Orbits and attainable Hamiltonian diffeomorphisms of mechanical Liouville equations

Bettina Kazandjian, Eugenio Pozzoli, Mario Sigalotti Sorbonne Université, Université Paris Cité, CNRS, Inria, Laboratoire Jacques Louis-Lions, Paris, France (bettina.kazandjian@sorbonne-universite.fr, mario.sigalotti@inria.fr)Univ Rennes, CNRS, IRMAR - UMR 6625, F-35000 Rennes, France (eugenio.pozzoli@univ-rennes.fr)
Abstract

We study the approximate controllability problem for Liouville transport equations along a mechanical Hamiltonian vector field. Such PDEs evolve inside the orbit

𝒪(ρ0):={ρ0ΦΦDHam(TM)},ρ0Lr(TM,),r[1,),\mathcal{O}(\rho_{0}):=\left\{\rho_{0}\circ\Phi\mid\Phi\in{\rm DHam}(T^{*}M)\right\},\quad\rho_{0}\in L^{r}(T^{*}M,{\mathbb{R}}),\quad r\in[1,\infty),

where ρ0\rho_{0} is the initial density and DHam(TM){\rm DHam}(T^{*}M) is the group of Hamiltonian diffeomorphisms of the cotangent bundle manifold TMT^{*}M. The approximately reachable densities from ρ0\rho_{0} are thus contained in 𝒪(ρ0)¯\overline{\mathcal{O}(\rho_{0})}, where the closure is taken with respect to the LrL^{r}-topology. Our first result is a characterization of 𝒪(ρ0)¯\overline{\mathcal{O}(\rho_{0})} when the manifold MM is the Euclidean space d{\mathbb{R}}^{d} or the torus 𝕋d{\mathbb{T}}^{d} of arbitrary dimension: 𝒪(ρ0)¯\overline{\mathcal{O}(\rho_{0})} is the set of all the densities whose sub- and super-level sets have the same measure as those of ρ0\rho_{0}. This result is an approximate version, in the case of DHam(TM){\rm DHam}(T^{*}M), 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 M=dM={\mathbb{R}}^{d} and 𝕋d{\mathbb{T}}^{d}, where the small-time approximately attainable diffeomorphisms coincide with DHam(TM){\rm DHam}(T^{*}M), 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

1 Introduction

1.1 The model

Let MM be a smooth dd-dimensional boundaryless Riemannian manifold. Its cotangent bundle TMT^{*}M is equipped with canonical symplectic and volume forms, that in local position and momentum coordinates (q,p)=(q1,,qd,p1,,pd)(q,p)=(q_{1},\dots,q_{d},p_{1},\dots,p_{d}) can be expressed, respectively, as ω=i=1ddpidqi\omega=\sum_{i=1}^{d}dp_{i}\wedge dq_{i} and ωω\omega\wedge\dots\wedge\omega (dd-times). For any measurable set ATMA\subset T^{*}M we denote by Vol(A)0\operatorname{Vol}(A)\geq 0 its volume.

In this article we study the controllability of Liouville equations describing the transportation of a density along a Hamiltonian vector field. Let 𝒞(TM,){\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}), Ham(TM){\rm Ham}(T^{*}M), DHam(TM),{\rm DHam}(T^{*}M), be, respectively, the space of smooth (𝒞{\mathcal{C}^{\infty}}) functions, Hamiltonian vector fields, and Hamiltonian diffeomorphisms. The space of real-valued functions 𝒞(TM,){\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}) is a Lie algebra equipped with the Poisson bracket {f,g}=j=1dpjfqjgqjgpjf.\left\{f,g\right\}=\sum_{j=1}^{d}\partial_{p_{j}}f\partial_{q_{j}}g-\partial_{q_{j}}g\partial_{p_{j}}f. With any smooth function ff one can associate the corresponding Hamiltonian vector field f={f,}\overrightarrow{f}=\left\{f,\cdot\right\}. In particular, in this work we shall restrict ourselves to mechanical Hamiltonians, i.e. , Hamiltonians of the form

Hu(t)(q,p)=|p|22+V0(q)+j=1muj(t)Vj(q),(q,p)TM,H_{u(t)}(q,p)=\frac{|p|^{2}}{2}+V_{0}(q)+\sum_{j=1}^{m}u_{j}(t)V_{j}(q),\quad(q,p)\in T^{*}M, (1)

which are the sum of a drift term, denoted as H0(q,p)=|p|22+V0(q)H_{0}(q,p)=\frac{|p|^{2}}{2}+V_{0}(q) (the sum of kinetic and potential energy), and a potential j=1muj(t)Vj(q)\sum_{j=1}^{m}u_{j}(t)V_{j}(q) controlled through u()=(u1(),,um())PWC(+,m)u(\cdot)=(u_{1}(\cdot),\dots,u_{m}(\cdot))\in\operatorname{PWC}({\mathbb{R}}_{+},{\mathbb{R}}^{m}), were PWC\operatorname{PWC} is used to denote piecewise constant functions. We suppose in the following that for each umu\in{\mathbb{R}}^{m} the vector field Hu={Hu,}\overrightarrow{H}_{u}=\left\{H_{u},\cdot\right\} is complete.

Let us fix r[1,)r\in[1,\infty). Given u()PWC(+,m)u(\cdot)\in\operatorname{PWC}({\mathbb{R}}_{+},{\mathbb{R}}^{m}), we consider the transportation of an initial density ρ0Lr(TM,)\rho_{0}\in L^{r}(T^{*}M,{\mathbb{R}}) along the piecewise constant Hamiltonian vector field tHu(t)Ham(TM)t\mapsto\overrightarrow{H}_{u(t)}\in{\rm Ham}(T^{*}M), given by the Liouville equation

{tρ(q,p,t)=Hu(t)(q,p)ρ(q,p,t),ρ(q,p,0)=ρ0(q,p).\begin{cases}\partial_{t}\rho(q,p,t)=\overrightarrow{H}_{u(t)}(q,p)\rho(q,p,t),&\\ \rho(q,p,0)=\rho_{0}(q,p).&\end{cases} (2)

Equation (2) more explicitly reads as the partial differential equation

tρ(q,p,t)=pqρ(q,p,t)qV0(q)pρ(q,p,t)j=1muj(t)qVj(q)pρ(q,p,t).\partial_{t}\rho(q,p,t)=p\cdot\nabla_{q}\rho(q,p,t)-\nabla_{q}V_{0}(q)\cdot\nabla_{p}\rho(q,p,t)-\sum_{j=1}^{m}u_{j}(t)\nabla_{q}V_{j}(q)\cdot\nabla_{p}\rho(q,p,t).

This is a bilinear (hence, non-linear) control problem, with infinite-dimensional state space Lr(TM,)L^{r}(T^{*}M,{\mathbb{R}}).

Since the vector field Hu\overrightarrow{H}_{u} is complete for every umu\in{\mathbb{R}}^{m}, the Liouville equation is globally-in-time well posed, meaning that for any ρ0Lr(TM,),u()PWC(+,m)\rho_{0}\in L^{r}(T^{*}M,{\mathbb{R}}),u(\cdot)\in\operatorname{PWC}({\mathbb{R}}_{+},{\mathbb{R}}^{m}) there exists a unique mild solution (tρ(t)=ρ(t;u(),ρ0))𝒞0(,Lr(TM,))(t\mapsto\rho(t)=\rho(t;u(\cdot),\rho_{0}))\in\mathcal{C}^{0}({\mathbb{R}},L^{r}(T^{*}M,{\mathbb{R}})) 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 ρ1Lr(TM)\rho_{1}\in L^{r}(T^{*}M) is said to be small-time reachable for (2) if for every time T>0T>0 there exist a smaller time τ[0,T]\tau\in[0,T] and a control u()PWC([0,τ],m)u(\cdot)\in\operatorname{PWC}([0,\tau],{\mathbb{R}}^{m}) such that ρ(τ;u(),ρ0)=ρ1\rho(\tau;u(\cdot),\rho_{0})=\rho_{1}.

By the method of characteristics, the solution of (2) can be written as

ρ(t)=ρ0ΦHut,\rho(t)=\rho_{0}\circ\Phi^{t}_{H_{u}}, (3)

where ΦHutDHam(TM)\Phi^{t}_{H_{u}}\in{\rm DHam}(T^{*}M) denotes the flow on TMT^{*}M of the non-autonomous Hamilton equation, which in local coordinates reads as

{q˙=pHu(t)(q,p),p˙=qHu(t)(q,p).\left\{\begin{array}[]{ll}\dot{q}=\nabla_{p}H_{u(t)}(q,p),\\ \dot{p}=-\nabla_{q}H_{u(t)}(q,p).\end{array}\right. (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 LrL^{r}-spaces is clearly impossible. We shall thus focus on the LrL^{r}-approximate controllability problem.

Definition 2 (Small-time approximately reachable densities).

A density ρ1Lr(TM)\rho_{1}\in L^{r}(T^{*}M) is said to be small-time approximately reachable for (2) if for any ε>0\varepsilon>0 there exist a time τ[0,ε]\tau\in[0,\varepsilon] and a control u()PWC([0,τ],m)u(\cdot)\in\operatorname{PWC}([0,\tau],{\mathbb{R}}^{m}) such that ρ(τ;u(),ρ0)ρ1Lr(TM)<ε\|\rho(\tau;u(\cdot),\rho_{0})-\rho_{1}\|_{L^{r}(T^{*}M)}<\varepsilon. In this case, we write ρ1Rst¯(ρ0)\rho_{1}\in\overline{R_{\rm st}}(\rho_{0}).

According to (3), the solutions of (2) evolve in the set

𝒪(ρ0)={ρ0ΦΦDHam(TM)},\mathcal{O}(\rho_{0})=\left\{\rho_{0}\circ\Phi\mid\Phi\in{\rm DHam}(T^{*}M)\right\}, (5)

that is, the orbit of the group action of DHam(TM){\rm DHam}(T^{*}M) on Lr(TM)L^{r}(T^{*}M) passing through ρ0\rho_{0}. As a consequence, the solution of system (2) can approximately reach at most any density belonging to the LrL^{r}-closure of 𝒪(ρ0)\mathcal{O}(\rho_{0}): that is,

Rst¯(ρ0)𝒪(ρ0)¯.\overline{R_{\rm st}}(\rho_{0})\subset\overline{\mathcal{O}(\rho_{0})}.

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 (LrL^{r}-closure of the orbits).

Let MM be 𝕋d=d/2πd{\mathbb{T}}^{d}={\mathbb{R}}^{d}/2\pi{\mathbb{Z}}^{d} or d{\mathbb{R}}^{d} and r[1,)r\in[1,\infty). Then, for any ρ0Lr(TM)\rho_{0}\in L^{r}(T^{*}M),

𝒪(ρ0)¯={ρ1Lr(TM)μ<ν,Vol({μ<ρ1<ν})=Vol({μ<ρ0<ν})},\overline{\mathcal{O}(\rho_{0})}=\left\{\rho_{1}\in L^{r}(T^{*}M)\mid\forall\mu<\nu,\operatorname{Vol}(\{\mu<\rho_{1}<\nu\})=\operatorname{Vol}(\{\mu<\rho_{0}<\nu\})\right\}, (6)

where the closure is taken in the LrL^{r}-topology.

Theorem 3 can be seen as the approximate version for Hamiltonian diffeomorphisms of a theorem by J. Moser on the group of diffeomorphism Diff(M){\rm Diff}(M) of a connected boundaryless compact manifold MM [17]. More precisely, for any ρ0𝒞(M,(0,))\rho_{0}\in{\mathcal{C}^{\infty}}(M,(0,\infty)) one can define the orbit of a group action of Diff(M){\rm Diff}(M) on 𝒞(M,(0,)){\mathcal{C}^{\infty}}(M,(0,\infty)) passing through ρ0\rho_{0} as follows

𝒪Diff(ρ0)={det(DP)(ρ0P)PDiff(M)}.\mathcal{O}_{\rm Diff}(\rho_{0})=\left\{\det(DP)(\rho_{0}\circ P)\mid P\in{\rm Diff}(M)\right\}.

Then, Moser’s theorem states that

𝒪Diff(ρ0)={ρ1𝒞(M,(0,))Mρ1(x)𝑑x=Mρ0(x)𝑑x}.\mathcal{O}_{\rm Diff}(\rho_{0})=\left\{\rho_{1}\in{\mathcal{C}^{\infty}}(M,(0,\infty))\mid\int_{M}\rho_{1}(x)dx=\int_{M}\rho_{0}(x)dx\right\}.

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 ρ0\rho_{0}, 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 Lr(TM)L^{r}(T^{*}M) if, for every ρ0Lr(TM)\rho_{0}\in L^{r}(T^{*}M) ,

Rst¯(ρ0)=𝒪(ρ0)¯.\overline{R_{\rm st}}(\rho_{0})=\overline{\mathcal{O}(\rho_{0})}.

In some situations it is easier to look at the approximate controllability problem lifted on the group DHam(TM){\rm DHam}(T^{*}M). Since the simplectic 2-form on TMT^{*}M is the differential of the Liouville 1-form i=1dpidqi\sum_{i=1}^{d}p_{i}dq_{i}, it is exact. As a consequence, the flux group of TMT^{*}M is trivial [21, Proposition 3.47], implying that DHam(TM){\rm DHam}(T^{*}M) is 𝒞1\mathcal{C}^{1}-closed [18]. This motivates the following notion of approximate reachability and controllability at the level of the group DHam(TM){\rm DHam(T^{*}M)}.

Definition 5 (Small-time approximately reachable diffeomorphisms and controllability).

A Hamiltonian diffeomorphism ΨDHam(TM)\Psi\in{\rm DHam}(T^{*}M) is said to be small-time approximately reachable for (4) if for every ε>0\varepsilon>0, every compact KTMK\subset T^{*}M, and every \ell\in{\mathbb{N}} there exist a time τ[0,ε]\tau\in[0,\varepsilon] and a control u()PWC([0,τ],m)u(\cdot)\in\operatorname{PWC}([0,\tau],{\mathbb{R}}^{m}) such that ΦHuτΨ𝒞(K)<ε\|\Phi^{\tau}_{H_{u}}-\Psi\|_{\mathcal{C}^{\ell}(K)}<\varepsilon. In this case, we write Ψst¯\Psi\in\overline{\mathcal{R}_{\rm st}}.
Equation (4) is said to be small-time approximately controllable in DHam(TM){\rm DHam}(T^{*}M) if

st¯=DHam(TM).\overline{\mathcal{R}_{\rm st}}={\rm DHam}(T^{*}M).

This is a non-linear control problem, with infinite-dimensional state space DHam(TM){\rm DHam}(T^{*}M).

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 Hu\overrightarrow{H_{u}} 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 ΨDHam(TM)\Psi\in\operatorname{DHam}(T^{*}M) is such that Ψst¯\Psi\in\overline{\mathcal{R}_{\rm st}}, then ρ0ΨRst¯(ρ0){\rho}_{0}\circ\Psi\in\overline{R_{\rm st}}(\rho_{0}) for every ρ0Lr(TM){\rho}_{0}\in L^{r}(T^{*}M).

1.3 Examples of approximate controllability in DHam(Td){\rm DHam}(T^{*}{\mathbb{R}}^{d}) and Lr(T𝕋d,)L^{r}(T^{*}{\mathbb{T}}^{d},{\mathbb{R}})

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 d{\mathbb{R}}^{d}

We consider the mechanical Hamiltonian

Hu(q,p)=|p|22+V0(q)+j=1dujqj+ud+1e|q|2/2,(q,p)Td=qd×pd,H_{u}(q,p)=\frac{|p|^{2}}{2}+V_{0}(q)+\sum_{j=1}^{d}u_{j}q_{j}+u_{d+1}e^{-|q|^{2}/2},\qquad(q,p)\in T^{*}{\mathbb{R}}^{d}={\mathbb{R}}^{d}_{q}\times{\mathbb{R}}^{d}_{p}, (7)

with V0𝒞(qd,)V_{0}\in{\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{q},{\mathbb{R}}) such that V0\nabla V_{0} is globally Lipschitz continuous (this last hypothesis is used to guarantee that each Hamiltonian vector field Hu\overrightarrow{H_{u}} is complete). Here and in the following 𝒞(qd,){\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{q},{\mathbb{R}}) is used to denote the space of functions on TdT^{*}{\mathbb{R}}^{d} that depend only on the state variable qq. Similarly, we will consider later 𝒞(pd,){\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{p},{\mathbb{R}}) and 𝒞(𝕋qd,){\mathcal{C}^{\infty}}({\mathbb{T}}^{d}_{q},{\mathbb{R}}). Since the vector fields q1=p1,,qd=pd\overrightarrow{q_{1}}=-\partial_{p_{1}},\dots,\overrightarrow{q_{d}}=-\partial_{p_{d}}, e|q|2/2=e|q|2/2j=1dqjpj\overrightarrow{e^{-|q|^{2}/2}}=e^{-|q|^{2}/2}\sum_{j=1}^{d}q_{j}\partial_{p_{j}} 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, qi\overrightarrow{q_{i}} stays for φ\overrightarrow{\varphi} where φ:(q,p)qi\varphi:(q,p)\mapsto q_{i} and similarly for e|q|2/2\overrightarrow{e^{-|q|^{2}/2}}.

For the system associated with the Hamiltonian introduced in (7) we prove the following result.

Theorem 7.

The Hamilton equations (4), (7) are small-time approximately controllable in DHam(Td){\rm DHam}(T^{*}{\mathbb{R}}^{d}).

As a consequence of Lemma 6 and Theorem 7, we also get the following result.

Corollary 8.

The Liouville equation (2), (7) is small-time approximately controllable in Lr(Td)L^{r}(T^{*}{\mathbb{R}}^{d}), r[1,)r\in[1,\infty).

1.3.2 A system in 𝕋d{\mathbb{T}}^{d}

We consider the mechanical Hamiltonian

Hu=|p|22+V0(q)+j=1du2j1cos(kjq)+u2jsin(kjq),(q,p)T𝕋d=𝕋qd×pd,H_{u}=\frac{|p|^{2}}{2}+V_{0}(q)+\sum_{j=1}^{d}u_{2j-1}\cos(k_{j}\cdot q)+u_{2j}\sin(k_{j}\cdot q),\quad(q,p)\in T^{*}{\mathbb{T}}^{d}={\mathbb{T}}^{d}_{q}\times{\mathbb{R}}^{d}_{p}, (8)

where V0𝒞(𝕋qd,)V_{0}\in{\mathcal{C}^{\infty}}({\mathbb{T}}^{d}_{q},{\mathbb{R}}) and

k1=(1,0,,0),,kd1=(0,,0,1,0),kd=(1,,1).k_{1}=(1,0,\dots,0),\dots,k_{d-1}=(0,\dots,0,1,0),\quad k_{d}=(1,\dots,1). (9)

For such system we prove the following result.

Theorem 9.

The Liouville equation (2), (8) is small-time approximately controllable in Lr(T𝕋d)L^{r}(T^{*}{\mathbb{T}}^{d}), r[1,)r\in[1,\infty).

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 T𝕋dT^{*}{\mathbb{T}}^{d}. 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 𝒪(ρ0)¯(ρ0)\overline{\mathcal{O}(\rho_{0})}\subset\mathcal{L}(\rho_{0}) readily follows from the fact that Hamiltonian diffeomorphisms are measure-preserving, in the sense that det(DΦ)1\det(D\Phi)\equiv 1 (where DΦD\Phi denotes the Jacobian matrix of a Hamiltonian diffeomorphism Φ\Phi). We show the converse inclusion (ρ0)𝒪(ρ0)¯\mathcal{L}(\rho_{0})\subset\overline{\mathcal{O}(\rho_{0})} by explicitly constructing an approximating Hamiltonian diffeomorphism. This is done in two steps: first, we approximate a rearrangement of any ρ1\rho_{1} 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 DHam(TM){\rm DHam}(T^{*}M), 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 |p|2/2+V0(q)|p|^{2}/2+V_{0}(q)) 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 DHam(Td){\rm DHam}(T^{*}{\mathbb{R}}^{d}), roughly thanks to the existence in this setting of a Hamiltonian function quadratic in pp (namely, the harmonic oscillator Hamiltonian (|p|2+|q|2)/2(|p|^{2}+|q|^{2})/2) whose flow is periodic in the group DHam(Td){\rm DHam}(T^{*}{\mathbb{R}}^{d}), since its period does not depend on the initial configuration (q,p)(q,p)). 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 DHam(T𝕋d){\rm DHam}(T^{*}{\mathbb{T}}^{d}), 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 DHam(T𝕋d){\rm DHam}(T^{*}{\mathbb{T}}^{d}) on Lr(T𝕋d)L^{r}(T^{*}{\mathbb{T}}^{d})-densities are approximately attainable. This is a weaker result than the one showed on DHam(Td){\rm DHam}(T^{*}{\mathbb{R}}^{d}). 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 DHam(T𝕋d){\rm DHam}(T^{*}{\mathbb{T}}^{d}) (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 DHam(T𝕋d){\rm DHam}(T^{*}{\mathbb{T}}^{d})).

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 DHam(TM){\rm DHam}(T^{*}M) and at the level of the densities in Lr(TM,)L^{r}(T^{*}M,{\mathbb{R}}), whereas in the quantum study only the controllability at the level of the wave functions in L2(M,)L^{2}(M,{\mathbb{C}}) 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 DHam(TM){\rm DHam}(T^{*}M) for M=d,𝕋dM={\mathbb{R}}^{d},{\mathbb{T}}^{d}. 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 MM denotes d{\mathbb{R}}^{d} or 𝕋d{\mathbb{T}}^{d}. Given ρ0,ρ1Lr(TM)\rho_{0},\rho_{1}\in L^{r}(T^{*}M), we shall write ρ0ρ1\rho_{0}\sim\rho_{1} if there exists ΦDHam(V)\Phi\in{\rm DHam}(V) such that ρ0=ρ1Φ\rho_{0}=\rho_{1}\circ\Phi (it is an equivalence relation). Let

(ρ0):={ρ1Lr(TM)μ<ν,Vol({μ<ρ1<ν})=Vol({μ<ρ0<ν})},\mathcal{L}(\rho_{0}):=\left\{\rho_{1}\in L^{r}(T^{*}M)\mid\forall\mu<\nu,\operatorname{Vol}(\{\mu<\rho_{1}<\nu\})=\operatorname{Vol}(\{\mu<\rho_{0}<\nu\})\right\},

and we have to prove that

𝒪(ρ0)¯=(ρ0),ρ0Lr(TM).\overline{\mathcal{O}(\rho_{0})}=\mathcal{L}(\rho_{0}),\qquad\forall\rho_{0}\in L^{r}(T^{*}M). (10)

2.1 Proof of the inclusion 𝒪(ρ0)¯(ρ0)\overline{\mathcal{O}(\rho_{0})}\subset\mathcal{L}(\rho_{0})

Given ρ1𝒪(ρ0)¯{\rho}_{1}\in\overline{\mathcal{O}({\rho}_{0})}, there exists a sequence (ϕn)n(\phi_{n})_{n\in{\mathbb{N}}} in DHam(V)\operatorname{DHam}(V) such that ρ0ϕnρ1{\rho}_{0}\circ\phi_{n}\rightarrow{\rho}_{1} in LrL^{r} as nn\to\infty. Let μ<ν\mu<\nu be in {\mathbb{R}}. Up to extracting a sub-sequence, ρ0ϕnρ1{\rho}_{0}\circ\phi_{n}\rightarrow{\rho}_{1} almost everywhere. In particular, for a.e. x{yμ<ρ1(y)<ν}x\in\{y\mid\mu<{\rho}_{1}(y)<\nu\}, ρ0ϕn(x)ρ1(x){\rho}_{0}\circ\phi_{n}(x)\rightarrow{\rho}_{1}(x) as nn\to\infty. So there exists a zero-measure set 𝒩V\mathcal{N}\subset V such that

{μ<ρ1<ν}NnN{μ<ρ0ϕn<ν}𝒩.\{\mu<{\rho}_{1}<\nu\}\subset\cup_{N\in{\mathbb{N}}}\cap_{n\geq N}\{\mu<{\rho}_{0}\circ\phi_{n}<\nu\}\cup\mathcal{N}.

Then,

Vol({μ<ρ1<ν})\displaystyle\operatorname{Vol}(\{\mu<{\rho}_{1}<\nu\}) Vol(NnN{μ<ρ0ϕn<ν})+Vol(𝒩)\displaystyle\leq\operatorname{Vol}(\cup_{N\in{\mathbb{N}}}\cap_{n\geq N}\>\{\mu<{\rho}_{0}\circ\phi_{n}<\nu\})+\operatorname{Vol}(\mathcal{N})
=limNVol(nN{μ<ρ0ϕn<ν})\displaystyle=\lim_{N\rightarrow\infty}\operatorname{Vol}(\cap_{n\geq N}\>\{\mu<{\rho}_{0}\circ\phi_{n}<\nu\})
lim supNVol({μ<ρ0ϕN<ν})\displaystyle\leq\limsup_{N\rightarrow\infty}\operatorname{Vol}(\{\mu<{\rho}_{0}\circ\phi_{N}<\nu\})
=Vol({μ<ρ0<ν}).\displaystyle=\operatorname{Vol}(\{\mu<{\rho}_{0}<\nu\}).

Conversely, since a Hamiltonian change of variables preserves the LrL^{r}-norm, we also have that ρ1ϕn1ρ0{\rho}_{1}\circ\phi_{n}^{-1}\rightarrow{\rho}_{0} in LrL^{r} as nn\to\infty. Hence, the previous reasoning also shows that Vol({μ<ρ0<ν})Vol({μ<ρ1<ν})\operatorname{Vol}(\{\mu<{\rho}_{0}<\nu\})\leq\operatorname{Vol}(\{\mu<{\rho}_{1}<\nu\}). This implies that Vol({μ<ρ0<ν})=Vol({μ<ρ1<ν})\operatorname{Vol}(\{\mu<{\rho}_{0}<\nu\})=\operatorname{Vol}(\{\mu<{\rho}_{1}<\nu\}). \square

Remark 10.

The proof of the inclusion 𝒪(ρ0)¯(ρ0)\overline{\mathcal{O}(\rho_{0})}\subset\mathcal{L}(\rho_{0}) given in this section works for any choice of the manifold MM, and not only for M=dM={\mathbb{R}}^{d} and M=𝕋dM={\mathbb{T}}^{d}.

2.2 Proof of the inclusion (ρ0)𝒪(ρ0)¯\mathcal{L}(\rho_{0})\subset\overline{\mathcal{O}(\rho_{0})}

2.2.1 Approximation with a permutation of the mesh

Lemma 11.

Let ρ0,ρ1Lr(TM){\rho}_{0},{\rho}_{1}\in L^{r}(T^{*}M) be such that ρ0ρ1{\rho}_{0}\sim{\rho}_{1}. For every μ<ν\mu<\nu in {\mathbb{R}} with μ0\mu\neq 0, we have

Vol({μρ0<ν})=Vol({μρ1<ν}).\operatorname{Vol}(\{\mu\leq{\rho}_{0}<\nu\})=\operatorname{Vol}(\{\mu\leq{\rho}_{1}<\nu\}).
Proof.

It is sufficient to prove that Vol({ρ0=μ})=Vol({ρ1=μ})\operatorname{Vol}(\{{\rho}_{0}=\mu\})=\operatorname{Vol}(\{{\rho}_{1}=\mu\}) for every μ0\mu\neq 0. The densities ρ0,ρ1\rho_{0},\rho_{1} are in LrL^{r}, so if μ0\mu\neq 0, there exists n0n_{0}\in{\mathbb{N}}^{*} such that Vol({μ1n0<ρi<μ+1n0})<+\operatorname{Vol}(\{\mu-\frac{1}{n_{0}}<{\rho}_{i}<\mu+\frac{1}{n_{0}}\})<+\infty, for i{0,1}i\in\{0,1\}, and then

Vol({ρ1=μ})\displaystyle\operatorname{Vol}(\{{\rho}_{1}=\mu\}) =Vol(n{μ1n<ρ1<μ+1n})\displaystyle=\operatorname{Vol}\left(\cap_{n\in{\mathbb{N}}^{*}}\{\mu-\frac{1}{n}<{\rho}_{1}<\mu+\frac{1}{n}\}\right)
=limnVol({μ1n<ρ1<μ+1n})\displaystyle=\lim_{n\rightarrow\infty}\operatorname{Vol}\left(\left\{\mu-\frac{1}{n}<{\rho}_{1}<\mu+\frac{1}{n}\right\}\right)
=limnVol({μ1n<ρ0<μ+1n})\displaystyle=\lim_{n\rightarrow\infty}\operatorname{Vol}\left(\left\{\mu-\frac{1}{n}<{\rho}_{0}<\mu+\frac{1}{n}\right\}\right)
=Vol({ρ0=μ}).\displaystyle=\operatorname{Vol}(\{{\rho}_{0}=\mu\}).

In the following we introduce a regular mesh of TMT^{*}M, i.e., we cover TMT^{*}M by a set of distinct cubes of volume h2d,h>0h^{2d},h>0. In order to work with cubes we use the norm

x=maxi\llbracket1,2d\rrbracket|xi|,xTM,\|x\|=\max_{i\in\llbracket 1,2d\rrbracket}|x^{i}|,\qquad x\in T^{*}M,

where the notation \llbracket1,2d\rrbracket\llbracket 1,2d\rrbracket stands for {1,,2d}\left\{1,\dots,2d\right\}. For xTMx\in T^{*}M and h>0h>0, the open and the closed cubes of center xx and radius hh are given by

C(x,h)={yTMxy<h},C¯(x,h)={yTMxyh}.C(x,h)=\{y\in T^{*}M\mid\|x-y\|<h\},\qquad\overline{C}(x,h)=\{y\in T^{*}M\mid\|x-y\|\leq h\}.

We denote by Mh={mnn}M_{h}=\left\{m_{n}\mid n\in{\mathbb{N}}\right\} a regular grid of points in VV such that

TM=nC¯(mn,h),T^{*}M=\cup_{n\in{\mathbb{N}}}\overline{C}(m_{n},h),

while the open cubes C(mn,h)C(m_{n},h), nn\in{\mathbb{N}}, are pairwise disjoint. We call such a MhM_{h} a mesh of size hh.

Definition 12.

Given h>0h>0 and a mesh MhM_{h} of size hh, a map F:TMTMF:T^{*}M\to T^{*}M is said to be a permutation of MhM_{h} if FF is bijective and FF translates every open cube of the mesh to another one, that is, for every nn\in{\mathbb{N}} there exists \ell\in{\mathbb{N}} such that, for every xC(mn,h)x\in C(m_{n},h), F(x)=x+mmnF(x)=x+m_{\ell}-m_{n}.

Lemma 13.

Let ρ0,ρ1Lr(TM){\rho}_{0},{\rho}_{1}\in L^{r}(T^{*}M) be such that ρ0ρ1{\rho}_{0}\sim{\rho}_{1}. For every ε>0\varepsilon>0, there exists h0>0h_{0}>0 such that for every h(0,h0)h\in(0,h_{0}) and every mesh MhM_{h} of size hh, there exists a permutation Fε,hF_{\varepsilon,h} of MhM_{h} such that

ρ0Fε,hρ1Lr<ε.\|{\rho}_{0}\circ F_{\varepsilon,h}-{\rho}_{1}\|_{L^{r}}<\varepsilon.
Proof.

First step: Let 0<a<A0<a<A. For i{0,1}i\in\{0,1\}, by dominated convergence we have

ρi𝟙{a<|ρi|<A}ρiLra0,A+0.\|{\rho}_{i}{\mathbb{1}}_{\{a<|{\rho}_{i}|<A\}}-{\rho}_{i}\|_{L^{r}}\underset{a\rightarrow 0,A\rightarrow+\infty}{\longrightarrow}0.

Moreover we still have

ρ0𝟙{a<|ρ0|<A}ρ1𝟙{a<|ρ1|<A},{\rho}_{0}{\mathbb{1}}_{\{a<|{\rho}_{0}|<A\}}\sim{\rho}_{1}{\mathbb{1}}_{\{a<|{\rho}_{1}|<A\}},

so in the following we can consider without loss of generality that there exist 0<a<A0<a<A and a set WTMW\subset T^{*}M of finite measure, such that a<|ρ0(x)|,|ρ1(x)|<Aa<|{\rho}_{0}(x)|,|{\rho}_{1}(x)|<A for xWx\in W and ρ0(x)=ρ1(x)=0{\rho}_{0}(x)={\rho}_{1}(x)=0 for xTMWx\in T^{*}M\setminus W.

Second step: Let us approximate the densities ρ0\rho_{0} and ρ1\rho_{1} by step functions. Given NN\in{\mathbb{N}} we set ξ0=A\xi_{0}=-A and ξk=ξk1+2AN\xi_{k}=\xi_{k-1}+\frac{2A}{N} for every k\llbracket1,N\rrbracketk\in\llbracket 1,N\rrbracket. For i{0,1}i\in\{0,1\}, let

INi=k=1Nξk𝟙{ξk1ρi<ξk},I_{N}^{i}=\sum_{k=1}^{N}\xi_{k}{\mathbb{1}}_{\left\{\xi_{k-1}\leq{\rho}_{i}<\xi_{k}\right\}},

and thanks to the previous step, we deduce that INiNρiI_{N}^{i}\underset{N\rightarrow\infty}{\longrightarrow}{\rho}_{i} in Lr\text{in }L^{r}. Notice that for NN large enough, by the previous step, either ξk10\xi_{k-1}\neq 0 (and Vol({ξk1ρi<ξk})<+\operatorname{Vol}(\{\xi_{k-1}\leq{\rho}_{i}<\xi_{k}\})<+\infty) or ρi0\rho_{i}\equiv 0 on {ξk1ρi<ξk}\{\xi_{k-1}\leq{\rho}_{i}<\xi_{k}\}.

Third step: Let us prove now that the sub-level sets of ρ0\rho_{0} and ρ1\rho_{1} can be approximately covered by the same number of cubes. Consider i{0,1}i\in\{0,1\}, ξ<ξ\xi<\xi^{\prime} in {\mathbb{R}}, set

Vi:={xTMξρi(x)<ξ},V^{i}:=\left\{x\in T^{*}M\mid\xi\leq{\rho}_{i}(x)<\xi^{\prime}\right\},

and assume that Vol(Vi)<+\operatorname{Vol}(V^{i})<+\infty.
Let WiTMW^{i}\subset T^{*}M be open and such that ViWiV^{i}\subset W^{i}, Vol(Wi)<Vol(Vi)+ε\operatorname{Vol}(W^{i})<\operatorname{Vol}(V^{i})+\varepsilon, where ε>0\varepsilon>0 is arbitrary. Set, moreover, for η>0{\eta}>0,

Wηi:={xWid(x,Wi)>η},W_{{\eta}}^{i}:=\left\{x\in W^{i}\mid d(x,\partial W^{i})>{\eta}\right\},

where d(x,Wi)=infyWixyd(x,\partial W^{i})=\inf_{y\in\partial W^{i}}\|x-y\|. By dominated convergence, 𝟙Wηi𝟙Wi{\mathbb{1}}_{W^{i}_{{\eta}}}\rightarrow{\mathbb{1}}_{W^{i}} in LrL^{r} as η{\eta} goes to zero. In particular, 0<Vol(Wi)Vol(Wηi)<ε0<\operatorname{Vol}(W^{i})-\operatorname{Vol}(W^{i}_{{\eta}})<\varepsilon, for η{\eta} small enough, and

Vol(WηiΔVi)=Vol(WηiVi)+Vol(ViWηi)Vol(WiVi)+Vol(WiWηi)<2ε,\operatorname{Vol}(W^{i}_{\eta}\Delta V^{i})=\operatorname{Vol}(W^{i}_{\eta}\setminus V^{i})+\operatorname{Vol}(V^{i}\setminus W^{i}_{\eta})\leq\operatorname{Vol}(W^{i}\setminus V^{i})+\operatorname{Vol}(W^{i}\setminus W^{i}_{\eta})<2\varepsilon,

where we used Δ\Delta to denote the symmetric difference. In particular, |Vol(Wηi)Vol(Vi)|Vol(WηiΔVi)<2ε|\operatorname{Vol}(W^{i}_{\eta})-\operatorname{Vol}(V^{i})|\leq\operatorname{Vol}(W^{i}_{\eta}\Delta V^{i})<2\varepsilon. Given h>0h>0 and a mesh Mh={mnn}M_{h}=\left\{m_{n}\mid n\in{\mathbb{N}}\right\} of size hh, consider the minimal set JiJ_{i}\subset{\mathbb{N}} such that W2hinJiC¯(mn,h)W^{i}_{2h}\subset\cup_{n\in J_{i}}\overline{C}(m_{n},h). Notice that if nJin\in J_{i} then C¯(mn,h)Wi\overline{C}(m_{n},h)\subset W^{i}. Hence, for hh small enough,

Vol(Wi)ε<Vol(W2hi)Vol(nJiC(mn,h))Vol(Wi)\operatorname{Vol}(W^{i})-\varepsilon<\operatorname{Vol}(W^{i}_{2h})\leq\operatorname{Vol}(\cup_{n\in J_{i}}C(m_{n},h))\leq\operatorname{Vol}(W^{i})

from which we deduce that

Vol(ViΔ(nJiC(mn,h)))Vol(WiVi)+Vol(WinJiC(mn,h)))<2ε.\operatorname{Vol}(V^{i}\Delta(\cup_{n\in J_{i}}C(m_{n},h)))\leq\operatorname{Vol}(W^{i}\setminus V^{i})+\operatorname{Vol}(W^{i}\setminus\cup_{n\in J_{i}}C(m_{n},h)))<2\varepsilon. (11)

Now, denote by J^i\hat{J}_{i} the set of indices nJin\in J_{i} such that Vol(C(mn,h)Vi)>12Vol(C(mn,h))\operatorname{Vol}(C(m_{n},h)\cap V_{i})>\frac{1}{2}\operatorname{Vol}(C(m_{n},h)), so that

ε>Vol(WiVi)nJiJ^iVol(C(mn,h)Vi)12Vol(nJiJ^iC(mn,h)).\varepsilon>\operatorname{Vol}(W^{i}\setminus V^{i})\geq\sum_{n\in J_{i}\setminus\hat{J}_{i}}\operatorname{Vol}(C(m_{n},h)\setminus V^{i})\geq\frac{1}{2}\operatorname{Vol}\left(\cup_{n\in J_{i}\setminus\hat{J}_{i}}C(m_{n},h)\right). (12)

According to (11) and (12), we then have

Vol(Vi)+ε>Vol(nJ^iC(mn,h))>Vol(nJiC(mn,h))2ε>Vol(Vi)4ε.\operatorname{Vol}(V^{i})+\varepsilon>\operatorname{Vol}(\cup_{n\in\hat{J}_{i}}C(m_{n},h))>\operatorname{Vol}(\cup_{n\in J_{i}}C(m_{n},h))-2\varepsilon>\operatorname{Vol}(V^{i})-4\varepsilon.

Hence |Vol(nJ^iC(mn,h))Vol(Vi)|<4ε|\operatorname{Vol}(\cup_{n\in\hat{J}_{i}}{C}(m_{n},h))-\operatorname{Vol}(V^{i})|<4\varepsilon. In particular, since Vol(V0)=Vol(V1)<+\operatorname{Vol}(V^{0})=\operatorname{Vol}(V^{1})<+\infty and denoting by J^i\sharp\hat{J}_{i} the cardinal of J^i\hat{J}_{i}, it follows that h2d|J^0J^1|<8εh^{2d}|\sharp\hat{J}_{0}-\sharp\hat{J}_{1}|<8\varepsilon.
Let us prove that we can approximate V0V^{0} and V1V^{1} by the same number of cubes. Pick J~1J^1\tilde{J}_{1}\subset\hat{J}_{1}, J~2J^2\tilde{J}_{2}\subset\hat{J}_{2} such that J~1=J~2=min(J^1,J^2)\sharp\tilde{J}_{1}=\sharp\tilde{J}_{2}=\min(\sharp\hat{J}_{1},\sharp\hat{J}_{2}). Then, for i{0,1}i\in\{0,1\},

|Vol(Vi)Vol(nJ~iC(mn,h))||Vol(Vi)Vol(nJ^iC(mn,h))|+(2h)2d|J^0J^1|<12ε.|\operatorname{Vol}(V^{i})-\operatorname{Vol}(\cup_{n\in\tilde{J}_{i}}C(m_{n},h))|\leq|\operatorname{Vol}(V^{i})-\operatorname{Vol}(\cup_{n\in\hat{J}_{i}}C(m_{n},h))|+(2h)^{2d}|\sharp\hat{J}_{0}-\sharp\hat{J}_{1}|<12\varepsilon.


Fourth step: We fix ε>0\varepsilon>0, NN\in{\mathbb{N}} such that INiρiLr<ε\|I^{i}_{N}-\rho_{i}\|_{L^{r}}<\varepsilon for i{0,1}i\in\{0,1\}, and we apply the previous step to the each of the sub-level sets

Vki={ξk1ρi<ξk},i{0,1},k\llbracket1,N\rrbracket,V_{k}^{i}=\{\xi_{k-1}\leq{\rho}_{i}<\xi_{k}\},\qquad i\in\{0,1\},\ k\in\llbracket 1,N\rrbracket,

with 0[ξk1,ξk)0\not\in[\xi_{k-1},\xi_{k}). We recall that by construction, Vol(Vki)<+\operatorname{Vol}(V^{i}_{k})<+\infty for 0[ξk1,ξk)0\not\in[\xi_{k-1},\xi_{k}), so we are indeed allowed to apply the third step.
Then, given η=η(ε)>0{\eta}={\eta}(\varepsilon)>0 to be fixed later, there exists h0>0h_{0}>0 such that for every h(0,h0)h\in(0,h_{0}), every mesh MhM_{h} of size hh, and every k\llbracket1,N\rrbracketk\in\llbracket 1,N\rrbracket with 0[ξk1,ξk)0\not\in[\xi_{k-1},\xi_{k}), there exist Jk0,Jk1J_{k}^{0},J_{k}^{1}\subset{\mathbb{N}} such that Jk0=Jk1\sharp J_{k}^{0}=\sharp J_{k}^{1} and

|Vol(Vki)Vol(nJkiC(mn,h))|<η,i{0,1}.|\operatorname{Vol}(V_{k}^{i})-\operatorname{Vol}(\cup_{n\in J_{k}^{i}}C(m_{n},h))|<{\eta},\qquad i\in\{0,1\}.

Moreover, the cubes indexed by JkiJ_{k}^{i} intersect VkiV_{k}^{i} for more than half their volume, so, for ii fixed, the sets JkiJ_{k}^{i} are disjoint. We can construct a permutation Fε,hF_{\varepsilon,h} of the mesh MhM_{h} that translates the cubes of Jk0J_{k}^{0} to the cubes of Jk1J_{k}^{1}. Then, since Fε,hF_{\varepsilon,h} preserves the volume,

TM|𝟙Vk0Fε,h𝟙Vk1|\displaystyle\int_{T^{*}M}|{\mathbb{1}}_{V^{0}_{k}}\circ F_{\varepsilon,h}-{\mathbb{1}}_{V^{1}_{k}}| =Vol(Vk1Fε,h(Vk0))+Vol(Fε,h(Vk0)Vk1)\displaystyle=\operatorname{Vol}(V^{1}_{k}\setminus F_{\varepsilon,h}(V^{0}_{k}))+\operatorname{Vol}(F_{\varepsilon,h}(V^{0}_{k})\setminus V^{1}_{k})
=Vol(Vk1Fε,h(Vk0))+Vol(Vk0Fε,h1(Vk1))\displaystyle=\operatorname{Vol}(V^{1}_{k}\setminus F_{\varepsilon,h}(V^{0}_{k}))+\operatorname{Vol}(V^{0}_{k}\setminus F_{\varepsilon,h}^{-1}(V^{1}_{k}))
Vol(Vk1nJk1C(mn,h))+Vol(Vk0nJk0C(mn,h))<2η.\displaystyle\leq\operatorname{Vol}(V^{1}_{k}\setminus\cup_{n\in J_{k}^{1}}C(m_{n},h))+\operatorname{Vol}(V^{0}_{k}\setminus\cup_{n\in J^{0}_{k}}C(m_{n},h))<2{\eta}.

Therefore,

TM|IN0Fε,hIN1|r\displaystyle\int_{T^{*}M}|I^{0}_{N}\circ F_{\varepsilon,h}-I^{1}_{N}|^{r} k\llbracket1,N\rrbracket0[ξk1,ξk)TMAr|𝟙Vk0Fε,h𝟙Vk1|<2NArη,\displaystyle\leq\sum_{\footnotesize\begin{array}[]{c}k\in\llbracket 1,N\rrbracket\\ 0\not\in[\xi_{k-1},\xi_{k})\end{array}}\int_{T^{*}M}A^{r}|{\mathbb{1}}_{V^{0}_{k}}\circ F_{\varepsilon,h}-{\mathbb{1}}_{V^{1}_{k}}|<2NA^{r}{\eta},

so that

IN0Fε,hIN1Lr<A(2Nη)1/r.\|I_{N}^{0}\circ F_{\varepsilon,h}-I^{1}_{N}\|_{L^{r}}<A(2N{\eta})^{1/r}.

In conclusion, taking η=12N(εA)r{\eta}=\frac{1}{2N}(\frac{\varepsilon}{A})^{r},

ρ0Fε,hρ1Lr\displaystyle\|{\rho}_{0}\circ F_{\varepsilon,h}-{\rho}_{1}\|_{L^{r}} ρ0Fε,hIN0Fε,hLr+IN0Fε,hIN1Lr+IN1ρ1Lr\displaystyle\leq\|{\rho}_{0}\circ F_{\varepsilon,h}-I^{0}_{N}\circ F_{\varepsilon,h}\|_{L^{r}}+\|I_{N}^{0}\circ F_{\varepsilon,h}-I_{N}^{1}\|_{L^{r}}+\|I_{N}^{1}-{\rho}_{1}\|_{L^{r}}
=ρ0IN0Lr+IN0Fε,hIN1Lr+IN1ρ1Lr<3ε.\displaystyle=\|{\rho}_{0}-I^{0}_{N}\|_{L^{r}}+\|I_{N}^{0}\circ F_{\varepsilon,h}-I_{N}^{1}\|_{L^{r}}+\|I_{N}^{1}-{\rho}_{1}\|_{L^{r}}<3\varepsilon.

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 ρLr(TM){\rho}\in L^{r}(T^{*}M) and ε>0\varepsilon>0, there exists h0>0h_{0}>0 such that for every h(0,h0)h\in(0,h_{0}), every mesh MhM_{h} of size hh, and every permutation FhF_{h} of MhM_{h}, there exists ϕDHam(TM)\phi\in\operatorname{DHam}(T^{*}M) such that

ρFhρϕLr<ε.\|{\rho}\circ F_{h}-{\rho}\circ\phi\|_{L^{r}}<\varepsilon.
Proof.

First step: Since the set of smooth functions with compact support 𝒞c(TM)\mathcal{C}^{\infty}_{c}(T^{*}M) is dense in Lr(TM)L^{r}(T^{*}M), there exists a smooth function ρ~\tilde{{\rho}} with compact support such that ρ~ρLr<ε\|\tilde{{\rho}}-{\rho}\|_{L^{r}}<\varepsilon. So in the following we assume that ρ𝒞c(TM){\rho}\in\mathcal{C}^{\infty}_{c}(T^{*}M) and we denote its compact support by KVK\subset V.

Second step: By compactness, for every h>0h>0 and every mesh MhM_{h}, KK can be covered by a finite number of cubes of MhM_{h}, that is, there exists NN\in{\mathbb{N}} (depending on hh and MhM_{h}) such that Kn=0NC¯(mn,h)K\subset\cup_{n=0}^{N}\overline{C}(m_{n},h). Since ρ{\rho} is continuous over the compact KK, then it is uniformly continuous. So for every h>0h>0 small enough there exists η>0\eta>0 such that for every mesh MhM_{h} of size hh,

ρn=0Nρ(mn)𝟙C(mn,hη)Lr<ε.\left\|{\rho}-\sum_{n=0}^{N}{\rho}(m_{n}){\mathbb{1}}_{C(m_{n},h-\eta)}\right\|_{L_{r}}<\varepsilon.


Third step: Given h>0h>0, a mesh MhM_{h} of size hh, and a permutation FhF_{h} of MhM_{h}, we look for a Hamiltonian diffeomorphism ϕ\phi such that 𝟙C(mn,hη)ϕ{\mathbb{1}}_{C(m_{n},h-\eta)}\circ\phi approximates 𝟙C(mn,hη)Fh{\mathbb{1}}_{C(m_{n},h-\eta)}\circ F_{h} in Lr(TM)L^{r}(T^{*}M) for every n\llbracket0,N\rrbracketn\in\llbracket 0,N\rrbracket. The image of mn=(qn,pn)m_{n}=(q^{n},p^{n}) by FhF_{h} is denoted by m¯n=(q¯n,p¯n)\bar{m}_{n}=(\bar{q}^{n},\bar{p}^{n}), so that Fh(C(mn,h))=C(m¯n,h)F_{h}(C(m_{n},h))=C(\bar{m}_{n},h), see Figure 1.

[Uncaptioned image]
Figure 1: The permutation FhF_{h}

As a preliminary step, let us construct a Hamiltonian diffeomorphism translating each cube C(mn,hη)C(m_{n},h-\eta), n\llbracket0,N\rrbracketn\in\llbracket 0,N\rrbracket to a cube of center (qn,on)({q}^{n},o^{n}), where on1on22h\|o^{n_{1}}-o^{n_{2}}\|\geq 2h for every n1n2n_{1}\neq n_{2} in \llbracket0,N\rrbracket\llbracket 0,N\rrbracket, see Figure 2. The cubes of the mesh that cover KK are organized in columns: we choose J\llbracket0,N\rrbracketJ\subset\llbracket 0,N\rrbracket of minimal cardinality such that {qjjJ}={qnn\llbracket0,N\rrbracket}\{q^{j}\mid j\in J\}=\{q^{n}\mid n\in\llbracket 0,N\rrbracket\} and we set Cj={n\llbracket0,N\rrbracketqn=qj}C_{j}=\{n\in\llbracket 0,N\rrbracket\mid q^{n}=q^{j}\} for jJj\in J. Given {ajjJ}\{a^{j}\mid j\in J\} in d{\mathbb{R}}^{d}, we consider a function f𝒞(TM,)f\in{\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}) with compact support and such that f(q,p)=ajqf(q,p)=a^{j}\cdot q if qqj<hη\|q-q^{j}\|<h-\eta and pp is at distance at most |aj|+hη|a^{j}|+h-\eta from {pnnCj}\{p^{n}\mid n\in C_{j}\}. Then

ef(q,p)=(q,paj),(q,p)C(mn,hη),nCj.e^{\overrightarrow{f}}(q,p)=(q,p-a^{j}),\qquad\forall\>(q,p)\in C(m_{n},h-\eta),\;\forall\>n\in C_{j}.

So, up to a suitable choice of the vectors aja^{j}, efe^{\overrightarrow{f}} is the required Hamiltonian diffeomorphism translating each cube C(mn,hη)C(m_{n},h-\eta), n\llbracket0,N\rrbracketn\in\llbracket 0,N\rrbracket, to a cube of center (qn,on)({q}^{n},o^{n}), where on1on22h\|o^{n_{1}}-o^{n_{2}}\|\geq 2h for every n1n2n_{1}\neq n_{2} in \llbracket0,N\rrbracket\llbracket 0,N\rrbracket.

[Uncaptioned image]
Figure 2: The Hamiltonian transformation efe^{\overrightarrow{f}}

Then we want to translate every cube C((qn,on),hη)C((q^{n},o^{n}),h-\eta), n\llbracket0,N\rrbracketn\in\llbracket 0,N\rrbracket, to the cube C((q¯n,on),hη)C((\bar{q}^{n},o^{n}),h-\eta) by a localized horizontal translation, see Figure 3. We consider a function g𝒞(TM,)g\in{\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}) with compact support and such that g(q,p)=(q¯nqn)pg(q,p)=(\bar{q}^{n}-q^{n})\cdot p for every (q,p)(q,p) such that pon<hη\|p-o^{n}\|<h-\eta and qq is at distance at most hηh-\eta from the segment connecting qnq^{n} and q¯n\bar{q}^{n}.

Then

eg(q,p)=(q+q¯nqn,p),(q,p)C((qn,on),hη).e^{\overrightarrow{g}}(q,p)=(q+\bar{q}^{n}-q^{n},p),\qquad\forall\>(q,p)\in C((q^{n},o^{n}),h-\eta).
[Uncaptioned image]
Figure 3: The Hamiltonian transformation ege^{\overrightarrow{g}}

With the two previous transformations we obtained a new grid {(q¯n,on)n\llbracket0,N\rrbracket}\{(\bar{q}^{n},o^{n})\mid n\in\llbracket 0,N\rrbracket\} with the good qq-components. We look for a Hamiltonian diffeomorphism that re-shuffles the cubes whose center have the same qq-component q¯n\bar{q}^{n} without altering the others. We have already seen how to approximate localized translations in the pp-coordinate, see Figure 4. In the case where d2d\geq 2 the desired re-shuffle can be obtained by iterating sufficiently many such transitions. In the case d=1d=1 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]
Figure 4: Localized vertical translations

Let us consider two consecutive cubes of the same column, with centers (q~,on1)(\tilde{q},o^{n_{1}}) and (q~,on2)(\tilde{q},o^{n_{2}}), where q~=q¯n1=q¯n2\tilde{q}=\bar{q}^{n_{1}}=\bar{q}^{n_{2}}. Consecutive means that the segment between (q~,on1)(\tilde{q},o^{n_{1}}) and (q~,on2)(\tilde{q},o^{n_{2}}) does not contain any other element of the grid. Moreover, we can assume that the other cubes with qq-component equal to q~\tilde{q} are as much separated as required.
We are going to permute the cubes with a rotation of center (q~,p~)(\tilde{q},\tilde{p}), where p~=on1+on22\tilde{p}=\frac{o^{n_{1}}+o^{n_{2}}}{2}, and of angle π2\frac{\pi}{2}. We consider a function hw𝒞(TM,)h_{w}\in{\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}) such that

hw(q,p)=|pp~|22+|qq~|22w2h_{w}(q,p)=\frac{|p-\tilde{p}|^{2}}{2}+\frac{|q-\tilde{q}|^{2}}{2w^{2}}

in

Ωw={(q,p)|qq~|<h,|pp~|<hw},\Omega_{w}=\left\{(q,p)\mid|q-\tilde{q}|<h,\;|p-\tilde{p}|<\frac{h}{w}\right\},

with w>0w>0 to be fixed.

The Hamiltonian system associated with hwh_{w} has equations

{ddt(q(t)q~)=p(t)p~,ddt(p(t)p~)=1w2(q(t)q~),\left\{\begin{array}[]{ll}\frac{d}{dt}(q(t)-\tilde{q})=p(t)-\tilde{p},\\ \frac{d}{dt}(p(t)-\tilde{p})=-\frac{1}{w^{2}}(q(t)-\tilde{q}),\end{array}\right. (13)

in Ωw\Omega_{w}. As long as they stay in Ωw\Omega_{w} its solutions have the expression q(t)=q~+r0sin(tw+θ0)q(t)=\tilde{q}+r_{0}\sin(\frac{t}{w}+\theta_{0}), p(t)=p~+r0wcos(tw+θ0)p(t)=\tilde{p}+\frac{r_{0}}{w}\cos(\frac{t}{w}+\theta_{0}). Hence, for ww small enough, each solution of (13) with initial condition in C((q~,on1),hη)C((q~,on2),hη)C((\tilde{q},o^{n_{1}}),h-\eta)\cup C((\tilde{q},o^{n_{2}}),h-\eta) stays in Ωw\Omega_{w} forever. Moreover, we can assume that the support of hwh_{w} is compact and does not intersect the cubes C((q¯n,on),hη)C((\bar{q}^{n},o^{n}),h-\eta) for nn1,n2n\neq n_{1},n_{2}. Then ewπhwe^{w\pi\overrightarrow{h_{w}}} permutes the cubes C((q~,on1),hη)C((\tilde{q},o^{n_{1}}),h-\eta) and C((q~,on2),hη)C((\tilde{q},o^{n_{2}}),h-\eta) and is the identity on C((q¯n,on),hη)C((\bar{q}^{n},o^{n}),h-\eta) for nn1,n2n\neq n_{1},n_{2}.

[Uncaptioned image]
Figure 5: Localized Hamiltonian rotation of angle π2\frac{\pi}{2}

Fourth step: For every h>η>0h>\eta>0 and every permutation FhF_{h} of a mesh MhM_{h}, we have constructed a Hamiltonian diffeomorphism ϕ\phi such that ϕ(C(mn,hη))=Fh(C(mn,hη))\phi(C(m_{n},h-\eta))=F_{h}(C(m_{n},h-\eta)) for every n\llbracket0,N\rrbracketn\in\llbracket 0,N\rrbracket, so that

n=0Nρ(mn)𝟙C(mn,hη)Fh=n=0Nρ(mn)𝟙C(mn,hη)ϕ.\sum_{n=0}^{N}{\rho}(m_{n}){\mathbb{1}}_{C(m_{n},h-\eta)}\circ F_{h}=\sum_{n=0}^{N}{\rho}(m_{n}){\mathbb{1}}_{C(m_{n},h-\eta)}\circ\phi.

Therefore, according to the second step, for hh smal enough,

ρFhρϕLr<2ε.\|{\rho}\circ F_{h}-{\rho}\circ\phi\|_{L^{r}}<2\varepsilon.

Proof of Theorem 3. Given ρ1(ρ0){\rho}_{1}\in{\mathcal{L}}({\rho}_{0}), we must prove that ρ1𝒪(ρ0)¯{\rho}_{1}\in\overline{\mathcal{O}({\rho}_{0})}. According to Lemma 13, for every ε>0\varepsilon>0, every h>0h>0 small enough, and every mesh MhM_{h} of size hh, there exists a permutation Fε,hF_{\varepsilon,h} of MhM_{h} such that

ρ0Fε,hρ1Lr<ε.\|{\rho}_{0}\circ F_{\varepsilon,h}-{\rho}_{1}\|_{L^{r}}<\varepsilon.

Moreover, according to Lemma 14, up to further reducing hh, there exists ϕhDHam(TM)\phi_{h}\in\operatorname{DHam}(T^{*}M) such that

ρ0Fε,hρ0ϕhLr<ε,\|{\rho}_{0}\circ F_{\varepsilon,h}-{\rho}_{0}\circ\phi_{h}\|_{L^{r}}<\varepsilon,

whence ρ1ρ0ϕhLr<2ε.\|{\rho}_{1}-{\rho}_{0}\circ\phi_{h}\|_{L^{r}}<2\varepsilon. \square

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, Vec(M){\rm Vec}(M) denotes the space of 𝒞{\mathcal{C}^{\infty}} vector fields on the manifold MM, which can be endowed with the compact-open topology, that is, the topology identified by the family of semi-norms ,K=C(K)\|\cdot\|_{\ell,K}=\|\cdot\|_{C^{\ell}(K)}, where \ell\in{\mathbb{N}} and KK is compact in MM. The group Diff(M){\rm Diff}(M) is also endowed with the compact-open topology.

Proposition 15.

Let fnVec(M)f_{n}\in{\rm Vec}(M), nn\in{\mathbb{N}}, and fVec(M)f\in{\rm Vec}(M) be complete and such that fnff_{n}\to f for the compact-open topology of Vec(M){\rm Vec}(M). Then, for any tt\in{\mathbb{R}}, etfnetfe^{tf_{n}}\to e^{tf} for the compact-open topology of Diff(M){\rm Diff}(M).

The following property is the analog for vector fields of the Lie product formula.

Proposition 16.

Let f,gVec(M)f,g\in{\rm Vec}(M) be complete and such that f+gf+g is complete. Then,

limn(ef/neg/n)n=ef+g,\lim_{n\to\infty}(e^{f/n}e^{g/n})^{n}=e^{f+g},

for the compact-open topology of Diff(M){\rm Diff}(M).

In order to approximate some ϕDHam(TM)\phi\in\operatorname{DHam}(T^{*}M) in the compact-open topology, we can apply a diagonal argument (based on the exhaustion of MM by compact sets) and reduce the problem to that of approximating ϕ\phi in the CC^{\ell}-topology on a given compact for a given \ell\in{\mathbb{N}}. This classical fact is recalled in the following lemma.

Lemma 17.

Let ϕDiff(TM)\phi\in{\rm Diff}(T^{*}M) and 𝒟Diff(TM)\mathcal{D}\subset{\rm Diff}(T^{*}M). Assume that for every compact set KTMK\subset T^{*}M and every \ell\in{\mathbb{N}} there exists a sequence (φn)n(\varphi_{n})_{n\in{\mathbb{N}}} in 𝒟\mathcal{D} such that

φnϕ,Kn0.\|\varphi_{n}-\phi\|_{\ell,K}\underset{n\rightarrow\infty}{\longrightarrow}0.

Then there exists a sequence (ϕn)n(\phi_{n})_{n\in{\mathbb{N}}} in 𝒟\mathcal{D} such that

ϕnϕ,Kn0,KTMcompact.\|\phi_{n}-\phi\|_{\ell,K}\underset{n\rightarrow\infty}{\longrightarrow}0\qquad\forall\>\ell\in{\mathbb{N}},\>\forall K\subset T^{*}M\>\text{compact.}

Given XVec(TM)X\in\operatorname{Vec}(T^{*}M) its associated adjoint operator is denoted by

adX:Vec(TM)Vec(TM),adXY=[X,Y].\operatorname{ad}_{X}:\operatorname{Vec}(T^{*}M)\rightarrow\operatorname{Vec}(T^{*}M),\quad\operatorname{ad}_{X}Y=[X,Y].

Given ϕDiff(TM)\phi\in\operatorname{Diff}(T^{*}M), we denote its pushforward action on vector fields as

ϕ:Vec(TM)Vec(TM),(ϕX)(x)=Dϕ(ϕ1(x))X(ϕ1(x)).\phi_{*}:\operatorname{Vec}(T^{*}M)\rightarrow\operatorname{Vec}(T^{*}M),\quad(\phi_{*}X)(x)=D\phi(\phi^{-1}(x))X(\phi^{-1}(x)).

Recall that if ϕ\phi is a syplectomorphism and h𝒞(TM,)h\in{\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}), then

ϕehϕ1=eϕh=ehϕ.\phi e^{\overrightarrow{h}}\phi^{-1}=e^{\phi_{*}\overrightarrow{h}}=e^{\overrightarrow{h\circ\phi}}. (14)

The following estimate can be found in [6, Section 8] (see also [4]).

Proposition 18.

Let \ell\in{\mathbb{N}} and KTMK\subset T^{*}M be compact. There exists a compact neighborhood KK^{\prime} of KK and a constant c=c(K,)>0c=c(K^{\prime},\ell)>0 such that, for every X,YVec(TM)X,Y\in\operatorname{Vec}(T^{*}M),

(eX)Yj=0N1adXjYj!,KcecX+1,KX+N,KNN!Y+N,K.\left\|(e^{X})_{*}Y-\sum_{j=0}^{N-1}\frac{\operatorname{ad}^{j}_{X}Y}{j!}\right\|_{\ell,K}\leq ce^{c\|X\|_{\ell+1,K^{\prime}}}\frac{\|X\|^{N}_{\ell+N,K^{\prime}}}{N!}\|Y\|_{\ell+N,K^{\prime}}.

4 Some properties of the approximately reachable sets

4.1 Some properties of st¯\overline{\mathcal{R}_{\rm st}}

We show in the next proposition that the small-time approximately reachable set st¯\overline{\mathcal{R}_{\rm st}} is a closed semi-group.

Proposition 19.

The following holds:

  • (i)

    If ϕ,ψst¯\phi,\psi\in\overline{\mathcal{R}_{\rm st}}, then ψϕst¯\psi\circ\phi\in\overline{\mathcal{R}_{\rm st}}.

  • (ii)

    If (ϕn)nst¯(\phi_{n})_{n\in{\mathbb{N}}}\subset\overline{\mathcal{R}_{\rm st}} is such that ϕnϕ\phi_{n}\to\phi, then ϕst¯\phi\in\overline{\mathcal{R}_{\rm st}}.

Proof.

Proof of (i): Let \ell\in{\mathbb{N}} and KTMK\subset T^{*}M be compact. Given ε>0\varepsilon>0 there exist u(),v()PWC([0,ε],m)u(\cdot),v(\cdot)\in\operatorname{PWC}([0,\varepsilon],{\mathbb{R}}^{m}) and τu,τv[0,ε]\tau_{u},\tau_{v}\in[0,\varepsilon] such that

ΦHuτuϕ,K<εandΦHvτvψ,ϕ(K)<ε.\|\Phi_{H_{u}}^{\tau_{u}}-\phi\|_{\ell,K}<\varepsilon\qquad\text{and}\qquad\|\Phi_{H_{v}}^{\tau_{v}}-\psi\|_{\ell,\phi(K)}<\varepsilon.

Consider w()PWC([0,τu+τv],m)w(\cdot)\in\operatorname{PWC}([0,\tau_{u}+\tau_{v}],{\mathbb{R}}^{m}) such that

{w(t)=u(t)if t[0,τu]w(t)=v(tτu)if t(τu,τu+τv].\left\{\begin{array}[]{ll}w(t)=u(t)\quad\text{if }t\in[0,\tau_{u}]\\ w(t)=v(t-\tau_{u})\quad\text{if }t\in(\tau_{u},\tau_{u}+\tau_{v}].\end{array}\right.

Then ΦHwτu+τu=ΦHvτvΦHuτu\Phi_{H_{w}}^{\tau_{u}+\tau_{u}}=\Phi_{H_{v}}^{\tau_{v}}\circ\Phi_{H_{u}}^{\tau_{u}} and

ΦHwτu+τvψϕ,K\displaystyle\|\Phi_{H_{w}}^{\tau_{u}+\tau_{v}}-\psi\circ\phi\|_{\ell,K} ΦHvτvΦHuτuΦHvτvϕ,K+ΦHvτvϕψϕ,K\displaystyle\leq\|\Phi_{H_{v}}^{\tau_{v}}\circ\Phi_{H_{u}}^{\tau_{u}}-\Phi_{H_{v}}^{\tau_{v}}\circ\phi\|_{\ell,K}+\|\Phi_{H_{v}}^{\tau_{v}}\circ\phi-\psi\circ\phi\|_{\ell,K}
=ΦHvτvΦHuτuΦHvτvϕ,K+ΦHvτvψ,ϕ(K)\displaystyle=\|\Phi_{H_{v}}^{\tau_{v}}\circ\Phi_{H_{u}}^{\tau_{u}}-\Phi_{H_{v}}^{\tau_{v}}\circ\phi\|_{\ell,K}+\|\Phi_{H_{v}}^{\tau_{v}}-\psi\|_{\ell,\phi(K)}
<ΦHvτvΦHuτuΦHvτvϕ,K+ε.\displaystyle<\|\Phi_{H_{v}}^{\tau_{v}}\circ\Phi_{H_{u}}^{\tau_{u}}-\Phi_{H_{v}}^{\tau_{v}}\circ\phi\|_{\ell,K}+\varepsilon.

There exists a compact K~V\tilde{K}\subset V such that ΦHuτu(K)ϕ(K)K~\Phi_{H_{u}}^{\tau_{u}}(K)\cup\phi(K)\subset\tilde{K} independently of ε\varepsilon. Moreover, we can assume that there exists C>0C>0 depending on K~\tilde{K}, \ell, and ϕ\phi (and independent of ε\varepsilon) such that ΦHvτv+1,K~<C\|\Phi_{H_{v}}^{\tau_{v}}\|_{\ell+1,\tilde{K}}<C. By applying the mean value theorem, we obtain that

ΦHvτvΦHuτuΦHvτvϕ,K\displaystyle\|\Phi_{H_{v}}^{\tau_{v}}\circ\Phi_{H_{u}}^{\tau_{u}}-\Phi_{H_{v}}^{\tau_{v}}\circ\phi\|_{\ell,K} ΦHvτv+1,K~ΦHuτuϕ,K<Cε.\displaystyle\leq\|\Phi_{H_{v}}^{\tau_{v}}\|_{\ell+1,\tilde{K}}\|\Phi_{H_{u}}^{\tau_{u}}-\phi\|_{\ell,K}<{C}\varepsilon.

Proof of (ii): Let ε>0\varepsilon>0, \ell\in{\mathbb{N}}, and KTMK\subset T^{*}M be compact. Let nn be such that ϕnϕ,K<ε\|\phi_{n}-\phi\|_{\ell,K}<\varepsilon. Since ϕnst¯\phi_{n}\in\overline{\mathcal{R}_{\rm st}}, there exist τ[0,ε]\tau\in[0,\varepsilon] and u()PWC([0,τ],m)u(\cdot)\in\operatorname{PWC}([0,\tau],{\mathbb{R}}^{m}) such that ΦHuτϕn,K<ε\|\Phi_{{H_{u}}}^{\tau}-\phi_{n}\|_{\ell,K}<\varepsilon. Hence,

ΦHuτϕ,KΦHuτϕn,K+ϕnϕ,K<2ε.\|\Phi_{{H_{u}}}^{\tau}-\phi\|_{\ell,K}\leq\|\Phi_{{H_{u}}}^{\tau}-\phi_{n}\|_{\ell,K}+\|\phi_{n}-\phi\|_{\ell,K}<2\varepsilon.

Definition 20.
  • A smooth function f𝒞(TM,)f\in{\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}) is said to be STAR (small-time approximately reachable) if esfst¯e^{s\overrightarrow{f}}\in\overline{\mathcal{R}_{\rm st}} for all ss\in{\mathbb{R}}.

  • A smooth function f𝒞(TM,)f\in{\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}) is said to be STAR at the level of densities if ρ0esfRst¯(ρ0)\rho_{0}\circ e^{s\overrightarrow{f}}\in\overline{{R}_{\rm st}}(\rho_{0}) for every ρ0Lr(TM)\rho_{0}\in L^{r}(T^{*}M) and every ss\in{\mathbb{R}}.

Remark 21.

If a function f𝒞(TM,)f\in{\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}) 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 𝒞(TM,){\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}).

Proof.

The fact that if ff is STAR and ss\in{\mathbb{R}} then sfsf is STAR is obvious, and the fact that f+gf+g is STAR if ff and gg are STAR is a direct consequence of Propositions 16 and 19.

Let us assume that f,gf,g are STAR and prove that {f,g}\left\{f,g\right\} is STAR. According to Proposition 19,

eτfe1τgeτfst¯τ>0.e^{\tau\overrightarrow{f}}e^{\frac{1}{\tau}\overrightarrow{g}}e^{-\tau\overrightarrow{f}}\in\overline{\mathcal{R}_{\rm st}}\qquad\forall\>\tau>0.

According to Proposition 18,

eτfe1τgeτf\displaystyle e^{\tau\overrightarrow{f}}e^{\frac{1}{\tau}\overrightarrow{g}}e^{-\tau\overrightarrow{f}} =exp(1τ(eτf)g)=exp(1τg+{f,g}+w(τ)),\displaystyle=\exp(\frac{1}{\tau}({e^{\tau\overrightarrow{f}}})_{*}\overrightarrow{g})=\exp(\frac{1}{\tau}\overrightarrow{g}+\overrightarrow{\left\{f,g\right\}}+\overrightarrow{w(\tau)}),

where, for every \ell\in{\mathbb{N}} and KTMK\subset T^{*}M compact, limτ0w(τ),K=0\lim_{\tau\to 0}\|\overrightarrow{w(\tau)}\|_{\ell,K}=0. Since g{g} is STAR, applying Proposition 16 we have

e{f,g}+w(τ)st¯τ>0.e^{\overrightarrow{\left\{f,g\right\}}+\overrightarrow{w(\tau)}}\in\overline{\mathcal{R}_{\rm st}}\qquad\forall\>\tau>0.

Finally, e{f,g}+w(τ)e{f,g}e^{\overrightarrow{\left\{f,g\right\}}+\overrightarrow{w(\tau)}}\to e^{\overrightarrow{\left\{f,g\right\}}} as τ0\tau\rightarrow 0 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), V1,,VmV_{1},\dots,V_{m} are STAR.

Proof.

Let i\llbracket1,m\rrbracketi\in\llbracket 1,m\rrbracket and ss\in{\mathbb{R}}. Consider the constant control u=(u1,,um)u=(u_{1},\dots,u_{m}) such that ui=sτu_{i}=\frac{s}{\tau} for τ>0\tau>0 and uj=0u_{j}=0 for jij\neq i. Then ΦHuτ\Phi_{H_{u}}^{\tau} is reachable in time τ\tau and, according to Proposition 15,

ΦHuτ=eτH0+sViτ0esVi.\Phi_{H_{u}}^{\tau}=e^{\tau\overrightarrow{H_{0}}+s\overrightarrow{V_{i}}}\underset{\tau\rightarrow 0}{\longrightarrow}e^{s\overrightarrow{V_{i}}}.

4.2 Some properties of Rst¯(ρ0)\overline{R_{\rm st}}(\rho_{0})

Lemma 24.

Let ϕDiff(TM)\phi\in\operatorname{Diff}(T^{*}M) be such that |detJϕ|1|\det J_{\phi}|\equiv 1, where JϕJ_{\phi} denotes the Jacobian matrix of ϕ\phi. If ρϕRst¯(ρ){\rho}\circ\phi\in\overline{R_{\rm st}}({\rho}) for every ρ𝒞c(TM){\rho}\in\mathcal{C}_{c}^{\infty}(T^{*}M), then ρϕRst¯(ρ){\rho}\circ\phi\in\overline{R_{\rm st}}({\rho}) for every ρLr(TM){\rho}\in L^{r}(T^{*}M).

Proof.

Let ε>0\varepsilon>0 and ρLr(TM){\rho}\in L^{r}(T^{*}M). There exist ρ0𝒞c(TM){\rho}_{0}\in\mathcal{C}^{\infty}_{c}(T^{*}M), τ[0,ε]\tau\in[0,\varepsilon], and u()PWC([0,τ],m)u(\cdot)\in\operatorname{PWC}([0,\tau],{\mathbb{R}}^{m}) such that ρρ0Lr<ε\|{\rho}-{\rho}_{0}\|_{L^{r}}<\varepsilon and ρ0ΦHuτρ0ϕLr<ε.\|{\rho}_{0}\circ\Phi_{H_{u}}^{\tau}-{\rho}_{0}\circ\phi\|_{L^{r}}<\varepsilon. Then

ρΦHuτρϕLr\displaystyle\|{\rho}\circ\Phi_{H_{u}}^{\tau}-{\rho}\circ\phi\|_{L^{r}} (ρρ0)ΦHuτLr+ρ0ΦHuτρ0ϕLr+(ρρ0)ϕLr\displaystyle\leq\|({\rho}-{\rho}_{0})\circ\Phi_{H_{u}}^{\tau}\|_{L^{r}}+\|{\rho}_{0}\circ\Phi_{H_{u}}^{\tau}-{\rho}_{0}\circ\phi\|_{L^{r}}+\|({\rho}-{\rho}_{0})\circ\phi\|_{L^{r}}
=2ρρ0Lr+ρ0ΦHuτρ0ϕLr<3ε.\displaystyle=2\|{\rho}-{\rho}_{0}\|_{L^{r}}+\|{\rho}_{0}\circ\Phi_{H_{u}}^{\tau}-{\rho}_{0}\circ\phi\|_{L^{r}}<3\varepsilon.

Lemma 25.
  • Let ϕ,ψDiff(TM)\phi,\psi\in\operatorname{Diff}(T^{*}M) be such that |detJϕ|1|detJψ||\det J_{\phi}|\equiv 1\equiv|\det J_{\psi}|. If ρϕRst¯(ρ){\rho}\circ\phi\in\overline{R_{\rm st}}({\rho}) and ρψRst¯(ρ){\rho}\circ\psi\in\overline{R_{\rm st}}({\rho}) for every ρLr(TM){\rho}\in L^{r}(T^{*}M), then ρϕψRst¯(ρ){\rho}\circ\phi\circ\psi\in\overline{R_{\rm st}}({\rho}) for every ρLr(TM){\rho}\in L^{r}(T^{*}M).

  • If (ρn)nRst¯(ρ)(\rho_{n})_{n}\subset\overline{R_{\rm st}}(\rho) and ρnρ\rho_{n}\to\rho_{\infty} in Lr(TM)L^{r}(T^{*}M), then ρRst¯(ρ)\rho_{\infty}\in\overline{R_{\rm st}}(\rho).

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 DHam(TM){\rm DHam}(T^{*}M) of the Hamilton equation (4) implies the small-time approximate controllability in Lr(TM,)L^{r}(T^{*}M,{\mathbb{R}}) of the Liouville equation (2) (cf. Lemma 6).

Proof of Lemma 6.

According to Lemma 24, it is sufficient to prove the result for ρ0𝒞c(TM){\rho}_{0}\in\mathcal{C}^{\infty}_{c}(T^{*}M). Denote the support of ρ0{\rho}_{0} by KK. For xKx\in K and nn\in{\mathbb{N}}^{*} denote the sphere of center Ψ1(x)\Psi^{-1}(x) and radius 1n\frac{1}{n} by S(Ψ1(x),1n)S(\Psi^{-1}(x),\frac{1}{n}). The distance between xx and the image of the previous sphere by Ψ\Psi is strictly positive, i.e. , d(x,Ψ(S(Ψ1(x),1/n)))>0d(x,\Psi(S(\Psi^{-1}(x),{1}/{n})))>0. By compactness, the minimum of the previous distance as xx varies in KK has to be positive:

δn:=minxKd(x,Ψ(S(Ψ1(x),1n)))>0.\delta_{n}:=\min_{x\in K}d\left(x,\Psi\left(S\left(\Psi^{-1}(x),\frac{1}{n}\right)\right)\right)>0.

Moreover δn0\delta_{n}\to 0 as n+n\to+\infty. Notice that if xKx\in K and yTMy\in T^{*}M satisfy d(x,y)<δnd(x,y)<\delta_{n} then

d(Ψ1(x),Ψ1(y))<1n.d(\Psi^{-1}(x),\Psi^{-1}(y))<\frac{1}{n}.

Fix a compact neighborhood K~\tilde{K} of KK. Since Ψst¯\Psi\in\overline{\mathcal{R}_{\rm st}}, for every nn\in{\mathbb{N}} there exist τn[0,1/n]\tau_{n}\in[0,1/n] and un()PWC([0,τn],m)u_{n}(\cdot)\in\operatorname{PWC}([0,\tau_{n}],{\mathbb{R}}^{m}) such that

ΦHunτnΨ0,K~<δn.\|\Phi_{H_{u_{n}}}^{\tau_{n}}-\Psi\|_{0,\tilde{K}}<\delta_{n}.

Then, in particular, d((ΦHunτn)1(x),Ψ1(x))<1/nd((\Phi_{H_{u_{n}}}^{\tau_{n}})^{-1}(x),\Psi^{-1}(x))<1/n for all xKx\in K, which implies that (ΦHunτn)1(K)(\Phi_{H_{u_{n}}}^{\tau_{n}})^{-1}(K) is contained in K~\tilde{K} for nn large enough.

TM|ρ0ΦHunτnρ0Ψ|r\displaystyle\int_{T^{*}M}|{\rho}_{0}\circ\Phi_{H_{u_{n}}}^{\tau_{n}}-{\rho}_{0}\circ\Psi|^{r} =K~|ρ0ΦHunτnρ0Ψ|rDρ00,KrK~d(ΦHunτn(x),Ψ(x))r𝑑x\displaystyle=\int_{\tilde{K}}|{\rho}_{0}\circ\Phi_{H_{u_{n}}}^{\tau_{n}}-{\rho}_{0}\circ\Psi|^{r}\leq\|D{\rho}_{0}\|_{0,K}^{r}\int_{\tilde{K}}d(\Phi_{H_{u_{n}}}^{\tau_{n}}(x),\Psi(x))^{r}dx
Dρ00,KrVol(K~)δnr.\displaystyle\leq\|D{\rho}_{0}\|_{0,K}^{r}\operatorname{Vol}(\tilde{K})\delta_{n}^{r}.

Remark 26.

In the proof of Lemma 6 we actually proved that, if a sequence ϕn\phi_{n} converges to ϕ\phi in Diff(TM){\rm Diff}(T^{*}M) for the compact-open topology and ρ0Lr(TM)\rho_{0}\in L^{r}(T^{*}M), r[1,)r\in[1,\infty), then ρ0ϕnρ0ϕ\rho_{0}\circ\phi_{n}\to\rho_{0}\circ\phi in Lr(TM)L^{r}(T^{*}M).

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 𝒞(TM,){\mathcal{C}^{\infty}}(T^{*}M,{\mathbb{R}}).

5 Vertical and horizontal shears

We introduce two abelian subgroups of DHam(TM){\rm DHam}(T^{*}M) which shall play a key role.

Definition 28 (Vertical and horizontal shears on M=dM={\mathbb{R}}^{d} or 𝕋d{\mathbb{T}}^{d}).
  • For f𝒞(qd,)f\in{\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{q},{\mathbb{R}}) or 𝒞(𝕋qd,){\mathcal{C}^{\infty}}({\mathbb{T}}^{d}_{q},{\mathbb{R}}), the Hamiltonian diffeomorphism

    ef(q,p)=(q,pqf(q)),e^{\overrightarrow{f}}(q,p)=(q,p-\nabla_{q}f(q)),

    is called a vertical shear. Vertical shears form an abelian subgroup of DHam(TM){\rm DHam}(T^{*}M), denoted by 𝒱\mathcal{V}.

  • For g𝒞(pd,)g\in{\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{p},{\mathbb{R}}), the Hamiltonian diffeomorphism

    eg(q,p)=(q+pg(p),p),e^{\overrightarrow{g}}(q,p)=(q+\nabla_{p}g(p),p),

    is called a horizontal shear. Horizontal shears form an abelian subgroup of DHam(TM){\rm DHam}(T^{*}M), denoted by 𝒯\mathcal{T}.

Berger and Turaev recently proved the following useful density property.

Theorem 29.

([10, Corollary 1.1]) Let M=dM={\mathbb{R}}^{d} or 𝕋d{\mathbb{T}}^{d} and ΦDHam(TM)\Phi\in\operatorname{DHam}(T^{*}M). Then, for every ε>0\varepsilon>0, \ell\in{\mathbb{N}}, and KTMK\subset T^{*}M compact there exist NN\in{\mathbb{N}} and S1,,SN𝒱𝒯S_{1},\dots,S_{N}\in\mathcal{V}\cup\mathcal{T} such that

ΦS1SN,K<ε.\|\Phi-S_{1}\cdots S_{N}\|_{\ell,K}<\varepsilon.

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 𝒱st¯\mathcal{V}\subset\overline{\mathcal{R}_{\rm st}} and 𝒯st¯\mathcal{T}\subset\overline{\mathcal{R}_{\rm st}} then st¯=DHam(TM)\overline{\mathcal{R}_{\rm st}}={\rm DHam}(T^{*}M).

6 Proof of Theorem 7

6.1 Vertical shears on TdT^{*}{\mathbb{R}}^{d}

We prove that vertical shears are small-time approximately reachable for system (4), (7).

Theorem 31.

System (4), (7) satisfies 𝒱st¯\mathcal{V}\subset\overline{\mathcal{R}_{\rm st}}.

Proof.

First step: Let us prove that the Hamiltonians p1,,pdp_{1},\dots,p_{d} are STAR. The functions q1,,qdq_{1},\dots,q_{d} are STAR according to Proposition 23. Applying (14), we have

evτqieτH0evτqi=exp(τH0(evτqi)),\displaystyle e^{-\frac{v}{\tau}\overrightarrow{q_{i}}}e^{\tau\overrightarrow{H_{0}}}e^{\frac{v}{\tau}\overrightarrow{q_{i}}}=\exp(\tau\overrightarrow{H_{0}(e^{-\frac{v}{\tau}\overrightarrow{q_{i}}})}),

and τH0(evτqi(q,p))=τ(|p|2/2+V0(q))+vpi+v2/(2τ)\tau H_{0}(e^{-\frac{v}{\tau}\overrightarrow{q_{i}}}(q,p))=\tau(|p|^{2}/2+V_{0}(q))+vp_{i}+v^{2}/(2\tau). So τH0evτpi=τH0+vpi\tau\overrightarrow{H_{0}\circ e^{-\frac{v}{\tau}\overrightarrow{p_{i}}}}=\tau\overrightarrow{H_{0}}+v\overrightarrow{p_{i}}, because the constant function has null contribution to the Hamiltonian vector field. By taking the limit as τ0\tau\rightarrow 0 we get that evpie^{v\overrightarrow{p_{i}}} is small-time approximately reachable for any vv\in{\mathbb{R}}.

Second step: Let us show that the functions q1f,,qdf\partial_{q_{1}}f,\dots,\partial_{q_{d}}f are STAR if ff is STAR. Applying (14), we have that for every vv\in{\mathbb{R}}, τ0\tau\neq 0, and i\llbracket1,d\rrbracketi\in\llbracket 1,d\rrbracket the diffeomorphism

evτfeτpievτf\displaystyle e^{-\frac{v}{\tau}\overrightarrow{f}}e^{\tau\overrightarrow{p_{i}}}e^{\frac{v}{\tau}\overrightarrow{f}} =exp(τpi+vqif)\displaystyle=\exp(\tau\overrightarrow{p_{i}}+v\overrightarrow{\partial_{q_{i}}f})

is small-time approximately reachable. By taking the limit as τ0\tau\to 0 we obtain the desired property.

Third step: Let us show that every f𝒞(qd,)f\in{\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{q},{\mathbb{R}}) is STAR. Notice that it is enough to consider ff with compact support, since the restriction of the flow esfe^{s\overrightarrow{f}} on a given compact KdK\subset{\mathbb{R}}^{d} coincides with esg|Ke^{s\overrightarrow{g}}|_{K}, where g𝒞(qd,)g\in{\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{q},{\mathbb{R}}) coincides with ff on a compact set K~K\tilde{K}\supset K and has compact support. In particular ff can be taken in Hs(d)H^{s}({\mathbb{R}}^{d}) for every s0s\geq 0. The set of linear combinations of Hermite functions is dense in Hs(d)H^{s}({\mathbb{R}}^{d}) for any s0s\geq 0, and hence approximates ff in 𝒞{\mathcal{C}^{\infty}} 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 (j)j(\mathcal{H}_{j})_{j\in{\mathbb{N}}} in 𝒞(qd,){\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{q},{\mathbb{R}}) by 0=Span{qe|q|2/2}\mathcal{H}_{0}=\operatorname{Span}_{{\mathbb{R}}}\left\{q\mapsto e^{-|q|^{2}/2}\right\} and, for every jj\in{\mathbb{N}}^{*},

j:=Span{f0+k=1dqkfkf0,,fdj1}.\mathcal{H}_{j}:=\operatorname{Span}_{{\mathbb{R}}}\left\{f_{0}+\sum_{k=1}^{d}\partial_{q_{k}}f_{k}\mid f_{0},\dots,f_{d}\in\mathcal{H}_{j-1}\right\}.

Thanks to the second step, Proposition 23, and the fact that linear combinations of STAR functions is STAR (cf. Proposition 22), any f:=jjf\in\mathcal{H}_{\infty}:=\cup_{j\in{\mathbb{N}}}\mathcal{H}_{j} is STAR. Recall that the Hermite functions of one variable (ψn)n(\psi_{n})_{n\in{\mathbb{N}}} satisfy the recurrence relations

ψ0(x)=ex22,ψ0=12ψ1,ψn=n2ψn1n+12ψn+1,n1.\psi_{0}(x)=e^{-\frac{x^{2}}{2}},\qquad\psi_{0}^{\prime}=\sqrt{\frac{1}{2}}\psi_{1},\qquad\psi_{n}^{\prime}=\sqrt{\frac{n}{2}}\psi_{n-1}-\sqrt{\frac{n+1}{2}}\psi_{n+1},\quad n\geq 1.

Since each Hermite function in d{\mathbb{R}}^{d} can be written as qψj1(q1)ψjd(qd)q\mapsto\psi_{j_{1}}(q_{1})\cdots\psi_{j_{d}}(q_{d}) with j1,,jdj_{1},\dots,j_{d}\in{\mathbb{N}}, we conclude that \mathcal{H}_{\infty} contains all Hermite functions. ∎

6.2 Quadratic Hamiltonians on TdT^{*}{\mathbb{R}}^{d}

This section contains some preliminary results about an auxiliary quadratic Hamiltonian of the form

Hu(q,p)=|p|22+u|q|22,u.H_{u}(q,p)=\frac{|p|^{2}}{2}+u\frac{|q|^{2}}{2},\qquad u\in{\mathbb{R}}. (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.

Proposition 32.

The function |p|2|p|^{2} is STAR for system (4), (15).

Proof.

First recall that the function |q|2|q|^{2} is STAR according to Proposition 23. Since the diffeomorphism eτ|p|22e^{\tau\frac{\overrightarrow{|p|^{2}}}{2}} is reachable in time τ>0\tau>0 with control u0u\equiv 0, it follows that, for every vv\in{\mathbb{R}},

ev2τ|q|2eτ|p|22ev2τ|q|2e^{\frac{v}{2\tau}\overrightarrow{|q|^{2}}}e^{\tau\frac{\overrightarrow{|p|^{2}}}{2}}e^{-\frac{v}{2\tau}\overrightarrow{|q|^{2}}}

is approximately reachable in time τ\tau. Applying (14), we have that

ev2τ|q|2eτ|p|22ev2τ|q|2\displaystyle e^{\frac{v}{2\tau}\overrightarrow{|q|^{2}}}e^{\tau\frac{\overrightarrow{|p|^{2}}}{2}}e^{-\frac{v}{2\tau}\overrightarrow{|q|^{2}}} =exp(τ|p|22vpq+v22τ|q|2).\displaystyle=\exp(\tau\frac{\overrightarrow{|p|^{2}}}{2}-v\overrightarrow{p\cdot q}+\frac{v^{2}}{2\tau}\overrightarrow{|q|^{2}}).

The function |q|2|q|^{2} being STAR, Proposition 16 implies that eτ|p|22vpqe^{\tau\frac{\overrightarrow{|p|^{2}}}{2}-v\overrightarrow{p\cdot q}} is approximately reachable in time τ\tau. Taking the limit as τ0\tau\rightarrow 0, we get that the function pqp\cdot q is STAR.

As a consequence, for every s>0s>0, the dilation

Ds(q,p):=eln(s)pq(q,p)=(sq,1sp)D_{s}(q,p):=e^{\ln(s)\overrightarrow{p\cdot q}}(q,p)=\left(sq,\frac{1}{s}p\right)

is small-time approximately reachable for every s>0s>0. For v0v\geq 0 and τ>0\tau>0, eτv|p|22e^{\tau v\frac{\overrightarrow{|p|^{2}}}{2}} is approximately reachable in time τv\tau v. Thus, the element

D1τeτv|p|22DτD_{\frac{1}{\sqrt{\tau}}}e^{\tau v\frac{\overrightarrow{|p|^{2}}}{2}}D_{\sqrt{\tau}}

is approximately reachable in time τv\tau v. We have

eτv|p|22Dτ(q,p)=(τq+τvp,pτ),e^{\tau v\frac{\overrightarrow{|p|^{2}}}{2}}D_{\sqrt{\tau}}(q,p)=\left(\sqrt{\tau}q+\sqrt{\tau}vp,\frac{p}{\sqrt{\tau}}\right),

and finally

(D1τeτv|p|22Dτ)(q,p)\displaystyle\left(D_{\frac{1}{\sqrt{\tau}}}e^{\tau v\frac{\overrightarrow{|p|^{2}}}{2}}D_{\sqrt{\tau}}\right)(q,p) =(q+vp,p)=ev|p|22(q,p).\displaystyle=(q+vp,p)=e^{v\frac{\overrightarrow{|p|^{2}}}{2}}(q,p).

So for every τ>0\tau>0,

D1τeτv|p|22Dτ=ev|p|22.D_{\frac{1}{\sqrt{\tau}}}e^{\tau v\frac{\overrightarrow{|p|^{2}}}{2}}D_{\sqrt{\tau}}=e^{v\frac{\overrightarrow{|p|^{2}}}{2}}.

Taking τ\tau arbitrarily small, we deduce that ev|p|22e^{v\frac{\overrightarrow{|p|^{2}}}{2}} is small-time approximately reachable for every constant v0v\geq 0. Applying Proposition 16, the diffeomorphism ev(|p|22+|q|22)e^{v(\frac{\overrightarrow{|p|^{2}}}{2}+\frac{\overrightarrow{|q|^{2}}}{2})} is small-time approximately reachable for every v0v\geq 0. Notice that the latter element is periodic, hence for every w<0w<0, there exists v0v\geq 0 such that ew(|p|22+|q|22)=ev(|p|22+|q|22)e^{w(\frac{\overrightarrow{|p|^{2}}}{2}+\frac{\overrightarrow{|q|^{2}}}{2})}=e^{v(\frac{\overrightarrow{|p|^{2}}}{2}+\frac{\overrightarrow{|q|^{2}}}{2})}. Thus the Hamiltonian |p|2+|q|2|p|^{2}+|q|^{2} is STAR. Since |q|2|q|^{2} is also STAR, we deduce from Proposition 22 that |p|2|p|^{2} is STAR. ∎

6.3 Horizontal shears on TdT^{*}{\mathbb{R}}^{d}

We prove that horizontal shears are small-time approximately reachable for system (4), (7).

Theorem 33.

System (4), (7) satisfies 𝒯st¯\mathcal{T}\subset\overline{\mathcal{R}_{\rm st}}.

Proof.

According to Proposition 32, |p|2|p|^{2} is STAR. Hence, combining Theorem 31 and Proposition 22 we have that ad|p|2/2kf{\rm ad}^{k}_{|p|^{2}/2}f is STAR for any f𝒞(q,)f\in{\mathcal{C}^{\infty}}({\mathbb{R}}_{q},{\mathbb{R}}), kk\in{\mathbb{N}}.

Let P(p)=p1m1pdmdP(p)=p_{1}^{m_{1}}\dots p_{d}^{m_{d}} be a monomial. Setting m=m1++mdm=m_{1}+\dots+m_{d} and

f(q)=1m!q1m1qdmd,f(q)=\frac{1}{m!}q_{1}^{m_{1}}\dots q_{d}^{m_{d}},

we have that ad|p|22mf=P\operatorname{ad}^{m}_{\frac{|p|^{2}}{2}}f=P, showing that PP is STAR. By density of polynomials in 𝒞(pd,){\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{p},{\mathbb{R}}), the proof is concluded. ∎

By virtue of Theorems 31, 33 and Corollary 30, the proof of Theorem 7 is concluded.

7 Proof of Theorem 9

7.1 Vertical shears on T𝕋dT^{*}{\mathbb{T}}^{d}

As in the case of d{\mathbb{R}}^{d}, we now prove that vertical shears are small-time approximately reachable for system (4), (8).

Theorem 34.

System (4), (8) satisfies 𝒱st¯\mathcal{V}\subset\overline{\mathcal{R}_{\rm st}}.

Proof.

We first claim that, if f𝒞(𝕋qd,)f\in{\mathcal{C}^{\infty}}({\mathbb{T}}^{d}_{q},{\mathbb{R}}) is STAR, then eu|qf|2st¯e^{u\overrightarrow{|\nabla_{q}f|^{2}}}\in\overline{\mathcal{R}_{\rm st}} for every u0u\geq 0. Applying (14) we have that the diffeomorphism

euτfeτH0euτf\displaystyle e^{\frac{\sqrt{u}}{\sqrt{\tau}}\overrightarrow{f}}e^{\tau\overrightarrow{H_{0}}}e^{-\frac{\sqrt{u}}{\sqrt{\tau}}\overrightarrow{f}} =exp(τH02τupqf+u|qf|2)\displaystyle=\exp(\tau\overrightarrow{H_{0}}-2\sqrt{\tau u}\overrightarrow{p\cdot\nabla_{q}f}+u|\nabla_{q}f|^{2})

is small-time approximately reachable in time τ>0\tau>0. By letting τ0\tau\to 0, the claim is proved.

We define an increasing sequence of vector spaces (j)j(\mathcal{H}_{j})_{j\in{\mathbb{N}}} by setting

0=Span{1,cos(k1),sin(k1),,cos(kd),sin(kd)k1,,kd as in (9)}𝒞(𝕋qd,)\mathcal{H}_{0}=\operatorname{Span}_{\mathbb{R}}\left\{1,\cos(k_{1}\cdot),\sin(k_{1}\cdot),\dots,\cos(k_{d}\cdot),\sin(k_{d}\cdot)\mid k_{1},\dots,k_{d}\text{ as in }\eqref{eq:frequencies}\right\}\subset{\mathcal{C}^{\infty}}({\mathbb{T}}^{d}_{q},{\mathbb{R}})

and, by induction, letting j\mathcal{H}_{j}, jj\in{\mathbb{N}}^{*}, be the largest vector space whose elements can be written as

φ0+k=1N|φk|2,with N,φ0,,φNj1.\varphi_{0}+\sum_{k=1}^{N}|\nabla\varphi_{k}|^{2},\qquad\text{with }N\in{\mathbb{N}},\ \varphi_{0},\dots,\varphi_{N}\in\mathcal{H}_{j-1}.

Let :=jj\mathcal{H}_{\infty}:=\cup_{j\in{\mathbb{N}}}\mathcal{H}_{j}. Thanks to the claim and Propositions 22, 23, any ff\in\mathcal{H}_{\infty} is STAR. Moreover, the proof of [13, Proposition 2.6] shows that \mathcal{H}_{\infty} contains all trigonometric polynomials. In particular, \mathcal{H}_{\infty} is dense in 𝒞(𝕋qd,){\mathcal{C}^{\infty}}({\mathbb{T}}^{d}_{q},{\mathbb{R}}), 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 T𝕋dT^{*}{\mathbb{T}}^{d} 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 SS be the symmetry defined by

S(q,p)=(q,p).S(q,p)=(q,-p).

Then ρSRst¯(ρ){\rho}\circ S\in\overline{R_{\rm st}}({\rho}) for every ρLr(T𝕋d){\rho}\in L^{r}(T^{*}{\mathbb{T}}^{d}).

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 τ>0\tau>0, the density ρSeτ|p|22S{\rho}\circ S\circ e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}}\circ S is approximately reachable in time τ\tau from ρ{\rho}. Thanks to the relation

Seτ|p|22S=eτ|p|22,S\circ e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}}\circ S=e^{-\tau\overrightarrow{\frac{|p|^{2}}{2}}}, (16)

we thus get that ρeτ|p|22{\rho}\circ e^{-\tau\overrightarrow{\frac{|p|^{2}}{2}}} is approximately reachable in time τ\tau from ρ{\rho}. 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 T𝕋dT^{*}{\mathbb{T}}^{d}.

Proof of Lemma 35.

According to Lemma 24, it is sufficient to prove the result for ρ𝒞c(TM){\rho}\in\mathcal{C}^{\infty}_{c}(T^{*}M). As in the proof of Theorem 3, we introduce a cubic mesh Mh={mnn}M_{h}=\left\{m_{n}\mid n\in{\mathbb{N}}\right\} of TMT^{*}M of size hh. Recall that, by definition, TM=nC¯(mn,h)T^{*}M=\cup_{n\in{\mathbb{N}}}\overline{C}(m_{n},h) and nC(mn,h)\cup_{n\in{\mathbb{N}}}C(m_{n},h) has zero-measure complement in TMT^{*}M. Let KTMK\subset T^{*}M be a compact set containing the support of ρ{\rho}. Given ε>0\varepsilon>0, since ρ{\rho} is uniformly continuous over KK, for hh small enough there exist NN\in{\mathbb{N}} and η(0,h)\eta\in(0,h) such that

ρn=0Nρ(mn)𝟙C(mn,hη)Lr<ε.\left\|{\rho}-\sum_{n=0}^{N}{\rho}(m_{n}){\mathbb{1}}_{C(m_{n},h-\eta)}\right\|_{L^{r}}<\varepsilon.

For every volume-preserving change of variables ϕ:TMTM\phi:T^{*}M\to T^{*}M (including ϕ=S\phi=S), we have

(ρn=NNρ(mn)𝟙C(mn,hη))ϕLr<ε.\left\|\left({\rho}-\sum_{n=N}^{N}{\rho}(m_{n}){\mathbb{1}}_{C(m_{n},h-\eta)}\right)\circ\phi\right\|_{L^{r}}<\varepsilon.

So it is sufficient to find ϕDHam(TM)\phi\in\operatorname{DHam}(T^{*}M) that is small-time approximately reachable and such that 𝟙C(mn,hη)ϕ=𝟙C(mn,hη)S{\mathbb{1}}_{C(m_{n},h-\eta)}\circ\phi={\mathbb{1}}_{C(m_{n},h-\eta)}\circ S for every n\llbracket0,N\rrbracketn\in\llbracket 0,N\rrbracket. The image by SS of a cube of center m=(q,p)m=(q,p) is a cube of center (q,p)(q,-p). We will emulate the action of SS, which is the symmetry with respect to the space {p=0}\left\{p=0\right\}, by a rotation of center (q,0)(q,0) and angle π2\frac{\pi}{2} on each plane (qi,pi)(q_{i},p_{i}). The two transformations differ pointwise, but their images of a cube of center mm coincide, see Figure 6.

[Uncaptioned image]
Figure 6: Emulation of SS by a Hamiltonian rotation

As in the proof of Theorem 3, we organize the N+1N+1 cubes of radius hηh-\eta into columns 𝒞j\mathcal{C}_{j}, jJj\in J, such that the centers of all cubes in 𝒞j\mathcal{C}_{j} have the same qq-component qjq^{j}. We consider jJj\in J, w>0w>0 (to be fixed later) and f𝒞(𝕋qd,)f\in{\mathcal{C}^{\infty}}({\mathbb{T}}^{d}_{q},{\mathbb{R}}) such that

f(q)={|qqj|22w2if |qqj|<h,0if |qqj|>h+η.f(q)=\begin{cases}\frac{|q-q^{j}|^{2}}{2w^{2}}&\mbox{if }|q-q^{j}|<h,\\ 0&\mbox{if }|q-q^{j}|>h+{\eta}.\end{cases}

In particular, if (q,p)(q,p) belongs to a cube in 𝒞j\mathcal{C}_{j^{\prime}} with jJ{j}j^{\prime}\in J\setminus\{j\}, then f(q)=0f(q)=0.

According to Proposition 16 and Theorem 34, ρeT(|p|22+f){\rho}\circ e^{T\overrightarrow{(\frac{|p|^{2}}{2}+f)}}, T>0T>0, is approximately reachable in time TT.

Integrating the Hamiltonian vector field |p|22+|qqj|22w2\frac{|p|^{2}}{2}+\frac{|q-q^{j}|^{2}}{2w^{2}} on Ωj={(q,p)V|qqj|<h}\Omega^{j}=\{(q,p)\in V\mid|q-q^{j}|<h\}, it turns out that for every (q,p)(q,p) such that |qqj|2+|p|2w2<h2|q-q^{j}|^{2}+\frac{|p|^{2}}{w^{2}}<h^{2} and every T>0T>0, eT(|p|22+f)(q,p)Ωje^{T\overrightarrow{(\frac{|p|^{2}}{2}+f)}}(q,p)\in\Omega^{j}, with eπw(|p|22+f)(qj+q~,p)=(qjq~,p)e^{\pi w\overrightarrow{(\frac{|p|^{2}}{2}+f})}(q^{j}+\tilde{q},p)=(q^{j}-\tilde{q},-p). Choosing ww small enough, the ellipsoid

={(q,p)|qqj|2+|p|2w2<h2}\mathcal{E}=\left\{(q,p)\mid|q-q^{j}|^{2}+\frac{|p|^{2}}{w^{2}}<h^{2}\right\}

contains all cubes in 𝒞j\mathcal{C}_{j}, see Figure 7. Set φj=eπw(|p|22+f)\varphi_{j}=e^{\pi w\overrightarrow{(\frac{|p|^{2}}{2}+f})}.

[Uncaptioned image]
Figure 7: Permutation of two cubes in 𝒞j\mathcal{C}_{j}

After time πw\pi w, every cube in 𝒞j\mathcal{C}_{j} is sent by φj\varphi_{j} to its image through SS, while φj\varphi_{j} preserves all cubes in 𝒞j\mathcal{C}_{j^{\prime}} for jJ{j}j^{\prime}\in J\setminus\{j\}. Moreover, φj\varphi_{j} is approximately reachable in time πw\pi w, which is arbitrarily small if ww is.

The proof is concluded by considering as Hamiltonian diffeomorphism ϕ\phi emulating the action of SS on the cubes C(mn,hη)C(m_{n},h-\eta), n\llbracket0,N\rrbracketn\in\llbracket 0,N\rrbracket, the composition of all φj\varphi_{j} for jJj\in J. ∎

7.3 Horizontal shears on T𝕋dT^{*}{\mathbb{T}}^{d} at the level of densities

Theorem 36.

System (2), (8), satisfies, for any ρ0Lr(T𝕋d)\rho_{0}\in L^{r}(T^{*}{\mathbb{T}}^{d}),

{ρ0ϕϕ𝒯}Rst¯(ρ0).\left\{\rho_{0}\circ\phi\mid\phi\in\mathcal{T}\right\}\subset\overline{R_{\rm st}}(\rho_{0}).
Proof.

First step. Let us show that, if f𝒞(T𝕋d,)f\in{\mathcal{C}^{\infty}}(T^{*}{\mathbb{T}}^{d},{\mathbb{R}}) is STAR at the level of densities, then the same is true for {|p|2,f}\left\{|p|^{2},f\right\}. According to Lemma 35 and relation (16), for every τ>0\tau>0 the density

ρ0eτ|p|22e1τfeτ|p|22\rho_{0}\circ e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}}e^{\frac{1}{\tau}\overrightarrow{f}}e^{-\tau\overrightarrow{\frac{|p|^{2}}{2}}}

is approximately reachable in time 2τ2\tau from ρ0\rho_{0}. Moreover, by Proposition 18, given KTMK\subset T^{*}M compact and \ell\in{\mathbb{N}},

eτ|p|22e1τfeτ|p|22=exp(1τf+{|p|22,f}+w(τ)),with w(τ),K=O(τ).e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}}e^{\frac{1}{\tau}\overrightarrow{f}}e^{-\tau\overrightarrow{\frac{|p|^{2}}{2}}}=\exp(\frac{1}{\tau}\overrightarrow{f}+\overrightarrow{\left\{\frac{|p|^{2}}{2},f\right\}}+\overrightarrow{w(\tau)}),\qquad\mbox{with }\|\overrightarrow{w(\tau)}\|_{\ell,K}=O(\tau).

Since ff is STAR, we deduce that ρ0e{|p|22,f(q)}+w(τ)\rho_{0}\circ e^{\overrightarrow{\left\{\frac{|p|^{2}}{2},f(q)\right\}}+\overrightarrow{w(\tau)}} is approximately reachable in time 2τ2\tau from ρ0\rho_{0}. By letting τ0\tau\to 0, we obtain the desired property.

Second step. Consider j\llbracket1,d\rrbracketj\in\llbracket 1,d\rrbracket and kk\in{\mathbb{N}}^{*} and let us show that the function pjkp_{j}^{k} is STAR at the level of densities. According to Theorem 34, the functions cos(qj)\cos(q_{j}) and sin(qj)\sin(q_{j}) are STAR. Thus, by the previous step, for every \ell\in{\mathbb{N}} the following functions are STAR at the level of densities:

ad|p|222cos(qj)=(1)pj2cos(qj),ad|p|222+1cos(qj)=(1)pj2+1sin(qj),\operatorname{ad}_{\frac{|p|^{2}}{2}}^{2\ell}\cos(q_{j})=(-1)^{\ell}p_{j}^{2\ell}\cos(q_{j}),\qquad\operatorname{ad}_{\frac{|p|^{2}}{2}}^{2\ell+1}\cos(q_{j})=(-1)^{\ell}p_{j}^{2\ell+1}\sin(q_{j}),
ad|p|222sin(qj)=(1)pj2sin(qj),ad|p|222+1sin(qj)=(1)+1pj2+1cos(qj).\operatorname{ad}_{\frac{|p|^{2}}{2}}^{2\ell}\sin(q_{j})=(-1)^{\ell}p_{j}^{2\ell}\sin(q_{j}),\qquad\operatorname{ad}_{\frac{|p|^{2}}{2}}^{2\ell+1}\sin(q_{j})=(-1)^{\ell+1}p_{j}^{2\ell+1}\cos(q_{j}).

In particular, writing k=2k=2\ell or k=2+1k=2\ell+1 depending on the parity of kk, pjkcos(qj)p_{j}^{k}\cos(q_{j}) and pjksin(qj)p_{j}^{k}\sin(q_{j}) are STAR at the level of densities. Thus, by Proposition 27, the Hamiltonian functions

1k+1{pjk+1cos(qj),sin(qj)}=pjkcos2(qj),1k+1{pjk+1sin(qj),cos(qj)}=pjksin2(qj),\frac{1}{k+1}\left\{p_{j}^{k+1}\cos(q_{j}),\sin(q_{j})\right\}=p_{j}^{k}\cos^{2}(q_{j}),\qquad\frac{1}{k+1}\left\{p_{j}^{k+1}\sin(q_{j}),-\cos(q_{j})\right\}=p_{j}^{k}\sin^{2}(q_{j}),

are STAR at the level of the densities as well. By taking their sum, pjkp_{j}^{k} is STAR at the level of densities.

Third step. Let us show now that every monomial p1m1pdmd,m1,,md,p_{1}^{m_{1}}\dots p_{d}^{m_{d}},m_{1},\dots,m_{d}\in{\mathbb{N}}, is STAR at the level of densities. We show that p1mp2kp_{1}^{m}p_{2}^{k} 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

1(m+2)(m+1){{p1m+2,sin(q1)cos(q2)},sin(q1)}=p1mcos2(q1)cos(q2)\frac{1}{(m+2)(m+1)}\left\{\left\{p_{1}^{m+2},\sin(q_{1})\cos(q_{2})\right\},\sin(q_{1})\right\}=p_{1}^{m}\cos^{2}(q_{1})\cos(q_{2})

is STAR at the level of densities, and similarly one gets that the same is true for p1msin2(q1)cos(q2)p_{1}^{m}\sin^{2}(q_{1})\cos(q_{2}). Taking a linear combination of the two functions, p1mcos(q2)p_{1}^{m}\cos(q_{2}) is STAR at the level of densities. Similarly, p1msin(q2)p_{1}^{m}\sin(q_{2}) is also STAR at the level of densities. Then,

1(k+2)(k+1){{p2k+2,p1msin(q2)},sin(q2)}=p1mp2kcos2(q2),\frac{1}{(k+2)(k+1)}\left\{\left\{p_{2}^{k+2},p_{1}^{m}\sin(q_{2})\right\},\sin(q_{2})\right\}=p_{1}^{m}p_{2}^{k}\cos^{2}(q_{2}),

is STAR at the level of densities and similarly one proves the same for p1mp2ksin2(q2)p_{1}^{m}p_{2}^{k}\sin^{2}(q_{2}). Finally, p1mp2kp_{1}^{m}p_{2}^{k} is STAR at the level of densities, concluding the proof of the third step.

The conclusion follows from the density of the polynomials in 𝒞(pd,){\mathcal{C}^{\infty}}({\mathbb{R}}^{d}_{p},{\mathbb{R}}). ∎

8 Small-time exact controllability of finite ensembles of points in TdT^{*}{\mathbb{R}}^{d} and T𝕋dT^{*}{\mathbb{T}}^{d}

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, MM denotes either d{\mathbb{R}}^{d} or 𝕋d{\mathbb{T}}^{d} and V=TMV=T^{*}M. Given any NN\in{\mathbb{N}}, let ΔNVN\Delta^{N}\subset V^{N} be the set of NN-uples (γ1,,γN)(\gamma_{1},\dots,\gamma_{N}) with (at least) two coinciding components γi=γj\gamma_{i}=\gamma_{j} for some iji\neq j. The space V(N):=VNΔNV^{(N)}:=V^{N}\setminus\Delta^{N} has a structure of a smooth manifold. For each γV(N)\gamma\in V^{(N)} the tangent space TγV(N)T_{\gamma}V^{(N)} is isomorphic to Tγ1V××TγNV.T_{\gamma_{1}}V\times\dots\times T_{\gamma_{N}}V. First we consider a lift of the controlled systems with controlled Hamiltonian (7) and (8) defined respectively on (Td)(N)(T^{*}{\mathbb{R}}^{d})^{(N)} and (T𝕋d)(N)(T^{*}{\mathbb{T}}^{d})^{(N)}. These systems are of the form (4). The lift on V(N)V^{(N)} is then defined by

γ˙=u(t)(γ),γ=(γ1,,γN)V(N),\dot{\gamma}=\overrightarrow{\mathcal{H}_{u(t)}}(\gamma),\qquad\gamma=(\gamma_{1},\dots,\gamma_{N})\in V^{(N)}, (17)

where u=j=1NHu(γj)\mathcal{H}_{u}=\sum_{j=1}^{N}H_{u}(\gamma_{j}), the Hamiltonian HuH_{u} is defined in (1), and u(γ)=(Hu(γ1),,Hu(γN)).\overrightarrow{\mathcal{H}_{u}}(\gamma)=(\overrightarrow{H_{u}}(\gamma_{1}),\dots,\overrightarrow{H_{u}}(\gamma_{N})). We then define the following family of Hamiltonians on V(N)V^{(N)}

={f(N):(q1,,qN)j=1Nf(qj)f𝒞(Mq,)}.\mathcal{F}=\left\{f^{(N)}:(q^{1},\dots,q^{N})\mapsto\sum_{j=1}^{N}f(q^{j})\mid f\in{\mathcal{C}^{\infty}}(M_{q},{\mathbb{R}})\right\}.

As a direct consequence of Theorems 31 and 34, we obtain the following Lie extension property.

Proposition 37.

Consider the system

γ˙=X(γ),X,γV(N).\dot{\gamma}=\overrightarrow{X}(\gamma),\qquad X\in\mathcal{F},\ \gamma\in V^{(N)}. (18)

For every τ>0\tau>0, the reachable set of system (18) is contained in the small-time approximately reachable set of system (17).

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 V(N)V^{(N)}, i.e. , for any NN distinct initial configurations (q1,p1),,(qN,pN)V(q^{1},p^{1}),\dots,(q^{N},p^{N})\in V, and NN distinct final configurations (q¯1,p¯1),,(q¯N,p¯N)V(\bar{q}^{1},\bar{p}^{1}),\dots,(\bar{q}^{N},\bar{p}^{N})\in V, and ε>0\varepsilon>0, then there exist τ[0,ε]\tau\in[0,\varepsilon] and u()PWC([0,τ],m)u(\cdot)\in\operatorname{PWC}([0,\tau],{\mathbb{R}}^{m}) such that

ΦHuτ(qi,pi)(q¯i,p¯i)<ε\|\Phi_{H_{u}}^{\tau}(q^{i},p^{i})-(\bar{q}^{i},\bar{p}^{i})\|<\varepsilon

for every i\llbracket1,N\rrbracketi\in\llbracket 1,N\rrbracket.

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 NN distinct positions q1,,qNMq^{1},\dots,q^{N}\in M and vectors a1,,aNda^{1},\dots,a^{N}\in\mathbb{R}^{d}, there exists a smooth function f𝒞(M,)f\in\mathcal{C}^{\infty}(M,{\mathbb{R}}) such that qf(qj)=aj\frac{\partial}{\partial q}f(q^{j})=a^{j} for every j\llbracket1,N\rrbracketj\in\llbracket 1,N\rrbracket.

Proof of Proposition 38.

For the case of M=dM={\mathbb{R}}^{d}, the result directly follows from the small-time approximated controllability in the group DHam(Td)\operatorname{DHam}(T^{*}{\mathbb{R}}^{d}).

Let us consider the case M=𝕋dM={\mathbb{T}}^{d}. The pairs of initial configurations (qi,pi),i\llbracket1,N\rrbracket,(q^{i},p^{i}),i\in\llbracket 1,N\rrbracket, are distinct so, for every δ>0\delta>0, there exists τ[0,δ)\tau\in[0,\delta) such that the points q1+τp1,,qN+τpNq^{1}+\tau p^{1},\dots,q^{N}+\tau p^{N} are pairwise distinct on the torus. For every τ>0\tau>0, the diffeomorphism eτ|p|22e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}} is approximately reachable in time τ\tau and eτ|p|22(qi,pi)=(qi+τpi,pi)e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}}(q^{i},p^{i})=(q^{i}+\tau p^{i},p^{i}) for i\llbracket1,N\rrbracketi\in\llbracket 1,N\rrbracket. So we can assume without loss of generality that the initial positions q1,,qNq^{1},\dots,q^{N} are distinct. Similarly, we can assume without loss of generality that the final positions q¯1,,q¯N\bar{q}^{1},\dots,\bar{q}^{N} are distinct, up to replace them by q¯iτp¯i\bar{q}^{i}-\tau\bar{p}^{i} with τ>0\tau>0 arbitrarily small.

For each i\llbracket1,N\rrbracketi\in\llbracket 1,N\rrbracket, let p^id\hat{p}^{i}\in{\mathbb{R}}^{d} be such that q¯iqi=p^i\bar{q}^{i}-q^{i}=\hat{p}^{i} modulo 2π2\pi. Applying Lemma 39, there exists a function fτ𝒞(𝕋d,)f_{\tau}\in{\mathcal{C}^{\infty}}({\mathbb{T}}^{d},{\mathbb{R}}) such that qfτ(qi)=pi1τp^i\frac{\partial}{\partial q}f_{\tau}(q^{i})=p^{i}-\frac{1}{\tau}\hat{p}^{i} for every i\llbracket1,N\rrbracketi\in\llbracket 1,N\rrbracket. As a consequence, efτ(qi,pi)=(qi,piqfτ(qi))e^{\overrightarrow{f_{\tau}}}(q^{i},p^{i})=(q^{i},p^{i}-\frac{\partial}{\partial q}f_{\tau}(q^{i})) and then eτ|p|22efτ(qi,pi)=(qi+p^i,1τpi^)=(q¯i,1τp^i)e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}}e^{\overrightarrow{f_{\tau}}}(q^{i},p^{i})=(q^{i}+\hat{p}^{i},\frac{1}{\tau}\hat{p^{i}})=(\bar{q}^{i},\frac{1}{\tau}\hat{p}^{i}). Applying again Lemma 39, there exists a smooth function gτ𝒞(𝕋d,)g_{\tau}\in{\mathcal{C}^{\infty}}({\mathbb{T}}^{d},{\mathbb{R}}) such that qgτ(q¯i)=p¯i+1τp^i\frac{\partial}{\partial q}g_{\tau}(\bar{q}^{i})=-\bar{p}^{i}+\frac{1}{\tau}\hat{p}^{i} for every i\llbracket1,N\rrbracketi\in\llbracket 1,N\rrbracket. Then egτ(q¯i,1τp^i)=(q¯i,p¯i)e^{\overrightarrow{g_{\tau}}}(\bar{q}^{i},\frac{1}{\tau}\hat{p}^{i})=(\bar{q}^{i},\bar{p}^{i}). Finally, thanks to Theorem 34, the diffeomorphism egτeτ|p|22efτe^{\overrightarrow{g_{\tau}}}e^{\tau\overrightarrow{\frac{|p|^{2}}{2}}}e^{\overrightarrow{f_{\tau}}} is approximately reachable in time τ\tau for every τ>0\tau>0. ∎

Theorem 40.

System (17) is small-time exactly controllable in V(N)V^{(N)}.

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 γ=(q1,p1,,qN,pN)V(N)\gamma=(q^{1},p^{1},\dots,q^{N},p^{N})\in V^{(N)}. Assume for now that the positions q1,,qNq^{1},\dots,q^{N} are distinct in MM. Let v1,,vdv^{1},\dots,v^{d} be a basis of d{\mathbb{R}}^{d}. Similarly to what is done for Lemma 39, by the Whitney extension theorem there exist f1,,fd𝒞(M,)f^{1},\dots,f^{d}\in{\mathcal{C}^{\infty}}(M,{\mathbb{R}}) such that

qfi(q1)=vi,qfi(qj)=0d,Hessqfi(q1)=0d×d,Hessqfi(qj)=0d×d,-\nabla_{q}f^{i}(q^{1})=v^{i},\quad\nabla_{q}f^{i}(q^{j})=0_{d},\quad{\rm Hess}_{q}f^{i}(q^{1})=0_{d\times d},\quad{\rm Hess}_{q}f^{i}(q^{j})=0_{d\times d}, (19)

for i\llbracket1,d\rrbracketi\in\llbracket 1,d\rrbracket and j\llbracket2,N\rrbracketj\in\llbracket 2,N\rrbracket. Thanks to (19), we have (fi)(N)(γ)=(0d,vi,0d,,0d){\overrightarrow{(f^{i})^{(N)}}}(\gamma)=(0_{d},v_{i},0_{d},\dots,0_{d}). Moreover, using again (19), for i\llbracket1,d\rrbracketi\in\llbracket 1,d\rrbracket we compute

[(fi)(N),0](γ)=(j=1Npjqjfi)(γ)=(vi,0d,,0d),\Bigl[{\overrightarrow{(f^{i})^{(N)}}},\overrightarrow{\mathcal{H}_{0}}\Bigr](\gamma)=\overrightarrow{(\sum_{j=1}^{N}-p^{j}\nabla_{q^{j}}f^{i})}(\gamma)=(v_{i},0_{d},\dots,0_{d}),

where 0(γ)=j=1NH0(γj)\mathcal{H}_{0}(\gamma)=\sum_{j=1}^{N}H_{0}(\gamma_{j}) is the lifted drift. Replacing q1q^{1} by q2,,qNq^{2},\dots,q^{N} in the above argument, we can generate 2dN2dN vectors that form a basis of TγV(N)T_{\gamma}V^{(N)}.

Since the Lie algebra Lie{Huum}\operatorname{Lie}\left\{{H_{u}}\mid u\in{\mathbb{R}}^{m}\right\} is dense in \mathcal{F} for the compact-open topology (cf. Theorems 31 and 34), it follows by continuity that system (17) satisfies the Lie algebra rank condition at γ\gamma.

We are left to consider the case where q1,,qNq^{1},\dots,q^{N} are not pairwise distinct. Since γVNΔN\gamma\in V^{N}\setminus\Delta^{N}, the pairs (q1,p1),,(qN,pN)(q^{1},p^{1}),\dots,(q^{N},p^{N}) are distinct in VV. Then there exists δ>0\delta>0 such that the positions q1+δp1,,qN+δpNq^{1}+\delta p^{1},\dots,q^{N}+\delta p^{N} are pairwise distinct in MM. Then γ=(q1+δp1,p1,,qN+δpN,pN)\gamma^{\prime}=(q^{1}+\delta p^{1},p^{1},\dots,q^{N}+\delta p^{N},p^{N}) is approximately reachable from γ\gamma and (17) satisfies the Lie algebra rank condition in a neighborhood of γ\gamma^{\prime}. Since each Hamiltonian HuH_{u}, umu\in{\mathbb{R}}^{m}, is analytic, it follows that the Lie algebra rank condition is satisfied at every point of the orbit through γ\gamma^{\prime} for system (17), and in particular at γ\gamma. ∎

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, LpL^{p} 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).