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

Coisotropic rigidity and C0C^{0}–symplectic geometry

Vincent Humilière, Rémi Leclercq, Sobhan Seyfaddini VH: Institut de Mathématiques de Jussieu, Université Pierre et Marie Curie, 4 place Jussieu, 75005 Paris, France vincent.humiliere@imj-prg.fr RL: Université Paris-Sud, Département de Mathématiques, Bat. 425, 91405 Orsay Cedex, France remi.leclercq@math.u-psud.fr SS: Département de Mathématiques et Applications de l’École Normale Supérieure, 45 rue d’Ulm, F 75230 Paris cedex 05 sobhan.seyfaddini@ens.fr
(Date: September 5, 2025)
Abstract.

We prove that symplectic homeomorphisms, in the sense of the celebrated Gromov–Eliashberg Theorem, preserve coisotropic submanifolds and their characteristic foliations. This result generalizes the Gromov–Eliashberg Theorem and demonstrates that previous rigidity results (on Lagrangians by Laudenbach–Sikorav, and on characteristics of hypersurfaces by Opshtein) are manifestations of a single rigidity phenomenon. To prove the above, we establish a C0C^{0}–dynamical property of coisotropic submanifolds which generalizes a foundational theorem in C0C^{0}–Hamiltonian dynamics: Uniqueness of generators for continuous analogs of Hamiltonian flows.

Key words and phrases:
symplectic manifolds, coisotropic submanifolds, characteristic foliation, C0C^{0}–symplectic topology, spectral invariants
2010 Mathematics Subject Classification:
Primary 53D40; Secondary 37J05

1. Introduction and main results

A submanifold CC of a symplectic manifold (M,ω)(M,\omega) is called coisotropic if for all pCp\in C, (TpC)ωTpC(T_{p}C)^{\omega}\subset T_{p}C where (TpC)ω(T_{p}C)^{\omega} denotes the symplectic orthogonal of TpCT_{p}C. For instance, hypersurfaces and Lagrangians are coisotropic. A coisotropic submanifold carries a natural foliation \mathcal{F} which integrates the distribution (TC)ω(TC)^{\omega}; \mathcal{F} is called the characteristic foliation of CC. Coisotropic submanifolds and their characteristic foliations have been studied extensively in symplectic topology. The various rigidity properties that they exhibit have been of particular interest. For example, in [7] Ginzburg initiated a program for studying rigidity of coisotropic intersections. In this paper, we prove that coisotropic submanifolds, along with their characteristic foliations, are C0C^{0}–rigid in the spirit of the Gromov–Eliashberg Theorem.

This celebrated theorem states that a diffeomorphism which is a C0C^{0}–limit of symplectomorphisms is symplectic. Motivated by this, symplectic homeomorphisms are defined as C0C^{0}–limits of symplectomorphisms (see Definition 9 and Remark 9). Area preserving homeomorphisms, and their products, are examples of symplectic homeomorphisms. Here is our main result.

Theorem 1.

Let CC be a smooth coisotropic submanifold of a symplectic manifold (M,ω)(M,\omega). Let UU be an open subset of MM and θ:UV\theta\colon\thinspace U\rightarrow V be a symplectic homeomorphism. If θ(CU)\theta(C\cap U) is smooth, then it is coisotropic. Furthermore, θ\theta maps the characteristic foliation of CUC\cap U to that of θ(CU)\theta(C\cap U).

An important feature of the above theorem is its locality: CC is not assumed to be necessarily closed and θ\theta is not necessarily globally defined. Here is an immediate, but surprising, consequence of Theorem 1.

Corollary 2.

If the image of a coisotropic submanifold via a symplectic homeomorphism is smooth, then so is the image of its characteristic foliation.

Theorem 1 uncovers a link between two previous rigidity results and demonstrates that they are in fact extreme cases of a single rigidity phenomenon.

One extreme case, where CC is a hypersurface, was established by Opshtein [23]. Clearly, in this case, the interesting part is the assertion on rigidity of characteristics, as the first assertion is trivially true.

Lagrangians constitute the other extreme case. When CC is Lagrangian, its characteristic foliation consists of one leaf, CC itself. In this case the theorem reads: If θ\theta is a symplectic homeomorphism and θ(C)\theta(C) is smooth, then θ(C)\theta(C) is Lagrangian. In [11], Laudenbach–Sikorav proved a similar result: Let LL be a closed manifold and ιk\iota_{k} denote a sequence of Lagrangian embeddings L(M,ω)L\rightarrow(M,\omega) which C0C^{0}–converges to an embedding ι\iota. If ι(L)\iota(L) is smooth, then (under some technical assumptions) ι(L)\iota(L) is Lagrangian. On one hand, their result only requires convergence of embeddings while Theorem 1 requires convergence of symplectomorphisms. On the other hand, Theorem 1 is local: It does not require the Lagrangian nor the symplectic manifold to be closed.

The above discussion raises the following question. Question. What can one say about C0C^{0}–limits of coisotropic embeddings and their characteristic foliations?

We would like to point out that Theorem 1 is a coisotropic generalization of the Gromov–Eliashberg Theorem. Indeed, it implies that if the graph of a symplectic homeomorphism is smooth, then it is Lagrangian.

As we shall see, the proof of Theorem 1 relies on dynamical properties of coisotropic submanifolds. In particular, we use C0C^{0}–Hamiltonian dynamics as defined by Müller and Oh [22]. To the best of our knowledge, this is one of the first extrinsic applications of this recent, yet promising, theory.

Following [22], we call a path of homeomorphisms ϕt\phi^{t} a hameotopy if there exists a sequence of smooth Hamiltonian functions HkH_{k} such that the isotopies ϕHkt\phi^{t}_{H_{k}} C0C^{0}–converge to ϕ\phi and the Hamiltonians HkH_{k} C0C^{0}–converge to a continuous function HH (see Definition 11). Then, HH is said to generate the hameotopy ϕt\phi^{t}, and to emphasize this we write ϕHt\phi^{t}_{H}; the set of such generators will be denoted CHam0C^{0}_{\mathrm{Ham}}. A foundational result of C0C^{0}–Hamiltonian dynamics is the uniqueness of generators Theorem (see [26, 3]) which states that the trivial hameotopy, ϕt=Id,\phi^{t}=\mathrm{Id}, can only be generated by those functions in CHam0C^{0}_{\mathrm{Ham}} which solely depend on time (see also Corollary 4 below).

Let HC([0,1]×M)H\in C^{\infty}([0,1]\times M). Recall the following two dynamical properties of a coisotropic submanifold CC: Assume that CC is closed as a subset,

1. H|CH|_{C} is a function of time if and only if ϕH\phi_{H} (preserves CC and) flows along the characteristic foliation of CC. By flowing along characteristics we mean that for any point pCp\in C and any time t0t\geqslant 0, ϕHt(p)(p)\phi_{H}^{t}(p)\in\mathcal{F}(p), where (p)\mathcal{F}(p) stands for the characteristic leaf through pp.

2. For each pC,H|(p)p\in C,\;H|_{\mathcal{F}(p)} is a function of time if and only if the flow ϕH\phi_{H} preserves CC.

We will show that the above two properties hold for continuous Hamiltonians. The C0C^{0}–analog of the first property, stated below, plays an important role in the proof of Theorem 1.

Theorem 3.

Denote by CC a connected coisotropic submanifold of a symplectic manifold (M,ω)(M,\omega) which is closed as a subset111 It is our convention that submanifolds have no boundary. Note that a submanifold is closed as a subset if and only if it is properly embedded. of MM. Let HCHam0H\in C^{0}_{\mathrm{Ham}} with induced hameotopy ϕH\phi_{H}. The restriction of HH to CC is a function of time if and only if ϕH\phi_{H} preserves CC and flows along the leaves of its characteristic foliation.

This result answers a question raised by Buhovsky and Opshtein who asked if the above holds in the particular case where CC is a smooth hypersurface. It also drastically generalizes the aforementioned uniqueness of generators Theorem. Indeed, if CC is taken to be MM, then the characteristic foliation consists of the points of MM and the theorem follows immediately:

Corollary 4.

HCHam0H\in C^{0}_{\mathrm{Ham}} is a function of time if and only if ϕHt=Id\phi^{t}_{H}=\mathrm{Id}.

After the first draft of this article was written, we were asked by Opshtein if the second of the aforementioned properties holds for C0C^{0} Hamiltonians. Our next result provides an affirmative answer to Opshtein’s question.

Theorem 5.

Denote by CC a connected coisotropic submanifold of a symplectic manifold (M,ω)(M,\omega) which is closed as a subset of MM. Let HCHam0H\in C^{0}_{\mathrm{Ham}} with induced hameotopy ϕH\phi_{H}. The restriction of HH to each leaf of the characteristic foliation of CC is a function of time if and only if the flow ϕH\phi_{H} preserves CC.

When CC is a Lagrangian, Theorems 3 and 5 coincide and both state that: The restriction of HH to LL is a function of time if and only if ϕHt(L)=L\phi_{H}^{t}(L)=L for all tt. In an interesting manifestation of Weinstein’s creed, “Everything is a Lagrangian submanifold!”, the general case of Theorems 1, 3 and 5 will be essentially deduced from the a priori particular case of Lagrangians.

The results of this paper establish C0C^{0}–rigidity of coisotropic submanifolds together with their characteristic foliations. It would be interesting to see if isotropic or symplectic submanifolds exhibit similar rigidity properties: If a smooth submanifold is the image of an isotropic (respectively symplectic) submanifold under a symplectic homeomorphism, is it isotropic (respectively symplectic)? Note that if in these questions one considers, instead of symplectic homeomorphisms, C0C^{0}–limits of isotropic (respectively symplectic) embeddings then Gromov’s results on the hh-principle provide negative answers in general. In short, isotropic and symplectic embeddings are not C0C^{0}–rigid. (See [9, Section 3.4.2], or [5, Theorems 12.1.1 and 12.4.1].)

Defining C0C^{0}–coisotropic submanifolds

As we will see in Section 6, an interesting feature of Theorem 1 is that it allows us to define C0C^{0}–coisotropic submanifolds along with their C0C^{0}–characteristic foliations. Roughly speaking, a C0C^{0}–coisotropic will be defined to be a C0C^{0}–submanifold of a symplectic manifold which is locally symplectic homeomorphic to a smooth coisotropic. The well-definedness of this notion is a consequence of Theorem 1. Furthermore, from the same theorem we conclude that a C0C^{0}–coisotropic submanifold admits a unique C0C^{0}–foliation which will be referred to as its C0C^{0}–characteristic foliation.

As a consequence, we obtain a definition for C0C^{0}–Lagrangian submanifolds as C0C^{0}–coisotropic submanifolds of dimension nn. Graphs of symplectic homeomorphisms and graphs of C0C^{0} 1–forms, closed in the sense of distributions, constitute examples of C0C^{0}–Lagrangians. For further details, we refer the interested reader to Section 6.

Main tools: Lagrangian spectral invariants

In order to prove the main results, we use the theory of Lagrangian spectral invariants. One consequence of this theory is the existence of the spectral distance γ\gamma on the space of Lagrangians Hamiltonian isotopic to the 0–section in cotangent bundles introduced by Viterbo in [25].222One of the main features of γ\gamma is that it is bounded from above by Hofer’s distance on Lagrangians. In particular, Lemmas 6 and 7 also hold with γ\gamma replaced by Hofer’s distance.

More precisely, we establish inequalities comparing γ\gamma to a capacity recently defined by Lisi–Rieser [15]. This capacity, which we denote by cLRc_{\mathrm{LR}}, is a relative (to a fixed Lagrangian) version of the Hofer–Zehnder capacity. We will now define cLRc_{\mathrm{LR}}. Fix a Lagrangian LL. Recall that a Hamiltonian chord of a Hamiltonian HH, of length TT, is a path x:[0,T]Mx\colon\thinspace[0,T]\rightarrow M such that x(0)x(0), x(T)Lx(T)\in L and for all t[0,T]t\in[0,T], x˙(t)=XHt(x(t))\dot{x}(t)=X_{H}^{t}(x(t)). A Hamiltonian is said to be LL–slow if all of its Hamiltonian chords of length at most 1 are constant. We denote by (U)\mathcal{H}(U) the set of admissible Hamiltonians, that is, smooth time-independent functions with compact support included in UU, which are non-negative and reach their maximum at a point of LL. For an open set UU which intersects LL, the relative capacity of UU with respect to LL is defined as

cLR(U;L)=sup{maxf|f(U) L–slow}.c_{\mathrm{LR}}(U;L)=\sup\{\max f\,|\,f\in\mathcal{H}(U)\text{ $L$--slow}\}\,.

For instance, if BB is the open ball of radius rr in 2n\mathbb{R}^{2n} and 0=n×{0}\mathcal{L}_{0}=\mathbb{R}^{n}\times\{0\}, then cLR(B;0)=πr22c_{\mathrm{LR}}(B;\mathcal{L}_{0})=\frac{\pi r^{2}}{2}; see [15].

In what follows, we denote by L0L_{0} the 0–section of TLT^{*}L. The first energy-capacity inequality used in this paper is the following:

Lemma 6.

Let LL be a smooth closed manifold and UU_{-} and U+U_{+} be open subsets of TLT^{*}L, so that U±L0U_{\pm}\cap L_{0}\neq\emptyset. If a compactly supported Hamiltonian HH satisfies H|U±=±C±H|_{U_{\pm}}=\pm C_{\pm} with C±C_{\pm}\in\mathbb{R} so that C±>cLR(U±;L0)C_{\pm}>c_{\mathrm{LR}}(U_{\pm};L_{0}), then γ(ϕH1(L0),L0)min{cLR(U;L0),cLR(U+;L0)}\gamma(\phi_{H}^{1}(L_{0}),L_{0})\geqslant\min\{c_{\mathrm{LR}}(U_{-};L_{0}),c_{\mathrm{LR}}(U_{+};L_{0})\}.

This is the Lagrangian analog of the energy-capacity inequality proven for the Hamiltonian spectral distance in [10, Corollary 12]. Then, as in [10], we will derive a similar inequality for Hamiltonians (not necessarily constant but) with controlled oscillations on U±U_{\pm}, see Corollary 14.

Lemma 6 can also be established on compact manifolds for weakly exact Lagrangians via Leclercq [12] and for monotone Lagrangians via Leclercq–Zapolsky [13].

The second energy-capacity inequality is due to Lisi–Rieser [15]. This is a relative version of the standard energy-capacity inequality, see for example Viterbo [25].

Lemma 7.

Let LL be a smooth closed manifold. Suppose that UU is an open subset of TLT^{*}L, with L0UL_{0}\cap U\neq\emptyset. Assume that LL^{\prime} is a Lagrangian Hamiltonian isotopic to L0L_{0} such that LU=L^{\prime}\cap U=\emptyset. Then γ(L,L0)cLR(U;L0)\gamma(L^{\prime},L_{0})\geqslant c_{\mathrm{LR}}(U;L_{0}).

A special case of this specific inequality appears in Barraud–Cornea [1] and Charette [4]. A similar inequality is worked out in Borman–McLean [2].

Finally, we will need an inequality which provides an upper bound for the spectral distance. Let gg denote a Riemannian metric on a closed manifold LL and denote by TrL={(q,p)TL|pgr}T^{*}_{r}L=\{(q,p)\in T^{*}L\,|\,\|p\|_{g}\leqslant r\} the cotangent ball bundle of radius rr. Suppose that ϕHt(L0)TrL\phi^{t}_{H}(L_{0})\subset T^{*}_{r}L for all t[0,1]t\in[0,1]. Viterbo has conjectured [27] that there exists a constant C>0C>0, depending on gg, such that γ(ϕH1(L0),L0)Cr\gamma(\phi^{1}_{H}(L_{0}),L_{0})\leqslant Cr. This conjecture has many important ramifications; see [18, 27]. Lemma 8 below is a special case of Viterbo’s conjecture; a more precise version of the lemma appears in [21, Theorem 9.7].

Lemma 8.

Let LL be a smooth closed manifold, 𝒱\mathcal{V} a proper open subset of LL, and V=π1(𝒱)TLV=\pi^{-1}(\mathcal{V})\subset T^{*}L, where π:TLL\pi\colon\thinspace T^{*}L\rightarrow L is the standard projection. There exists C>0C>0, depending on the set 𝒱\mathcal{V}, such that: For all r>0r>0, if HH is a smooth, compactly supported Hamiltonian on TLT^{*}L such that H|V=0H|_{V}=0, and ϕHt(L0)TrL\phi^{t}_{H}(L_{0})\subset T^{*}_{r}L for all t[0,1]t\in[0,1] then γ(ϕH1(L0),L0)Cr.\gamma(\phi^{1}_{H}(L_{0}),L_{0})\leqslant Cr.

Organization of the paper

In Section 2, we review the preliminaries on C0C^{0}–Hamiltonian dynamics and Lagrangian spectral invariants. In Section 3, we prove energy-capacity inequalities (Lemmas 6 and 7) as well as the upper bound on the spectral distance (Lemma 8). In Section 4, we use these inequalities in order to prove localized versions of Theorem 3 in the special case of Lagrangians. In Section 5, we prove Theorems 1, 3, and 5 using the results of Section 4. In Section 6, we define C0C^{0}–coisotropic submanifolds and their characteristic foliations. In the same section, we provide examples of such C0C^{0}–objects.

In Appendix A, we provide relatively simple, and hopefully enlightening, proofs of Theorems 1 and 3 in the special case of closed Lagrangians in cotangent bundles. We hope that this appendix will give the reader an idea of the proofs of the main results while avoiding the technicalities of Sections 4 and 5.

Aknowledgements

We thank Samuel Lisi and Tony Rieser for sharing their work with us before it was completed and for related discussions. The inspiration for this paper came partly from an unpublished work by Lev Buhovsky and Emmanuel Opshtein, whom we also thank. We are especially grateful to Emmanuel Opshtein for generously sharing his ideas and insights with us through many stimulating discussions. We are also grateful to Alan Weinstein for interesting questions and suggestions. Lastly we would like to thank the anonymous referees for pointing out several inaccuracies and for many helpful suggestions which have improved the exposition.

This work is partially supported by the grant ANR-11-JS01-010-01. The research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP/2007-2013) / ERC Grant Agreement 307062.

2. Preliminaries

2.1. Symplectic and Hamiltonian homeomorphisms

In this section we give precise definitions for symplectic and Hamiltonian homeomorphisms and recall a few basic properties of the theory of continuous Hamiltonian dynamics developed by M ller and Oh [22].

Given two manifolds M1M_{1}, M2M_{2}, a compact subset KM1K\subset M_{1}, a Riemannian distance dd on M2M_{2}, and two maps f,g:M1M2f,g\colon\thinspace M_{1}\to M_{2}, we denote

dK(f,g)=supxKd(f(x),g(x)).d_{K}(f,g)=\sup_{x\in K}d(f(x),g(x)).

We say that a sequence of maps fi:M1M2f_{i}\colon\thinspace M_{1}\to M_{2} C0C^{0}–converges to some map f:M1M2f\colon\thinspace M_{1}\to M_{2}, if for every compact subset KM1K\subset M_{1}, the sequence dK(fi,f)d_{K}(f_{i},f) converges to 0. This notion does not depend on the choice of the Riemannian metric.

Definition 9.

Let (M1,ω1)(M_{1},\omega_{1}) and (M2,ω2)(M_{2},\omega_{2}) be symplectic manifolds. A continuous map θ:UM2\theta\colon\thinspace U\to M_{2}, where UM1U\subset M_{1} is open, is called symplectic if it is the C0C^{0}–limit of a sequence of symplectic diffeomorphisms θi:Uθi(U)\theta_{i}\colon\thinspace U\to\theta_{i}(U).

Let U1M1U_{1}\subset M_{1} and U2M2U_{2}\subset M_{2} be open subsets. If a homeomorphism θ:U1U2\theta\colon\thinspace U_{1}\to U_{2} and its inverse θ1\theta^{-1} are both symplectic maps, we call θ\theta a symplectic homeomorphism.

Clearly, if θ\theta is a symplectic homeomorphism, so is θ1\theta^{-1}. By the Gromov–Eliashberg Theorem a symplectic homeomorphism which is smooth is a symplectic diffeomorphism.

Remark 10. More generally, one can define a symplectic homeomorphism to be a homeomorphism which is locally a C0C^{0}–limit of symplectic diffeomorphisms. For simplicity, we do not use this more general definition, however it is evident from the proof of Theorem 1 that it does hold for such homeomorphisms as well.

We now turn to the definition of Hamiltonian homeomorphisms (called hameomorphisms) introduced by M ller and Oh [22].

Definition 11.

Let (M,ω)(M,\omega) be a symplectic manifold and II\subset\mathbb{R} an interval. An isotopy (ϕt)tI(\phi^{t})_{t\in I} is called a hameotopy if there exist a compact subset KMK\subset M and a sequence of smooth Hamiltonians HiH_{i} supported in KK such that:

  1. (1)

    The sequence of flows ϕHit\phi_{H_{i}}^{t} C0C^{0}–converges to ϕt\phi^{t} uniformly in tt on every compact subset of II,

  2. (2)

    the sequence Hi(t,)H_{i}(t,\cdot) C0C^{0}–converges to a continuous function H(t,)H(t,\cdot) uniformly in tt on every compact subset of II.

We say that HH generates ϕt\phi^{t}, denote ϕt=ϕHt\phi^{t}=\phi_{H}^{t}, and call HH a continuous Hamiltonian. We denote by CHam0(M,ω)C^{0}_{\mathrm{Ham}}(M,\omega) (or just CHam0C^{0}_{\mathrm{Ham}}) the set of all continuous functions H:[0,1]×MH\colon\thinspace[0,1]\times M\to\mathbb{R} which generate hameotopies parametrized by [0,1][0,1]. A homeomorphism is called a hameomorphism if it is the time–11 map of some hameotopy parametrized by [0,1][0,1].

A continuous function HCHam0H\in C^{0}_{\mathrm{Ham}} generates a unique hameotopy [22]. Conversely, Viterbo [26] and Buhovsky–Seyfaddini [3] proved that a hameotopy has a unique (up to addition of a function of time) continuous generator.

One can easily check that generators of hameotopies satisfy the same composition formulas as their smooth counterparts. Namely, if ϕHt\phi_{H}^{t} is a hameotopy, then (ϕHt)1(\phi_{H}^{t})^{-1} is a hameotopy generated by H(t,ϕHt(x))-H(t,\phi^{t}_{H}(x)); given another hameotopy ϕKt\phi_{K}^{t}, the isotopy ϕHtϕKt\phi_{H}^{t}\phi_{K}^{t} is also a hameotopy, generated by H(t,x)+K(t,(ϕHt)1(x))H(t,x)+K(t,(\phi^{t}_{H})^{-1}(x)).

Moreover, we will repeatedly use the following simple fact: If HCHam0(V)H\in C^{0}_{\mathrm{Ham}}(V) for some open set VV in a symplectic manifold (M,ω)(M,\omega) and if θ:UV\theta\colon\thinspace U\to V is a symplectic homeomorphism, then HθH\circ\theta belongs to CHam0(U)C^{0}_{\mathrm{Ham}}(U) and generates the hameotopy θ1ϕHtθ\theta^{-1}\phi_{H}^{t}\theta. This, in particular, holds for smooth H:[0,1]×MH\colon\thinspace[0,1]\times M\to\mathbb{R} supported in VV.

2.2. Lagrangian spectral invariants

In [25], Viterbo defined Lagrangian spectral invariants on 2n\mathbb{R}^{2n} and cotangent bundles via generating functions. Then Oh [19] defined similar invariants via Lagrangian Floer homology in cotangent bundles which have been proven to coincide with Viterbo’s invariants by Milinković [16]. They have been adapted to the compact case by Leclercq [12] for weakly exact Lagrangians and Leclercq–Zapolsky [13] for monotone Lagrangians. However, for the type of problems which we consider here (C0C^{0}–convergence of Lagrangians), we can restrict ourselves to Weinstein neighborhoods and thus work only in cotangent bundles. We briefly outline below the construction of these invariants via Lagrangian Floer homology and collect their main properties in this situation. We refer to Monzner–Vichery–Zapolsky [18] which gives a very nice exposition of the theory.

Let LL be a smooth compact manifold, L0L_{0} denote the 0–section in TLT^{*}L, and λ\lambda the Liouville 1–form. To a compactly supported smooth time-dependent Hamiltonian HCc([0,1]×TL)H\in C^{\infty}_{c}([0,1]\times T^{*}L) is associated the action functional

𝒜H:Ω(TL),γ01Ht(γ(t))𝑑tγλ\displaystyle\mathcal{A}_{H}\colon\thinspace\Omega(T^{*}L)\rightarrow\mathbb{R}\;,\quad\gamma\mapsto\int_{0}^{1}H_{t}(\gamma(t))\,dt-\int\gamma^{*}\lambda

where Ω(TL)={γ:[0,1]TL|γ(0)L0,γ(1)L0}\Omega(T^{*}L)=\{\gamma\colon\thinspace[0,1]\rightarrow T^{*}L\,|\,\gamma(0)\in L_{0},\;\gamma(1)\in L_{0}\}. The critical points of 𝒜H\mathcal{A}_{H} are the chords of the Hamiltonian vector field XHX_{H} which start and end on L0L_{0}. The spectrum of 𝒜H\mathcal{A}_{H}, denoted by Spec(𝒜H)\mathrm{Spec}(\mathcal{A}_{H}), consists of the critical values of 𝒜H\mathcal{A}_{H}. It is a nowhere dense subset of \mathbb{R} which only depends on the time–1 map ϕH1\phi_{H}^{1}, hence we sometimes denote it by Spec(ϕH1)\mathrm{Spec}(\phi_{H}^{1}).

Following Floer’s construction, for a generic choice of Hamiltonian function, crit(𝒜H)\mathrm{crit}(\mathcal{A}_{H}) is finite and one can form a chain complex (CF(H),H,J)(CF_{*}(H),\partial_{H,J}) whose generators are the critical chords and whose differential counts the elements of the 0–dimensional component of moduli spaces of Floer trajectories (i.e pseudo-holomorphic curves perturbed by HH) which run between the critical chords (with boundary conditions on L0L_{0}). The differential relies on the additional data of a generic enough almost complex structure, JJ.

This complex is filtered by the values of the action, that is, for aa\in\mathbb{R} a regular value of 𝒜H\mathcal{A}_{H}, one can consider only chords of action less than aa. Such chords generate a subcomplex of the total complex CFa(H)CF^{a}_{*}(H) (because the action decreases along Floer trajectories). We denote by iai^{a}_{*} the inclusion CFa(H)CF(H)CF^{a}_{*}(H)\rightarrow CF_{*}(H). By considering homotopies between pairs (H,J)(H,J) and (H,J)(H^{\prime},J^{\prime}), one can canonically identify the homology induced by the respective Floer complexes H(CF(H),H,J)H_{*}(CF(H),\partial_{H,J}) and H(CF(H),H,J)H_{*}(CF(H^{\prime}),\partial_{H^{\prime},J^{\prime}}) and by considering C2C^{2}–small enough Hamiltonian functions, one can see that the resulting object HF(L0)HF_{*}(L_{0}) is canonically isomorphic to the singular homology of LL.

Thus, one can consider spectral invariants associated to any non-zero homology class α\alpha of LL, defined as the smallest action level which detects α\alpha:

(α;H)=inf{a|αim(H(ia))}\displaystyle\ell(\alpha;H)=\inf\{a\in\mathbb{R}\,|\,\alpha\in\mathrm{im}(H_{*}(i^{a}))\}

In what follows we will only be interested in the spectral invariants associated to the class of a point and the fundamental class which will be respectively denoted by (H)=([pt];H)\ell_{-}(H)=\ell([\mathrm{pt}];H) and +(H)=([L];H)\ell_{+}(H)=\ell([L];H).

These invariants were proven to be continuous with respect to the C0C^{0}–norm on Hamiltonian functions so that they are defined for any (not necessarily generic) Hamiltonian. Moreover, they only depend on the time–1 map ϕH1\phi^{1}_{H} induced by the flow of HH; hence they are well-defined on Hamc(TL,dλ).\mathrm{Ham}^{c}(T^{*}L,d\lambda).

Their main properties are collected in the following theorem, which corresponds to [18, Theorem 2.20], except for (7) which we prove below. Note that, except for (6) and (7), these properties already appear in Viterbo [25].

Theorem 12.

Let LL be a smooth closed connected manifold. Let L0L_{0} denote the 0–section of TLT^{*}L. There exist two maps ±:Hamc(TL,dλ)\ell_{\pm}\colon\thinspace\mathrm{Ham}^{c}(T^{*}L,d\lambda)\rightarrow\mathbb{R} with the following properties:

  1. (1)

    For any ϕHamc(TL,dλ)\phi\in\mathrm{Ham}^{c}(T^{*}L,d\lambda), ±(ϕ)\ell_{\pm}(\phi) lie in Spec(ϕ)\mathrm{Spec}(\phi).

  2. (2)

    +\ell_{-}\leqslant\ell_{+}.

  3. (3)

    For any two Hamiltonian functions HH and KK,
    01min(HtKt)𝑑t(ϕH1)(ϕK1)01max(HtKt)𝑑t\int_{0}^{1}\min(H_{t}-K_{t})\,dt\leqslant\ell_{-}(\phi^{1}_{H})-\ell_{-}(\phi^{1}_{K})\leqslant\int_{0}^{1}\max(H_{t}-K_{t})\,dt, and 01min(HtKt)𝑑t+(ϕH1)+(ϕK1)01max(HtKt)𝑑t\int_{0}^{1}\min(H_{t}-K_{t})\,dt\leqslant\ell_{+}(\phi^{1}_{H})-\ell_{+}(\phi^{1}_{K})\leqslant\int_{0}^{1}\max(H_{t}-K_{t})\,dt.

  4. (4)

    For any ϕ\phi and ϕHamc(TL,dλ)\phi^{\prime}\in\mathrm{Ham}^{c}(T^{*}L,d\lambda), +(ϕϕ)+(ϕ)++(ϕ)\ell_{+}(\phi\phi^{\prime})\leqslant\ell_{+}(\phi)+\ell_{+}(\phi^{\prime}).

  5. (5)

    For any ϕHamc(TL,dλ)\phi\in\mathrm{Ham}^{c}(T^{*}L,d\lambda), ±(ϕ)=(ϕ1)\ell_{\pm}(\phi)=-\ell_{\mp}(\phi^{-1}).

  6. (6)

    If H|L0cH|_{L_{0}}\leqslant c (respectively H|L0cH|_{L_{0}}\geqslant c or H|L0=cH|_{L_{0}}=c), then ±(ϕH1)c\ell_{\pm}(\phi^{1}_{H})\leqslant c (respectively ±(ϕH1)c\ell_{\pm}(\phi^{1}_{H})\geqslant c or ±(ϕH1)=c\ell_{\pm}(\phi^{1}_{H})=c).

  7. (7)

    If ff is a L0L_{0}–slow admissible Hamiltonian, then +(ϕf1)=max(f|L0)\ell_{+}(\phi^{1}_{f})=\max(f|_{L_{0}}) and (ϕf1)=0\ell_{-}(\phi_{f}^{1})=0.

  8. (8)

    For any ϕ\phi and ϕHamc(TL,dλ)\phi^{\prime}\in\mathrm{Ham}^{c}(T^{*}L,d\lambda) such that ϕ(L0)=ϕ(L0)\phi(L_{0})=\phi^{\prime}(L_{0}), +(ϕ)(ϕ)=+(ϕ)(ϕ)\ell_{+}(\phi)-\ell_{-}(\phi)=\ell_{+}(\phi^{\prime})-\ell_{-}(\phi^{\prime}).

Proof of item (7).

Since ff is L0L_{0}–slow, Spec(ϕf1)\mathrm{Spec}(\phi_{f}^{1}) consists of critical values of ff corresponding to critical points lying in L0L_{0}. Now for all s[0,1]s\in[0,1], since ff is autonomous, Spec(ϕfs)=Spec(ϕsf1)=sSpec(ϕf1)\mathrm{Spec}(\phi_{f}^{s})=\mathrm{Spec}(\phi_{sf}^{1})=s\cdot\mathrm{Spec}(\phi_{f}^{1}). Since in cotangent bundles spectral invariants lie in the spectrum regardless of degeneracy of ff, by continuity of ±\ell_{\pm} there exist p±crit(f)L0p_{\pm}\in\mathrm{crit}(f)\cap L_{0} such that ±(ϕfs)=sf(p±)\ell_{\pm}(\phi_{f}^{s})=s\cdot f(p_{\pm}). Now we claim that for small times ss, +(ϕsf1)\ell_{+}(\phi_{sf}^{1}) (respectively (ϕsf1)\ell_{-}(\phi_{sf}^{1})) is the maximum (respectively the minimum) of sfsf so that f(p+)=max(f)f(p_{+})=\max(f) (respectively f(p)=min(f)f(p_{-})=\min(f)) which concludes the proof.

It remains to prove the claim. Since ff is autonomous, the path of Lagrangians it generates is a geodesic with respect to Hofer’s distance for small times, see Milinković [17, Theorem 8]. That is, for s(0,1)s\in(0,1) small enough

(1) max(sf)min(sf)=length({Lt}t[0,s])=dHof(L0,Ls)\displaystyle\max(sf)-\min(sf)=\mathrm{length}(\{L_{t}\}_{t\in[0,s]})=d_{\mathrm{Hof}}(L_{0},L_{s})

with Lt=ϕft(L0)L_{t}=\phi_{f}^{t}(L_{0}). (As ff reaches its extrema on L0L_{0}, the extrema of sfsf over TLT^{*}L and L0L_{0} coincide and we remove them from the notation.) On the other hand, by choosing ss small enough we can ensure that Ls=ΓdSL_{s}=\Gamma_{\!dS}, the graph of the differential of some smooth function S:LS\colon\thinspace L\rightarrow\mathbb{R}. Now in general dHof(L0,ΓdS)osc(S)d_{\mathrm{Hof}}(L_{0},\Gamma_{\!dS})\leqslant\mathrm{osc}(S) and by [17, Proposition 1.(6)]: osc(S)=+(ϕS~1)(ϕS~1)\mathrm{osc}(S)=\ell_{+}(\phi_{\tilde{S}}^{1})-\ell_{-}(\phi_{\tilde{S}}^{1}) where S~\tilde{S} lifts SS to TLT^{*}L by pullback via the natural projection and cutoff far from the Lagrangian isotopy. Since ϕS~1(L0)=ΓdS\phi_{\tilde{S}}^{1}(L_{0})=\Gamma_{\!dS}, by Property (8), we can replace ϕS~1\phi_{\tilde{S}}^{1} by ϕsf1\phi_{sf}^{1} thus (1) leads to the inequality max(sf)min(sf)+(ϕsf1)(ϕsf1)\max(sf)-\min(sf)\leqslant\ell_{+}(\phi_{sf}^{1})-\ell_{-}(\phi_{sf}^{1}). Finally, since min(sf)(ϕsf1)+(ϕsf1)max(sf)\min(sf)\leqslant\ell_{-}(\phi_{sf}^{1})\leqslant\ell_{+}(\phi_{sf}^{1})\leqslant\max(sf), we deduce that (ϕsf1)=min(sf)\ell_{-}(\phi_{sf}^{1})=\min(sf) and +(ϕsf1)=max(sf)\ell_{+}(\phi_{sf}^{1})=\max(sf). ∎

In view of Property (8), Viterbo (followed by Oh) derived an invariant γ\gamma of Lagrangians Hamiltonian isotopic to the 0–section, defined as follows.

Definition 13.

For a Hamiltonian diffeomorphism ϕHamc(TL,dλ)\phi\in\mathrm{Ham}^{c}(T^{*}L,d\lambda) we set γ(ϕ)=+(ϕ)(ϕ)\gamma(\phi)=\ell_{+}(\phi)-\ell_{-}(\phi) and for a Lagrangian LL Hamiltonian isotopic to the 0–section, we set γ(L,L0)=γ(ϕ)\gamma(L,L_{0})=\gamma(\phi) for any ϕHamc(TL,dλ)\phi\in\mathrm{Ham}^{c}(T^{*}L,d\lambda) such that ϕ(L0)=L\phi(L_{0})=L.

From the properties of spectral invariants, it is immediate that for all ϕ\phi and ψHamc(TL,dλ)\psi\in\mathrm{Ham}^{c}(T^{*}L,d\lambda), 0γ(ϕψ)γ(ϕ)+γ(ψ)0\leqslant\gamma(\phi\psi)\leqslant\gamma(\phi)+\gamma(\psi), and that γ(ϕ)=γ(ϕ1)\gamma(\phi)=\gamma(\phi^{-1}). Moreover, γ(L,L0)=0\gamma(L,L_{0})=0 implies L=L0L=L_{0} as proven in [25].

One of the main properties of γ\gamma which will be used in what follows is the fact that

(2) γ(ϕH1(L0),L0)=γ(ϕH1)maxt[0,1](osc(Ht|L0))\displaystyle\gamma(\phi_{H}^{1}(L_{0}),L_{0})=\gamma(\phi_{H}^{1})\leqslant\max_{t\in[0,1]}(\mathrm{osc}(H_{t}|_{L_{0}}))

where osc(Ht|L0)=maxL0(Ht)minL0(Ht)\displaystyle\mathrm{osc}(H_{t}|_{L_{0}})=\max_{L_{0}}(H_{t})-\min_{L_{0}}(H_{t}). This inequality can be directly derived from Property (6) of Theorem 12. Note that this yields

t0,γ(ϕHt0(L0),L0)t0maxt[0,t0](osc(Ht|L0)).\displaystyle\forall t_{0}\in\mathbb{R},\;\;\gamma(\phi_{H}^{t_{0}}(L_{0}),L_{0})\leqslant t_{0}\cdot\max_{t\in[0,t_{0}]}(\mathrm{osc}(H_{t}|_{L_{0}})).

3. Energy-capacity inequalities

3.1. The energy-capacity inequality for Hamiltonians constant on open sets

We prove Lemma 6 by mimicking the proof of the corresponding inequality in [10] (here for Lagrangians, in the easier world of aspherical objects). Then we prove a corollary which will be used in the proof of the main result.

Proof of Lemma 6.

First, assume that (ϕH1)0\ell_{-}(\phi_{H}^{1})\leqslant 0. Then for any admissible L0L_{0}–slow function with support in U+U_{+}, f(U+)f\in\mathcal{H}(U_{+}), we define the 11–parameter family of Hamiltonians Hs(t,x)=H(t,x)sf(x)H_{s}(t,x)=H(t,x)-sf(x) with s[0,1]s\in[0,1]. Since HH is constant on U+U_{+}, HsH_{s} generates ϕHs1=ϕH1ϕfs\phi_{H_{s}}^{1}=\phi_{H}^{1}\phi_{f}^{-s}. By triangle inequality and duality (i.e Properties (4) and (5)) of spectral invariants, we get

(3) +(ϕfs)+(ϕHs1)++(ϕH1)=+(ϕH1)(ϕHs1).\displaystyle\ell_{+}(\phi_{f}^{s})\leqslant\ell_{+}(\phi_{H_{s}}^{-1})+\ell_{+}(\phi_{H}^{1})=\ell_{+}(\phi_{H}^{1})-\ell_{-}(\phi_{H_{s}}^{1})\;.

Then notice that

Spec(𝒜Hs)=Spec(𝒜H){C+sf(p)|pcrit(f)U+}.\displaystyle\mathrm{Spec}(\mathcal{A}_{H_{s}})=\mathrm{Spec}(\mathcal{A}_{H})\cup\{C_{+}-sf(p)\,|\,p\in\mathrm{crit}(f)\cap U_{+}\}\;.

Since for all pcrit(f)U+p\in\mathrm{crit}(f)\cap U_{+}, sf(p)max(f)cLR(U+;L0)<C+sf(p)\leqslant\max(f)\leqslant c_{\mathrm{LR}}(U_{+};L_{0})<C_{+}, there exists ε>0\varepsilon>0, such that Specε(𝒜Hs)\mathrm{Spec}_{\varepsilon}(\mathcal{A}_{H_{s}}), defined as Spec(𝒜Hs)(,ε)\mathrm{Spec}(\mathcal{A}_{H_{s}})\cap(-\infty,\varepsilon), does not depend on ss and coincides with Specε(𝒜H)\mathrm{Spec}_{\varepsilon}(\mathcal{A}_{H}) which is totally discontinuous. Since the map s(ϕHs1)s\mapsto\ell_{-}(\phi_{H_{s}}^{1}) is continuous and maps 0 to Specε(𝒜H)\mathrm{Spec}_{\varepsilon}(\mathcal{A}_{H}), it has to be constant so that (ϕH11)=(ϕH1)\ell_{-}(\phi_{H_{1}}^{1})=\ell_{-}(\phi_{H}^{1}) and, from (3), Property (7) of spectral invariants immediately leads to

max(f)=+(ϕf1)+(ϕH1)(ϕH1)=γ(ϕH1(L0),L0).\displaystyle\max(f)=\ell_{+}(\phi_{f}^{1})\leqslant\ell_{+}(\phi_{H}^{1})-\ell_{-}(\phi_{H}^{1})=\gamma(\phi_{H}^{1}(L_{0}),L_{0}).

Since this holds for any L0L_{0}–slow function in (U+)\mathcal{H}(U_{+}), we get γ(ϕH1(L0),L0)cLR(U+;L0)\gamma(\phi_{H}^{1}(L_{0}),L_{0})\geqslant c_{\mathrm{LR}}(U_{+};L_{0}).

Now, assume that (ϕH1)0\ell_{-}(\phi_{H}^{1})\geqslant 0 and consider H¯(t,x)=H(t,ϕHt(x))\bar{H}(t,x)=-H(t,\phi_{H}^{t}(x)). By assumption, ϕHt\phi_{H}^{t} is the identity on UU_{-} and H¯|U=H|U=C\bar{H}|_{U_{-}}=-H|_{U_{-}}=C_{-}. Since H¯\bar{H} generates ϕH¯1=(ϕH1)1\phi_{\bar{H}}^{1}=(\phi_{H}^{1})^{-1},

(ϕH¯1)=+(ϕH1)(ϕH1)0.\displaystyle\ell_{-}(\phi_{\bar{H}}^{1})=-\ell_{+}(\phi_{H}^{1})\leqslant-\ell_{-}(\phi_{H}^{1})\leqslant 0.

Then the first case gives that γ(ϕH¯1)cLR(U;L0)\gamma(\phi_{\bar{H}}^{1})\geqslant c_{\mathrm{LR}}(U_{-};L_{0}). Since γ(ϕH¯1)=γ(ϕH1)=γ(ϕH1(L0),L0)\gamma(\phi_{\bar{H}}^{1})=\gamma(\phi_{H}^{1})=\gamma(\phi_{H}^{1}(L_{0}),L_{0}), we get that when (ϕH1)0\ell_{-}(\phi_{H}^{1})\geqslant 0, γ(ϕH1)cLR(U;L0)\gamma(\phi_{H}^{1})\geqslant c_{\mathrm{LR}}(U_{-};L_{0}).

So γ(ϕH1)min{cLR(U;L0),cLR(U+;L0)}\gamma(\phi_{H}^{1})\geqslant\min\{c_{\mathrm{LR}}(U_{-};L_{0}),c_{\mathrm{LR}}(U_{+};L_{0})\} regardless the sign of (ϕH1)\ell_{-}(\phi_{H}^{1}) which concludes the proof. ∎

From this (and as in [10]) we infer the same result but for Hamiltonian functions which are allowed to have controlled oscillations on U±U_{\pm}.

Corollary 14.

Let LL be a smooth closed manifold and U±U_{\pm} be open subsets of TLT^{*}L such that U±L0U_{\pm}\cap L_{0}\neq\emptyset. Let HH be a Hamiltonian so that for all t[0,1]t\in[0,1]

  1. (1)

    Ht|U+>cLR(U+;L0)H_{t}|_{U_{+}}>c_{\mathrm{LR}}(U_{+};L_{0}), and Ht|U<cLR(U;L0)H_{t}|_{U_{-}}<-c_{\mathrm{LR}}(U_{-};L_{0}),

  2. (2)

    osc(Ht|U±)<ε\mathrm{osc}(H_{t}|_{U_{\pm}})<\varepsilon, for some ε>0\varepsilon>0.

Then, γ(ϕH1(L0),L0)min{cLR(U;L0),cLR(U+;L0)}2ε\gamma(\phi_{H}^{1}(L_{0}),L_{0})\geqslant\min\{c_{\mathrm{LR}}(U_{-};L_{0}),c_{\mathrm{LR}}(U_{+};L_{0})\}-2\varepsilon.

Proof.

Fix η>0\eta>0. We choose disjoint open subsets V±V_{\pm} such that U±V±U_{\pm}\Subset V_{\pm} (with \Subset denoting compact containment) and333In the next few lines we loosely denote maxt[0,1]osc(ft|U)\max_{t\in[0,1]}\mathrm{osc}(f_{t}|_{U}) by oscU(f)\mathrm{osc}_{U}(f) for readability. oscV±(H)<oscU±(H)+η\mathrm{osc}_{V_{\pm}}(H)<\mathrm{osc}_{U_{\pm}}(H)+\eta. We also choose cut-off functions ρ±\rho_{\pm} with support in V±V_{\pm}, such that 0ρ±10\leqslant\rho_{\pm}\leqslant 1 and ρ±|U±=1\rho_{\pm}|_{U_{\pm}}=1. Then we define

h=Hρ+(HC+)ρ(H+C)\displaystyle h=H-\rho_{+}(H-C_{+})-\rho_{-}(H+C_{-})

with C+=inf(H|[0,1]×U+)C_{+}=\inf(H|_{[0,1]\times U_{+}}) and C=sup(H|[0,1]×U)C_{-}=-\sup(H|_{[0,1]\times U_{-}}). By triangle inequality, we get γ(ϕH1)γ(ϕh1)γ((ϕH1)1ϕh1)\gamma(\phi_{H}^{1})\geqslant\gamma(\phi_{h}^{1})-\gamma((\phi_{H}^{1})^{-1}\phi_{h}^{1}) and we now bound the right-hand side terms.

First, notice that hh satisfies the requirements of Lemma 6: h|U±=±C±h|_{U_{\pm}}=\pm C_{\pm} with by assumption C+>cLR(U+;L0),andC>cLR(U;L0)C_{+}>c_{\mathrm{LR}}(U_{+};L_{0}),\;\mbox{and}\;C_{-}>c_{\mathrm{LR}}(U_{-};L_{0}). Thus we immediately get that γ(ϕh1)min{cLR(U;L0),cLR(U+;L0)}\gamma(\phi_{h}^{1})\geqslant\min\{c_{\mathrm{LR}}(U_{-};L_{0}),c_{\mathrm{LR}}(U_{+};L_{0})\}. Now by Property (3) of spectral invariants,

γ((ϕH1)1ϕh1)oscTL((Hh)(ϕH)1)oscVV+(Hh)\displaystyle\gamma((\phi_{H}^{1})^{-1}\phi_{h}^{1})\leqslant\mathrm{osc}_{T^{*}L}\big{(}(H-h)\circ(\phi_{H})^{-1}\big{)}\leqslant\mathrm{osc}_{V_{-}\cup V_{+}}(H-h)

so that

γ((ϕH1)1ϕh1)\displaystyle\gamma((\phi_{H}^{1})^{-1}\phi_{h}^{1}) oscV+(ρ+(HC+))+oscV(ρ(H+C))\displaystyle\leqslant\mathrm{osc}_{V_{+}}\big{(}\rho_{+}(H-C_{+})\big{)}+\mathrm{osc}_{V_{-}}\big{(}\rho_{-}(H+C_{-})\big{)}
oscV+(H)+oscV(H)oscU+(H)+oscU(H)+2η\displaystyle\leqslant\mathrm{osc}_{V_{+}}(H)+\mathrm{osc}_{V_{-}}(H)\leqslant\mathrm{osc}_{U_{+}}(H)+\mathrm{osc}_{U_{-}}(H)+2\eta
2ε+2η.\displaystyle\leqslant 2\varepsilon+2\eta.

Finally, γ(ϕH1(L0),L0)min{cLR(U;L0),cLR(U+;L0)}2ε2η\gamma(\phi_{H}^{1}(L_{0}),L_{0})\geqslant\min\{c_{\mathrm{LR}}(U_{-};L_{0}),c_{\mathrm{LR}}(U_{+};L_{0})\}-2\varepsilon-2\eta for any η>0\eta>0. ∎

3.2. The energy-capacity inequality for Lagrangians displaced from an open set

We give a proof of Lemma 7 from Lisi–Rieser [15], for the reader’s convenience. The method of proof is now classical and goes back to Viterbo [25] (see also Usher’s proof of the analogous result [24] for compact manifolds, itself heavily influenced by Frauenfelder–Ginzburg–Schlenk [6]). Recall that in cotangent bundles the spectral invariants only depend on the endpoints of Hamiltonian isotopies; this drastically simplifies the proof.

Proof of Lemma 7.

Assume that ϕH1(L0)U=\phi_{H}^{1}(L_{0})\cap U=\emptyset. Choose a L0L_{0}–slow function ff in (U)\mathcal{H}(U). For s[0,1]s\in[0,1], consider the Hamiltonian diffeomorphism ϕsf1ϕH1\phi_{sf}^{1}\phi_{H}^{1}. It is the end of the isotopy defined as the concatenation

θt={ϕH2t,t[0,1/2],ϕsf2t1ϕH1,t[1/2,1].\displaystyle\theta^{t}=\left\{\begin{array}[]{cl}\phi_{H}^{2t}\,,&{t\in[0,1/2]},\\ \phi_{sf}^{2t-1}\phi_{H}^{1}\,,&{t\in[1/2,1]}.\end{array}\right.

A Hamiltonian chord of θ\theta is a path tγ(t)t\mapsto\gamma(t) such that γ(0)L0\gamma(0)\in L_{0}, γ(1)L0\gamma(1)\in L_{0} and for all tt, γ(t)=θt(γ(0))\gamma(t)=\theta^{t}(\gamma(0)). In particular, for such a chord, ϕsf1ϕH1(γ(0))L0\phi_{sf}^{1}\phi_{H}^{1}(\gamma(0))\in L_{0}. However, since by assumption ϕH1(L0)supp(sf)=\phi_{H}^{1}(L_{0})\cap\mathrm{supp}(sf)=\emptyset, necessarily ϕH1(γ(0))\phi_{H}^{1}(\gamma(0)) is not in the support of ff and γ(t)=ϕH2t(γ(0))\gamma(t)=\phi_{H}^{2t}(\gamma(0)) for all t1/2t\leqslant 1/2 and remains constant γ(t)=ϕH1(γ(0))\gamma(t)=\phi_{H}^{1}(\gamma(0)) for t1/2t\geqslant 1/2.

This means that for all ss, the set of Hamiltonian chords remains constant and so does Spec(ϕsf1ϕH1)\mathrm{Spec}(\phi_{sf}^{1}\phi_{H}^{1}). Since this set is nowhere dense and +\ell_{+} is continuous (and takes its values in the action spectrum), the function s+(ϕsf1ϕH1)s\mapsto\ell_{+}(\phi_{sf}^{1}\phi_{H}^{1}) is constant so that +(ϕH1)=+(ϕf1ϕH1)\ell_{+}(\phi_{H}^{1})=\ell_{+}(\phi_{f}^{1}\phi_{H}^{1}). Thus,

+(ϕf1)+(ϕf1ϕH1)++((ϕH1)1)=+(ϕH1)(ϕH1)=γ(ϕH1)\displaystyle\ell_{+}(\phi_{f}^{1})\leqslant\ell_{+}(\phi_{f}^{1}\phi_{H}^{1})+\ell_{+}((\phi_{H}^{1})^{-1})=\ell_{+}(\phi_{H}^{1})-\ell_{-}(\phi_{H}^{1})=\gamma(\phi_{H}^{1})

by Properties (4) and (5) of spectral invariants. Since this holds for any L0L_{0}–slow function f(U)f\in\mathcal{H}(U) and since for such a function max(f)=+(ϕf1)\max(f)=\ell_{+}(\phi_{f}^{1}) by Property (7) of spectral invariants, the result follows. ∎

Remark 15. Lemma 7 can be reformulated as follows:

Let L,L0,L,L_{0}, and UU be as in Lemma 7. Suppose that HH is a smooth Hamiltonian such that L0ϕH1(U)=L_{0}\cap\phi^{1}_{H}(U)=\emptyset. Then, γ(ϕH1(L0),L0)cLR(U;L0)\gamma(\phi^{1}_{H}(L_{0}),L_{0})\geqslant c_{\mathrm{LR}}(U;L_{0}).

The equivalence of the above statement to Lemma 7 follows easily from the fact that γ(ϕH1(L0),L0)=γ((ϕH1)1(L0),L0)\gamma(\phi^{1}_{H}(L_{0}),L_{0})=\gamma((\phi^{1}_{H})^{-1}(L_{0}),L_{0}).

Remark 16. Let L,L0,L,L_{0}, and UU be as in Lemma 7. Suppose that HH is a smooth Hamiltonian such that either ϕH1(L0)U=\phi^{1}_{H}(L_{0})\cap U=\emptyset or L0ϕH1(U)=L_{0}\cap\phi^{1}_{H}(U)=\emptyset. Then, it follows from Lemma 7, the previous remark, and Inequality (2) that maxt[0,1](osc(Ht|L0))cLR(U;L0).\max_{t\in[0,1]}(\mathrm{osc}(H_{t}|_{L_{0}}))\geqslant c_{\mathrm{LR}}(U;L_{0}).

3.3. An upper bound for the spectral distance

In this section we prove Lemma 8 which establishes Viterbo’s conjecture in a special case. The fact that Viterbo’s conjecture holds under the additional assumptions of Lemma 8 seems to be well known to experts; we provide a proof here for the sake of completeness.

Proof of Lemma 8.

Note that since ϕHt(L0)TrL\phi^{t}_{H}(L_{0})\subset T^{*}_{r}L for all t[0,1]t\in[0,1], modifying HH outside of TrLT^{*}_{r}L leaves ϕH1(L0)\phi^{1}_{H}(L_{0}), and hence γ(ϕH1(L0),L0)\gamma(\phi^{1}_{H}(L_{0}),L_{0}), unchanged. Therefore, by cutting HH off outside of TrLT^{*}_{r}L and replacing rr with 2r2r we may assume that HH is supported inside TrLVT^{*}_{r}L\setminus V.

Pick f:Lf\colon\thinspace L\rightarrow\mathbb{R} to be a Morse function on LL whose critical points are all contained in the open set 𝒱\mathcal{V}. Because ff has no critical points inside L𝒱L\setminus\mathcal{V}, we may assume, by rescaling, that

(4) df|L𝒱g1.\|df|_{L\setminus\mathcal{V}}\|_{g}\geqslant 1.

Let β:TL\beta\colon\thinspace T^{*}L\rightarrow\mathbb{R} denote a non-negative cutoff function such that β=1\beta=1 on TRLT^{*}_{R}L, where RR is picked so that RrR\gg r. Let F=βπf:TL.F=\beta\pi^{*}f\colon\thinspace T^{*}L\rightarrow\mathbb{R}. By picking RR to be sufficiently large, we can ensure that, for t[0,2r]t\in[0,2r] and (q,p)(q,p) in a neighborhood of TrLT^{*}_{r}L, the Hamiltonian flow of FF is given by the formula

ϕFt(q,p)=(q,p+tdf(q)).\phi^{t}_{F}(q,p)=(q,p+tdf(q)).

This, combined with (4), implies that ϕFt(q,0)TrL\phi^{t}_{F}(q,0)\notin T^{*}_{r}L for r<t2rr<t\leqslant 2r and any point qL𝒱q\in L\setminus\mathcal{V}. Hence, we see that ϕFt(L0)\phi^{t}_{F}(L_{0}) is outside the support of HH for r<t2rr<t\leqslant 2r. Therefore, ϕH1ϕF2r(L0)=ϕF2r(L0),\phi^{1}_{H}\phi^{2r}_{F}(L_{0})=\phi^{2r}_{F}(L_{0}), and so

γ(ϕH1ϕF2r(L0),L0)=γ(ϕF2r(L0),L0).\gamma(\phi^{1}_{H}\phi^{2r}_{F}(L_{0}),L_{0})=\gamma(\phi^{2r}_{F}(L_{0}),L_{0}).

Thus we get,

γ(ϕH1(L0),L0)=γ(ϕH1)\displaystyle\gamma(\phi_{H}^{1}(L_{0}),L_{0})=\gamma(\phi_{H}^{1}) γ(ϕH1ϕF2r)+γ((ϕF2r)1)=γ(ϕF2r)+γ(ϕF2r)\displaystyle\leqslant\gamma(\phi^{1}_{H}\phi^{2r}_{F})+\gamma((\phi^{2r}_{F})^{-1})=\gamma(\phi^{2r}_{F})+\gamma(\phi^{2r}_{F})
4rosc(F)=4rosc(f)\displaystyle\leqslant 4r\,\mathrm{osc}(F)=4r\,\mathrm{osc}(f)

and the result follows with C=4osc(f).C=4\,\mathrm{osc}(f).

4. Localized results for Lagrangians

The main goal of this section is to establish a suitable localized version of Theorem 3 for Lagrangians; since we seek localized statements we do not assume that the Lagrangians in question are necessarily closed. Not surprisingly, in this new setting Theorem 3 does not hold as stated. The localized results of this section, which have more complicated statements and proofs, are more powerful and they constitute the main technical steps towards proving Theorems 1 and 3.

We prove the analog of the direct implication of Theorem 3 in Section 4.1. The analog of the converse implication is proven in Section 4.2. Since LL is a Lagrangian, its characteristic foliation has a single leaf, LL itself, and thus in this section we make no mention of characteristic foliations.

4.1. C0C^{0}–Hamiltonians constant on a Lagrangian preserve it

In this subsection, we show that if the restriction of HCHam0H\in C^{0}_{\mathrm{Ham}} to a Lagrangian LL is a function of time, then the associated hameotopy ϕH\phi_{H} preserves LL, locally. More precisely,

Proposition 17.

Let LML\subset M denote a Lagrangian (not necessarily closed) and HCHam0H\in C^{0}_{\mathrm{Ham}} with associated hameotopy ϕH\phi_{H}. If H|L=c(t)H|_{L}=c(t), is a function of time, then for any point pLp\in L there exists ε>0\varepsilon>0 such that ϕHt(p)L\phi_{H}^{t}(p)\in L for all t[0,ε]t\in[0,\varepsilon].

Our proof of the above proposition will use the following simple lemma on the local structure of Lagrangians.

Lemma 18.

Let LML\subset M denote a Lagrangian. Around each point pLp\in L there exists a neighborhood LpLL_{p}\subset L such that LpL_{p} is contained in a closed Lagrangian torus TMT\subset M.

Proof.

Let UU denote a Darboux ball around pp, equipped with the standard coordinates (xi,yi),(x_{i},y_{i}), such that LU={(xi,yi)|a<xi<a,yi=0}.L\cap U=\{(x_{i},y_{i})\,|\,-a<x_{i}<a,y_{i}=0\}. Let Lp={(xi,yi)|a2xia2,yi=0}.L_{p}=\{(x_{i},y_{i})\,|\,-\frac{a}{2}\leqslant x_{i}\leqslant\frac{a}{2},y_{i}=0\}. Then, for all kk, the projection of LpL_{p} onto the xk,yk\langle x_{k},y_{k}\rangle plane is the segment [a2,a2]×{0}[-\frac{a}{2},\frac{a}{2}]\times\{0\} which can be completed to a smooth embedded loop, say TkT_{k}, in the xk,yk\langle x_{k},y_{k}\rangle plane. Here, we can make sense of the projection onto a coordinate plane by identifying UU with a standard ball in 2n.\mathbb{R}^{2n}. We set T=T1××TnT=T_{1}\times\cdots\times T_{n}. This is a Lagrangian torus inside UU containing LpL_{p}. ∎

Proof of Proposition 17.

By replacing HH by Hc(t)H-c(t) we can suppose that H|L=0H|_{L}=0. Apply Lemma 18 to obtain LpL_{p} and TT as described in the lemma and note that by replacing LL with LpL_{p} we may make the following simplifying assumption: there exists a Lagrangian torus TT in MM such that LTL\subset T.

We will now prove the proposition under this simplifying assumption. Let Hi:[0,1]×MH_{i}\colon\thinspace[0,1]\times M\to\mathbb{R} denote a sequence of smooth Hamiltonians such that HiH_{i} converges uniformly to HH and ϕHi\phi_{H_{i}} converges to ϕH\phi_{H} in C0C^{0}–topology. Take WVW\Subset V to be open subsets of MM such that

(5) pWLandV¯(TL)=.p\in W\cap L\quad\text{and}\quad\overline{V}\cap(T\setminus L)=\emptyset.

Recall that the symbol \Subset denotes compact containment and V¯\overline{V} denotes the closure of VV. The second condition in (5) allows us to pick a cutoff function β:M\beta\colon\thinspace M\to\mathbb{R} such that β|V=1\beta|_{V}=1 and β|TL=0\beta|_{T\setminus L}=0. By shrinking VV, if needed, we may assume that β\beta is supported in a Weinstein neighborhood of TT.

Let Gi=βHiG_{i}=\beta H_{i} and G=βHG=\beta H. Observe that GiG_{i} converges uniformly to GG and G|T=0G|_{T}=0. We pick ε>0\varepsilon>0 such that

(6) t[0,ε],ϕHt(W)V.\forall t\in[0,\varepsilon],\;\;\phi^{t}_{H}(W)\subset V.

For ii large enough ϕHit(W)V\phi^{t}_{H_{i}}(W)\subset V for all t[0,ε]t\in[0,\varepsilon]. Since, Gi|V=Hi|VG_{i}|_{V}=H_{i}|_{V} we conclude that, for large i,i,

(7) (t,x)[0,ε]×W,ϕGit=ϕHit.\forall(t,x)\in[0,\varepsilon]\times W,\;\;\phi^{t}_{G_{i}}=\phi^{t}_{H_{i}}.

For a contradiction, suppose that ϕHt0(p)\phi^{t_{0}}_{H}(p) is not contained in LL for some t0[0,ε].t_{0}\in[0,\varepsilon]. From (5) and (6) we conclude that ϕHt0(p)T\phi^{t_{0}}_{H}(p)\notin T. Hence, we can find a small ball BWB\subset W around pp which intersects TT non-trivially and such that ϕHt0(B)¯T=\overline{\phi^{t_{0}}_{H}(B)}\cap T=\emptyset. Hence, for ii large enough, we have ϕHit0(B)T=.\phi^{t_{0}}_{H_{i}}(B)\cap T=\emptyset. From (7) we get that ϕGit0(B)T=.\phi^{t_{0}}_{G_{i}}(B)\cap T=\emptyset.

We picked β\beta such that the Hamiltonians GiG_{i} all have support in a Weinstein neighborhood of TT. Therefore, we can pass to TTT^{*}T, apply Lemma 7 and its consequences as stated in Remarks 3.2 and 3.2 and conclude that

t0maxt[0,t0](osc(Gi(t,)|T0))cLR(B;T0),t_{0}\cdot\max_{t\in[0,t_{0}]}(\mathrm{osc}(G_{i}(t,\cdot)|_{T_{0}}))\geqslant c_{\mathrm{LR}}(B;T_{0}),

where T0T_{0} stands for the 0–section in TTT^{*}T. Here, we have used the fact that ϕGit0\phi^{t_{0}}_{G_{i}} is the time–1 map of the flow of the Hamiltonian t0Gi(t0t,x)t_{0}G_{i}(t_{0}t,x). Since GiG_{i} converges uniformly to GG, the same inequality must hold for GG but this contradicts the fact that G|T=0.G|_{T}=0.

4.2. C0C^{0}–Hamiltonians preserving a Lagrangian are constant on it

In this subsection, we show that if HCHam0H\in C^{0}_{\mathrm{Ham}} generates a hameotopy ϕH\phi_{H} which (locally) preserves a Lagrangian LL then, the restriction of HH to LL is (locally) a function of time. More precisely,

Proposition 19.

Let LML\subset M denote a Lagrangian, 𝒰\mathcal{U} an open subset of LL, and HCHam0H\in C^{0}_{\mathrm{Ham}} with associated hameotopy ϕH\phi_{H}. Suppose that ϕHt(𝒰)L\phi^{t}_{H}(\mathcal{U})\subset L for all t[0,1]t\in[0,1] and let 𝒱\mathcal{V} denote the interior of t[0,1]ϕHt(𝒰)\cap_{t\in[0,1]}\phi^{t}_{H}(\mathcal{U}). Then, H(t,)|𝒱H(t,\cdot)|_{\mathcal{V}} is a locally constant function for each t[0,1]t\in[0,1].

We were recently informed by Y.-G. Oh that it is possible to extract the above proposition from [20, Theorem 4.9]; the techniques of [20] are different than ours.

Our proof of the proposition uses the following consequence of Corollary 14. This result can be viewed as a Lagrangian analog of the uniqueness of generators Theorem [10, Theorem 2]. The argument presented here is similar to the proof of the mentioned uniqueness theorem.

Proposition 20.

Let LL be a smooth closed manifold and {Hk}k\{H_{k}\}_{k} a sequence of smooth, uniformly compactly supported Hamiltonian functions on TLT^{*}L (that is, there exists a compact KK such that ksupp(Hk)K\cup_{k}\mathrm{supp}(H_{k})\subset K), so that

  1. (1)

    for all t[0,1]t\in[0,1], γ(ϕHkt(L0),L0)\gamma(\phi^{t}_{H_{k}}(L_{0}),L_{0}) converges to 0, and

  2. (2)

    {Hk}k\{H_{k}\}_{k} uniformly converges to a continuous function HH.

Then, HH restricted to L0L_{0} is a function of time.

The proofs of this section repeatedly use the following simple fact: Let HH denote a Hamiltonian, C0C^{0} or smooth. The time–tt flow of H~(t,x)=aH(t0+at,x)\tilde{H}(t,x)=aH(t_{0}+at,x) is given by the expression: ϕH~t=ϕHt0+at(ϕHt0)1\phi_{\tilde{H}}^{t}=\phi^{t_{0}+at}_{H}(\phi_{H}^{t_{0}})^{-1}.

Proof of Proposition 20.

If H|[0,1]×L0H|_{[0,1]\times L_{0}} is not a function of time, there exist t0[0,1)t_{0}\in[0,1) and x+x_{+}, xL0x_{-}\in L_{0} such that Ht0(x+)>Ht0(x)H_{t_{0}}(x_{+})>H_{t_{0}}(x_{-}). Up to a shift (and cutoff far from KL0K\cup L_{0}), we can assume that Ht0(x+)=Ht0(x)=Δ>0H_{t_{0}}(x_{+})=-H_{t_{0}}(x_{-})=\Delta>0.

Now, let ε=Δ/4\varepsilon=\Delta/4 and notice that there exist δ(0,1]\delta\in(0,1] and r>0r>0 such that [t0,t0+δ][0,1][t_{0},t_{0}+\delta]\subset[0,1] and that there exist symplectically embedded balls, centered at x±x_{\pm}, B±=ι±(Bn(0,r))B_{\pm}=\iota_{\pm}(B_{\mathbb{C}^{n}}(0,r)), with real part mapped to L0L_{0}, which are disjoint and such that

osc[t0,t0+δ]×B±H<ε.\mathrm{osc}_{[t_{0},t_{0}+\delta]\times B_{\pm}}H<\varepsilon.

By shrinking either δ\delta or rr, we can assume that the Lisi-Rieser capacity of the balls with respect to L0L_{0} satisfy

cLR(B±;L0)=δ(Δε).c_{\mathrm{LR}}(B_{\pm};L_{0})=\delta(\Delta-\varepsilon).

Then set Fk(t,x)=δHk(t0+δt,x)F_{k}(t,x)=\delta H_{k}(t_{0}+\delta t,x) and define FF accordingly. By construction, the function FF satisfies

(t,x)[0,1]×B+,F(t,x)>cLR(B±;L0),\forall(t,x)\in[0,1]\times B_{+},\ F(t,x)>c_{\mathrm{LR}}(B_{\pm};L_{0}),
(t,x)[0,1]×B,F(t,x)<cLR(B±;L0)\forall(t,x)\in[0,1]\times B_{-},\ F(t,x)<-c_{\mathrm{LR}}(B_{\pm};L_{0})

and

osc[0,1]×B±(F)<δε.\mathrm{osc}_{[0,1]\times B_{\pm}}(F)<\delta\varepsilon.

Since the Hamiltonians FkF_{k} converge uniformly to FF, they satisfy the same inequalities for kk large enough. Thus, by Corollary 14,

γ(ϕFk1(L0),L0)cLR(B±;L0)2δε=δ(Δ3ε)=14δΔ>0.\gamma(\phi_{F_{k}}^{1}(L_{0}),{L_{0}})\geqslant c_{\mathrm{LR}}(B_{\pm};L_{0})-2\delta\varepsilon=\delta(\Delta-3\varepsilon)=\tfrac{1}{4}\delta\Delta>0.

Hence, γ(ϕFk1(L0),L0)\gamma(\phi_{F_{k}}^{1}(L_{0}),{L_{0}}) is uniformly bounded away from 0. However, ϕFk1=ϕHkt0+δ(ϕHkt0)1\phi_{F_{k}}^{1}=\phi_{H_{k}}^{t_{0}+\delta}(\phi_{H_{k}}^{t_{0}})^{-1} so that, by Properties (4) and (5) of spectral invariants,

γ(ϕFk1(L0),L0)=γ(ϕFk1)\displaystyle\gamma(\phi_{F_{k}}^{1}(L_{0}),{L_{0}})=\gamma(\phi_{F_{k}}^{1}) γ(ϕHkt0+δ)+γ((ϕHkt0)1)=γ(ϕHkt0+δ)+γ(ϕHkt0)\displaystyle\leqslant\gamma(\phi_{H_{k}}^{t_{0}+\delta})+\gamma((\phi_{H_{k}}^{t_{0}})^{-1})=\gamma(\phi_{H_{k}}^{t_{0}+\delta})+\gamma(\phi_{H_{k}}^{t_{0}})
γ(ϕHkt0+δ(L0),L0)+γ(ϕHkt0(L0),L0)\displaystyle\leqslant\gamma(\phi_{H_{k}}^{t_{0}+\delta}(L_{0}),{L_{0}})+\gamma(\phi_{H_{k}}^{t_{0}}(L_{0}),{L_{0}})

which goes to 0 when kk goes to infinity because of Assumption (1) and we get a contradiction. ∎

Proof of Proposition 19.

Assume, for a contradiction, that the conclusion of the proposition fails to hold. We can therefore find p𝒱p\in\mathcal{V} and t0[0,1)t_{0}\in[0,1) such that H(t0,)H(t_{0},\cdot) is not constant on any neighborhood of pp in 𝒱\mathcal{V}. First, note that, up to time reparametrization, we may assume that t0=0t_{0}=0. Indeed, replace HH with H~(t,x)=aH(t0+at,x)\tilde{H}(t,x)=aH(t_{0}+at,x), where a=1t0,a=1-t_{0}, and 𝒰\mathcal{U} with 𝒰~=ϕHt0(𝒰)\tilde{\mathcal{U}}=\phi^{t_{0}}_{H}(\mathcal{U}). Then, 𝒱\mathcal{V} is contained in the interior of t[0,1]ϕH~t(𝒰~)\cap_{t\in[0,1]}\phi^{t}_{\tilde{H}}(\tilde{\mathcal{U}}) and H~(0,)\tilde{H}(0,\cdot) is not constant on any neighborhood of pp.

Apply Lemma 18 to obtain LpL_{p} and TT as described in the lemma. By shrinking LpL_{p}, if needed, we may assume that Lp𝒱.L_{p}\Subset\mathcal{V}. Let UMU\subset M denote a small open set around pp which is contained in a Weinstein neighborhood of TT and such that LpU=LU=TULp.L_{p}\cap U=L\cap U=T\cap U\Subset L_{p}. Furthermore, towards the end of this proof we will need to apply Lemma 8, and so we pick UU such that the projection of UU to TT along the cotangent fibers in the Weinstein neighborhood is a proper subset of TT whose complement contains a ball.

Since LpULp𝒱L_{p}\cap U\Subset L_{p}\Subset\mathcal{V}, it follows that there exists a small ε>0\varepsilon>0 such that ϕHt(LpU),(ϕHt)1(LpU)Lp\phi^{t}_{H}(L_{p}\cap U),(\phi^{t}_{H})^{-1}(L_{p}\cap U)\Subset L_{p} for all t[0,ε].t\in[0,\varepsilon]. Replacing HH with H~(t,x)=εH(εt,x)\tilde{H}(t,x)=\varepsilon H(\varepsilon t,x), we may assume that

ϕHt(LpU),(ϕHt)1(LpU)Lp for all t[0,1].\phi^{t}_{H}(L_{p}\cap U),(\phi^{t}_{H})^{-1}(L_{p}\cap U)\Subset L_{p}\text{ for all }t\in[0,1].

Next, let BB denote an open neighborhood of pp which is compactly contained in UU. Once again, as in the previous paragraph, by a reparametrization in time, where HH is replaced with H~(t,x)=εH(εt,x)\tilde{H}(t,x)=\varepsilon H(\varepsilon t,x) for a sufficiently small ε\varepsilon, we may assume that ϕHt(B)\phi^{t}_{H}(B), (ϕHt)1(B)U(\phi^{t}_{H})^{-1}(B)\Subset U for all t[0,1].t\in[0,1].

Pick qBq\in B such that H(0,p)H(0,q)H(0,p)\neq H(0,q) and take a symplectomorphism ψ\psi supported in BB such that ψ\psi preserves TT and ψ(p)=q\psi(p)=q. Consider the continuous Hamiltonian G=(HψH)ϕHG=(H\circ\psi-H)\circ\phi_{H}. It is supported in t[0,1](ϕHt)1(B)U\cup_{t\in[0,1]}(\phi_{H}^{t})^{-1}(B)\Subset U, and moreover, the flow of GG is (ϕHt)1ψ1ϕHtψ(\phi_{H}^{t})^{-1}\psi^{-1}\phi_{H}^{t}\psi. We will now prove that this flow preserves TT globally. Note that the flow is supported in UU and pick xTULpx\in T\cap U\subset L_{p}. Since ϕHt(TU)T\phi^{t}_{H}(T\cap U)\subset T, and ψ(x)TU\psi(x)\in T\cap U, we see that ϕHtψ(x)T\phi^{t}_{H}\psi(x)\in T. First, suppose that ϕHtψ(x)B\phi^{t}_{H}\psi(x)\notin B. Then, ϕHtψ(x)\phi^{t}_{H}\psi(x) is outside the support of ψ1\psi^{-1} and so (ϕHt)1ψ1ϕHtψ(x)=ψ(x)(\phi_{H}^{t})^{-1}\psi^{-1}\phi_{H}^{t}\psi(x)=\psi(x) which is in TT. Next, suppose that ϕHtψ(x)BT\phi^{t}_{H}\psi(x)\in B\cap T. Then, ψ1ϕHtψ(x)BT\psi^{-1}\phi^{t}_{H}\psi(x)\in B\cap T, and so it suffices to check that (ϕHt)1(BT)T(\phi_{H}^{t})^{-1}(B\cap T)\subset T: this is because BTUTULpB\cap T\subset U\cap T\subset U\cap L_{p} and (ϕHt)1(LpU)Lp(\phi^{t}_{H})^{-1}(L_{p}\cap U)\Subset L_{p} for all t[0,1]t\in[0,1]. We have proven that the flow of GG preserves TT globally.

Note that G|TUG|_{T\cap U} is not a function of time only: G(0,p)=H(0,q)H(0,p)0G(0,p)=H(0,q)-H(0,p)\neq 0 and G=0G=0 near the boundary of UU. Hence, we have obtained a C0C^{0}–Hamiltonian GG, supported in UU, whose flow ϕGt\phi^{t}_{G} preserves TT globally, but G(0,)G(0,\cdot) is not constant on TT. Because GCHam0G\in C^{0}_{\mathrm{Ham}} there exist smooth Hamiltonians GiG_{i} such that {Gi}\{G_{i}\} converges uniformly to GG and {ϕGi}\{\phi_{G_{i}}\} converges to ϕG\phi_{G}. Furthermore, we can ensure that all GiG_{i}’s are supported in UU. This can be achieved by picking a corresponding sequence of smooth Hamiltonians HiH_{i} for HH and defining Gi=(HiψHi)ϕHiG_{i}=(H_{i}\circ\psi-H_{i})\circ\phi_{H_{i}}. For large ii, GiG_{i} is supported in UU.

Since UU is contained in a Weinstein neighborhood of TT, we can pass to TTT^{*}T and work with the Lagrangian spectral invariants of the 0–section T0T_{0} associated to the Hamiltonians GiG_{i}. Recall that in the second paragraph of the proof we picked the set UU so that Lemma 8 could be applied. For any fixed r>0r>0, because ϕGt\phi^{t}_{G} preserves T0,T_{0}, we have ϕGit(T0)TrT\phi^{t}_{G_{i}}(T_{0})\subset T^{*}_{r}T for sufficiently large ii. The Hamiltonians GiG_{i} are all supported in UU and hence using Lemma 8 we conclude that γ(ϕGi1(T0),T0)Cr\gamma(\phi^{1}_{G_{i}}(T_{0}),T_{0})\leqslant Cr, i.e. γ(ϕGi1(T0),T0)0\gamma(\phi^{1}_{G_{i}}(T_{0}),T_{0})\to 0. Of course, by the same reasoning we obtain that γ(ϕGit(T0),T0)0\gamma(\phi^{t}_{G_{i}}(T_{0}),T_{0})\to 0 for all t[0,1]t\in[0,1]. Then, Proposition 20 implies that G|T=c(t)G|_{T}=c(t), which contradicts the fact that G|TG|_{T} is not a function of time only. ∎

5. C0C^{0}–rigidity of coisotropic submanifolds and their characteristic foliations

This section is devoted to the proofs of Theorems 1, 3 and 5. We begin by proving Theorem 3 and then deduce Theorems 1 and 5 from it.

Before going into the proof, recall (see [14, Proposition 13.7] and [8]) that coisotropic submanifolds admit coisotropic charts, that is, for every point pCp\in C, there is a pair (θ,U)(\theta,U) where UU is an open neighborhood of pp and θ:UV2n\theta\colon\thinspace U\to V\subset\mathbb{R}^{2n} is a symplectic diffeomorphism which maps pp to 0 and CC to the standard coisotropic linear subspace

𝒞0={(x1,,xn,y1,,yn)|(ynk+1,,yn)=(0,,0)}.\displaystyle\mathcal{C}_{0}=\{(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})\,|\,(y_{n-k+1},\ldots,y_{n})=(0,\ldots,0)\}.

Such a diffeomorphism sends the characteristic foliation of CC to that of 𝒞0\mathcal{C}_{0}, whose leaf through a point q=(a1,,an,b1,,bnk,0,,0)𝒞0q=(a_{1},\ldots,a_{n},b_{1},\ldots,b_{n-k},0,\ldots,0)\in\mathcal{C}_{0} is the affine subspace

0(q)={(a1,,ank,xnk+1,,xn,b1,,\displaystyle\mathcal{F}_{0}(q)=\{(a_{1},\ldots,a_{n-k},x_{n-k+1},\ldots,x_{n},b_{1},\ldots, bnk,0,,0)\displaystyle b_{n-k},0,\ldots,0)
|(xnk+1,,xn)k}.\displaystyle\,|\,(x_{n-k+1},\ldots,x_{n})\in\mathbb{R}^{k}\}\,.

The first step of the proof, is establishing the next lemma which is a version of the first implication of Theorem 3 that does not require the coisotropic submanifold to be a closed subset but holds only for small times.

Lemma 21.

Let (M,ω)(M,\omega) be a symplectic manifold and CC a coisotropic submanifold of MM. Let HCHam0(M,ω)H\in C^{0}_{\mathrm{Ham}}(M,\omega) with induced hameotopy ϕH\phi_{H}. Assume that the restriction of HH to CC only depends on time. Then, for every pCp\in C, there exists ε>0\varepsilon>0 such that for all t[0,ε]t\in[0,\varepsilon], ϕHt(p)\phi_{H}^{t}(p) belongs to (p)\mathcal{F}(p), the characteristic leaf of CC through pp.

Before going into the details of the proof of Lemma 21, we make the following observation. The lemma holds for coisotropic submanifolds of arbitrary codimension but its proof will follow from the particular case of Lagrangians. As mentioned in the introduction, this is not surprising in view of Weinstein’s creed: “Everything is a Lagrangian submanifold!” [28].

Proof.

Let pCp\in C and let (U,θ)(U,\theta) be a coisotropic chart as defined above. For i{1,,nk}i\in\{1,\ldots,n-k\} consider the Lagrangian linear subspaces

Λi={(x1,,xn,y1,,yn)|xi=0 and ji,yj=0},\Lambda_{i}=\{(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})\,|\,x_{i}=0\text{ and }\forall j\neq i,y_{j}=0\},

and their pull backs Li=θ1(Λi)L_{i}=\theta^{-1}(\Lambda_{i}). Clearly, for all i{1,,nk},LiCUi\in\{1,\ldots,n-k\},L_{i}\subset C\cap U and

(p)U=i=1nkLi.\mathcal{F}(p)\cap U=\bigcap_{i=1}^{n-k}L_{i}.

Let HH be as in the statement of Lemma 21. Then for any ii, the restriction of HH to LiL_{i} is a function of time since LiL_{i} is included in CC. Thus by Proposition 17 there exists εi>0\varepsilon_{i}>0 such that for all t[0,εi]t\in[0,\varepsilon_{i}], ϕHt(p)Li\phi_{H}^{t}(p)\in L_{i}. Taking ε=min{ε1,,εnk}\varepsilon=\min\{\varepsilon_{1},\ldots,\varepsilon_{n-k}\}, we get

t[0,ε],ϕHt(p)i=1nkLi(p).\forall t\in[0,\varepsilon],\;\phi_{H}^{t}(p)\in\bigcap_{i=1}^{n-k}L_{i}\subset\mathcal{F}(p).

We can now prove Theorem 3.

Proof of Theorem 3.

Let HCHam0H\in C^{0}_{\mathrm{Ham}} such that H|CH|_{C} is a function of time only and pick pCp\in C. For a contradiction, assume that for some t>0t>0, ϕHt(p)(p)\phi_{H}^{t}(p)\notin\mathcal{F}(p) and set t0=inf{t>0|ϕHt(p)(p)}t_{0}=\inf\{t>0\,|\,\phi_{H}^{t}(p)\notin\mathcal{F}(p)\}. Note that since CC is a closed subset, the point ϕHt0(p)\phi_{H}^{t_{0}}(p) belongs to CC. Then, consider the Hamiltonian Kt=Ht+t0K_{t}=-H_{-t+t_{0}}, so that ϕKt=ϕHt+t0(ϕHt0)1\phi_{K}^{t}=\phi_{H}^{-t+t_{0}}(\phi_{H}^{t_{0}})^{-1}. Its restriction to CC is also a function of time. Lemma 21 applied to KK at the point ϕHt0(p)\phi_{H}^{t_{0}}(p) implies that for some small t>0t>0, ϕHt0t(p)(ϕHt0(p))\phi_{H}^{t_{0}-t}(p)\in\mathcal{F}(\phi_{H}^{t_{0}}(p)). But by definition of t0t_{0}, we also have ϕHt0t(p)(p)\phi_{H}^{t_{0}-t}(p)\in\mathcal{F}(p), hence ϕHt0(p)(p)\phi_{H}^{t_{0}}(p)\in\mathcal{F}(p). Now apply Lemma 21 again to Ht+t0H_{t+t_{0}} at the point ϕHt0(p)\phi_{H}^{t_{0}}(p). We get that for some ε>0\varepsilon^{\prime}>0 and all t[t0,t0+ε]t\in[t_{0},t_{0}+\varepsilon^{\prime}], ϕHt(p)(p)\phi_{H}^{t}(p)\in\mathcal{F}(p) which contradicts the definition of t0t_{0}. Thus, ϕHt(p)(p)\phi_{H}^{t}(p)\in\mathcal{F}(p) and the direct implication of Theorem 3 follows.

We now prove the converse. Assume that the flow of HCHam0H\in C^{0}_{\mathrm{Ham}} preserves each leaf of the characteristic foliation. We are going to show first that the function H0H_{0} is locally constant.

Let pCp\in C and θ:UV\theta\colon\thinspace U\to V be a coisotropic chart around pp, with θ(p)=0\theta(p)=0. For σ>0\sigma>0 small enough, the set t[0,σ](ϕHt)1(U)\cap_{t\in[0,\sigma]}(\phi_{H}^{t})^{-1}(U) contains pp in its interior. Denote by UU^{\prime} this interior for some fixed σ\sigma. Similarly, for s(0,σ]s\in(0,\sigma] small enough, t[0,s]ϕHt(U)\cap_{t\in[0,s]}\phi_{H}^{t}(U^{\prime}) contains pp in its interior. Let U′′U^{\prime\prime} be an open neighborhood of pp contained in this interior, and with the property that θ(U′′)\theta(U^{\prime\prime}) is convex. Let qq be any other point in U′′U^{\prime\prime} and Λ\Lambda be a linear Lagrangian subspace included in 𝒞0\mathcal{C}_{0}, containing θ(q)\theta(q) and the standard leaf 0(0)\mathcal{F}_{0}(0). The subspace Λ\Lambda can be written as the union of the leaves 0(x)\mathcal{F}_{0}(x) for all xΛx\in\Lambda.

Now, consider the Lagrangian L=θ1(ΛV)L=\theta^{-1}(\Lambda\cap V). Let 𝒰=LU\mathcal{U}=L\cap U^{\prime} and 𝒱=LU′′\mathcal{V}=L\cap U^{\prime\prime}. By construction, q𝒱q\in\mathcal{V}. By assumption ϕHt(𝒰)L\phi_{H}^{t}(\mathcal{U})\subset L for all t[0,s]t\in[0,s]. We may apply Proposition 19 to LL and the continuous Hamiltonian Kt(x)=sHst(x)K_{t}(x)=sH_{st}(x) which generates the hameotopy ϕHst\phi_{H}^{st}. We get that for any t[0,1]t\in[0,1], KtK_{t} is locally constant on 𝒱\mathcal{V}. Equivalently, for any t[0,s]t\in[0,s], HtH_{t} is locally constant on 𝒱\mathcal{V}. Now since θ(U′′)\theta(U^{\prime\prime}) is convex and Λ\Lambda is linear, θ(U′′)Λ\theta(U^{\prime\prime})\cap\Lambda is connected. It follows that 𝒱\mathcal{V} is also connected and therefore Ht(p)=Ht(q)H_{t}(p)=H_{t}(q). To summarize, we proved that for tt small enough, HtH_{t} is constant on U′′CU^{\prime\prime}\cap C. In particular, H0H_{0} is locally constant on CC.

Since CC is assumed to be connected, this means that H0H_{0} is constant on CC. The argument we followed for t=0t=0 applies for any other initial time. Thus, HtH_{t} must be constant on CC for any tt. ∎

The proof of Theorem 1 relies on the first implication of Theorem 3 and the following characterization of coisotropic submanifolds and their characteristic foliations:

A submanifold is coisotropic if and only if the flow of every autonomous Hamiltonian constant on it preserves it. Moreover, the leaf through a point pp is locally the union of the orbits of pp under the flows of all such Hamiltonians.

The next lemma is based on this characterization.

Lemma 22.

Let CC be a submanifold in a symplectic manifold (M,ω)(M,\omega). Assume that every point pCp\in C admits an open neighborhood VV such that any HCc(V)H\in C_{c}^{\infty}(V), with H|C0H|_{C}\equiv 0, satisfies ϕHt(p)C\phi_{H}^{t}(p)\in C for every t[0,+)t\in[0,+\infty). Then CC is coisotropic.

Moreover, for such a neighborhood VV, there exists a smaller neighborhood WVW\Subset V such that, the leaf (p)\mathcal{F}(p) of the characteristic foliation of CC passing through pp satisfies

W(p)=W{ϕHt(p)|t[0,+),HCc(V),H|C0}.W\cap\mathcal{F}(p)=W\cap\{\phi_{H}^{t}(p)\,|\,t\in[0,+\infty),H\in C_{c}^{\infty}(V),H|_{C}\equiv 0\}.
Proof.

Let pCp\in C and let VV be an open subset as in the statement of the lemma. Assume that CC coincides locally with f11(0)fk1(0)f_{1}^{-1}(0)\cap\ldots\cap f_{k}^{-1}(0) for some smooth functions f1,,fkf_{1},\ldots,f_{k} whose differentials are linearly independent at pp. By multiplying by an appropriate cutoff function, we can assume that these functions are defined everywhere on MM, have compact support in VV, and vanish on CC.

The Hamiltonian vector fields at pp of f1,,fkf_{1},\ldots,f_{k} span (TpC)ω(T_{p}C)^{\omega}, and by assumption belong to TpCT_{p}C. Thus (TpC)ωTpC(T_{p}C)^{\omega}\subset T_{p}C and CC is coisotropic.

Now, since the characteristic leaves are preserved by smooth Hamiltonians constant on CC, we have the inclusion

(p){ϕHt(p)|t[0,+),HCc(V),H|C0}.\mathcal{F}(p)\supset\{\phi_{H}^{t}(p)\,|\,t\in[0,+\infty),H\in C_{c}^{\infty}(V),H|_{C}\equiv 0\}.

Conversely, consider the map

F:kC,(v1,,vk)ϕi=1kvifi1(p).F\colon\thinspace\mathbb{R}^{k}\to C,\quad(v_{1},\ldots,v_{k})\mapsto\phi^{1}_{\sum_{i=1}^{k}v_{i}f_{i}}(p).

Since, i=1kvifi\sum_{i=1}^{k}v_{i}f_{i} is constant on CC, its flow preserves the characteristics, hence FF takes values in the characteristic leaf (p)\mathcal{F}(p) through pp. The partial derivatives of FF at 0 are viF(0)=Xfi(p)\partial_{v_{i}}F(0)=X_{f_{i}}(p) and in particular they are linearly independent and span Tp(p)=(TpC)ωT_{p}\mathcal{F}(p)=(T_{p}C)^{\omega}. The inverse function theorem then shows that FF is a diffeomorphism from a neighborhood of 0 to a neighborhood of pp in (p)\mathcal{F}(p). This shows

W(p)W{ϕHt(p)|t[0,+),HCc(V),H|C0}W\cap\mathcal{F}(p)\subset W\cap\{\phi_{H}^{t}(p)\,|\,t\in[0,+\infty),H\in C_{c}^{\infty}(V),H|_{C}\equiv 0\}

for some neighborhood of pp in MM and finishes the proof of Lemma 22. ∎

We are now ready to prove Theorem 1.

Proof of Theorem 1.

Let CC be a smooth coisotropic submanifold, and θ:UV\theta\colon\thinspace U\to V be a symplectic homeomorphism. Assume C=θ(UC)C^{\prime}=\theta(U\cap C) is smooth. Let pCp^{\prime}\in C^{\prime} and p=θ1(p)p=\theta^{-1}(p^{\prime}). By passing to an appropriate Darboux chart around pp, we may assume that U2nU\subset\mathbb{R}^{2n}, p=02np=0\in\mathbb{R}^{2n}, and C=𝒞0C=\mathcal{C}_{0}. We are going to prove that any function HCc(V)H\in C_{c}^{\infty}(V), with H|C0H|_{C^{\prime}}\equiv 0, satisfies ϕHt(p)C\phi_{H}^{t}(p^{\prime})\in C^{\prime} for all t[0,+)t\in[0,+\infty). According to Lemma 22, this will imply that CC^{\prime} is coisotropic.

Let HH be such a function and consider the function HθH\circ\theta. It is compactly supported in UU and can be extended by 0 outside UU to a continuous compactly supported function K:2nK\colon\thinspace\mathbb{R}^{2n}\to\mathbb{R}. Since HH is smooth and θ\theta is a symplectic homeomorphism, KCHam0(2n,ω0)K\in C^{0}_{\mathrm{Ham}}(\mathbb{R}^{2n},\omega_{0}). Since K|𝒞0=0K|_{\mathcal{C}_{0}}=0, Theorem 3 yields ϕKt(0)0(0)\phi_{K}^{t}(0)\in\mathcal{F}_{0}(0) for any t0t\geqslant 0. Since KK has support in UU, we have

(8) t0,ϕKt(0)0(0)U𝒞0U.\forall t\geqslant 0,\ \phi_{K}^{t}(0)\in\mathcal{F}_{0}(0)\cap U\subset\mathcal{C}_{0}\cap U.

Since ϕHt=θϕKtθ1\phi_{H}^{t}=\theta\phi_{K}^{t}\theta^{-1}, we deduce ϕHt(p)C\phi_{H}^{t}(p^{\prime})\in C^{\prime} as desired and hence that CC^{\prime} is coisotropic.

Denote \mathcal{F}^{\prime} the characteristic foliation of CC^{\prime}. From (8), we deduce that for any HCc(V)H\in C_{c}^{\infty}(V), H|C0H|_{C^{\prime}}\equiv 0,

t0,ϕHt(p)θ(0(0)U).\forall t\geqslant 0,\ \phi_{H}^{t}(p^{\prime})\in\theta(\mathcal{F}_{0}(0)\cap U).

Now according to Lemma 22, there exists a neighborhood WVW\subset V such that

W(p)=W{ϕHt(p)|t[0,+),HCc(V),H|C0}.W\cap\mathcal{F}^{\prime}(p^{\prime})=W\cap\{\phi_{H}^{t}(p^{\prime})\,|\,t\in[0,+\infty),H\in C_{c}^{\infty}(V),H|_{C^{\prime}}\equiv 0\}.

Thus,

W(p)Wθ(0(0)).W\cap\mathcal{F}^{\prime}(p^{\prime})\subset W\cap\theta(\mathcal{F}_{0}(0)).

We get the reverse inclusion by switching the roles of CC and CC^{\prime}, and we see that θ\theta sends locally 0(0)\mathcal{F}_{0}(0) onto (p)\mathcal{F}^{\prime}(p^{\prime}). ∎

Let us now turn to the proof of Theorem 5. The proof has three main ingredients: the Lagrangian case in Theorem 3 (i.e., Propositions 17 and 19), Theorem 1, and the fact that the graph of the characteristic foliation, given by

Γ()={(x,x)M×M|xC,x(x)},\Gamma(\mathcal{F})=\{(x,x^{\prime})\in M\times M\,|\,x\in C,x^{\prime}\in\mathcal{F}(x)\},

is Lagrangian in the product M×MM\times M endowed with the symplectic form ω(ω)\omega\oplus(-\omega), as long as it is a submanifold.

Proof of Theorem 5.

Let pCp\in C and (U,θ)(U,\theta) be a coisotropic chart of CC around pp sending pp to 0. The symplectic diffeomorphism Θ=θ×θ\Theta=\theta\times\theta, defined on U×UU\times U maps Γ()\Gamma(\mathcal{F}) to the graph of the standard characteristic foliation

Γ(0)={(x1,,xn,y1,,yn,x1,\displaystyle\Gamma(\mathcal{F}_{0})=\{(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n},x_{1}^{\prime}, ,xn,y1,,yn)2n×2n|\displaystyle\ldots,x_{n}^{\prime},y_{1}^{\prime},\ldots,y_{n}^{\prime})\in\mathbb{R}^{2n}\times\mathbb{R}^{2n}\,|
i{nk+1,\displaystyle\forall i\in\{n-k+1, ,n},yi=yi=0 and\displaystyle\ldots,n\},y_{i}=y_{i}^{\prime}=0\text{ and }
j{1,,nk},xi=xi,yi=yi}.\displaystyle\forall j\in\{1,\ldots,n-k\},x_{i}=x_{i}^{\prime},y_{i}=y_{i}^{\prime}\}\,.

Since Γ(0)\Gamma(\mathcal{F}_{0}) is a Lagrangian submanifold of 2n×2n\mathbb{R}^{2n}\times\mathbb{R}^{2n}, then Λ=Θ1(Γ(0))=(U×U)Γ()\Lambda=\Theta^{-1}(\Gamma(\mathcal{F}_{0}))=(U\times U)\cap\Gamma(\mathcal{F}) is a Lagrangian submanifold of M×MM\times M.

Now note that if HCHam0H\in C^{0}_{\mathrm{Ham}} then the function K:[0,1]×M×MK:[0,1]\times M\times M\to\mathbb{R} given by Kt(x,x)=Ht(x)Ht(x)K_{t}(x,x^{\prime})=H_{t}(x)-H_{t}(x^{\prime}) is a continuous Hamiltonian generating the hameotopy ϕH×ϕH\phi_{H}\times\phi_{H} (recall that the symplectic form on M×MM\times M is ω(ω)\omega\oplus(-\omega)).

Assume for a contradiction that HH is a function of time on every leaf of the characteristic foliation \mathcal{F} of CC and that for some point qCq\in C and some time t<1t<1, ϕHt(q)C\phi_{H}^{t}(q)\notin C. Set t0=sup{t0|ϕHt(q)C}t_{0}=\sup\{t\geqslant 0\,|\,\phi_{H}^{t}(q)\in C\}. Since CC is a closed subset of MM, p=ϕHt0(q)Cp=\phi_{H}^{t_{0}}(q)\in C and we can assume that the above construction yielding the construction of Λ\Lambda is performed in the neighborhood of this point. Consider the ”time-reparametrized” Hamiltonians H~\tilde{H}, K~\tilde{K} given by H~(t,x)=(1t0)H(t0+(1t0)t,x)\tilde{H}(t,x)=(1-t_{0})H(t_{0}+(1-t_{0})t,x) and K~(t,x,x)=H~(t,x)H~(t,x)\tilde{K}(t,x,x^{\prime})=\tilde{H}(t,x)-\tilde{H}(t,x^{\prime}). The fact that HH is a function of time on any leaf implies that the restriction of K~\tilde{K} to Γ()\Gamma(\mathcal{F}) is identically 0. In particular, it vanishes on the Lagrangian Λ\Lambda and according to Proposition 17 there exists ε>0\varepsilon>0 such that for all t[0,ε]t\in[0,\varepsilon], ϕK~t(p,p)=(ϕH~t×ϕH~t)(p,p)Λ\phi_{\tilde{K}}^{t}(p,p)=(\phi_{\tilde{H}}^{t}\times\phi_{\tilde{H}}^{t})(p,p)\in\Lambda. This implies that

ϕHt0+ε(1t0)(q)=ϕH~ε(p)C,\phi_{H}^{t_{0}+\varepsilon(1-t_{0})}(q)=\phi_{\tilde{H}}^{\varepsilon}(p)\in C,

which contradicts the maximality of t0t_{0}.

Conversely, assume that the flow ϕHt\phi_{H}^{t} preserves CC. By Theorem 1, ϕHt\phi_{H}^{t} sends leaves to leaves and in particular, it preserves the graph of the foliation. Therefore, for any point pCp\in C, we may apply Proposition 19 to the Lagrangian Λ\Lambda and the continuous Hamiltonian KK. We get that on a neighborhood of (p,p)(p,p), and for small times tt, KtK_{t} is constant. Since K0(p,p)=0K_{0}(p,p)=0 we get that K0K_{0} vanishes in a neighborhood of (p,p)(p,p). But this implies that H0H_{0} is constant on a neighborhood of pp in the leaf (p)\mathcal{F}(p). The argument can be performed for any pCp\in C and at any initial time tt instead of 0. It shows that HtH_{t} is locally constant, hence constant, on leaves. ∎

6. Defining C0C^{0}–Coisotropic submanifolds and their characteristic foliations

In this section we will use Theorem 1 to define C0C^{0}–coisotropic submanifolds and their characteristic foliations. Below, we assume that 2n\mathbb{R}^{2n} is equipped with the standard symplectic structure. Recall from the beginning of Section 5 that every coisotropic submanifold of codimension kk is locally symplectomorphic to

𝒞0={(x1,,xn,y1,,yn)|(ynk+1,,yn)=(0,,0)}2n,\displaystyle\mathcal{C}_{0}=\{(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})\,|\,(y_{n-k+1},\ldots,y_{n})=(0,\ldots,0)\}\subset\mathbb{R}^{2n}\;,

and that the leaf of its characteristic foliation, 0\mathcal{F}_{0}, passing through p=(a1,,an,b1,,bnk,0,,0)p=(a_{1},\ldots,a_{n},b_{1},\ldots,b_{n-k},0,\ldots,0) is given by

0(p)={(a1,,ank,xnk+1,,xn,b1,,\displaystyle\mathcal{F}_{0}(p)=\{(a_{1},\ldots,a_{n-k},x_{n-k+1},\ldots,x_{n},b_{1},\ldots, bnk,0,,0)|\displaystyle b_{n-k},0,\ldots,0)\,|\,
(xnk+1,,xn)k}.\displaystyle(x_{n-k+1},\ldots,x_{n})\in\mathbb{R}^{k}\}\;.
Definition 23.

A codimension–kk C0C^{0}–submanifold CC of a symplectic manifold (M,ω)(M,\omega) is C0C^{0}–coisotropic if around each point pCp\in C there exists a C0C^{0}–coisotropic chart, that is, a pair (U,θ)(U,\theta) with UU an open neighborhood of pp and θ:UV2n\theta\colon\thinspace U\rightarrow V\subset\mathbb{R}^{2n} a symplectic homeomorphism, such that θ(CU)=𝒞0V\theta(C\cap U)=\mathcal{C}_{0}\cap V.

A codimension–nn C0C^{0}–coisotropic submanifold is called a C0C^{0}–Lagrangian.

Example. Graphs of symplectic homeomorphisms are C0C^{0}–Lagrangians. Graphs of differentials of C1C^{1} functions and, more generally, graphs of C0C^{0} 1–forms, closed in the sense of distributions, provide a family of non trivial examples; see Proposition 26 for a proof.

Conversely, we could ask whether every continuous 1–form whose graph is a C0C^{0}–Lagrangian is closed in the sense of distributions. An affirmative answer in a particular case appears in Viterbo [26, Corollary 22].

As a consequence of Theorem 1, C0C^{0}–coisotropic submanifolds carry (C0C^{0}–) characteristic foliations in the following sense.

Proposition 24.

Any C0C^{0}–coisotropic submanifold CC admits a unique C0C^{0}–foliation \mathcal{F} which is mapped to 0\mathcal{F}_{0} by any C0C^{0}–coisotropic chart.

Proof.

If such a foliation exists it has to coincide with θ1(0)\theta^{-1}(\mathcal{F}_{0}) on the domain of any C0C^{0}–coisotropic chart φ\varphi. The only thing to check is that for any two C0C^{0}–coisotropic charts θ1:U1V1\theta_{1}\colon\thinspace U_{1}\to V_{1} and θ2:U2V2\theta_{2}\colon\thinspace U_{2}\to V_{2}, the foliations θ11(0)\theta_{1}^{-1}(\mathcal{F}_{0}) and θ21(0)\theta_{2}^{-1}(\mathcal{F}_{0}) coincide on U1U2U_{1}\cap U_{2}. But this follows immediately from Theorem 1 applied to C=𝒞0C=\mathcal{C}_{0} and θ=θ1θ21:θ2(U1U2)θ1(U1U2)\theta=\theta_{1}\theta_{2}^{-1}\colon\thinspace\theta_{2}(U_{1}\cap U_{2})\to\theta_{1}(U_{1}\cap U_{2}).∎

Theorem 1 states that a smooth C0C^{0}–coisotropic submanifold is coisotropic and its natural C0C^{0}–foliation coincides with its characteristic foliation.

Example. If C=θ(C)C=\theta(C^{\prime}), with CC^{\prime} a smooth coisotropic submanifold and θ\theta a symplectic homeomorphism, then =θ()\mathcal{F}=\theta(\mathcal{F}^{\prime}) where \mathcal{F}^{\prime} is the characteristic foliation of CC^{\prime}.

One may wonder if every topological hypersurface is C0C^{0}–coisotropic. It is possible to show, via an application of Proposition 24, that the boundary of the standard cube in 4\mathbb{R}^{4} does not possess a C0C^{0}–characteristic foliation, and hence, it is not C0C^{0}–coisotropic.

The following proposition tells us that Theorems 3 and 5 hold for C0C^{0}–coisotropic submanifolds.

Proposition 25.

Denote by CC a connected C0C^{0}–coisotropic submanifold of a symplectic manifold (M,ω)(M,\omega) which is closed as a subset of MM. Let HCHam0H\in C^{0}_{\mathrm{Ham}} with induced hameotopy ϕH\phi_{H}.

  1. (1)

    The restriction of HH to CC is a function of time if and only if ϕH\phi_{H} preserves CC and flows along the leaves of its (C0C^{0}–)characteristic foliation.

  2. (2)

    The restriction of HH to each leaf of the characteristic foliation of CC is a function of time if and only if the flow ϕH\phi_{H} preserves CC.

The above can be proven by adapting the proofs of Theorems 3 and 5 to C0C^{0}–coisotropics. We will not provide a proof for Proposition 25 here, and we only mention that to adapt the proofs one would have to introduce C0C^{0}–coisotropic charts and use the following simple fact: if θ\theta is a symplectic homeomorphism and HCHam0H\in C^{0}_{Ham} then HθCHam0H\circ\theta\in C^{0}_{Ham} and ϕHθt=θ1ϕHtθ.\phi^{t}_{H\circ\theta}=\theta^{-1}\phi^{t}_{H}\theta.

Finally, we provide a family of non trivial examples of C0C^{0}–Lagrangians.

Proposition 26.

Let α\alpha be a C0C^{0} 1–form on a smooth manifold NN which is closed in the sense of distributions. Then, its graph, graph(α)TN\mathrm{graph}(\alpha)\subset T^{*}N, is a C0C^{0}–Lagrangian.

Proof.

Since the statement is local, it is sufficient to prove it when NN is an open set in n\mathbb{R}^{n}. Then α\alpha can be written as α=i=1npi(x)dxi\alpha=\sum_{i=1}^{n}p_{i}(x)dx_{i}, where x1,,xnx_{1},\ldots,x_{n} are the canonical coordinates in n\mathbb{R}^{n} and p1,,pnp_{1},\ldots,p_{n} continuous functions on NN. The fact that α\alpha is closed is equivalent to the equations

(9) i,j{1,,n},jpi=ipj,\forall i,j\in\{1,\ldots,n\},\;\partial_{j}p_{i}=\partial_{i}p_{j},

where ipj\partial_{i}p_{j} is the ii–th partial derivative of pjp_{j} in the sense of distributions.

We use convolution to approximate α\alpha. To that end, take a compactly supported smooth function ρ\rho such that ρ0\rho\geqslant 0, and Nρ(x)𝑑x=1\int_{N}\rho(x)dx=1 and set ρε(x)=1εnρ(xε)\rho_{\varepsilon}(x)=\frac{1}{\varepsilon^{n}}\rho\!\left(\frac{x}{\varepsilon}\right) for every ε>0\varepsilon>0. For any continuous function ff on NN, the functions

fρε(x)=Nf(y)ρε(xy)𝑑yf\ast\rho_{\varepsilon}(x)=\int_{N}f(y)\rho_{\varepsilon}(x-y)dy

are well-defined on any compact subset of NN for ε\varepsilon small enough. Moreover, for any ε\varepsilon, fρεf\ast\rho_{\varepsilon} is smooth, converges locally uniformly to ff as ε\varepsilon goes to 0, its differential satisfies d(fρε)=(df)ρεd(f\ast\rho_{\varepsilon})=(df)\ast\rho_{\varepsilon} and converges in the sense of distributions to dfdf.

Let UNU\Subset N be an open subset of NN. Then, for ε\varepsilon small enough,

αε=i=1npiρεdxi\alpha_{\varepsilon}=\sum_{i=1}^{n}p_{i}\ast\rho_{\varepsilon}\,dx_{i}

is a well-defined 1–form on UU. It satisfies Equations (9) and thus is closed. Moreover, it converges uniformly to α\alpha on UU.

Now let ϕε\phi_{\varepsilon} be the family of symplectic diffeomorphisms of TUT^{\ast}U defined by ϕε(x,p)=(x,p+αε(x))\phi_{\varepsilon}(x,p)=(x,p+\alpha_{\varepsilon}(x)). They converge uniformly on UU to the symplectic homeomorphism ϕ:TUTU\phi\colon\thinspace T^{\ast}U\to T^{\ast}U, (x,p)(x,p+α(x))(x,p)\mapsto(x,p+\alpha(x)) and graph(α)\mathrm{graph}(\alpha) restricted to TUT^{\ast}U is ϕ(U)\phi(U). This shows that graph(α)\mathrm{graph}(\alpha) is locally the image of a smooth Lagrangian by a symplectic homeomorphism. ∎

Appendix A The main results for closed Lagrangian

In this section we provide relatively simple proofs for Theorems 1, 3, and 5 in an enlightening and important special case. We suppose that M=TLM=T^{*}L equipped with its canonical symplectic structure for some closed smooth manifold LL. Denote by θ\theta a symplectic homeomorphism of TLT^{*}L. And let L=θ(L0)L^{\prime}=\theta(L_{0}), where L0L_{0} denotes the 0–section of TLT^{*}L.

Below, we will prove Theorems 1, 3, and 5 in the special case where the coisotropic CC is taken to be the zero section L0L_{0}. In this case Theorem 1 states the following:

Theorem 27.

If LL^{\prime} is smooth, then it is Lagrangian.

In the settings considered in this appendix, Theorems 3 and 5 coincide and state the following:

Theorem 28.

Let HCHam0H\in C^{0}_{\mathrm{Ham}} with induced hameotopy ϕH\phi_{H}. The restriction of HH to L0L_{0} is a function of time if and only if ϕH\phi_{H} preserves L0L_{0}.

We believe that the above special cases provide the reader with the opportunity to get an idea of the proofs of our main results without having to go through the technical details of Sections 4 and 5.

We will first show that Theorem 27 follows from Theorem 28. In order to do so we will need the following dynamical characterizations of isotropic and coisotropic submanifolds, respectively.

Lemma 29.

Let II denote a (smooth) submanifold of a symplectic manifold (M,ω)(M,\omega). The following are equivalent:

  • II is isotropic,

  • For every smooth Hamiltonian HH, if ϕH\phi_{H} preserves II, then H|IH|_{I} is a function of time only.

Lemma 30.

Let CC denote a (smooth) submanifold of a symplectic manifold (M,ω)(M,\omega). The following are equivalent:

  • CC is coisotropic,

  • For every smooth Hamiltonian HH, if H|CH|_{C} is a function of time only, then ϕH\phi_{H} preserves CC.

We leave the proofs of the above lemmas, which follow from symplectic linear algebra, to the reader. In the proof of Theorem 1, we use Lemma 22 which is a variation of the second of the above two lemmas.

Proof of Theorem 27.

Each of Lemmas 29 and 30 gives a different proof. We provide both proofs here.

First proof: Suppose that HH is any smooth Hamiltonian whose flow ϕH\phi_{H} preserves LL^{\prime}. Then HθCHam0H\circ\theta\in C^{0}_{\mathrm{Ham}} and its flow, θ1ϕHtθ\theta^{-1}\phi^{t}_{H}\theta, preserves L0L_{0}. It follows from Theorem 28 that the restriction of HθH\circ\theta to L0L_{0} is a function of time only. Therefore, H|LH|_{L^{\prime}} depends on time only, and so using Lemma 29 we conclude that LL^{\prime} is isotropic.

Second proof: Suppose that HH is any smooth Hamiltonian whose restriction to LL^{\prime} is a function of time only. Then HθCHam0H\circ\theta\in C^{0}_{\mathrm{Ham}} and its restriction to L0L_{0} depends only on time. It follows from Theorem 28 that the flow of HθH\circ\theta, which is θ1ϕHtθ\theta^{-1}\phi^{t}_{H}\theta, preserves L0L_{0} and so the flow of HH preserves LL^{\prime}. Using Lemma 30 we conclude that LL^{\prime} is coisotropic. ∎

Proof of Theorem 28.

To prove the direct implication suppose that Ht|L0=c(t)H_{t}|_{L_{0}}=c(t), where c(t)c(t) is a function of time only. For a contradiction assume that ϕH\phi_{H} does not preserve L0L_{0}, then for some t0t_{0} we have ϕHt0(L0)L0\phi^{t_{0}}_{H}(L_{0})\not\subset L_{0}, and after the time reparametrization tt0tt\mapsto t_{0}t we may assume that t0=1t_{0}=1, that is, ϕH1(L0)L0\phi_{H}^{1}(L_{0})\not\subset L_{0}.

Since HCHam0H\in C^{0}_{\mathrm{Ham}} there exists a sequence of smooth Hamiltonians Hi:[0,1]×MH_{i}\colon\thinspace[0,1]\times M\to\mathbb{R} such that HiH_{i} converges uniformly to HH and ϕHi\phi_{H_{i}} converges to ϕH\phi_{H} in C0C^{0}–topology.

Because ϕH1(L0)L0\phi^{1}_{H}(L_{0})\not\subset L_{0}, there exists a ball BB such that BL0B\cap L_{0}\neq\emptyset and ϕH1(B)L0=\phi^{1}_{H}(B)\cap L_{0}=\emptyset. It follows that ϕHi1(L0)B=\phi^{-1}_{H_{i}}(L_{0})\cap B=\emptyset for large ii. And so,

γ(ϕHi1(L0),L0)=γ(ϕHi1(L0),L0)cLR(B;L0)>0.\gamma(\phi_{H_{i}}^{1}(L_{0}),L_{0})=\gamma(\phi_{H_{i}}^{-1}(L_{0}),L_{0})\geqslant c_{\mathrm{LR}}(B;L_{0})>0.

Inequality (2) from Section 2.2 implies that

maxt[0,1](osc(Hi(t,)|L0))cLR(B;L0),\max_{t\in[0,1]}(\mathrm{osc}(H_{i}(t,\cdot)|_{L_{0}}))\geqslant c_{\mathrm{LR}}(B;L_{0}),

contradicting the fact that H|L0H|_{L_{0}} is a function of time. We conclude that ϕH\phi_{H} preserves L0L_{0}.

Next, to prove the converse implication suppose that ϕH\phi_{H} preserves L0L_{0}. We will show that H(0,)|L0H(0,\cdot)|_{L_{0}} is constant. A time reparametrization argument, where H(t,x)H(t,x) is replaced with H~(t,x)=(1s)H(s+(1s)t,x)\tilde{H}(t,x)=(1-s)H(s+(1-s)t,x), would then show that H(s,)|L0H(s,\cdot)|_{L_{0}} is constant for any choice of s[0,1)s\in[0,1). This in turn would imply that H|L0H|_{L_{0}} is a function of time.

Let BB denote an open ball intersecting L0L_{0} and UU a small open neighborhood of B¯\overline{B}, such that L0π(U)L_{0}\setminus\pi(U) has a non-empty interior, where π:TLL0\pi\colon\thinspace T^{*}L\to L_{0} is the natural projection. (Picking UU in this way enables us to apply Lemma 8.) Let ψ\psi be any symplectomorphism supported in BB and preserving L0L_{0}. Next, we pick ε>0\varepsilon>0 such that ϕHt(B),(ϕHt)1(B)U\phi^{t}_{H}(B),(\phi^{t}_{H})^{-1}(B)\Subset U for all t[0,ε]t\in[0,\varepsilon]. By a reparametrization in time, where H(t,x)H(t,x) is replaced with εH(εt,x)\varepsilon H(\varepsilon t,x), we may assume that ϕHt(B)\phi^{t}_{H}(B), (ϕHt)1(B)U(\phi^{t}_{H})^{-1}(B)\Subset U for all t[0,1]t\in[0,1].

Consider the C0C^{0}–Hamiltonian G=(HψH)ϕHG=(H\circ\psi-H)\circ\phi_{H}. We will now show that G|L0=0G|_{L_{0}}=0. The support of GG is included in t[0,1](ϕHt)1(B)U\cup_{t\in[0,1]}(\phi_{H}^{t})^{-1}(B)\subset U, and moreover, its flow is (ϕHt)1ψ1ϕHtψ(\phi_{H}^{t})^{-1}\psi^{-1}\phi_{H}^{t}\psi. Because ψ\psi and ϕH\phi_{H} preserve L0L_{0} the flow of GG also preserves L0L_{0}.

Since GCHam0G\in C^{0}_{\mathrm{Ham}} there exist smooth Hamiltonians GiG_{i} such that {Gi}\{G_{i}\} converges uniformly to GG and {ϕGi}\{\phi_{G_{i}}\} converges to ϕG\phi_{G}. Furthermore, we can require that all GiG_{i}’s are supported in UU. This can be achieved by picking a corresponding sequence of smooth Hamiltonians HiH_{i} for HH and defining Gi=(HiψHi)ϕHiG_{i}=(H_{i}\circ\psi-H_{i})\circ\phi_{H_{i}}. For large ii, GiG_{i} is supported in UU.

Fix a small r>0r>0. Because ϕGt(L0)=L0\phi^{t}_{G}(L_{0})=L_{0} for any t[0,1],t\in[0,1], for sufficiently large ii we have ϕGit(L0)TrL0\phi^{t}_{G_{i}}(L_{0})\subset T^{*}_{r}L_{0}. Furthermore, the Hamiltonians GiG_{i} are all supported in UU and hence we can apply Lemma 8 and conclude that γ(ϕGi1(L0),L0)Cr\gamma(\phi^{1}_{G_{i}}(L_{0}),L_{0})\leqslant Cr, i.e γ(ϕGi1(L0),L0)0\gamma(\phi^{1}_{G_{i}}(L_{0}),L_{0})\to 0. Of course, the same reasoning yields γ(ϕGit(L0),L0)0\gamma(\phi^{t}_{G_{i}}(L_{0}),L_{0})\to 0 for all t[0,1].t\in[0,1]. Then, Proposition 20 implies that G|L0=c(t)G|_{L_{0}}=c(t). Since it has support in UU, we conclude that G|L0=0G|_{L_{0}}=0.

In particular, G|L0=0G|_{L_{0}}=0 at time 0. Now since the ball BB can contain any chosen pair of points x1x_{1}, x2L0x_{2}\in L_{0}, and ψ\psi can be chosen so that ψ(x1)=x2\psi(x_{1})=x_{2}, we conclude that the restriction H(0,)|L0H(0,\cdot)|_{L_{0}} is constant. ∎

References

  • [1] J.-F. Barraud and O. Cornea. Lagrangian intersections and the Serre spectral sequence. Ann. of Math. (2), 166(3):657–722, 2007.
  • [2] M. S. Borman and M. McLean. Bounding Lagrangian widths via geodesic paths. Compos. Math., 150(12):2143–2183, 2014.
  • [3] L. Buhovsky and S. Seyfaddini. Uniqueness of generating Hamiltonians for topological Hamiltonian flows. J. Symplectic Geom., 11(1):37–52, 2013.
  • [4] F. Charette. A geometric refinement of a theorem of Chekanov. J. Symplectic Geom., 10(3):475–491, 2012.
  • [5] Y. Eliashberg and N. Mishachev. Introduction to the hh-principle, volume 48 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2002.
  • [6] U. Frauenfelder, V. L. Ginzburg, and F. Schlenk. Energy capacity inequalities via an action selector. In Geometry, spectral theory, groups, and dynamics, volume 387 of Contemp. Math., pages 129–152. Amer. Math. Soc., Providence, RI, 2005.
  • [7] V. L. Ginzburg. Coisotropic intersections. Duke Math. J., 140(1):111–163, 2007.
  • [8] M. J. Gotay. On coisotropic imbeddings of presymplectic manifolds. Proc. Amer. Math. Soc., 84(1):111–114, 1982.
  • [9] M. Gromov. Partial differential relations, volume 9 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1986.
  • [10] V. Humilière, R. Leclercq, and S. Seyfaddini. New energy-capacity-type inequalities and uniqueness of continuous Hamiltonians. Comment. Math. Helv., to appear (arXiv:1209.2134).
  • [11] F. Laudenbach and J.-C. Sikorav. Hamiltonian disjunction and limits of Lagrangian submanifolds. Internat. Math. Res. Notices, (4):161 ff., approx. 8 pp. (electronic), 1994.
  • [12] R. Leclercq. Spectral invariants in Lagrangian Floer theory. J. Mod. Dyn., 2(2):249–286, 2008.
  • [13] R. Leclercq and F. Zapolsky. Spectral invariants for monotone Lagrangian submanifolds. (in preparation).
  • [14] P. Libermann and C.-M. Marle. Symplectic geometry and analytical mechanics, volume 35 of Mathematics and its Applications. D. Reidel Publishing Co., Dordrecht, 1987. Translated from the French by Bertram Eugene Schwarzbach.
  • [15] S. Lisi and A. Rieser. Coisotropic Hofer–Zehnder capacities and non-squeezing for relative embeddings. ArXiv:1312.7334, 2013.
  • [16] D. Milinković. On equivalence of two constructions of invariants of lagrangian submanifolds. Pacific J. of Math., 195:371–415, 2000.
  • [17] D. Milinković. Geodesics on the space of Lagrangian submanifolds in cotangent bundles. Proc. Amer. Math. Soc., 129(6):1843–1851 (electronic), 2001.
  • [18] A. Monzner, N. Vichery, and F. Zapolsky. Partial quasimorphisms and quasistates on cotangent bundles, and symplectic homogenization. J. Mod. Dyn., 6(2):205–249, 2012.
  • [19] Y.-G. Oh. Symplectic topology as the geometry of action functional. I. Relative Floer theory on the cotangent bundle. J. Differential Geom., 46(3):499–577, 1997.
  • [20] Y.-G. Oh. Locality of continuous Hamiltonian flows and Lagrangian intersections with the conormal of open subsets. J. Gökova Geom. Topol. GGT, 1:1–32, 2007.
  • [21] Y.-G. Oh. Geometry of generating functions and Lagrangian spectral invariants. ArXiv:1206.4788, June 2012.
  • [22] Y.-G. Oh and S. Müller. The group of Hamiltonian homeomorphisms and C0C^{0}–symplectic topology. J. Symplectic Geom., 5(2):167–219, 2007.
  • [23] E. Opshtein. C0C^{0}–rigidity of characteristics in symplectic geometry. Ann. Sci. Éc. Norm. Supér. (4), 42(5):857–864, 2009.
  • [24] M. Usher. The sharp energy-capacity inequality. Commun. Contemp. Math., 12(3):457–473, 2010.
  • [25] C. Viterbo. Symplectic topology as the geometry of generating functions. Math. Annalen, 292:685–710, 1992.
  • [26] C. Viterbo. On the uniqueness of generating Hamiltonian for continuous limits of Hamiltonians flows. Int. Math. Res. Not., pages Art. ID 34028, 9, 2006.
  • [27] C. Viterbo. Symplectic homogenization. ArXiv:0801.0206, Dec. 2008.
  • [28] A. Weinstein. Symplectic geometry. Bull. Amer. Math. Soc. (N.S.), 5(1):1–13, 1981.