2007
DOI: 10.3842/sigma.2007.026
|View full text |Cite
|
Sign up to set email alerts
|

Quantum Deformations and Superintegrable Motions on Spaces with Variable Curvature

Abstract: Abstract. An infinite family of quasi-maximally superintegrable Hamiltonians with a common set of (2N − 3) integrals of the motion is introduced. The integrability properties of all these Hamiltonians are shown to be a consequence of a hidden non-standard quantum sl(2, R) Poisson coalgebra symmetry. As a concrete application, one of this Hamiltonians is shown to generate the geodesic motion on certain manifolds with a non-constant curvature that turns out to be a function of the deformation parameter z. Moreov… Show more

Help me understand this report
View preprint versions

Search citation statements

Order By: Relevance

Paper Sections

Select...
2
1
1
1

Citation Types

0
5
0

Year Published

2008
2008
2023
2023

Publication Types

Select...
4

Relationship

1
3

Authors

Journals

citations
Cited by 4 publications
(5 citation statements)
references
References 42 publications
(99 reference statements)
0
5
0
Order By: Relevance
“…• And, finally, the last possible generalization is to consider Poisson-Hopf algebra deformations of sl(2, R) [28,62,63] which convey an additional quantum deformation parameter q = e z giving rise to a deformed classical Hamiltonian H λ,N . In this case, the deformation parameter z would determine superintegrable perturbations of the initial (underformed) Hamiltonian (6.3).…”
Section: Discussionmentioning
confidence: 99%
“…• And, finally, the last possible generalization is to consider Poisson-Hopf algebra deformations of sl(2, R) [28,62,63] which convey an additional quantum deformation parameter q = e z giving rise to a deformed classical Hamiltonian H λ,N . In this case, the deformation parameter z would determine superintegrable perturbations of the initial (underformed) Hamiltonian (6.3).…”
Section: Discussionmentioning
confidence: 99%
“…and the common integrals (2.2) are just the mth (left and right) coproducts of the Casimir of sl(2, R). This set of (2N − 3) integrals is "universal" for any Hamiltonian function defined by H = H(q 2 , p 2 , q · p) so that this always provides, at least, a quasi-MS system [32,33]. Therefore the Hamiltonians shown in Table 1 are distinguished systems since they have "additional" symmetries.…”
Section: 1mentioning
confidence: 99%
“…Since both of them are central potentials, the angular momentum integrals (2.2) are valid for both cases, that is, S (m) U ≡ S (m) and S U,(m) ≡ S (m) in (1.2). We recall that, in fact, the spherical symmetry of a central potential on E N directly provides such (2N − 3) independent angular momentum integrals, so they characterize a quasi-MS system [32,33]. However what makes rather special the harmonic oscillator and KC systems is the existence of one more independent integral, which is extracted from a new set of integrals that ensure their MS property and is related to the fact that these two systems are the only ones fulfilling the classical Bertand's theorem [34].…”
Section: Harmonic Oscillator and Kepler Potentials On Euclidean Spacementioning
confidence: 99%
“…, N) define two sets of N integrals in involution. Proofs, technical details and further generalizations can be found in [1,2] but, at this point, some remarks concerning the symmetry and superintegrability properties of H (N) are in order.…”
Section: Introductionmentioning
confidence: 99%
“…The Poisson-coalgebraic "dynamical" symmetry underlying all the superintegrable Hamiltonian systems that we shall present in the sequel can be summarized as the following quite general result [1,2]: Let (q, p) = (q 1 , . .…”
Section: Introductionmentioning
confidence: 99%