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

Topological Necessary Conditions for Control Dynamics

EFTHIMIOS KAPPOS
School of General Sciences,111Submitted for publication, 2009, under revision
Department of Mathematics,
Faculty of Engineering,
Aristotle University of Thessaloniki, Greece

1 Introduction

Suppose a smooth feedback control has been found so that the controlled dynamics have an asymptotically stable attractor at some point x in state space. Then local Lyapunov functions exist for the dynamics; these functions must all have a unique minimum at the point x, but are otherwise arbitrary. On a compact level set of any such Lyapunov function, the controlled dynamics point inwards, in other words in the direction of the negative of the gradient of the Lyapunov function. As maps to the unit sphere, the two vector fields thus have the same degree. But the degree of the negative gradient vector field is known —it is exactly (1)n0(−1)^{n}\neq 0. Hence the map of the controlled dynamics, restricted to a level set, to the sphere must be onto. This means roughly that all control directions must be available near a point that is to be stabilized by control. In control theory, this is known as the Brockett condition, but such simple de- gree results were widely known before (Krasnosel’skii’s name is mentioned in conjunction with that of Brockett.) Such degree-theoretic arguments have been used for some time in topology and were eventually adopted by control theorists to derive a number of related necessary conditions for design of controlled dynamics using continuous feed- back.

In this paper, we give an account of this theory that has two distinguishing features: first, there is really no reason to limit ourselves to local results: a global theory is straightforward to obtain. Secondly, we point out that such necessary conditions are in some funda- mental sense of limited value; this is because they involve maps from manifolds to spheres of the same dimension. By the Hopf theory, homotopy equivalence classes of such maps are completely classified by a single integer (traditionally called the degree, but better interpreted in terms of homology groups.)

Several comments are in order: generalizations of the Brockett necessary condition have been obtained by Coron and others. These suffer to some extent by the problem mentioned above, but have helped clarify the fact that surjectivity is not enough: the vector field cannot ‘twist’ too much either (examples use the degree kk maps z˙=zk\dot{z}=z^{k} as part of a control system decomposition.) Moreover, extensions to the case of dynamic feedback have been derived.

Perhaps more importantly, many systems cannot be stabilized using continuous feedback, but can be easily stabilized with discontinuous feedback. The discontinuity is rather mild: it is usually limited to a ‘thin’ subset of state space. Recent work of Sontag, Clarke, Subbotin and others has led to a theory of discontinuous feedback controls and a methodology for obtaining nonsmooth Lyapunov functions. Now it is possible to interpret this theory in a hamiltonian context: The discontinuities correspond to jumps between locally nonsingularly projected lagrangian levels. This way of examining possibly discontinuous feedback controls is conceptually easier to understand and is in step with the philosophy of the book [3] which is to give, as far as is possible, geometric accounts of analytical points.

We begin by giving an outline of the algebraic topological machinery needed for a discussion of necessary conditions. Here, we depart from the practice of delegating mathematical background to an appendix, because we believe that this theory is quite accessible and elegant.

A collection of global necessary conditions is then given, directly based on the topological results. Essentially, it is argued that if certain dynamics are achieved, then index-theoretic conditions can be deduced by counting the equilibrium points and their stability (Euler-type arguments) and degree-theoretic results are obtained by the Hopf theorems using the Gauss maps of the dynamics and the gradient vector field of Lyapunov functions.

Finally, let us point out that this paper by no means exhausts the theory of necessary conditions in control design. Much more crucial limitations on achievable control dynamics arise from the theory of feedback invariant objects, a theory that is developed in the book [3].

2 Some Background and Methods from Algebraic Topology

A thumb-nail sketch of a number of concepts and methods from algebraic topology will now be given. There is no effort to be rigorous, but we do hope to explain enough about the computational methods so that a non-expert reader can use them in concrete situations.

Algebraic topology is based on a simple principle: attach algebraic objects to topological spaces that are invariants of the homotopy type of the space. Thus, more precisely we assign algebraic objects to homotopy equivalence classes of spaces and this assignment is ‘functorial’ in the sense that maps of spaces induce homomorphisms of the algebraic objects. This already gives useful tests: since homotopy equivalent spaces have isomorphic algebraic objects, two spaces are definitely not homotopy equivalent if their algebraic objects are not isomorphic. The bulk of algebraic topology consists of deriving finer and finer such objects so as to be able to better distinguish spaces and in making clever use of its basic constructions to aid the analysis of global aspects of other subjects (such as complex analysis, pdes, geometry etc.)

Singular homology

The easiest algebraic object we can attach to a space is the graded abelian group H(X)H_{*}(X) called the singular homology group of XX (with integer coefficients.) One can get quite far with only a vague understanding of what the singular homology measures and the reason is that powerful and effective methods for the computation of H(X)H_{*}(X) exist. We outline the Mayer-Vietoris sequence and explain the concept of a long exact sequence of a pair and its relation to excision.

A graded abelian group G=GkG=\oplus G_{k} is a direct sum of groups GkG_{k}, k+k\in{\mathbb{Z}}_{+}, such that the group addition is ‘component-wise’, in other words we add elements belonging to the same graded component together. The notation

g=+g0+g1+g=\ldots+g^{0}+g^{1}+\ldots

for an element of GG, where gkGkg^{k}\in G_{k} is therefore unambiguous.

For a topological space, the kkth homology group Hk(X;)H_{k}(X;{\mathbb{Z}}) measures in some sense the ‘holes’ of XX that are like kk spheres SkS^{k} (think of a boundary-less space, like a sphere, that does not actually bound anything itself in XX.) The 0th group H0(X)H_{0}(X) is equal to {\mathbb{Z}} if XX is path connected. (Recall that a 0-sphere is the boundary of an interval, i.e. the union of two points.) There is a way of defining reduced homology groups H~k(X)\tilde{H}_{k}(X) so that H~0(X)=0\tilde{H}_{0}(X)=0 for a connected space and so that all higher dimensional groups coincide with the non-reduced ones.

Let us give some examples (we omit the zeroth homology group.) The singular homology of the circle S1S^{1} is H1(S1)H_{1}(S^{1})\simeq{\mathbb{Z}} and zero for k>1k>1. Since π1(S1)=\pi_{1}(S^{1})={\mathbb{Z}} also, the homology group contains the same information as the fundamental group of the circle. Note the difference in interpretation, though: In the former case (for π1\pi_{1}), we are thinking of maps from the circle to itself, classifed by the number of net encirclements In the latter, we are thinking of a fixed circle —coinciding in this case with the whole space S1S^{1}— as the generator of a free abelian group; in this sense, we can write

H1(S1)=[S1].H_{1}(S^{1})={\mathbb{Z}}[S^{1}]\simeq{\mathbb{Z}}.

For the sphere SmS^{m} of dimension m>1m>1, Hm(Sm)H_{m}(S^{m})\simeq{\mathbb{Z}} is the only nonzero homology group in positive dimension. Since we also have that the mmth homotopy group of an mm sphere is {\mathbb{Z}}, we have not yet obtained anything new, compared with homotopy theory. This is a little misleading: homotopy is both subtler than homology and far more difficult to compute: we do not, even today, have a complete list of the homotopy groups of spheres. Moreover, πm+k(Sm)\pi_{m+k}(S^{m}) may very well be nonzero for some k>0k>0, while Hm+k(Sm)=0H_{m+k}(S^{m})=0 always.

The ‘coincidence’ is really due to a nontrivial theorem, the Hurewicz isomorphism that states that homotopy and homology groups are isomorphic at the first level when one, and hence both, are nontrivial (the abelianization of the possibly nonabelian fundamental group is to be considered, if this happens at the first level.)

A quick check that homology theory does indeed give something new is to compute the homology of the torus T2T^{2}. We have that H2(T2)H_{2}(T^{2})\simeq{\mathbb{Z}} even though π2(T2)=0\pi_{2}(T^{2})=0! (it may be profitable to spend a minute or two pondering the difference.)

The Mayer-Vietoris sequence

Supose a space XX is the union of two opne subsets, X=ABX=A\cup B, with A,BA,B opne and ABA\cap B\neq\emptyset. Then there is a long exact sequence involving the homology groups of the three spaces

Hk(AB)Hk(A)Hk(B)Hk(X)Hk1(AB)\ldots\to H_{k}(A\cap B)\to H_{k}(A)\oplus H_{k}(B)\to H_{k}(X)\to H_{k-1}(A\cap B)\to\ldots (1)

This gives a surprisingly powerful tool for the computation of homology. Even without knowing what the maps at each stage are (for which we refer the reader to standard accounts such as [2]) the exactness allows the computation in concrete cases, such as that of the spheres. For this, decompose an mm-sphere into two slightly overlapping hemispheres AA and BB so that their intersection ABA\cap B is deformable to a sphere of dimension m1m-1. We can start an induction with dimension m=1m=1 and use the Mayer-Vietoris sequence to obtain

Hk(Sm1)Hk(A)Hk(B)Hk(Sm)Hk1(Sm1)\ldots\to H_{k}(S^{m-1})\to H_{k}(A)\oplus H_{k}(B)\to H_{k}(S^{m})\to H_{k-1}(S^{m-1})\to\ldots (2)

yielding, for k=mk=m, and since disks have no homology

000Hm(Sm)0.\ldots\to 0\to 0\oplus 0\to H_{m}(S^{m})\to{\mathbb{Z}}\to 0\to\ldots. (3)

We conclude that Hm(Sm)H_{m}(S^{m})\simeq{\mathbb{Z}}, since any exact sequence of the form

0CD00\to C\to D\to 0

implies that the middle map is one-to-one and onto, i.e. an isomorphism.

Long exact sequence of a pair and excision

A second very useful method for the computation of homology comes from considering pairs (X,A)(X,A), where AXA\subset X is a subspace. One gets the long exact sequence for the pair

Hk(A)Hk(X)Hk(X,A)Hk1(A)\ldots\to H_{k}(A)\to H_{k}(X)\to H_{k}(X,A)\to H_{k-1}(A)\to\ldots (4)

where the groups Hk(X,A)H_{k}(X,A) are the relative homology groups. Without giving the exact definition, found in the standard texts, let us mention that in many important cases, these relative groups are isomorphic to the homology groups of the quotient space X/AX/A (see chapter 3 for the definition.) The long exact sequence for a pair is thus extremely useful for the computation of the homological Conley index.

As an example, let us show that the quotient Dn/Sn1D^{n}/S^{n-1} of a closed ball by its bounding sphere has the homology of the nn-sphere SnS^{n}. The long exact sequence of the pair (Dn,Sn1)(D^{n},S^{n-1}) is

Hk(Dn)Hk(Dn,Sn1)Hk1(Sn1)Hk1(Dn)\ldots\to H_{k}(D^{n})\to H_{k}(D^{n},S^{n-1})\to H_{k-1}(S^{n-1})\to H_{k-1}(D^{n})\to\ldots (5)

and so, at k=nk=n, we get

0Hk(Dn,Sn1)0\ldots\to 0\to H_{k}(D^{n},S^{n-1})\to{\mathbb{Z}}\to 0\to\ldots (6)

hence Hk(Dn,Sn1)Hk(Dn/Sn1)H_{k}(D^{n},S^{n-1})\simeq H_{k}(D^{n}/S^{n-1})\simeq{\mathbb{Z}}. Similarly, one finds that, for knk\neq n (and nonzero), Hk(Dn/Sn1)=0H_{k}(D^{n}/S^{n-1})=0.

Maps and homomorphisms

Given a continuous map f:XYf:X\to Y, there is an induced map in homology, which we shall denote by H(f)H_{*}(f) or ff_{*}

H(f):H(X)H(Y)H_{*}(f):H_{*}(X)\to H_{*}(Y)

and one checks that homology is a covariant functor from the category Top =( Top,C0)=(\text{ Top},C^{0}) of topological spaces and continuous maps to the category Ab=(Ab,Hom)=(\text{Ab},\text{Hom}) of abelian groups and homomorphisms between them. Since H(X)H_{*}(X) is graded, the above homomorphism is understood to mean that it consists of homomorphisms at each level of homology:

Hk(f):Hk(X)Hk(Y)H_{k}(f):H_{k}(X)\to H_{k}(Y)

for all kk.

In fact, it would be more precise to say that the functor goes from the category hTop of homotopy equivalence classes of spaces and homotopic maps to the category Ab, since

Proposition 1.

If the maps ff and gg are homotopic, then the maps in homology coincide: f=gf_{*}=g_{*}.

and

Proposition 2.

If two spaces XX and YY are homotopy equivalent and the map ff has a homotopy inverse, then ff_{*} is an isomorphism and the homology groups of XX and YY, H(X)H_{*}(X) and H(Y)H_{*}(Y) are isomorphic.

Corollary 1.

If ff is a homeomorphism of spaces, then ff_{*} is an isomorphism.

When the space XX is a finite-dimensional manifold, the homology groups Hk(X;)H_{k}(X;{\mathbb{Z}}) are finitely generated; thus, in this case, the basic structure theorem for finitely generated abelian groups is applicable.

Theorem 1 (Structure Theorem).

Any finitely generated abelian group GG decomposes uniquely as the direct sum

G=FτG=F\oplus\tau

where the abelian group FF is free and the group τ\tau is a torsion subgroup.

In fact, one can describe the torsion group τ\tau is more detail (see, for example, [4].)

The dimension of the free part of the homology group HkH_{k} is called the kkth-Betti number, bk=dimHk(X;)b_{k}=\dim H_{k}(X;{\mathbb{Z}}). The Euler characteristic χ(X)\chi(X) is the alternating sum of the Betti numbers

χ(X)=k(1)kbk.\chi(X)=\sum_{k}{(-1)}^{k}b_{k}.

3 Collections of topological necessary conditions

The definition of certain Gauss maps is helpful in the statement of our results. We shall assume that Mn=nM^{n}={\mathbb{R}}^{n} or is an open subset of it.

Definition 3.1.
  1. 1.

    Suppose the vector field XX is nowhere zero in MnM^{n}. Then the Gauss map GX:MnSn1G_{X}:M^{n}\to S^{n-1} is defined by

    xX(x)|X(x)|.x\mapsto\frac{X(x)}{|X(x)|}.
  2. 2.

    Suppose that Nn1MnN^{n-1}\subset M^{n} is a submanifold such that the restriction of the vector field XX to NN is nowhere zero. Then the Gauss map GX|N:Nn1Sn1G_{X|N}:N^{n-1}\to S^{n-1} is obtained by restricting the Gauss map GXG_{X} to NN. Note that this is a map between two manifolds of the same dimension, one of which is a sphere.

  3. 3.

    If the submanifold Nn1MnN^{n-1}\subset M^{n} is orientable, we define the Gauss map GN:Nn1Sn1G_{N}:N^{n-1}\to S^{n-1} by mapping xNx\in N to the unit normal vector to NN at xx (where an ‘outward’ direction is fixed by choosing an oriented basis on NN and completing it to a basis of n{\mathbb{R}}^{n} consistent with an orientation of n{\mathbb{R}}^{n}.) Note again that the Gauss map is a map from an (n1)(n-1) dimensional space to the (n1)(n-1)-sphere.

3.1 Index-Theoretic Necessary Conditions

The topological index of equilibrium points leads to a number of necessary conditions for achieving dynamics with equilibrium points of given stability. These are global results and are rather classical; our only novelty is in trying to use as modern an algebraic topological framework as we can to express them.

If eMne\in M^{n} is an isolated equilibrium point of the vector field XX, take a ball neighborhood UU of ee (an open set homeomorphic to a ball) such that ee is the only equilibrium of XX in UU and its boundary N=UN=\partial U is a closed submanifold homeomorphic to a sphere. Then the Gauss map GX|NG_{X|N} gives a map from the sphere Sn1S^{n-1} to itself

Sn1hNGX|NSn1S^{n-1}\stackrel{{\scriptstyle h}}{{\rightarrow}}N\stackrel{{\scriptstyle G_{X|N}}}{{\rightarrow}}S^{n-1}

where h1h^{-1} is the homeomorphism from NN to the sphere.

At the level of homology, we thus get a homomorphism ψ=GX|Nh\psi=G_{X|N}\circ h from Hn1(Sn1)H_{n-1}(S^{n-1}) to itself. Since this group is isomorphic to {\mathbb{Z}}, we get a homomorphism from to{\mathbb{Z}}to{\mathbb{Z}}. Since {\mathbb{Z}} is a principal ideal domain, such maps are specified by the image of the generator, say αHn1(Sn1)\alpha\in H_{n-1}(S^{n-1}). If, say, ψ(α)=kα\psi(\alpha)=k\alpha, then kk is the topological index of the equilibrium ee.222Confusingly, we are about to give a theorem where the term ‘degree’ is used instead of ‘toplogical index’; the two terms are equivalent. We shall try use the qualifier ‘topological’ to avoid confusion with other uses of the term index. It does not depend on the precise UU chosen.

The classical theorem of Hopf describes maps from the sphere to itself.

Theorem 2 (Hopf’s Classification Theorem).

Homotopy equivalence classes of maps from Sn1S^{n-1} to itself are in a one-to-one correspondence with the integers. For each integer kk, the class of maps corresponding to it is called the class of maps of degree kk.

For a hyperbolic equilibrium point of stability index kk, the topological index (or degree) is equal to (1)nk{(-1)}^{n-k}. Degree kk maps are easily obtained from the degenerate equilibria at the origin of the system in complex form: z˙=zk\dot{z}=z^{k}, for k0k\neq 0.

The Hopf classification of maps from the sphere to itself has a crucial generalization to maps of an arbitrary compact manifold of dimension n1n-1 to a sphere of dimension n1n-1 (see Whitehead, [7], p.244)

Theorem 3 (Hopf-Whitney).

The homotopy equivalence classes of maps of an (n1)(n-1)-dimensional compact manifold Nn1N^{n-1} to the sphere Sn1S^{n-1} are in one-to-one correspondence with the elements of the cohomology group Hn1(Nn1;)H^{n-1}(N^{n-1};{\mathbb{Z}}).

Corollary 2.

If NN is orientable, then the homotopy equivalence classes of maps from Nn1N^{n-1} to Sn1S^{n-1} are in one-to-one correspondence with the integers; they are thus again classified by ‘degree.’

This is, of course, because, for any orientable manifold, Hn1(Nn1;)H^{n-1}(N^{n-1};{\mathbb{Z}})\simeq{\mathbb{Z}}. If NN is not orientable, then this group is 2{\mathbb{Z}}_{2} and two maps are homotopic iff they have the same mod-22 degree.

The global version of the Hopf index classification result is the following theorem of Poincaré-Hopf

Theorem 4 (Poincaré-Hopf).
  1. 1.

    Suppose WnnW^{n}\subset{\mathbb{R}}^{n} is a compact subset with nonempty interior such that its boundary is an (n1)(n-1)-dimensional submanifold of n{\mathbb{R}}^{n}. Suppose XX is a vector field on n{\mathbb{R}}^{n} that is nowhere zero on the boundary W\partial W and has a finite set of equilibrium points EE. Then

    degGX|W=eiEWindei\deg G_{X|\partial W}=\sum_{e_{i}\in E\cap W}\operatorname{ind}e_{i} (7)
  2. 2.

    Suppose MnM^{n} is a compact manifold and XX is a vector field on MnM^{n} with a finite number of isolated equilibria. If the boundary of MnM^{n} is not empty, we require the vector field to point inwards at all points. The we have

    eiEindei=(1)nχ(Mn)\sum_{e_{i}\in E}\operatorname{ind}e_{i}={(-1)}^{n}\chi(M^{n}) (8)

    where χ(Mn)\chi(M^{n}) is the Euler characteristic of the manifold MnM^{n} and EE is the set of equilibrium points.

    In particular, the sum of the topological indices of the equilibria is a topological invariant of the manifold and thus is independent of the vector field chosen.

  3. 3.

    Suppose WkW^{k} is any submanifold of n{\mathbb{R}}^{n}, with 0kn10\leq k\leq n-1. Consider a tubular neighborhood Nϵ(Wk)N_{{\epsilon}}(W^{k}) so that Nϵ(Wk)\partial N_{{\epsilon}}(W^{k}) is an (n1)(n-1)-dimensional submanifold of n{\mathbb{R}}^{n}. If XX is any vector field on n{\mathbb{R}}^{n} such that, on WkW^{k}, XX has a finite number of nondegenerate equilibria, then

    eiEW=degGX|Nϵ(Wk)\sum_{e_{i}\in E\cap W}=\deg G_{X|N_{{\epsilon}}(W^{k})} (9)

We have collected different versions of this important theorem to help the reader find the most convenient form for extracting topological information in applications. Milnor [5] proves versions (2) and (3) and contains a nice discussion.

Remark.

The index already contains considerable topological information for the purposes of extracting necessary conditions. For the case of an asymptotically attracting equilibrium, for example, the topological index is equal to (1)n{(-1)}^{n}, which means that the generator of Hn1(Sn1)H_{n-1}(S^{n-1}) is mapped to itself or its negative, depending on the parity of nn. As a result, the Gauss map is an isomorphism in homology and we conclude that it must then be surjective and injective. The surjectivity is essentially the Krasnosel’skii-Brockett condition and the injectivity was derived by Coron. The form we have given is, however, considerably more general.

Remark.

It must be emphasized that the index is ‘blind’ to all other dynamical features except equilibria. Looking at the same point from the other side of the equalities in Theorem 4, the topological type of the Gauss map in the large (on the boundary of an enclosing set) affects the configuration of equilibria inside—and fixes the sum of their indices.

A few examples as simple illustrations of the statements of the theorem:

Example 1.

In n{\mathbb{R}}^{n}, a ball with a vector field pointing inwards at the boundary must contain equilibria whose index sum is (1)n{(-1)}^{n}. If these are all hyperbolic, then the options are

  • A single attracting equilibrium.

  • Two attractors and a one-saddle.

  • If nn is even, a single repeller is not ruled out; notice that the two cases can be distinguished using the Conley index, since the exit set differs for the two cases.

  • Any other configuration of equilibria with the same net index sum.

Example 2.

In 3{\mathbb{R}}^{3}, an embedded torus T2T^{2} gives possible Gauss maps of arbitrary degree, since its top homology is equal to {\mathbb{Z}}. If, however, we know that there are no enclosed equilibria, as for example in the case where the torus isolates a limit cycle, then the degree must be zero, by part (1) of Theorem 4, independently of the stability type of the limit cycle.

This means that the Gauss map is homotopic to the constant map and hence does not have to be onto (it is not onto in general, for a small enough torus around the limit cycle). Thus, no necessary condition is derivable in this case, whether the limit cycle is stable or not.

Example 3.

On the torus T2T^{2} we have, by part (2) of the Theorem, that any vector field must have total index sum equal to zero, since the Euler characteristic of the torus is zero. Thus, vector fields that everywhere nonzero are permissible topologically, as are vector fields with one attractor and one saddle, one repeller and a saddle, one attractor, one repeller and two saddles etc.

3.2 Necessary conditions using the topological index

It should be clear from the examples how to derive necessary conditions for achieving global dynamics from the index theorems.

Suppose given a Morse specification of gradient type, =(E,h0){\mathcal{M}}=(E,h_{0}), with Morse-lyapunov functions (){\mathcal{F}}({\mathcal{M}}). In the state space manifold, any choice of an oriented hypersurface that avoids |E||E| has a Gauss map degree fixed by the sum of the indices of the enclosed ‘equilibria’. If this is non-zero, this implies that there must exist control sections such that the controlled dynamics give a Gauss map with the desired property. In particular, if the index sum is equal to plus or minus one, then the Gauss map is onto. Let us remark that the conditions obtained can iether be used locally to check, for example, local stabilizability by requiring the map to have degree (1)n{(-1)}^{n} for an arbitrarily small neighborhood of the equilibrium, or globally, since the only relevant information is the position and stability of the desired equilibria and hence the index/degree results hold for any compact hypersurface avoiding |E||E|.

A more elegant algebraic topological way of checking simultaneously all necessary conditions is the following (this does not make it easier to check in concrete cases):

We do first the case of local asymptotic stabilizability.

Suppose a control section UΓ(D)U\in\Gamma(D) is found that locally stabilizes the origin 0 in some neighborhood BB; it will be helpful to consider the set, for ϵ>0\epsilon>0,

ΣB={(x,v)D|B;X(x)+v=0}\Sigma_{B}=\{(x,v)\in D{|}_{B}\;;\;X(x)+v=0\}

and the sequence

B{0}graphUB×mΣBιTn|B{0}\displaystyle B\setminus\{0\}\stackrel{{\scriptstyle\operatorname{graph}U}}{{\longrightarrow}}B\times{\mathbb{R}}^{m}\setminus\Sigma_{B}\stackrel{{\scriptstyle\iota}}{{\hookrightarrow}}T{\mathbb{R}}^{n}{|}_{B}\setminus\{0\}
GSn|BπSn1\displaystyle\stackrel{{\scriptstyle G}}{{\longrightarrow}}S{\mathbb{R}}^{n}{|}_{B}\stackrel{{\scriptstyle\pi}}{{\longrightarrow}}S^{n-1}

where graphU(x)=(x,U(x))\operatorname{graph}U(x)=(x,U(x)), ι\iota is the inclusion map, GG is the Gauss map and π\pi is the obvious projection in the trivial local sphere bundle.

Since 0 is an isolated equilibrium of X+UX+U, X+U0X+U\neq 0 in B{0}B\setminus\{0\} and the above is well-defined.

3.3 A reinterpretation of Coron’s condition

With the tools we have at our disposal, it is now easy to give a more geometric interpretation of the necessary condition for local feedback stabilization given in [1]: We start by noticing that, if BB is a ball neighborhood of the equilibrium 0, B{0}B\setminus\{0\} is homotopically equivalent to Sn1S^{n-1} (it actually retracts to the sphere). Thus the composed map defined by the above sequence, call it ϕ\phi,

ϕ:B{0}Sn1\phi:B\setminus\{0\}\to S^{n-1}

has a well-defined degree, since 0 is asymptotically stable for X+UX+U and this degree is equal to (1)n{(-1)}^{n}. This means that, at the level of, for example, homology (or homotopy), the generator, call it α\alpha, of Hn1(Sn1)H_{n-1}(S^{n-1})\simeq{\mathbb{Z}} is in the image of ϕ\phi. In other words, if 0 is LAS, then there is some local section such that the degree of the above map is defined and the image of the corresponding homomorphism at the level of homology is the whole of Hn1(Sn1)H_{n-1}(S^{n-1}). This is essentially Coron’s result: Consider the commutative diagram

B{0}Sn1DΣV\begin{array}[]{ccc}B\setminus\{0\}&\rightarrow&S^{n-1}\\ \downarrow&\nearrow&\\ D_{\setminus}\Sigma_{V}&&\end{array} (10)

where the vertical map is inclusion and the map from DBΣBD_{B}\setminus\Sigma_{B} to Sn1S^{n-1} will be denoted also by X+UX+U and is given by the composition (x,v)X(x)+vG(X(x)+v)(x,v)\mapsto X(x)+v\mapsto G(X(x)+v). We have that

ϕ(Hn1(B{0}))=Hn1(Sn1).\phi_{*}(H_{n-1}(B\setminus\{0\}))=H_{n-1}(S^{n-1}).
Theorem 5 (Coron, 1990).

If the system (X,D)(X,D) is locally asymptotically stabilizable, then

(X+U)(Hn1(DBΣB))=Hn1(Sn1).{(X+U)}_{*}(H_{n-1}(D_{B}\setminus\Sigma_{B}))=H_{n-1}(S^{n-1}).

3.4 Generalizations

The simple reasoning that led to Coron’s result can be generalized to equilibrium points that are not attractors, but have a well-defined stability index.

Theorem 6.

Let 0 be an equilibrium of the state dynamics XX of the control pair (X,D)(X,D). If there is a continuous local feedback that yields dynamics X+UX+U with 0 an equilibrium of index k, 0knk\,,\,0\leq k\leq n, then

(X+U)(Hn1(DBΣB))=Hn1(Sn1).{(X+U})_{*}(H_{n-1}(D_{B}\setminus\Sigma_{B}))=H_{n-1}(S^{n-1}).

Finally, necessary conditions applicable to an arbitrary compact, connected IIS 𝒮{\mathcal{S}}, isolated by the set BMnB\subset M^{n} can be given. More explicitly, we assume that there is a local feedback U:BDU:B\to D such that X+UX+U has an IIS 𝒮{\mathcal{S}}, whose dynamical structure is known (for example, 𝒮{\mathcal{S}} as a set consists of a number of equilibria and limit cycles and their connecting orbits.) Notice that X+U0X+U\neq 0 in B𝒮B\setminus{\mathcal{S}}.

Lemma 1.

Endow MnM^{n} with a Riemannian metric. There is a function hh defined on B𝒮B\setminus{\mathcal{S}} such that its gradient vector field h\nabla h is topologically equivalent to X+UX+U and such that the Gauss maps of h\nabla h and XUX_{U} induce the same homomorphisms on homology, both GhG_{\nabla h} and GXUG_{X_{U}} mapping

H(V𝒮)H(SMn|V𝒮).H_{*}(V\setminus{\mathcal{S}})\rightarrow H_{*}(SM^{n}{|}_{V\setminus{\mathcal{S}}}).

Now, as we did for the case of local stabilization, we have the map ϕ\phi defined by the composite map below

DB𝒮ΣX+vTMnΣGSMnD_{B\setminus{\mathcal{S}}}\setminus\Sigma\stackrel{{\scriptstyle X+v}}{{\rightarrow}}TM^{n}\setminus\Sigma\stackrel{{\scriptstyle G}}{{\rightarrow}}SM^{n} (11)

which induces the map ϕ\phi_{*} in homology

H(DB𝒮)H(SMn|B𝒮).H_{*}(D_{B\setminus{\mathcal{S}}})\rightarrow H_{*}(SM^{n}{|}_{B\setminus{\mathcal{S}}}).

We now have the result

Theorem 7.

If the control pair (X,D)(X,D) can achieve dynamics with IIS 𝒮{\mathcal{S}} isolated by the set BB, then the images of the maps ϕ\phi_{*} and GhG_{-\nabla h} in H(SMn|B𝒮)H_{*}(SM^{n}{|}_{B\setminus{\mathcal{S}}}) coincide.

4 Homotopy equivalence and homotopic results

An elementary, but fundamental result forms the key to an alternative approach to the derivation of necessary conditions. It concerns the Gauss maps of a gradient vector field of a Lyapunov function for the dynamics XX and the Gauss map of the dynamics on level sets of the Lyapunov function.

Theorem 8.

Let 𝒱n1{\mathcal{V}}^{n-1} be a compact regular level set of some Lyapunov function VV for the dynamics XX on MnnM^{n}\subset{\mathbb{R}}^{n}. Then, the Gauss maps GX|𝒱G_{X|{\mathcal{V}}} and GV|𝒱G_{-\nabla V|{\mathcal{V}}} os the vector field XX and of the gradient vector field of VV with respect to any riemannian metric are homotopy equivalent.

Proof.

Decompose the tangent space TMn|𝒱TM^{n}{|}_{{\mathcal{V}}} into the tangent space of 𝒱{\mathcal{V}} and the span of the gradient vector field V\nabla V. If XnX_{n} is the projection of XX to the span of V\nabla V,we have that XnX_{n} is nowhere zero on 𝒱{\mathcal{V}}.

Consider the isotopy of vector field

Yt(x)=(1t)Xn(x)+tX(x), 0t1.Y_{t}(x)=(1-t)X_{n}(x)+tX(x),\;0\leq t\leq 1.

We have that Y0=XnY_{0}=X_{n} and Y1=XY_{1}=X.

Now notice that this gives an isotopy for the corresponding Gauss maps as well: this is because Yt(x)0Y_{t}(x)\neq 0 on 𝒱{\mathcal{V}} and for all tt. To see this, write YtY_{t} as

Yt=Xn+t(XXn)Y_{t}=X_{n}+t(X-X_{n})

and notice that the vector field XXnX-X_{n} is orthogonal to XnX_{n}, which is everywhere nonzero.

Define the Gauss maps parametrized by tt

Gt:𝒱n1Sn1,xYt(x)|Yt(x)|.G_{t}:{\mathcal{V}}^{n-1}\to S^{n-1},\;x\mapsto\frac{Y_{t}(x)}{|Y_{t}(x)|}.

Since Yt(x)Y_{t}(x) is everywhere nonzero, this is well defined and gives an isotopy between

G0=Xn|Xn|=V|V|=GVG_{0}=\frac{X_{n}}{|X_{n}|}=\frac{-\nabla V}{|-\nabla V|}=G_{-\nabla V}

and

G1=X|X|=GX.G_{1}=\frac{X}{|X|}=G_{X}.

For reference purposes, let us denote the set of homotopy equivalence classes of maps between two spaces Ω\Omega and Ω\Omega^{\prime} by

[Ω,Ω][\Omega,\Omega^{\prime}]

according to the standard notation. Given a map f:ΩΩf:\Omega\to\Omega^{\prime}, we write [f][f] for its equivalence classs. We thus have, in this notation, that

[GX]=[GV],in[𝒱n1,Sn1].[G_{X}]=[G_{-\nabla V}],\text{in}\;[{\mathcal{V}}^{n-1},S^{n-1}].

Relations to the index

Since the spaces involved are of the same dimension and the target space is a sphere, we have, by the Hopf theory, that these homotopy equivalence classes are classified by degree.

Limit Cycles

In the case of a limit cycle γ\gamma, we saw that the Gauss map always has degree zero. Additional necessary conditions are obtained by examining the Gauss map in more detail.

Theorem 9.

Suppose γ\gamma is a limit cycle for the dynamics XX on n{\mathbb{R}}^{n}. then

  1. 1.

    For any ϵ>0{\epsilon}>0, there is a neighborhood Nδ(γ)N_{\delta}(\gamma) such that

    GX(Nδ(γ))Nϵ(GX(γ)).G_{X}(N_{\delta}(\gamma))\subset N_{{\epsilon}}(G_{X}(\gamma)).
  2. 2.

    The image GX(γ)G_{X}(\gamma) is not contained in any hemisphere: in other words, for any hyperplane 𝒫n{\mathcal{P}}\subset{\mathbb{R}}^{n}, GX(γ)𝒫G_{X}(\gamma)\cap{\mathcal{P}}\neq\emptyset. Moreover, for generic 𝒫{\mathcal{P}}, |GX(γ)𝒫||G_{X}(\gamma)\cap{\mathcal{P}}| is even (here the bars denote cadinality of a finite set.)

Proof.

The first part is proved by continuity and the long flow box (see [6].)

The second part is by contradiction: suppose there exists a hyperplane 𝒫a={vn;a(v)=0}{\mathcal{P}}_{a}=\{v\in{\mathbb{R}}^{n}\;;\;a(v)=0\}, for some a(n)a\in{({\mathbb{R}}^{n})}^{*} and is such that GX(γ)𝒫a=G_{X}(\gamma)\cap{\mathcal{P}}_{a}=\emptyset. Since any hyperplane separates Sn1S^{n-1} into two parts, we must have that a(GX(x))a(G_{X}(x)) is of uniform sign, say negative, for all xγx\in\gamma.

Choose a basis b1,,bnb_{1},\ldots,b_{n} of n{\mathbb{R}}^{n} such that aa is the dual basis vector of b1b_{1}, i.e. a(b1)=1a(b_{1})=1 and a(bi)=0a(b_{i})=0 for all i1i\neq 1. Write x1,,xnx_{1},\ldots,x_{n} for the coordinates in this basis.

Claim.

The function V(x)=12x12V(x)=\frac{1}{2}x_{1}^{2} is a Lyapunov function for XX in some open neighhborhood of γ\gamma.

This is shown by computing dVdt|γ\frac{dV}{dt}{|}_{\gamma}. We have

dVdt=(x1,0,,0)γ˙\frac{dV}{dt}=(x_{1},0,\ldots,0)\cdot\dot{\gamma}

and, since GX=X|X|G_{X}=\frac{X}{|X|}, this is just a(X)<0a(X)<0.

The claim now establishes a contradiction that proves the theorem, since GX(γ)G_{X}(\gamma) is a closed curve. The last part also follows from this fact and an elementary transversality argument. ∎

Theorem 9 says roughly that, even though the image of the Gauss map of a limt cycle is ‘thin,’ still it must curve sufficiently in the target sphere so as to intersect all possible hyperplanes.

As for the Lyapunov level sets near a limit cycle, we have

Theorem 10.

Suppose γ\gamma is a stable limit cycle for some controlled dynamics. then, on each level set of a Lyapunov function near γ\gamma, each direction (i.e. element of the unit sphere) appears at least twice, in other words, for each vSn1v\in S^{n-1},

|GV1(v)|2.|G_{-\nabla V}^{-1}(v)|\geq 2.

(The proof is a basic topolgical facts about tori and is omitted.) Thus, even though the Gauss map of the gradient vector field of Lyapunov functions is of degree zero on any level (as it should be by the Poincaré-Hopf theorem 4), it covers the unit sphere at least twice.

We see, therefore, that members of the same homotopy equivalence class of maps can have widely different Gauss images. The trick, as far as control is concerned, is to find a representative arising from a control section (see [3].)

5 Summary

We have presented ways of deriving collections of necessary conditions for achieving dynamics of a given type and we also pointed out the limitations of such topological conditions (due to the simplicity of the Hopf theory of maps to a sphere.) The basic aim of any analysis is, of course, to arrive at constructive methodologies. In the treatment of this subject in [3], we find that conditions that are both necessary and sufficient can be found for achieving dynamics in a certain class. In this light, the fundamental source of necessary conditions is the class of control-transverse sections and the resulting feedback-invariant dynamics.

References

  • [1] J. M. Coron. A necessary condition for feedback stabilization. System and Control Letters, 14:227–232, 1990.
  • [2] M. Greenberg and J. Harper. Algebraic topology: a first course. Benjamin/Cummins, Reading, Mass., 1981.
  • [3] E. Kappos. Global Controlled Dynamics. Draft Manuscript, 2007.
  • [4] S. Lang. Algebra. Addison-Wesley, Reading, Mass., 1971.
  • [5] John W. Milnor. Topology from a Differentiable Viewpoint. The University Press of Virginia, Charlottsville, 1965.
  • [6] J. Palis and W. de Melo. Geometric theory of dynamical systems: an introduction. Springer-Verlag, 1982.
  • [7] George W. Whitehead. Elements of homotopy theory, volume 61 of Graduate Texts in Mathematics. Springer-Verlag, 1978.