Some Remarks on Controllability of the Liouville Equation
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 and a state-space diffeomorphism , design a feedback control to steer an arbitrary initial state to 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 . 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 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 , where the focus is not on the evolution of the system state per se, but rather on the evolution of its probability density starting from a given initial density . 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
(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 and 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 . One can also formulate optimal control problems in this setting, e.g., minimizing a finite-horizon performance index of the form
over appropriately chosen controls , where the initial density is given and the integration is over both the (finite) time interval and the state space . 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 versus closed-loop (or feedback) controls . 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 and assume that the target density can be expressed as the pushforward of the initial density by an orientation-preserving diffeomorphism , i.e.,
for all bounded, continuous functions . In this case, the two densities are related via the Monge–Ampère equation
where denotes the Jacobian of . It is important to note that not any pair of densities can be joined in this way; for example, it will not be possible if vanishes on some open set while 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 and ask whether there exists a feedback control law that transfers the state of the system from to , for every . 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 that pushes forward to 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 on the space of everywhere positive probability densities on .
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 and be two probability densities on , 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 , such that its gradient pushes forward to . Moreover, this mapping is optimal in the sense that
where the minimum is over all that push forward to . 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 , such that the trajectory of the Liouville equation
joins to , and the performance index
is minimized. On the other hand, the corresponding controlled system allows for maximum “control authority,” in the sense that all directions of motion are available to the controller at each time . The more difficult case of a general controlled system 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 indexed by the elements of the control set .
2 The case of a controllable linear system
The case of a linear system
(2) |
with -dimensional state and -dimensional input has been considered in [Bro07] under the assumption that the system is controllable, i.e., the columns of span . Let an orientation-preserving diffeomorphism be given. By controllability, we can steer (2) from any fixed initial state to using the open-loop control
where
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 to ensure that the map
(3) |
is injective for all . If this is the case, then the closed-loop (feedback) control
(4) |
can be used to carry out the transfer from to for every initial condition .
In [Bro07], Brockett stated that one sufficient condition to ensure the above result is for the map 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
is invertible for every . This, in turn, relies on the claim that, since is positive definite for every by controllability, it follows that . However, the latter claim is not valid: Since the matrix 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 with
(5) |
From this, it follows readily that the mapping
is invertible since (multiply the matrix inequality on the left and on the right by ). The desired conclusion then follows from the fact that
The above assumption on is fairly restrictive. For an arbitrary , one could first represent it as a composition such that each satisfies (modified) Brockett’s condition and then apply the above construction for each . However, the number 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 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 with symmetric positive-semidefinite and symmetric positive-definite [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 is monotone if
Note that this definition makes no assumptions on differentiability of (in fact, it can be extended to set-valued mappings); however, when is differentiable, monotonicity is equivalent to the symmetric part of the Jacobian being everywhere positive-semidefinite. In particular, any given by the gradient of a convex function is monotone.
Theorem 1.
Proof.
We will show that the map defined in (3) is injective for , thus the feedback control given by (4) will do the job.
To that end, we claim that, for any ,
(6) |
which will imply that is injective since the matrices are nonsingular for . Using the definition of in (5), we can write
For any , let and . Then
If , then the first term on the right-hand side is strictly positive by controllability, while the second term is nonnegative since is monotone. This proves (6), so that is indeed injective and in (4) gives the desired feedback control law. ∎
To connect Theorem 1 to the problem of controlling a given initial density to a given final density using (2), let be the pushforward of by the invertible linear transformation and, similarly, let be the pushforward of by the invertible linear transformation . By Brenier’s theorem, there is a monotone map such that . Then the map related to via (5) evidently pushes forward to , and in that case Theorem 1 tells us how to construct the desired feedback control . In particular, the Benamou–Brenier dynamic formulation of optimal transportation is a special case corresponding to and with , cf. [HPR11, CGP17].
3 Controllability on the group of diffeomorphisms
Let us now consider the case of a nonlinear system
(7) |
whose state space is a smooth (say, ) closed finite-dimensional manifold. A smooth diffeomorphism is given, and the problem is to determine whether there exists a control law that can steer every initial state to , for a fixed finite . 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
(8) |
where are smooth vector fields on and are real-valued control inputs. Suppose that is a bracket-generating family, i.e., for each the set , where is the Lie algebra generated by , coincides with the tangent space to at . Then, for any diffeomorphism isotopic to the identity there exists a time-dependent feedback control law that transfers to for every . (Two diffeomorphisms are isotopic if there exists a smooth map , such that for each is a diffeomorphism, , and [Ban97]. We will denote by the family of all diffeomorphisms on that are isotopic to the identity.)
The general setting considered in [AC09] is as follows: Let a family of smooth vector fields on be given. The control group of , defined by
is a subgroup of the group of smooth diffeomorphisms of . Suppose that acts transitively on , i.e., for any there exists some such that . Then we have the following [AC09]:
Theorem 2.
There exist a neighborhood of the identity in and a positive integer that depends only on , such that every can be represented as
(9) |
for some and some .
Note that the vector fields in (9) are rescaled by smooth real-valued functions on , which explains the origin of time-varying feedback controls in the above discussion of control-affine systems. It is also evident that these controls will be piecewise constant in .
With this in mind, let us consider the system (8), where is bracket-generating. Let be given. We will apply Theorem 2 to —because is bracket-generating, acts transitively on . Following [Cap11], we first establish a fragmentation property of relative to , i.e., show that there exist some such that . Since , there exists a smooth path such that and . For every , the maps
belong to , and . We can ensure that each by choosing large enough. Applying Theorem 2 to each , we conclude that there exist smooth functions , such that can be represented as
for some choice of indices , cf. [Cap11, Prop. 4]. From this representation, it is straightforward to derive feedback controls that are piecewise constant in , with the number of pieces (switchings) equal to . Thus, the integer 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 .
To that end, we first make some assumptions on and . We take to be a smooth compact manifold of dimension isometrically embedded in for some , and we will equip with the ambient metric , where is the Euclidean norm on . We further assume that has positive reach , where the reach of a set is defined as the largest value of , such that any point at a distance from has a unique nearest point in [Fed59]. For example, the unit sphere has reach . We will assume that the vector fields are uniformly bounded in the seminorm, in the following sense [AS04, Sec. 2.2]. For , the seminorm of a function is defined by
where are the orthogonal projections on of the standard basis vector fields on . The seminorm of a vector field on is defined as
where is the Lie derivative of along . In terms of these definitions, we will assume that
We then have the following:
Theorem 3.
Let and satisfy the above assumptions. Then Theorem 2 holds with
(10) |
for some sufficiently small , where is the -dimensional Euclidean ball of radius centered at the origin.
Remark 1.
The lower bound on 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 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 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 containing the origin, we will denote by the space of functions , such that . When , we will simply write . Also, will denote the Euclidean ball of radius centered at .
The first key ingredient in the proof of Theorem 2 is the following:
Proposition 1.
Let be vector fields on with for all , such that are linearly independent. Then there exist constants and an open set
such that every can be written as
for some .
Remark 2.
By multiplying each by a suitable bump function, the everywhere boundedness assumption on can be relaxed to 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 , we can represent every in a certain open neighborhood of the identity in as a composition of 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 be vector fields on satisfying the conditions of Proposition 1. Let be a given neighborhood of the identity in . Then there exist constants , such that any with can be represented as a composition of the form
for some , where each preserves the integral curves of , i.e., for any there exists some such that .
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 , let be the matrix with columns . Since span , the smallest singular value of is positive:
The function is continuous as a consequence of the continuity of ’s and Weyl’s perturbation theorem [Bha97]. Therefore, there exists some , such that
Consider the function given by
Then, denoting by and the partial derivatives of w.r.t. and respectively, we have
Hence, there exists some , such that
so, by the classical inverse function theorem, the map is invertible on for each . We claim that, for any , the image of the ball under contains the ball , and thus for any there exists a unique , such that . To establish the claim, fix a unit vector and define a vector field on by for all . Since on , the curve is well-defined and remains in for all , as long as . Moreover, since by the chain rule we have
it follows that for . Since was arbitrary, the claim follows.
Now, fix some to be chosen later and let be the set of all , such that
(11) |
By the preceding discussion, if , then for any and for any there exists a unique , such that . The chain rule then gives
and, since is invertible for , we can solve for :
Consider the curve , , given by
where, for ,
Then , and we can express the derivative as the solution of the matrix-valued variational equation
For small enough, we will have
for some constant that depends only on and on the seminorms of the ’s. It follows that, for with small enough, we will have
i.e., the map is -Lipschitz.
Now consider, following [AC09], the maps
where and . Since , it is a diffeomorphism with domain by the preceding discussion, and, in fact, for small enough we will have for some constant . Finally, define, for , the maps
so that . Each evidently preserves the integral curves of . Moreover, from (11) it follows that for all , and thus we can ensure that each by taking sufficiently small. ∎
The second key ingredient is a fragmentation lemma for the elements of 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 :
In other words, the support of is the closure of the set on which differs from the identity [Ban97].
Lemma 2.
Let be a neighborhood of the identity in . Then, for any , where is the reach of , there exist points with
(12) |
such that the sets cover and such that any can be written in the form , where each is an element of with for all .
Remark 3.
Since the group is path-connected, it is generated by any neighborhood of the identity. Thus, is generated by the set .
Proof.
Since is compact, for any it can be covered by finitely many sets of the form for some . From [BJPR22, Corollary 10], the smallest value of for a given , i.e., the covering number of at resolution w.r.t. the ambient metric, can be estimated from above and from below as in (12). Let be such a covering of . 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 , the control group generated by
and define the isotropy subgroup of at :
For an open set and , we will denote by the family of all smooth maps , such that .
Lemma 3.
There exists a constant such that, for any and for , the set
(13) |
has nonempty interior in which contains the identity. Moreover, there exist vector fields depending on , such that any element of the interior of (13) can be represented as a composition of the form
(14) |
for some .
Proof.
We follow the proof of Lemma 5.1 and Corollary 5.2 in [AC09], but with some steps made more explicit. Since acts transitively on , Sussmann’s orbit theorem [Sus73] implies that the tangent space to at is equal to the span of the set , where denotes the action of the tangent map of on . Thus, there exist vector fields of the form for some and , such that span . Moreover, for all smooth functions that vanish at , the diffeomorphism
(15) |
belongs to the isotropy group . Now, using the techniques of [FMN16], we can show that, for any small enough, the set can be locally coordinatized using the elements of as
for some , where is a map from into , such that , , and the Jacobian is -Lipschitz, where is some constant that depends only on . In particular, the map is invertible. (In the terminology of [FMN16], the set
is a patch over of radius , centered at and tangent to at .) 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 . Moreover, since the vector fields in are uniformly bounded in seminorm, it follows from the proof of Lemma 1 that the constant can be chosen independently of .
To show that the interior of the set (13) contains the identity map, we let be an open subset of contained in the interior of (13). Then, for any such that , the set is a neighborhood of the identity contained in (13).
Finally, since each in (15) belongs to the control group , it can be expressed as
for some integer , times , and vector fields . Substituting this into (15) and relabeling the functions and the vector fields as needed, we arrive at the representation (14) for some and , where
This completes the proof. ∎
Lemma 4.
Let be a neighborhood of the identity in . Then for any and any open set containing ,
Proof.
See the proof of Lemma 5.3 in [AC09]. ∎
Now, let be a finite covering of by sets of the form for some , where is the constant from Lemma 3. By Lemma 3, for each there exists a neighborhood of the identity in , such that any with belongs to . Moreover, by Lemma 4 we may assume that , i.e., . Taking , Lemma 2 says that every can be written as , where each . This, together with the lower estimate for in (12), completes the proof of Theorem 3. The extra factor of 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 according to Lemma 2 and for the fact that 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 , find a complete vector field on , such that for all . 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 use density feedback of the form , where is the density of [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 . 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.