This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Some Remarks on Controllability of the Liouville Equation

Maxim Raginsky
[email protected]
University of Illinois, Urbana, IL 61801, USA
Abstract

We revisit the work of Roger Brockett on controllability of the Liouville equation, with a particular focus on the following problem: Given a smooth controlled dynamical system of the form x˙=f(x,u)\dot{x}=f(x,u) and a state-space diffeomorphism ψ\psi, design a feedback control u(t,x)u(t,x) to steer an arbitrary initial state x0x_{0} to ψ(x0)\psi(x_{0}) in finite time. This formulation of the problem makes contact with the theory of optimal transportation and with nonlinear controllability. For controllable linear systems, Brockett showed that this is possible under a fairly restrictive condition on ψ\psi. We prove that controllability suffices for a much larger class of diffeomorphisms. For nonlinear systems defined on smooth manifolds, we review a recent result of Agrachev and Caponigro regarding controllability on the group of diffeomorphisms. A corollary of this result states that, for control-affine systems satisfying a bracket generating condition, any ψ\psi in a neighborhood of the identity can be implemented using a time-varying feedback control law that switches between finitely many time-invariant flows. We prove a quantitative version which allows us to describe the implementation complexity of the Agrachev–Caponigro construction in terms of a lower bound on the number of switchings.

In a series of papers [Bro97, Bro07, Bro12], Roger Brockett drew attention to a class of problems involving a smooth controlled dynamical system x˙=f(x,u)\dot{x}=f(x,u), where the focus is not on the evolution of the system state x(t)x(t) per se, but rather on the evolution of its probability density ρ(t,)\rho(t,\cdot) starting from a given initial density ρ(0,)=ρ0()\rho(0,\cdot)=\rho_{0}(\cdot). This shift of perspective leads to questions pertaining to control of the so-called Liouville equation [Mac92], i.e., the first-order partial differential equation

ρ(t,x)t=div(ρ(t,x)f(x,u)),\displaystyle\frac{\partial\rho(t,x)}{\partial t}=-{\rm div}\big{(}\rho(t,x)f(x,u)\big{)}, (1)

also referred to as the transport or the continuity equation. For example, we may be interested in the questions of controllability, where two densities ρ0\rho_{0} and ρ1\rho_{1} are given and the objective is to determine whether they can be joined by a curve lying along the trajectory of (1) for some choice of the control u()u(\cdot). One can also formulate optimal control problems in this setting, e.g., minimizing a finite-horizon performance index of the form

J(ρ0;u())=0TML(x,u,t)ρ(t,x)dxdt+Mr(x)ρ(T,x)dx,\displaystyle J(\rho_{0};u(\cdot))=\int^{T}_{0}\int_{M}L(x,u,t)\rho(t,x)\operatorname{d\!}x\operatorname{d\!}t+\int_{M}r(x)\rho(T,x)\operatorname{d\!}x,

over appropriately chosen controls u()u(\cdot), where the initial density ρ(0,)=ρ0()\rho(0,\cdot)=\rho_{0}(\cdot) is given and the integration is over both the (finite) time interval [0,T][0,T] and the state space MM. As pointed out by Brockett, because of the presence of the divergence operator on the right-hand side of (1), there will in general be a nontrivial difference between what can be achieved with open-loop controls u(t)u(t) versus closed-loop (or feedback) controls u(t,x)u(t,x). See, e.g., [Pog16, BBFR19] for further developments of these ideas.

In this paper, we will focus on the questions of controllability of (1). For the most part, we will stick with the Euclidean setting M=nM={\mathbb{R}}^{n} and assume that the target density ρ1\rho_{1} can be expressed as the pushforward of the initial density ρ0\rho_{0} by an orientation-preserving diffeomorphism ψ:nn\psi\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}^{n}, i.e.,

nh(x)ρ1(x)dx=nh(ψ(x))ρ0(x)dx\displaystyle\int_{{\mathbb{R}}^{n}}h(x)\rho_{1}(x)\operatorname{d\!}x=\int_{{\mathbb{R}}^{n}}h(\psi(x))\rho_{0}(x)\operatorname{d\!}x

for all bounded, continuous functions h:nh\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}. In this case, the two densities are related via the Monge–Ampère equation

detDψ(x)=ρ0(x)ρ1(ψ(x)),\displaystyle\det D\psi(x)=\frac{\rho_{0}(x)}{\rho_{1}(\psi(x))},

where DψD\psi denotes the Jacobian of ψ\psi. It is important to note that not any pair of densities ρ0,ρ1\rho_{0},\rho_{1} can be joined in this way; for example, it will not be possible if ρ0\rho_{0} vanishes on some open set while ρ1\rho_{1} does not. Nevertheless, this restricted setting already exposes several nontrivial aspects of the problem of controlling the Liouville equation, and an added benefit is that in this instance we can also phrase everything in terms of the original system x˙=f(x,u)\dot{x}=f(x,u) and ask whether there exists a feedback control law u(t,x)u(t,x) that transfers the state of the system from x(0)=x0x(0)=x_{0} to x(1)=ψ(x0)x(1)=\psi(x_{0}), for every x0x_{0}. This, in turn, makes contact with the problem of implementing orientation-preserving diffeomorphisms using a given controlled system [AC09, Cap11]. On the other hand, the question of existence of a diffeomorphism ψ\psi that pushes ρ0\rho_{0} forward to ρ1\rho_{1} makes contact with the theory of optimal transportation [Vil03]. This aspect had not been explored by Brockett, apart from a brief discussion of the classical theorem of Moser [Mos65] on the transitivity of the action of orientation-preserving diffeomorphisms of a compact differentiable manifold MM on the space of everywhere positive probability densities on MM.

The remainder of the paper is structured as follows. In Section 1, we take a quick look at controllability in the space of densities through the lens of optimal transportation. We then consider the case of controllable linear systems in Section 2, where we present an extension of a controllability result from [Bro07] and relate it to optimal transportation. Moving on to nonlinear systems, in Section 3 we present a quantitative version of a structural result of Agrachev and Caponigro [AC09] on controllability on the group of diffeomorphisms of a smooth compact manifold. Some concluding remarks are presented in Section 4.

1 The relation to optimal transportation

Let ρ0\rho_{0} and ρ1\rho_{1} be two probability densities on n{\mathbb{R}}^{n}, which we will assume to be everywhere positive to keep things simple. Then a theorem of Brenier [Vil03, Thm. 2.12] guarantees the existence of a strictly convex function h:nh\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}, such that its gradient ψ=h\psi=\nabla h pushes ρ0\rho_{0} forward to ρ1\rho_{1}. Moreover, this mapping is optimal in the sense that

n|xh(x)|2ρ0(x)dx=min{n|xψ(x)|2ρ0(x)dx:ρ1=ψρ0},\displaystyle\int_{{\mathbb{R}}^{n}}|x-\nabla h(x)|^{2}\rho_{0}(x)\operatorname{d\!}x=\min\left\{\int_{{\mathbb{R}}^{n}}|x-\psi(x)|^{2}\rho_{0}(x)\operatorname{d\!}x\mathrel{\mathop{\mathchar 58\relax}}\rho_{1}=\psi_{*}\rho_{0}\right\},

where the minimum is over all ψ:nn\psi\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} that push ρ0\rho_{0} forward to ρ1\rho_{1}. There is also a complementary dynamic formulation due to Benamou and Brenier [Vil03, Sec. 8.1]; in control-theoretic language, it guarantees the existence of a smooth feedback control law u:[0,1]×nnu\mathrel{\mathop{\mathchar 58\relax}}[0,1]\times{\mathbb{R}}^{n}\to{\mathbb{R}}^{n}, such that the trajectory of the Liouville equation

ρ(t,x)t=div(ρ(t,x)u(t,x))\displaystyle\frac{\partial\rho(t,x)}{\partial t}=-{\rm div}\big{(}\rho(t,x)u(t,x)\big{)}

joins ρ(0,)=ρ0()\rho(0,\cdot)=\rho_{0}(\cdot) to ρ(1,)=ρ1()\rho(1,\cdot)=\rho_{1}(\cdot), and the performance index

J(ρ0;u())=1201n|u(t,x)|2ρ(t,x)dxdt\displaystyle J(\rho_{0};u(\cdot))=\frac{1}{2}\int^{1}_{0}\int_{{\mathbb{R}}^{n}}|u(t,x)|^{2}\rho(t,x)\operatorname{d\!}x\operatorname{d\!}t

is minimized. On the other hand, the corresponding controlled system x˙=u\dot{x}=u allows for maximum “control authority,” in the sense that all directions of motion are available to the controller at each time t[0,1]t\in[0,1]. The more difficult case of a general controlled system x˙=f(x,u)\dot{x}=f(x,u) corresponds to the so-called nonholonomic setting [AL09, KL09], where the key question is how to relate the question of controllability (or obstructions to controllability) in the space of densities to structural properties of the family of vector fields {f(,u)}uU\{f(\cdot,u)\}_{u\in U} indexed by the elements of the control set UU.

2 The case of a controllable linear system

The case of a linear system

x˙=Ax+Bu\displaystyle\dot{x}=Ax+Bu (2)

with nn-dimensional state xx and mm-dimensional input uu has been considered in [Bro07] under the assumption that the system is controllable, i.e., the columns of B,AB,,An1BB,AB,\dots,A^{n-1}B span n{\mathbb{R}}^{n}. Let an orientation-preserving diffeomorphism ψ:nn\psi\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} be given. By controllability, we can steer (2) from any fixed initial state x(0)=x0x(0)=x_{0} to x(T)=ψ(x0)x(T)=\psi(x_{0}) using the open-loop control

ux0(t)=B𝖳eA𝖳tW(0,T)1(eATψ(x0)x0),\displaystyle u_{x_{0}}(t)=B^{\mathsf{T}}e^{-A^{\mathsf{T}}t}W(0,T)^{-1}\big{(}e^{-AT}\psi(x_{0})-x_{0}\big{)},

where

W(0,t):=0teAτBB𝖳eA𝖳τdτ\displaystyle W(0,t)\mathrel{\mathop{\mathchar 58\relax}}=\int^{t}_{0}e^{-A\tau}BB^{\mathsf{T}}e^{-A^{\mathsf{T}}\tau}\operatorname{d\!}\tau

is the controllability Gramian of (2), which is positive definite since the system is controllable. The main idea in [Bro07] is to impose appropriate regularity conditions on eATψe^{-AT}\psi to ensure that the map

Kt(x0):=eAt(x0+0teAτBux0(τ)dτ)=eAt(x0+W(0,t)W(0,T)1(eATψ(x0)x0))\displaystyle\begin{split}K_{t}(x_{0})&\mathrel{\mathop{\mathchar 58\relax}}=e^{At}\left(x_{0}+\int^{t}_{0}e^{-A\tau}Bu_{x_{0}}(\tau)\operatorname{d\!}\tau\right)\\ &=e^{At}\left(x_{0}+W(0,t)W(0,T)^{-1}\big{(}e^{-AT}\psi(x_{0})-x_{0}\big{)}\right)\end{split} (3)

is injective for all 0<t<T0<t<T. If this is the case, then the closed-loop (feedback) control

u(t,x)=B𝖳eA𝖳tW(0,T)1(eATψ(Kt1(x))Kt1(x))\displaystyle u(t,x)=B^{\mathsf{T}}e^{-A^{\mathsf{T}}t}W(0,T)^{-1}\big{(}e^{-AT}\psi(K^{-1}_{t}(x))-K^{-1}_{t}(x)\big{)} (4)

can be used to carry out the transfer from x(0)=x0x(0)=x_{0} to x(T)=ψ(x0)x(T)=\psi(x_{0}) for every initial condition x0nx_{0}\in{\mathbb{R}}^{n}.

In [Bro07], Brockett stated that one sufficient condition to ensure the above result is for the map eATψide^{-AT}\psi-{\rm id} to be a contraction (i.e., Lipschitz-continuous with constant strictly smaller than one). However, as pointed out recently by Abdelgalil and Georgiou [AG24], this is not enough. Indeed, the crux of the argument in [Bro07] is that the above contraction condition implies that the map

eAtKt=id+W(0,t)W(0,T)1(eATψid)e^{-At}K_{t}={\rm id}+W(0,t)W(0,T)^{-1}\big{(}e^{-AT}\psi-{\rm id}\big{)}

is invertible for every tt. This, in turn, relies on the claim that, since W(0,T)W(0,t)W(0,T)-W(0,t) is positive definite for every 0<t<T0<t<T by controllability, it follows that W(0,t)W(0,T)11\|W(0,t)W(0,T)^{-1}\|\leq 1. However, the latter claim is not valid: Since the matrix W(0,t)W(0,T)1W(0,t)W(0,T)^{-1} need not be symmetric, we can only guarantee that its spectral radius is bounded above by one. Following the line of argument in [AG24], we can easily show that Brockett’s argument goes through if we modify his contraction condition by replacing ψ\psi with

ψ^(x):=W(0,T)1/2eATψ(W(0,T)1/2x).\displaystyle\hat{\psi}(x)\mathrel{\mathop{\mathchar 58\relax}}=W(0,T)^{-1/2}e^{-AT}\psi\big{(}W(0,T)^{1/2}x\big{)}. (5)

From this, it follows readily that the mapping

K~t:=id+W(0,T)1/2W(0,t)W(0,T)1/2(ψ^id)\displaystyle\tilde{K}_{t}\mathrel{\mathop{\mathchar 58\relax}}={\rm id}+W(0,T)^{-1/2}W(0,t)W(0,T)^{-1/2}\big{(}\hat{\psi}-{\rm id}\big{)}

is invertible since W(0,T)1/2W(0,t)W(0,T)1/21\|W(0,T)^{-1/2}W(0,t)W(0,T)^{-1/2}\|\leq 1 (multiply the matrix inequality W(0,t)W(0,T)W(0,t)\leq W(0,T) on the left and on the right by W(0,T)1/2W(0,T)^{-1/2}). The desired conclusion then follows from the fact that

Kt(x)=eAtW(0,T)1/2K~t(W(0,T)1/2x).\displaystyle K_{t}(x)=e^{At}W(0,T)^{1/2}\tilde{K}_{t}(W(0,T)^{-1/2}x).

The above assumption on ψ\psi is fairly restrictive. For an arbitrary ψ\psi, one could first represent it as a composition ψkψ1\psi_{k}\circ\dots\circ\psi_{1} such that each ψi\psi_{i} satisfies (modified) Brockett’s condition and then apply the above construction for each ii. However, the number kk will, in general, be very large, resulting in controls of very high complexity (we will come back to the issue of complexity later in the broader context of nonlinear systems). It was later shown by Hindawi et al. [HPR11] and Chen et al. [CGP17] that controllability is sufficient for any ψ\psi that can be written as the gradient of a convex function after a certain change of coordinates. These results can also be phrased in the setting of optimal transportation, as the resulting constructions are optimal for quadratic costs of the form L(x,u)=x𝖳Qx+u𝖳RuL(x,u)=x^{\mathsf{T}}Qx+u^{\mathsf{T}}Ru with symmetric positive-semidefinite QQ and symmetric positive-definite RR [HPR11].

Here, we will give a result of the same flavor that preserves the spirit of Brockett’s proof. The key concept we will need is that of a monotone mapping [RW98]: A mapping ψ:nn\psi\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} is monotone if

xy,ψ(x)ψ(y)0,for all x,yn.\displaystyle\langle x-y,\psi(x)-\psi(y)\rangle\geq 0,\qquad\text{for all }x,y\in{\mathbb{R}}^{n}.

Note that this definition makes no assumptions on differentiability of ψ\psi (in fact, it can be extended to set-valued mappings); however, when ψ\psi is differentiable, monotonicity is equivalent to the symmetric part of the Jacobian DψD\psi being everywhere positive-semidefinite. In particular, any ψ\psi given by the gradient of a convex function is monotone.

Theorem 1.

If the system (2) is controllable and if ψ:nn\psi\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} is such that the mapping ψ^\hat{\psi} defined in (5) is monotone, then there exists a feedback control law u(t,x)u(t,x) that steers the state of (2) from x(0)=x0x(0)=x_{0} to x(T)=ψ(x0)x(T)=\psi(x_{0}) for every x0nx_{0}\in{\mathbb{R}}^{n}.

Proof.

We will show that the map KtK_{t} defined in (3) is injective for 0t<T0\leq t<T, thus the feedback control u(t,x)u(t,x) given by (4) will do the job.

To that end, we claim that, for any 0<t<T0<t<T,

x0y0x0y0,W(0,t)1eAt(Kt(x0)Kt(y0))>0,\displaystyle x_{0}\neq y_{0}\,\Rightarrow\,\langle x_{0}-y_{0},W(0,t)^{-1}e^{-At}(K_{t}(x_{0})-K_{t}(y_{0}))\rangle>0, (6)

which will imply that KtK_{t} is injective since the matrices W(0,T)W(0,t)1eAtW(0,T)W(0,t)^{-1}e^{-At} are nonsingular for t>0t>0. Using the definition of ψ^\hat{\psi} in (5), we can write

Kt(x0)\displaystyle K_{t}(x_{0}) =eAt((IW(0,t)W(0,T)1)x0\displaystyle=e^{At}\Big{(}\big{(}I-W(0,t)W(0,T)^{-1}\big{)}x_{0}
+W(0,t)W(0,T)1/2ψ^(W(0,T)1/2x0)).\displaystyle\qquad\qquad+W(0,t)W(0,T)^{-1/2}\hat{\psi}(W(0,T)^{-1/2}x_{0})\Big{)}.

For any x0,y0x_{0},y_{0}, let x^0:=W(0,T)1/2x0\hat{x}_{0}\mathrel{\mathop{\mathchar 58\relax}}=W(0,T)^{-1/2}x_{0} and y^0:=W(0,T)1/2y0\hat{y}_{0}\mathrel{\mathop{\mathchar 58\relax}}=W(0,T)^{-1/2}y_{0}. Then

x0y0,W(0,t)1eAt(Kt(x0)Kt(y0))\displaystyle\langle x_{0}-y_{0},W(0,t)^{-1}e^{-At}(K_{t}(x_{0})-K_{t}(y_{0}))\rangle
=x0y0,(W(0,t)1W(0,T)1)(x0y0)+x^0y^0,ψ^(x^0)ψ^(y^0).\displaystyle=\langle x_{0}-y_{0},(W(0,t)^{-1}-W(0,T)^{-1})(x_{0}-y_{0})\rangle+\langle\hat{x}_{0}-\hat{y}_{0},\hat{\psi}(\hat{x}_{0})-\hat{\psi}(\hat{y}_{0})\rangle.

If x0y0x_{0}\neq y_{0}, then the first term on the right-hand side is strictly positive by controllability, while the second term is nonnegative since ψ^\hat{\psi} is monotone. This proves (6), so that KtK_{t} is indeed injective and u(t,x)u(t,x) in (4) gives the desired feedback control law. ∎

To connect Theorem 1 to the problem of controlling a given initial density ρ0\rho_{0} to a given final density ρ1\rho_{1} using (2), let ρ^0\hat{\rho}_{0} be the pushforward of ρ0\rho_{0} by the invertible linear transformation xW(0,T)1/2xx\mapsto W(0,T)^{1/2}x and, similarly, let ρ^1\hat{\rho}_{1} be the pushforward of ρ1\rho_{1} by the invertible linear transformation xW(0,T)1/2eATxx\mapsto W(0,T)^{-1/2}e^{-AT}x. By Brenier’s theorem, there is a monotone map ψ^\hat{\psi} such that ρ^1=ψ^ρ^0\hat{\rho}_{1}=\hat{\psi}_{*}\hat{\rho}_{0}. Then the map ψ\psi related to ψ^\hat{\psi} via (5) evidently pushes ρ0\rho_{0} forward to ρ1\rho_{1}, and in that case Theorem 1 tells us how to construct the desired feedback control u(t,x)u(t,x). In particular, the Benamou–Brenier dynamic formulation of optimal transportation is a special case corresponding to A=0A=0 and B=IB=I with m=nm=n, cf. [HPR11, CGP17].

3 Controllability on the group of diffeomorphisms

Let us now consider the case of a nonlinear system

x˙=f(x,u)\displaystyle\dot{x}=f(x,u) (7)

whose state space MM is a smooth (say, CC^{\infty}) closed finite-dimensional manifold. A smooth diffeomorphism ψ:MM\psi\mathrel{\mathop{\mathchar 58\relax}}M\to M is given, and the problem is to determine whether there exists a control law that can steer every initial state x(0)=x0Mx(0)=x_{0}\in M to x(T)=ψ(x0)x(T)=\psi(x_{0}), for a fixed finite T>0T>0. By analogy with the linear case, we hope to capitalize on some form of controllability of (7). Here, however, apart from the usual complications arising in the context of nonlinear controllability [HK77], a major difficulty is the lack of explicit expressions for control laws that transfer a given initial state to a given final state. Nevertheless, we can still aim for a structural result of some form.

One such result was obtained by Agrachev and Caponigro [AC09]. We will discuss it shortly in full generality, but for now we mention its corollary for driftless control-affine systems of the form

x˙=i=1muifi(x),\displaystyle\dot{x}=\sum^{m}_{i=1}u_{i}f_{i}(x), (8)

where f1,,fmf_{1},\dots,f_{m} are smooth vector fields on MM and u1,,umu_{1},\dots,u_{m} are real-valued control inputs. Suppose that {f1,,fm}\{f_{1},\dots,f_{m}\} is a bracket-generating family, i.e., for each xMx\in M the set {g(x):gLie(f1,,fm)}\{g(x)\mathrel{\mathop{\mathchar 58\relax}}g\in{\rm Lie}(f_{1},\dots,f_{m})\}, where Lie(f1,,fm){\rm Lie}(f_{1},\dots,f_{m}) is the Lie algebra generated by f1,,fmf_{1},\dots,f_{m}, coincides with the tangent space TxMT_{x}M to MM at xx. Then, for any diffeomorphism ψ:MM\psi\mathrel{\mathop{\mathchar 58\relax}}M\to M isotopic to the identity there exists a time-dependent feedback control law u(t,x)=(u1(t,x),,um(t,x))u(t,x)=(u_{1}(t,x),\dots,u_{m}(t,x)) that transfers x(0)=x0x(0)=x_{0} to x(1)=ψ(x0)x(1)=\psi(x_{0}) for every x0Mx_{0}\in M. (Two diffeomorphisms ψ,ψ:MM\psi,\psi^{\prime}\mathrel{\mathop{\mathchar 58\relax}}M\to M are isotopic if there exists a smooth map H:[0,1]×MMH\mathrel{\mathop{\mathchar 58\relax}}[0,1]\times M\to M, such that H(t,):MMH(t,\cdot)\mathrel{\mathop{\mathchar 58\relax}}M\to M for each t[0,1]t\in[0,1] is a diffeomorphism, H(0,)=ψ()H(0,\cdot)=\psi(\cdot), and H(1,)=ψ()H(1,\cdot)=\psi^{\prime}(\cdot) [Ban97]. We will denote by Diff0(M){\rm Diff}_{0}(M) the family of all diffeomorphisms on MM that are isotopic to the identity.)

The general setting considered in [AC09] is as follows: Let a family \mathcal{F} of smooth vector fields on MM be given. The control group of \mathcal{F}, defined by

𝔾():={etkfket1f1:ti,fi,k},\displaystyle{\mathbb{G}}(\mathcal{F})\mathrel{\mathop{\mathchar 58\relax}}=\left\{e^{t_{k}f_{k}}\circ\dots\circ e^{t_{1}f_{1}}\mathrel{\mathop{\mathchar 58\relax}}t_{i}\in{\mathbb{R}},\,f_{i}\in\mathcal{F},\,k\in{\mathbb{N}}\right\},

is a subgroup of the group Diff(M){\rm Diff}(M) of smooth diffeomorphisms of MM. Suppose that 𝔾(){\mathbb{G}}(\mathcal{F}) acts transitively on MM, i.e., for any x,yMx,y\in M there exists some ψ𝔾()\psi\in{\mathbb{G}}(\mathcal{F}) such that y=ψ(x)y=\psi(x). Then we have the following [AC09]:

Theorem 2.

There exist a neighborhood 𝒪\mathcal{O} of the identity in Diff0(M){\rm Diff}_{0}(M) and a positive integer kk that depends only on \mathcal{F}, such that every ψ𝒪\psi\in\mathcal{O} can be represented as

ψ=eakfkea1f1\displaystyle\psi=e^{a_{k}f_{k}}\circ\dots\circ e^{a_{1}f_{1}} (9)

for some f1,,fkf_{1},\dots,f_{k}\in\mathcal{F} and some a1,,akC(M)a_{1},\ldots,a_{k}\in C^{\infty}(M).

Note that the vector fields in (9) are rescaled by smooth real-valued functions on MM, which explains the origin of time-varying feedback controls ui(t,x)u_{i}(t,x) in the above discussion of control-affine systems. It is also evident that these controls will be piecewise constant in tt.

With this in mind, let us consider the system (8), where ={f1,,fm}\mathcal{F}=\{f_{1},\dots,f_{m}\} is bracket-generating. Let ψDiff0(M)\psi\in{\rm Diff}_{0}(M) be given. We will apply Theorem 2 to \mathcal{F}—because \mathcal{F} is bracket-generating, 𝔾(){\mathbb{G}}(\mathcal{F}) acts transitively on MM. Following [Cap11], we first establish a fragmentation property of ψ\psi relative to 𝒪\mathcal{O}, i.e., show that there exist some ψ1,,ψN𝒪\psi_{1},\dots,\psi_{N}\in\mathcal{O} such that ψ=ψNψ1\psi=\psi_{N}\circ\dots\circ\psi_{1}. Since ψDiff0(M)\psi\in{\rm Diff}_{0}(M), there exists a smooth path [0,1]tφtDiff0(M)[0,1]\ni t\mapsto\varphi_{t}\in{\rm Diff}_{0}(M) such that φ1=ψ\varphi_{1}=\psi and φ0=id\varphi_{0}={\rm id}. For every NN\in{\mathbb{N}}, the maps

ψi:=φi/Nφ(i1)/N1,i=1,,N\displaystyle\psi_{i}\mathrel{\mathop{\mathchar 58\relax}}=\varphi_{i/N}\circ\varphi_{(i-1)/N}^{-1},\qquad i=1,\dots,N

belong to Diff0(M){\rm Diff}_{0}(M), and ψ=ψNψN1ψ1\psi=\psi_{N}\circ\psi_{N-1}\circ\dots\circ\psi_{1}. We can ensure that each ψi𝒪\psi_{i}\in\mathcal{O} by choosing NN large enough. Applying Theorem 2 to each ψi\psi_{i}, we conclude that there exist K=kNK=kN smooth functions a1,,aKC(M)a_{1},\dots,a_{K}\in C^{\infty}(M), such that ψ\psi can be represented as

ψ=eaKfiKea1fi1,\displaystyle\psi=e^{a_{K}f_{i_{K}}}\circ\dots\circ e^{a_{1}f_{i_{1}}},

for some choice of indices i1,,iK{1,,m}i_{1},\dots,i_{K}\in\{1,\dots,m\}, cf. [Cap11, Prop. 4]. From this representation, it is straightforward to derive mm feedback controls ui:[0,1]×Mu_{i}\mathrel{\mathop{\mathchar 58\relax}}[0,1]\times M\to{\mathbb{R}} that are piecewise constant in tt, with the number of pieces (switchings) equal to KK. Thus, the integer kk in Theorem 2 is a lower bound on the number of switchings, which in turn is a natural measure of implementation complexity of a control law. Since the result of [AC09] has been used as a black-box device in subsequent works [Cap11, EZOP23], it is of interest to provide some quantitative estimates of kk.

To that end, we first make some assumptions on MM and \mathcal{F}. We take MM to be a smooth compact manifold of dimension nn isometrically embedded in d{\mathbb{R}}^{d} for some d>nd>n, and we will equip MM with the ambient metric d(x,y)=|xy|d(x,y)=|x-y|, where |||\cdot| is the Euclidean norm on d{\mathbb{R}}^{d}. We further assume that MM has positive reach τ>0\tau>0, where the reach of a set AdA\subset{\mathbb{R}}^{d} is defined as the largest value of τ\tau, such that any point at a distance 0<r<τ0<r<\tau from AA has a unique nearest point in AA [Fed59]. For example, the unit sphere Sd1S^{d-1} has reach 11. We will assume that the vector fields ff\in\mathcal{F} are uniformly bounded in the C1(M)C^{1}(M) seminorm, in the following sense [AS04, Sec. 2.2]. For r=0,1,r=0,1,\dots, the Cr(M)C^{r}(M) seminorm of a function aC(M)a\in C^{\infty}(M) is defined by

aCr(M):=sup{|Di1Dia(x)|:xM, 1i1,,id, 0r},\displaystyle\|a\|_{C^{r}(M)}\mathrel{\mathop{\mathchar 58\relax}}=\sup\left\{|D_{i_{1}}\dots D_{i_{\ell}}a(x)|\mathrel{\mathop{\mathchar 58\relax}}x\in M,\,1\leq i_{1},\dots,i_{\ell}\leq d,\,0\leq\ell\leq r\right\},

where D1,,DdD_{1},\dots,D_{d} are the orthogonal projections on MM of the standard basis vector fields x1,,xd\frac{\partial}{\partial x_{1}},\dots,\frac{\partial}{\partial x_{d}} on d{\mathbb{R}}^{d}. The Cr(M)C^{r}(M) seminorm of a vector field ff on MM is defined as

fCr(M):=sup{faCr(M):aC(M),aCr+1(M)=1},\displaystyle\|f\|_{C^{r}(M)}\mathrel{\mathop{\mathchar 58\relax}}=\sup\left\{\|fa\|_{C^{r}(M)}\mathrel{\mathop{\mathchar 58\relax}}a\in C^{\infty}(M),\,\|a\|_{C^{r+1}(M)}=1\right\},

where faC(M)fa\in C^{\infty}(M) is the Lie derivative of aa along ff. In terms of these definitions, we will assume that

supffC1(M)<.\displaystyle\sup_{f\in\mathcal{F}}\|f\|_{C^{1}(M)}<\infty.

We then have the following:

Theorem 3.

Let MM and \mathcal{F} satisfy the above assumptions. Then Theorem 2 holds with

kvol(M)vol(Bn(0,1))n216n(1r)n\displaystyle k\geq\frac{{\rm vol}(M)}{{\rm vol}(B_{{\mathbb{R}}^{n}}(0,1))}\frac{n^{2}}{16^{n}}\left(\frac{1}{r}\right)^{n} (10)

for some sufficiently small r(0,τ)r\in(0,\tau), where Bn(0,1)B_{{\mathbb{R}}^{n}}(0,1) is the nn-dimensional Euclidean ball of radius 11 centered at the origin.

Remark 1.

The lower bound on kk in (10) is not optimal, but rather an artifact of the specific construction used in [AC09]. Moreover, as will become evident from the proof of the theorem, the actual number kk will be much larger than the right-hand side of (10). At any rate, the exponential dependence of the lower bound on the dimension of the state space MM should be kept in mind when using the result of [AC09].

3.1 The proof of Theorem 3

We will follow the logic of [AC09], but with some of the steps replaced by more explicit quantitative arguments.

For any subset VnV\subseteq{\mathbb{R}}^{n} containing the origin, we will denote by C0(V,k)C^{\infty}_{0}(V,{\mathbb{R}}^{k}) the space of CC^{\infty} functions F:VkF\mathrel{\mathop{\mathchar 58\relax}}V\to{\mathbb{R}}^{k}, such that F(0)=0F(0)=0. When k=1k=1, we will simply write C0(V)C^{\infty}_{0}(V). Also, Bk(z,r)B_{{\mathbb{R}}^{k}}(z,r) will denote the Euclidean ball of radius rr centered at zkz\in{\mathbb{R}}^{k}.

The first key ingredient in the proof of Theorem 2 is the following:

Proposition 1.

Let X1,,XnX_{1},\dots,X_{n} be vector fields on n{\mathbb{R}}^{n} with XiC1(n)<\|X_{i}\|_{C^{1}({\mathbb{R}}^{n})}<\infty for all ii, such that X1(0),,Xn(0)X_{1}(0),\dots,X_{n}(0) are linearly independent. Then there exist constants r,ε>0r,\varepsilon>0 and an open set

𝒱{FC0(Bn(0,r),n):FidC1<ε},\displaystyle\mathcal{V}\subseteq\left\{F\in C^{\infty}_{0}(B_{{\mathbb{R}}^{n}}(0,r),{\mathbb{R}}^{n})\mathrel{\mathop{\mathchar 58\relax}}\|F-{\rm id}\|_{C^{1}}<\varepsilon\right\},

such that every F𝒱F\in\mathcal{V} can be written as

F=eanXnea1X1|Bn(0,r)\displaystyle F=e^{a_{n}X_{n}}\circ\dots\circ e^{a_{1}X_{1}}\big{|}_{B_{{\mathbb{R}}^{n}}(0,r)}

for some a1,,anC0(Bn(0,r))a_{1},\dots,a_{n}\in C^{\infty}_{0}(B_{{\mathbb{R}}^{n}}(0,r)).

Remark 2.

By multiplying each XiX_{i} by a suitable bump function, the everywhere C1C^{1} boundedness assumption on X1,,XnX_{1},\dots,X_{n} can be relaxed to C1C^{1} boundedness on a ball of sufficiently large radius centered at the origin.

The proof of Proposition 1 makes use of a lemma which states that, for some r>0r>0, we can represent every FF in a certain open neighborhood of the identity in C0(Bn(0,r),n)C^{\infty}_{0}(B_{{\mathbb{R}}^{n}}(0,r),{\mathbb{R}}^{n}) as a composition of nn smooth maps with certain additional properties. We give a quantitative version here; given the lemma, Proposition 1 can be proved exactly as in [AC09].

Lemma 1.

Let X1,,XnX_{1},\dots,X_{n} be vector fields on n{\mathbb{R}}^{n} satisfying the conditions of Proposition 1. Let 𝒰0\mathcal{U}_{0} be a given neighborhood of the identity in C0(n,n)C^{\infty}_{0}({\mathbb{R}}^{n},{\mathbb{R}}^{n}). Then there exist constants r,ε>0r,\varepsilon>0, such that any FC0(Bn(0,r),n)F\in C^{\infty}_{0}(B_{{\mathbb{R}}^{n}}(0,r),{\mathbb{R}}^{n}) with FidC1<ε\|F-{\rm id}\|_{C^{1}}<\varepsilon can be represented as a composition of the form

F=ψnψn1ψ1|Bn(0,r)\displaystyle F=\psi_{n}\circ\psi_{n-1}\circ\dots\circ\psi_{1}\big{|}_{B_{{\mathbb{R}}^{n}}(0,r)}

for some ψ1,,ψn𝒰0\psi_{1},\dots,\psi_{n}\in\mathcal{U}_{0}, where each ψi\psi_{i} preserves the integral curves of XiX_{i}, i.e., for any xBn(0,r)x\in B_{{\mathbb{R}}^{n}}(0,r) there exists some ti=ti(x)t_{i}=t_{i}(x)\in{\mathbb{R}} such that ψi(x)=etiXi(x)\psi_{i}(x)=e^{t_{i}X_{i}}(x).

Proof (of Lemma 1).

We first establish a quantitative inverse function theorem following the ideas in [Chr85, Sec. 8]; one could also use a less direct argument appealing to the Lyusternik–Graves theorem [DR09], with similar estimates. For every xnx\in{\mathbb{R}}^{n}, let 𝑿(x)\boldsymbol{X}(x) be the n×nn\times n matrix with columns Xn(x),Xn1(x),,X1(x)X_{n}(x),X_{n-1}(x),\dots,X_{1}(x). Since X1(0),,Xn(0)X_{1}(0),\dots,X_{n}(0) span n{\mathbb{R}}^{n}, the smallest singular value of 𝑿(0)\boldsymbol{X}(0) is positive:

σn(𝑿(0))=c>0.\displaystyle\sigma_{n}(\boldsymbol{X}(0))=c>0.

The function xσn(𝑿(x))x\mapsto\sigma_{n}(\boldsymbol{X}(x)) is continuous as a consequence of the continuity of XiX_{i}’s and Weyl’s perturbation theorem [Bha97]. Therefore, there exists some r>0r>0, such that

σn(𝑿(x))c/2,xV:=Bn(0,r).\displaystyle\sigma_{n}(\boldsymbol{X}(x))\geq c/2,\qquad x\in V\mathrel{\mathop{\mathchar 58\relax}}=B_{{\mathbb{R}}^{n}}(0,r).

Consider the function G:n×nnG\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\times{\mathbb{R}}^{n}\to{\mathbb{R}}^{n} given by

G(x,y):=eynXney1X1(x).\displaystyle G(x,y)\mathrel{\mathop{\mathchar 58\relax}}=e^{y_{n}X_{n}}\circ\dots\circ e^{y_{1}X_{1}}(x).

Then, denoting by D1GD_{1}G and D2GD_{2}G the partial derivatives of GG w.r.t. xx and yy respectively, we have

D2G(x,0)=𝑿(x).\displaystyle D_{2}G(x,0)=\boldsymbol{X}(x).

Hence, there exists some ε0>0\varepsilon_{0}>0, such that

σn(D2G(x,y¯))c/4,xV,y¯W:=Bn(0,ε0)\displaystyle\sigma_{n}\left(D_{2}G(x,\bar{y})\right)\geq c/4,\qquad x\in V,\,\bar{y}\in W\mathrel{\mathop{\mathchar 58\relax}}=B_{{\mathbb{R}}^{n}}(0,\varepsilon_{0})

so, by the classical inverse function theorem, the map yG(x,y)y\mapsto G(x,y) is invertible on WW for each xVx\in V. We claim that, for any 0εε00\leq\varepsilon\leq\varepsilon_{0}, the image of the ball Bn(0,ε)B_{{\mathbb{R}}^{n}}(0,\varepsilon) under G(x,)G(x,\cdot) contains the ball Bn(x,cε/4)B_{{\mathbb{R}}^{n}}(x,c\varepsilon/4), and thus for any zBn(x,cε/4)z\in B_{{\mathbb{R}}^{n}}(x,c\varepsilon/4) there exists a unique yBn(0,ε)y\in B_{{\mathbb{R}}^{n}}(0,\varepsilon), such that G(x,y)=zG(x,y)=z. To establish the claim, fix a unit vector vnv\in{\mathbb{R}}^{n} and define a vector field YY on WW by D2G(x,y)Y(y)=vD_{2}G(x,y)Y(y)=v for all yWy\in W. Since |Y(y)|4/c|Y(y)|\leq 4/c on WW, the curve ξ(t):=etY(0)\xi(t)\mathrel{\mathop{\mathchar 58\relax}}=e^{tY}(0) is well-defined and remains in Bn(0,ε)B_{{\mathbb{R}}^{n}}(0,\varepsilon) for all 0tcε/40\leq t\leq c\varepsilon/4, as long as εε0\varepsilon\leq\varepsilon_{0}. Moreover, since by the chain rule we have

ddtG(x,ξ(t))=D2G(x,ξ(t))Y(ξ(t))=v,\displaystyle\frac{\operatorname{d\!}}{\operatorname{d\!}t}G(x,\xi(t))=D_{2}G(x,\xi(t))Y(\xi(t))=v,

it follows that G(x,ξ(t))=x+tvBn(x,cε/4)G(x,\xi(t))=x+tv\in B_{{\mathbb{R}}^{n}}(x,c\varepsilon/4) for 0tcε/40\leq t\leq c\varepsilon/4. Since vv was arbitrary, the claim follows.

Now, fix some ε>0\varepsilon>0 to be chosen later and let 𝒰ε\mathcal{U}_{\varepsilon} be the set of all FC0(V,n)F\in C^{\infty}_{0}(V,{\mathbb{R}}^{n}), such that

FidC1<cε/4.\displaystyle\|F-{\rm id}\|_{C^{1}}<c\varepsilon/4. (11)

By the preceding discussion, if εε0\varepsilon\leq\varepsilon_{0}, then for any F𝒰εF\in\mathcal{U}_{\varepsilon} and for any xVx\in V there exists a unique y(x)Bn(0,ε)Wy(x)\in B_{{\mathbb{R}}^{n}}(0,\varepsilon)\subseteq W, such that F(x)=G(x,y(x))F(x)=G(x,y(x)). The chain rule then gives

DF(x)\displaystyle DF(x) =D1G(x,y(x))+D2G(x,y(x))Dy(x),\displaystyle=D_{1}G(x,y(x))+D_{2}G(x,y(x))Dy(x),

and, since D2G(x,y(x))D_{2}G(x,y(x)) is invertible for xVx\in V, we can solve for Dy(x)Dy(x):

Dy(x)=D2G(x,y(x))1(DF(x)D1G(x,y(x))).\displaystyle Dy(x)=D_{2}G(x,y(x))^{-1}\big{(}DF(x)-D_{1}G(x,y(x))\big{)}.

Consider the curve η(t)\eta(t), 0tn0\leq t\leq n, given by

η˙(t)=Z(η(t),t),η(0)=x\displaystyle\dot{\eta}(t)=Z(\eta(t),t),\qquad\eta(0)=x

where, for 0tn0\leq t\leq n,

Z(x,t):=yiXi(x),i1t<i,i=1,,n.\displaystyle Z(x,t)\mathrel{\mathop{\mathchar 58\relax}}=y_{i}X_{i}(x),\qquad i-1\leq t<i,\,i=1,\dots,n.

Then G(x,y)=η(n)G(x,y)=\eta(n), and we can express the derivative D1G(x,y)D_{1}G(x,y) as the t=nt=n solution of the matrix-valued variational equation

Λ˙(t)=Zx(η(t),t)Λ(t),Λ(0)=I.\displaystyle\dot{\Lambda}(t)=\frac{\partial Z}{\partial x}(\eta(t),t)\Lambda(t),\qquad\Lambda(0)=I.

For |y||y| small enough, we will have

D1G(x,y)IC|y|\displaystyle\|D_{1}G(x,y)-I\|\leq C|y|

for some constant CC that depends only on nn and on the C1C^{1} seminorms of the XiX_{i}’s. It follows that, for F𝒰εF\in\mathcal{U}_{\varepsilon} with ε\varepsilon small enough, we will have

supxVDy(x)4c(c4+C)ε=:Cε,\displaystyle\sup_{x\in V}\left\|Dy(x)\right\|\leq\frac{4}{c}\left(\frac{c}{4}+C\right)\varepsilon=\mathrel{\mathop{\mathchar 58\relax}}C^{\prime}\varepsilon,

i.e., the map xy(x)x\mapsto y(x) is CεC^{\prime}\varepsilon-Lipschitz.

Now consider, following [AC09], the maps

Φi(x):=eyi(x)Xiey1(x)X1(x),i=0,,n\displaystyle\Phi_{i}(x)\mathrel{\mathop{\mathchar 58\relax}}=e^{y_{i}(x)X_{i}}\circ\dots\circ e^{y_{1}(x)X_{1}}(x),\qquad i=0,\dots,n

where Φ0=id\Phi_{0}={\rm id} and Φn=F\Phi_{n}=F. Since Φi(x)=G(x,(y1(x),,yi(x),0,,0))\Phi_{i}(x)=G(x,(y_{1}(x),\dots,y_{i}(x),0,\dots,0)), it is a diffeomorphism with domain VV by the preceding discussion, and, in fact, for ε\varepsilon small enough we will have ΦiidC1C′′ε\|\Phi_{i}-{\rm id}\|_{C^{1}}\leq C^{\prime\prime}\varepsilon for some constant C′′>0C^{\prime\prime}>0. Finally, define, for i=1,,ni=1,\dots,n, the maps

ψi(x):=eyi(Φi11(x))Xi(x),\displaystyle\psi_{i}(x)\mathrel{\mathop{\mathchar 58\relax}}=e^{y_{i}(\Phi^{-1}_{i-1}(x))X_{i}}(x),

so that F=ψnψ1F=\psi_{n}\circ\dots\circ\psi_{1}. Each ψi\psi_{i} evidently preserves the integral curves of XiX_{i}. Moreover, from (11) it follows that |yi(Φi11(x))|ε|y_{i}(\Phi^{-1}_{i-1}(x))|\leq\varepsilon for all xVx\in V, and thus we can ensure that each ψi𝒰0\psi_{i}\in\mathcal{U}_{0} by taking ε\varepsilon sufficiently small. ∎

The second key ingredient is a fragmentation lemma for the elements of Diff0(M){\rm Diff}_{0}(M) along the lines of [PS70, Lemma 3.1], see also [Ban97, Lemma 2.1.8]. Again, we give a quantitative version. Recall the definition of the support of a diffeomorphism ψDiff(M)\psi\in{\rm Diff}(M):

supp(ψ):={xM:ψ(x)x}¯.\displaystyle{\rm supp}(\psi)\mathrel{\mathop{\mathchar 58\relax}}=\overline{\left\{x\in M\mathrel{\mathop{\mathchar 58\relax}}\psi(x)\neq x\right\}}.

In other words, the support of ψ\psi is the closure of the set on which ψ\psi differs from the identity [Ban97].

Lemma 2.

Let 𝒪\mathcal{O} be a neighborhood of the identity in Diff(M){\rm Diff}(M). Then, for any r(0,τ)r\in(0,\tau), where τ\tau is the reach of MM, there exist NN points z1,,zNMz_{1},\dots,z_{N}\in M with

vol(M)vol(Bn(0,1))n(116)n(1r)nNvol(M)vol(Bn(0,1))n(π2)n(1r)n,\displaystyle\frac{{\rm vol}(M)}{{\rm vol}(B_{{\mathbb{R}}^{n}}(0,1))}n\left(\frac{1}{16}\right)^{n}\left(\frac{1}{r}\right)^{n}\leq N\leq\frac{{\rm vol}(M)}{{\rm vol}(B_{{\mathbb{R}}^{n}}(0,1))}n\left(\frac{\pi}{2}\right)^{n}\left(\frac{1}{r}\right)^{n}, (12)

such that the sets Ui=MBd(zi,r)U_{i}=M\cap B_{{\mathbb{R}}^{d}}(z_{i},r) cover MM and such that any ψ𝒪Diff0(M)\psi\in\mathcal{O}\cap{\rm Diff}_{0}(M) can be written in the form ψ=ψNψ1\psi=\psi_{N}\circ\dots\circ\psi_{1}, where each ψi\psi_{i} is an element of 𝒪\mathcal{O} with supp(ψi)Ui{\rm supp}(\psi_{i})\subseteq U_{i} for all ii.

Remark 3.

Since the group Diff0(M){\rm Diff}_{0}(M) is path-connected, it is generated by any neighborhood 𝒪\mathcal{O} of the identity. Thus, Diff0(M){\rm Diff}_{0}(M) is generated by the set {ψ𝒪:supp(ψ)Ui for some 1iN}\{\psi\in\mathcal{O}\mathrel{\mathop{\mathchar 58\relax}}{\rm supp}(\psi)\subseteq U_{i}\text{ for some $1\leq i\leq N$}\}.

Proof.

Since MM is compact, for any r>0r>0 it can be covered by finitely many sets of the form Ui=MBd(zi,r)U_{i}=M\cap B_{{\mathbb{R}}^{d}}(z_{i},r) for some z1,,zNMz_{1},\dots,z_{N}\in M. From [BJPR22, Corollary 10], the smallest value of NN for a given r<τr<\tau, i.e., the covering number of MM at resolution rr w.r.t. the ambient metric, can be estimated from above and from below as in (12). Let U1,,UNU_{1},\dots,U_{N} be such a covering of MM. We can now follow the steps in the proof of [AC09, Lemma 5.4] to finish. ∎

After these preparations, we can proceed essentially along the same lines as in [AC09]. Let 𝒫:=𝔾(~)\mathcal{P}\mathrel{\mathop{\mathchar 58\relax}}={\mathbb{G}}(\tilde{\mathcal{F}}), the control group generated by

~:={af:aC(M),f},\displaystyle\tilde{\mathcal{F}}\mathrel{\mathop{\mathchar 58\relax}}=\{af\mathrel{\mathop{\mathchar 58\relax}}a\in C^{\infty}(M),\,f\in\mathcal{F}\},

and define the isotropy subgroup of 𝒫\mathcal{P} at xMx\in M:

𝒫x:={ψ𝒫:ψ(x)=x}.\displaystyle\mathcal{P}_{x}\mathrel{\mathop{\mathchar 58\relax}}=\{\psi\in\mathcal{P}\mathrel{\mathop{\mathchar 58\relax}}\psi(x)=x\}.

For an open set UMU\subseteq M and xUx\in U, we will denote by Cx(U,M)C^{\infty}_{x}(U,M) the family of all smooth maps F:UMF\mathrel{\mathop{\mathchar 58\relax}}U\to M, such that F(x)=xF(x)=x.

Lemma 3.

There exists a constant r(0,τ)r\in(0,\tau) such that, for any xMx\in M and for U:=MBd(x,r)U\mathrel{\mathop{\mathchar 58\relax}}=M\cap B_{{\mathbb{R}}^{d}}(x,r), the set

{ψ|U:ψ𝒫x}\displaystyle\{\psi|_{U}\mathrel{\mathop{\mathchar 58\relax}}\psi\in\mathcal{P}_{x}\} (13)

has nonempty interior in Cx(U,M)C^{\infty}_{x}(U,M) which contains the identity. Moreover, there exist knk\geq n vector fields f1,,fkf_{1},\dots,f_{k} depending on xx, such that any element of the interior of (13) can be represented as a composition of the form

eakfkea1f1\displaystyle e^{a_{k}f_{k}}\circ\dots\circ e^{a_{1}f_{1}} (14)

for some a1,,akC(M)a_{1},\dots,a_{k}\in C^{\infty}(M).

Proof.

We follow the proof of Lemma 5.1 and Corollary 5.2 in [AC09], but with some steps made more explicit. Since 𝔾(){\mathbb{G}}(\mathcal{F}) acts transitively on MM, Sussmann’s orbit theorem [Sus73] implies that the tangent space TxMT_{x}M to MM at xx is equal to the span of the set {ψf(x):ψ𝔾(),f}\{\psi_{*}f(x)\mathrel{\mathop{\mathchar 58\relax}}\psi\in{\mathbb{G}}(\mathcal{F}),f\in\mathcal{F}\}, where ψf\psi_{*}f denotes the action of the tangent map of ψ\psi on ff. Thus, there exist vector fields X1,,XnX_{1},\dots,X_{n} of the form Xi=(ψi)fiX_{i}=(\psi_{i})_{*}f_{i} for some ψi𝔾()\psi_{i}\in{\mathbb{G}}(\mathcal{F}) and fif_{i}\in\mathcal{F}, such that X1(x),,Xn(x)X_{1}(x),\dots,X_{n}(x) span TxMT_{x}M. Moreover, for all smooth functions a1,,anC(M)a_{1},\dots,a_{n}\in C^{\infty}(M) that vanish at xx, the diffeomorphism

eanXnea1X1=ψne(anψn)fnψn1ψ1e(a1ψ1)f1ψ11\displaystyle\begin{split}&e^{a_{n}X_{n}}\circ\dots\circ e^{a_{1}X_{1}}\\ &\qquad=\psi_{n}\circ e^{(a_{n}\circ\psi_{n})f_{n}}\circ\psi^{-1}_{n}\circ\dots\circ\psi_{1}\circ e^{(a_{1}\circ\psi_{1})f_{1}}\circ\psi^{-1}_{1}\end{split} (15)

belongs to the isotropy group 𝒫x\mathcal{P}_{x}. Now, using the techniques of [FMN16], we can show that, for any r>0r>0 small enough, the set U=MBd(x,r)U=M\cap B_{{\mathbb{R}}^{d}}(x,r) can be locally coordinatized using the elements of TxMT_{x}M as

U={x+v+Ψ(v):vBn(0,r)}Bd(x,r)\displaystyle U=\{x+v+\Psi(v)\mathrel{\mathop{\mathchar 58\relax}}v\in B_{{\mathbb{R}}^{n}}(0,r^{\prime})\}\cap B_{{\mathbb{R}}^{d}}(x,r)

for some r>rr^{\prime}>r, where Ψ\Psi is a C2C^{2} map from Bn(0,r)B_{{\mathbb{R}}^{n}}(0,r^{\prime}) into dn{\mathbb{R}}^{d-n}, such that Ψ(0)=0\Psi(0)=0, DΨ(0)=0D\Psi(0)=0, and the Jacobian DΨD\Psi is Cτ\frac{C}{\tau}-Lipschitz, where CC is some constant that depends only on nn. In particular, the map vx+v+Ψ(v)v\mapsto x+v+\Psi(v) is invertible. (In the terminology of [FMN16], the set

{x+v+Ψ(v):vBn(0,r)}\{x+v+\Psi(v)\mathrel{\mathop{\mathchar 58\relax}}v\in B_{{\mathbb{R}}^{n}}(0,r^{\prime})\}

is a patch over TxMT_{x}M of radius rr^{\prime}, centered at xx and tangent to TxMT_{x}M at xx.) Hence, by reparametrizing and rescaling, we end up in the situation of Proposition 1. It then follows that the set (13) has nonempty interior in Cx(U,M)C^{\infty}_{x}(U,M). Moreover, since the vector fields in \mathcal{F} are uniformly bounded in C1C^{1} seminorm, it follows from the proof of Lemma 1 that the constant rr can be chosen independently of xx.

To show that the interior of the set (13) contains the identity map, we let 𝒜\mathcal{A} be an open subset of Cx(U,M)C^{\infty}_{x}(U,M) contained in the interior of (13). Then, for any ψ0\psi_{0} such that ψ0|U𝒜\psi_{0}|_{U}\in\mathcal{A}, the set 𝒪:=ψ01𝒜\mathcal{O}\mathrel{\mathop{\mathchar 58\relax}}=\psi^{-1}_{0}\circ\mathcal{A} is a neighborhood of the identity contained in (13).

Finally, since each ψi\psi_{i} in (15) belongs to the control group 𝔾(){\mathbb{G}}(\mathcal{F}), it can be expressed as

ψi=eti,kifi,kieti,1fi,1\displaystyle\psi_{i}=e^{t_{i,k_{i}}f_{i,k_{i}}}\circ\dots e^{t_{i,1}f_{i,1}}

for some integer ki1k_{i}\geq 1, times ti,1,,ti,kit_{i,1},\dots,t_{i,k_{i}}\in{\mathbb{R}}, and vector fields fi,1,,fi,kif_{i,1},\dots,f_{i,k_{i}}\in\mathcal{F}. Substituting this into (15) and relabeling the functions aa and the vector fields ff as needed, we arrive at the representation (14) for some a1,,akC(M)a_{1},\dots,a_{k}\in C^{\infty}(M) and f1,,fkf_{1},\dots,f_{k}\in\mathcal{F}, where

k=n+2i=1nkin.\displaystyle k=n+2\sum^{n}_{i=1}k_{i}\geq n.

This completes the proof. ∎

Lemma 4.

Let 𝒪\mathcal{O} be a neighborhood of the identity in Diff(M){\rm Diff}(M). Then for any xMx\in M and any open set UMU\subset M containing xx,

xint{ψ(x):ψ𝒪𝒫,supp(ψ)U}.\displaystyle x\in{\rm int}\{\psi(x)\mathrel{\mathop{\mathchar 58\relax}}\psi\in\mathcal{O}\cap\mathcal{P},{\rm supp}(\psi)\subset U\}.
Proof.

See the proof of Lemma 5.3 in [AC09]. ∎

Now, let {Ui}1iN\{U_{i}\}_{1\leq i\leq N} be a finite covering of MM by sets of the form Ui=MBd(zi,r)U_{i}=M\cap B_{{\mathbb{R}}^{d}}(z_{i},r) for some z1,,zNMz_{1},\dots,z_{N}\in M, where r>0r>0 is the constant from Lemma 3. By Lemma 3, for each ii there exists a neighborhood 𝒪i\mathcal{O}_{i} of the identity in Diff(M){\rm Diff}(M), such that any ψ𝒪i\psi\in\mathcal{O}_{i} with supp(ψ)Ui{\rm supp}(\psi)\subseteq U_{i} belongs to 𝒫\mathcal{P}. Moreover, by Lemma 4 we may assume that ψ(zi)=zi\psi(z_{i})=z_{i}, i.e., ψ𝒫zi\psi\in\mathcal{P}_{z_{i}}. Taking 𝒪=1iN𝒪i\mathcal{O}=\cup_{1\leq i\leq N}\mathcal{O}_{i}, Lemma 2 says that every ψ𝒪\psi\in\mathcal{O} can be written as ψ=ψNψ1\psi=\psi_{N}\circ\dots\circ\psi_{1}, where each ψi𝒪i\psi_{i}\in\mathcal{O}_{i}. This, together with the lower estimate for NN in (12), completes the proof of Theorem 3. The extra factor of nn in (10) comes from Lemma 3. Moreover, the lower bound in (10) is rather generous since it only accounts for the lower bound on the number of fragments of ψ\psi according to Lemma 2 and for the fact that knk\geq n in Lemma 3.

4 Conclusion

While we were able to touch upon only some of the aspects of the controllability problem for the Liouville equation, we can highlight certain key ideas: First of all, the use of time-varying feedback controls seems inevitable. Indeed, the situations where a given diffeomorphism can be realized as a finite-time flow map of some time-invariant vector field are rather rare, as can be seen from the relative dearth of explicit results on the so-called embedding problem [Gro91, AG02]: Given a diffeomorphism ψ:nn\psi\mathrel{\mathop{\mathchar 58\relax}}{\mathbb{R}}^{n}\to{\mathbb{R}}^{n}, find a complete vector field ff on n{\mathbb{R}}^{n}, such that ψ(x)=ef(x)\psi(x)=e^{f}(x) for all xnx\in{\mathbb{R}}^{n}. Second, even if the requisite feedback controls can be obtained by switching between finitely many time-invariant flows, the resulting implementation complexity as measured by the number of switchings can be quite high, at least given the available techniques in the nonlinear setting, which rely on some combination of Sussmann’s orbit theorem, smooth fragmentation, and covering arguments. Hence, the construction of more parsimonious controls is an interesting open problem. Another promising direction is to consider the possibility of feedback control in the space of densities, i.e., instead of state feedback of the form u(t,x(t))u(t,x(t)) use density feedback of the form u(t,ρ(t,))u(t,\rho(t,\cdot)), where ρ(t,)\rho(t,\cdot) is the density of x(t)x(t) [AG08, PR13].

Acknowledgments

The author would like to thank Joshua Hanson, Borjan Geshkovski, Yury Polyanskiy, and an anonymous reviewer for their comments on the manuscript, and Tryphon Georgiou for bringing to his attention the flaw in the controllability argument in [Bro07] (and consequently in an earlier version of this work). This work was supported by the NSF under awards CCF-2348624 (“Towards a control framework for neural generative modeling”) and CCF-2106358 (“Analysis and Geometry of Neural Dynamical Systems”), and by the Illinois Institute for Data Science and Dynamical Systems (iDS2), an NSF HDR TRIPODS institute, under award CCF-1934986.

References

  • [AC09] A.A. Agrachev and M. Caponigro. Controllability on the group of diffeomorphisms. Annales de l’Institut Henri Poincaré C, Analyse non linéaire, 26(6):2503–2509, 2009.
  • [AG02] J. Arango and A. Gómez. Diffeomorphisms as time one maps. Aequationes Mathematicae, 64:304–314, 2002.
  • [AG08] L. Ambrosio and W. Gangbo. Hamiltonian ODEs in the Wasserstein space of probability measures. Communications on Pure and Applied Mathematics, 61(1):18–53, 2008.
  • [AG24] M. Abdelgalil and T.T. Georgiou. Collective steering in finite time: controllability on GL+(n,)GL^{+}(n,{\mathbb{R}}). arXiv.org preprint https://arxiv.org/abs/2411.18766, 2024.
  • [AL09] A. Agrachev and P. Lee. Optimal transportation under nonholonomic constraints. Transactions of the American Mathematical Society, 361(11):6019–6047, 2009.
  • [AS04] A. A. Agrachev and Yu. L. Sachkov. Control Theory from the Geometric Viewpoint. Springer, 2004.
  • [Ban97] A. Banyaga. The Structure of Classical Diffeomorphism Groups. Springer, 1997.
  • [BBFR19] J. Bartsch, A. Borzì, F. Fanelli, and S. Roy. A theoretical investigation of Brockett’s ensemble optimal control problems. Calculus of Variations and Partial Differential Equations, 58(5):162, 2019.
  • [Bha97] R. Bhatia. Matrix Analysis. Springer, 1997.
  • [BJPR22] A. Block, Z. Jia, Y. Polyanskiy, and A. Rakhlin. Intrinsic dimension estimation using Wasserstein distances. Journal of Machine Learning Research, 23:1–37, 2022.
  • [Bro97] R. W. Brockett. Minimum attention control. In Proceedings of the IEEE International Conference on Decision and Control, pages 2628–2632, 1997.
  • [Bro07] R. W. Brockett. Optimal control of the Liouville equation. AMS/IP Studies in Advanced Mathematics, 39:23–35, 2007.
  • [Bro12] R. W. Brockett. Notes on the control of the Liouville equation. In P. Cannarsa and J.-M. Coron, editors, Control of Partial Differential Equations, pages 101–129. Springer, 2012.
  • [Cap11] M. Caponigro. Orientation preserving diffeomorphisms and flows of control-affine systems. IFAC Proceedings Volumes, 44(1):8016–8021, 2011.
  • [CGP17] Y. Chen, T. T. Georgiou, and M. Pavon. Optimal transport over a linear dynamical system. IEEE Transactions on Automatic Control, 62(5):2137–2152, 2017.
  • [Chr85] M. Christ. Hilbert transforms along curves I. Nilpotent groups. Annals of Mathematics, 122:575–596, 1985.
  • [DR09] A. L. Dontchev and R. T. Rockafellar. Implicit Functions and Solution Mappings: A View From Variational Analysis. Springer, 2009.
  • [EZOP23] K. Elamvazhuthi, X. Zhang, S. Oymak, and F. Pasqualetti. Learning on manifolds: Universal approximations properties using geometric controllability conditions for neural ODEs. In Proceedings of The 5th Annual Learning for Dynamics and Control Conference, pages 1–11, 2023.
  • [Fed59] H. Federer. Curvature measures. Transactions of the American Mathematical Society, 93(418-491), 1959.
  • [FMN16] C. Fefferman, S. Mitter, and H. Narayanan. Testing the manifold hypothesis. Journal of the American Mathematical Society, 29(4):983–1049, 2016.
  • [Gro91] D. Gronau. On the structure of the solutions of the Jabotinsky equations in Banach spaces. Zeitschrift für Analysis und ihre Anwendungen, 10(3):335–343, 1991.
  • [HK77] R. Hermann and A. J. Krener. Nonlinear controllability and observability. IEEE Transactions on Automatic Control, AC-22(5):728–740, 1977.
  • [HPR11] A. Hindawi, J.-B. Pomet, and L. Rifford. Mass transportation with LQ cost functions. Acta Applicandae Mathematicae, 113(2):215–229, 2011.
  • [KL09] B. Khesin and P. Lee. A nonholonomic Moser theorem and optimal transport. Journal of Symplectic Geometry, 7(4):381–414, 2009.
  • [Mac92] M. C. Mackey. Time’s Arrow: The Origins of Thermodynamic Behavior. Springer, 1992.
  • [Mos65] J. Moser. On the volume elements on a manifold. Transactions of the American Mathematical Society, 120:286–294, 1965.
  • [Pog16] N. Pogodaev. Optimal control of continuity equations. Nonlinear Differential Equations and Applications NoDEA, 23(2):21, 2016.
  • [PR13] B. Piccoli and F. Rossi. Transport equation with nonlocal velocity in Wasserstein spaces: convergence of numerical schemes. Acta Applicandae Mathematicae, 124:73–105, 2013.
  • [PS70] J. Palis and S. Smale. Structural stability theorems. Proceedings of Symposia in Pure Mathematics, 14:223–231, 1970.
  • [RW98] R. T. Rockafellar and R. J.-B. Wets. Variational Analysis. Springer, 1998.
  • [Sus73] H. J. Sussmann. Orbits of families of vector fields and integrability of distributions. Transactions of the American Mathematical Society, 180:171–188, 1973.
  • [Vil03] C. Villani. Topics in Optimal Transportation. American Mathematical Society, 2003.