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

C0C^{0}-rigidity of Poisson diffeomorphisms

Dušan Joksimović
Abstract.

We prove the Poisson version of the Eliashberg-Gromov C0C^{0}-rigidity. More precisely, we prove that the group of Poisson diffeomorphisms is closed with respect to the C0C^{0} topology inside the group of all diffeomorphisms. The proof relies on the Poisson version of the energy-capacity inequality.

1. Introduction and main result

The famous C0C^{0}-rigidity theorem of Y. Eliashberg and M. Gromov states that the group of symplectic diffeomorphisms forms a closed subset of the group of all diffeomorphisms equipped with the C0C^{0} topology (i.e. the compact-open topology). This result led to definitions of symplectic homeomorphisms and topological symplectic manifolds, and is considered as the beginning of the subfield called C0C^{0}-symplectic geometry. Roughly, it investigates non-smooth symplectic objects and the behaviour of smooth symplectic objects with respect to the C0C^{0} topology. The aim of this article is to extend this philosophy to Poisson manifolds. As the main result, we will prove the Poisson analog of the Eliashberg-Gromov theorem.

A Poisson structure on a smooth manifold MM is a Lie bracket {,}\{\cdot,\cdot\} on the space C(M)C^{\infty}(M) which satisfies the Leibniz identity

(1) {fg,h}=f{g,h}+g{f,h},f,g,hC(M).\{fg,h\}=f\{g,h\}+g\{f,h\},\quad\forall f,g,h\in C^{\infty}(M).

Note that every manifold carries the trivial Poisson structure {,}0.\{\cdot,\cdot\}\equiv 0. Moreover, examples of Poisson manifolds include all symplectic manifolds and duals of Lie algebras.

Definition.

A Poisson diffeomorphism is a smooth diffeomorphism ψ:(M,{,})(M,{,})\psi:(M,\{\cdot,\cdot\})\rightarrow(M,\{\cdot,\cdot\}) that satisfies

(2) {f,g}ψ={fψ,gψ},\{f,g\}\circ\psi=\{f\circ\psi,g\circ\psi\},

for all f,gC(M).f,g\in C^{\infty}(M). We denote the group of all Poisson diffeomorphisms by Poiss(M,{,}).\operatorname{Poiss}(M,\{\cdot,\cdot\}).

Remark 1.

Notice that being a Poisson map is a local condition. Therefore to check that a diffeomorphism ψ:(M,{,})(M,{,})\psi:(M,\{\cdot,\cdot\})\rightarrow(M,\{\cdot,\cdot\}) is Poisson it is enough to verify condition (2) for all compactly supported functions ff and g.g. We will exploit this fact in the proof of Theorem 1.

Let (M,{,})(M,\{\cdot,\cdot\}) be a Poisson manifold and let fCc([0,1]×M)f\in C_{c}^{\infty}([0,1]\times M) be a compactly supported time-dependent Hamiltonian function. We define the (time-dependent) Hamiltonian vector field XftX_{f}^{t} associated to ff by

Xft:={ft,}𝔛(M),X_{f}^{t}:=\{f_{t},\cdot\}\in\mathfrak{X}(M),

where ft:=f(t,)C(M),f_{t}:=f(t,\cdot)\in C^{\infty}(M), t[0,1].t\in[0,1]. 111Note that the Leibniz identity (1) implies that {ft,}\{f_{t},\cdot\} is a derivation of C(M)C^{\infty}(M) and as such it defines a (time-dependent) vector field on M.M. The flow {φft}\{\varphi_{f}^{t}\} of XftX_{f}^{t} is called the Hamiltonian flow (or the Hamiltonian isotopy) generated by f.f. The Hamiltonian group of (M,{,})(M,\{\cdot,\cdot\}) is

Ham(M,{,}):={φf1|fCc([0,1]×M)}.\mathrm{Ham}(M,\{\cdot,\cdot\}):=\big{\{}\varphi^{1}_{f}\hskip 2.84526pt|\hskip 2.84526ptf\in C^{\infty}_{c}([0,1]\times M)\big{\}}.

The orbits of the standard action of Ham(M,{,})\operatorname{Ham}(M,\{\cdot,\cdot\}) on MM induce a foliation of MM which is called the symplectic foliation and its leaves are called symplectic leaves.

Although the condition of being a Poisson map includes the derivative of the map it turns out that Poisson diffeomorphisms behave nicely with respect to the C0C^{0} limits. The main result of this article is the following.

Theorem 1 (C0C^{0} rigidity of Poisson diffeomorphisms).

Let (M,{,})(M,\{\cdot,\cdot\}) be a Poisson manifold. Then Poiss(M,{,})\operatorname{Poiss}(M,\{\cdot,\cdot\}) is a closed subset of Diff(M)\operatorname{Diff}(M) with respect to the C0C^{0} topology.

Theorem 1 generalizes the Eliashberg-Gromov theorem to general Poisson manifolds, in the sense that any symplectic manifold (M,ω)(M,\omega) carries the natural Poisson structure {,}:={,}ω\{\cdot,\cdot\}:=\{\cdot,\cdot\}_{\omega} given by

{f1,f2}:=ω(Xf1,Xf2),f1,f2C(M),\{f_{1},f_{2}\}:=\omega(X_{f_{1}},X_{f_{2}}),\quad\forall f_{1},f_{2}\in C^{\infty}(M),

where XfiX_{f_{i}} are the corresponding Hamiltonian vector fields (i.e. unique vector fields that satisfy dfi=ω(Xfi,),df_{i}=\omega(X_{f_{i}},\cdot), i=1,2i=1,2). Poisson structures generalize symplectic structures by relaxing the non-degeneracy condition (which is algebraic), but still keeping the closedness condition222Note that for symplectic manifolds, the closedness of the symplectic form is equivalent with the Jacobi identity for the corresponding Poisson bracket. (which is geometric). Thus Theorem 1 shows that it is the closedness condition which is crucial for the C0C^{0}-rigidity rather than non-degeneracy.

It is known that in the symplectic setting the Poisson bracket exhibits C0C^{0}-rigid behaviour. The first result in this direction is due to F. Cardin and C. Viterbo [CV08] who proved that for all sequences fk,gkCc(M)f_{k},g_{k}\in C_{c}^{\infty}(M) which converge in the C0C^{0} topology to smooth functions ff and g,g, the following holds: if {fk,gk}C00\{f_{k},g_{k}\}\overset{C^{0}}{\longrightarrow}0 then {f,g}=0.\{f,g\}=0. This result was improved later by M. Entov, L. Polterovich, and F. Zapolsky [EP10, Zap07] where they proved that the Poisson bracket of a pair of functions is lower semicontinuous with respect to the C0C^{0} topology, while L. Buhovsky [Buh10] further improved this result by giving a sharp estimate on the rate of convergence. Sufficient conditions for {fk,gk}\{f_{k},g_{k}\} to converge to {f,g}\{f,g\} (in the C0C^{0} topology) provided that (fk,gk)C0(f,g),(f_{k},g_{k})\overset{C^{0}}{\longrightarrow}(f,g), were given by V. Humilière [Hum09] and M.-C. Arnaud [Arn15].

The proof of Theorem 1 exploits a general principle that many rigidity phenomena in symplectic geometry arise from the energy-capacity inequality (see e.g. [LM95]). More precisely, in the proof of Theorem 1 we use the following consequence of the energy-capacity inequality: if a sequence of Hamiltonian diffeomorphisms converges in both C0C^{0} and Hofer topology then the limits coincide (see Proposition 2 below). For the definition of the Hofer norm for Poisson structures we refer the reader to Section 2 or [JM21].

1.1. Poisson homeomorphisms and some open questions

The Eliashberg-Gromov theorem led to the definition of symplectic homeomorphisms as the C0C^{0} limits of symplectic diffeomorphisms. The behavior of symplectic homeomorphisms drew a lot of attention lately, as they play an important role in symplectic geometry and dynamics (see e.g. [CGHS20, Hum17, BO16, OM07] and references therein). Following the same principle we introduce the following.

Definition (Poisson homeomorphisms).

Let (M,{,})(M,\{\cdot,\cdot\}) be a Poisson manifold. A map φHomeo(M)\varphi\in\operatorname{Homeo}(M) is a Poisson homeomorphism if there exists a sequence φkPoiss(M,{,}),\varphi_{k}\in\operatorname{Poiss}(M,\{\cdot,\cdot\}), kk\in\mathbb{N} such that φkC0φ.\varphi_{k}\overset{C^{0}}{\longrightarrow}\varphi.

From Theorem 1 it follows that a Poisson homeomorphism which is also a diffeomorphism is a Poisson diffeomorphism.

The following general question naturally arises.

Question.

How much Poisson geometry do Poisson homeomorphisms remember?

There are many interesting instances of the above question. Here we will state just some of them.

Question 1 (Poisson homeomorphisms and symplectic leaves).

Do Poisson homeomorphisms map symplectic leaves to symplectic leaves homeomorphically? If so, does the restriction of a Poisson homeomorphism to a symplectic leaf induce a symplectic homeomorphism between the leaf and its image?

Another interesting problem is to understand how various classes of submanifolds behave with respect to Poisson homeomorphisms. It is known that symplectic homeomorphisms express very interesting behavior in this sense. For example, coisotropic submanifolds are preserved by symplectic homeomorphisms (see [HLS15] for a precise statement), while there is an example of a symplectic homeomorphism which maps a symplectic submanifold to an isotropic submanifold (see [BO16]).

Let (M,{,})(M,\{\cdot,\cdot\}) be a Poisson manifold and NM.N\subseteq M. The vanishing ideal of NN is

(N):={fC(M)|f(x)=0, xN}.\mathcal{I}(N):=\{f\in C^{\infty}(M)\hskip 2.84526pt|\hskip 2.84526ptf(x)=0,\text{ }\forall x\in N\}.

A submanifold N(M,{,})N\subseteq(M,\{\cdot,\cdot\}) is coisotropic if the vanishing ideal (N)\mathcal{I}(N) is a Lie subalgebra, i.e. for every f,g(N)f,g\in\mathcal{I}(N) it holds that {f,g}(N).\{f,g\}\in\mathcal{I}(N). Following [HLS15] one could ask the following.

Question 2 (Poisson coisotropic C0C^{0}-rigidity).

Let (M,{,})(M,\{\cdot,\cdot\}) be a Poisson manifold, φ\varphi be a Poisson homeomorphism, and NMN\subseteq M be a coisotropic submanifold such that φ(N)\varphi(N) is a smooth submanifold of M.M. Is φ(N)\varphi(N) coisotropic?

From an abstract point of view, Poisson manifolds can be seen as certain “quotients” of symplectic manifolds. In this sense it is natural to ask when Poisson homeomorphisms induce symplectic homeomorphisms of the associated symplectic manifold.

To make this more precise let (M,{,})(M,\{\cdot,\cdot\}) be a Poisson manifold. A symplectic realization of (M,{,})(M,\{\cdot,\cdot\}) is a pair ((S,ω),μ)((S,\omega),\mu) where (S,ω)(S,\omega) is a smooth symplectic manifold and μ:(S,ω)(M,{,})\mu:(S,\omega)\rightarrow(M,\{\cdot,\cdot\}) is a surjective submersion which is a Poisson map. In practice, constructed symplectic realizations (S,ω)(S,\omega) have additional structure of a (symplectic) Lie groupoid.

A Lie groupoid is a tuple (𝒢,M,t,s,m,u,i)(\mathcal{G},M,t,s,m,u,i) where 𝒢\mathcal{G} (the space of “arrows”) and MM (the space of “objects”) are smooth manifolds, s,t:𝒢Ms,t:\mathcal{G}\rightarrow M (the “source” and “target” maps) are smooth submersions, and u:M𝒢,i:𝒢G,m:𝒢(2)Gu:M\rightarrow\mathcal{G},i:\mathcal{G}\rightarrow G,m:\mathcal{G}^{(2)}\rightarrow G are smooth maps, where 𝒢(2)\mathcal{G}^{(2)} denotes the space of so-called “composable arrows”. For more details on (symplectic) Lie groupoids and symplectic realizations we refer to the book [CFM21].

Definition (Integrable Poisson manifolds).

A Poisson manifold (M,{,})(M,\{\cdot,\cdot\}) is called integrable if there exists a symplectic groupoid (Σ,Ω)(\Sigma,\Omega) such that ((Σ,Ω),t)((\Sigma,\Omega),t) is a symplectic realization of (M,{,})(M,\{\cdot,\cdot\}) where tt is the target map and such that the canonical embedding (using the unit map) u:MΣu:M\rightarrow\Sigma is a Lagrangian embedding.

It is known that if (M,{,})(M,\{\cdot,\cdot\}) is integrable then there is a unique integration (Σ,Ω)(\Sigma,\Omega) with simply-connected source-fibers and this one we call the canonical integration. For more details on the problem of integrability of Poisson manifolds we refer to [CF04, MX00, CFM21] and references therein.

Integration of Poisson manifolds can be seen as a generalization of the fact that for every smooth manifold there is a canonical way of associating a symplectic manifold by taking its cotangent bundle. It turns out that the cotangent bundles represent integrations of the trivial Poisson structures {,}0,\{\cdot,\cdot\}\equiv 0, where the target map tt is the canonical projection and the unit embedding uu is the canonical embedding MTMM\rightarrow T^{*}M as the zero-section. For more examples we refer to [CFM21, Chapters 12-14].

Assume that (M,{,})(M,\{\cdot,\cdot\}) is an integrable Poisson manifold and denote by ((Σ,Ω),t)((\Sigma,\Omega),t) the corresponding (canonical) integration. One could ask whether a Poisson homeomorphism of (M,{,})(M,\{\cdot,\cdot\}) induce a symplectic homeomorphisms on the integration (Σ,Ω).(\Sigma,\Omega).

Question 3.

Let φ\varphi be a Poisson homeomorphism (M,{,}).(M,\{\cdot,\cdot\}). Does there exist a symplectic homeomorphism φ~\widetilde{\varphi} of the symplectic groupoid (Σ,Ω)(\Sigma,\Omega) such that tφ=φ~tt\circ\varphi=\widetilde{\varphi}\circ t?

This question is already interesting for the trivial Poisson structure since it is not clear whether a homeomorphism of the base induce a symplectic homeomorphism of the cotangent bundle. Note that a diffeomorphism of the base always induces a (smooth) symplectomorphism of the cotangent bundle in the above sense. Namely, for every φDiff(M)\varphi\in\operatorname{Diff}(M) the map given by

TM(x,p)(φ(x),p(dφ1()))TMT^{*}M\ni(x,p)\mapsto\left(\varphi(x),p(d\varphi^{-1}(\cdot))\right)\in T^{*}M

is a symplectomorphism of TMT^{*}M with respect to the canonical symplectic structure.

Remark 2 (C0C^{0} rigidity of Lagrangian bisections).

Let (Σ,Ω)(\Sigma,\Omega) be a symplectic groupoid integrating a closed333compact and without boundary Poisson manifold (M,{,}).(M,\{\cdot,\cdot\}). A Lagrangian bisection is a section444Here by section we mean with respect to the target map, i.e. tb=idM:MMt\circ b=\operatorname{id}_{M}:M\rightarrow M b:MΣb:M\rightarrow\Sigma such that bΩ=0b^{*}\Omega=0 and such that sb:MMs\circ b:M\rightarrow M is a diffeomorphism where ss is the source map. It is not hard to check that for a given Lagrangian bisection bb the induced diffeomorphism sbs\circ b is a Poisson diffeomorphism. Such maps define a subgroup Γ(Σ,Ω)\Gamma(\Sigma,\Omega) of Poiss(M,{,}).\operatorname{Poiss}(M,\{\cdot,\cdot\}).

By the Laudenbach-Sikorav’s theorem [LS94] about the C0C^{0}-rigidity of Lagrangian embeddings it follows that the space of Lagrangian bisections is closed with respect to the C0C^{0} topology, and hence the induced group Γ(Σ,Ω)\Gamma(\Sigma,\Omega) is closed in the C0C^{0} topology inside Diff(M)\operatorname{Diff}(M). Note that in general Γ(Σ,Ω)Poiss(M,{,})\Gamma(\Sigma,\Omega)\subsetneq\operatorname{Poiss}(M,\{\cdot,\cdot\}) and hence Theorem 1 gives a stronger result.

On the other hand, C0C^{0} rigidity of Lagrangian bisections could carry some additional information (not necessarily related to the group Poiss(M,{,})\operatorname{Poiss}(M,\{\cdot,\cdot\})). For example, consider the case where (M,{,}0).(M,\{\cdot,\cdot\}\equiv 0). Then (Σ,Ω)=(TM,ωcan)(\Sigma,\Omega)=(T^{*}M,\omega_{can}) and the Lagrangian bisections are exactly closed 1-forms. Hence from the C0C^{0}-rigidity of Lagrangian bisections it follows that the space of closed differential 1-forms is closed inside the space of all differential 1-forms equipped with the C0C^{0} topology. Note that in this case Γ(Σ,Ω)={idM},\Gamma(\Sigma,\Omega)=\{\operatorname{id}_{M}\}, and therefore the C0C^{0} closedness of Γ(Σ,Ω)\Gamma(\Sigma,\Omega) trivially follows.

Acknowledgement

I would like to thank Ioan Mărcu\cbt for carefully reading and providing a very useful feedback on a preliminary version of the article and for suggesting Questions 1 and 3, and to Marius Crainic for an interesting discussion and suggesting Remark 2. The idea for the proof of Theorem 1 arose while I was preparing a talk for the UGC seminar at Utrecht University and therefore I would like to thank Fabian Ziltener and Álvaro del Pino Gómez for inviting me to give a talk.

The work on this project was funded by Agence Nationale de la Recherche through “ANR COSY: New challenges in contact and symplectic topology” grant (decision ANR-21-CE40-0002).

2. Proof of Theorem 1

First, we recall the definition of the Hofer metric on Ham(M,{,}).\operatorname{Ham}(M,\{\cdot,\cdot\}).

Let (M,{,})(M,\{\cdot,\cdot\}) be a Poisson manifold and fCc([0,1]×M).f\in C_{c}^{\infty}([0,1]\times M). The length of the Hamiltonian isotopy generated by ff is defined as

l(f):=01(supxMft(x)infxMft(x))𝑑t.l(f):=\int_{0}^{1}\big{(}\sup_{x\in M}f_{t}(x)-\inf_{x\in M}f_{t}(x)\big{)}\ dt.

Notice that, contrary to the symplectic case, the length of the Hamiltonian isotopy depends on the choice of a Hamiltonian function that generates the isotopy since Casimir functions need not be constant in general. We define a norm on Ham(M,{,})\mathrm{Ham}(M,\{\cdot,\cdot\}) by

(3) φHof:=inf{l(f)|fCc([0,1]×M),φf1=φ},||\varphi||_{Hof}:=\inf\big{\{}l(f)\ |\ f\in C^{\infty}_{c}([0,1]\times M),\ \varphi_{f}^{1}=\varphi\big{\}},

which we call the Hofer norm.

The Hofer norm was first introduced on symplectic manifolds. It is not hard to check that (3) defines a conjugation invariant pseudo-norm on Ham(M,{,}),\operatorname{Ham}(M,\{\cdot,\cdot\}), while the proof of the non-degeneracy relies on hard methods from symplectic topology.

In the symplectic case, H. Hofer [Hof90] proved non-degeneracy for the standard symplectic structure on 2n.\mathbb{R}^{2n}. Later L. Polterovich [Pol93] extended it to a larger class of symplectic manifolds, and F. Lalonde and D. McDuff [LM95] provided a proof for all symplectic manifolds.

In the Poisson setting the non-degeneracy of the Hofer norm was proven by I. Mărcu\cbt and the author in [JM21], reducing the setting to the symplectic case by restricting to a symplectic leaf. Before that, the non-degeneracy was known for Poisson manifolds whose symplectic leaves are closed embedded submanifolds due to D. Sun and Z. Zhang [SZ14]555Actually, in [SZ14] the non-degeneracy of the Hofer norm was claimed for regular Poisson manifolds, but in the proof they do not use regularity, but the assumption that the restriction of a compactly supported function to a leaf is compactly supported, however, without stating this explicitly., and for Poisson manifolds whose closed leaves form a dense set due to T. Rybicki [Ryb16].

We refer the reader to the book by L. Polterovich [Pol93] for a detailed overview on Hofer geometry and to [BIP08] for a more general discussion on the importance of conjugation-invariant norms on various symmetry groups.

The main ingredient of the proof of Theorem 1 is the following proposition which is the Poisson analog of [OM07, Prop. 3.6].

Proposition 2.

Let φkHam(M,{,}),\varphi_{k}\in\operatorname{Ham}(M,\{\cdot,\cdot\}), kk\in\mathbb{N} be a sequence such that φkHofφHam(M,{,})\varphi_{k}\overset{Hof}{\longrightarrow}\varphi\in\operatorname{Ham}(M,\{\cdot,\cdot\}) and φkC0ψHomeo(M).\varphi_{k}\overset{C^{0}}{\longrightarrow}\psi\in\operatorname{Homeo}(M). Then φ=ψ.\varphi=\psi.

The proof uses the Poisson version of the energy-capacity inequality: let φHam(M,{,})\varphi\in\operatorname{Ham}(M,\{\cdot,\cdot\}) and BLB\subseteq L be an open ball such that φ(B)B=.\varphi(B)\cap B=\emptyset. Then we have

(4) φHofe(B,L),||\varphi||_{Hof}\geq e(B,L),

where e(B,L)e(B,L) is the displacement energy of BB inside the symplectic manifold L.L. The proof follows from the standard energy-capacity inequality after restricting to a symplectic leaf, see [JM21] for the details.

Proof of Proposition 2.

Assume on the contrary that φψ.\varphi\neq\psi. Then there exists xMx\in M such that φ1ψ(x)x.\varphi^{-1}\psi(x)\neq x. Denote by LL the symplectic leaf through x.x. Then there exists an open ball BLB\subseteq L which contains xx and which is displaced by φ1ψ.\varphi^{-1}\psi. Hence there exists k0k_{0}\in\mathbb{N} such that for every kk0k\geq k_{0} it holds that

φ1φk(B)B=.\varphi^{-1}\varphi_{k}(B)\cap B=\emptyset.

Note that Hamiltonian isotopies preserve symplectic leaves and hence φ1φk(B)L.\varphi^{-1}\varphi_{k}(B)\subseteq L. Now, from the energy-capacity inequality (4) we have that

φ1φkHofe(B,L).||\varphi^{-1}\varphi_{k}||_{Hof}\geq e(B,L).

Since the displacement energy of an open set of a symplectic manifold is always positive, we get a contradiction with the fact that φ1φkHof0.||\varphi^{-1}\varphi_{k}||_{Hof}\rightarrow 0. Hence φ=ψ.\varphi=\psi. This completes the proof of Proposition 2. ∎

Corollary 3.

Let ψkPoiss(M,{,}),\psi_{k}\in\operatorname{Poiss}(M,\{\cdot,\cdot\}), kk\in\mathbb{N} be a sequence which converges in C0C^{0} topology to a smooth map ψDiff(M).\psi\in\operatorname{Diff}(M). Then for every fCc(M)f\in C_{c}^{\infty}(M) it holds that φfψt=ψ1φftψ,\varphi_{f\circ\psi}^{t}=\psi^{-1}\varphi_{f}^{t}\psi, t[0,1].\forall t\in[0,1].

Proof.

It is enough to prove the statement for t=1,t=1, then the other cases follow after rescaling Hamiltonians. Note that fψkC0fψf\circ\psi_{k}\overset{C^{0}}{\longrightarrow}f\circ\psi implies φfψk1Hofφfψ1.\varphi_{f\circ\psi_{k}}^{1}\overset{Hof}{\longrightarrow}\varphi_{f\circ\psi}^{1}. Namely, denoting F:=fψ,F:=f\circ\psi, Fk:=fψkF_{k}:=f\circ\psi_{k} we have that the Hamiltonian

F#F¯k:=FFkφFt=(FFk)φFt,F\#\bar{F}_{k}:=F-F_{k}\circ\varphi_{F}^{-t}=(F-F_{k})\circ\varphi_{F}^{-t},

generates the flow (φfψtφfψkt)t[0,1],(\varphi^{t}_{f\circ\psi}\circ\varphi_{f\circ\psi_{k}}^{-t})_{t\in[0,1]}, and therefore

φfψ1φfψk1Hofl(F#F¯k)2fψfψkk0.||\varphi^{1}_{f\circ\psi}\circ\varphi_{f\circ\psi_{k}}^{-1}||_{Hof}\leq l(F\#\bar{F}_{k})\leq 2||f\circ\psi-f\circ\psi_{k}||_{\infty}\overset{k\rightarrow\infty}{\longrightarrow}0.

On the other hand

φfψk1=ψk1φf1ψkC0ψ1φf1ψ.\varphi^{1}_{f\circ\psi_{k}}=\psi_{k}^{-1}\varphi^{1}_{f}\psi_{k}\overset{C^{0}}{\longrightarrow}\psi^{-1}\varphi^{1}_{f}\psi.

Now Corollary 3 follows from Proposition 2 applied to the sequence {φfψk1}k.\{\varphi^{1}_{f\circ\psi_{k}}\}_{k\in\mathbb{N}}.

We are now ready for the proof of the main result.

Proof of Theorem 1.

Let ψkPoiss(M,{,}),k\psi_{k}\in\operatorname{Poiss}(M,\{\cdot,\cdot\}),k\in\mathbb{N} be a sequence which converges in the C0C^{0} topology to a smooth map ψDiff(M).\psi\in\operatorname{Diff}(M). We will show that ψ\psi preserves the Poisson bracket of all compactly supported functions f,gCc(M),f,g\in C_{c}^{\infty}(M), see Remark 1.

Let f,gCc(M).f,g\in C_{c}^{\infty}(M). We define F,Fk,G,GkCc(M)F,F_{k},G,G_{k}\in C_{c}^{\infty}(M) by

Fk:=fψk,\displaystyle F_{k}:=f\circ\psi_{k}, F:=fψ,\displaystyle\quad F:=f\circ\psi,
Gk:=gψk,\displaystyle G_{k}:=g\circ\psi_{k}, G:=gψ.\displaystyle\quad G:=g\circ\psi.

Notice that FkC0F,F_{k}\overset{C^{0}}{\longrightarrow}F, GkC0G.G_{k}\overset{C^{0}}{\longrightarrow}G. From Corollary 3 we get that

(5) φGkt=ψk1φgtψkC0ψ1φgtψ=φGt,t[0,1].\varphi^{t}_{G_{k}}=\psi_{k}^{-1}\varphi_{g}^{t}\psi_{k}\overset{C^{0}}{\longrightarrow}\psi^{-1}\varphi_{g}^{t}\psi=\varphi_{G}^{t},\quad\forall t\in[0,1].

Therefore

FFφGt\displaystyle F-F\circ\varphi_{G}^{t} =limkFkFkφGkt\displaystyle=\lim_{k\rightarrow\infty}F_{k}-F_{k}\circ\varphi_{G_{k}}^{t}
=limk0t{Fk,Gk}φGks𝑑s\displaystyle=\lim_{k\rightarrow\infty}\int_{0}^{t}\{F_{k},G_{k}\}\circ\varphi_{G_{k}}^{s}ds
=limk0t{fψk,gψk}φGks𝑑s.\displaystyle=\lim_{k\rightarrow\infty}\int_{0}^{t}\{f\circ\psi_{k},g\circ\psi_{k}\}\circ\varphi_{G_{k}}^{s}ds.

Using that ψkPoiss(M,{,})\psi_{k}\in\operatorname{Poiss}(M,\{\cdot,\cdot\}) we get

FFφGt=limk0t{f,g}ψkφGks𝑑s,F-F\circ\varphi_{G}^{t}=\lim_{k\rightarrow\infty}\int_{0}^{t}\{f,g\}\circ\psi_{k}\circ\varphi_{G_{k}}^{s}ds,

and hence from (5) and the fact that ψkC0ψ\psi_{k}\overset{C^{0}}{\longrightarrow}\psi it follows that

(6) FFφGt=0t{f,g}ψφGs𝑑s.F-F\circ\varphi_{G}^{t}=\int_{0}^{t}\{f,g\}\circ\psi\circ\varphi_{G}^{s}ds.

On the other hand

(7) FFφGt=0t{F,G}φGs𝑑s=0t{fψ,gψ}φGs𝑑s.F-F\circ\varphi_{G}^{t}=\int_{0}^{t}\{F,G\}\circ\varphi_{G}^{s}ds=\int_{0}^{t}\{f\circ\psi,g\circ\psi\}\circ\varphi_{G}^{s}ds.

Differentiating both expressions (6) and (7) with respect to t,t, and setting t=0,t=0, we get that

{fψ,gψ}={f,g}ψ,\{f\circ\psi,g\circ\psi\}=\{f,g\}\circ\psi,

for all f,gCc(M).f,g\in C_{c}^{\infty}(M). This completes the proof of Theorem 1.

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

References

  • [Arn15] M.-C. Arnaud. Rigidity in topology C0C^{0} of the Poisson bracket for Tonelli Hamiltonians. Nonlinearity, 28(8):2731–2742, 2015.
  • [BIP08] Dmitri Burago, Sergei Ivanov, and Leonid Polterovich. Conjugation-invariant norms on groups of geometric origin. In Groups of diffeomorphisms, volume 52 of Adv. Stud. Pure Math., pages 221–250. Math. Soc. Japan, Tokyo, 2008.
  • [BO16] Lev Buhovsky and Emmanuel Opshtein. Some quantitative results in C0{C}^{0} symplectic geometry. Invent. Math., 205(1):1–56, 2016.
  • [Buh10] Lev Buhovsky. The 2/32/3-convergence rate for the Poisson bracket. Geom. Funct. Anal., 19(6):1620–1649, 2010.
  • [CF04] Marius Crainic and Rui Loja Fernandes. Integrability of Poisson brackets. J. Differential Geom., 66(1):71–137, 2004.
  • [CFM21] Marius Crainic, Rui Loja Fernandes, and Ioan Mărcuţ. Lectures on Poisson geometry, volume 217 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, [2021] ©2021.
  • [CGHS20] Dan Cristofaro-Gardiner, Vincent Humilière, and Sobhan Seyfaddini. Proof of the simplicity conjecture. arXiv:2001.01792, 2020.
  • [CV08] Franco Cardin and Claude Viterbo. Commuting Hamiltonians and Hamilton-Jacobi multi-time equations. Duke Math. J., 144(2):235–284, 2008.
  • [EP10] Michael Entov and Leonid Polterovich. C0C^{0}-rigidity of Poisson brackets. In Symplectic topology and measure preserving dynamical systems, volume 512 of Contemp. Math., pages 25–32. Amer. Math. Soc., Providence, RI, 2010.
  • [HLS15] Vincent Humilière, Rémi Leclercq, and Sobhan Seyfaddini. Coisotropic rigidity and C0C^{0}-symplectic geometry. Duke Math. J., 164(4):767–799, 2015.
  • [Hof90] Helmut Hofer. On the topological properties of symplectic maps. Proc. Roy. Soc. Edinburgh Sect. A, 115(1-2):25–38, 1990.
  • [Hum09] Vincent Humilière. Hamiltonian pseudo-representations. Comment. Math. Helv., 84(3):571–585, 2009.
  • [Hum17] Vincent Humilière. Géométrie symplectique C0C^{0} et sélecteurs d’action. Diplôme d’habilitation à diriger des recherches en mathématiques de l’Université Pierre et Marie Curie, 2017.
  • [JM21] Dušan Joksimović and Ioan Mărcuţ. Non-degeneracy of the Hofer norm for Poisson structures. J. Symplectic Geom., 19(5):1095–1100, 2021.
  • [LM95] François Lalonde and Dusa McDuff. The geometry of symplectic energy. Ann. of Math. (2), 141(2):349–371, 1995.
  • [LS94] F. Laudenbach and J.-C. Sikorav. Hamiltonian disjunction and limits of Lagrangian submanifolds. Internat. Math. Res. Notices, (4):161 ff., approx. 8 pp.  1994.
  • [MX00] Kirill C. H. Mackenzie and Ping Xu. Integration of Lie bialgebroids. Topology, 39(3):445–467, 2000.
  • [OM07] Yong-Geun Oh and Stefan Müller. The group of Hamiltonian homeomorphisms and C0C^{0}-symplectic topology. J. Symplectic Geom., 5(2):167–219, 2007.
  • [Pol93] Leonid Polterovich. Symplectic displacement energy for Lagrangian submanifolds. Ergodic Theory Dynam. Systems, 13(2):357–367, 1993.
  • [Ryb16] Tomasz Rybicki. On the existence of a Hofer type metric for Poisson manifolds. Internat. J. Math., 27(9):1650075, 16, 2016.
  • [SZ14] Dawei Sun and Zhenxing Zhang. A Hofer-type norm of Hamiltonian maps on regular Poisson manifold. J. Appl. Math., pages Art. ID 879196, 9, 2014.
  • [Zap07] Frol Zapolsky. Quasi-states and the Poisson bracket on surfaces. J. Mod. Dyn., 1(3):465–475, 2007.

Dušan Joksimović
Université Paris-Saclay, F-91405 Orsay Cedex, France
e-mail: [email protected]