Orbits and attainable Hamiltonian diffeomorphisms of mechanical Liouville equations
Abstract
We study the approximate controllability problem for Liouville transport equations along a mechanical Hamiltonian vector field. Such PDEs evolve inside the orbit
where is the initial density and is the group of Hamiltonian diffeomorphisms of the cotangent bundle manifold . The approximately reachable densities from are thus contained in , where the closure is taken with respect to the -topology. Our first result is a characterization of when the manifold is the Euclidean space or the torus of arbitrary dimension: is the set of all the densities whose sub- and super-level sets have the same measure as those of . This result is an approximate version, in the case of , of a theorem by J. Moser (Trans. Am. Math. Soc. 120: 286-294, 1965) on the group of diffeomorphisms.
We then present two examples of systems, respectively on and , where the small-time approximately attainable diffeomorphisms coincide with , respectively at the level of the group and at the level of the densities.
The proofs are based on the construction of Hamiltonian diffeomorphisms that approximate suitable permutations of finite grids, and Poisson bracket techniques.
Keywords: Group of Hamiltonian diffeomorphisms, controllability, Liouville transport equation.
Mathematics Subject Classification 2020: 93C20, 35Q49, 58D05, 37J39
Contents
1 Introduction
1.1 The model
Let be a smooth -dimensional boundaryless Riemannian manifold. Its cotangent bundle is equipped with canonical symplectic and volume forms, that in local position and momentum coordinates can be expressed, respectively, as and (-times). For any measurable set we denote by its volume.
In this article we study the controllability of Liouville equations describing the transportation of a density along a Hamiltonian vector field. Let , , be, respectively, the space of smooth () functions, Hamiltonian vector fields, and Hamiltonian diffeomorphisms. The space of real-valued functions is a Lie algebra equipped with the Poisson bracket With any smooth function one can associate the corresponding Hamiltonian vector field . In particular, in this work we shall restrict ourselves to mechanical Hamiltonians, i.e. , Hamiltonians of the form
(1) |
which are the sum of a drift term, denoted as (the sum of kinetic and potential energy), and a potential controlled through , were is used to denote piecewise constant functions. We suppose in the following that for each the vector field is complete.
Let us fix . Given , we consider the transportation of an initial density along the piecewise constant Hamiltonian vector field , given by the Liouville equation
(2) |
Equation (2) more explicitly reads as the partial differential equation
This is a bilinear (hence, non-linear) control problem, with infinite-dimensional state space .
Since the vector field is complete for every , the Liouville equation is globally-in-time well posed, meaning that for any there exists a unique mild solution of (2). In this article we are interested in studying the small-time reachable sets of (2).
Definition 1 (Small-time reachable densities).
A density is said to be small-time reachable for (2) if for every time there exist a smaller time and a control such that .
By the method of characteristics, the solution of (2) can be written as
(3) |
where denotes the flow on of the non-autonomous Hamilton equation, which in local coordinates reads as
(4) |
The study of the reachable densities for (2) can then be naturally lifted to the study of the reachable Hamiltonian diffeomorphisms for (4). In fact, as one expects, controllability in the group of Hamiltonian diffeomorphisms implies controllability of the associated Liouville equation.
1.2 Approximate orbits of Hamiltonian diffeomorphisms on densities
Due to (3), smooth initial densities are preserved along the flow of (2), hence exact controllability in -spaces is clearly impossible. We shall thus focus on the -approximate controllability problem.
Definition 2 (Small-time approximately reachable densities).
A density is said to be small-time approximately reachable for (2) if for any there exist a time and a control such that . In this case, we write .
According to (3), the solutions of (2) evolve in the set
(5) |
that is, the orbit of the group action of on passing through . As a consequence, the solution of system (2) can approximately reach at most any density belonging to the -closure of : that is,
Our first result is a characterization of the orbits’ closure, when the underlying manifold is the torus or the Euclidean space of arbitrary dimension.
Theorem 3 (-closure of the orbits).
Let be or and . Then, for any ,
(6) |
where the closure is taken in the -topology.
Theorem 3 can be seen as the approximate version for Hamiltonian diffeomorphisms of a theorem by J. Moser on the group of diffeomorphism of a connected boundaryless compact manifold [17]. More precisely, for any one can define the orbit of a group action of on passing through as follows
Then, Moser’s theorem states that
An exact counterpart for Hamiltonian diffeomorphisms of Moser’s theorem should also take into account the topology of the sub- and super-level sets of , which plays no role in the approximate version stated in Theorem 3. This could be the subject of future research.
We now define the notion of approximate controllability for (2).
Definition 4 (Small-time approximate controllability).
Equation (2) is said to be small-time approximately controllable in if, for every ,
In some situations it is easier to look at the approximate controllability problem lifted on the group . Since the simplectic 2-form on is the differential of the Liouville 1-form , it is exact. As a consequence, the flux group of is trivial [21, Proposition 3.47], implying that is -closed [18]. This motivates the following notion of approximate reachability and controllability at the level of the group .
Definition 5 (Small-time approximately reachable diffeomorphisms and controllability).
This is a non-linear control problem, with infinite-dimensional state space .
We remark that the previous notion is equivalent to the notion of approximate controllability in the compact-open topology (whose definition is recalled in Section 3). We also notice that the definition can be extended to the case where the vector fields are not necessarily complete, up to considering flows on compacts.
As previously announced, approximate controllability at the level of the group implies approximate controllability at the level of the densities.
Lemma 6.
If is such that , then for every .
1.3 Examples of approximate controllability in and
The scope of the second part of the paper is to present two examples of systems for which the previously introduced controllability properties hold. The first system that we consider is posed in an Euclidean space of arbitrary dimension, the second one in a torus of arbitrary dimension (both equipped with the standard simplectic form).
1.3.1 A system in
We consider the mechanical Hamiltonian
(7) |
with such that is globally Lipschitz continuous (this last hypothesis is used to guarantee that each Hamiltonian vector field is complete). Here and in the following is used to denote the space of functions on that depend only on the state variable . Similarly, we will consider later and . Since the vector fields , are also globally Lipschitz continuous, system (4), (7) (and hence also (2), (7)) is globally-in-time well posed. Here and in similar situations in the rest of the paper, with a slight abuse of notation, stays for where and similarly for .
For the system associated with the Hamiltonian introduced in (7) we prove the following result.
1.3.2 A system in
We consider the mechanical Hamiltonian
(8) |
where and
(9) |
For such system we prove the following result.
To the best of our knowledge, it is an open problem whether the analogue of Theorem 7 for the Hamilton equations (4), (8) holds. However, it is at least possible to control finite ensembles of points of arbitrary cardinality in . Let us stress that such controllability property is not only approximate but also exact, as shown in Section 8.
1.4 Proof strategy
Concerning Theorem 3, the inclusion readily follows from the fact that Hamiltonian diffeomorphisms are measure-preserving, in the sense that (where denotes the Jacobian matrix of a Hamiltonian diffeomorphism ). We show the converse inclusion by explicitly constructing an approximating Hamiltonian diffeomorphism. This is done in two steps: first, we approximate a rearrangement of any in the right-hand side of (6) with a permutation acting on a finite grid of cubes (see Lemma 13); then, we approximate the permutation of cubes with localized Hamiltonian diffeomorphisms (see Lemma 14). The first step of our construction reminds of a result by P.D. Lax on the approximation of measure-preserving diffeomorphisms by permutations [16] (for similar constructions, see also the work by Y. Brenier and W. Gangbo [11]). Note, however, that our result is conceptually reciprocal, meaning that we approximate permutations with Hamiltonian (hence, in particular, measure-preserving) diffeomorphisms.
Our study of the approximately attainable Hamiltonian diffeomorphisms, and densities, is based on the identification of some finite families of mechanical Hamiltonians generating the group of Hamiltonian maps. It has been recently proved by Berger and Turaev [10] that any Hamiltonian diffeomorphism, on the torus or Euclidean space’s cotangent bundles, is generated by two abelian subgroups of , namely horizontal and vertical shears. We thus focus on the approximation of the elements of such subgroups.
However, a major constraint is present: due to the presence of the uncontrolled drift term (the kinetic and potential energy ) in (1), not all Poisson brackets are (a priori) attainable Hamiltonians, but only those corresponding to forward flows along the drift. We can circumvent this difficulty on , roughly thanks to the existence in this setting of a Hamiltonian function quadratic in (namely, the harmonic oscillator Hamiltonian ) whose flow is periodic in the group , since its period does not depend on the initial configuration ). This property allows to approximate also backward propagation along the drift kinetic energy, and hence all the needed Poisson brackets (for the details, see Proposition 32).
On , the harmonic oscillator Hamiltonian is not globally defined and hence we need to work locally. In particular, in this setting we are able to show that the orbits of the group action of on -densities are approximately attainable. This is a weaker result than the one showed on . An important point here is that, at the level of densities, we can approximate the action of some non-Hamiltonian diffeomorphisms, which are not allowed at the level of the group (see Lemma 35). Such additional non-Hamiltonian movement allows to approximate also backwards propagation along the drift kinetic energy, but only at the level of the densities (hence not at the level of the group ).
1.5 Some additional literature
Attainable diffeomorphisms by dynamical systems is the subject of a rich literature in various domains. The problem of attainability for volume-preserving diffeomorphisms appears e.g. in fluidodynamics [14, 20], as well as in shape optimization [8], where they furnish the natural configuration space. More in general, controlling diffeomorphisms is also useful in understanding which output can be approximated in deep learning architectures described by neural ODEs [3, 15, 19], since diffeomorphisms act transitivily on any finite set of training data.
This paper fits in a series of previous work on the controllability of diffeomorphisms group, initiated in [1] and [7] on local exact controllability, and continued in the recent works [6, 3] on global approximate controllability, which correspond to the properties studied in Theorems 7 and 9.
This work also represents the classical counterpart of the quantum study recently developed in [9], concerning attainable diffeomorphisms and Schrödinger equations. The results are similar, but here we are able to study both controllability at the level of the group and at the level of the densities in , whereas in the quantum study only the controllability at the level of the wave functions in was performed.
The control problem for Liouville transport equations was previously considered also in [12], but in a different setting (non-Hamiltonian, and with space-dependent controls).
To the best of our knowledge, this is the first work establishing the approximate controllability of the group of Hamiltonian diffeomorphisms and Liouville equations for mechanical systems (that is, for systems of the form (4),(1) or (2),(1)). Furthermore, the results are proved to hold in arbitrarily small times (notice that, due to the presence of a drift, small-time approximate controllability is not equivalent to approximate controllability).
Section 2 is dedicated to the proof of Theorem 3. In Section 3 we recall some tools on vector fields. In Section 4 we collect some properties of the attainable sets. In Section 5 we introduce two abelian subgroups that generate for . Sections 6 and 7 are dedicated, respectively, to the proofs of Theorems 7 and 9. In the last Section 8 we show the small-time exact controllability of finite ensembles of points for systems (7) and (8).
2 Proof of Theorem 3
In this section denotes or . Given , we shall write if there exists such that (it is an equivalence relation). Let
and we have to prove that
(10) |
2.1 Proof of the inclusion
Given , there exists a sequence in such that in as . Let be in . Up to extracting a sub-sequence, almost everywhere. In particular, for a.e. , as . So there exists a zero-measure set such that
Then,
Conversely, since a Hamiltonian change of variables preserves the -norm, we also have that in as . Hence, the previous reasoning also shows that . This implies that .
Remark 10.
The proof of the inclusion given in this section works for any choice of the manifold , and not only for and .
2.2 Proof of the inclusion
2.2.1 Approximation with a permutation of the mesh
Lemma 11.
Let be such that . For every in with , we have
Proof.
It is sufficient to prove that for every . The densities are in , so if , there exists such that , for , and then
∎
In the following we introduce a regular mesh of , i.e., we cover by a set of distinct cubes of volume . In order to work with cubes we use the norm
where the notation stands for . For and , the open and the closed cubes of center and radius are given by
We denote by a regular grid of points in such that
while the open cubes , , are pairwise disjoint. We call such a a mesh of size .
Definition 12.
Given and a mesh of size , a map is said to be a permutation of if is bijective and translates every open cube of the mesh to another one, that is, for every there exists such that, for every , .
Lemma 13.
Let be such that . For every , there exists such that for every and every mesh of size , there exists a permutation of such that
Proof.
First step: Let . For , by dominated convergence we have
Moreover we still have
so in the following we can consider without loss of generality that there exist and a set of finite measure, such that
for and
for .
Second step: Let us approximate the densities and by step functions.
Given we set and for every . For , let
and thanks to
the previous step, we deduce that .
Notice that for large enough, by the previous step, either (and ) or on .
Third step: Let us prove now that the sub-level sets of and can be approximately covered by the same number of cubes.
Consider , in ,
set
and assume that .
Let be open and such that , , where is arbitrary.
Set, moreover, for ,
where . By dominated convergence, in as goes to zero. In particular, , for small enough, and
where we used to denote the symmetric difference. In particular, . Given and a mesh of size , consider the minimal set such that . Notice that if then . Hence, for small enough,
from which we deduce that
(11) |
Now, denote by the set of indices such that , so that
(12) |
According to (11) and (12), we then have
Hence . In particular, since and
denoting by the cardinal of ,
it follows that .
Let us prove that we can approximate and by the same number of cubes.
Pick , such that . Then, for ,
Fourth step: We fix , such that for , and we apply the previous step to the each of the sub-level sets
with . We recall that by construction, for , so we are indeed allowed to apply the third step.
Then, given to be fixed later, there exists such that for every , every mesh of size , and every with , there exist such that and
Moreover, the cubes indexed by intersect for more than half their volume, so, for fixed, the sets are disjoint. We can construct a permutation of the mesh that translates the cubes of to the cubes of . Then, since preserves the volume,
Therefore,
so that
In conclusion, taking ,
∎
2.2.2 Approximation by a Hamiltonian diffeomorphism
The following lemma guarantees that the previously introduced permutation of the mesh can be approximated arbitrarily well by a Hamiltonian diffeomorphism.
Lemma 14.
Given and , there exists such that for every , every mesh of size , and every permutation of , there exists such that
Proof.
First step: Since the set of smooth functions with compact support is dense in , there exists a smooth function with compact support such that .
So in the following we assume that and we denote its compact support by .
Second step:
By compactness, for every and every mesh , can be covered by a finite number of cubes of , that is, there exists (depending on and ) such that . Since is continuous over the compact , then it is uniformly continuous. So
for every small enough there exists such that for every mesh of size ,
Third step: Given , a mesh of size , and a permutation of , we look for a Hamiltonian diffeomorphism such that approximates in for every . The image of by is denoted by , so that , see Figure 1.
![[Uncaptioned image]](x1.png)
As a preliminary step, let us construct a Hamiltonian diffeomorphism translating each cube , to a cube of center , where for every in , see Figure 2. The cubes of the mesh that cover are organized in columns: we choose of minimal cardinality such that and we set for . Given in , we consider a function with compact support and such that if and is at distance at most from . Then
So, up to a suitable choice of the vectors , is the required Hamiltonian diffeomorphism translating each cube , , to a cube of center , where for every in .
![[Uncaptioned image]](x2.png)
Then we want to translate every cube , , to the cube by a localized horizontal translation, see Figure 3. We consider a function with compact support and such that for every such that and is at distance at most from the segment connecting and .
Then
![[Uncaptioned image]](x3.png)
With the two previous transformations we obtained a new grid with the good -components. We look for a Hamiltonian diffeomorphism that re-shuffles the cubes whose center have the same -component without altering the others. We have already seen how to approximate localized translations in the -coordinate, see Figure 4. In the case where the desired re-shuffle can be obtained by iterating sufficiently many such transitions. In the case it is sufficient to prove that there exists a Hamiltonian diffeomorphism that permutes two consecutive cubes in the same column without altering the other cubes, see Figure 5.
![[Uncaptioned image]](x4.png)
Let us consider two consecutive cubes of the same column, with centers and , where .
Consecutive means that the segment between and does not contain any other element of the grid. Moreover, we can assume that the other cubes with -component equal to are as much separated as required.
We are going to permute the cubes with a rotation of center , where , and of angle . We consider a function such that
in
with to be fixed.
The Hamiltonian system associated with has equations
(13) |
in . As long as they stay in its solutions have the expression , . Hence, for small enough, each solution of (13) with initial condition in stays in forever. Moreover, we can assume that the support of is compact and does not intersect the cubes for . Then permutes the cubes and and is the identity on for .
![[Uncaptioned image]](x5.png)
Fourth step: For every and every permutation of a mesh , we have constructed a Hamiltonian diffeomorphism such that for every , so that
Therefore, according to the second step, for smal enough,
∎
3 Some properties of vector fields
We collect here some facts from the theory of ODEs that we will extensively use later in the article. The next two propositions (see, e.g., [5, Theorem 8.7 and Lemma 8.10]) allow to deduce the convergence of flows from that of vector fields. In what follows, denotes the space of vector fields on the manifold , which can be endowed with the compact-open topology, that is, the topology identified by the family of semi-norms , where and is compact in . The group is also endowed with the compact-open topology.
Proposition 15.
Let , , and be complete and such that for the compact-open topology of . Then, for any , for the compact-open topology of .
The following property is the analog for vector fields of the Lie product formula.
Proposition 16.
Let be complete and such that is complete. Then,
for the compact-open topology of .
In order to approximate some in the compact-open topology, we can apply a diagonal argument (based on the exhaustion of by compact sets) and reduce the problem to that of approximating in the -topology on a given compact for a given . This classical fact is recalled in the following lemma.
Lemma 17.
Let and . Assume that for every compact set and every there exists a sequence in such that
Then there exists a sequence in such that
Given its associated adjoint operator is denoted by
Given , we denote its pushforward action on vector fields as
Recall that if is a syplectomorphism and , then
(14) |
Proposition 18.
Let and be compact. There exists a compact neighborhood of and a constant such that, for every ,
4 Some properties of the approximately reachable sets
4.1 Some properties of
We show in the next proposition that the small-time approximately reachable set is a closed semi-group.
Proposition 19.
The following holds:
-
(i)
If , then .
-
(ii)
If is such that , then .
Proof.
Proof of (i): Let and be compact. Given there exist and such that
Consider such that
Then and
There exists a compact such that independently of . Moreover, we can assume that there exists depending on , , and (and independent of ) such that . By applying the mean value theorem, we obtain that
Proof of (ii): Let , , and be compact. Let be such that . Since , there exist and such that . Hence,
∎
Definition 20.
-
•
A smooth function is said to be STAR (small-time approximately reachable) if for all .
-
•
A smooth function is said to be STAR at the level of densities if for every and every .
Remark 21.
If a function is STAR then according to Lemma 6 it is also STAR at the level of densities.
Proposition 22.
The set of STAR functions is a Lie subalgebra of .
Proof.
The fact that if is STAR and then is STAR is obvious, and the fact that is STAR if and are STAR is a direct consequence of Propositions 16 and 19.
Let us assume that are STAR and prove that is STAR. According to Proposition 19,
According to Proposition 18,
where, for every and compact, . Since is STAR, applying Proposition 16 we have
Finally, as in the compact-open topology by Proposition 15 and we conclude thanks to point (ii) of Proposition 19. ∎
Proposition 23.
For a mechanical Hamiltonian of the form (1), are STAR.
Proof.
Let and . Consider the constant control such that for and for . Then is reachable in time and, according to Proposition 15,
∎
4.2 Some properties of
Lemma 24.
Let be such that , where denotes the Jacobian matrix of . If for every , then for every .
Proof.
Let and . There exist , , and such that and Then
∎
Lemma 25.
-
•
Let be such that . If and for every , then for every .
-
•
If and in , then .
The proof follows the same arguments of Proposition 19 and is omitted.
We conclude this section by proving that, as already announced, the small-time approximate controllability in of the Hamilton equation (4) implies the small-time approximate controllability in of the Liouville equation (2) (cf. Lemma 6).
Proof of Lemma 6.
According to Lemma 24, it is sufficient to prove the result for . Denote the support of by . For and denote the sphere of center and radius by . The distance between and the image of the previous sphere by is strictly positive, i.e. , . By compactness, the minimum of the previous distance as varies in has to be positive:
Moreover as . Notice that if and satisfy then
Fix a compact neighborhood of . Since , for every there exist and such that
Then, in particular, for all , which implies that is contained in for large enough.
∎
Remark 26.
In the proof of Lemma 6 we actually proved that, if a sequence converges to in for the compact-open topology and , , then in .
In analogy to Proposition 22, we have the following property.
Proposition 27.
The set of STAR functions at the level of the densities is a Lie subalgebra of .
5 Vertical and horizontal shears
We introduce two abelian subgroups of which shall play a key role.
Definition 28 (Vertical and horizontal shears on or ).
-
•
For or , the Hamiltonian diffeomorphism
is called a vertical shear. Vertical shears form an abelian subgroup of , denoted by .
-
•
For , the Hamiltonian diffeomorphism
is called a horizontal shear. Horizontal shears form an abelian subgroup of , denoted by .
Berger and Turaev recently proved the following useful density property.
Theorem 29.
([10, Corollary 1.1]) Let or and . Then, for every , , and compact there exist and such that
As a consequence of Proposition 19 and Theorem 29, we get the following sufficient condition for small-time approximate reachability of all Hamiltonian diffeomorphisms.
Corollary 30.
If and then .
6 Proof of Theorem 7
6.1 Vertical shears on
Proof.
First step: Let us prove that the Hamiltonians are STAR. The functions are STAR according to Proposition 23. Applying (14), we have
and . So , because the constant function has null contribution to the Hamiltonian vector field. By taking the limit as we get that is small-time approximately reachable for any .
Second step: Let us show that the functions are STAR if is STAR. Applying (14), we have that
for every , , and
the diffeomorphism
is small-time approximately reachable. By taking the limit as we obtain the desired property.
Third step: Let us show that every is STAR.
Notice that it is enough to consider with compact support, since the restriction of the flow on a given compact
coincides with , where coincides with on a compact set and has compact support.
In particular can be taken in for every . The set of linear combinations of Hermite functions is dense in for any , and hence approximates in by Sobolev embeddings.
By Propositions 15 and 19,
we are thus left to prove that any linear combinations of Hermite functions is STAR. We define by induction an increasing sequence of sets in by
and, for every ,
Thanks to the second step, Proposition 23, and the fact that linear combinations of STAR functions is STAR (cf. Proposition 22), any is STAR. Recall that the Hermite functions of one variable satisfy the recurrence relations
Since each Hermite function in can be written as with , we conclude that contains all Hermite functions. ∎
6.2 Quadratic Hamiltonians on
This section contains some preliminary results about an auxiliary quadratic Hamiltonian of the form
(15) |
The following proposition states that for the control system associated with such a Hamiltonian any backward propagation along the drift is small-time approximately reachable. Such a property already appeared in [2]. We recall here its proof for completeness.
Proof.
First recall that the function is STAR according to Proposition 23. Since the diffeomorphism is reachable in time with control , it follows that, for every ,
is approximately reachable in time . Applying (14), we have that
The function being STAR, Proposition 16 implies that is approximately reachable in time . Taking the limit as , we get that the function is STAR.
As a consequence, for every , the dilation
is small-time approximately reachable for every . For and , is approximately reachable in time . Thus, the element
is approximately reachable in time . We have
and finally
So for every ,
Taking arbitrarily small, we deduce that is small-time approximately reachable for every constant . Applying Proposition 16, the diffeomorphism is small-time approximately reachable for every . Notice that the latter element is periodic, hence for every , there exists such that . Thus the Hamiltonian is STAR. Since is also STAR, we deduce from Proposition 22 that is STAR. ∎
6.3 Horizontal shears on
7 Proof of Theorem 9
7.1 Vertical shears on
As in the case of , we now prove that vertical shears are small-time approximately reachable for system (4), (8).
Proof.
We first claim that, if is STAR, then for every . Applying (14) we have that the diffeomorphism
is small-time approximately reachable in time . By letting , the claim is proved.
We define an increasing sequence of vector spaces by setting
and, by induction, letting , , be the largest vector space whose elements can be written as
Let . Thanks to the claim and Propositions 22, 23, any is STAR. Moreover, the proof of [13, Proposition 2.6] shows that contains all trigonometric polynomials. In particular, is dense in , and the conclusion follows from Proposition 19. ∎
7.2 A non-Hamiltonian symmetry on densities
To the best of our knowledge, it is an open problem whether horizontal shears on can be approximately reached by system (4), (8) or not. We thus turn our attention to the weaker property of approximately controlling the Liouville equation (2), (8). At the level of densities, the system is less rigid and we can approximately reach the following non-Hamiltonian diffeomorphism.
Lemma 35.
Let be the symmetry defined by
Then for every .
Before going through the proof of Lemma 35, let us explain why it is useful. According to Lemma 35 and reasoning as in Lemma 25, for , the density is approximately reachable in time from . Thanks to the relation
(16) |
we thus get that is approximately reachable in time from . Hence, at the level of the densities, system (2), (8) can be approximately made behave as the time-reversion of the drift. It is not clear whether this can be done also at the level of Hamiltonian diffeomorphisms on .
Proof of Lemma 35.
According to Lemma 24, it is sufficient to prove the result for . As in the proof of Theorem 3, we introduce a cubic mesh of of size . Recall that, by definition, and has zero-measure complement in . Let be a compact set containing the support of . Given , since is uniformly continuous over , for small enough there exist and such that
For every volume-preserving change of variables (including ), we have
So it is sufficient to find that is small-time approximately reachable and such that for every . The image by of a cube of center is a cube of center . We will emulate the action of , which is the symmetry with respect to the space , by a rotation of center and angle on each plane . The two transformations differ pointwise, but their images of a cube of center coincide, see Figure 6.
![[Uncaptioned image]](x6.png)
As in the proof of Theorem 3, we organize the cubes of radius into columns , , such that the centers of all cubes in have the same -component . We consider , (to be fixed later) and such that
In particular, if belongs to a cube in with , then .
Integrating the Hamiltonian vector field on , it turns out that for every such that and every , , with . Choosing small enough, the ellipsoid
contains all cubes in , see Figure 7. Set .
![[Uncaptioned image]](x7.png)
After time , every cube in is sent by to its image through , while preserves all cubes in for . Moreover, is approximately reachable in time , which is arbitrarily small if is.
The proof is concluded by considering as Hamiltonian diffeomorphism emulating the action of on the cubes , , the composition of all for . ∎
7.3 Horizontal shears on at the level of densities
Proof.
First step. Let us show that, if is STAR at the level of densities, then the same is true for . According to Lemma 35 and relation (16), for every the density
is approximately reachable in time from . Moreover, by Proposition 18, given compact and ,
Since is STAR,
we deduce that is approximately reachable in time from . By letting , we obtain the desired property.
Second step.
Consider and
and let us
show that
the function is STAR at the level of densities.
According to Theorem 34, the functions and are STAR. Thus, by the previous step, for every
the following functions are STAR at the level of densities:
In particular, writing or depending on the parity of , and are STAR at the level of densities. Thus, by Proposition 27, the Hamiltonian functions
are STAR at the level of the densities as well. By taking their sum, is STAR at the level of densities.
Third step. Let us show now that every monomial is STAR at the level of densities. We show that is STAR at the level of densities and the method can be easily generalized to an arbitrary number of variables.
By Proposition 27,
the function
is STAR at the level of densities, and similarly one gets that the same is true for . Taking a linear combination of the two functions, is STAR at the level of densities. Similarly, is also STAR at the level of densities. Then,
is STAR at the level of densities and similarly one proves the same for . Finally, is STAR at the level of densities, concluding the proof of the third step.
The conclusion follows from the density of the polynomials in . ∎
8 Small-time exact controllability of finite ensembles of points in and
In this section, we detail that the small-time controllability of finite ensembles of points for systems (7) and (8) can be proved not only approximately, but also in the exact sense. This is a consequence of the fact that a finite-dimensional control systems that is approximately controllable and Lie bracket generating, is also controllable (see, e.g., [5, Corollary 8.3]).
In what follows, denotes either or and . Given any , let be the set of -uples with (at least) two coinciding components for some . The space has a structure of a smooth manifold. For each the tangent space is isomorphic to First we consider a lift of the controlled systems with controlled Hamiltonian (7) and (8) defined respectively on and . These systems are of the form (4). The lift on is then defined by
(17) |
where , the Hamiltonian is defined in (1), and We then define the following family of Hamiltonians on
As a direct consequence of Theorems 31 and 34, we obtain the following Lie extension property.
Proposition 37.
Let us now prove the small-time approximate controllability of system (17), which is an intermediate step towards the small-time exact controllability proved later in the section.
Proposition 38.
System (17) is small-time approximately controllable in , i.e. , for any distinct initial configurations , and distinct final configurations , and , then there exist and such that
for every .
The proof of Proposition 38 is based on the following technical result, which can be deduced for instance from Whitney extension theorem.
Lemma 39.
For distinct positions and vectors , there exists a smooth function such that for every .
Proof of Proposition 38.
For the case of , the result directly follows from the small-time approximated controllability in the group .
Let us consider the case . The pairs of initial configurations are distinct so, for every , there exists such that the points are pairwise distinct on the torus. For every , the diffeomorphism is approximately reachable in time and for . So we can assume without loss of generality that the initial positions are distinct. Similarly, we can assume without loss of generality that the final positions are distinct, up to replace them by with arbitrarily small.
For each , let be such that modulo . Applying Lemma 39, there exists a function such that for every . As a consequence, and then . Applying again Lemma 39, there exists a smooth function such that for every . Then . Finally, thanks to Theorem 34, the diffeomorphism is approximately reachable in time for every . ∎
Theorem 40.
System (17) is small-time exactly controllable in .
Proof.
It is a well-known consequence of Krener’s theorem (see, e.g., [5, Corollary 8.3]111the result is not written in the small-time case, but the proof readily extends to such a case) that if a finite-dimensional control system is small-time approximately controllable and satisfies the Lie algebra rank condition, then it is small-time controllable. Hence, according to Proposition 38, we are left to check that (17) satisfies the Lie algebra rank condition.
Let . Assume for now that the positions are distinct in . Let be a basis of . Similarly to what is done for Lemma 39, by the Whitney extension theorem there exist such that
(19) |
for and . Thanks to (19), we have . Moreover, using again (19), for we compute
where is the lifted drift. Replacing by in the above argument, we can generate vectors that form a basis of .
Since the Lie algebra is dense in for the compact-open topology (cf. Theorems 31 and 34), it follows by continuity that system (17) satisfies the Lie algebra rank condition at .
We are left to consider the case where are not pairwise distinct. Since , the pairs are distinct in . Then there exists such that the positions are pairwise distinct in . Then is approximately reachable from and (17) satisfies the Lie algebra rank condition in a neighborhood of . Since each Hamiltonian , , is analytic, it follows that the Lie algebra rank condition is satisfied at every point of the orbit through for system (17), and in particular at . ∎
Acknowledgments. The authors wish to thank Ivan Beschastnyi for inspiring conversations at the origin of this project, and Andrei Agrachev, Sylvain Arguillère, Pierre Berger, Borjan Geshkovski, Emmanuel Trélat, and Claude Viterbo for enlightening discussions.
E.P. thanks the SMAI for supporting and the CIRM for hosting the BOUM project ”Small-time controllability of Liouville transport equations along an Hamil- tonian field”, where some ideas of this work were conceived.
This work has been partly supported by the ANR-DFG project CoRoMo ANR-22-CE92-0077-01 and the ANR project QuBiCCS ANR-24-CE40-3008-01. This project has received financial support from the CNRS through the MITI interdisciplinary programs.
References
- [1] A. Agrachev and M. Caponigro, Controllability on the group of diffeomorphisms, Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 26 (2009), pp. 2503–2509.
- [2] A. Agrachev, B. Kazandjian, and E. Pozzoli, Good Lie brackets for classical and quantum harmonic oscillators, Systems & Control Letters, (2025).
- [3] A. Agrachev and A. Sarychev, Control on the manifolds of mappings with a view to the deep learning, Journal of Dynamical and Control Systems, 28 (2022), pp. 989–1008.
- [4] A. A. Agrachev and R. V. Gamkrelidze, The exponential representation of flows and the chronological calculus, Math. USSR, Sb., 35 (1979), pp. 727–785.
- [5] A. A. Agrachev and Y. L. Sachkov, Control theory from the geometric viewpoint, vol. 87 of Encyclopaedia of Mathematical Sciences, Springer-Verlag, Berlin, 2004. Control Theory and Optimization, II.
- [6] A. A. Agrachev and A. V. Sarychev, Control in the spaces of ensembles of points, SIAM J. Control. Optim., 58 (2019), pp. 1579–1596.
- [7] S. Arguillère and E. Trélat, Sub-Riemannian structures on groups of diffeomorphisms, J. Inst. Math. Jussieu, 16 (2017), pp. 745–785.
- [8] S. Arguillère, E. Trélat, A. Trouvé, and L. Younes, Shape deformation analysis from the optimal control viewpoint, J. Math. Pures Appl. (9), 104 (2015), pp. 139–178.
- [9] K. Beauchard and E. Pozzoli, Small-time approximate controllability of bilinear Schrödinger equations and diffeomorphisms, Annales de l’Institut Henri Poincaré Analyse Non-Linéaire, (2025).
- [10] P. Berger and D. Turaev, Generators of groups of Hamiltonian maps, Israel J. Math., 267 (2025), pp. 237–252.
- [11] Y. Brenier and W. Gangbo, approximation of maps by diffeomorphisms, Calc. Var. Partial Differ. Equ., 16 (2003), pp. 147–164.
- [12] R. W. Brockett, Optimal control of the Liouville equation, in Proceedings of the international conference on complex geometry and related fields, Shanghai, China, 2004, Providence, RI: American Mathematical Society (AMS); Somerville, MA: International Press, 2007, pp. 23–35.
- [13] A. Duca and V. Nersesyan, Bilinear control and growth of Sobolev norms for the nonlinear Schrödinger equation, J. Eur. Math. Soc. (JEMS), 27 (2025), pp. 2603–2622.
- [14] D. G. Ebin and J. Marsden, Groups of diffeomorphisms and the motion of an incompressible fluid, Ann. Math. (2), 92 (1970), pp. 102–163.
- [15] K. Elamvazhuthi, B. Gharesifard, A. L. Bertozzi, and S. Osher, Neural ODE control for trajectory approximation of continuity equation, IEEE Control Syst. Lett., 6 (2022), pp. 3152–3157, https://doi.org/10.1109/lcsys.2022.3182284, https://doi.org/10.1109/lcsys.2022.3182284.
- [16] P. D. Lax, Approximation of measure preserving transformations, Commun. Pure Appl. Math., 24 (1971), pp. 133–135.
- [17] J. Moser, On the volume elements on a manifold, Trans. Am. Math. Soc., 120 (1965), pp. 286–294.
- [18] K. Ono, Floer-Novikov cohomology and the flux conjecture, Geom. Funct. Anal., 16 (2006), pp. 981–1020.
- [19] D. Ruiz-Balet and E. Zuazua, Control of neural transport for normalising flows, J. Math. Pures Appl. (9), 181 (2024), pp. 58–90.
- [20] A. I. Shnirelman, Attainable diffeomorphisms, Geom. Funct. Anal., 3 (1993), pp. 279–294.
- [21] C. Viterbo, Symplectic topology in the cotangent bundle through generating functions. Lecture notes (2021).