• No results found

Stackel transform of Lax equations

N/A
N/A
Protected

Academic year: 2021

Share "Stackel transform of Lax equations"

Copied!
16
0
0

Loading.... (view fulltext now)

Full text

(1)

Stackel transform of Lax equations

Maciej Blaszak and Krzysztof Marciniak

The self-archived postprint version of this journal article is available at Linköping

University Institutional Repository (DiVA):

http://urn.kb.se/resolve?urn=urn:nbn:se:liu:diva-167405

N.B.: When citing this work, cite the original publication.

Blaszak, M., Marciniak, K., (2020), Stackel transform of Lax equations, Studies in applied

mathematics (Cambridge). https://doi.org/10.1111/sapm.12315

Original publication available at:

https://doi.org/10.1111/sapm.12315

Copyright: Wiley

(2)

St¨

ackel transform of Lax equations

Maciej B laszak

Faculty of Physics, Division of Mathematical Physics and Computer Modeling,

A. Mickiewicz University, Uniwersytetu Pozna´

nskiego 2, 61-614 Pozna´

n, Poland

blaszakm@amu.edu.pl

Krzysztof Marciniak

Department of Science and Technology

Campus Norrk¨

oping, Link¨

oping University

601-74 Norrk¨

oping, Sweden

krzma@itn.liu.se

May 6, 2020

Abstract

We construct a map between Lax equations for pairs of Liouville integrable Hamiltonian systems related by a mutli-parameter St¨ackel transform. Using this map we construct Lax representation for a wide class of separable systems by applying the multi-parameter St¨ackel transform to Lax equations of suitably chosen systems from a seed class. For a given separable system, we obtain in this way a set of non-equivalent Lax equations parametrized by an arbitrary function of the spectral parameter, as it is in the case of a related seed system.

Keywords and phrases: Hamiltonian systems, completely integrable systems, St¨ackel systems, St¨ackel transform, Lax representation

1

Introduction

The theory of integrable systems is an important part of mathematical physics. In this paper we focus on some class of Liouville integrable (completely integrable) Hamiltonian systems on a 2n-dimensional Poisson manifold (M, π), where π is a Poisson tensor of maximal rank. In particular M = T∗Q is a cotangent bundle of some pseudo-Riemannian configuration space Q. Then, Liouville integrable system is defined by n functionally independent functions (Hamiltonians) hi : M → R on M, such that hi

mutually commute: {hi, hj} = 0 for all i, j = 1, . . . , n, where {·, ·} : C∞(M ) × C∞(M ) → C∞(M ) is

the Poisson bracket defined by the Poisson tensor π through {f, g} = π (df, dg). In consequence, the Hamiltonian equations

dξ dti

= πdhi≡ Xi, i = 1, . . . , n (1.1)

(ξ ∈ M denote throughout the paper a given/arbitrary point of the manifold M ) mutually commute and the system of differential equations (1.1) can be solved in quadratures. The Liouivlle system (1.1) has often a (one or many) Lax representation

Ltj = [Uj, L] , j = 1, . . . , n (1.2) where L = L(x, ξ) and Uj = Uj(x, ξ) are some matrices depending on a spectral parameter x. It means

that Lax equations (1.2) are differential consequences of (1.1) and in some cases can be actually equivalent to (1.1). Lax representations such as (1.2) play an important role in the integrability theory. Given a Lax matrix L we can reconstruct the constants of motion of the system, and the Lax representation Lt = [U, L] of a Hamiltonian system dξdt = πdh often makes it possible to find separation variables for

(3)

this system [12], that is variables in which the system separates in the sense of Hamilton-Jacobi theory. The system dξdt = πdh is then called a separable Liouville system.

If the Liouville integrable system (1.1) depends on some additional parameters, then it is possible to apply to it the St¨ackel transform. Theory of multi-parameter St¨ackel trasforms has been developed in [14, 15, 11, 5, 7] as a generalization of the concept of coupling constant metamorphosis [10, 9]. Any multi-parameter St¨ackel transform maps a Liuouville integrable system into a new Liouville integrale system [6] (see Section 2 for details)

dξ d˜ti

= πd˜hi≡ ˜Xi, i = 1, . . . , n (1.3)

and if the Lax representation (1.2) for the first system is known there arises a question how to find a Lax representation

˜

L˜tj = [ ˜Uj, ˜L], j = 1, . . . , n (1.4) for the transformed system (1.3). Thus, the first aim of this paper is to prove a theorem (Theorem 2) describing how to find the Lax representation (1.4) of the system (1.3) given the Lax representation (1.2) of the original Liouville integrable system (1.1). This is done in Section 3.

A convenient way of constructing a separable Liouville system is through a spectral curve (separation curve). A spectral curve is any relation of the form

ϕ(x, y, c1, . . . , cn) = 0 (1.5)

such that the system of algebraic equations

ϕ(λi, µi, c1, . . . , cn) = 0, i = 1, . . . , n (1.6)

(where (λi, µi) are pairs of canonically conjugate Darboux coordinates (canonical coordinates) for π) is

globally solvable (except possibly for a union of lower dimensional submanifolds) with respect to the parameters cj∈ R. The relations (1.6) are then called separation relations [12] generated by the spectral

curve (1.5). Solving (1.6) with respect to cj yields

cj = hj(λ1, . . . , λn, µ1, . . . , µn), j = 1, . . . , n

where the functions hj : M → R on right hand sides are easily proven to be in involution with respect

to the Poisson bracket. Moreover, the corresponding Hamiltonian systems are separable (in the sense of Hamilton-Jacobi theory) in the Darboux coordinates (λ, µ) = (λ1, . . . , λn, µ1, . . . , µn), which explains

why (1.6) are called separation relations. Thus, each spectral curve (1.5) generates a separable Liouville system (1.1).

The second aim of this paper is to apply Theorem 2 in order to find Lax representation for wide classes of so called St¨ackel systems [13]. More specifically, the aim is to find Lax representation for Liovulle separable systems generated by hyperelliptic spectral curves of the form

n X j=1 ˜ Hjxγj = 1 2f (x)y 2 − σ(x), (1.7)

(here ˜Hj play the role of cj) where γj ∈ N with the normalization γ1> γ2> · · · > γn = 0 and where σ,

f are Laurent polynomials in x. The corresponding separation relations are in this case a system of n linear equations for ˜Hj and their solution yields n commuting, in the sense of canonical Poisson bracket

{λi, µj} = δij, St¨ackel Hamiltonians ˜Hj(λ, µ). Such systems constitute a fairly general subclass in the

class of all possible St¨ackel systems. Each choice of the constants γ = {γ1, · · · , γn} fixes the St¨ackel

matrix Sij = λ γj

i generated by (1.7). Unless we specify the functions σ and f the curve (1.7) yields a

family of St¨ackel systems that we will call a γ-class.

In paper [6] we demonstrated how to generate the Hamiltonians ˜Hi of the γ-class (1.7) through a

multi-parameter St¨ackel transform of Hamiltonians Hi from the class n X j=1 Hjxn−j= 1 2f (x)y 2 − σ(x), (1.8)

(4)

which is a particular γ-class of (1.7) obtained by setting γj = n − j, j = 1, . . . , n in (1.7) (in such a

case the corresponding St¨ackel matrix becomes simply Sij = λn−ji i.e. a Vandermonde matrix). We

call it Benenti class or seed class [1, 2, 3]. Also, in the recent paper [8] the authors found Lax pairs (L(x), Ui(x)), i = 1, . . . , n for equations generated by the Hamiltonians given by (1.8). The Lax matrices

found there are of Mumford class [16]. Combining the ideas of these papers and using Theorem 2 we will find Lax pairs ( ˜L(x), ˜Ui(x)), i = 1, . . . , n for all the Hamiltonian systems of a given γ-class (1.7).

In order to do this we will make an appropriate extension of Hamiltonians from both classes (1.7) and (1.8) by a number of parameters αi and ˜αi so that the extended systems will be related by a

multi-parameter St¨ackel transform. Applying this St¨ackel transform to the Lax pairs (L(x, a), Ui(x, a)) of the

parameter-dependent systems from the seed class, we will obtain the Lax pairs ( ˜L(x, ˜a), ˜Ui(x, ˜a)) of the

parameter-dependent Hamiltonians of γ-class. Finally, by setting all the parameters to zero we will obtain the sought Lax pairs ( ˜L(x), ˜Ui(x)) for (1.7).

The paper is organized as follows. In Section 2 we remind the main concepts of the multi-parameter St¨ackel transform. In Section 3 we prove Theorem 2 describing how Lax pairs transform under the St¨ackel transform. Section 4 is devoted to Hamiltonian systems from the seed class (1.8) and to their Lax representations (L(x), Ui(x)). In Section 5 we remind the St¨ackel transform relating systems from

the seed class (1.8) and these from arbitrary γ-class (1.7). Finally, in Section 6, we apply the results of Section 3 and Section 5 in order to construct the Lax pairs for the St¨ackel systems from γ-classes (1.7). The paper is also furnished with some examples.

2

St¨

ackel transform of integrable Hamiltonian systems

Let us consider a Liouville integrable Hamiltonian system on a 2n-dimensional Poisson manifold (M, π) defined by n Hamiltonians hi: M → R on M, each depending on k ≤ n parameters α1, . . . , αk so that

hi= hi(ξ, α1, . . . , αk), i = 1, . . . , n (2.1)

where ξ ∈ M . From n functions in (2.1) we choose k functions hsi, i = 1, . . . , k, where {s1, . . . , sk} ⊂ {1, . . . , n}. Solving (we assume it is globally possible) the system of equations

hsi(ξ, α1, . . . , αk) = ˜αi, i = 1, . . . , k (2.2) (where ˜αi is another set of k parameters) with respect to αi yields

αi= ˜hsi(ξ, ˜α1, . . . , ˜αk), i = 1, . . . , k, (2.3) where the right-hand sides of these solutions define k new functions ˜hsi on M , each depending on k parameters ˜αi. Let us also define n − k functions ˜hi by substituting ˜hsi instead of αi in hi for all i /∈ {s1, . . . , sk}

˜

hi= hi|α1→˜h

s1,...,αk→˜hsk, i = 1, . . . , n, i /∈ {s1, . . . , sk}. (2.4) The functions ˜hi = ˜hi(ξ, ˜α1, . . . , ˜αk), i = 1, . . . , n, defined through (2.3) and (2.4) are called the

k-parameter St¨ackel transform of the functions (2.1). If we perform again the St¨ackel transform on the functions ˜hi with respect to ˜hsi we will receive back the functions hiin (2.1). One can prove [11, 6] that St¨ackel transform preserves functional independence as well as involutivity with respect to π.

The Hamiltonians hi yield n commuting Hamiltonian systems on M

dξ dti

= πdhi≡ Xi, i = 1, . . . , n (2.5)

depending on k parameters αi, while ˜hi define n commuting systems

dξ d˜ti

= πd˜hi≡ ˜Xi, i = 1, . . . , n (2.6)

(5)

Observe that as soon as we fix the values of both all αi and all ˜αi the relation (2.2) defines the

(2n − k)-dimensional submanifold Mα, ˜α given by (2.2):

Mα, ˜α= {ξ ∈ M : hsi(ξ, α1, . . . , αk) = ˜αi, i = 1, . . . k} (2.7) or equivalently by (2.3) Mα, ˜α= n ξ ∈ M : ˜hsi(ξ, ˜α1, . . . , ˜αk) = αi, i = 1, . . . k o (2.8) Remark 1 Through each point ξ in M there passes infinitely many submanifolds Mα, ˜α. If we fix the

values of all the parameters αi we can for any ξ always find some values of the parameters ˜αi so that

ξ ∈ Mα, ˜α and vice versa, if we fix ˜αi, for any given ξ we can find αi so that ξ ∈ Mα, ˜α.

As it follows from (2.2), (2.3) and (2.4) the following identities are valid on the whole M and for all values of parameters ˜αi:

hsi(ξ, ˜hs1(ξ, ˜α1, . . . , ˜αn), . . . , ˜hsk(ξ, ˜α1, . . . , ˜αn)) ≡ ˜αi, i = 1, . . . , k (2.9) ˜

hi(ξ, ˜α1, . . . , ˜αn) ≡ hi(ξ, ˜hs1(ξ, ˜α1, . . . , ˜αn), . . . , ˜hsk(ξ, ˜α1, . . . , ˜αn)), i = 1, . . . , n, i /∈ {s1, . . . , sk} (2.10) Differentiating (2.9) with respect to ξ we find that on each Mα, ˜α

dhsi= − k X j=1 ∂hsi ∂αj d˜hsj, i = 1, . . . , k (2.11)

while differentiation of (2.10) gives that on Mα, ˜α we have

dhi = d˜hi− k X j=1 ∂hi ∂αj d˜hsj, i = 1, . . . , n, i /∈ {s1, . . . , sk} . (2.12)

The transformations (2.11), (2.12) on Mα, ˜αcan be written in a matrix form as

dh = Ad˜h (2.13)

where we denote dh = (dh1, . . . , dhn)T and d˜h = (d˜h1, . . . , d˜hn)T and where the n × n matrix A is given

by

Aij = δij for j /∈ {s1, . . . , sk}, Aisj = − ∂hi

∂αj

for j = 1, . . . , k (2.14) From the structure of the matrix A it follows that

det A = ± det ∂hsi ∂αj



so that det A 6= 0 due to our assumptions. Thus, the relation (2.13) can be inverted yielding d˜h = A−1dh. The relation (2.13) and its inverse can be used to show the functional independence of ˜hi for all values

of ˜αi from the functional independence of hi for all values of αi. Moreover, the same relations are used

to prove the involutivity of ˜hi from involutivity of hi. See [11, 6] for details.

Since Xi = πdhi and ˜Xi = πd˜hi we obtain from (2.11)-(2.12) that the Hamiltonian vector fields

Xi= πdhi and ˜Xi= πd˜hi are on the appropriate Mα, ˜α related by the following transformation

Xsi = − k X j=1 ∂hsi ∂αj ˜ Xsj, i = 1, . . . , k (2.15) Xi = ˜Xi− k X j=1 ∂hi ∂αj ˜ Xsj, i = 1, . . . , n, i /∈ {s1, . . . , sk} (2.16)

(6)

This means that the Hamiltonian vector fields Xi and ˜Xi span on each Mα, ˜αthe same n-dimensional

distribution and also that the vector fields Xsi and ˜Xsi span on each Mα, ˜α the same k-dimensional subdistribution of the above distribution. The transformation (2.15)-(2.16) on Mα, ˜α can be written in

matrix form as

X = A ˜X (2.17)

where we denote X = (X1, . . . , Xn)T and ˜X = ( ˜X1, . . . , ˜Xn)T and where the n × n matrix A is given

above.

All the vector fields Xiand ˜Xiare naturally tangent to the corresponding Mα, ˜αso that if ξ0∈ Mα, ˜α

then the multi-parameter (simultaneous) solution

ξ = ξ(t1, . . . , tn, ξ0) (2.18)

of all equations in (2.5) starting at ξ0 for t = 0, will always remain in Mα, ˜αand the same is also true for

multi-parameter solutions of (2.6).

The relations (2.15)-(2.16) can be reformulated in the dual language, that of reciprocal (multi-time) transformations. The reciprocal transformation ˜ti= ˜ti(t1, . . . , tn, ξ), i = 1, . . . , n given on Mα, ˜α by

d˜t = ATdt (2.19)

where dt = (dt1, . . . , dtn)T and d˜t = (d˜t1, . . . , d˜tn)T, transforms the k-parameter solutions (2.18) of the

system (2.5) to the k-parameter solutions ˜ξ = ˜ξ(˜t1, . . . , ˜tn, ξ0) of the system (2.6) (with the same initial

condition ξ(0) = ξ0∈ Mα, ˜α) in the sense that for any ξ0∈ Mα, ˜αwe have

˜

ξ(˜t1(t1, . . . , tn, ξ0), . . . , ˜tn(t1, . . . , tn, ξ0), ξ0) = ξ(t1, . . . , tn, ξ0)

for all values of ti sufficiently close to zero.

The transformation (2.19) is well defined since the right-hand side of (2.19) is an exact differential, as it follows from the above construction. It means that it is possible (at least locally) to integrate (2.19) and obtain an explicit transformation ˜ti= ˜ti(t1, . . . , tn, ξ) that takes multi-time (simultaneous) solutions

of all Hamiltonian systems (2.5) to multi-time solutions of all the systems in (2.6).

3

St¨

ackel transform of Lax equations

In the theorem below, we establish a connection between the Lax pairs of the systems related by a multi-parameter St¨ackel transform.

Theorem 2 Suppose that the Liouville integrable system (2.5) has the Lax representation

Ltj = [Uj, L] , j = 1, . . . , n (3.1) where L = L(x, ξ, α) and Uj= Uj(x, ξ, α) are some matrices depending on the spectral parameter x. Then

the Liouville integrable system (2.6) has the Lax representation ˜ L˜tj = [ ˜Uj, ˜L], j = 1, . . . , n (3.2) where ˜ L(x, ξ, ˜α) = L(x, ξ, ˜h(ξ, ˜α)) ˜ Uj(x, ξ, ˜α) = n X i=1 (A−1)ji(ξ, ˜h(ξ, ˜α))Ui(x, ξ, ˜h(ξ, ˜α)) (3.3)

Thus, in order to obtain the Lax matrix ˜L(x, ξ, ˜α) of the system (2.6) it is enough to replace each αi

in L(x, ξ, α) by the corresponding ˜hsi(ξ, ˜α); the same substitutions are performed in (A

−1)

ji(ξ, α) and in

Ui(x, ξ, α) in the second formula in (3.3) in order to obtain ˜Uj(x, ξ, ˜α).

Proof. Fix arbitrary values of the parameters ˜αi and choose a point ξ ∈ M . According to Remark 1

we can then find values of the parameters αi so that ξ ∈ Mα, ˜αand then αi= ˜hsi(ξ, ˜α) for i = 1, . . . , k. Obviously

∂t˜j ˜

(7)

for all i, j and moreover, due to (2.17), we have on Mα, ˜α ∂ ∂˜tj = n X i=1 (A−1)ji ∂ ∂ti , j = 1, . . . , n. (3.5)

In consequence, at the chosen (and thus arbitrary) ξ ∈ M ˜ Lt˜j(x, ξ, ˜α) = L˜tj(x, ξ, ˜h(ξ, ˜α)) (3.4) = L˜tj(x, ξ, α) (3.5) = n X i=1 A−1(ξ, α)jiLti(x, ξ, α) (3.1) = = n X i=1 A−1(ξ, α)ji[Ui(x, ξ, α), L(x, ξ, α)] =h ˜Uj(x, ξ, ˜α), ˜L(x, ξ, ˜α) i

Let us make two comments on the above theorem. Firstly, the Lax pairs (3.1) and (3.2) are understood as differential-algebraic consequences of the systems (2.5) and (2.6) respectively, i.e. we do not require that the Lax pairs (3.1) and (3.2) actually reconstruct the systems themselves (see also Remark 3). Further, this theorem is a global result, not just restricted to some submanifold Mα, ˜α.

4

Hamiltonian systems from the seed class and their Lax

rep-resentations

Let us now consider separable systems generated by separation curves (spectral curves) in the form

n X j=1 Hjxn−j= 1 2f (x)y 2 − σ(x) ≡ F (x, y) (4.1)

Solving the system of n copies of (4.1), with λi and µi substituted for x and y, i = 1, . . . , n, with respect

to Hj we obtain n separable (and thus Liouville integrable) Hamiltonians Hj(λ, µ) and n related vector

fields Xj(λ, µ),

Hj(λ, µ) = Ej(λ, µ) + V (σ)

j (λ), Xj(λ, µ) = πdHj(λ, µ), j = 1, ..., n (4.2)

on the Poisson manifold (M, π), where Ej(λ, µ) represent geodesic part of the Hamiltonians and V (σ) j (λ)

represent the potential part. Throughout the article (λ, µ) = (λ1, . . . , λn, µ1, . . . , µn) denotes Darboux

(canonical) coordinates on (M, π) which are also separation coordinates for all Hj[3]. The functions V (σ) j

are linear combinations of so called basic separable potentials Vj(β), generated by monomials σ(x) = xβ, β ∈ Z. The potentials Vj(β) can be obtained by solving the system of n copies of the equation

xβ+

n

X

j=1

Vj(β)xn−j= 0, β ∈ Z (4.3)

with x replaced by λi, i = 1, . . . , n. Explicitly the functions V (β)

j can be calculated from the formula [5]

V(β)= RβV(0), V(β)= (V1(β), ..., Vn(β))T, (4.4) where R =       −ρ1(λ) 1 0 0 .. . 0 . .. 0 .. . 0 0 1 −ρn(λ) 0 0 0       , V(0)= (0, 0, ..., −1)T, (4.5)

(8)

ρi(λ) = (−1)isi(λ) and si(λ) are elementary symmetric polynomials. Notice that for β = 0, ..., n − 1

Vk(β)= −δk,n−β. (4.6)

It has been proved in [8] that each system (4.2) has a family of Lax representations d

dtk

L(λ) = [Uk(x), L(x)], k = 1, . . . , n, (4.7)

parametrized by nonvanishing smooth functions g(x). The Lax matrix in the Darboux variables (λ, µ), has the form [8]

L(x) = v(x) u(x) w(x) −v(x)



, (4.8)

(thus attaining the form of systems from the Mumford class [16]) while the auxiliary matrices Uk(x) in

(4.7) are of the form

Uk(x) =  Bk(x) u(x)  + , Bk(x) = 1 2 f (x) g(x)  u(x) xn−k+1  + L(x) (4.9) where u(x) ≡ n Y k=1 (x − λk) = n X k=0 ρkxn−k, ρ0≡ 1 (4.10)

(notice that u(λi) = 0), v(x) is a polynomial of degree n−1 determined by the n conditions v(λi) = g(λi)µi

and is therefore given by

v(x) = n X i=1 g(λi)µi Y k6=i x − λk λi− λk = − n−1 X k=0 " n X i=1 ∂ρn−k ∂λi g(λi)µi ∆i # xk (4.11) while w(x) = −2g 2(x) f (x)  F (x, v(x)/g(x)) u(x)  + . (4.12)

The symbol [ ]+ denotes a polynomial part (Laurent polynomial part) of the quotient, i.e. if P (x) is a

polynomial or a Laurent polynomial and if Q(x) is a polynomial then: P (x) Q(x) =  P (x) Q(x)  + +R(x) Q(x) (4.13)

where R(x) is a reminder of the quotient, so deg R < deg Q (see [8] for details). In particular, for positive basic separable potentials Vj(n+s), s ∈ N we have

 xn+s u(x)  + = − s X r=0 V1(n+r−1)xs−r= − s X r=0 V1(n+s−r−1)xr (4.14)

while for basic negative separable potentials Vj(−s), s ∈ N

 x−s u(x)  + = s X r=1 V1(−r)x−s+r−1= s X r=1 V1(−s+r−1)x−r. (4.15)

Notice also that the function w(x) in (4.12) splits into kinetic part wE(x) and potential part wV(x)

respectively: w(x) = wE(x) + wV(x) = − g2(x) f (x)  f (x)v2(x) u(x)g2(x)  + + 2g 2(x) f (x)  σ(x) u(x)  + . (4.16)

(9)

Remark 3 The Lax matrix L(x) (4.8) reconstructs the separation curve (4.1) in the sense that [8] 0 = det [L(x) − g(x)yI] = −2g 2(x) f (x)   n X j=1 Hjxn−j− 1 2f (x)x 2+ σ(x)  . (4.17)

The Lax matrices for different choices of g(x) are not equivalent.

The Lax pairs (4.7), although given here in the separation coordinates (λ, µ), are invariant with respect to any point transformation on the manifold M = T∗Q. In particular, we will use so called Viet´e coordinates defined as qi= (−1)isi(λ), pi= − n X k=1 λn−ik µk ∆k , ∆i= Y k6=i (λi− λk), i = 1, ..., n. (4.18) In these coordinates [8] u(x; q) = n X k=0 qkxn−k, q0≡ 1 (4.19) and v(x; q, p) = n X k=1   k−1 X s=0 qs   n X j=1 Vj(r+k−s−1)pj    x n−k, r ∈ Z (4.20)

5

St¨

ackel transform of the seed class

In the previous sections we have discussed systems from the seed class, generated by separation curves (4.1), and their Lax pairs (4.7). In this section we will demonstrate, using this knowledge, how to generate systems from γ-classes (1.7), defined by the separation curves of the form

n X j=1 ˜ Hjxγj = 1 2f (x)y 2− σ(x), (5.1)

where γi ∈ N and γ1 > γ2 > · · · > γn = 0. Actually, our goal is to demonstrate how to construct the

St¨ackel systems of a given γ-class (5.1) by applying the multi-parameter St¨ackel transform (2.3)-(2.4) to an appropriate system from the seed class (4.1).

In order to be able to relate the systems from classes (4.1) and (5.1) by a St¨ackel transform, we need to extend both of them to appropriate multi-parameter systems.

Theorem 4 Assume that γ1 > γ2 > · · · > γk > n − 1 and S = {s1, . . . , sk} ⊂ {1, . . . , n − 1} with

s1< · · · < sk. Then, the St¨ackel transform (2.3)-(2.4) transforms the Hamiltonians

hj(λ, µ, α) = Hj(λ, µ) + k X i=1 αiV (γi) j , (5.2)

defined by the separation curve from the seed class

k X j=1 αjxγj + n X j=1 hjxn−j= 1 2f (x)y 2 − σ(x), (5.3) to the Hamiltonians ˜ hj(λ, µ, α) = ˜Hj(λ, µ) + k X i=1 ˜ αiV˜ (n−si) j . (5.4)

(10)

defined by the separation curve from the γ-class γ = {γ1, . . . , γk} ∪ {n − i : i /∈ S} n X j=1 ˜ hjxγj + k X j=1 ˜ αjxn−sj = 1 2f (x)y 2− σ(x), (5.5)

Moreover, the explicit transform between Hamiltonians from both classes takes the form ˜

h = −A−1γ H + A−1γ α,˜ (5.6)

and

h = H + Aγα

where h = (h1, . . . , hn)T, H = (H1, . . . , Hn)T, α = (α1, ..., αk, 0, ..., 0)T and likewise for ˜h, ˜H, ˜α. The

n × n matrix Aγ is given by

(Aγ)ij = V (γj)

i (5.7)

and in particular we obtain the explicit map between H and ˜H: ˜

H = −A−1γ H. (5.8)

This theorem follows from Theorem 3 in [6]. A careful inspection of this theorem reveals that the enumeration of Hamiltonians ˜hi in this theorem has been changed in comparison with the general

con-struction presented in Section 2. We still have that hsi = ˜αi for i = 1, . . . , k but now ˜hi = αi for i = 1, . . . , k (and not ˜hsi = αi as in the general construction) while the remaining transformed

Hamil-tonians ˜hi (for i = k + 1, . . . , n) are obtained by substituting, in the consecutive hj for j /∈ S, all αi,

i = 1, . . . , k, with the corresponding ˜hi, as the general idea of St¨ackel transform stipulates. This is done

in order to obtain a convenient enumeration of ˜hi in the transformed system (5.5).

Notice that due to (5.2) the matrix Aγ is simply the matrix A (2.14) written in the particular settings

of this theorem. Notice also that (5.8) is valid on the whole M , in contrast with the relations (2.13) that are valid only on Mα, ˜α. However, by explanations in Section 2, the solutions of (5.2) and (5.4) are related

only on the appropriate submanifolds Ma, ˜α. Thus, although the St¨ackel transform (5.8) transforms the

parameter-free Liouville-integrable system (4.1) into another parameter-free Liouville integrable system (5.1), the solutions of these systems are not globally related by any reciprocal transformation.

Let us demonstrate the whole procedure on two examples, both involving one-parameter St¨ackel transform. We restrict ourselves to one-parameter examples as the examples involving two parameters lead to large and complicated expressions not very suitable to be presented in a printed form.

Example 5 As a first example, consider the H´enon-Heiles system given by the separation curve H1x + H2=

1 2xy

2+ x4 (5.9)

(so that n = 2, σ(x) = −x4 and f (x) = x) and its one-parameter (so that k = 1) St¨ackel transform

(”one-hole deformation” [3])) with γ1= 2 and with respect to the first Hamiltonian, i.e. S = {1}. Thus

γ = {2, 0} and the extended H´enon-Heiles system is generated by separation curve αx2+ h1x + h2=

1 2xy

2+ x4 (5.10)

The St¨ackel transform (5.6) yields the γ-system generated by ˜

h1x2+ ˜αx + eh2=

1 2xy

2+ x4 (5.11)

Setting ˜α = 0 we obtain the system generated by ˜

H1x2+ ˜H2=

1 2xy

(11)

The matrix Aγ given by (5.7) is Aγ = V1(2) V1(0) V2(2) V2(0) ! =  −λ1− λ2 0 λ1λ2 −1 

and thus the parameter-free St¨ackel transform (5.8) between both systems is ˜ H1= − 1 V1(2) H1, H˜2= H2− V2(2) V1(2) H1.

We will illustrate the structure of the above objects in the flat orthogonal coordinates (x1, x2, y1, y2) of the

system (the reader should not confuse the flat coordinates xi and yi with the spectral parameters x and

y). They are given by

x1= λ1+ λ2, x2= 2p−λ1λ2 (5.13)

with the conjugate momenta given by y1= λ1µ1 λ1− λ2 + λ2µ2 λ2− λ1 , y2=p−λ1λ2  µ1 λ1− λ2 + µ2 λ2− λ1  (5.14) In the flat coordinates the Hamiltonians Hi attain the form

H1= 1 2y 2 1+ 1 2y 2 2+ x 3 1+ 1 2x1x 2 2 H2= 1 2x2y1y2− 1 2x1y 2 2+ 1 16x 4 2+ 1 4x 2 1x 2 2 (5.15)

while their St¨ackel transform becomes ˜ H1= 1 2 1 x1 y21+1 2 1 x1 y22+ x21+1 2x 2 2 ˜ H2= − 1 8 x2 2 x1 y21+ 1 2x 2 1y1y2− 1 8 x2 2 x1 y22− 1 16x 4 2 (5.16)

Example 6 Consider now the one-parameter (again k = 1) St¨ackel transform of the system defined by separation curve (4.1)

H1x2+ H2x + H3=

1 2y

2− x5 (5.17)

(so that n = 3, σ(x) = x5 and f (x) = 1), with γ

1 = 3 and with respect to the second Hamiltonian, i.e.

S = {2}. Thus γ = {3, 2, 0} so the extended system is defined by separation curve (5.3) αx3+ h1x2+ h2x + h3=

1 2y

2− x5 (5.18)

and the St¨ackel transform (5.6) yields the γ-system defined by (5.5) ˜

h1x3+ ˜h2x2+ ˜αx + ˜h3=

1 2y

2− x5 (5.19)

Setting ˜α = 0 we obtain a new system defined by separation curve ˜ H1x3+ ˜H2x2+ ˜H3= 1 2y 2− x5 (5.20) As Aγ =    V1(3) V1(2) V1(0) V2(3) V2(2) V2(0) V3(3) V3(2) −V3(0)   =   −λ1− λ2− λ3 −1 0 λ1λ2+ λ1λ3+ λ2λ3 0 0 −λ1λ2λ3 0 −1  ,

(12)

so the parameter-free St¨ackel transform (5.8) between both systems attains the form: ˜ H1= − 1 V2(3) H2, H˜2= H1− V1(3) V2(3) H2, H˜3= H3− V3(3) V2(3) H2.

As in the previous example, we will explicitly illustrate the structure of both systems in another coordinates. This time we make a point transformation to Viet´e coordinates (4.18). In these coordinates all the Hamiltonians Hi are polynomials [4]. Explicitly

H1= 1 2p 2 2+ q1p2p3+ p1p3+ 1 2q2p 2 3+ q 3 1− 2 q1q2+ q3 H2= q1p1p3+ q1p22+ p1p2+ 1 2( q1q2− q3) p 2 3+ q 2 1p2p3+ q12q2− q1q3− q22 (5.21) H3= 1 2p 2 1+ q2p1p3+ 1 2q 2 1p 2 2+ 2q1p1p2+ 1 2 q 2 2− q1q3 p23+ (q1q2− q3) p2p3+ q12q3− q2q3 while ˜ H1= − 1 q2 p1p2− q1 q2 p1p3− q1 q2 p22−q 2 1 q2 p2p3+ 1 2  q3 q2 − q1  p23− q2 1+ q1q3 q2 + q2 ˜ H2= − q1 q2 p1p2+  1 − q 2 1 q2  p1p3+  1 2 − q2 1 q2  p22+  q1− q3 1 q2  p2p3 +1 2  q3q1 q2 + q2− q12  p23+q3q 2 1 q2 − q1q2+ q3 (5.22) ˜ H3= 1 2p 2 1+  q1− q3 q2  p1p2+  q2− q1q3 q2  p1p3+  1 2q 2 1− q1q3 q2  p22 +  q1q2− q3− q2 1q3 q2  p2p3+  1 2q 2 2+ 1 2 q2 3 q2 − q1q3  p23+q1q 2 3 q2 .

6

Lax representation of γ-classes

In this section we apply the results from Section 3 and Section 5 in order to construct the Lax pairs for the St¨ackel system (5.1) from a given γ-class. In order to do this we start from the Lax pairs (L(x, α), Uj(x, α)) for the extended systems from the seed class (5.3), where

L(x, α) =  v(x) u(x) w(x, α) −v(x)  (6.1) with w(x, α) = w(x) + 2g 2(x) f (x) Xk j=1αj  xγj u(x)  +

and Uj(x, α) are given by (4.9) with L(x) replaced by L(x, α). From Theorem 2 we obtain the following

corollary.

Corollary 7 For any smooth nonvanishing function g(x) the matrices  ˜L(x, ˜α), ˜Uj(x, ˜α)

 given by ˜ L(x, ˜α) =  v(x) u(x) ˜ w(x, ˜α) −v(x)  (6.2) where ˜ w(x, ˜α) = w(x) + 2g 2(x) f (x) Xk j=1 ˜ hj(ξ, ˜α)  xγj u(x)  + and with ˜ Uj(x, ˜α) = n X i=1 A−1γ  ji(ξ)Ui(x, ˜h(ξ, ˜α)).

(13)

The matrices Ui(x, ˜h(ξ, ˜α)) can effectively be calculated by the formulas (4.9) with L(x) replaced

by ˜L(x, ˜α). Notice that the matrix A in the above formula does not depend on α (it still does depend on ξ). Finally, the Lax pairs  ˜L(x), ˜Uj(x)



for the system (5.5) are obtained by letting ˜αi = 0 in

 ˜L(x, ˜α), ˜Uj(x, ˜α) .

Example 8 (Example 5 continued) The H´enon-Heiles system (5.9) has the Lax pairs given by (4.8-4.16). Then for the extended system (5.10) we get

L(x, α) =    v(x) u(x) w(x) + 2αg2x(x)hu(x)x2 i + −v(x)   

(and Uj as given by (4.9)) and thus, for the extended γ-system (5.11) the Lax matrices are

˜ L(x, ˜α) =    v(x) u(x) w(x) + 2˜h1g 2(x) x h x2 u(x) i + −v(x)    and ˜ U1(x) = − 1 V1(2) U1(x, α = ˜h1), U˜2(x) = U2(x, α = ˜h1) − V2(2) V1(2) U1(x, α = ˜h1)

so that the Lax matrices for the transformed system (5.12) (or (5.16)) are

˜ L(x) =    v(x) u(x) w(x) + 2 ˜H1g 2(x) x h x2 u(x) i + −v(x)    and ˜ U1(x) = − 1 V1(2) U1(x, α = ˜H1), U˜2(x) = U2(x, α = ˜H1) − V2(2) V1(2) U1(x, α = ˜H1).

As in Example 5, we present the explicit form of these matrices for a particular function g(x) and in the flat coordinates (x1, y2, y1, y2), given by (5.13)-(5.14). Thus, for the H´enon-Heiles system (5.9) (or

equivalently (5.15)), the Lax matrix L(x) for g(x) = 1 takes the form [8]

L(x) =    2y2 x2x + y1− 2x1y2 x2 x 2− x 1x − 14x22 −2x −4y22 x2 2 + 2x1  +4x1y22 x2 2 −4y1y2 x2 − 2x 2 1−12x 2 2  x−1 −2y2 x2 x − y1+ 2x1y2 x2    while U1(x) =   y2 x2 1 2x −1 −y2 x2  , U2(x) =    y2 x2x − x1y2 x2 + 1 2y1 1 2x 21 2x1x −x −2y22 x2 2 − x1 −yx2 2x + x1y2 x2 − 1 2y1   .

The St¨ackel transform (3.3) of these Lax pairs yields

˜ L(x) =     2y2 x2x − 2x 1y2 x2 − y1  x2− x 1x − 14x22 −2x −2x1+ 4y22 x2 2  +12x2 2+ 4x1y22 x2 2 −4y1y2 x2 + y12 x1 + y22 x1  x−1 −2y2 x2x +  2x1y2 x2 − y1      and ˜ U1(x) =   y2 x1x2 1 2 1 x1x −1 x1 − y2 x1x2  , U˜2(x) =     y2 x2x − x 1y2 x2 − 1 2y1+ 1 4 x2y2 x1  x2− x 1x −x −2y22 x2 2 + x1−14 x2 2 x1  −y2 x2x + x 1y2 x2 − 1 2y1+ 1 4 x2y2 x1      ,

(14)

i.e. the g = 0 Lax representation for the system (5.12) (or (5.16)). In a similar way, the g(x) = x Lax representation for the H´enon-Heiles system takes the form

L(x) =   y1x +12x2y2 x2− x1x − 14x 2 2 −2x3− 2x 1x2− 2x21+ 1 2x 2 2 x + y22 −y1x −12x2y2  , U1(x) =   0 12 −x − 2x1 0  , U2(x) =   1 2y1 1 2x − 1 2x1 −x2− x 1x − x21− 1 2x 2 2 − 1 2y1  ,

while the St¨ackel transform (3.3) gives the g(x) = x Lax representation

˜ L(x) =    y1x +12x2y2 x2− x1x − 14x 2 2 −2x3− 2x 1x2+  1 2x 2 2+ y2 1 x1 + y2 2 x1  x + y22 −y1x −12x2y2   , ˜ U1(x) =   0 12x1 1 −1 x1x − 2 0  , U˜2(x) =    1 2y2 1 2x − 1 2x1− 1 8 x2 2 x1 −x2x 1−14 x2 2 x1  x + 12y21 x1 + y2 2 x1 + x 2 2  −1 2y2   

for the system (5.12) (or (5.16)).

Example 9 (Example 6 continued). The system (5.17) has the Lax pairs given by (4.8) and (4.9) with f (x) = 1; let us now also choose g(x) = 1. In the first step we construct the Lax pairs (L(x, α), Uj(x, α))

for the extended system (5.18). We obtain

L(x, α) =    v(x) u(x) w(x) + αhu(x)2x3i + −v(x)    (6.3)

while Uj(x, α) are then given by (4.9) with L(x, α) given by (6.3). Thus,

˜ L(x) =    v(x) u(x) w(x) + ˜H1 h 2x3 u(x) i + −v(x)    and ˜ U1(x) = − 1 V2(3) U2(x, α = ˜H1), U˜2(x) = U1(x, α = ˜H1) − V1(3) V2(3) U2(x, α = ˜H1), ˜ U3(x) = U3(x, α = ˜H1) − V3(3) V2(3) U2(x, α = ˜H1).

As in Example 6, we present the explicit form of these formulas in the Viet´e coordinates (4.18). For seed system, generated by Hamiltonians (5.21), the g(x) = 1 Lax operator L(x) becomes

L(x) =   −p3x2− (q1p3+ p2) x − q1p2− q2p3− p1 x3+ x2q1+ x q2+ q3 2x2− p2 3+ 2q1 x − q1p23− 2 p2p3+ 2q12− 2q2 p3x2+ (q1p3+ p2) x + q1p2+ q2p3+ p1   while U1(x) =   0 1 2 0 0  , U2(x) =   −1 2p3 x + q1 1 12p3  ,

(15)

U3(x) =   −1 2p3x − 1 2q1p3− 1 2q2 1 2x 2+1 2q1x + 1 2q2 x − 12p23− q1 12p3x +12q1p3+12q2  .

The St¨ackel transform (3.3) of the above Lax pairs yields

˜ L(x) =     −p3x2− (q1p3+ p2) x − q1p2− q2p3− p1 x3+ x2q1+ x q2+ q3 2x2− p2 3+ 2q1 x −q22( q21p2p3+ q1q2p23+ q1p1p3 +q1p22+ q2p2p3−12q2q3p23+ p1p2− q1q3) p3x2+ (q1p3+ p2) x + q1p2+ q2p3+ p1     and ˜ U1(x) =   1 2 p3 q2 − 1 2 1 q2x − 1 2 q1 q2 −1 q2 − 1 2 p3 q2  , U˜2(x) =    1 2 q1p3 q2 − 1 2 q1 q2x + 1 2− 1 2 q21 q2 −q1 q2 − 1 2 q1p3 q2   , ˜ U3(x) =   −1 2p3x − 12q1p3−12p2+12 q3p3 q2 1 2x 2+1 2(q1−qq32)x + 1 2q2−12 q1q3 q2 x − 12p2 3− q1−qq3 2 1 2p3x + 1 2q1p3+ 1 2p2− 1 2 q3p3 q2  ,

i.e. the respective Lax pairs for transformed system generated by Hamiltonians (5.22).

7

Acknowledgments

MB wishes to express his gratitude for Department of Science, Link¨oping, University, Sweden, for their kind hospitality.

References

[1] S. Benenti, Inertia tensors and St¨ackel systems in the Euclidean spaces. Rend. Sem. Mat. Univ. Politec. Torino 50 (1992), 315–341.

[2] S. Benenti, Intrinsic characterization of the variable separation in the Hamilton-Jacobi equation. J. Math. Phys. 38 (1997), 6578–6602.

[3] M. B laszak, Separable systems with quadratic in momenta first integrals, J. Phys. A 38 (2005), 1667–1685.

[4] M. B laszak, A. Sergyeyev, Natural coordinates for a class of Benenti systems, Phys. Lett. A 365 (2007), 28–33.

[5] M. B laszak, A. Sergyeyev, Generalized St¨ackel systems. Phys. Lett. A 375 (2011), 2617–2623. [6] M. B laszak, and K. Marciniak, On reciprocal equivalence of St¨ackel systems, Stud. Appl. Math. 129

(2012), 26–50.

[7] M. B laszak, and K. Marciniak, Classical and quantum superintegrability of St¨ackel systems, SIGMA: Symmetry Integrability Geom. Methods Appl. 13 (2017), paper No. 008.

[8] M. B laszak, and Z. Domanski, Lax representation for separable systems from Benenti class, SIGMA: Symmetry Integrability Geom. Methods Appl. 13 (2019), paper No. 045.

[9] C.P. Boyer, E.G. Kalnins and W.Jr. Miller, St¨ackel-equivalent integrable Hamiltonian systems, SIAM J. Math. Anal., 17 (1986), 778–797.

(16)

[10] J. Hietarinta, B. Grammaticos, B. Dorizzi and A. Ramani, Coupling-Constant Metamorphosis and Duality between Integrable Hamiltonian Systems, Phys. Rev. Lett. 53 (1984), 1707–1710.

[11] A. Sergyeyev, M. B laszak, Generalized St¨ackel transform and reciprocal transformations for finite-dimensional integrable systems, J. Phys. A 41 (2008), no. 10, 105205, 20 pp.

[12] E.K. Sklyanin, Separation of variables—new trends. Progr. Theoret. Phys. Suppl. 118 (1995) 35–60 [13] P. St¨ackel, Die Integration der Hamilton-Jacobischen Differentialgleichung mittelst Separation der

Variablen. Habilitationsschrift, Halle 1891.

[14] A.V. Tsiganov, Canonical transformations of the extended phase space, Toda lattices and the St¨ackel family of integrable systems, J. Phys. A: Math. Gen. 33 (2000), 4169–4182.

[15] A.V. Tsiganov, The Maupertuis Principle and Canonical Transformations of the Extended Phase Space, J. Nonlin. Math. Phys. 8 (2001), 157–182.

[16] P. Vanhaecke, Integrable systems in the realm of algebraic geometry, 2-nd edition. Lecture Notes in Mathematics, 1638. Springer-Verlag, Berlin, 2001.

References

Related documents

Arealerna för detta i övre Klarälvsdalen är fortfarande relativt stora, framförallt om man ställer den mot att antalet lekfiskar som transporteras till lekområdena idag endast

The dominating direc- tions (gradient of image function) in the directional textures (spatial domain) correspond to the large magnitude of frequency

As a preschool teacher in the Swedish preschool context, I have personally experienced challenges regarding the participation and to what extent are children a part in the

Similar to Section VI, the simulations for fading are focus- ing on three aspects, namely the sensing efficiency, the SU throughput and the bandwidth loss. The sensing operations of

var det huvudsakliga motivet till bedömningen att sträckan ej bedömdes lämplig för vare sig laxfiskars reproduktion eller uppväxt.. 7_8 Låg Låg Låg Vattendjupet

In the block marked RATIO record the ratio o f the absolute value of the coefficient o f each term over the number o f variables, both original and y„ in that term....

tural property of M-invariant transformations enable us to perform a rounding error analysis on the weighted QR decomposition and the algorithm that uses this

fångsten som noteras på märkningsgrupper som satts ut i Klarälven beror på att en del lax och öring då kan fångas även i avelsfis- ket, vilket inte är fallet efter