Abstract

A Lie group acting on finite-dimensional space is generated by its infinitesimal transformations and conversely, any Lie algebra of vector fields in finite dimension generates a Lie group (the first fundamental theorem). This classical result is adjusted for the infinite-dimensional case. We prove that the (local, 𝐶 smooth) action of a Lie group on infinite-dimensional space (a manifold modelled on ) may be regarded as a limit of finite-dimensional approximations and the corresponding Lie algebra of vector fields may be characterized by certain finiteness requirements. The result is applied to the theory of generalized (or higher-order) infinitesimal symmetries of differential equations.

1. Preface

In the symmetry theory of differential equations, the generalized (or: higher-order, Lie-Bäcklund) infinitesimal symmetries 𝑧𝑍=𝑖𝜕𝜕𝑥𝑖+𝑧𝑗𝐼𝜕𝜕𝑤𝑗𝐼𝑖=1,,𝑛;𝑗=1,,𝑚;𝐼=𝑖1𝑖𝑛;𝑖1,,𝑖𝑛,=1,,𝑛(1.1) where the coefficients𝑧𝑖=𝑧𝑖,𝑥𝑖,𝑤𝑗𝐼,,𝑧𝑗𝐼=𝑧𝑗𝐼,𝑥𝑖,𝑤𝑗𝐼,(1.2) are functions of independent variables 𝑥𝑖, dependent variables 𝑤𝑗 and a finite number of jet variables 𝑤𝑗𝐼=𝜕𝑛𝑤𝑗/𝜕𝑥𝑖1𝜕𝑥𝑖𝑛 belong to well-established concepts. However, in spite of this matter of fact, they cause an unpleasant feeling. Indeed, such vector fields as a rule do not generate any one-parameter group of transformations𝑥𝑖=𝐺𝑖𝜆;,𝑥𝑖,𝑤𝑗𝐼,,𝑤𝑗𝐼=𝐺𝑗𝐼𝜆;,𝑥𝑖,𝑤𝑗𝐼,(1.3) in the underlying infinite-order jet space since the relevant Lie system𝜕𝐺𝑖𝜕𝜆=𝑧𝑖,𝐺𝑖,𝐺𝑗𝐼,,𝜕𝐺𝑗𝐼𝜕𝜆=𝑧𝑗𝐼,𝐺𝑖,𝐺𝑗𝐼𝐺,𝑖||𝜆=0=𝑥𝑖,𝐺𝑗𝐼||𝜆=0=𝑤𝑗𝐼(1.4) need not have any reasonable (locally unique) solution. Then 𝑍 is a mere formal concept [17] not related to any true transformations and the term “infinitesimal symmetry 𝑍" is misleading, no 𝑍-symmetries of differential equations in reality appear.

In order to clarify the situation, we consider one-parameter groups of local transformations in . We will see that they admit “finite-dimensional approximations" and as a byproduct, the relevant infinitesimal transformations may be exactly characterized by certain “finiteness requirements" of purely algebraical nature. With a little effort, the multidimensional groups can be easily involved, too. This result was briefly discussed in [8, page 243] and systematically mentioned at several places in monograph [9], but our aim is to make some details more explicit in order to prepare the necessary tools for systematic investigation of groups of generalized symmetries. We intend to continue our previous articles [1013] where the algorithm for determination of all individual generalized symmetries was already proposed.

For the convenience of reader, let us transparently describe the crucial approximation result. We consider transformations (2.1) of a local one-parameter group in the space with coordinates 1,2,. Equations (2.1) of transformations 𝐦(𝜆) can be schematically represented by Figure 1(a).

We prove that in appropriate new coordinate system 𝐹1,𝐹2,on , the same transformations 𝐦(𝜆) become block triangular as in Figure 1(b). It follows that a certain hierarchy of finite-dimensional subspaces of is preserved which provides the “approximation" of 𝐦(𝜆). The infinitesimal transformation 𝑍=d𝐦(𝜆)/d𝜆|𝜆=0 clearly preserves the same hierarchy which provides certain algebraical “finiteness" of 𝑍.

If the primary space is moreover equipped with an appropriate structure, for example, the contact forms, it turns into the jet space and the results concerning the transformation groups on become the theory of higher-order symmetries of differential equations. Unlike the common point symmetries which occupy a number of voluminous monographs (see, e.g., [14, 15] and extensive references therein) this higher-order theory was not systematically investigated yet. We can mention only the isolated article [16] which involves a direct proof of the “finiteness requirements" for one-parameter groups (namely, the result (𝜄) of Lemma 5.4 below) with two particular examples and monograph [7] involving a theory of generalized infinitesimal symmetries in the formal sense.

Let us finally mention the intentions of this paper. In the classical theory of point or Lie's contact-symmetries of differential equations, the order of derivatives is preserved (Figure 2(a)). Then the common Lie's and Cartan's methods acting in finite dimensional spaces given ahead of calculations can be applied. On the other extremity, the generalized symmetries need not preserve the order (Figure 2(c)) and even any finite-dimensional space and then the common classical methods fail. For the favourable intermediate case of groups of generalized symmetries, the invariant finite-dimensional subspaces exist, however, they are not known in advance (Figure 2(b)). We believe that the classical methods can be appropriately adapted for the latter case, and this paper should be regarded as a modest preparation for this task.

2. Fundamental Approximation Results

Our reasonings will be carried out in the space with coordinates 1,2,[9] and we introduce the structural family of all real-valued, locally defined and 𝐶-smooth functions 𝑓=𝑓(1,,𝑚(𝑓)) depending on a finite number of coordinates. In future, such functions will contain certain 𝐶-smooth real parameters, too.

We are interested in (local) groups of transformations 𝐦(𝜆) in defined by formulae𝐦(𝜆)𝑖=𝐻𝑖𝜆;1,,𝑚(𝑖),𝜀𝑖<𝜆<𝜀𝑖,𝜀𝑖>0(𝑖=1,2,),(2.1) where 𝐻𝑖 if the parameter 𝜆 is kept fixed. We suppose𝐦(0)=id.,𝐦(𝜆+𝜇)=𝐦(𝜆)𝐦(𝜇)(2.2) whenever it makes a sense. An open and common definition domain for all functions 𝐻𝑖 is tacitly supposed. (In more generality, a common definition domain for every finite number of functions 𝐻𝑖 is quite enough and the germ and sheaf terminology would be more adequate for our reasonings, alas, it looks rather clumsy.)

Definition 2.1. For every 𝐼=1,2, and 0<𝜀<min{𝜀1,,𝜀𝐼}, let (𝐼,𝜀) be the subset of all composed functions 𝜆𝐹=𝐹,𝐦𝑗𝑖,=𝐹,𝐻𝑖𝜆𝑗;1,,𝑚(𝑖),,(2.3) where 𝑖=1,,𝐼;  𝜀<𝜆𝑗<𝜀;  𝑗=1,,𝐽=𝐽(𝐼)=max{𝑚(1),,𝑚(𝐼)} and 𝐹 is arbitrary 𝐶-smooth function (of 𝐼𝐽 variables). In functions 𝐹(𝐼,𝜀), variables 𝜆1,,𝜆𝐽 are regarded as mere parameters.

Functions (2.3) will be considered on open subsets of where the rank of the Jacobi (𝐼𝐽×𝐽)-matrix𝜕𝜕𝑗𝐻𝑖𝜆𝑗;1,,𝑚(𝑖)𝑖=1,,𝐼;𝑗,𝑗=1,,𝐽(2.4) of functions 𝐻𝑖(𝜆𝑗;1,,𝑚(𝑖)) locally attains the maximum (for appropriate choice of parameters). This rank and therefore the subset (𝐼,𝜀) does not depend on 𝜀 as soon as 𝜀=𝜀(𝐼) is close enough to zero. This is supposed from now on and we may abbreviate (𝐼)=(𝐼,𝜀).

We deal with highly nonlinear topics. Then the definition domains cannot be kept fixed in advance. Our results will be true locally, near generic points, on certain open everywhere dense subsets of the underlying space . With a little effort, the subsets can be exactly characterized, for example, by locally constant rank of matrices, functional independence, existence of implicit function, and so like. We follow the common practice and as a rule omit such routine details from now on.

Lemma 2.2 (approximation lemma). The following inclusion is true: 𝐦(𝜆)(𝐼)(𝐼).(2.5)

Proof. Clearly 𝐦(𝜆)𝐻𝑖𝜆𝑗;=𝐦(𝜆)𝐦𝜆𝑗𝑖=𝐦𝜆+𝜆𝑗𝑖=𝐻𝑖𝜆+𝜆𝑗;(2.6) and therefore 𝐦(𝜆)𝐹=𝐹,𝐻𝑖𝜆+𝜆𝑗;1,,𝑚(𝑖),(𝐼).(2.7)

Denoting by 𝐾(𝐼) the rank of matrix (2.4), there exist basical functions 𝐹𝑘=𝐹𝑘,𝐻𝑖𝜆𝑗;1,,𝑚(𝑖),(𝐼)(𝑘=1,,𝐾(𝐼))(2.8) such that rank(𝜕𝐹𝑘/𝜕𝑗)=𝐾(𝐼). Then a function 𝑓 lies in (𝐼) if and only if 𝑓=𝑓(𝐹1,,𝐹𝐾(𝐼)) is a composed function. In more detail𝐹=𝐹𝜆1,,𝜆𝐽;𝐹1,,𝐹𝐾(𝐼)(𝐼)(2.9) is such a composed function if we choose 𝑓=𝐹 given by (2.3). Parameters 𝜆1,,𝜆𝐽 occurring in (2.3) are taken into account here. It follows that𝜕𝐹𝜕𝜆𝑗=𝜕𝐹𝜕𝜆𝑗𝜆1,,𝜆𝐽;𝐹1,,𝐹𝐾(𝐼)(𝐼)(𝑗=1,,𝐽)(2.10) and analogously for the higher derivatives.

In particular, we also have 𝐻𝑖𝜆;1,,𝑚(𝑖)=𝐻𝑖𝜆;𝐹1,,𝐹𝐾(𝐼)(𝐼)(𝑖=1,,𝐼)(2.11) for the choice 𝐹=𝐻𝑖(𝜆;) in (2.9) whence𝜕𝑟𝐻𝑖𝜕𝜆𝑟=𝜕𝑟𝐻𝑖𝜕𝜆𝑟𝜆;𝐹1,,𝐹𝐾(𝐼)(𝐼)(𝑖=1,,𝐼;𝑟=0,1,).(2.12) The basical functions can be taken from the family of functions 𝐻𝑖(𝜆;)  (𝑖=1,,𝐼) for appropriate choice of various values of 𝜆. Functions (2.12) are enough as well even for a fixed value 𝜆, for example, for 𝜆=0, see Theorem 3.2 below.

Lemma 2.3. For any basical function, one has 𝐦(𝜆)𝐹𝑘=𝐹𝑘𝜆;𝐹1,,𝐹𝐾(𝐼)(𝑘=1,,𝐾(𝐼)).(2.13)

Proof. 𝐹𝑘(𝐼) implies 𝐦(𝜆)𝐹𝑘(𝐼) and (2.9) may be applied with the choice 𝐹=𝐦(𝜆)𝐹𝑘 and 𝜆1==𝜆𝐽=𝜆.

Summary 1. Coordinates 𝑖=𝐻𝑖(0;)  (𝑖=1,,𝐼) were included into the subfamily (𝐼) which is transformed into itself by virtue of (2.13). So we have a one-parameter group acting on (𝐼). One can even choose 𝐹1=1,,𝐹𝐼=𝐼 here and then, if 𝐼 is large enough, formulae (2.13) provide a “finite-dimensional approximation" of the primary mapping 𝐦(𝜆). The block-triangular structure of the infinite matrix of transformations 𝐦(𝜆) mentioned in Section 1 appears if 𝐼 and the system of functions 𝐹1,𝐹2, is succesively completed.

3. The Infinitesimal Approach

We introduce the vector field 𝑧𝑍=𝑖𝜕𝜕𝑖=d𝐦(𝜆)|||d𝜆𝜆=0𝑧𝑖=𝜕𝐻𝑖𝜕𝜆0;1,,𝑚(𝑖);𝑖=1,2,,(3.1) the infinitesimal transformation (𝑇) of group 𝐦(𝜆). Let us recall the celebrated Lie system 𝜕𝐦𝜕𝜆(𝜆)𝑖=𝜕𝐻𝑖𝜕𝜆(𝜆;)=𝜕𝐻𝑖||||𝜕𝜇(𝜆+𝜇;)𝜇=0=𝜕𝜕𝜇𝐦(𝜆+𝜇)𝑖||||𝜇=0=𝐦(𝜆)𝜕𝜕𝜇𝐦(𝜇)𝑖||||𝜇=0=𝐦(𝜆)𝑍𝑖=𝐦(𝜆)𝑧𝑖.(3.2) In more explicit (and classical) transcription𝜕𝐻𝑖𝜕𝜆𝜆;1,,𝑚(𝑖)=𝑧𝑖𝐻1𝜆;1,,𝑚(1),,𝐻𝑚(𝑖)𝜆;1,,𝑚(𝑚(𝑖)).(3.3) One can also check the general identity𝜕𝑟𝜕𝜆𝑟𝐦(𝜆)𝑓=𝐦(𝜆)𝑍𝑟𝑓(𝑓;𝑟=0,1,)(3.4) by a mere routine induction on 𝑟.

Lemma 3.1 (finiteness lemma). For all 𝑟, 𝑍𝑟(𝐼)(𝐼).

Proof. Clearly 𝑍𝐹=𝐦(𝜆)||𝑍𝐹𝜆=0=𝜕𝜕𝜆𝐦(𝜆)𝐹|||𝜆=0(𝐼)(3.5) for any function (2.3) by virtue of (2.10): induction on 𝑟.

Theorem 3.2 (finiteness theorem). Every function 𝐹(𝐼) admits (locally, near generic points) the representation 𝐹𝜕𝐹=,𝑟𝐻𝑖𝜕𝜆𝑟0;1,,𝑚(𝑖),(3.6) in terms of a composed function where 𝑖=1,,𝐼 and 𝐹 is a -smooth function of a finite number of variables.

Proof. Let us temporarily denote 𝐻𝑖𝑟=𝜕𝑟𝐻𝑖𝜕𝜆𝑟𝜕(𝜆;)=𝑟𝜕𝜆𝑟𝐦(𝜆)𝑖,𝑖𝑟=𝐻𝑖𝑟(0;)=𝑍𝑟𝑖,(3.7) where the second equality follows from (3.4) with 𝑓=𝑖,  𝜆=0. Then 𝐻𝑖𝑟=𝐦(𝜆)𝑖𝑟=𝐦(𝜆)𝑍𝑟𝑖(3.8) by virtue of (3.4) with general 𝜆.
If 𝑗=𝑗(𝑖) is large enough, there does exist an identity 𝑖𝑗+1=𝐺𝑖(𝑖0,,𝑖𝑗). Therefore 𝜕𝑗+1𝐻𝑖𝜕𝜆𝑗+1=𝐻𝑖𝑗+1=𝐺𝑖𝐻𝑖0,,𝐻𝑖𝑗=𝐺𝑖𝐻𝑖𝜕,,𝑗𝐻𝑖𝜕𝜆𝑗(3.9) by applying 𝐦(𝜆). This may be regarded as ordinary differential equation with initial values 𝐻𝑖||𝜆=0=𝑖0𝜕,,𝑗𝐻𝑖𝜕𝜆𝑗||||𝜆=0=𝑖𝑗.(3.10) The solution 𝐻𝑖=𝐻𝑖(𝜆;𝑖0,,𝑖𝑗) expressed in terms of initial values reads 𝐻𝑖𝜆;1,,𝑚(𝑖)=𝐻𝑖𝜆;𝐻𝑖0;1,,𝑚(𝑖)𝜕,,𝑗𝐻𝑖𝜕𝜆𝑗0;1,,𝑚(𝑖)(3.11) in full detail. If 𝜆 is kept fixed, this is exactly the identity (3.6) for the particular case 𝐹=𝐻𝑖(𝜆;1,,𝑚(𝑖)). The general case follows by a routine.

Definition 3.3. Let 𝔾 be the set of (local) vector fields 𝑧𝑍=𝑖𝜕𝜕𝑖𝑧𝑖,innitesum(3.12) such that every family of functions {𝑍𝑟𝑖}𝑟   (𝑖 fixed but arbitrary) can be expressed in terms of a finite number of coordinates.

Remark 3.4. Neither 𝔾+𝔾𝔾 nor [𝔾,𝔾]𝔾 as follows from simple examples. However, 𝔾 is a conical set (over ): if 𝑍𝔾 then 𝑓𝑍𝔾 for any 𝑓. Easy direct proof may be omitted here.

Summary 2. If 𝑍 is  𝒯 of a group then all functions 𝑍𝑟𝑖  (𝑖=1,,𝐼; 𝑟=0,1,) are included into family (𝐼) hence 𝑍𝔾. The converse is clearly also true: every vector field 𝑍𝔾 generates a local Lie group since the Lie system (3.3) admits finite-dimensional approximations in spaces (𝐼).

Let us finally reformulate the last sentence in terms of basical functions.

Theorem 3.5 (approximation theorem). Let 𝑍𝔾 be a vector field locally defined on and 𝐹1,,𝐹𝐾(𝐼) be a maximal functionally independent subset of the family of all functions 𝑍𝑟𝑖(𝑖=1,,𝐼;𝑟=0,1,).(3.13) Denoting 𝑍𝐹𝑘=𝐹𝑘(𝐹1,,𝐹𝐾(𝐼)), then the system 𝜕𝜕𝜆𝐦(𝜆)𝐹𝑘=𝐦(𝜆)𝑍𝐹𝑘=𝐹𝑘𝐦(𝜆)𝐹1,,𝐦(𝜆)𝐹𝐾(𝐼)(𝑘=1,,𝐾(𝐼))(3.14) may be regarded as a “finite-dimensional approximation" to the Lie system (3.3) of the one-parameter local group 𝐦(𝜆) generated by 𝑍.

In particular, assuming 𝐹1=1,,𝐹𝐼=𝐼, then the the initial portion dd𝜆𝐦(𝜆)𝐹𝑖=dd𝜆𝐦(𝜆)𝑖=d𝐻d𝜆𝑖=𝑧𝑖𝐻1,,𝐻𝑚(𝑖)(𝑖=1,,𝐼)(3.15) of the above system transparently demonstrates the approximation property.

4. On the Multiparameter Case

The following result does not bring much novelty and we omit the proof.

Theorem 4.1. Let 𝑍1,,𝑍𝑑 be commuting local vector fields in the space . Then 𝑍1,,𝑍𝑑𝔾 if and only if the vector fields 𝑍=𝑎1𝑍1++𝑎𝑑𝑍𝑑  (𝑎1,,𝑎𝑑) locally generate an abelian Lie group.

In full non-Abelian generality, let us consider a (local) multiparameter group formally given by the same equations (2.1) as above where 𝜆=(𝜆1,,𝜆𝑑)𝑑 are parameters close to the zero point 0=(0,,0)𝑑. The rule (2.2) is generalized as 𝐦(0)=id.,𝐦(𝜑(𝜆,𝜇))=𝐦(𝜆)𝐦(𝜇),(4.1) where 𝜆=(𝜆1,,𝜆𝑑),  𝜇=(𝜇1,,𝜇𝑑) and 𝜑=(𝜑1,,𝜑𝑑) determine the composition of parameters. Appropriately adapting the space (𝐼) and the concept of basical functions 𝐹1,,𝐹𝐾(𝐼), Lemma 2.2 holds true without any change.

Passing to the infinitesimal approach, we introduce vector fields 𝑍1,,𝑍𝑑 which are 𝒯 of the group. We recall (without proof) the Lie equations [17] 𝜕𝜕𝜆𝑗𝐦(𝜆)𝑎𝑓=𝑗𝑖(𝜆)𝐦(𝜆)𝑍𝑗𝑓(𝑓;𝑗=1,,𝑑)(4.2) with the initial condition 𝐦(0)=id. Assuming 𝑍1,,𝑍𝑑 linearly independent over , coefficients 𝑎𝑗𝑖(𝜆) may be arbitrarily chosen and the solution 𝐦(𝜆) always is a group transformation (the first fundamental theorem). If basical functions 𝐹1,,𝐹𝐾(𝐼) are inserted for 𝑓, we have a finite-dimensional approximation which is self-contained in the sense that 𝑍𝑗𝐹𝑘=𝐹𝑘𝑗𝐹1,,𝐹𝐾(𝐼)(𝑗=1,,𝑑;𝑘=1,,𝐾(𝐼))(4.3) are composed functions in accordance with the definition of the basical functions.

Let us conversely consider a Lie algebra of local vector fields 𝑍=𝑎1𝑍1++𝑎𝑑𝑍𝑑  (𝑎𝑖) on the space . Let moreover 𝑍1,,𝑍𝑑𝔾  uniformly in the sense that there is a universal space (𝐼) with 𝑍𝑖(𝐼)(𝐼) for all 𝑖=1,,𝑑. Then the Lie equations may be applied and we obtain reasonable finite-dimensional approximations.

Summary 3. Theorem 4.1 holds true even in the non-Abelian and multidimensional case if the inclusions 𝑍1,,𝑍𝑑𝔾 are uniformly satisfied.

As yet we have closely simulated the primary one-parameter approach, however, the results are a little misleading: the uniformity requirement in Summary 3 may be completely omitted. This follows from the following result [9, page 30] needless here and therefore stated without proof.

Theorem 4.2. Let 𝒦 be a finite-dimensional submodule of the module of vector fields on such that [𝒦,𝒦]𝒦. Then 𝒦𝔾 if and only if there exist generators (over ) of submodule 𝒦 that are lying in 𝔾.

5. Symmetries of the Infinite-Order Jet Space

The previous results can be applied to the groups of generalized symmetries of partial differential equations. Alas, some additional technical tools cannot be easily explained at this place, see the concluding Section 11 below. So we restrict ourselves to the trivial differential equations, that is, to the groups of generalized symmetries in the total infinite-order jet space which do not require any additional preparations.

Let 𝐌(𝑚,𝑛) be the jet space of 𝑛-dimensional submanifolds in 𝑚+𝑛 [913]. We recall the familiar (local) jet coordinates 𝑥𝑖,𝑤𝑗𝐼𝐼=𝑖1𝑖𝑟;𝑖,𝑖1,,𝑖𝑟.=1,,𝑛;𝑟=0,1,;𝑗=1,,𝑚(5.1) Functions 𝑓=𝑓(,𝑥𝑖,𝑤𝑗𝐼,) on 𝐌(𝑚,𝑛) are 𝐶-smooth and depend on a finite number of coordinates. The jet coordinates serve as a mere technical tool. The true jet structure is given just by the module Ω(𝑚,𝑛) of contact forms 𝑎𝜔=𝑗𝐼𝜔𝑗𝐼nitesum,𝜔𝑗𝐼=d𝑤𝑗𝐼𝑤𝑗𝐼𝑖d𝑥𝑖(5.2) or, equivalently, by the “orthogonal" module (𝑚,𝑛)=Ω(𝑚,𝑛) of formal derivatives 𝑎𝐷=𝑖𝐷𝑖𝐷𝑖=𝜕𝜕𝑥𝑖+𝑤𝑗𝐼𝑖𝜕𝜕𝑤𝑗𝐼;𝑖=1,,𝑛;𝐷𝜔𝑗𝐼=𝜔𝑗𝐼(𝐷)=0.(5.3) Let us state useful formulae 𝐷d𝑓=𝑖𝑓d𝑥𝑖+𝜕𝑓𝜕𝑤𝑗𝐼𝜔𝑗𝐼,𝐷𝑖d𝜔𝑗𝐼=𝜔𝑗𝐼𝑖,𝐷𝑖𝜔𝑗𝐼=𝜔𝑗𝐼𝑖,(5.4) where 𝐷𝑖=𝐷𝑖d+d𝐷𝑖 denotes the Lie derivative.

We are interested in (local) one-parameter groups of transformations 𝐦(𝜆) given by certain formulae𝐦(𝜆)𝑥𝑖=𝐺𝑖𝜆;,𝑥𝑖,𝑤𝑗𝐼,,𝐦(𝜆)𝑤𝑗𝐼=𝐺𝑗𝐼𝜆;,𝑥𝑖,𝑤𝑗𝐼,(5.5) and in vector fields𝑧𝑍=𝑖,𝑥𝑖,𝑤𝑗𝐼𝜕,𝜕𝑥𝑖+𝑧𝑗𝐼,𝑥𝑖,𝑤𝑗𝐼𝜕,𝜕𝑤𝑗𝐼(5.6) locally defined on the jet space 𝐌(𝑚,𝑛); see also (1.1) and (1.2).

Definition 5.1. We speak of a group of morphisms (5.5)of the jet structure if the inclusion 𝐦(𝜆)Ω(𝑚,𝑛)Ω(𝑚,𝑛) holds true. We speak of a (universal) variation (5.6) of the jet structure if𝑍Ω(𝑚,𝑛)Ω(𝑚,𝑛). If a variation (5.6) moreover generates a group, speaks of a (generalized or higher-order) infinitesimal symmetry of the jet structure.

So we intentionally distinguish between true infinitesimal transformations generating a group and the formal concepts; this point of view and the terminology are not commonly used in the current literature.

Remark 5.2. A few notes concerning this unorthodox terminology are useful here. In actual literature, the vector fields (5.6) are as a rule decomposed into the “trivial summand 𝐷" and the so-called “evolutionary form 𝑉" of the vector field 𝑍, explicitly 𝑧𝑍=𝐷+V𝐷=𝑖𝐷𝑖𝑄(𝑚,𝑛),𝑉=𝑗𝐼𝜕𝜕𝑤𝑗𝐼,𝑄𝑗𝐼=𝑧𝑗𝐼𝑤𝑗𝐼𝑖𝑧𝑖.(5.7) The summand 𝐷 is usually neglected in a certain sense [37] and the “essential" summand 𝑉 is identified with the evolutional system 𝜕𝑤𝑗𝐼𝜕𝜆=𝑄𝑗𝐼,𝑥𝑖,𝑤𝑗𝐼𝑤,𝑗𝐼=𝜕𝑛𝑤𝑗𝜕𝑥𝑖1𝜕𝑥𝑖𝑛𝜆,𝑥1,,𝑥𝑛(5.8) of partial differential equations (the finite subsystem with 𝐼=𝜙 empty is enough here since the remaining part is a mere prolongation). This evolutional system is regarded as a “virtual flow" on the “space of solutions" 𝑤𝑗=𝑤𝑗(𝑥1,,𝑥𝑛), see [7, especially page 11]. In more generality, some differential constraints may be adjoint. However, in accordance with the ancient classical tradition, functions 𝛿𝑤𝑗=𝜕𝑤𝑗/𝜕𝜆 are just the variations. (There is only one novelty: in classical theory, 𝛿𝑤𝑗 are introduced only along a given solution while the vector fields 𝑍 are “universally" defined on the space.) In this “evolutionary approach", the properties of the primary vector field 𝑍 are utterly destroyed. It seems that the true sense of this approach lies in the applications to the topical soliton theory. However, then the evolutional system is always completed with boundary conditions and embedded into some normed functional spaces in order to ensure the existence of global “true flows". This is already quite a different story and we return to our topic.

In more explicit terms, morphisms (5.5) are characterized by the (implicit) recurrence𝐺𝑗𝐼𝑖𝐷𝑖𝐺𝑖=𝐷𝑖𝐺𝑗𝐼𝑖=1,,𝑛,(5.9) where det(𝐷𝑖𝐺𝑖)0 is supposed and vector field (5.6) is a variation if and only if𝑧𝑗𝐼𝑖=𝐷𝑖𝑧𝑗𝐼𝑤𝑗𝐼𝑖𝐷𝑖𝑧𝑖.(5.10) Recurrence (5.9) easily follows from the inclusion 𝐦(𝜆)𝜔𝑗𝐼Ω(𝑚,𝑛) and we omit the proof. Recurrence (5.10) follows from the identity 𝑍𝜔𝑗𝐼=𝑍d𝑤𝑗𝐼𝑤𝑗𝐼𝑖d𝑥𝑖=d𝑧𝑗𝐼𝑧𝑗𝐼𝑖d𝑥𝑖𝑤𝑗𝐼𝑖d𝑧𝑖𝐷𝑖z𝑗𝐼𝑧𝑗𝐼𝑖𝑤𝑗𝐼𝑖𝐷𝑖𝑧𝑖d𝑥𝑖(modΩ(𝑚,𝑛))(5.11) and the inclusion 𝑍𝜔𝑗𝐼Ω(𝑚,𝑛). The obvious formula𝑍𝜔𝑗𝐼=𝜕𝑧𝑗𝐼𝜕𝑤𝑗𝐼𝑤𝑗𝐼𝑖𝜕𝑧𝑖𝜕𝑤𝑗𝐼𝜔𝑗𝐼(5.12) appearing on this occasion also is of a certain sense, see Theorem 5.5 and Section 10 below. It follows that the initial functions 𝐺𝑖, 𝐺𝑗, 𝑧𝑖, 𝑧𝑗 (empty 𝐼=𝜙) may be in principle arbitrarily prescribed in advance. This is the familiar prolongation procedure in the jet theory.

Remark 5.3. Recurrence (5.10) for the variation 𝑍 can be succintly expressed by 𝜔𝑗𝐼𝑖(𝑍)=𝐷𝑖𝜔𝑗𝐼(𝑍). This remarkable formula admits far going generalizations, see concluding Examples 11.3 and 11.4 below.

Let us recall that a vector field (5.6) generates a group (5.5) if and only if 𝑍𝔾 hence if and only if every family𝑍𝑟𝑥𝑖𝑟,𝑍𝑟𝑤𝑗𝐼𝑟(5.13) can be expressed in terms of a finite number of jet coordinates. We conclude with simple but practicable remark: due to jet structure, the infinite number of conditions (5.13) can be replaced by a finite number of requirements if 𝑍 is a variation.

Lemma 5.4. Let (5.6) be a variation of the jet structure. Then the inclusion 𝑍𝔾 is equivalent to any of the requirements (𝜄) every family of functions 𝑍𝑟𝑥𝑖𝑟,𝑍𝑟𝑤𝑗𝑟(𝑖=1,,𝑛;𝑗=1,,𝑚)(5.14) can be expressed in terms of a finite number of jet coordinates,(𝜄𝜄)every family of differential forms 𝑟𝑍d𝑥𝑖𝑟,𝑟𝑍d𝑤𝑗𝑟(𝑖=1,,𝑛;𝑗=1,,𝑚)(5.15) involves only a finite number of linearly independent terms,(𝜄𝜄𝜄)every family of differential forms 𝑟𝑍d𝑥𝑖𝑟,𝑟𝑍d𝑤𝑗𝐼𝑟(𝑖=1,,𝑛;𝑗=1,,𝑚;arbitrary𝐼)(5.16) involves only a finite number of linearly independent terms.

Proof. Inclusion 𝑍𝔾 is defined by using the families (5.13) and this trivially implies (𝜄) where only the empty multi-indice 𝐼=𝜙 is involved. Then (𝜄) implies (𝜄𝜄) by using the rule 𝑍d𝑓=d𝑍𝑓. Assuming (𝜄𝜄), we may employ the commutative rule 𝐷𝑖,𝑍=𝐷𝑖𝑍𝑍𝐷𝑖=𝑎𝑖𝑖𝐷𝑖𝑎𝑖𝑖=𝐷𝑖𝑧𝑖(5.17) in order to verify identities of the kind 𝑍d𝑤𝑗𝑖=Zd𝐷𝑖𝑤𝑗=𝑍𝐷𝑖d𝑤𝑖=𝐷𝑖𝑍d𝑤𝑖𝑎𝑖𝑖𝐷𝑖𝑤𝑗(5.18) and in full generality identities of the kind 𝑘𝑍d𝑤𝑗𝐼=𝑎𝐼𝐼,𝑘𝐷𝐼𝑘𝑍d𝑤𝑗sumwith𝑘||𝐼𝑘,||||𝐼||(5.19) with unimportant coefficients, therefore (𝜄𝜄𝜄) follows. Finally (𝜄𝜄𝜄) obviously implies the primary requirement on the families (5.13).

This is not a whole story. The requirements can be expressed only in terms of the structural contact forms. With this final result, the algorithms [1013] for determination of all individual morphisms can be closely simulated in order to obtain the algorithm for the determination of all groups 𝐦(𝜆) of morphisms, see Section 10 below.

Theorem 5.5 (technical theorem). Let (5.6) be a variation of the jet space. Then 𝑍𝔾 if and only if every family 𝑟𝑍𝜔𝑗𝑟(𝑗=1,,𝑚)(5.20) involves only a finite number of linearly independent terms.

Some nontrivial preparation is needful for the proof. Let Θ be a finite-dimensional module of 1-forms (on the space 𝐌(𝑚,𝑛) but the underlying space is irrelevant here). Let us consider vector fields 𝑋 such that 𝑓𝑋ΘΘ for all functions 𝑓. Let moreover AdjΘ be the module of all forms 𝜑 satisfying 𝜑(𝑋)=0 for all such 𝑋. Then AdjΘ has a basis consisting of total differentials of certain functions 𝑓1,,𝑓𝐾 (the Frobenius theorem), and there is a basis of module Θ which can be expressed in terms of functions 𝑓1,,𝑓𝐾. Alternatively saying, (an appropriate basis of) the Pfaffian system 𝜗=0 (𝜗Θ) can be expressed only in terms of functions 𝑓1,,𝑓𝐾. This result frequently appears in Cartan's work, but we may refer only to [9, 18, 19] and to the appendix below for the proof.

Module AdjΘ is intrinsically related to Θ: if a mapping 𝐦 preserves Θ then 𝐦 preserves AdjΘ. In particular, assuming 𝐦(𝜆)ΘΘ,then𝐦(𝜆)AdjΘAdjΘ(5.21) is true for a group 𝐦(𝜆). In terms of 𝑇 of the group 𝐦(𝜆), we have equivalent assertion 𝑍ΘΘimplies𝑍AdjΘAdjΘ(5.22) and therefore 𝑟𝑍AdjΘAdjΘ for all 𝑟. The preparation is done.

Proof. Let Θ be the module generated by all differential forms 𝑟𝑍𝜔𝑗 (𝑗=1,,𝑚; 𝑟=0,1,). Assuming finite dimension of module Θ, we have module AdjΘ and clearly 𝑍ΘΘ whence 𝑟𝑍AdjΘAdjΘ (𝑟=0,1,). However AdjΘ involves both the differentials d𝑥1,,d𝑥𝑛 (see below) and the forms 𝜔1,,𝜔𝑚. Point (𝜄𝜄) of previous Lemma 5.4 implies 𝑍𝔾. The converse is trivial.
In order to finish the proof, let us on the contrary assume that AdjΘ  does not contain all differentials d𝑥1,,d𝑥𝑛. Alternatively saying, the Pfaffian system 𝜗=0 (𝜗Θ) can be expressed in terms of certain functions 𝑓1,,𝑓𝐾 such that d𝑓1==d𝑓𝐾=0 does not imply d𝑥1==d𝑥𝑛=0. On the other hand, it follows clearly that maximal solutions of the Pfaffian system can be expressed only in terms of functions 𝑓1,,𝑓𝐾 and therefore we do not need all independent variables 𝑥1,,𝑥𝑛. This is however a contradiction: the Pfaffian system consists of contact forms and involves the equations 𝜔1==𝜔𝑛=0. All independent variables are needful if we deal with the common classical solutions 𝑤𝑗=𝑤𝑗(𝑥1,,𝑥𝑛).

The result can be rephrased as follows.

Theorem 5.6. Let Ω0Ω(𝑚,𝑛) be the submodule of all zeroth-order contact forms 𝑎𝜔=𝑗𝜔𝑗 and 𝑍 be a variation of the jet structure. Then 𝑍𝔾 if and only if dim𝑟𝑍Ω0<.

6. On the Multiparameter Case

Let us temporarily denote by 𝕍 the family of all infinitesimal variations (5.6) of the jet structure. Then 𝕍+𝕍𝕍, 𝑐𝕍𝕍 (𝑐), [𝕍,𝕍]𝕍, and it follows that 𝕍 is an infinite-dimensional Lie algebra (coefficients in ). On the other hand, if 𝑍𝕍 and 𝑓𝑍𝕍 for certain 𝑓 then 𝑓 is a constant. (Briefly saying: the conical variations of the total jet space do not exist. We omit easy direct proof.) It follows that only the common Lie algebras over are engaged if we deal with morphisms of the jet spaces 𝐌(𝑚,𝑛).

Theorem 6.1. Let 𝒢𝕍 be a finite-dimensional Lie subalgebra. Then 𝒢𝔾 if and only if there exists a basis of 𝒢 that is lying in 𝔾.

The proof is elementary and may be omitted. Briefly saying, Theorem 4.2 (coefficients in ) turns into quite other and much easier Theorem 6.1 (coefficients in ).

7. The Order-Preserving Groups in Jet Space

Passing to particular examples from now on, we will briefly comment some well-known classical results for the sake of completeness.

Let Ω𝑙Ω(𝑚,𝑛) be the submodule of all contact forms 𝑎𝜔=𝑗𝐼𝜔𝑗𝐼 (sum with |𝐼|𝑙) of the order 𝑙 at most. A morphism (5.5) and the infinitesimal variation (5.6) are called order preserving if𝐦(𝜆)Ω𝑙Ω𝑙,𝑍Ω𝑙Ω𝑙,(7.1) respectively, for a certain 𝑙=0,1,(equivalently: for all 𝑙, see Lemmas 9.1 and 9.2 below). Due to the fundamental Lie-Bäcklund theorem [1, 3, 6, 1013], this is possible only in the pointwise case or in the Lie's contact transformation case. In quite explicit terms: assuming (7.1) then either functions 𝐺𝑖, 𝐺𝑗, 𝑧𝑖, 𝑧𝑗 (empty 𝐼=𝜙) in formulae (5.5) and (5.6) are functions only of the zeroth-order jet variables 𝑥𝑖, 𝑤𝑗 or, in the second case, we have 𝑚=1 and all functions 𝐺𝑖, 𝐺1, 𝐺1𝑖, 𝑧𝑖, 𝑧1, 𝑧1𝑖 contain only the zeroth- and first-order variables 𝑥𝑖, 𝑤1, 𝑤1𝑖.

A somewhat paradoxically, short proofs of this fundamental result are not easily available in current literature. We recall a tricky approach here already applied in [1013], to the case of the order-preserving morphisms. The approach is a little formally improved and appropriately adapted to the infinitesimal case.

Theorem 7.1 (infinitesimal Lie-Bäcklund). Let a variation 𝑍 preserve a submodule Ω𝑙Ω(𝑚,𝑛) of contact forms of the order 𝑙 at most for a certain 𝑙. Then 𝑍𝔾 and either 𝑍 is an infinitesimal point transformation or 𝑚=1 and 𝑍 is the infinitesimal Lie's contact transformation.

Proof. We suppose 𝑍Ω𝑙Ω𝑙. Then 𝑟𝑍Ω0𝑟𝑍Ω𝑙Ω𝑙 therefore 𝑍𝔾 by virtue of Theorem 5.5. Moreover 𝑍Ω𝑙1Ω𝑙1,,𝑍Ω0Ω0 by virtue of Lemma 9.2 below. So we have 𝑍𝜔𝑗=𝑎𝑗𝑗𝜔𝑗𝑗,𝑗.=1,,𝑚(7.2) Assuming 𝑚=1, then (7.2) turns into the classical definition of Lie's infinitesimal contact transformation. Assume 𝑚2. In order to finish the proof we refer to the following result which implies that 𝑍 is indeed an infinitesimal point transformation.

Lemma 7.2. Let 𝑍 be a vector field on the jet space 𝐌(𝑚,𝑛) satisfying (7.2) and 𝑚2. Then 𝑍𝑥𝑖=𝑧𝑖,𝑥𝑖,𝑤𝑗,,𝑍𝑤𝑗=𝑧𝑗,𝑥𝑖,𝑤𝑗,(𝑖=1,,𝑛;𝑗=1,,𝑚)(7.3) are functions only of the point variables.

Proof. Let us introduce module Θ of (𝑚+2𝑛)-forms generated by all forms of the kind 𝜔1𝜔𝑚d𝜔𝑗1𝑛1d𝜔𝑗𝑘𝑛𝑘=d𝑤1d𝑤𝑚d𝑥1d𝑥𝑛±d𝑤𝑗1𝑖1d𝑤𝑗𝑛𝑖𝑛,(7.4) where 𝑛𝑘=𝑛. Clearly Θ=(Ω0)𝑚(dΩ0)𝑛. The inclusions 𝑍Ω0Ω0,𝑍dΩ0=d𝑍Ω0+Ω0dΩ0+Ω0(7.5) are true by virtue of (7.2) and imply 𝑍ΘΘ.
Module Θ vanishes when restricted to certain hyperplanes, namely, just to the hyperplanes of the kind 𝑎𝜗=𝑖d𝑥𝑖+𝑎𝑗d𝑤𝑗=0(7.6) (use 𝑚2 here). This is expressed by Θ𝜗=0 and it follows that 0=𝑍(Θ𝜗)=𝑍Θ𝜗+Θ𝑍𝜗=Θ𝑍𝜗.(7.7) Therefore 𝑍𝜗 again is such a hyperplane: 𝑍𝜗0 (mod all d𝑥𝑖 and d𝑤𝑗). On the other hand, 𝑍𝑎𝜗𝑖d𝑧𝑖+𝑎𝑗d𝑧𝑗modalld𝑥𝑖andd𝑤𝑗(7.8) and it follows that d𝑧𝑖, d𝑧𝑗0.

There is a vast literature devoted to the pointwise transformations and symmetries so that any additional comments are needless. On the other hand, the contact transformations are more involved and less popular. They explicitly appear on rather peculiar and dissimilar occasions in actual literature [20, 21]. However, in reality the groups of Lie contact transformations are latently involved in the classical calculus of variations and provide the core of the Hilbert-Weierstrass extremality theory of variational integrals.

8. Digression to the Calculus of Variations

We establish the following principle.

Theorem 8.1 (metatheorem). The geometries of nondegenerate local one-parameter groups of Lie contact transformations (𝒞𝒯) and of nondegenerate first-order one-dimensional variational integrals (𝒱) are identical. In particular, the orbits of a given 𝒞𝒯 group are extremals of appropriate 𝒱 and conversely.

Proof. The 𝒞𝒯 groups act in the jet space 𝐌(1,𝑛) equipped with the contact module Ω(1,𝑛). Then the abbreviations 𝑤𝐼=𝑤1𝐼,𝜔𝐼=𝜔1𝐼=d𝑤𝐼𝑤𝐼𝑖d𝑥𝑖𝑧𝑍=𝑖𝜕𝜕𝑥𝑖+𝑧1𝐼𝜕𝜕𝑤𝐼(8.1) are possible. Let us recall the classical approach [22, 23]. The Lie contact transformations defined by certain formulae 𝐦𝑥𝑖=𝐺𝑖(),𝐦𝑤=𝐺1(),𝐦𝑤𝑖=𝐺1𝑖𝑥()()=1,,𝑥𝑛,𝑤,𝑤1,,𝑤𝑛(8.2) preserve the Pfaffian equation 𝑤𝜔=d𝑤𝑖d𝑥𝑖=0 or (equivalently) the submodule Ω0Ω(1,𝑛) of zeroth-order contact forms. Explicit formulae are available in literature. We are interested in one-parameter local 𝒞𝒯 groups of transformations 𝐦(𝜆)(𝜀<𝜆<𝜀) which are “nondegenerate" in a sense stated below and then the explicit formulae are not available yet. On the other hand, our 𝒱 with smooth Lagrangian Ł Ł𝑡,𝑦1,,𝑦𝑛,𝑦1,,𝑦𝑛𝑦𝑑𝑡𝑖=𝑦𝑖(𝑡),=𝑑𝜕𝑑𝑡,det2Ł𝜕𝑦𝑖𝜕𝑦𝑗0(8.3) to appear later, involves variables from quite other jet space 𝐌(𝑛,1) with coordinates denoted 𝑡 (the independent variable), 𝑦1,,𝑦𝑛 (the dependent variables) and higher-order jet variables like 𝑦𝑖, 𝑦𝑖 and so on.
We are passing to the topic proper. Let us start in the space 𝐌(1,𝑛) with 𝒞𝒯 groups. One can check that vector field (5.6) is infinitesimal 𝒞𝒯 if and only if 𝑄𝑍=𝑤𝑖𝜕𝜕𝑥𝑖+𝑤𝑄𝑖𝑄𝑤𝑖𝜕+𝑄𝜕𝑤𝑥𝑖+𝑤𝑖𝑄𝑤𝜕𝜕𝑤𝑖+,(8.4) where the function 𝑄=𝑄(𝑥1,,𝑥𝑛,𝑤,𝑤1,,𝑤𝑛) may be arbitrarily chosen.
“Hint: we have, by definition 𝑍𝑧𝜔=𝑍d𝜔+d𝜔(𝑍)=𝑖𝜔𝑖𝜔𝑖(𝑍)d𝑥𝑖+d𝑄Ω0,(8.5) where 𝑄=𝑄(𝑥1,,𝑥𝑛,𝑤,𝑤1,,𝑤𝑛,)=𝜔(𝑍)=𝑧1𝑤𝑖𝑧𝑖, 𝐷d𝑄=𝑖𝑄d𝑥𝑖+𝜕𝑄𝜕𝑤𝜔+𝜕𝑄𝜕𝑤𝑖𝜔𝑖(8.6) whence immediately 𝑧𝑖=𝜕𝑄/𝜕𝑤𝑖, 𝑧1𝑤=𝑄+𝑖𝑧𝑖𝑤=𝑄𝑖𝜕𝑄/𝜕𝑤𝑖,   𝜕𝑄/𝜕𝑤𝐼=0 if |𝐼|1 and formula (8.4) follows.”
Alas, the corresponding Lie system (not written here) is not much inspirational. Let us however consider a function 𝑤=𝑤(𝑥1,,𝑥𝑛) implicitly defined by an equation 𝑉(𝑥1,,𝑥𝑛,𝑤)=0. We may suppose that the transformed function 𝐦(𝜆)𝑤 satisfies the equation 𝑉𝑥1,,𝑥𝑛,𝐦(𝜆)𝑤=𝜆(8.7) without any loss of generality. In infinitesimal terms 1=𝜕(𝑉𝜆)𝑄𝜕𝜆=𝑍(𝑉𝜆)=𝑤𝑖𝑉𝑥𝑖+𝑤𝑄𝑖𝑄𝑤𝑖𝑉𝑤.(8.8) However 𝑤𝑖=𝜕𝑤/𝜕𝑥𝑖=𝑉𝑥𝑖/𝑉𝑤 may be inserted here, and we have the crucial Jacobi equation 𝑥1=𝑄1,,𝑥𝑛𝑉,𝑤,𝑥1𝑉𝑤𝑉,,𝑥𝑛𝑉𝑤𝑉𝑤(8.9) (not involving 𝑉) which can be uniquely rewritten as the Hamilton-Jacobi (𝒥) equation 𝑉𝑤𝑥+1,,𝑥𝑛,𝑤,𝑝1,,𝑝𝑛𝑝𝑖=𝑉𝑥𝑖(8.10) in the “nondegenerate" case 𝑄𝑤𝑖𝑉𝑥𝑖1. Let us recall the characteristic curves [22, 23] of the 𝒥 equation given by the system d𝑤1=d𝑥𝑖𝑝𝑖=d𝑝𝑖𝑥𝑖=d𝑉𝑝+𝑖𝑝𝑖.(8.11) The curves may be interpreted as the orbits of the group 𝐦(𝜆). (Hint: look at the well-known classical construction of the solution 𝑉 of the Cauchy problem [22, 23] in terms of the characteristics. The initial Cauchy data are transferred just along the characteristics, i.e., along the group orbits.) Assume moreover the additional condition det(𝜕2/𝜕𝑝𝑖𝜕𝑝𝑗)0. We may introduce variational integral (8.3) with the Lagrange function Ł given by the familiar identities 𝑝Ł+=𝑖𝑦𝑖(8.12) with interrelations 𝑡=𝑤,𝑦𝑖=𝑥𝑖,𝑦𝑖=𝑝𝑖,𝑝𝑖=Ł𝑦𝑖(𝑖=1,,𝑛)(8.13) between variables 𝑡, 𝑦𝑖, 𝑦𝑖 of the space 𝐌(𝑛,1) and variables 𝑥𝑖, 𝑤, 𝑤𝑖 of the space 𝐌(1,𝑛). Since (8.11) may be regarded as a Hamiltonian system for the extremals of 𝒱, the metatheorem is clarified.

Remark 8.2. Let us recall the Mayer fields of extremals for the 𝒱 since they provide the true sense of the above construction. The familiar Poincaré-Cartan form Ł̆𝜑=Łd𝑡+𝑦𝑖d𝑦𝑖𝑦𝑖𝑝d𝑡=d𝑡+𝑖d𝑦𝑖(8.14) is restricted to appropriate subspace 𝑦𝑖=𝑔𝑖(𝑡,𝑦1,,𝑦𝑛) (𝑖=1,,𝑛;the slope field) in order to become a total differential ||̆𝜑𝑦𝑖=𝑔𝑖=d𝑉𝑡,𝑦1,,𝑦𝑛=𝑉𝑡𝑉d𝑡+𝑦𝑖d𝑦𝑖(8.15) of the action 𝑉. We obtain the requirements 𝑉𝑡=, 𝑉𝑦𝑖=𝑝𝑖 identical with (8.10). In geometrical terms: transformations of a hypersurface 𝑉=0 by means of 𝒞𝒯 group may be identified with the level sets 𝑉=𝜆 (𝜆) of the action of a Mayer fields of extremals.
The last statement is in accordance with (8.11) where 𝑝d𝑉=+𝑖𝑝𝑖𝑝d𝑤=+𝑖𝑦𝑖d𝑡=Łd𝑡,(8.16) use the identifications (8.13) of coordinates. This is the classical definition of the action 𝑉 in a Mayer field. We have moreover clarified the additive nature of the level sets 𝑉=𝜆: roughly saying, the composition with 𝑉=𝜇 provides 𝑉=𝜆+𝜇 (see Figure 3(c)) and this is caused by the additivity of the integral Ł𝑑𝑡 calculated along the orbits.

On this occasion, the wave enveloping approach to 𝒞𝒯 groups is also worth mentioning.

Lemma 8.3 (see [1013]). Let 𝑊(𝑥1,,𝑥𝑛,𝑤,𝑥1,,𝑥𝑛,𝑤) be a function of 2𝑛+2 variables. Assume that the system 𝑊=𝐷1𝑊==𝐷𝑛𝑊=0 admits a unique solution 𝑥𝑖=𝐹𝑖,𝑥𝑖,𝑤,𝑤𝑖,,𝑤=𝐹1,𝑥𝑖,𝑤,𝑤𝑖,(8.17) by applying the implicit function theorem and analogously the system 𝑊=𝐷1𝑊==𝐷𝑛𝑊=0   (where 𝐷𝑖=𝜕/𝜕𝑥𝑖+𝑤𝑖𝜕/𝜕𝑤) admits a certain solution 𝑥𝑖=𝐹𝑖,𝑥𝑖,𝑤,𝑤𝑖,,𝑤=𝐹1,𝑥𝑖,𝑤,𝑤𝑖.,(8.18) Then 𝐦𝑥𝑖=𝐹𝑖,   𝐦𝑤=𝐹1 provides a Lie 𝒞𝒯 and (𝐦1)𝑥𝑖=𝐹𝑖, (𝐦1)𝑤=𝐹1 is the inverse.

In more generality, if function 𝑊 in Lemma 8.3 moreover depends on a parameter 𝜆, we obtain a mapping 𝐦(𝜆) which is a certain 𝒞𝒯 involving a parameter 𝜆 and the inverse 𝐦(𝜆)1. In favourable case (see below) this 𝐦(𝜆) may be even a 𝒞𝒯 group. The geometrical sense is as follows. Equation 𝑊=0 with 𝑥𝑖, 𝑤 kept fixed represents a wave in the space 𝑥𝑖, 𝑤 (Figure 3(a)).

The total system 𝑊=𝐷1𝑊==𝐷𝑛𝑊=0 provides the intersection (envelope) of infinitely close waves (Figure 3(b)) with the resulting transform, the focus point 𝐦 (or 𝐦(𝜆) if the parameter 𝜆 is present). The reverse waves with the role of variables interchanged gives the inversion. Then the group property holds true if the waves can be composed (Figure 3(c)) within the parameters 𝜆, 𝜇, but this need not be in general the case.

Let us eventually deal with the condition ensuring the group composition property. Without loss of generality, we may consider the 𝜆-depending wave𝑊𝑥1,,𝑥𝑛,𝑤,𝑥1,,𝑥𝑛,𝑤𝜆=0.(8.19) If 𝑥𝑖, 𝑤 are kept fixed, the previous results may be applied. We obtain a group if and only if the 𝒥 equation (8.10) holds true, therefore𝑊𝑤𝑥+1,,𝑥𝑛,𝑤,𝑊𝑥1,,𝑊𝑥𝑛=0.(8.20) The existence of such function means that functions 𝑊𝑤,𝑊𝑥1,,𝑊𝑥𝑛  of dashed variables are functionally dependent whence𝑊det𝑤𝑤𝑊𝑤𝑥𝑖𝑊𝑥𝑖𝑤𝑊𝑥𝑖𝑥𝑖𝑊=0,det𝑥𝑖𝑥𝑖0.(8.21) The symmetry 𝑥𝑖,𝑤𝑥𝑖,𝑤 is not surprising here since the change 𝜆𝜆 provides the inverse mapping: equations 𝑊,𝑥𝑖,𝑤,,𝑥𝑖,𝑤=𝜆,𝑊,𝑥𝑖,𝑤,,𝑥𝑖,𝑤=𝜆(8.22) are equivalent. In particular, it follows that 𝑊,𝑥𝑖,𝑤,,𝑥𝑖,𝑤=𝑊,𝑥𝑖,𝑤,,𝑥𝑖,𝑤,𝑊,𝑥𝑖,𝑤,,𝑥𝑖,𝑤=0(8.23) and the wave 𝑊𝜆=0 corresponds to the Mayer central field of extremals.

Summary 4. Conditions (8.21) ensure the existence of𝒥 equation (8.20) for the 𝜆-wave (8.19) and therefore the group composition property of waves (8.19) in the nondegenerate case det(𝜕2/𝜕𝑝𝑖𝜕𝑝𝑗)0.

Remark 8.4. A reasonable theory of Mayer fields of extremals and Hamilton-Jacobi equations can be developed also for the constrained variational integrals (the Lagrange problem) within the framework of jet spaces, that is, without the additional Lagrange multipliers [9, Chapter 3]. It follows that there do exist certain groups of generalized Lie's contact transformations with differential constraints.

9. On the Order-Destroying Groups in Jet Space

We recall that in the order-preserving case, the filtrationΩ(𝑚,𝑛)Ω0Ω1Ω(𝑚,𝑛)=Ω𝑙(9.1) of module Ω(𝑚,𝑛) is preserved (Figure 4(a)). It follows that certain invariant submodules Ω𝑙Ω(𝑚,𝑛) are a priori prescribed which essentially restricts the store of the symmetries (the Lie-Bäcklund theorem). The order-destroying groups also preserve certain submodules of Ω(𝑚,𝑛) due to approximation results, however, they are not known in advance (Figure 4(b)) and appear after certain saturation (Figure 4(c)) described in technical theorem 5.1.

The saturation is in general a toilsome procedure. It may be simplified by applying two simple principles.

Lemma 9.1 (going-up lemma). Let a group of morphisms 𝐦(𝜆) preserve a submodule ΘΩ(𝑚,𝑛). Then also the submodule Θ+𝐷𝑖ΘΩ(𝑚,𝑛)(9.2) is preserved.

Proof. We suppose 𝑍ΘΘ. Then 𝑍Θ+𝐷𝑖Θ=𝑍Θ+𝐷𝑖𝑍𝐷Θ𝑖𝑧𝑖𝐷𝑖ΘΘ+𝐷𝑖Θ(9.3) by using the commutative rule (5.17).

Lemma 9.2 (going-down lemma). Let the group of morphisms 𝐦(𝜆) preserve a submodule ΘΩ(𝑚,𝑛). Let ΘΘ be the submodule of all 𝜔Θ satisfying 𝐷𝑖𝜔Θ (𝑖=1,,𝑛). Then Θ is preserved, too.

Proof. Assume 𝜔Θ hence 𝐷𝑖𝜔Θ. Then 𝐷𝑖𝑍𝜔=𝑍𝐷𝑖𝜔+𝐷𝑖𝑧𝑖𝐷𝑖𝜔Θ hence 𝑍𝜔Θ and Θ is preserved.

We are passing to illustrative examples.

Example 9.3. Let us consider the vector field (the variation of jet structure) 𝑧𝑍=𝑗𝐼𝜕𝜕𝑤𝑗𝐼𝑧𝑗𝐼=𝐷𝐼𝑧𝑗,𝐷𝐼=𝐷𝑖1𝐷𝑖𝑛,(9.4) see (5.6) and (5.10) for the particular case 𝑧𝑖=0. Then 𝑍𝑟𝑥𝑖=0  (𝑖=1,,𝑛) and the sufficient requirement 𝑍2𝑤𝑗=0 (𝑗=1,,𝑚) ensures 𝑍𝔾, see (𝜄) of Lemma 5.4. We will deal with the linear case where 𝑧𝑗=𝑎𝑗𝑗𝑖𝑤𝑗𝑖𝑎𝑗𝑗𝑖(9.5) is supposed. Then 𝑍2𝑤𝑗=𝑍𝑧𝑗=𝑎𝑗𝑗𝑖𝑧𝑗𝑖=𝑎𝑗𝑗𝑖𝑎𝑗𝑗𝑖𝑤𝑗𝑖𝑖=0(9.6) identically if and only if 𝑗𝑎𝑗𝑗𝑖𝑎𝑗𝑗𝑖+𝑎𝑗𝑗𝑖𝑎𝑗𝑗𝑖𝑖=0,𝑖=1,,𝑛;𝑗,𝑗,𝑗.=1,,𝑚(9.7) This may be expressed in terms of matrix equations 𝐴𝑖𝐴𝑖=0𝑖,𝑖=1,,𝑛;𝐴𝑖=𝑎𝑗𝑗𝑖(9.8) or, in either of more geometrical transcriptions 𝐴2𝜆=0,Im𝐴Ker𝐴𝐴=𝑖𝐴𝑖,𝜆𝑖,(9.9) where 𝐴 is regarded as (a matrix of an) operator acting in 𝑚-dimensional linear space and depending on parameters 𝜆1,,𝜆𝑛. We do not know explicit solutions 𝐴 in full generality, however, solutions 𝐴 such that Ker𝐴 does not depend on the parameters 𝜆1,,𝜆𝑛 can be easily found (and need not be stated here). The same approach can be applied to the more general sufficient requirement 𝑍𝑟𝑤𝑗=0  (𝑗=1,,𝑚; fixed 𝑟) ensuring 𝑍𝔾. If 𝑟𝑛, the requirement is equivalent to the inclusion 𝑍𝔾.

Example 9.4. Let us consider vector field (5.6) where 𝑧1==𝑧𝑚=0. In more detail, we take 𝑧𝑍=𝑖𝜕𝜕𝑥𝑖+𝑧𝑗𝑖𝜕𝜕𝑤𝑗𝑖𝑧+𝑗𝑖𝑤=𝑗𝑖𝐷𝑖𝑧𝑖.(9.10) Then 𝑍𝑟𝑤𝑗=0 and we have to deal with functions 𝑍𝑟𝑥𝑖 in order to ensure the inclusion 𝑍𝔾. This is a difficult task. Let us therefore suppose 𝑧1=𝑧,𝑥𝑖,𝑤𝑗,𝑤𝑗1,,𝑧𝑘=𝑐𝑘(𝑘=2,,𝑛).(9.11) Then 𝑍𝑥𝑘=0  (𝑘=2,,𝑛) and 𝑍2𝑥1=𝑍𝑧=𝜕𝑧𝜕𝑥𝑖𝑧𝑖+𝜕𝑧𝜕𝑤𝑗1𝑧𝑗1,(9.12) where 𝑧𝑗1=𝑤𝑗1𝐷1𝑧=𝑤𝑗1𝜕𝑧𝜕𝑥1+𝜕𝑧𝜕𝑤𝑗𝑤𝑗1+𝜕𝑧𝜕𝑤𝑗1𝑤𝑗11.(9.13) The second-order summand 𝑍2𝑥1=+𝜕𝑧𝜕𝑤𝑗1𝑧𝑗1=𝜕𝑧𝜕𝑤𝑗1𝑤𝑗1𝜕𝑧𝜕𝑤𝑗1𝑤𝑗11(9.14) identically vanishes for the choice 𝑧=𝑓,𝑥𝑖,𝑤𝑗,𝑢𝑙𝑢,𝑙=𝑤𝑙1𝑤11;𝑙=2,,𝑚(9.15) as follows by direct verification. Quite analogously 𝑍𝑢𝑙𝑤=𝑍𝑙1𝑤11=𝑧𝑙11𝑤11𝑧11𝑤𝑙1𝑤112=𝑤𝑙11𝑤11+𝑤11𝑤𝑙1𝑤112𝐷1𝑧=0.(9.16) It follows that all functions 𝑍𝑟𝑥𝑖, 𝑍𝑟𝑤𝑗 can be expressed in terms of the finite family of functions 𝑥𝑖 (𝑖=1,,𝑛), 𝑤𝑗 (𝑗=1,,𝑚), 𝑢𝑙 (𝑙=2,,𝑚) and therefore 𝑍𝔾.

Remark 9.5. On this occasion, let us briefly mention the groups generated by vector fields 𝑍 of the above examples. The Lie system of the vector field (9.4) and (9.5) reads d𝐺𝑖d𝜆=0,d𝐺𝑗=𝑎d𝜆𝑗𝑗𝑖𝐺𝑗𝑖(𝑖=1,,𝑛;𝑗=1,,𝑚),(9.17) where we omit the prolongations. It is resolved by 𝐺𝑖=𝑥𝑖,𝐺𝑗=𝑤𝑗𝑎+𝜆𝑗𝑗𝑖𝑤𝑗𝑖(𝑖=1,,𝑛;𝑗=1,,𝑚)(9.18) as follows either by direct verification or, alternatively, from the property 𝑍2𝑥𝑖=𝑍𝑧𝑖=0  (𝑖=1,,𝑛) which implies d𝑎𝑗𝑗𝑖𝐺𝑗𝑖𝑎d𝜆=0,𝑗𝑗𝑖𝐺𝑗𝑖=𝑎𝑗𝑗𝑖𝐺𝑗𝑖|||𝜆=0=𝑎𝑗𝑗𝑖𝑤𝑗𝑖.(9.19) Quite analogously, the Lie system of the vector field (9.10), (9.11), (9.15) reads d𝐺1d𝜆=𝑓,𝐺𝑖,𝐺𝑗,𝐺𝑙1𝐺11,,d𝐺𝑘d𝜆=𝑐𝑘,d𝐺𝑗d𝜆=0(𝑘=2,,𝑛;𝑗=1,,𝑚)(9.20) and may be completed with the equations d𝐺𝑙1/𝐺11d𝜆=0(𝑙=2,,𝑚)(9.21) following from (9.16). This provides a classical self-contained system of ordinary differential equations where the common existence theorems can be applied.
The above Lie systems admit many nontrivial first integrals 𝐹, that is, functions 𝐹 that are constant on the orbits of the group. Conditions 𝐹=0 may be interpreted as differential equations in the total jet space, and the above transformation groups turn into the external generalized symmetries of such differential equations, see Section 11 below.

10. Towards the Main Algorithm

We briefly recall the algorithm [1013] for determination of all individual automorphisms 𝐦 of the jet space 𝐌(𝑚,𝑛) in order to compare it with the subsequent calculation of vector field 𝑍𝔾.

Morphisms 𝐦 of the jet structure were defined by the property 𝐦Ω(𝑚,𝑛)Ω(𝑚,𝑛). The inverse 𝐦1 exists if and only ifΩ0𝐦Ω(𝑚,𝑛),equivalentlyΩ0𝐦Ω𝑙(𝑙=𝑙(𝐦))(10.1) for appropriate term Ω𝑙(𝐦) of filtration (9.1). However𝐦Ω𝑙+1=𝐦Ω𝑙+𝐷𝑖𝐦Ω𝑙(10.2) and it follows that criterion (10.1) can be verified by repeated use of operators 𝐷𝑖. In more detail, we start with equations𝐦𝜔𝑗=𝑎𝑗𝑗𝐼𝜔𝑗𝐼=d𝐦𝑤𝑗𝐦𝑤𝑗𝑖d𝐦𝑥𝑖(10.3) with uncertain coefficients. Formulae (10.3) determine the module 𝐦Ω0. Then we search for lower-order contact forms, especially forms from Ω0, lying in 𝐦Ω𝑙 with the use of (10.2). Such forms are ensured if certain linear relations among coefficients exist. The calculation is finished on a certain level 𝑙=𝑙(𝐦) and this is the algebraic part of the algorithm. With this favourable choice of coefficients 𝑎𝑗𝑗𝐼, functions 𝐦𝑥𝑖,  𝐦𝑤𝑗 (and therefore the invertible morphism 𝐦) can be determined by inspection of the bracket in (10.3). This is the analytic part of algorithm.

Let us turn to the infinitesimal theory. Then the main technical tool is the rule (5.17) in the following transcription:𝑍𝐷𝑖=𝐷𝑖𝑍𝐷𝑖𝑧𝑖𝐷𝑖(10.4) or, when applied to basical forms𝑍𝜔𝑗𝐼𝑖=𝐷𝑖𝑍𝜔𝑗𝐼𝐷𝑖𝑧𝑖𝜔𝑗𝐼𝑖.(10.5) We are interested in vector fields 𝑍𝔾. They satisfy the recurrence (5.10) together with requirementsdim𝑟𝑍Ω0<,equivalently𝑟𝑍Ω0Ω𝑙(𝑍)(𝑟=0,1,)(10.6) for appropriate 𝑙(𝑍). Due to the recurrence (10.5) these requirements can be effectively investigated. In more detail, we start with equations𝑍𝜔𝑗=𝑎𝑗𝑗𝐼𝜔𝑗𝐼=d𝑧𝑗𝑧𝑗𝑖d𝑥𝑖𝑤𝑗𝑖d𝑧𝑖.(10.7) Formulae (10.7) determine module 𝑍Ω0. Then, choosing 𝑙(𝑍), operator 𝑍 is to be repeatedly applied and requirements (10.6) provide certain polynomial relations for the coefficients by using (10.5). This is the algebraical part of the algorithm. With such coefficients 𝑎𝑗𝑗𝐼 available, functions 𝑧𝑖=𝑍𝑥𝑖,  𝑧𝑗=𝑍𝑤𝑗 (and therefore the vector field 𝑍𝔾) can be determined by inspection of the bracket in (10.7) or, alternatively, with the use of formulae (5.12) for the particular case 𝐼=𝜙 empty𝑍𝜔𝑗=𝜕𝑧𝑗𝜕𝑤𝑗𝐼𝑤𝑗𝑖𝜕𝑧𝑖𝜕𝑤𝑗𝐼𝜔𝑗𝐼.(10.8) This is the analytic part of the algorithm.

Altogether taken, the algorithm is not easy and the conviction [7, page 121] that the “exhaustive description of integrable 𝐶-fields (fields 𝑍 in our notation) is given in [16]" is disputable. We can state only one optimistic result at this place.

Theorem 10.1. The jet spaces 𝐌(1,𝑛) do not admit any true generalized infinitesimal symmetries 𝑍𝔾.

Proof. We suppose 𝑚=1 and then (10.7) reads 𝑍𝜔1=𝑎𝐼11𝜔1𝐼=+𝑎𝐼11𝜔1𝐼𝑎𝐼110,(10.9) where we state a summand of maximal order. Assuming 𝐼=𝜙, the Lie-Bäcklund theorem can be applied and we do not have the true generalized symmetry 𝑍. Assuming 𝐼𝜙, then 𝑟𝑍𝜔1=+𝑎𝐼11𝜔1𝐼𝐼𝑟terms𝐼(10.10) by using rule (10.5) where the last summand may be omitted. It follows that (10.6) is not satisfied hence 𝑍𝔾.

Example 10.2. We discuss the simplest possible but still a nontrivial particular example. Assume 𝑚=2, 𝑛=1 and 𝑙(𝑍)=1. Let us abbreviate 𝑥=𝑥1,𝐷=𝐷1𝜕,𝑍=𝑧+𝑧𝜕𝑥𝑗𝐼𝜕𝜕𝑤𝑗𝐼(𝑗=1,2;𝐼=11).(10.11) Then, due to 𝑙(𝑍)=1, requirement (10.6) reads 𝑟𝑍Ω0Ω1(𝑟=0,1,).(10.12) In particular (if 𝑟=1) we have (10.7) written here in the simplified notation 𝑍𝜔𝑗=𝑎𝑗1𝜔1+𝑎𝑗2𝜔2+𝑏𝑗1𝜔11+𝑏𝑗2𝜔21(𝑗=1,2).(10.13) The next requirement (𝑟=2) implies the (only seemingly) stronger inclusion 2𝑍Ω0𝑍Ω0+Ω0(10.14) which already ensures (10.12) for all 𝑟 and therefore 𝑍𝔾 (easy). We suppose (10.14) from now on.
“Hint for proof of (10.14): assuming (10.12) and moreover the equality 2𝑍Ω0+𝑍Ω0+Ω0=Ω1,(10.15) it follows that 𝑍Ω13𝑍Ω0+2𝑍Ω0+𝑍Ω0Ω1(10.16) and Lie-Bäcklund theorem can be applied whence 𝑍Ω0Ω0,  𝑙(𝑍)=0 which we exclude. It follows that necessarily dim2𝑍Ω0+𝑍Ω0+Ω0<dimΩ1=4.(10.17) On the other hand dim(𝑍Ω0+Ω0)3 and the inclusion (10.14) follows.”
After this preparation, we are passing to the proper algebra. Clearly 2𝑍𝜔𝑗=+𝑏𝑗1𝑍𝜔11+𝑏𝑗2𝑍𝜔21=+𝑏𝑗1𝑏11𝜔111+𝑏12𝜔211+𝑏𝑗2𝑏21𝜔111+𝑏22𝜔211(10.18) by using the commutative rule (10.5). Due to “weaker" inclusion (10.12) with 𝑟=2, we obtain identities 𝑏𝑗1𝑏11+𝑏𝑗2𝑏21=0,𝑏𝑗1𝑏12+𝑏𝑗2𝑏22=0(𝑗=1,2).(10.19) Omitting the trivial solution, they are satisfied if either 𝑏11+𝑏22=0,𝑏12=𝑐𝑏11,𝑏11+𝑐𝑏21=0(10.20) for appropriate factor 𝑐 (where 𝑏110 and either 𝑏120 or 𝑏210 is supposed) or 𝑏11=𝑏22=0,either𝑏12=0or𝑏21=0.(10.21) We deal only with the (more interesting) identities (10.20) here. Then 𝑍𝜔1=𝑎11𝜔1+𝑎12𝜔2𝜔𝑐𝑏11+𝑐𝜔21,𝑍𝜔2=𝑎21𝜔1+𝑎22𝜔2𝜔+𝑏11+𝑐𝜔21(10.22) (abbreviation 𝑏=𝑏21) by inserting (10.20) into (10.13). It follows that 𝑍𝜔1+𝑐𝜔2=𝑎1𝜔1+𝑎2𝜔2𝑎1=𝑎11+𝑐𝑎21,𝑎2=𝑎12+𝑐𝑎22.+𝑍𝑐(10.23) It may be seen by direct calculation of 2𝑍𝜔2 that the “stronger" inclusion (10.14) is equivalent to the identity 𝑐𝑎1=𝑎2, that is, 𝑍𝜔1+𝑐𝜔2𝜔=𝑎1+𝑐𝜔2(10.24) (abbreviation 𝑎=𝑎1). Alternatively, (10.24) can be proved by using Lemma 9.2.
“Hint: denoting Θ=𝑍Ω0+Ω0, (10.14) implies 𝑍ΘΘ. Moreover 𝐷(𝜔1+𝑐𝜔2)Θ by using (10.22). Lemma 9.2 can be applied: 𝜔1+𝑐𝜔2Θ and Θ involves just all multiples of form 𝜔1+𝑐𝜔2. Therefore 𝑍(𝜔1+𝑐𝜔2)Θ is a multiple of 𝜔1+𝑐𝜔2.”
The algebraical part is concluded. We have congruences 𝑍𝜔1𝜔𝑐𝑏11+𝑐𝜔21,𝑍𝜔2𝜔𝑏11+𝑐𝜔21modΩ0(10.25) and equality 𝑍𝜔1+𝑐𝑍𝜔2+𝑍𝑐𝜔2𝜔=𝑎1+𝑐𝜔2.(10.26) If 𝑍 is a variation then these three conditions together ensure the “stronger inclusion" (10.14) hence 𝑍𝔾.
We turn to analysis. Abbreviating 𝑍𝑗𝑗𝐼=𝜕𝑧𝑗𝜕𝑤𝑗𝐼𝑤𝑗1𝜕𝑧𝜕𝑤𝑗𝐼𝑗,𝑗=1,2;𝐼=11(10.27) and employing (10.8), the above conditions (10.25) and (10.26) read 𝑍1𝑗𝐼𝜔𝑗𝐼𝜔=𝑐𝑏11+𝑐𝜔21,𝑍2𝑗𝐼𝜔𝑗𝐼𝜔=𝑏11+𝑐𝜔21||𝐼||,𝑍11𝑗𝐼+𝑐𝑍2𝑗𝐼𝜔𝑗𝐼+𝑍𝑐𝜔2𝜔=𝑎1+𝑐𝜔2.(10.28) We compare coefficients of forms 𝜔𝑗𝐼 on the level 𝑠=|𝐼|𝑠=0𝑍11+𝑐𝑍21=𝑎,𝑍12+𝑐𝑍22+𝑍𝑐=𝑎𝑐,(10.29)𝑠=1𝑍111=𝑐𝑏,𝑍112=(𝑐)2𝑏,𝑍121=𝑏,𝑍122=𝑏𝑐,𝑍11𝑗+𝑐𝑍12𝑗=0,(10.30)𝑠2𝑍𝑗𝑗𝐼=0,𝑍1𝑗𝐼+𝑐𝑍2𝑗𝐼=0.(10.31) We will successively delete the coefficients 𝑎, 𝑏, 𝑐 in order to obtain interrelations only for variables 𝑍𝑗𝑗𝐼. Clearly 𝑠=0𝑍12+𝑐𝑍22𝑍+𝑍𝑐=11+𝑐𝑍21𝑐,𝑠=1𝑍111+𝑍122=0,𝑍111𝑍122=𝑍112𝑍121,(10.32) and we moreover have three compatible equations 𝑍𝑐=111𝑍121𝑍=112𝑍122,(𝑐)2𝑍=112𝑍121(10.33) for the coefficient 𝑐. To cope with levels 𝑠2, we introduce functions 𝑄𝑗=𝜔𝑗(𝑍)=𝑧𝑗𝑤𝑗1𝑧(𝑗=1,2).(10.34) Then substitution into (10.27) with the help of (10.31) gives 𝜕𝑄𝑗𝜕𝑤𝑗𝐼=0𝑗,𝑗||𝐼=1,2;||.2(10.35) It follows moreover easily that 𝑍𝑗𝑗1=𝜕𝑄𝑗𝜕𝑤𝑗1𝑗𝑗,𝑍1𝑗𝑗=𝑧+𝜕𝑄𝑗𝜕𝑤𝑗1,𝑍𝑗𝑗=𝜕𝑄𝑗𝜕𝑤𝑗(10.36) and we have the final differential equations 𝑠=0𝜕𝑄1𝜕𝑤2+𝑐𝜕𝑄2𝜕𝑤2+𝑍𝑐=𝜕𝑄1𝜕𝑤1+𝑐𝜕𝑄2𝜕𝑤1𝑐,(10.37)𝑠=12𝑧+𝜕𝑄1𝜕𝑤11+𝜕𝑄2𝜕𝑤21=0,𝑧+𝜕𝑄1𝜕𝑤11𝑧+𝜕𝑄2𝜕𝑤21=𝜕𝑄1𝜕𝑤21𝜕𝑄2𝜕𝑤11(10.38)for the unknown functions 𝑧=𝑧𝑥,𝑤1,𝑤2,𝑤11,𝑤21,𝑄𝑗=𝑄𝑗𝑥,𝑤1,𝑤2,𝑤11,𝑤21.(10.39)The coefficient 𝑐 is determined by (10.33) and (10.36) in terms of functions 𝑄𝑗. This concludes the analytic part of the algorithm since trivially 𝑧𝑗=𝑤𝑗1𝑧+𝑄𝑗 and the vector field 𝑍 is determined.
The system is compatible: particular solutions with functions 𝑄𝑗 quadratic in jet variables and 𝑐=const. can be found as follows. Assume 𝑄𝑗=𝐴𝑗𝑤112+2𝐵𝑗𝑤11𝑤21+𝐶𝑗𝑤212(𝑗=1,2)(10.40) with constant coefficients 𝐴𝑗,𝐵𝑗,𝐶𝑗. We also suppose 𝑐 and then (10.37) is trivially satisfied.
On the other hand, (10.33) provide the requirements 𝑧+𝜕𝑄1𝜕𝑤11+𝑐𝜕𝑄2𝜕𝑤11=𝜕𝑄1𝜕𝑤21+𝑐𝑧+𝜕𝑄2𝜕𝑤21=𝜕𝑄1𝜕𝑤21+(𝑐)2𝜕𝑄2𝜕𝑤11=0(10.41) by using (10.36). If we put 𝑧=𝜕𝑄1𝜕𝑤11𝜕𝑄2𝜕𝑤21𝐴=1+𝐵1𝑤11𝐵1+𝐶2𝑤21,(10.42) then (10.38) is satisfied (a clumsy direct verification).
The above requirements turn to a system of six homogeneous linear equations (not written here) for the six constants 𝐴𝑗, 𝐵𝑗, 𝐶𝑗 (𝑗=1,2) with determinant Δ=𝑐2(𝑐28) if the values 𝑧, 𝑄1, 𝑄2 are inserted and the coefficients of 𝑤11 and 𝑤21 are compared. The roots 𝑐=0 and 𝑐=±22 of the equation Δ=0 provide rather nontrivial infinitesimal transformation 𝑍, however, we can state only the simplest result for the trivial root 𝑐=0 for obvious reason. It reads 𝑄1=𝐴1𝑤112,𝑄2=𝐴2𝑤112,𝑧=𝐴1𝑤11,𝑧1=0,𝑧2=𝑤11𝐴2𝑤11+𝐴1𝑤21,(10.43) where 𝐴1, 𝐴2 are arbitrary constants.

Remark 10.3. It follows that investigation of vector fields 𝑍𝔾 cannot be regarded for easy task and some new powerful methods are necessary, for example, better use of differential forms (involutive systems) with pseudogroup symmetries of the problem (moving frames).

11. A Few Notes on the Symmetries of Differential Equations

The external theory deals with (systems of) differential equations (𝒟) that are firmly localized in the jet spaces. This is the common approach and it runs as follows. A given finite system of 𝒟 is infinitely prolonged in order to ensure the compatibility. In general, this prolongation is a toilsome and delicate task, in particular the “singular solutions" are tacitly passed over. The prolongation procedure is expressed in terms of jet variables and as a result a fixed subspace of the (infinite-order) jet space appears which represents the 𝒟 under consideration. Then the external symmetries [2, 3, 6, 7] are such symmetries of the ambient jet space which preserve the subspace. In this sense we may speak of classical symmetries (point and contact transformations) and higher-order symmetries (which destroy the order of derivatives).

The internal theory of 𝒟 is irrelevant to the jet localization, in particular to the choice of the hierarchy of independent and dependent variables. This point of view is due to E. Cartan and actually the congenial term “diffiety" was introduced in [6, 7]. Alas, these diffieties were defined as objects locally identical with appropriate external 𝒟 restricted to the corresponding subspace of the ambient total jet space. This can hardly be regarded as a coordinate-free (or jet theory-free) approach since the model objects (external 𝒟) and the intertwining mappings (higher-order symmetries) essentially need the use of the above hard jet theory mechanisms and concepts.

In reality, the final result of prolongation, the infinitely prolonged 𝒟, can be alternatively characterized by three simple axioms as follows [8, 9, 2427].

Let 𝐌 be a space modelled on (local coordinates 1,2, as in Sections 1 and 2 above). Denote by (𝐌) the structural module of all smooth functions 𝑓 on 𝐌 (locally depending on a finite number 𝑚(𝑓) of coordinates). Let Φ(𝐌), 𝒯(𝐌) be the (𝐌)-modules of all differential 1-forms and vector fields on 𝐌, respectively. For every submodule ΩΦ(𝐌), we have the “orthogonal" submodule Ω=𝒯(𝐌) of all 𝑋 such that Ω(𝑋)=0.

Then an (𝐌)-submodule ΩΦ(𝐌) is called a diffiety if the following three requirements are locally satisfied. (𝒜)Ω is of codimension 𝑛<, equivalent is of dimension 𝑛<.Here 𝑛 is the number of independent variables. The independent variables provide the complementary module to Ω in Φ(𝐌) which is not prescribed in advance. ()dΩ0 (mod Ω), equivalent ΩΩ, equivalently: [,]. This Frobenius condition ensures the classical passivity requirement: we deal with the compatible infinite prolongation of differential equations. (𝒞)There exists filtration ΩΩ0Ω1Ω=Ω𝑙 by finite-dimensional submodules Ω𝑙Ω such that Ω𝑙Ω𝑙+1(all𝑙),Ω𝑙+1=Ω𝑙+Ω𝑙(𝑙largeenough).(11.1) This condition may be expressed in terms of a -polynomial algebra on the graded module Ω𝑙/Ω𝑙1 (the Noetherian property) and ensures the finite number of dependent variables. Filtration Ω may be capriciously modified. In particular, various localizations of Ω in jet spaces Ω(𝑚,𝑛) can be easily obtained.

The internal symmetries naturally appear. For instance, a vector field 𝑍𝒯(𝐌) is called a (universal) variation of diffiety Ω if 𝑍ΩΩ and infinitesimal symmetry if moreover 𝑍 generates a local group, that is, if and only if 𝑍𝔾.

Theorem 11.1 (technical theorem). Let 𝑍 be a variation of diffiety Ω. Then 𝑍𝔾 if and only if there is a finite-dimensional (M)-submodule ΘΩ such that 𝑟Θ=Ω,dim𝑟𝑍Θ<.(11.2)

This is exactly counterpart to Theorem 5.6: submodule ΘΩ stands here for the previous submodule Ω0Ω(𝑚,𝑛). We postpone the proof of Theorem 11.1 together with applications to some convenient occasion.

Remark 11.2. There may exist conical symmetries 𝑍 of a diffiety Ω, however, they are all lying in   and generate just the Cauchy characteristics of the diffiety [9, page 155].

We conclude with two examples of internal theory of underdetermined ordinary differential equations. The reasonings to follow can be carried over quite general diffieties without any change.

Example 11.3. Let us deal with the Monge equation 𝑑𝑥𝑑𝑡=𝑓𝑡,𝑥,𝑦,𝑑𝑦𝑑𝑡.(11.3) The prolongation can be represented as the Pfaffian system d𝑥𝑓𝑡,𝑥,𝑦,𝑦d𝑡=0,d𝑦𝑦d𝑡=0,d𝑦𝑦d𝑡=0,.(11.4) Within the framework of diffieties, we introduce space 𝐌 with coordinates 𝑡,𝑥0,𝑦0,𝑦1,𝑦2,(11.5) and submodule ΩΦ(𝐌) with generators d𝑥0𝜔𝑓d𝑡,𝑟=d𝑦𝑟𝑦𝑟+1d𝑡𝑟=0,1,;𝑓=𝑓𝑡,𝑥0,𝑦0,𝑦1.(11.6) Clearly =Ω𝒯(𝐌) is one-dimensional subspace including the vector field 𝜕𝐷=𝜕𝜕𝑡+𝑓𝜕𝑥0+𝑦𝑟+1𝜕𝜕𝑦𝑟.(11.7) One can easily find that we have a diffiety. (𝒜 and are trivially satisfied. The common order preserving filtrations where Ω𝑙 involves d𝑥0𝑓d𝑡 and 𝜔𝑟 with 𝑟𝑙 is enough for 𝒞.)
We introduce a new (standard [9]) filtration Ω where the submodule Ω𝑙Ω is generated by the forms 𝜗0=d𝑥0𝑓d𝑡𝜕𝑓𝜕𝑦1𝜔0,𝜔𝑟(𝑟𝑙1).(11.8) This is indeed a filtration since 𝐷𝜗0=d𝑓𝐷𝑓d𝑡𝐷𝜕𝑓𝜕𝑦1𝜔0𝜕𝑓𝜕𝑦1𝜔1=𝜕𝑓𝜕𝑥0d𝑥0+𝑓d𝑡𝜕𝑓𝜕𝑦0𝐷𝜕𝑓𝜕𝑦1𝜔0=𝜕𝑓𝜕𝑥0𝜗0+𝐴𝜔0𝐴=𝜕𝑓𝜕𝑦0+𝜕𝑓𝜕𝑥0𝜕𝑓𝜕𝑦1𝐷𝜕𝑓𝜕𝑦1(11.9) and (trivially) 𝐷𝜔𝑟=𝜔𝑟+1. Assuming 𝐴0 from now on (this is satisfied if 𝑓𝑦1𝑦10) every module Ω𝑙 is generated by the forms 𝜗𝑟=𝑟𝐷𝜗0 (𝑟𝑙).
The forms 𝜗𝑟 satisfy the recurrence 𝐷𝜗𝑟=𝜗𝑟+1. Then the formula 𝜗𝑟+1=𝐷𝜗𝑟=𝐷d𝜗𝑟+d𝜗𝑟(𝐷)=𝐷d𝜗𝑟(11.10) implies the congruence d𝜗𝑟d𝑡𝜗𝑟+1 (mod ΩΩ). Let 𝜕𝑍=𝑧𝜕𝑡+𝑧0𝜕𝜕𝑥0+𝑧𝑟𝜕𝜕𝑦𝑟(11.11) be a variation of Ω in the common sense 𝑍ΩΩ. This inclusion is equivalent to the congruence 𝑍𝜗𝑟=𝑍d𝜗𝑟+d𝜗𝑟(𝑍)𝜗𝑟+1(𝑍)d𝑡+𝐷𝜗𝑟(𝑍)d𝑡=0(modΩ)(11.12) whence to the recurrence 𝜗𝑟+1(𝑍)=𝐷𝜗𝑟(𝑍)(11.13) quite analogous to the recurrence (5.10), see Remark 5.3. It follows that the functions 𝑧=𝑍𝑡=d𝑡(𝑍),𝑔=𝜗0(𝑍)(11.14) can be quite arbitrarily chosen. Then functions 𝜗𝑟(𝑍)=𝐷𝑟𝑔 are determined and we obtain quite explicit formulae for the variation 𝑍. In more detail 𝑔=𝜗0(𝑍)=d𝑥0𝑓d𝑡𝜕𝑓𝜕𝑦1𝜔0(𝑍)=𝑧0𝑓𝑧𝜕𝑓𝜕𝑦1𝑧0𝑦1𝑧,𝐷𝑔=𝜗1(𝑍)𝜕𝑓𝜕𝑥0𝜗0+𝐴𝜔0(𝑍)=𝜕𝑓𝜕𝑥0𝑧𝑔+𝐴0𝑦1𝑧(11.15) and these equations determine coefficients 𝑧0 and 𝑧0 in terms of functions 𝑧 and 𝑔. Coefficients 𝑧𝑟 (𝑟1) follow by prolongation (not stated here). If moreover dim{𝑟𝑍𝜗0}𝑟<(11.16) we have infinitesimal symmetry 𝑍𝔾, see Theorem 11.1.

Example 11.4. Let us deal with the Hilbert-Cartan equation [3] 𝑑𝑦=𝑑𝑑𝑡2𝑥𝑑𝑡22.(11.17) Passing to the diffiety, we introduce space 𝐌 with coordinates 𝑡,𝑥0,𝑥1,𝑦0,𝑦1,𝑦2,(11.18) and submodule ΩΦ(𝐌) generated by forms d𝑥0𝑥1d𝑡,d𝑥1𝑦1𝜔d𝑡,𝑟=d𝑦𝑟𝑦𝑟+1d𝑡(𝑟=0,1,).(11.19) The submodule =Ω𝒯(𝐌) is generated by the vector field 𝜕𝐷=𝜕𝑡+𝑥1𝜕𝜕𝑥0+𝑦1𝜕𝜕𝑥1+𝑦𝑟+1𝜕𝜕𝑦𝑟.(11.20) We introduce the form 𝜗0=d𝑥0𝑥1d𝑡+𝐵d𝑥1𝑦11d𝑡2𝑦1𝜔0𝐵=1/𝑦1𝐷1/𝑦1(11.21) and moreover the forms 𝜗1=𝐷𝜗0𝜗=(1+𝐷𝐵){},2=𝐷𝜗1=𝐷2𝐵{}𝐶𝜔01𝐶=(1+𝐷𝐵)𝐷2𝑦1,𝜗3=+𝐶𝜔1,𝜗4=+𝐶𝜔2,(11.22) Assuming 𝐶0, we have a standard filtration Ω where the submodules Ω𝑙Ω are generated by forms 𝜗𝑟 (𝑟𝑙). Explicit formulae for variations 𝜕𝑍=𝑧𝜕𝑡+𝑧0𝜕𝜕𝑥0+𝑧1𝜕𝜕𝑥1+𝑧𝑟𝜕𝜕𝑦𝑟(11.23) can be obtained analogously as in Example 11.3 (and are omitted here). Functions 𝑧 and 𝑔=𝜗0(𝑍) can be arbitrarily chosen. Condition (11.16) ensures 𝑍𝔾.

Appendix

For the convenience of reader, we survey some results [9, 18, 19] on the modules Adj. Our reasonings are carried out in the space 𝑛 and will be true locally near generic points.

Let Θ be a given module of 1-forms and 𝐴(Θ) the module of all vector fields 𝑋 such that 𝑓𝑋ΘΘ for all functions 𝑓, see [9]. Clearly [𝑋,𝑍]Θ=𝑋𝑍𝑍𝑋ΘΘ(𝑋,𝑍𝐴(Θ))(A.1) and it follows that identity 𝑓[]=[]𝑋,𝑌𝑋,𝑍+𝑋𝑓𝑌(𝑋,𝑌𝐴(Θ);𝑍=𝑓𝑌)(A.2) implies 𝑓[𝑋,𝑌]ΘΘ whence [𝐴(Θ),𝐴(Θ)]𝐴(Θ).

Let Θ be of a finite dimension 𝐼. The Frobenius theorem can be applied, and it follows that module AdjΘ=𝐴(Θ) (of all forms 𝜑 satisfying 𝜑(𝐴(Θ))=0) has a certain basis d𝑓1,,d𝑓𝐾 (𝐾𝐼).

On the other hand, identity 𝑓𝑋𝜗=𝑓𝑋d𝜗+d(𝑓𝜗(𝑋))=𝑓𝑋𝜗+𝜗(𝑋)𝜗(A.3) implies that 𝑋𝐴(Θ) if and only if 𝜗(𝑋)=0,𝑋d𝜗Θ(𝜗Θ)(A.4) which is the classical definition, see [2]. In particular ΘAdjΘ so we may suppose the generators 𝜗𝑖=d𝑓𝑖+𝑔𝑖𝐼+1d𝑓𝐼+1++𝑔𝑖𝐾d𝑓𝐾Θ(𝑖=1,,𝐼)(A.5) of module Θ. Recall that 𝑋𝑓𝑘=0   (𝑘=1,,𝐾;𝑋𝐴(Θ)) whence 𝑋𝜗𝑖=𝑋𝑔𝑖𝐼+1d𝑓𝐼+1++𝑋𝑔𝑖𝐾d𝑓𝐾Θ(A.6) and this implies 𝑋𝑔𝑖𝐼+1==𝑋𝑔𝑖𝐾=0. It follows that d𝑔𝑖𝐼+1,,d𝑔𝑖𝐾AdjΘ(𝑖=1,,𝐼)(A.7) and therefore all coefficients 𝑔𝑖𝑘 depend only on variables 𝑓1,,𝑓𝐾.

Acknowledgment

This research has been conducted at the Department of Mathematics as part of the research project CEZ: Progressive reliable and durable structures, MSM 0021630519.