• 検索結果がありません。

The aim of the paper is to announce some recent results con- cerning Hamiltonian theory for higher order variational problems on fibered manifolds

N/A
N/A
Protected

Academic year: 2022

シェア "The aim of the paper is to announce some recent results con- cerning Hamiltonian theory for higher order variational problems on fibered manifolds"

Copied!
21
0
0

読み込み中.... (全文を見る)

全文

(1)

on Differential Geometry, 25–30 July, 2000, Debrecen, Hungary

HAMILTONIAN FIELD THEORY REVISITED: A GEOMETRIC APPROACH TO REGULARITY

OLGA KRUPKOV ´A

Abstract. The aim of the paper is to announce some recent results con- cerning Hamiltonian theory for higher order variational problems on fibered manifolds. A reformulation, generalization and extension of basic concepts such as Hamiltonian system, Hamilton equations, regularity, and Legendre transformation, is presented. The theory is based on the concept ofLepagean (n+ 1)-form(wherenis the dimension of the base manifold). Contrary to the classical approach, where Hamiltonian theory is related to a single Lagrangian, within the present setting a Hamiltonian system is associated with an Euler–

Lagrange form, i.e., with theclass of all equivalent Lagrangians. Hamilton equations are introduced to be equations for integral sections of an exterior differential system, defined by a Lepagean (n+ 1)-form. Relations between extremals and solutions of Hamilton equations are studied in detail. A revi- sion of the concepts ofregularity and Legendre transformation is proposed, reflecting geometric properties of the related exterior differential system. The new look is shown to lead to new regularity conditions and Legendre trans- formation formulas, and provides a procedure ofregularizationof variational problems. Relations to standard Hamilton–De Donder theory, as well as to multisymplectic geometry are studied. Examples of physically interesting La- grangian systems which are traditionally singular, but regular in this revised sense, are discussed.

1. Introduction

Hamiltonian theory belongs to the most important parts of the calculus of vari- ations. The idea goes back to the first half of the 19th century and is due to Sir William Rowan Hamilton and Carl Gustav Jacob Jacobi who, for the case ofclas- sical mechanics, developed a method to pass from the Euler–Lagrange equations to another set of differential equations, now called Hamilton equations, which are

“better adapted” to integration. This celebrated procedure, however, is applicable

1991Mathematics Subject Classification. 35A15, 49L10, 49N60, 58Z05.

Key words and phrases. Lagrangian system, Poincar´e-Cartan form, Lepagean form, Hamil- tonian system, Hamilton extremals, Hamilton–De Donder theory, Hamilton equations, regularity, Legendre transformation.

Research supported by Grants MSM:J10/98:192400002 and VS 96003 of the Czech Ministry of Education, Youth and Sports, and GACR 201/00/0724 of the Czech Grant Agency. The author also wishes to thank Professors L. Kozma and P. Nagy for kind hospitality during the Colloquium on Differential Geometry, Debrecen, July 2000.

187

(2)

only to a certain class of variational problems, called regular. Later the method was formally generalized to higer order mechanics, and both first and higher order field theory, and became one of the constituent parts of the classical variational theory (cf. [8], [4]). In spite of this fact, it has been clear that this generaliza- tion of Hamiltonian theory suffers from a principal defect: allmost all physically interesting field Lagrangians (gravity, Dirac field, electromagnetic field, etc.) are non-regular, hence they cannot be treated within this approach.

Since the second half of the 20’th century, together with an increasing interest to bring the more or less heuristic classical variational theory to a modern framework of differential geometry, an urgent need to understand the geometric meaning of the Hamiltonian theory has been felt, in order to develop its proper generalizations as well as global aspects. There appeared many papers dealing with this task in different ways, with results which are in no means complete: from the most important ones let us mention here at least Goldschmidt and Sternberg [14], Aldaya and Azc´arraga [1], Dedecker [5], [7], Shadwick [37], Krupka [21]–[23], Ferraris and Francaviglia [9], Krupka and ˇStˇep´ankov´a [26], Gotay [16], Garcia and Mu˜noz [10], [11], together with a rather pessimistic Dedecker’s paper [6] summarizing main problems and predicting that a way-out should possibly lead through some new understanding of such fundamental concepts asregularity,Legendre transformation, or even the Hamiltonian theory as such.

The purpose of this paper is to announce some very recent results, partially pre- sented in [30]–[32] and [38], which, in our opinion, open a new way for understanding the Hamiltonian field theory. We work within the framework of Krupka’s theory of Lagrange structures on fibered manifolds where the so called Lepagean form is a central concept ([18], [19], [21], [24], [25]). Inspired by fresh ideas and inter- esting, but, unfortunately, not very wide-spread “nonclassical” results of Dedecker [5] and Krupka and ˇStˇep´ankov´a [26], the present geometric setting means a direct

“field generalization” of the corresponding approach to higher order Hamiltonian mechanics as developed in [27] and [28] (see [29] for review). The key point is the concept of aHamiltonian system, which, contrary to the usual approach, is not re- lated with a single Lagrangian, but rather with an Euler–Lagrange form (i.e., with theclass of equivalent Lagrangians), as well as ofregularity, which is understood to be ageometric property of Hamilton equations. It turns out that “classical” results are incorporated as a special case in this scheme. Moreover, for many variational systems which appear singular within the standard approach, one obtains here a regular Hamiltonian counterpart (Hamiltonian, independent momenta which can be considered a part of certainLegendre coordinates,Hamilton equations equivalent with the Euler–Lagrange equations). This concerns, among others, such important physical systems as, eg., gravity, electromagnetism or the Dirac field, mentioned above.

(3)

2. Notations and preliminaries

All manifolds and mappings throughout the paper are smooth. We use standard notations as, eg.,T for the tangent functor,Jr for ther-jet prolongation functor, dfor the exterior derivative of differential forms,iξ for the contraction by a vector field ξ, and∗for the pull-back.

We consider a fibered manifold (i.e., surjective submersion)π:Y →X, dimX= n, dimY =m+n, its r-jet prolongationπr:JrY →X, r≥1, and canonical jet projectionsπr,k:JrY →JkY, 0≤k < r (with an obvious notationJ0Y =Y). A fibered chart onY is denoted by (V, ψ),ψ= (xi, yσ), the associated chart onJrY by (Vr, ψr),ψr= (xi, yσ, yjσ1, . . . , yjσ1...jr).

A vector field ξ on JrY is called πr-vertical (respectively, πr,k-vertical) if it projects onto the zero vector field onX (respectively, onJkY). We denote byV πr

the distribution onJrY spanned by theπr-vertical vector fields.

Aq-formρonJrY is calledπr,k-projectableif there is aq-form ρ0 onJkY such that πr,kρ0 = ρ. A q-form ρ on JrY is called πr-horizontal (respectively, πr,k- horizontal) ifiξρ= 0 for everyπr-vertical (respectively,πr,k-vertical) vector fieldξ onJrY.

The fibered structure ofY induces a morphism,h, of exterior algebras, defined by the condition Jrγρ=Jr+1γhρ for every sectionγ of π, and called the hor- izontalization. Apparently, horizontalization applied to a function, f, and to the elements of the canonical basis of 1-forms, (dxi, dyσ, dyjσ

1, . . . , dyjσ

1...jr), on JrY gives

hf=f◦πr+1,r, hdxi=dxi, hdyσ =yσldxl, . . . , hdyσj1...jr =yjσ1...jrldxl. A q-form ρon JrY is calledcontactif hρ= 0. On JrY, behind the canonical basis of 1-forms, we have also the basis (dxi, ωσ, ωσj

1, . . . , ωσj

1...jr−1, dyσj

1...jr) adapted to the contact structure, where in place of thedy’s one has the contact 1-forms

ωσ=dyσ−ylσdxl, . . . , ωjσ1...jr−1=dyjσ1...jr−1−yjσ1...jr−1ldxl.

Sections ofJrY which areintegral sections of the contact idealare calledholonomic.

Apparently, a section δ : U → JrY is holonomic if and only if δ = Jrγ where γ:U →Y is a section of π.

Notice that every p-form on JrY, p > n, is contact. Let q > 1. A contact q-form ρ onJrY is called 1-contact if for everyπr-vertical vector field ξ on JrY the (q−1)-formiξρis horizontal. Recurrently, ρis calledi-contact, 2 ≤i ≤q, if iξρis (i−1)-contact. Everyq-form onJrY admits aunique decomposition

πr+1,r ρ=hρ+p1ρ+p2ρ+· · ·+pqρ,

wherepiρ, 1≤i≤q, is ani-contact form onJr+1Y, called thei-contact partofρ.

It is helpful to notice that the chart expression ofpiρin any fibered chart contains exactly iexterior factors ωjσ1...jl wherel is admitted to run from 0 tor. For more details on jet prolongations of fibered manifolds, and the calculus of horizontal and contact forms the reader can consult eg. [18], [19], [24], [25], [29], [33], [34].

(4)

Finally, throughout the paper the following notation is used:

ω0=dx1∧dx2∧...∧dxn, ωi=i∂/∂xiω0, ωij =i∂/∂xjωi, etc.

3. Hamiltonian systems

In this section we discuss the concept of a Hamiltonian system and of a La- grangian system as introduced in [30], and the relation between Hamiltonian and Lagrangian systems.

Let s ≥ 0, and put n = dimX. A closed (n+ 1)-form α on JsY is called a Lepagean (n+ 1)-formifp1αisπs+1,0-horizontal. Ifαis a Lepagean (n+ 1)-form and E =p1αwe also say that αis a Lepagean equivalent of E. By definition, in every fiber chart (V, ψ),ψ= (xi, yσ), onY,

E=Eσωσ∧ω0,

whereEσare functions onVs+1⊂Js+1Y. A Lepagean (n+ 1)-formαonJsY will be also called aHamiltonian system of order s. A sectionδof the fibered manifold πswill be called aHamilton extremal of αif

(3.1) δiξα= 0 for everyπs-vertical vector field ξonJsY . The equations (3.1) will be then calledHamilton equations.

Hamiltonian systems are closely related with Lagrangians and Euler–Lagrange forms. The relation follows from the properties of Lepageann-forms (see eg. [21], [24], [25] for review). Recall that an n-form ρ on JsY is said to be a Lepagean n-form ifhiξdρ = 0 for every πs,0-vertical vector field ξon JsY [18], [21]. Thus, every Lepagean (n+ 1)-form locally equals to dρ where ρis a Lepagean n-form.

Consequently, ifαis a Lepagean (n+ 1)-form then its 1-contact partE is alocally variational form. In other words, there exists an open covering of Js+1Y such that, on each set of this covering, E coincides with the Euler–Lagrange form of a Lagrangian of order r≤s, i.e.,

E=∂L

∂yσ

r

X

l=1

(−1)ldp1dp2. . . dpl

∂L

∂yσp1p2...p

l

ωσ∧ω0.

This suggests the following definition of aLagrangian system: Lepagean (n+ 1)- forms (possibly of different orders) are said to be equivalent if their one-contact parts coincide (up to a possible projection). In what follows, we denote the equiv- alence class of a Lepagean (n+ 1)-formαby [α], and call it aLagrangian system.

The minimum of the set of orders of the elements in the class [α] will then be called the(dynamical) orderof the Lagrangian system [α].

Every Lagrangian system is locally characterized by Lagrangians of all orders starting from a certain minimal one, denoted byr0, and called theminimal order for [α].

TheEuler–Lagrange equationscorresponding to a Lagrangian system [α] of order snow read

(3.2) JsγiJsξα= 0 for every π-vertical vector fieldξonY ,

(5)

whereαis any representative of ordersof the class [α]. Notice that (3.2) are PDE of orders+ 1 for sectionsγof the fibered manifoldπ, their solutions areextremals ofE=p1α.

Let us stop for a moment to discuss relations between Hamiltonian systems and Lagrangian systems.

Every Hamiltonian system α of order s has a unique associated Lagrangian system [α]; its order isr≤s, and it is represented by the locally variational form E =p1α(or, alternatively, by the family of all, generally only local, Lagrangians whose Euler–Lagrange forms locally coincide with E). Comparing the Hamilton equations (3.1) with the Euler–Lagrange equations (3.2) one can see immediately that the sets of extremals and Hamilton extremals generally are not in one-to-one correspondence, in other words,Hamilton equations need not be equivalent with the Euler–Lagrange equations. However, the s-jet prolongation of every extremal is a Hamilton extremal; in this sense there is an inclusion of the set of extremals into the set of Hamilton extremals. More precisely, there is abijectionbetween the set of extremals and the set ofholonomic Hamilton extremals.

On the other hand, a Lagrangian system [α] of order r has many associated Hamiltonian systems, each of an order s ≥ r. Consequently, to a given set of Euler–Lagrange equations one has many sets of Hamilton equations. Behind the given Euler–Lagrange expressions (respectively, a Lagrangian), Hamilton equations depend also upon “free functions” which come from the at least 2-contact part of πs+1,sα. Since α is locally the exterior derivative of a Lepagean n-form, it is convenient to discuss different possibilities on the level of Lepagean equivalents of a corresponding Lagrangian. To this purpose, let us first recall an important result, due to Krupka [25]: ρis a Lepageann-form of ordersiff in any fibered chart (V, ψ), ψ= (xi, yσ), onY,

(3.3) πs+1,sρ=θλ+dν+µ,

where

(3.4) θλ=Lω0+

r−1

X

k=0

r−k−1X

l=0

(−1)ldp1dp2. . . dpl

∂L

∂yjσ

1...jkp1...pli

ωjσ1...j

k∧ωi, ν is an arbitrary contact (n−1)-form, and µ is an arbitrary at least 2-contact n-form; in the formula (3.4),rdenotes the order of hρ. It should be stressed that the decomposition (3.3) is generally not invariant with respect to transformations of fibered coordinates. θλ is called the local Poincar´e–Cartan equivalent of the Lagrangianλ=hρ, each of the (invariant) forms Θ =θλ+p1dν is then called the Poincar´e–Cartan form. Now, let α be a Hamiltonian system. If locally α =dρ where the Lepagean n-form ρ is at most i-contact (1 ≤ i ≤ n) we speak about Hamilton pi-theory and call the corresponding Hamilton equations (3.1)Hamilton pi-equations[30].

(6)

In particular,Hamilton p1-equationsare based upon the Poincar´e–Cartan form Θ. In an usual approach to Hamiltonian field theory only these equations are con- sidered (cf. [1], [3], [7]–[17], [21]–[23], [26], [33], [35]–[37] and references therein);

they are often called Hamilton–De Donder equations. Obviously, behind a La- grangian, they depend upon ν. (With the exception of the case of first order La- grangians whenν = 0, i.e.,θλis invariant, and the Hamilton–De Donder equations are unique and completely determined by the Lagrangian).

Hamilton p2-equationsare based upon a Lepagean form ρ= Θ +ν where ν is 2-contact; for first order Lagrangians they have been studied in [31], [32], a second order generalization is due to [38]. Hamiltonpn-equations are based upon a general Lepageann-form; the first oder case was considered by Dedecker [5].

Each of these Hamiltonian systems can be viewed as adifferent extensionof the original variational problem. In this way, in any concrete situation, one can utilize a possibility to apply additional (geometric or physical) requirements to choose from many alternative Hamiltonian systems the “best one”. A deeper insight into this question is subject of the next sections.

Comments 1. Let us mention main differences between the presented approach and the usual one.

(i) Hamiltonian systems, Hamilton equations. Roughly speaking, there are two main geometric ways approaching Hamiltonian field theory. One, close to the clas- sical calculus of variations, declares the philosophy to assign auniqueHamiltonian system to a singleLagrangian. This task is represented by theHamilton–De Don- der theory, based upon the Poincar´e–Cartan form of a (global) Lagrangianλ, which gives “good” results for first order Lagrangians, and is considered problematic in the higher order case (cf. eg. [6]). The second approach is more or less axiomatic, and is based upon the so called multisymplectic form(eg., [3] and references therein).

(Recall that an (n+ 1)-form Ω on a manifold M is called multisymplectic if it is closed, and the “musical map” ξ →iξΩ, mapping vector fields on M to n-forms, is injective.)

Our approach is close to both of them, but different. This can be seen im- mediately if the definitions of the multisymplectic and Lepagean (n+ 1)-form are compared. For the “zero order” case one gets that every multisymplectic form on Y is a Lepagean (n+ 1)-form, however, a Lepagean (n+ 1)-form need not be multisymplectic. For higher order the difference becomes even sharper, since a multisymplectic form on JsY,s≥1, need not be Lepagean. Apparently, our moti- vation was to define a Hamiltonian system to be, contrary to the multisymplectic definition, sufficiently generalon the one hand (covering allLagrangians without any a priori restriction), and, on the other hand,directly related with a variational problem (defined by a locally variational form). Among others, this also means that our Hamiltonian system is assigned not to a particular Lagrangian (as al- ways done),but to the whole class of all equivalent Lagrangianscorresponding to a given Euler–Lagrange form. Other differences are connected with the concepts of regularityandLegendre transformation, and will be discussed in the next sections.

(7)

(ii) Lagrangian systems. One should notice that also in the definition of a La- grangian system and of the order of a Lagrangian system we differ from other authors. While usually by a “Lagrangian system of order r” one means a global Lagrangian onJrY, by our definition a Lagrangian system is theequivalence class of Lagrangiansgiving rise to an Euler–Lagrange form; the order of a Lagrangian system is then, as a property of the whole equivalence class, determined via the order of the Euler–Lagrange form. In this way, only properties directly connected with dynamics, hence common to all the equivalent Lagrangians, enter in this def- inition, while distinct properties of particular Lagrangians which are not essential for the dynamics are eliminated (the latter are namely connected with the fact that the family of equivalent Lagrangians contains Lagrangians of all orders start- ing from a minimal one, which, as functions, may look completely different from each other, and whose domains usually are open subsets of the corresponding jet prolongations of the underlying fibered manifold; a global Lagrangian often even does not exist at all, obstructions lie in the topology of Y). In the sense of our definition, for example, the Dirac fieldis a Lagrangian system of orderzero (and not of order one); indeed, in this case the class [α] is represented by the global form dθλprojectable ontoY, since the corresponding Lagrangianλis a global first order Lagrangian affine in the first order derivatives. Similarly, theEinstein gravitational field, which is usually defined by the scalar curvature Lagrangian (global second order Lagrangian), is a Lagrangian system of order one, since the corresponding Poincar´e–Cartan form is projectable ontoJ1Y.

4. Regular Hamiltonian systems

A sectionδ ofπs is called aDedecker’s section[30] ifδµ= 0 for every at least 2-contact formµonJsY.

Consider a Hamiltonian system α on JsY. Denote E = p1α, F = p2α, G = πs+1,sα−E−F (i.e.,Gis the at least 3-contact part ofπs+1,sα), and

(4.1) αˆ=E+F.

We shall call the form ˆαtheprincipal part ofα.

A Dedecker’s section which is a Hamilton extremal of πs+1,s α will be called Dedecker–Hamilton extremalofα.

It is easy to obtain the following relation between the sets of extremals, Hamilton extremals, and Dedecker–Hamilton extremals of a Hamiltonian systemαon JsY [30]:

If γ is an extremal ofE = p1α then for every Lepagean equivalent α of E, α defined on JsY (s ≥ 0), the section δ = Jsγ is a Hamilton extremal of α, and δˆ=Js+1γis its Dedecker–Hamilton extremal.

For everyα∈[α], defined onJsY (s≥0), and for every its Dedecker–Hamilton extremalδ, the sectionˆ δ=πs+1,s◦ˆδis a Hamilton extremal of α.

Denote by Dαs and Dαs+1ˆ the family ofn-forms iξαandiξαˆ respectively, where ξ runs over allπs-vertical vector fields onJsY, respectively, over all πs+1-vertical

(8)

vector fields on Js+1Y. Notice that the rank of Ds+1αˆ is never maximal, since i∂/∂yν

Pαˆ = 0 for all multiindicesP of the lenghts+ 1.

Apparently, Hamilton extremals and Dedecker–Hamilton extremals ofαare in- tegral sections of the ideal generated by the family Dαs and Ds+1αˆ , respectively.

Hence, “regularity” can be understood to be aproperty of the idealDs+1αˆ as follows [30]. For convenience, let us consider the casess= 0 and s >0 separately.

We call a Hamiltonian systemαonY regularif rankD1αˆ is constant and equal to rankV π=m. Let s≥1 and r≥1. A Hamiltonian systemα onJsY will be calledregular of degree rif the system of local generators ofDs+1αˆ contains all the n-forms

(4.2) ωσ∧ωi, . . . , ωσj1...jr−1∧ωi, and rankDs+1αˆ =nrankV πr−1+ rankV πs−r.

We refer to (4.2) as localcanonical 1-contactn-forms of orderr.

Roughly speaking, regularity of degreermeans that the systemDs+1αˆ contains all the canonical contactn-forms onJrY, and the rank of the “remaining” subsystem ofDs+1αˆ is the greatest possible one.

Notice that by this definition, every Dedecker–Hamilton extremal of a regular Hamiltonian system with degree of regularityrisholonomic up to the orderr, i.e.,

πs+1,r◦ˆδ=Jrs+1,0◦ˆδ).

Moreover, we have the following main theorem (for the proof we refer to [30]).

Theorem 4.1. [30]Let αbe a Hamiltonian system onJsY,r0 the minimal order forE=p1α. Suppose thatαis regular of degreer0. Then every Dedecker–Hamilton extremalδˆofαis of the form

πs+1,r0◦ˆδ=Jr0γ whereγ is an extremal ofE.

Taking into account this result, a Hamiltonian system of order s≥1 which is regular of degreer0will be called simplyregular. Thus,regularHamilton equations and the corresponding Euler–Lagrange equations arealmost equivalentin the sense that extremals are in bijective correspondence with classes of Dedecker–Hamilton extremals,γ→[Js+1γ], where ˆδ∈[Js+1γ] iffπs+1,r0◦δˆ=Jr0γ.

We shall call a regular Hamiltonian systemstrongly regularif the Hamilton and Euler–Lagrange equations areequivalentin the sense thatextremals are in bijective correspondence with classes of Hamilton extremals,γ →[Jsγ], whereδ∈[Jsγ] iff πs,r0◦δˆ=Jr0γ. (Clearly, fors= 1 this precisely means a bijective correspondence between extremals and Hamilton extremals).

The concept of regularity of aLagrangian systemis now at hand: regularity can be viewed as the property that there exists at least one associated Hamiltonian system which is regular; obviously, the order of this Hamiltonian system may differ from the order of the Lagrangian system. Hence, in accordance with [30], we call a Lagrangian system [α] regular if the family of associated Hamiltonian systems

(9)

contains a regular Hamiltonian system. Similarly, we call a Lagrangian system strongly regularif the family of associated Hamiltonian systems contains a strongly regular Hamiltonian system.

5. Regularity conditions

The above geometric definition of regularity enables one to find explicitregularity conditions. Keeping notations introduced so far, we write ˆα = E +F, where E=Eσωσ∧ω0, and

(5.1) F =

s

X

|J|,|P|=0

FσνJ,P,iωJσ∧ωPν ∧ωi, FσνJ,P,i=−FνσP,J,i;

here J, P are multiindices of the lenght k and l, respectively, J = (j1j2. . . jk), P = (p1p2. . . pl), where 0≤ |J|,|P| ≤s, i.e., 0≤k, l≤s, and, as usual, 1≤i≤n, 1≤σ, ν ≤m. Sincedα= 0,E is an Euler–Lagrange formof order s+ 1, i.e., the functions Eσ satisfy the identities

(5.2) ∂Eσ

∂ypν

1...pl

s+1

X

k=l

(−1)k k

l

dpl+1dpl+2. . . dpk

∂Eν

∂yσp

1...pl

= 0, 0≤l≤s+ 1, calledAnderson–Duchamp–Krupka conditionsfor local variationality ofE[2], [20].

The conditiondα= 0 means that, in particular, (5.3) p2dα=p2dE+p2dF = 0.

In fibered coordinates this equation is equivalent with the following set of identities

(5.4) ∂Eσ

∂yν −∂Eν

∂yσ −diFσν,i= 0,

(5.5)

(Fσν0,S,i)sym(Si)= 1 2

∂Eσ

∂ySiν , (Fσν0,P,i)sym(P i)= 1

2

∂Eσ

∂yP iν −djFσν0,P i,j, 0≤ |P| ≤s−1, and

(5.6) (F

J,S,i

σν )sym(Si)= 0, 1≤ |J| ≤s, (FσνJ j,P,p)sym(P p)+ (FσνJ,P p,j)sym(J j)+diFσνJ j,P p,i= 0, 0≤ |J|,|P| ≤s−1, where sym means symmetrization in the indicated indices, and S = (p1p2. . . ps).

Denote

(5.7) fσνJ,P,i=FσνJ,P,i−(FσνJ,P,i)sym(P i).

(10)

Then from (5.5) we easily get

(5.8) (Fσν,p1...pl−1,pl)sym(p1...pl−1pl)=1 2

s+1−l

X

k=0

(−1)kdi1di2. . . dik

∂Eσ

∂ypν

1...pli1i2...ik

−difσν,p1...pl,i, 1≤l≤s+ 1,

and the recurrent formulas (5.6) give us the remainingF’s expressed by means of the (Fσν0,P,i)sym(P i)’s and the above f’s. As a result, one getsF (5.1)determined by the Euler–Lagrange expressions Eσ and the (free) functions fσνJ,P,i. Moreover, (5.4), and the antisymmetry conditions for theFσνJ,P,i’s, lead to the identities (5.2), as expected.

Now, we are prepared to find explicit regularity conditions forα. By definition, Ds+1αˆ is locally spanned by the followingn-forms:

(5.9)

ηνP =−i∂/∂yν

Pαˆ=

s

X

|J|=0

2FσνJ,P,iωJσ∧ωi, 1≤ |P| ≤s,

ην =−i∂/∂yναˆ=−Eνω0+

s

X

|J|=0

2FσνJ,0,iωσJ∧ωi.

One can see from (5.6) thatthe functionsFσνJ,P,iwhere|J|+|P| ≥s+ 1, depend only upon the f’s (5.7). The (invariant) choice

(5.10) fσνJ,P,i= 0, |J|+|P| ≥s+ 1 then leads to

(5.11) FσνJ,P,i= 0, |J|+|P| ≥s+ 1, and we obtain

(5.12)

ηSν = 2Fσν0,S,iωσ∧ωi,

ηνP = 2Fσν0,P,iωσ∧ωi+ 2Fσνj1,P,iωσj1∧ωi, |P|=s−1, . . .

ηνP =

r0−2

X

|J|=0

2FσνJ,P,iωσJ∧ωi+ 2Fσνj1...jr0−1,P,iωσj1...jr

0−1∧ωi, |P|=s−r0+ 1, . . .

ηνp1=

s

X

|J|=0

2FσνJ,p1,iωσJ∧ωi,

ην=−Eνω0+

s

X

|J|=0

2FσνJ,0,iωJσ∧ωi.

(11)

where (5.13)

FσνJ j,P,i= (−1)|J|+1(Fσν0,J jP,i)sym(J jP i)−(fσνJ,P i,j)sym(J j)+fσνJ j,P,i,

= (−1)|J|+11 2

∂Eσ

∂yJ jP iν −(fσνJ,P i,j)sym(J j)+fσνJ j,P,i, |J|+|P|=s−1.

The results can be summarized as follows.

Theorem 5.1. [30] Let α be Hamiltonian system of order s, let r0 denote the minimal order of the corresponding Lagrangians. Suppose that

(5.14) fσνJ,P,i= 0, |J|+|P| ≥s+ 1, and

(5.15)

rank(Fνσp1...ps,0,i) =mn, rank(Fνσp1...ps−1,j1,i) =mn2,

. . .

rank(Fνσp1...ps−r0 +1,j1...jr0−1,i) =nm n+r0−2 r0−1

! ,

rank

0 FνσP,J,i

−Eν 2Fνσ0,J,i

=maximal, 0≤ |J| ≤s, 1≤ |P| ≤s−r0, where in the above matrices, the (ν, P) label rows, and the (σ, J, i) label columns.

Thenαis regular.

For the most frequent cases of second and first order locally variational forms this result is reduced to the following:

Corollary 5.1. [30] Let s = 1. The following are necessary conditions for α be regular:

(5.16) r0= 1,

(5.17) fσνj,p,i= 0, fσνi,0,j =fνσj,0,i,

(5.18) det∂Eσ

∂yijν −2fσνi,0,j 6= 0,

where in the indicated(mn×mn)-matrix,(ν, j)label the rows and(σ, i)the columns.

Then

(5.19)

ˆ

α=Eσωσ∧ω0+1 4

∂Eσ

∂yiν −∂Eν

∂yiσ

−djfσνi,0,j

ωσ∧ων∧ωi

+∂Eσ

∂yνij −2fσνi,0,j

ωσ∧ωjν∧ωi.

(12)

In terms of a first order Lagrangian λ = Lω0 the regularity conditions (5.17) and (5.18) read

(5.20) det ∂2L

∂yjσ∂yνk −gσνij 6= 0,

where

(5.21) gijσν =−gσνji =−gijνσ, and (in the notations of (3.3)) it holds

(5.22) αˆ =dθλ+p2dµ, p2µ=1

4gσνijωσ∧ων∧ωij. If, in particular,E is projectable ontoJ1Y, i.e.,

(5.23) ∂Eσ

∂yijν = 0

for all (σ, ν, i, j), we can consider for E either a first order Hamiltonian system (s=1), or a zero order Hamiltonian system (s=0) (the latter follows from the fact that variationality conditions imply thatEσ are polynomials of ordernin the first derivatives). Taking into account the above corollary fors= 1, and the definition of regularity fors= 0, we obtain regularity conditions forfirst orderlocally variational forms as follows.

Corollary 5.2. [30] Every locally variational formE onJ1Y is regular.

(1) Let W ⊂J1Y be an open set, and gijσν be functions onW such that (5.24) −gσνij =gσνji =gijνσ, det(gσνij)6= 0.

Then every closed(n+ 1)-form αonW such that (5.25) αˆ=Eσωσ∧ω0+1

2 ∂Eσ

∂yνi +djgjiσν

ωσ∧ων∧ωi+gσνjiωσ∧ωνj ∧ωi

is a regular first order Hamiltonian system related toE.

(2) Suppose that

(5.26) rank∂Eσ

∂yiν

= rank ∂2L

∂yσ∂yνi − ∂2L

∂yσi∂yν

=m,

where in the indicated (m×mn)-matrices, (σ) label the rows and (ν, i) the columns, andλ=Lω0is any (local) first order Lagrangian forE. Then there exists a unique regular zero order Hamiltonian system related toE; it is given by the (n+ 1)-form αonY which in every fibered chart(V, ψ),ψ= (xi, yσ) is expressed as follows:

(5.27)

π1,0 α=Eσωσ∧ω0 +

n

X

k=1

1 (k+ 1)!

kEσ

∂yiν1

1 . . . ∂yνik

k

ωσ∧ων1∧ · · · ∧ωνk∧ωi1...ik.

(13)

6. Legendre transformation

Letα be a regular Hamiltonian system of orders≥1. Then all the canonical 1-contactn-forms

(6.1) ωσ∧ωi, . . . , ωjσ

1...jr0−1∧ωi,

where r0 is the minimal order for the locally variational form E = p1α, belong to the exterior differential system generated by Dαs+1ˆ . However, the generators of Ds+1αˆ naturally associated with the fibered coordinates (i.e., (5.9)), are of the form oflinear combinationsof (6.1), and, moreover, fors >1,nonzero generators are not linearly independent. In this sense the generators associated with fibered coordinates are not canonical. In what follows we shall discuss the existence of coordinates on JsY, canonically adapted to the system Ds+1αˆ , i.e., such that a part of the naturally associated generators coincides with the forms (6.1), and all superfluous generators vanish. More precisely, we shall be interested in the existence of local charts, (W, χ), χ = (xi, yJσ, pJ,iσ , zL), on JsY such that (xi, yJσ) are local fibered coordinates on Jr0−1Y, and the generators of Dαs+1ˆ naturally associated with the coordinates pJ,iσ coincide with the n-forms (6.1), and those associated withzL vanish. Hence,

(6.2)

i∂/∂zLαˆ= 0, ∀L, i∂/∂pJ,i

σ αˆ=ωσJ∧ωi, 0≤ |J| ≤r0−1.

Consequently, thenonzero generators are linarly independent. We shall call such

“canonical” cordinates on JsY Legendre coordinates associated with the regular Hamiltonian systemα.

In the following theorem we adopt the notations of (3.3) and (3.4).

Theorem 6.1. [30] Let α be a regular Hamiltonian system onJsY, let x∈JsY be a point. Suppose that in a neighborhood W of x

(6.3) α=dρ, ρ=θλ+dν+µ,

whereλis a Lagrangian of the minimal order,r0, forE=p1α, defined onπs,r0(W), andµ is such that

(6.4) p2µ=

r0−1

X

|J|,|K|=0

gσνJ,K,i1i2ωσJ∧ωKν ∧ωi1i2,

wheregσνJ,K,i1i2are functions onπs,r0−1(W), satisfying the antisymmetry conditions gσνJ,K,i1i2=−gJ,K,iσν 2i1 =−gνσK,J,i1i2. Put

(6.5) pJ,iσ =

r0−|J|−1

X

l=0

(−1)ldp1dp2. . . dpl

∂L

∂yσJ p1...p

li

+ 4gσνJ,K,ilyνKl, 0≤ |J| ≤r0−1.

Then for any suitable functions zL on W,(W, χ), whereχ= (xi, yσJ, pJ,iσ , zL), is a Legendre chart forα.

(14)

Using (6.5) we can write

(6.6) ρ=−Hω0+

r0−1

X

|J|=0

pJ,iσ dyJσ∧ωi+η+dν+µ3, where

(6.7) η=

r0−1

X

|J|,|K|=0

gJ,K,iσν 1i2dyJσ∧dyνK∧ωi1i2,

µ3 is at least 3-contact, and

(6.8) H =−L+pJ,iσ yJ iσ + 2gJ,K,iσν 1i2yσJ i

1yνKi

2.

We call the functions H (6.8) and pJ,iσ (6.5) a Hamiltonian and momenta of α.

Now,

(6.9) αˆ=−dH∧ω0+

r0−1

X

|J|=0

dpJ,iσ ∧dyσJ∧ωi+dη−p3dη, and computing i∂/∂zLα,ˆ i∂/∂pJ,i

σ α, andˆ i∂/∂yσ

Jαˆ we get that Dαs+1ˆ is spanned by the following nonzeron-forms:

(6.10)

− ∂H

∂pJ,iσ

ω0+dyJσ∧ωiσJ∧ωi, ∂H

∂yσJ −4∂gJ K,iσν 1i2

∂xi2 yνKi1−2

∂gKP,iνρ 1i2

∂yJσ +∂gσνJ K,i1i2

∂yρP +∂gP J,iρσ 1i2

∂yνK

yKiν 1yP iρ 2

ω0

+dpJ,iσ ∧ωi,

where 0≤ |J| ≤r0−1. Notice that if dη= 0 we getDαs+1ˆ spanned by (6.11) ωJσ∧ωi, ∂H

∂yσJ ω0+dpJ,iσ ∧ωi, 0≤ |J| ≤r0−1.

Hamilton equations in Legendre coordinates thus take the following form.

Theorem 6.2. [30]

(1) A section δ:U →W is a Dedecker–Hamilton extremal of α (6.3), (6.6) if, along δ,

(6.12)

∂yJσ

∂xi = ∂H

∂pJ,iσ

,

∂pJ,iσ

∂xi =− ∂H

∂yσJ + 4∂gJ K,iσν 1i2

∂xi2

∂H

∂pK,iν 1

+ 2∂gKP,iνρ 1i2

∂yσJ +∂gσνJ K,i1i2

∂yρP +∂gP J,iρσ 1i2

∂yKν

∂H

∂pK,iν 1

∂H

∂pP,iρ 2

.

(15)

If, in particular, dη= 0, (6.12) take the “classical” form (6.13) ∂yσJ

∂xi = ∂H

∂pJ,iσ

, ∂pJ,iσ

∂xi =−∂H

∂yJσ, 0≤ |J| ≤r0−1.

(2) Ifµ3isπs,r0−1-projectable then (6.12) (resp. (6.13)) are equations for Hamil- ton extremals ofα.

As a consequence of (2) we obtain thatextremals are in bijective correspondence with classes of Hamilton extremals(with the equivalence in the sense of Sec. 4, i.e., that δ1 is equivalent withδ2iffπs,r0◦δ1s,r0◦δ2). In other words,

Corollary 6.1. [30] Hamiltonian systems satisfying the condition (2) of Theorem 6.2 are strongly regular.

The above result shows another geometrical meaning of Legendre transforma- tion: Hamiltonian systems which are regular and admit Legendre transformation according to Theorem 6.1 either are strongly regular, or can be easily brought to a stongly regular form (by modifying the termµ3).

Comments 2. Let us compare our approach to regularity and Legendre transfor- mation with other authors.

(i) Standard Hamilton–DeDonder theory. In the usual formulation of Hamilton- ian field theory Legendre transformation is a map associated with a Lagrangian, definedby the following formulas:

(6.14) piσ= ∂L

∂yiσ if L=L(xi, yν, yνj), and

(6.15) pjσ1...jki=

r−k−1

X

l=0

(−1)ldp1dp2. . . dpl ∂L

∂yσj

1...jkp1...pli

, 0≤k≤r−1, for a Lagrangian of orderr≥2 ([8], [7], [16], [33], [35], [37], etc.). These formulas have their origin in the (noninvariant) decomposition of the Poincar´e–Cartan form θλ (3.4) in the canonical basis (dxi, dyJσ), 0≤ |J| ≤r−1, i.e.,

(6.16) θλ= (L−

r−1

X

|J|=0

pJ iσyJ iσ0+

r−1

X

|J|=0

pJ iσ dyσJ∧ωi.

However, for global Lagrangians of order r ≥ 2 the form (3.4) is neither unique nor globally defined. It is replaced by Θ =θλ+p1dν (cf. notations of (3.3)), and, consequently, (6.15) are replaced by more general formulas

(6.17) pjσ1...jki=

r−k−1

X

l=0

(−1)ldp1dp2. . . dpl

∂L

∂yjσ

1...jkp1...pli

+cjσ1...jkp1...pli , 0 ≤ k ≤ r−1, where cjσ1...jkp1...pli are auxiliary (free) functions (Krupka [23], Gotay [16]). In the Hamilton–De Donder theory, a Lagrangian is called regularif

(16)

the Legendre map defined by (6.15), resp. (6.17) is regular. In the case of a first order Lagrangian this means thatλsatisfies the condition

(6.18) det ∂2L

∂qjσ∂qkν 6= 0

at each point of J1Y. Since for higher order Lagrangians the Legendre transfor- mation (6.17) depends uponp1dν, one could expect that the corresponding regu- larity condition conserves this property. Surprizingly enough, it has been proved in [23] and [16] that the regularity conditions do not depend upon the functions cjσ1...jkp1...pli, and are of the form [37], [23], [16]

(6.19) rank 1

[j1. . . j2r−s(pr+1. . . ps] [p1. . . pr)]

2L

∂yjσ

1...j2r−s(pr+1...ps∂ypν

1...pr)

!

= max where [j1. . . j2r−spr+1. . . ps] and [p1. . . pr] denotes the number of all different se- quences arising by permuting the sequence j1. . . j2r−spr+1. . . ps and p1. . . pr, re- spectively; as usual,r ≤k≤2r−1, and, in the indicated matrices, σ, j1≤ · · · ≤ j2r−k label columns andν, p1≤ · · · ≤pk label rows, and the bracket denotes sym- metrization in the corresponding indices. (Notice that within the present approach, by (6.3), the nondependence of regularity conditions upon the c’s is trivial). As a result one obtains that if a Lagrangian satisfies (6.18) (respectively, (6.19)) then every solution δof the Hamilton–De Donder equationsδiξλ = 0 (respectively, δiξdΘ = 0), is of the formδ=J1γ (respectively,π2r−1,r◦δ=Jrγ) whereγis an extremal ofλ. However, while in the Legendre coordinates defined by (6.15) the lo- cal Hamilton–De Donder equations, i.e.,δiξλ= 0, take the familiar “canonical”

form,

∂yJσ

∂xi = ∂H

∂pJ,iσ

, ∂pJ,iσ

∂xi =−∂H

∂yJσ,

for the global Hamilton–De Donder equations, i.e., δiξdΘ = 0, using the “Le- gendre transformation” defined by (6.17), one does not generally obtain a similar

“canonical” representation.

(ii) A generalization of the concepts of regularity and Legendre transformation within the Hamilton–De Donder theory. In the paper [26] second order Lagrangians, affine in the second derivatives, and admitting first order Poincar´e–Cartan forms were studied. Notice that in the sense of the regularity condition (6.19), La- grangians of this kind are apparently singular. In [26], the definition of a regular Lagrangian is extended in the following way: a Lagrangian is calledregular if the solutions of the Euler–Lagrange and Hamilton–De Donder equations are equiva- lent (in the sense that the sets of solutions are in bijective correspondence). The following results were proved:

Theorem 6.3. [26] Consider a Lagrangian of the formλ=Lω0 where, in fibered coordinates,L admits an (obviously invariant) decomposition

(6.20) L=L0(xi, yσ, yjσ) +hpqν (xi, yσ)yνpq.

(17)

Then θλ is projectable onto J1Y, and, consequently, Hamilton–De Donder equa- tions are equations for sectionsδ:U →J1Y. If the condition

(6.21) det ∂2L0

∂yσi∂yρk −∂hikσ

∂yρ −∂hikρ

∂yσ 6= 0

is satisfied thenλis regular, i.e., the Euler–Lagrange and the Hamilton–De Donder equations ofλare equivalent, and the mapping

(6.22) (xi, yσ, yjσ)→(xi, yσ, pjσ), pjσ= ∂L0

∂yjσ −∂hjkσ

∂xk −∂hjkσ

∂yν +∂hjkν

∂yσ yνk

is a local coordinate transformation on J1Y.

The formula (6.22) comes from the following noninvariant decomposition of the Poincar´e–Cartan form

(6.23) θλ=−Hω0+pjσdyσ∧ωj+d(hijσyjσωi), where

(6.24) H =−L0+∂L0

∂yjσ −∂hjkσ

∂yν yjσykν.

The functionsH and pjσ were called in [26] theHamiltonian andmomenta of the Lagrangian (6.20), and (6.22) was calledLegendre transformation. In the Legendre coordinates (6.22) the Hamilton–De Donder equations read

(6.25) ∂H

∂yν =−∂piν

∂xi, ∂H

∂pkν = ∂yν

∂xk.

As pointed out by Krupka and ˇStˇep´ankov´a, the above results directly apply to the Einstein–Hilbert Lagrangian (scalar curvature) of the General Relativity Theory (for explicit computations see [26]). The above ideas were applied to study also some other kinds of higher order Lagrangians with projectable Poincar´e–Cartan forms by [10] (cf. also comments in [11]).

(iii) Dedecker’s approach to first order Hamiltonian field theory. In [5], Dedecker proposed a Hamilton theory for first order Lagrangians on contact elements. If transfered to fibered manifolds, it becomes a “nonstandard” Hamiltonian theory.

The core is to consider Hamilton equations of the form

(6.26) δiξdρ= 0,

whereρis a Lepagean equivalent of a first order Lagrangian of the form ρ=θλ+

n

X

k=2

gσi1...ik

1...σkωσ1∧ · · · ∧ωσk∧ωi1...ik. Dedecker showed that if the condition

(6.27) det ∂2L

∂yiσ∂yjν −gσνij 6= 0

参照

関連したドキュメント

In the second computation, we use a fine equidistant grid within the isotropic borehole region and an optimal grid coarsening in the x direction in the outer, anisotropic,

He thereby extended his method to the investigation of boundary value problems of couple-stress elasticity, thermoelasticity and other generalized models of an elastic

Keywords: continuous time random walk, Brownian motion, collision time, skew Young tableaux, tandem queue.. AMS 2000 Subject Classification: Primary:

Kilbas; Conditions of the existence of a classical solution of a Cauchy type problem for the diffusion equation with the Riemann-Liouville partial derivative, Differential Equations,

Here we continue this line of research and study a quasistatic frictionless contact problem for an electro-viscoelastic material, in the framework of the MTCM, when the foundation

Abstract The representation theory (idempotents, quivers, Cartan invariants, and Loewy series) of the higher-order unital peak algebras is investigated.. On the way, we obtain

This paper presents an investigation into the mechanics of this specific problem and develops an analytical approach that accounts for the effects of geometrical and material data on

7.1. Deconvolution in sequence spaces. Subsequently, we present some numerical results on the reconstruction of a function from convolution data. The example is taken from [38],