Schrödinger equations and generally they produce so-called quadratic algebras.[6]

In contrast, much less attention has been devoted to integrable and superintegrable systems with cubic and higher-order integrals of motion. Ten different potentials in a complex Euclidian plane, which admitted a cubic integral of motion are listed in Ref. [7]. Among them, seven are reducible for their cubic integrals can be written as the Poisson commutators of two quadratic integrals.[8] Systematic studies for superintegrable systems with at least one cubic integral was initiated in Ref. [9]. There the coexistence of first- and third-order integrals of motion in two-dimensional classical and quantum mechanics are considered. Corresponding potentials and integrals are identified. The above mentioned paper was followed by a series of publications. In Refs. [1011], superintegrable systems that are separable in Cartesian coordinates and admit a cubic integral in classical and quantum mechanics are presented. Polynomial Poisson algebras for classical superintegrable systems with a cubic integral are given in Ref. [12]. The complete classification of quantum and classical superintegrable systems allowing the separation of variables in polar coordinates and admitting an additional integral of motion of order three in the momentum is performed.[13] In Ref. [14], the conditions for superintegrable systems in two-dimensional Euclidean space admitting separation of variables in an orthogonal coordinate system and a functionally independent third-order integral are studied.

In Ref. [15], families of Hamiltonians of the form

is considered. There differential Galois theory was utilized to determine necessary conditions for complete integrability. When , eight integrable systems were isolated, four of which were superintegrable indeed. For each one among the four superintegrable cases, two first integrals, which were quadratic in momenta, were also exhibited. And it means that these systems are separable in two different coordinate systems. Based on this fact, shortly later these superintegrable systems were studied in more detail.[16] These systems were reproduced by change of coordinates. In addition, the identity of these systems was clarified and the corresponding Poisson algebras were given. Quite recently, superintegrable systems of above form with a position dependent mass were investigated.[17]

Here we deal with the Hamiltonian functions taking the form of

and study the superintegrability of Eq. (1) with one quadratic and one cubic integrals of motion. Unless specified otherwise, . In sec. 2, three geometric first integrals of the kinetic energy, related to the Killing vectors, are derived. Starting from the Killing vectors, cubic integrals are built and corresponding integrable systems are determined in sec. 3, since the metric implied by the kinetic energy of Eq. (1) is flat. And in sec. 4, further restriction of quadratic integrals yields superintegrable systems where the potentials are given explicitly. Section 5 is devoted to the integrability of Eq. (1) with n=2. And the last section contains some remarks and conclusion.

2 First Integrals of Kinetic Energy

Now we recall some known facts on Killing vectors. In a Riemannian manifold , a Killing vector field X is the infinitesimal generator of a symmetry of the metric g. Namely X is a generator of isometries. In geometric terms X satisfies the condition with the Lie derivative . For a manifold of dimension n, the metric possesses linearly independent Killing vectors at most. Note that when the space is either flat or constant curvature, it admits the maximal group of isometries, which is dimension.

For a Riemannian manifold , which corresponds to a 2n-dimensional configuration space, with coordinates (), of a system, g determines a kinetic Lagrangian such that the associated motion is just the geodesic motion with kinetic energy , where gij is the inverse of the metric tensor gij when gij is nonsingular. Note that is generalised position (this can be for example Cartesian coordinates, angles, arc lengths along a curve) and is corresponding generalised momentum. For a metric with isometries, the Killing vectors correspond to functions that are linear in momenta (the so-called Noether integrals) and that Poisson commute with the kinetic energy H0.

Following the standard approach, direct computation shows that the kinetic energy

has three first order integrals

which satisfy the following commutative relations

In view of the algebraic structure, the quadratic Casimir of the Euclidean algebra (3) is . And the corresponding Killing vectors can be written as

and .

Generally speaking, for the kinetic energy, the introduction of a potential will destroy its first integrals. Nevertheless, it is well-known that for a flat metric, the leading-order terms of all second- and higher-order integrals of motion are determined by corresponding Killing tensors built from tensor products of Killing vectors.[18]

3 Existence of a Cubic Integral

Applying the canonical transformation governed by the generating function

on Eqs. (2) and (3) leads to

and , , . In classical mechanics, the most generic cubic integrals of the Hamiltonian

is of the form

By straightforward computation, we find that the terms of odd and even orders of momenta can not exist simultaneously since they will commute independently. For the generality of this form, see Ref. [19]. Hence, we only need to search for cubic integrals of the form

where Aijk are constants.

And thus, for the Hamiltonian (1), we should consider the cubic integrals

where Ki are defined in Eq. (3). Here we restrict ourselves to the case

and construct corresponding integrable and superintegrable systems. The condition gives rise to three integrable systems with the following potentials

where ψ is an arbitrary function of its argument,

here and hereafter c1 is an arbitrary constant.

4 Superintegrable Systems

With integrable systems (4), (5), and (6) allowing one cubic first integrals in hand, we impose a second independent function, which is the second order of motion and Poisson commutes with its Hamiltonians. Here we consider superintegrability of systems (4), (5), and (6) separately.

4.1 Superintegrability of System (4)

(i) Superintegrable Restriction with

Given the second first integrals, the arbitrary function ψ in Eq. (4) is specified up to finite parameters. In the case of

the relation results in

With these potentials, we have and . is a quartic integral and it may be a polynomial combination of H, F1, and F2. However, it is not the case. We try to close the algebra at lowest dimension possible. Now we add two elements K2 and F3, with

According to the nonzero commutation relations

the integrals H, K2, F1, F2, and F3 form a polynomial Poisson algebra. Following the Poisson relations and without using the specific representation, the Casimir functions and can be obtained. Inserting the exact forms of the integrals, we find that

which are the algebraic relations among the five first integrals H, K2, F1, F2, and F3.

According to , the cubic integral F1 can be expressed as the commutator of two quadratic integrals F2 and F3. Consequently, the cubic integral is reducible. And the superintegrable system coincide with system (14) obtained in Ref. [16].

(ii) Superintegrable Restriction with

Assuming system (4) allows the following quadratic integral

leads to

It can be easily seen that . Then we find

which shows that the functions H, F1, F2, K2 generate a Poisson algebra. And of course, they are dependent and satisfy the relation .

(iii) Superintegrable Restriction with

Given a quadratic integral of the form

and presume , we arrive at

It can be verified that the integrals H, K2, F1, F2 give a Possion algebra because of

And for this algebra, the constraint holds.

4.2 Superintegrability of System (5)

Here for the system (5) with

we preselect a specific quadratic integral F2 and verify that the system (5) is superintegrable.

Assume

we have

With this potential, the following commutative relation

hold. As a consequence, H, F1, F2 bring about a Poisson algebra.

4.3 Superintegrability of System (6)

System (6) is superintegrable if the Hamiltonian there admits another independent first integral. Choosing , we have

Moreover . And thus H, F1, F2 generate a Poisson algebra.

5 Integrability of Eq. (1) When

Provided n = 2, the kinetic energy (2) reduces to . Its Killing vectors take the forms of

and they still satisfy the relations

Imposing the cubic first integral on

and solving the equations induced by the involutive relation, we have three integrable systems whose potentials are listed as below:

with the arbitrary function ψ,

We now look for superintegrable systems (8) with the potential given in Eq. (9) and prove that systems (8) corresponding to Eqs. (10) and (11) are superintegrable themselves.

To set and suppose it is admitted by Eq. (9) give

and . Let , it is obvious that

Following these relations, the first integrals K2, F1, F2, F3, H form a Poisson algebra with the algebraic relation and . In addition, choosing yields another superintegrable system where

With above potentials, H, K2, F1, F2 give a Poisson algebra where .

For system (10), it also possesses a quadratic first integral

Since , now we have a Poisson algebra spanned by and F2.

Analogously, the quadratic integral is admitted by the Hamiltonian (8) with (11) and the corresponding Poisson algebra is governed by H, F1, F2, which satisfy .

6 Concluding Remarks

In this paper we dealt with the super integrability of some Hamiltonian systems, namely for 2n-dimensional Hamiltonian equation, there exists more than n first integrals. Such a property is stronger compared with “complete integrability”, namely the solution structure of such systems has fewer arbitrary parameters since the motion is more restricted by the extra first integrals. Here we constructed some superintegrable systems with one quadratic and one cubic integrals of motion, and build up Poisson algebra for each superintegrable Hamiltonian system. Moreover, the algebraic dependence relations to each Poisson algebra are also given.

A generalization of integrable Hamiltonian equations is the study of the integrability of infinite-dimensional equations, namely integrable partial differential equations (also known as soliton equations) including a number of (1+1)- and (2+1)-dimensional equations, in which case an infinite number of independent first integrals being in involution is required in order to guarantee the integrability, see e.g. Adler,[20] Gel’fand and Dikii,[21] Magri,[22] Fuchsteiner and Fokas,[23] and TAH,[24] etc.

The theory of the completely integrability of soliton equations is also applicable to the so-called super soliton equations. We refer the reader to Zhang[2527] and references therein.

Reference
[1] Capel J. J. Kress J. M. J. Phys. A: Math. Theor. 47 2014 495202
[2] Capel J. J. Kress J. M. Post S. SIGMA 11 2015 038
[3] Evans N. W. Phys. Rev. A 41 1990 5666
[4] Kalnins E. G. Miller W. Jr. Hakobyan Ye. M. Pogosyan G. S. J. Math. Phys. 40 1999 2291
[5] Kalnins E. G. Kress J. M. Pogosyan G. S. Miller W. Jr. J. Phys. A: Math. Gen. 34 2001 4705
[6] Létourneau P. Vinet L. Ann. Phys. 243 1995 144
[7] Drach J. Compt. Rend. Acad. Sci. III 200 1935 22
[8] Rañada M. F. J. Math. Phys. 38 1997 4165
[9] Gravel S. Winternitz P. J. Math. Phys. 43 2002 5902
[10] Gravel S. Theor. Math. Phys. 137 2003 1439
[11] Gravel S. J. Math. Phys. 45 2004 1003
[12] Marquette I. Winternitz P. J. Math. Phys. 48 2007 012902
[13] Tremblay F. Winternitz P. J. Phys. A: Math. Theor. 43 2010 175206
[14] Marchesiello A. Post S. Şnobl L. J. Math. Phys. 56 2015 102104
[15] Szuminski W. Maciejewski A. J. Przbylska M. Phys. Lett. A 379 2015 2970
[16] Fordy A. J. Geom. Phys. 115 2017 98
[17] Rañada M. F. Phys. Lett. A 380 2016 2204
[18] Gilmore R. Lie Groups, Lie Algebras and Some of Their Applications New York Wiley 1974
[19] Post S. Winternitz P. Physics 41 2015 463
[20] Adler M. Invent. Math. 50 1979 219
[21] Gel’fand I. M. Dikii L. A. Collected Works New York Springer 1990
[22] Magri F. J. Math. Phys. 19 1978 1156
[23] Fuchssteiner B. Fokas A. S. Physica D 4 1981 47
[24] Tu G. Z. Andrushkiw R. I. Huang X. C. J. Math. Phys. 32 1991 1900
[25] Zhang Y. F. Rui W. J. Rep. On Math. Phys. 75 2015 231
[26] Zhang Y. F. Tam H. Mei J. Q. Z. Naturforsch. A 70 2015 791
[27] Zhang Y. F. Bai Y. Wu L. X. Int. J. Theor. Phys. 55 2016 2837