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

Hameomorphism groups of positive genus surfaces

Cheuk Yu Mak and Ibrahim Trifa
Abstract.

In their previous works [CGHM+21, CGHM+22], Cristofaro-Gardiner, Humilière, Mak, Seyfaddini and Smith defined links spectral invariants on connected compact surfaces and used them to show various results on the algebraic structure of the group of area-preserving homeomorphisms of surfaces, particularly in cases where the surfaces have genus zero. We show that on surfaces with higher genus, for a certain class of links, the invariants will satisfy a local quasimorphism property. Subsequently, we generalize their results to surfaces of any genus. This extension includes the non-simplicity of (i) the group of hameomorphisms of a closed surface, and (ii) the kernel of the Calabi homomorphism inside the group of hameomorphisms of a surface with non-empty boundary. Moreover, we prove that the Calabi homomorphism extends (non-canonically) to the C0C^{0}-closure of the set of Hamiltonian diffeomorphisms of any surface. The local quasimorphism property is a consequence of a quantitative Künneth formula for a connected sum in Heegaard Floer homology, inspired by results of Ozsváth and Szabó.

1. Introduction

Let Σ\Sigma be a compact connected orientable surface (possibly with boundary) equipped with an area from ω\omega. In 1980’s, Fathi defined the mass-flow homomorphism [Fat80]

Homeoc(Σ,ω)\operatorname{Homeo}_{c}(\Sigma,\omega)\to\mathbb{R}

from the group of area-preserving homeomorphisms supported in the interior of Σ\Sigma to \mathbb{R}. Whether its kernel is a simple group was an open question for a long time and has recently been resolved negatively using techniques from symplectic geometry. The case of the sphere was answered by [CGHS20] using periodic Floer homology, building on the work of [Hut11] and [CGHR15]. The case of positive genus surfaces was answered by [CGHM+21] using Lagrangian Floer theory, borrowing ideas from [OS04], [MS21] and [PS21].

Symplectic geometry enters the picture because ω\omega is a symplectic form and the kernel of the mass-flow homomorphism can be identified with the C0C^{0} closure of the group Ham(Σ)\operatorname{Ham}(\Sigma) of Hamiltonian diffeomorphisms supported in the interior of Σ\Sigma. The Hofer metric, a bi-invariant and non-degenerate metric, on Ham(Σ)\operatorname{Ham}(\Sigma) enables us to define two natural normal subgroups of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma), namely the group of hameomorphisms Hameo(Σ)\operatorname{Hameo}(\Sigma) and the group of finite energy homeomorphisms FHomeo(Σ)\operatorname{FHomeo}(\Sigma) (see Section 2.1 for the precise definitions, and also [OM07], [CGHS20] for more discussions). Indeed, the authors of [CGHM+21] show that the subgroup Hameo(Σ)\operatorname{Hameo}(\Sigma) is always a proper normal subgroup.

Since then, the method has been pushed further to answer more refined questions about the algebraic structure of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma), especially when Σ\Sigma has genus 0, using a property called the quasimorphism property. The goal of this paper is to generalize the results of [CGHM+22] to all surfaces even though we no longer have the quasimorphism property for positive genus surfaces.

1.1. Link spectral invariants and known results for genus zero surfaces

Link spectral invariants are introduced in [CGHM+21] as the main tool to study Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma). Given a Lagrangian link (i.e. a union of disjoint circles) L¯=L1Lk\underline{L}=L_{1}\cup...\cup L_{k} satisfying certain monotonicity conditions on a surface (Σ,ω)(\Sigma,\omega), we can associate a spectral invariant cL¯:C(S1×Σ,)c_{\underline{L}}:C^{\infty}(S^{1}\times\Sigma,\mathbb{R})\rightarrow\mathbb{R} which satisfies several useful properties (Proposition 13).

In particular, the homotopy invariance permits to define cL¯(φ)c_{\underline{L}}(\varphi) for φHam~(Σ)\varphi\in\widetilde{\operatorname{Ham}}(\Sigma), by the formula cL¯({φHt}t[0,1])=cL¯(H)c_{\underline{L}}(\{\varphi_{H}^{t}\}_{t\in[0,1]})=c_{\underline{L}}(H) for a mean normalized Hamiltonian function HH. The homogenization μL¯\mu_{\underline{L}} of cL¯c_{\underline{L}} is defined by

μL¯(φ)=limncL¯(φn)n.\mu_{\underline{L}}(\varphi)=\lim_{n\to\infty}\frac{c_{\underline{L}}(\varphi^{n})}{n}.

In the case of Σ=S2\Sigma=S^{2}, we have the following :

Theorem 1 (Theorem 7.7 of [CGHM+21]).

cL¯:Ham~(S2)c_{\underline{L}}:\widetilde{\operatorname{Ham}}(S^{2})\to\mathbb{R} is a quasimorphism with defect Dk+1kλD\leqslant\frac{k+1}{k}\lambda where λ\lambda is the monotonicity constant of L¯\underline{L}. Moreover, μL¯\mu_{\underline{L}} descends to a quasimorphism on Ham(S2)\operatorname{Ham}(S^{2}).

The fact that μL¯\mu_{\underline{L}} is a quasimorphism and that we can quantify its defect is the key ingredient to prove the following results (the definition of Cal\operatorname{Cal} will be recalled in Section 2.1):

  1. (1)

    The Calabi homomorphism Cal:Hameo(D2)\operatorname{Cal}:\operatorname{Hameo}(D^{2})\to\mathbb{R} can be extended to Ham¯(D2)=Homeo(D2,ω)\overline{\operatorname{Ham}}(D^{2})=\operatorname{Homeo}(D^{2},\omega). ([CGHM+22, Theorem 1.9])

  2. (2)

    Ker(Cal)Hameo(D2)\operatorname{Ker}(\operatorname{Cal})\cap\operatorname{Hameo}(D^{2}) is not simple. ([CGHM+22, Theorem 1.3])

  3. (3)

    Hameo(S2)\operatorname{Hameo}(S^{2}) is not simple. ([CGHM+22, Theorem 1.3])

1.2. Main results for positive genus surfaces

The purpose of this paper is to generalize (1) and (2) to any compact oriented surface Σ\Sigma (of any genus) with non-empty boundary:

Theorem 2.

The Calabi homomorphism Hameo(Σ)\operatorname{Hameo}(\Sigma)\to\mathbb{R} can be extended to Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma).

Theorem 3.

Ker(Cal)Hameo(Σ)\operatorname{Ker}(\operatorname{Cal})\cap\operatorname{Hameo}(\Sigma) is not simple.

and generalize (3) to any connected closed oriented surface (Σ,ω)(\Sigma,\omega):

Theorem 4.

Hameo(Σ,ω)\operatorname{Hameo}(\Sigma,\omega) is not simple.

Theorem 3 and 4 together answer a question in [OM07, Problem 4] for all surfaces.

There is a fundamental difference between the genus 0 and positive genus case: cL¯c_{\underline{L}} and μL¯\mu_{\underline{L}} are never quasimorphisms for positive genus surfaces for any L¯\underline{L} (cf. Proposition 16). To remedy this, we need to prove a local version of the quasimorphism property when Σ\Sigma has positive genus and combine it with the fragmentation technique. This requires a slightly different class of Lagrangian links (see Definition 14) than those in [CGHM+21]. We define the spectral invariants cL¯c_{\underline{L}} for this new class of links, show that they satisfy all the usual spectral invariant properties listed in Proposition 13, as well as the following local quasimorphism property.

Theorem 5.

Let L¯\underline{L} be an admissible link with kk contractible components, with monotonicity constant λ\lambda (see Definition 14). Let DΣD\subset\Sigma be a disk that does not intersect the non-contractible components of L¯\underline{L}, and denote by HamD(Σ)\operatorname{Ham}_{D}(\Sigma) the Hamiltonian diffeomorphisms supported in DD. Then, the restriction of cL¯c_{\underline{L}} to HamD(Σ)\operatorname{Ham}_{D}(\Sigma) is a quasimorphism with defect bounded by k+1k+gλ\frac{k+1}{k+g}\lambda.

The construction of cL¯c_{\underline{L}} and the proof of its local quasimorphism property relies on the following Künneth formula for connected sums in Heegaard Floer Homology, similar to the stabilization result of [OS04], which is proved by identifying moduli spaces of holomorphic maps under degeneration:

Theorem 6.

Consider two transverse η\eta-monotone admissible Lagrangian links L¯\underline{L} and K¯\underline{K} with kk components on a closed surface (Σ,ω)(\Sigma,\omega). Let (E,ωE)(E,\omega_{E}) denote the two-dimension torus, and α\alpha be a non-contractible circle on EE. Let α\alpha^{\prime} be a small Hamiltonian deformation of α\alpha, such that α\alpha and α\alpha^{\prime} are transverse. Then for an appropriate choice of almost complex structure, there is an isomorphism of filtered chain complexes

CF(SymL¯,SymK¯)CF(α,α)CF(Sym(L¯α),Sym(K¯α))CF^{*}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K})\otimes CF^{*}(\alpha,\alpha^{\prime})\xrightarrow{\sim}CF^{*}(\operatorname{Sym}(\underline{L}\cup\alpha),\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}))

where the LHS is computed considering the links L¯\underline{L} and K¯\underline{K} in (Σ,ω)(\Sigma,\omega), α\alpha and α\alpha^{\prime} in (E,ωE)(E,\omega_{E}), while in the RHS, L¯α\underline{L}\cup\alpha and K¯α\underline{K}\cup\alpha^{\prime} are links in the connected sum (Σ#E,ω)(\Sigma\#E,\omega^{\prime}) (where we perform the connected sum between a point σ1Σ\sigma_{1}\in\Sigma away from the links L¯\underline{L} and K¯\underline{K}, and a point σ2E\sigma_{2}\in E away from the isotopy between α\alpha and α\alpha^{\prime}).

If we forget the filtration, Thereom 6 is an identification of generators and differentials so it doesn’t depend on the symplectic form. To guarantee that the filtration also agrees, the symplectic form ω\omega^{\prime} on Σ#E\Sigma\#E is chosen such that it equals to ω\omega away from a neighborhood B(σ1)B(\sigma_{1}) of σ1\sigma_{1} which does not intersect L¯K¯\underline{L}\cup\underline{K}, equals to ωE\omega_{E} over the support KαK_{\alpha} of the Hamiltonian isotopy from α\alpha to α\alpha^{\prime}, and satisfies ω(Σ#E)=ω(Σ)\omega^{\prime}(\Sigma\#E)=\omega(\Sigma) (so we need to assume that ωE(Kα)<ω(B(σ1))\omega_{E}(K_{\alpha})<\omega(B(\sigma_{1})) for ω\omega^{\prime} to exist).

Structure of the paper

We collect some preliminaries in Section 2. The new class of Lagrangian links and the proof of its local quasimorphism property (Theorem 5) are given in Section 3. Section 4 is devoted to the proof of the main results, Theorem 2, 3 and 4. Theorem 6 is proved in Section 5.

Acknowledgement

The authors thank Sobhan Seyfaddini for suggesting the exploration of a local quasimorphism property. They also thank Ivan Smith for discussions regarding the Floer theory of symmetric products, Kristen Hendricks for discussions concerning Heegaard Floer homology, and Vincent Humilière for discussions about properties of spectral invariants. C.M. was supported by the Royal Society University Research Fellowship while working on this project. I.T. was supported by the École Normale Supérieure while working on this project. I.T. was also partially supported by the ERC Starting number 851701.

2. Preliminaries

2.1. Subgroups of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma)

Let Σ\Sigma be a compact connected surface equipped with an area form ω\omega. We start by introducing some conventions and notations, which we follow closely from [CGHM+21]:

  • Given a Hamiltonian H:S1×ΣH:S^{1}\times\Sigma\to\mathbb{R}, the Hamiltonian diffeomorphism ϕH1\phi^{1}_{H} is the time 1 flow of the Hamiltonian vector field XHtX_{H_{t}} defined by ιXHtω=dHt\iota_{X_{H_{t}}}\omega=dH_{t};

  • Given two Hamiltonians HH and KK, we define the composition by (H#K)t(x):=Ht(x)+Kt((ϕHt)1(x))(H\#K)_{t}(x):=H_{t}(x)+K_{t}((\phi^{t}_{H})^{-1}(x));

  • We denote by Ham(Σ)\operatorname{Ham}(\Sigma) the group of Hamiltonian diffeomorphisms of Σ\Sigma supported in the interior of Σ\Sigma (it is often denoted Hamc(Σ)\operatorname{Ham}_{c}(\Sigma) in the literature);

  • Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) its closure for the C0C^{0} distance;

  • the Hofer norm of a Hamiltonian is HtHof:=01oscHtdt=01(maxHtminHt)𝑑t||H_{t}||_{\operatorname{Hof}}:=\int_{0}^{1}\operatorname{osc}H_{t}dt=\int_{0}^{1}(\max H_{t}-\min H_{t})dt;

  • the Hofer norm of a Hamiltonian diffeomorphism is φHof:=infHt,φ=φHt1HtHof||\varphi||_{\operatorname{Hof}}:=\inf\limits_{H_{t},\varphi=\varphi^{1}_{H_{t}}}||H_{t}||_{\operatorname{Hof}};

  • the Hofer distance on Ham(Σ)\operatorname{Ham}(\Sigma) is dH(φ,ψ):=φψ1Hofd_{H}(\varphi,\psi):=||\varphi\psi^{-1}||_{\operatorname{Hof}};

We define some subgroups of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) (cf. [OM07] and [CGHS20]):

Definition 7.

φHam¯(Σ,ω)\varphi\in\overline{\operatorname{Ham}}(\Sigma,\omega) is called a finite energy homeomorphism if there exists a sequence of smooth Hamiltonians HiH_{i} such that :

  • ϕHi1C0φ\phi^{1}_{H_{i}}\xrightarrow{C^{0}}\varphi

  • There exists C0C\geqslant 0 such that for every ii,

    HiHof:=01osc(Hi,t)𝑑tC||H_{i}||_{\operatorname{Hof}}:=\int_{0}^{1}\operatorname{osc}(H_{i,t})dt\leqslant C
Definition 8.

φHam¯(Σ,ω)\varphi\in\overline{\operatorname{Ham}}(\Sigma,\omega) is called a hameomorphism if there exists an isotopy (ψt)t[0,1](\psi^{t})_{t\in[0,1]} in Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) from Id\operatorname{Id} to φ\varphi and a sequence of smooth Hamiltonians HiH_{i} supported in a compact subset KK of the interior of Σ\Sigma such that :

  • ϕHitC0ψt\phi^{t}_{H_{i}}\xrightarrow{C^{0}}\psi^{t} uniformly in t[0,1]t\in[0,1];

  • (Hi)(H_{i}) is a Cauchy sequence for the Hofer norm.

We denote the group of finite energy homeomorphisms by FHomeo(Σ,ω)\operatorname{FHomeo}(\Sigma,\omega), and the group of hameomorphisms by Hameo(Σ,ω)\operatorname{Hameo}(\Sigma,\omega). When Σ\Sigma has non-empty boundary, one can define the Calabi invariant Cal:Ham(Σ)\operatorname{Cal}:\operatorname{Ham}(\Sigma)\to\mathbb{R} as follow: let φHam(Σ)\varphi\in\operatorname{Ham}(\Sigma), and HtH_{t} be a Hamiltonian supported in the interior of Σ\Sigma such that φ=ϕHt1\varphi=\phi^{1}_{H_{t}}. Then,

Cal(φ)=01ΣHtω𝑑t\operatorname{Cal}(\varphi)=\int_{0}^{1}\int_{\Sigma}H_{t}\omega dt

This definition does not depend on the choice of the Hamiltonian HtH_{t}, and Cal\operatorname{Cal} is a group homomorphism.

As shown in [CGHM+21], Cal\operatorname{Cal} can be extended canonically to a group homomorphism Hameo(Σ)R\operatorname{Hameo}(\Sigma)\to R by the formula Cal(φ)=limiCal(φHi1)\operatorname{Cal}(\varphi)=\lim\limits_{i\to\infty}\operatorname{Cal}(\varphi^{1}_{H_{i}}), where we consider any sequence (Hi)(H_{i}) as in the definition of a hameomorphism.

The purpose of this paper is to study the algebraic structure of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) and its subgroups, for a general surface Σ\Sigma.

Here is what was known before this paper:

  1. (1)

    Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) is not simple since FHomeo(Σ)\operatorname{FHomeo}(\Sigma) is a proper normal subgroup ([CGHM+21]);

  2. (2)

    Hameo(S2)\operatorname{Hameo}(S^{2}) is not simple ([CGHM+22]);

  3. (3)

    FHomeo(S2)\operatorname{FHomeo}(S^{2}) is not simple since Hameo(S2)\operatorname{Hameo}(S^{2}) is a proper normal subgroup ([Buh22]);

  4. (4)

    when Σ\Sigma has non-empty boundary :

    1. (a)

      Hameo(Σ)\operatorname{Hameo}(\Sigma) is not simple since it contains the kernel of the (extended) Calabi homomorphism ([CGHM+21]);

    2. (b)

      FHomeo(Σ)\operatorname{FHomeo}(\Sigma) is not simple, since either Hameo(Σ)\operatorname{Hameo}(\Sigma) is a proper normal subgroup, or they coincide and by the previous point they are not simple ([CGHM+21]);

    3. (c)

      Hameo(D2)Ker(Cal)\operatorname{Hameo}(D^{2})\cap\operatorname{Ker}(\operatorname{Cal}) is not simple ([CGHM+22])

  5. (5)

    All normal subgroups of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) contain the commutator subgroup, which is perfect and simple.

We will extend this picture with a generalization of (2)(2), (3)(3) and (4)(c)(4)(c) respectively:

  • when Σ\Sigma is closed, Hameo(Σ)\operatorname{Hameo}(\Sigma) is not simple (Theorem 4);

  • when Σ\Sigma is closed, FHomeo(Σ)\operatorname{FHomeo}(\Sigma) is not simple, since either Hameo(Σ)\operatorname{Hameo}(\Sigma) is a proper normal subgroup, or they coincide and by the previous point they are not simple;

  • when Σ\Sigma has non-empty boundary, Hameo(Σ)Ker(Cal)\operatorname{Hameo}(\Sigma)\cap\operatorname{Ker}(\operatorname{Cal}) is not simple (Theorem 3).

2.2. Spectral invariants and Quasimorphisms

Let (M,ω)(M,\omega) be a closed symplectic manifold, and LML\subset M a monotone Lagrangian, i.e. ω|π2(M,L)=τμ|π2(M,L)\omega|_{\pi_{2}(M,L)}=\tau\mu|_{\pi_{2}(M,L)} for some constant τ>0\tau>0, where μ\mu is the Maslov homomorphism. Then, by [LZ18], for a Lagrangian LL^{\prime} Hamiltonian isotopic to LL, and a Hamiltonian HH such that φH1(L)L\varphi^{1}_{H}(L)\pitchfork L^{\prime}, the Floer cohomology HF(L,L,H)HF^{*}(L,L^{\prime},H) is well defined.

We follow the convention in [CGHM+21, Section 6] and define the Floer cohomology HF(L,L,H)HF^{*}(L,L^{\prime},H) is a vector space over [[T]][T1]\mathbb{C}[[T]][T^{-1}]. In particular, there is an action filtration on the Floer complex, by defining CFλ(L,L,H)CF_{\lambda}(L,L^{\prime},H), the subcomplex of CF(L,L,H)CF(L,L^{\prime},H) generated by capped Hamiltonian chords of action less than or equal to λ\lambda. The inclusion of this subcomplex gives rise to a map

iλ:HFλ(L,L,H)HF(L,L,H).i_{\lambda}:HF_{\lambda}(L,L^{\prime},H)\to HF(L,L^{\prime},H).

We assume that either L=LL^{\prime}=L or LLL^{\prime}\pitchfork L. In the former case, there is the PSS isomorphism QH(L)HF(L,L,H)QH(L)\to HF(L,L,H). In the latter case, there is the continuation isomorphism HF(L,L,0)HF(L,L,H)HF(L,L^{\prime},0)\to HF(L,L^{\prime},H). By an abuse of notation, we denote QH(L)QH(L) by HF(L,L,0)HF(L,L,0) and the isomorphism (in either case) by κ\kappa. Given a homology class aHF(L,L,0){0}a\in HF^{*}(L,L^{\prime},0)\setminus\{0\}, one can define a spectral invariant:

cL,L(a,H):=inf{λ|κ(a)Imiλ}c_{L,L^{\prime}}(a,H):=\inf\{\lambda|\kappa(a)\in\operatorname{Im}i_{\lambda}\}

When L=LL=L^{\prime} and a=eLa=e_{L} is the unit of QH(L)QH^{*}(L), we will simply denote cL(H):=cL,L(eL,H)c_{L}(H):=c_{L,L}(e_{L},H). This spectral invariant satisfies a homotopy invariance property, which enables us to define cLc_{L} on Ham~(M,ω)\widetilde{\operatorname{Ham}}(M,\omega), the universal cover of Ham(M,ω)\operatorname{Ham}(M,\omega).

We recall the definition of a quasimorphism:

Definition 9.

Let GG be a group. A quasimorphism on GG is a map μ:G\mu:G\to\mathbb{R} that satisfies:

D0,g,hG,|μ(gh)μ(g)μ(h)|D\exists D\geqslant 0,\forall g,h\in G,|\mu(gh)-\mu(g)-\mu(h)|\leqslant D

The infimal value of DD such that this property holds is called the defect of μ\mu.

Moreover, μ\mu is an homogeneous quasimorphism if it also satisfies

n,gG,μ(gn)=nμ(g)\forall n\in\mathbb{Z},\forall g\in G,\mu(g^{n})=n\mu(g)

When (M,ω)=(n,ωFS)(M,\omega)=(\mathbb{CP}^{n},\omega_{FS}) and LL is a monotone Lagrangian submanifold with HF(L)0HF(L)\neq 0, cLc_{L} is a quasimorphism on Ham~(M,ω)\widetilde{\operatorname{Ham}}(M,\omega). This is a consequence of the same result for the Hamiltonian spectral invariant cc (cf. [EP]), and the inequality cLcc_{L}\leqslant c (cf. [LZ18, Proposition 4]).

Proposition 10 (Homogenization).

Let μ:G\mu:G\to\mathbb{R} be a quasimorphism. Then,

μ~(g):=limnμ(gn)n\widetilde{\mu}(g):=\lim\limits_{n\to\infty}\frac{\mu(g^{n})}{n}

is well defined, and it is a homogeneous quasimorphism, called the homogenization of μ\mu.

Now we explain the construction of spectral invariants for Lagrangian links as defined in [CGHM+21].

Consider a closed symplectic surface (Σ,ω)(\Sigma,\omega), with a compatible complex structure jj. A Lagrangian link on Σ\Sigma is a disjoint union L¯=L1Lk\underline{L}=L_{1}\cup...\cup L_{k} of smooth simple curves in Σ\Sigma.

Definition 11.

Denote by BjB_{j}, 1js1\leqslant j\leqslant s, the connected components of ΣL¯\Sigma\setminus\underline{L}. Let kjk_{j} be the number of boundary components of BjB_{j}, and AjA_{j} the ω\omega-area of BjB_{j}. Let η0\eta\geqslant 0. We say that L¯\underline{L} is η\eta-monotone if

λ:=2η(kj1)+Aj\lambda:=2\eta(k_{j}-1)+A_{j}

does not depend on jj. λ\lambda is called the monotonicity constant of L¯\underline{L}.

A Lagrangian link L¯\underline{L} on a compact surface Σ0\Sigma_{0} with non-empty boundary is called η\eta-monotone if there exists a symplectic embedding of Σ0\Sigma_{0} into a closed surface Σ\Sigma such that L¯\underline{L} is η\eta-monotone inside Σ\Sigma.

Remark 12.

λ\lambda is equal to the area of the disks bounded by contractible components of the link. Therefore, if L¯\underline{L} has mm components bounding pairwise disjoint disks, then λ1m\lambda\leqslant\frac{1}{m}.

Let L¯=L1Lk\underline{L}=L_{1}\cup...\cup L_{k} be a Lagrangian link on Σ\Sigma. Denote by SymL¯\operatorname{Sym}\underline{L} the image of L1××LkL_{1}\times...\times L_{k} in the symmetric product Symk(Σ):=Σk/𝔖k\operatorname{Sym}^{k}(\Sigma):=\Sigma^{k}/\mathfrak{S}_{k}, where 𝔖k\mathfrak{S}_{k} is the permutation group permuting the factors. Suppose that L¯\underline{L} is η\eta-monotone and L¯\underline{L}^{\prime} is Hamiltonian isotopic to L¯\underline{L}. Let H:S1×ΣH:S^{1}\times\Sigma\to\mathbb{R} be a Hamiltonian and Symk(H):S1×Symk(Σ)\operatorname{Sym}^{k}(H):S^{1}\times\operatorname{Sym}^{k}(\Sigma)\to\mathbb{R} be given by Symk(H)t(x1,,xk):=i=1kHt(xi)\operatorname{Sym}^{k}(H)_{t}(x_{1},\dots,x_{k}):=\sum_{i=1}^{k}H_{t}(x_{i}). We recall in Section 5.1 how from such a link one can define a Floer cohomology111The function Symk(H)\operatorname{Sym}^{k}(H) is not smooth along the diagonal of Symk(Σ)\operatorname{Sym}^{k}(\Sigma) but it turns out that any smooth Hamiltonian that agrees with Symk(H)\operatorname{Sym}^{k}(H) outside a sufficiently small neighborhood of the diagonal will give the same Floer cohomology up to canonical isomorphisms as a filtered vector space. Therefore, HF(SymL¯,SymL¯,Symk(H))HF^{*}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}^{k}(H)) is defined to be the filtered vector space.

(1) HF(L¯,L¯,H):=HF(SymL¯,SymL¯,Symk(H)).\displaystyle HF(\underline{L},\underline{L}^{\prime},H):=HF^{*}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}^{k}(H)).

It was shown in [CGHM+21] that HF(SymL¯,SymL¯,Symk(H))H(SymL¯)HF^{*}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}^{k}(H))\cong H^{*}(\operatorname{Sym}\underline{L}) so a vector space (without filtration) so it is non-zero. Moreover, they show that Lagrangian spectral invariants cSymL¯,SymL¯(a,Symk(H))c_{\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime}}(a,\operatorname{Sym}^{k}(H)) are well-defined. Therefore, one can define link spectral invariants

cL¯:=1kcSymL¯=1kcSymL¯,SymL¯(eSymL¯,)c_{\underline{L}}:=\frac{1}{k}c_{\operatorname{Sym}\underline{L}}=\frac{1}{k}c_{\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}}(e_{\operatorname{Sym}\underline{L}},\cdot)
Proposition 13.

This invariant inherits all the properties of Lagrangian spectral invariants:

  • (spectrality) cL¯(H)c_{\underline{L}}(H) lies in the action spectrum Spec(H,L¯)\operatorname{Spec}(H,\underline{L})

  • (Hofer Lipschitz) |cL¯(H)cL¯(K)|HKHof\left|c_{\underline{L}}(H)-c_{\underline{L}}(K)\right|\leqslant||H-K||_{\operatorname{Hof}}

  • (monotonicity) If HKH\leqslant K then cL¯(H)cL¯(K)c_{\underline{L}}(H)\leqslant c_{\underline{L}}(K)

  • (Lagrangian control) If Ht|Li=si(t)H_{t}|_{L_{i}}=s_{i}(t) for each ii, then

    cL¯(H)=1ki=1ksi(t)𝑑tc_{\underline{L}}(H)=\frac{1}{k}\sum\limits_{i=1}^{k}\int s_{i}(t)dt

    Moreover,

    1ki=1kS1minLiHtdtcL¯(H)1ki=1kS1maxLiHtdt\frac{1}{k}\sum\limits_{i=1}^{k}\int_{S^{1}}\min\limits_{L_{i}}H_{t}dt\leqslant c_{\underline{L}}(H)\leqslant\frac{1}{k}\sum\limits_{i=1}^{k}\int_{S^{1}}\max\limits_{L_{i}}H_{t}dt
  • (triangle inequality) cL¯(H#K)cL¯(H)+cL¯(K)c_{\underline{L}}(H\#K)\leqslant c_{\underline{L}}(H)+c_{\underline{L}}(K)

  • (homotopy invariance) If H,KH,K are mean normalized, ϕH1=ϕK1\phi^{1}_{H}=\phi^{1}_{K} and (ϕHt)t[0,1](\phi^{t}_{H})_{t\in[0,1]} is homotopic to (ϕKt)t[0,1](\phi^{t}_{K})_{t\in[0,1]} relative to endpoints, then cL¯(H)=cL¯(K)c_{\underline{L}}(H)=c_{\underline{L}}(K).

  • (shift) cL¯(H+s(t))=cL¯(H)+s(t)𝑑tc_{\underline{L}}(H+s(t))=c_{\underline{L}}(H)+\int s(t)dt

The homotopy invariance permits to define cL¯({φt}t[0,1])c_{\underline{L}}(\{\varphi^{t}\}_{t\in[0,1]}) for {φt}t[0,1]Ham~(Σ)\{\varphi^{t}\}_{t\in[0,1]}\in\widetilde{\operatorname{Ham}}(\Sigma), by the formula cL¯({ϕHt}t[0,1])=cL¯(H)c_{\underline{L}}(\{\phi_{H}^{t}\}_{t\in[0,1]})=c_{\underline{L}}(H) for a mean normalized HH.

Moreover, cL¯c_{\underline{L}} a quasimorphism when Σ=S2\Sigma=S^{2} (i.e. Theorem 1). It is proved using the fact that Symk(S2)k\operatorname{Sym}^{k}(S^{2})\cong\mathbb{CP}^{k} (cf. [EP]).

3. Construction of the new invariants

Let (Σ,ω)(\Sigma,\omega) be a compact surface of genus gg. We suppose that Σ\Sigma has area 11. We introduce the following class of links, which is slightly different from the ones in [CGHM+21] (cf. [Che21], [Che22] for the study of this class of links in the cylindrical setting).

Definition 14.

A Lagrangian link L¯=L1Lkα1αg\underline{L}=L_{1}\cup...\cup L_{k}\cup\alpha_{1}\cup...\cup\alpha_{g} is called admissible if:

  • the circles L1,,Lk,α1,,αgL_{1},...,L_{k},\alpha_{1},...,\alpha_{g} are all disjoint;

  • α1,,αg\alpha_{1},...,\alpha_{g} are non-contractible;

  • there exists a decomposition of Σ\Sigma as a connected sum of a genus zero surface Σ0\Sigma_{0} and gg tori such that each αi\alpha_{i} lives in a different torus and LiL_{i} lives in Σ0\Sigma_{0};

  • L¯0:=L1LkΣ0\underline{L}_{0}:=L_{1}\cup...\cup L_{k}\subset\Sigma_{0} is η\eta-monotone for some η0\eta\geqslant 0, with respect to a symplectic form ω0\omega_{0} on Σ0\Sigma_{0} which coincides with ω\omega outside a small neighborhood of the connected sum region away from the link, and such that ω0(Σ0)=1\omega_{0}(\Sigma_{0})=1.

We define the monotonicity constant of L¯\underline{L} as the monotonicity constant of L¯0\underline{L}_{0} (see Definition 11).

Remark 15 (A remark on the third bullet of Definition 14).

Suppose that L¯=L1Lkα1αg\underline{L}=L_{1}\cup...\cup L_{k}\cup\alpha_{1}\cup...\cup\alpha_{g} satisfies the first two bullets of Definition 14. Let BB be the image of H1(Σ)H1(Σ)H_{1}(\partial\Sigma)\to H_{1}(\Sigma), VV be the image of H1(α1αg)H1(Σ)H_{1}(\alpha_{1}\cup\dots\cup\alpha_{g})\to H_{1}(\Sigma) and lil_{i} be the image of H1(Li)H1(Σ)H_{1}(L_{i})\to H_{1}(\Sigma). Topologically, if VV is a gg dimensional subspace which intersects BB only at 0 and liBl_{i}\subset B for all ii, then there is a decomposition of Σ\Sigma as a connected sum of a genus zero surface Σ0\Sigma_{0} and gg tori such that the third bullet of Definition 14 is satisfied.

To see this, for simplicity, we first assume that there is no LiL_{i} and Σ\Sigma is closed (so B=0B=0). Then VV is a Lagrangian subspace with respect to the intersection form Ω\Omega on H1(Σ)H_{1}(\Sigma). Let ai:=[αi]H1(Σ)a_{i}:=[\alpha_{i}]\in H_{1}(\Sigma). We can complete {ai}\{a_{i}\} to a basis {a1,,ag,b1,,bg}\{a_{1},\dots,a_{g},b_{1},\dots,b_{g}\} of H1(Σ;)H_{1}(\Sigma;\mathbb{Z}) such that Ω(ai,bi)=1\Omega(a_{i},b_{i})=1 and Ω(ai,bj)=0\Omega(a_{i},b_{j})=0 if iji\neq j, and Ω(bi,bj)=0\Omega(b_{i},b_{j})=0 for all i,ji,j. We can find circles βiΣ\beta_{i}\subset\Sigma, i=1,,gi=1,\dots,g, such that the geometric intersection number between any two circles in {αi,βj}\{\alpha_{i},\beta_{j}\} agrees with the homological intersection number. The regular neighborhood of αiβi\alpha_{i}\cup\beta_{i} gives the splitting of the ithi^{th} torus in the connected sum decomposition. The case when Σ\Sigma has boundary components can be proved by first embedding it to a closed surface by capping off the boundary components by disks (and choosing βi\beta_{i} to avoid the capping disks). The case when there is LiL_{i} can be reduced to the case with no LiL_{i} by running the argument above, for the positive genus components, in the complement of iLi\cup_{i}L_{i} (in particular, LiL_{i} are allowed to be non-contractible separating circles).

Refer to caption
Figure 1. An admissible link

We assume that Σ\Sigma is closed. Then, given an admissible link L¯\underline{L}, there exists a decomposition of Σ\Sigma as a connected sum Σ=S2#E1##Eg\Sigma=S^{2}\#E_{1}\#...\#E_{g}, where the EiE_{i} are copies of the 2-torus, such that L¯0:=L1LkS2\underline{L}_{0}:=L_{1}\cup...\cup L_{k}\subset S^{2} is η\eta-monotone, and for all 1ig1\leqslant i\leqslant g, αiEg\alpha_{i}\subset E_{g}. (Here, we inflate the symplectic form near the connected sum point in S2S^{2} so that S2S^{2} has area 11. This choice of symplectic form makes L¯0\underline{L}_{0} η\eta-monotone by the fourth bullet of Definition 14, and it is compatible with the one in Theorem 6.)

The authors of [CGHM+21] show that HF(L¯0,L¯0)HF^{*}(\underline{L}_{0},\underline{L}_{0}) is well-defined and non-zero.

By applying Theorem 6 gg times, we get that HF(L¯,L¯)HF^{*}(\underline{L},\underline{L}) is non-zero, and therefore for a non-degenerate Hamiltonian HH, HF(L¯,L¯,H)HF^{*}(\underline{L},\underline{L},H) is also non-zero. As a result, one can define spectral invariants

cL¯(H):=1k+gcSymL¯(Symk+g(H))c_{\underline{L}}(H):=\frac{1}{k+g}c_{\operatorname{Sym}\underline{L}}(\operatorname{Sym}^{k+g}(H))

for non-degenerate HH, and then extend it to all Hamiltonians by continuity (i.e. the Hofer Lipschitz property in Proposition 13).

If Σ0\Sigma_{0} has non-empty boundary, then one can embed Σ0\Sigma_{0} into a closed surface Σ\Sigma, such that L¯\underline{L} remains admissible in Σ\Sigma. Indeed, by the definition of η\eta-monotonicity for surfaces with boundary, there exists an embedding into a closed surface Σ\Sigma such that L¯\underline{L} is still η\eta-monotone inside Σ\Sigma. Then, one can define the link spectral invariant for Σ0\Sigma_{0} by restricting cL¯c_{\underline{L}} to Ham(Σ0)Ham(Σ)\operatorname{Ham}(\Sigma_{0})\subset\operatorname{Ham}(\Sigma).

The fact that this invariant satisfy all the properties listed in Proposition 13, as in [CGHM+21], is a straightforward consequence of the properties of Lagrangian spectral invariants (see [Hum17] for instance).

Before proving the local quasimorphism property 5, we show the following statement:

Proposition 16.

Let Σ\Sigma be a surface of genus g>0g>0. Let L¯=L1Lkα1αg\underline{L}=L_{1}\cup...\cup L_{k}\cup\alpha_{1}\cup...\cup\alpha_{g} be a monotone admissible Lagrangian link on Σ\Sigma, where α1\alpha_{1}, …, αg\alpha_{g} are the non-contractible components of L¯\underline{L}. Then cL¯c_{\underline{L}} is not a quasimorphism.

Proof.

It is enough to show that there exists a sequence of Hamiltonians (Hn)n(H_{n})_{n} such that γL¯(Hn):=cL¯(Hn)+cL¯(H¯n)\gamma_{\underline{L}}(H_{n}):=c_{\underline{L}}(H_{n})+c_{\underline{L}}(\overline{H}_{n}) is not bounded. We pick a non-contractible circle in Σ\Sigma that intersects L¯\underline{L} at a single point in α1\alpha_{1}. Such a circle always exists, take for instance β1\beta_{1} as in Remark 15. Then, we pick a small neighborhood UU of this circle, diffeomorphic to the annulus A=S1×(1;1)A=S^{1}\times(-1;1) (we denote by ψ:UA\psi:U\rightarrow A such a diffeomorphism), such that UL¯=Uα1U\cap\underline{L}=U\cap\alpha_{1} is connected and sent to a vertical {θ0}×(1;1)\{\theta_{0}\}\times(-1;1) by ψ\psi. Let H:(1;1)H:(-1;1)\rightarrow\mathbb{R} be a smooth function such that :

  • HH is compactly supported;

  • HH admits a single local maximum at 0 and no other critical point in the interior of its support;

  • H(0)=1H(0)=1

We define KnK_{n} on AA by Kn(θ,t)=nH(t)K_{n}(\theta,t)=nH(t), and HnH_{n} on Σ\Sigma by:

  • Hn(x)=Kn(ψ(x))H_{n}(x)=K_{n}(\psi(x)) if xUx\in U

  • Hn(x)=0H_{n}(x)=0 if xUx\notin U

Now, we compute the sequence (γL¯(Hn))n(\gamma_{\underline{L}}(H_{n}))_{n} for this choice of Hamiltonians.

We know that cL¯(Hn)c_{\underline{L}}(H_{n}) lies in 1k+gSpec(Sym(Hn))\frac{1}{k+g}\operatorname{Spec}(\operatorname{Sym}(H_{n})). In order to compute this spectrum, we consider critical points of the action that are in the same connected component as a chosen reference path in 𝒫(Sym(L¯),Sym(L¯))\mathcal{P}(\operatorname{Sym}(\underline{L}),\operatorname{Sym}(\underline{L})) (see Section 5.1 for a definition of the Heegaard Floer complex and the action functional). We pick x1L1,,xkLk,y1α1,,ygαgx_{1}\in L_{1},...,x_{k}\in L_{k},y_{1}\in\alpha_{1},...,y_{g}\in\alpha_{g} fixed by the flow of HnH_{n}, and take the constant path η:={x1,,xk,y1,,yg}\eta:=\{x_{1},...,x_{k},y_{1},...,y_{g}\} in Sym(L¯)\operatorname{Sym}(\underline{L}) as the reference path.

Then, the only critical points of the action that are in the same connected component as η\eta in 𝒫(SymL¯,SymL¯)\mathcal{P}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}) are symmetric products of points fixed by HnH_{n}. They all have zero action except when we choose in α1\alpha_{1} the point y1y^{\prime}_{1} for which HnH_{n} is maximal. For any choice of xiLi,1ikx^{\prime}_{i}\in L_{i},1\leqslant i\leqslant k, and yiαi,2igy^{\prime}_{i}\in\alpha_{i},2\leqslant i\leqslant g, the critical point {x1,,xk,y1,,yg}\{x^{\prime}_{1},...,x^{\prime}_{k},y^{\prime}_{1},...,y^{\prime}_{g}\} has action nn.

Hence, Spec(Sym(Hn))={0,n}\operatorname{Spec}(\operatorname{Sym}(H_{n}))=\{0,n\}. Similarly, Spec(Sym(H¯n))={n,0}\operatorname{Spec}(\operatorname{Sym}(\overline{H}_{n}))=\{-n,0\} because H¯n=Hn\overline{H}_{n}=-H_{n}. Therefore, γL¯(Hn){nk+g,0,nk+g}\gamma_{\underline{L}}(H_{n})\in\{-\frac{n}{k+g},0,\frac{n}{k+g}\}.

Since γL¯\gamma_{\underline{L}} is non-negative, we can rule out nk+g-\frac{n}{k+g}. Moreover, Sym(L¯)\operatorname{Sym}(\underline{L}) is not fixed by φHn\varphi_{H_{n}}, so γL¯(Hn)\gamma_{\underline{L}}(H_{n}) is non-zero.

Finally, we get that γL¯(Hn)=nk+g\gamma_{\underline{L}}(H_{n})=\frac{n}{k+g}, which is unbounded as nn goes to infinity. ∎

We now prove that this invariant satisfies Theorem 5.

Proof of Theorem 5.

We consider an admissible link L¯=L1Lkα1αg\underline{L}=L_{1}\cup...\cup L_{k}\cup\alpha_{1}\cup...\cup\alpha_{g}, and a disk DD that does not intersect α¯:=α1αg\underline{\alpha}:=\alpha_{1}\cup...\cup\alpha_{g}. Then, one can find a decomposition of Σ\Sigma as a connected sum Σ=S2#E1##Eg\Sigma=S^{2}\#E_{1}\#...\#E_{g} such that L¯0:=L1LkS2\underline{L}_{0}:=L_{1}\cup...\cup L_{k}\subset S^{2} is η\eta-monotone, for all 1ig1\leqslant i\leqslant g, αiEg\alpha_{i}\subset E_{g}, and DS2D\subset S^{2}.

Let HH be a Hamiltonian supported in DD, and let HϵH_{\epsilon} be an ϵ\epsilon-perturbation of HH in small neighborhoods of the link’s components so that HF(L¯,L¯,Hϵ)HF(\underline{L},\underline{L},H_{\epsilon}) is well defined (cf. (1)). We can assume that HϵH_{\epsilon} is chosen such that it is away from the connected sum neighborhoods of the decomposition Σ=S2#E1##Eg\Sigma=S^{2}\#E_{1}\#...\#E_{g}.

Then, by applying Theorem 6 gg times, we have that

CF(L¯,L¯,Hϵ)CF(L¯0,L¯0,Hϵ|S2)i=1gCF(αi,αi,Hϵ|Ei)\displaystyle CF^{*}(\underline{L},\underline{L},H_{\epsilon})\simeq CF^{*}(\underline{L}_{0},\underline{L}_{0},H_{\epsilon}|_{S^{2}})\otimes\bigotimes_{i=1}^{g}CF^{*}(\alpha_{i},\alpha_{i},H_{\epsilon}|_{E_{i}})

Therefore, representatives of κ(eSym(L¯))\kappa(e_{\operatorname{Sym}(\underline{L})}) in CF(L¯,L¯,Hϵ)CF^{*}(\underline{L},\underline{L},H_{\epsilon}) are in one-to-one correspondence with tensor products of representatives of unit classes in CF(L¯0,L¯0,Hϵ|S2)CF^{*}(\underline{L}_{0},\underline{L}_{0},H_{\epsilon}|_{S^{2}}) and CF(αi,αi,Hϵ|Ei)CF^{*}(\alpha_{i},\alpha_{i},H_{\epsilon}|_{E_{i}}). It follows from the proof of Theorem 6 that this one-to-one correspondence preserves the action (which is defined as the sum of the actions on the tensor product).

Since HϵH_{\epsilon} is ϵ\epsilon-small on EiE_{i}, we get that cSymL¯(Symk+g(H))=cSymL¯0(Symk(H))c_{\operatorname{Sym}\underline{L}}(\operatorname{Sym}^{k+g}(H))=c_{\operatorname{Sym}\underline{L}_{0}}(\operatorname{Sym}^{k}(H)) where cSymL¯0c_{\operatorname{Sym}\underline{L}_{0}} is computed inside Symk(S2)\operatorname{Sym}^{k}(S^{2}). Since L0L_{0} is η\eta-monotone inside S2S^{2}, with monotonicity constant λ\lambda, applying Theorem 1 gives that the restriction of cSymL¯c_{\operatorname{Sym}\underline{L}} to HamD(Σ)\operatorname{Ham}_{D}(\Sigma) is a quasimorphism with defect bounded by k+1k+gλ\frac{k+1}{k+g}\lambda. ∎

We define the homogenized spectral invariant μL¯\mu_{\underline{L}} by the formula:

μL¯(H):=limncL¯(H#n)n\mu_{\underline{L}}(H):=\lim\limits_{n\to\infty}\frac{c_{\underline{L}}(H^{\#n})}{n}

This is well defined by the triangle inequality and Fekete’s lemma.

Proposition 17.

The invariant μL¯\mu_{\underline{L}} satisfies the following properties:

  • (Hofer Lipschitz) |μL¯(H)μL¯(K)|HKHof\left|\mu_{\underline{L}}(H)-\mu_{\underline{L}}(K)\right|\leqslant||H-K||_{\operatorname{Hof}}

  • (Lagrangian control) Suppose HH is mean-normalized, and Ht|Li=si(t)H_{t}|_{L_{i}}=s_{i}(t). Then,

    μL¯(H)=1ki=1k01si(t)𝑑t\mu_{\underline{L}}(H)=\frac{1}{k}\sum\limits_{i=1}^{k}\int_{0}^{1}s_{i}(t)dt

    Moreover,

    1ki=1k01minLiHtdtμL¯(H)1ki=1k01maxLiHtdt\frac{1}{k}\sum\limits_{i=1}^{k}\int_{0}^{1}\min\limits_{L_{i}}H_{t}dt\leqslant\mu_{\underline{L}}(H)\leqslant\frac{1}{k}\sum\limits_{i=1}^{k}\int^{1}_{0}\max\limits_{L_{i}}H_{t}dt
  • (homotopy invariance) μL¯\mu_{\underline{L}} descends to a map Ham(Σ)\operatorname{Ham}(\Sigma)\to\mathbb{R}

  • (support control) If supp(φ)ΣL¯\operatorname{supp}(\varphi)\subset\Sigma\setminus\underline{L}, then μL¯(φ)=Cal(φ)\mu_{\underline{L}}(\varphi)=-\operatorname{Cal}(\varphi).

  • (conjugacy invariance) μL¯(ψφψ1)=μL¯(φ)\mu_{\underline{L}}(\psi\varphi\psi^{-1})=\mu_{\underline{L}}(\varphi)

Proof.

These are all straightforward consequences of the properties of cL¯c_{\underline{L}} (13) and the definition of μL¯\mu_{\underline{L}}. ∎

Moreover, we show the following:

Theorem 18.

Suppose that L¯\underline{L} and L¯\underline{L}^{\prime} are two admissible η\eta-monotone links with the same number of components kk, that share the same non-contractible components α¯\underline{\alpha}. Then the homogenized spectral invariants μL¯\mu_{\underline{L}} and μL¯\mu_{\underline{L}^{\prime}} coincide, and we denote by μk,η,α¯\mu_{k,\eta,\underline{\alpha}} their common value.

Proof.

Let * denote the pants product

HF(L¯,L¯)HF(L¯,L¯)HF(L¯,L¯).HF^{*}(\underline{L},\underline{L}^{\prime})\otimes HF^{*}(\underline{L}^{\prime},\underline{L})\to HF^{*}(\underline{L},\underline{L}).

Using Theorem 6, we can view it as a map

HF(L¯0,L¯0)HF(α¯,α¯)HF(L¯0,L¯0)HF(α¯,α¯)HF(L¯0,L¯0)HF(α¯,α¯)HF^{*}(\underline{L}_{0},\underline{L}_{0}^{\prime})\otimes HF^{*}(\underline{\alpha},\underline{\alpha})\otimes HF^{*}(\underline{L}^{\prime}_{0},\underline{L}_{0})\otimes HF^{*}(\underline{\alpha},\underline{\alpha})\to HF^{*}(\underline{L}_{0},\underline{L}_{0})\otimes HF^{*}(\underline{\alpha},\underline{\alpha})

Since L¯\underline{L} and L¯0\underline{L}_{0} are two η\eta-monotone links with the same number of components in S2S^{2}, there exist classes a0HF(L¯0,L¯0)a_{0}\in HF^{*}(\underline{L}_{0},\underline{L}_{0}^{\prime}) and b0HF(L¯0,L¯0)b_{0}\in HF^{*}(\underline{L}^{\prime}_{0},\underline{L}_{0}) such that a0b0=eSymL¯0HF(L¯0,L¯0)a_{0}*b_{0}=e_{\operatorname{Sym}\underline{L}_{0}}\in HF^{*}(\underline{L}_{0},\underline{L}_{0}).

Let aa be the image of a0eα¯a_{0}\otimes e_{\underline{\alpha}} in HF(L¯,L¯)HF(L¯0,L¯0)HF(α¯,α¯)HF^{*}(\underline{L},\underline{L}^{\prime})\cong HF^{*}(\underline{L}_{0},\underline{L}_{0}^{\prime})\otimes HF^{*}(\underline{\alpha},\underline{\alpha}), and bb the image of b0eα¯b_{0}\otimes e_{\underline{\alpha}} in HF(L¯,L¯)HF(L¯0,L¯0)HF(α¯,α¯)HF^{*}(\underline{L}^{\prime},\underline{L})\cong HF^{*}(\underline{L}_{0}^{\prime},\underline{L}_{0})\otimes HF^{*}(\underline{\alpha},\underline{\alpha}).

Then, aba*b is the image of (a0b0)(eα¯eα¯)=eSymL¯0eα¯(a_{0}*b_{0})\otimes(e_{\underline{\alpha}}*e_{\underline{\alpha}})=e_{\operatorname{Sym}\underline{L}_{0}}\otimes e_{\underline{\alpha}}, i.e. ab=eSymL¯a*b=e_{\operatorname{Sym}\underline{L}} is the unit of HF(L¯,L¯)HF^{*}(\underline{L},\underline{L}).

Then, by the subadditivity property of Lagrangian spectral invariants, we have for any Hamiltonian HH:

c(SymL¯,SymL¯,eSymL¯,H)\displaystyle c(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L},e_{\operatorname{Sym}\underline{L}},H)
\displaystyle\leqslant c(SymL¯,SymL¯,a,H)+c(SymL¯,SymL¯,b,0)\displaystyle c(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime},a,H)+c(\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}\underline{L},b,0)
\displaystyle\leqslant c(SymL¯,SymL¯,a,0)+c(SymL¯,SymL¯,eSymL¯,H)+c(SymL¯,SymL¯,b,0)\displaystyle c(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime},a,0)+c(\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}\underline{L}^{\prime},e_{\operatorname{Sym}\underline{L}^{\prime}},H)+c(\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}\underline{L},b,0)

i.e.

cL¯(H)cL¯(H)+1k(c(SymL¯,SymL¯,a,0)+c(SymL¯,SymL¯,b,0))c_{\underline{L}}(H)\leqslant c_{\underline{L}^{\prime}}(H)+\frac{1}{k}\left(c(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime},a,0)+c(\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}\underline{L},b,0)\right)

We get for all n>0n>0:

cL¯(H#n)ncL¯(H#n)n+c(SymL¯,SymL¯,a,0)+c(SymL¯,SymL¯,b,0)kn\frac{c_{\underline{L}}(H^{\#n})}{n}\leqslant\frac{c_{\underline{L}^{\prime}}(H^{\#n})}{n}+\frac{c(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{L}^{\prime},a,0)+c(\operatorname{Sym}\underline{L}^{\prime},\operatorname{Sym}\underline{L},b,0)}{kn}

and therefore μL¯(H)μL¯(H)\mu_{\underline{L}}(H)\leqslant\mu_{\underline{L}^{\prime}}(H). Swapping the roles of L¯\underline{L} and L¯\underline{L}^{\prime}, we get the other inequality and finally μL¯=μL¯\mu_{\underline{L}}=\mu_{\underline{L}^{\prime}}.

Since the homogenized spectral invariants are conjugacy invariant, μk,η,α¯=μk,η,α¯\mu_{k,\eta,\underline{\alpha}}=\mu_{k,\eta,\underline{\alpha}^{\prime}} when α¯\underline{\alpha} and α¯\underline{\alpha}^{\prime} are Hamiltonian isotopic, and therefore we can write μk,η,[α¯]:=μk,η,α¯\mu_{k,\eta,[\underline{\alpha}]}:=\mu_{k,\eta,\underline{\alpha}} where [α¯][\underline{\alpha}] is the class of α¯\underline{\alpha}.

We fix a decomposition of Σ\Sigma as a connected sum Σ=Σ0#E1##Eg\Sigma=\Sigma_{0}\#E_{1}\#...\#E_{g} where Σ0\Sigma_{0} is a genus zero surface, and the EiE_{i} are copies of the 2-torus. Recall that we modify the symplectic form in a neighborhood of the connected sum points so that Σ0\Sigma_{0} has area 1.

Let βj,θ1\beta^{1}_{j,\theta} be the circle {θ}×S1S1×S1Ej\{\theta\}\times S^{1}\subset S^{1}\times S^{1}\cong E_{j}, and βj,θ2\beta^{2}_{j,\theta} be the circle S1×{θ}EjS^{1}\times\{\theta\}\subset E_{j}.

For θ¯=(θ1,,θg)(S1)g\underline{\theta}=(\theta_{1},...,\theta_{g})\in(S^{1})^{g}, and ϵ¯=(ϵ1,,ϵg){1,2}g\underline{\epsilon}=(\epsilon_{1},...,\epsilon_{g})\in\{1,2\}^{g}, let α¯θ¯ϵ¯:=β1,θ1ϵ1βg,θgϵg\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}}:=\beta^{\epsilon_{1}}_{1,\theta_{1}}\cup...\cup\beta^{\epsilon_{g}}_{g,\theta_{g}}. When the components of α¯θ¯ϵ¯\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}} do not intersect the connected sum regions, this defines a Lagrangian link on Σ\Sigma.

Refer to caption
Figure 2. Two links α¯\underline{\alpha} and β¯\underline{\beta} inducing independent invariants μα¯\mu_{\underline{\alpha}} and μβ¯\mu_{\underline{\beta}}
Proposition 19.

The {μk,η,[α¯θ¯ϵ¯]}\{\mu_{k,\eta,[\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}}]}\} are linearly independent.

Proof.

Let EE be the vector space generated by the {μk,η,[α¯θ¯ϵ¯]}\{\mu_{k,\eta,[\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}}]}\}, and Ek,ηE_{k,\eta} the vector subspace generated by the {μk,η,[α¯θ¯ϵ¯]}\{\mu_{k,\eta,[\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}}]}\} where kk and η\eta are fixed. Following the argument in [CGHM+21], one can show that E=k,ηEk,ηE=\bigoplus\limits_{k,\eta}E_{k,\eta}.

Now we show that for fixed kk and η\eta, the {μk,η,[α¯θ¯ϵ¯]}\{\mu_{k,\eta,[\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}}]}\} are linearly independent. For 1jg1\leqslant j\leqslant g, let Ek,η,βj,θϵE_{k,\eta,\beta^{\epsilon}_{j,\theta}} be the subspace generated by the μk,η,[α¯θ¯ϵ¯]\mu_{k,\eta,[\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}}]} that satisfy ϵj=ϵ\epsilon_{j}=\epsilon and θj=θ\theta_{j}=\theta. We are going to show that for every j=1,,gj=1,\dots,g, we have

(2) Ek,η=ϵ,θEk,η,βj,θϵ.\displaystyle E_{k,\eta}=\bigoplus_{\epsilon,\theta}E_{k,\eta,\beta^{\epsilon}_{j,\theta}}.

Let ll and mm be non-negative integers. Pick ll different elements θ1,,θl\theta_{1},...,\theta_{l} in S1S^{1}, and let μi\mu_{i} be an element of Ek,η,βj,θi1E_{k,\eta,\beta^{1}_{j,\theta_{i}}}. We also pick mm different elements θl+1,,θl+m\theta_{l+1},...,\theta_{l+m} in S1S^{1}, and let μl+i\mu_{l+i} be an element of Ek,η,βj,θl+i2E_{k,\eta,\beta^{2}_{j,\theta_{l+i}}}. Let aia_{i} be real numbers such that

i=1l+maiμi=0.\sum\limits_{i=1}^{l+m}a_{i}\mu_{i}=0.

We want to show that for all ii, ai=0a_{i}=0.

Let VV be a small neighborhood of βj,θ11\beta^{1}_{j,\theta_{1}} that does not intersect the connected sum points and the βj,θi1\beta^{1}_{j,\theta_{i}} for 2il2\leqslant i\leqslant l. Let HH be a Hamiltonian supported in VV such that H|βj,θ111H|_{\beta^{1}_{j,\theta_{1}}}\equiv 1.

Let θ0S1{θ1,,θl}\theta_{0}\in S^{1}-\{\theta_{1},...,\theta_{l}\} be such that βj,θ01\beta^{1}_{j,\theta_{0}} is away from the connected sum points. For 0il0\leqslant i\leqslant l, let ρi\rho_{i} be the rotation of the torus defined by

ρi(θ,φ)=(θ+θiθ1,φ)\rho_{i}(\theta,\varphi)=(\theta+\theta_{i}-\theta_{1},\varphi)

We can assume that VV is small enough such that for any i=0,,li=0,\dots,l, ρi(V)\rho_{i}(V) is a neighborhood of βj,θi1\beta^{1}_{j,\theta_{i}} that does not intersect the connected sum points and the βj,θs1\beta^{1}_{j,\theta_{s}} for sis\neq i. Let Hi:=Hρi1H_{i}:=H\circ\rho_{i}^{-1}, which is supported in ρi(V)\rho_{i}(V).

Then, by the Lagrangian control property, for 0il0\leqslant i\leqslant l and 1pl1\leqslant p\leqslant l, μp(Hi)=1k+gδi,p\mu_{p}(H_{i})=\frac{1}{k+g}\delta_{i,p}. Moreover, for l+1pl+ml+1\leqslant p\leqslant l+m, since the ρi\rho_{i} stabilize the βj,θp2\beta^{2}_{j,\theta_{p}}, we get that μp(Hi)\mu_{p}(H_{i}) does not depend on ii.

Therefore, applying the equality i=1l+maiμi=0\sum\limits_{i=1}^{l+m}a_{i}\mu_{i}=0 at HiH_{i} for 1il1\leqslant i\leqslant l gives

aik+g=p=l+1l+mapμp(H1)\frac{a_{i}}{k+g}=-\sum\limits_{p=l+1}^{l+m}a_{p}\mu_{p}(H_{1})

and at H0H_{0} :

p=l+1l+mapμp(H1)=0\sum\limits_{p=l+1}^{l+m}a_{p}\mu_{p}(H_{1})=0

Thus, ai=0a_{i}=0 for 1il1\leqslant i\leqslant l. Since the β1\beta^{1} and β2\beta^{2} play symmetric roles, we can show in the same way that ai=0a_{i}=0 for l+1il+ml+1\leqslant i\leqslant l+m.

Therefore, we have (2) for all jj, and hence {μk,η,[α¯θ¯ϵ¯]}\{\mu_{k,\eta,[\underline{\alpha}^{\underline{\epsilon}}_{\underline{\theta}}]}\} are linearly independent.

4. Extending the Calabi homomorphism and simplicity

The proofs of Theorem 2, 3 and 4 will be given in the following three subsections respectively. The main idea is to replace the quasimorphism property (Theorem 1), which no longer exists for positive genus surfaces, by Theorem 5 and the fragmentation technique. Some of the estimates is a bit more delicate than the ones in [CGHM+22].

4.1. Proof of Theorem 2

We can now give a proof of Theorem 2, inspired by the proof of (1) found in [CGHM+22].

Definition 20.

A sequence of admissible Lagrangian links (L¯k)(\underline{L}^{k}) is called equidistributed if:

  • all the L¯k\underline{L}^{k} share the same contractible components α1\alpha_{1},…, αg\alpha_{g};

  • L¯k\underline{L}^{k} has kk contractible components L¯1k\underline{L}^{k}_{1},…, L¯kk\underline{L}^{k}_{k};

  • the LikL^{k}_{i} bound disjoint disks DikD^{k}_{i}, and diam(L¯k):=max(diamDik)k0\operatorname{diam}(\underline{L}^{k}):=\max(\operatorname{diam}D^{k}_{i})\xrightarrow[k\to\infty]{}0.

Given such a sequence, we get a sequence of link spectral invariants cL¯kc_{\underline{L}^{k}} which satisfies the Calabi property:

Proposition 21.

For any smooth Hamiltonian H:S1×ΣH:S^{1}\times\Sigma\rightarrow\mathbb{R},

cL¯k(H)=S1ΣHtω𝑑t+Ok(diam(L¯k))c_{\underline{L}^{k}}(H)=\int_{S^{1}}\int_{\Sigma}H_{t}\omega dt+\operatorname{O}\limits_{k\to\infty}\left(\operatorname{diam}(\underline{L}^{k})\right)

In particular, for any smooth Hamiltonian diffeomorphism φ\varphi,

cL¯k(φ)=Ok(diam(L¯k))c_{\underline{L}^{k}}(\varphi)=\operatorname{O}_{k\to\infty}\left(\operatorname{diam}(\underline{L}^{k})\right)
Proof.

We fix a point xikx^{k}_{i} in each of the disks DikD^{k}_{i}. Then, one can find smooth Hamiltonians GkG^{k} such that:

  • GtkHt(xik)G^{k}_{t}\equiv H_{t}(x^{k}_{i}) on DikD^{k}_{i}

  • Gtk=HtG^{k}_{t}=H_{t} on αi\alpha_{i}

  • GtkHtdiam(L¯k)supS1×ΣdHt||G^{k}_{t}-H_{t}||_{\infty}\leqslant\operatorname{diam}(\underline{L}^{k})\sup\limits_{S^{1}\times\Sigma}||dH_{t}||

Then, using the Hofer Lipschitz property and Lagrangian control (Proposition 13), we get:

|cL¯k(H)S1ΣHtω𝑑t|\displaystyle\left|c_{\underline{L}^{k}}(H)-\int_{S^{1}}\int_{\Sigma}H_{t}\omega dt\right|
\displaystyle\leqslant |cL¯k(H)cL¯k(Gk)|+|cL¯k(Gk)S1ΣGtkω𝑑t|+|S1Σ(GtkHt)ω𝑑t|\displaystyle\left|c_{\underline{L}^{k}}(H)-c_{\underline{L}^{k}}(G^{k})\right|+\left|c_{\underline{L}^{k}}(G^{k})-\int_{S^{1}}\int_{\Sigma}G^{k}_{t}\omega dt\right|+\left|\int_{S^{1}}\int_{\Sigma}(G^{k}_{t}-H_{t})\omega dt\right|
\displaystyle\leqslant HGkHof+|1k+gi=1kS1Gtk(xik)𝑑tS1ΣGtkω𝑑t|\displaystyle||H-G^{k}||_{\operatorname{Hof}}+\left|\frac{1}{k+g}\sum\limits_{i=1}^{k}\int_{S^{1}}G^{k}_{t}(x^{k}_{i})dt-\int_{S^{1}}\int_{\Sigma}G^{k}_{t}\omega dt\right|
+1k+gi=1gS1maxαi|Gtk|dt+GkHHof\displaystyle+\frac{1}{k+g}\sum\limits_{i=1}^{g}\int_{S^{1}}\max\limits_{\alpha_{i}}|G^{k}_{t}|dt+||G^{k}-H||_{\operatorname{Hof}}
\displaystyle\leqslant 2diam(L¯k)supS1×ΣdHt+|1k+gi=1kS1Gtk(xik)𝑑tS1ΣGtkω𝑑t|+O(1k)\displaystyle 2\operatorname{diam}(\underline{L}^{k})\sup\limits_{S^{1}\times\Sigma}||dH_{t}||+\left|\frac{1}{k+g}\sum\limits_{i=1}^{k}\int_{S^{1}}G^{k}_{t}(x^{k}_{i})dt-\int_{S^{1}}\int_{\Sigma}G^{k}_{t}\omega dt\right|+\operatorname{O}\left(\frac{1}{k}\right)

Let A:=Area(Dik)A:=\operatorname{Area}(D^{k}_{i}). We have 1k+1A1k\frac{1}{k+1}\leqslant A\leqslant\frac{1}{k}, so diam(L¯k)Ck\operatorname{diam}(\underline{L}^{k})\geqslant\frac{C}{\sqrt{k}} for some positive constant CC and therefore 1k=Ok(diam(L¯k))\frac{1}{k}=\operatorname{O}_{k\to\infty}\left(\operatorname{diam}(\underline{L}^{k})\right). Moreover,

|1k+gi=1kS1Gtk(xik)𝑑tS1ΣGtkω𝑑t|\displaystyle\left|\frac{1}{k+g}\sum\limits_{i=1}^{k}\int_{S^{1}}G^{k}_{t}(x^{k}_{i})dt-\int_{S^{1}}\int_{\Sigma}G^{k}_{t}\omega dt\right|
=\displaystyle= |1(k+g)Ai=1kS1DikGtkω𝑑tS1ΣGtkω𝑑t|\displaystyle\left|\frac{1}{(k+g)A}\sum\limits_{i=1}^{k}\int_{S^{1}}\int_{D^{k}_{i}}G^{k}_{t}\omega dt-\int_{S^{1}}\int_{\Sigma}G^{k}_{t}\omega dt\right|
=\displaystyle= |(1(k+g)A1)S1DikGtkω𝑑tS1Σ(Dik)Gtkω𝑑t|\displaystyle\left|\left(\frac{1}{(k+g)A}-1\right)\int_{S^{1}}\int_{\bigcup D^{k}_{i}}G^{k}_{t}\omega dt-\int_{S^{1}}\int_{\Sigma\setminus(\bigcup D^{k}_{i})}G^{k}_{t}\omega dt\right|
\displaystyle\leqslant |1(k+g)A1|Gk+GkΣ(Dik)ω\displaystyle\left|\frac{1}{(k+g)A}-1\right|||G^{k}||_{\infty}+||G^{k}||_{\infty}\int_{\Sigma\setminus(\bigcup D^{k}_{i})}\omega
\displaystyle\leqslant (|1(k+g)A1|+1kA)(H+HGk)\displaystyle\left(\left|\frac{1}{(k+g)A}-1\right|+1-kA\right)\left(||H||_{\infty}+||H-G^{k}||_{\infty}\right)
=\displaystyle= O(1k)\displaystyle\operatorname{O}\left(\frac{1}{k}\right)

Since the invariants cL¯kc_{\underline{L}^{k}} satisfy all the properties listed in Proposition 13, the same proof as in [CGHM+21] show that:

Proposition 22.

fL¯k:=cL¯k+Cal:Ham(Σ)f_{\underline{L}^{k}}:=c_{\underline{L}^{k}}+\operatorname{Cal}:\operatorname{Ham}(\Sigma)\rightarrow\mathbb{R} is uniformly continuous, and therefore extends continuously to Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma).

Now, we define the following relation on the space of real valued sequences \mathbb{R}^{\mathbb{N}} : we say that xyx\sim y if limxy=0\lim x-y=0. This is an equivalence relation, and the quotient /\mathbb{R}^{\mathbb{N}}/\sim is a real vector space. Then, we can define a map

f:\displaystyle f: Ham¯(Σ)/\displaystyle\overline{\operatorname{Ham}}(\Sigma)\to\mathbb{R}^{\mathbb{N}}/\sim
φ\displaystyle\varphi (fL¯1(φ),fL¯2(φ),)\displaystyle\mapsto(f_{\underline{L}^{1}}(\varphi),f_{\underline{L}^{2}}(\varphi),...)

We claim that ff is a group homomorphism, that is, for every φ\varphi, ψ\psi in Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma):

fL¯k(φψ)fL¯k(φ)fL¯k(ψ)0f_{\underline{L}^{k}}(\varphi\psi)-f_{\underline{L}^{k}}(\varphi)-f_{\underline{L}^{k}}(\psi)\to 0

Since the spectral invariants satisfy the triangle inequality, and since Cal\operatorname{Cal} is a group homomorphism, we have the following inequality for every φ\varphi, ψ\psi in Ham(Σ)\operatorname{Ham}(\Sigma) and kk in \mathbb{N} :

fL¯k(φψ)fL¯k(φ)fL¯k(ψ)0f_{\underline{L}^{k}}(\varphi\psi)-f_{\underline{L}^{k}}(\varphi)-f_{\underline{L}^{k}}(\psi)\leqslant 0

This inequality still holds for the extension of fL¯kf_{\underline{L}^{k}} to Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma). We also have, using the triangle inequality :

fL¯k(φψ)fL¯k(φ)fL¯k(ψ)\displaystyle f_{\underline{L}^{k}}(\varphi\psi)-f_{\underline{L}^{k}}(\varphi)-f_{\underline{L}^{k}}(\psi) fL¯k(φψ)fL¯k(φ)fL¯k(φψ)fL¯k(φ1)\displaystyle\geqslant f_{\underline{L}^{k}}(\varphi\psi)-f_{\underline{L}^{k}}(\varphi)-f_{\underline{L}^{k}}(\varphi\psi)-f_{\underline{L}^{k}}(\varphi^{-1})
(fL¯k(φ)+fL¯k(φ1))\displaystyle\geqslant-(f_{\underline{L}^{k}}(\varphi)+f_{\underline{L}^{k}}(\varphi^{-1}))

Hence, the following property is enough to show that ff is a group homomorphism:

Proposition 23.

For all φHam¯(Σ)\varphi\in\overline{\operatorname{Ham}}(\Sigma), γL¯k(φ):=fL¯k(φ)+fL¯k(φ1)\gamma_{\underline{L}^{k}}(\varphi):=f_{\underline{L}^{k}}(\varphi)+f_{\underline{L}^{k}}(\varphi^{-1}) goes to zero when kk goes to infinity.

Proof.

Fix φHam¯(Σ)\varphi\in\overline{\operatorname{Ham}}(\Sigma). Using a standard fragmentation result (see for instance [Ban97] or [Sey13] for a more quantitative version), one can find φ1,,φn\varphi_{1},...,\varphi_{n} supported in disks D1,,DnD_{1},...,D_{n} such that φ=φ1φn\varphi=\varphi_{1}\circ...\circ\varphi_{n}. For each 1in1\leqslant i\leqslant n, we pick a smooth ψi\psi_{i} sending DiD_{i} to a disk that does not intersect α¯:=α1..,αg\underline{\alpha}:=\alpha_{1}\cup..\cup,\alpha_{g}.

Then we have, using the triangle inequality:

0γL¯k(φ)\displaystyle 0\leqslant\gamma_{\underline{L}^{k}}(\varphi) γL¯k(φi)\displaystyle\leqslant\sum\gamma_{\underline{L}^{k}}(\varphi_{i})
(γL¯k(ψiφiψi1)+γL¯k(ψi)+γL¯k(ψi1))\displaystyle\leqslant\sum(\gamma_{\underline{L}^{k}}(\psi_{i}\varphi_{i}\psi_{i}^{-1})+\gamma_{\underline{L}^{k}}(\psi_{i})+\gamma_{\underline{L}^{k}}(\psi_{i}^{-1}))
(γL¯k(ψiφiψi1)+2γL¯k(ψi))\displaystyle\leqslant\sum(\gamma_{\underline{L}^{k}}(\psi_{i}\varphi_{i}\psi_{i}^{-1})+2\gamma_{\underline{L}^{k}}(\psi_{i}))

Since for all ii, ψiφiψi1\psi_{i}\varphi_{i}\psi_{i}^{-1} is supported in a disk away from α¯\underline{\alpha}, we can apply Theorem 5 and Remark 12 to get that γL¯k(ψiφiψi1)k+1k+gλk+1k(k+g)\gamma_{\underline{L}^{k}}(\psi_{i}\varphi_{i}\psi_{i}^{-1})\leqslant\frac{k+1}{k+g}\lambda\leqslant\frac{k+1}{k(k+g)}, and hence this term goes to zero.

As for the other terms, since the ψi\psi_{i} are smooth, the Calabi property (Proposition 21) implies that γL¯k(ψi)\gamma_{\underline{L}^{k}}(\psi_{i}) goes to Cal(ψi)+Cal(ψi1)=0\operatorname{Cal}(\psi_{i})+\operatorname{Cal}(\psi_{i}^{-1})=0. Thus, γk(φ)\gamma_{k}(\varphi) goes to zero for any φ\varphi, and hence ff is a group homomorphism. ∎

For φ\varphi smooth, we have f(φ)=(Cal(φ),Cal(φ),)f(\varphi)=(\operatorname{Cal}(\varphi),\operatorname{Cal}(\varphi),...) since fL¯k(φ)f_{\underline{L}^{k}}(\varphi) converges to Cal(φ)\operatorname{Cal}(\varphi). Let Δ\Delta denote the vector (1,1,1,.)(1,1,1,....) in /\mathbb{R}^{\mathbb{N}}/\sim. Using Zorn’s lemma, we complete this vector into a base (a1=Δ,a2,a3,)(a_{1}=\Delta,a_{2},a_{3},...) of /\mathbb{R}^{\mathbb{N}}/\sim. Now let ss be the following map :

/\displaystyle\mathbb{R}^{\mathbb{N}}/\sim \displaystyle\to\mathbb{R}
λiai\displaystyle\sum\lambda_{i}a_{i} λ1\displaystyle\mapsto\lambda_{1}

Then sfs\circ f is a group homomorphism from Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) to \mathbb{R} that extends Cal\operatorname{Cal}. This completes the proof of Theorem 2.

4.2. Proof of Theorem 3

We now give a proof of Theorem 3. Once again, it is inspired of the proof of (2) found in [CGHM+22].

We start by fixing an equidistributed sequence of links L¯k=L1kLkkα1αg\underline{L}^{k}=L^{k}_{1}\cup...\cup L^{k}_{k}\cup\alpha_{1}\cup...\cup\alpha_{g}, such that diam(L¯k)=O(1k)\operatorname{diam}(\underline{L}^{k})=\operatorname{O}\left(\frac{1}{\sqrt{k}}\right). We define

N((L¯k)k):={φHameo(Σ,ω)|k(fL¯kφ)Cal(φ)) is bounded}N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}):=\{\varphi\in\operatorname{Hameo}(\Sigma,\omega)|\sqrt{k}(f_{\underline{L}^{k}}\varphi)-\operatorname{Cal}(\varphi))\text{ is bounded}\}

We claim:

Proposition 24.

N((L¯k)k)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}) is a normal sub-group of Hameo(Σ,ω)\operatorname{Hameo}(\Sigma,\omega)

Proof.

Let φ,ψN((L¯k)k)\varphi,\psi\in N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}), and θHameo(Σ)\theta\in\operatorname{Hameo}(\Sigma). Then, by the triangle inequality and the fact that Cal\operatorname{Cal} is a homomorphism,

kγL¯k(φ)k(fL¯k(φψ)Cal(φψ))k(fL¯k(φ)Cal(φ))k(fL¯k(ψ)Cal(ψ))0-\sqrt{k}\gamma_{\underline{L}^{k}}(\varphi)\leqslant\sqrt{k}(f_{\underline{L}^{k}}(\varphi\psi)-\operatorname{Cal}(\varphi\psi))-\sqrt{k}(f_{\underline{L}^{k}}(\varphi)-\operatorname{Cal}(\varphi))-\sqrt{k}(f_{\underline{L}^{k}}(\psi)-\operatorname{Cal}(\psi))\leqslant 0

Moreover, we have that

k(fL¯k(φ1)Cal(φ1))=kγL¯k(φ)k(fL¯k(φ)Cal(φ))\sqrt{k}(f_{\underline{L}^{k}}(\varphi^{-1})-\operatorname{Cal}(\varphi^{-1}))=\sqrt{k}\gamma_{\underline{L}^{k}}(\varphi)-\sqrt{k}(f_{\underline{L}^{k}}(\varphi)-\operatorname{Cal}(\varphi))

and the triangle inequality also implies that

kγL¯k(θ)k(fL¯k(φ)fL¯k(θφθ1))kγL¯k(θ)-\sqrt{k}\gamma_{\underline{L}^{k}}(\theta)\leqslant\sqrt{k}(f_{\underline{L}^{k}}(\varphi)-f_{\underline{L}^{k}}(\theta\varphi\theta^{-1}))\leqslant\sqrt{k}\gamma_{\underline{L}^{k}}(\theta)

Therefore, the following lemma proves that φψ\varphi\psi, φ1\varphi^{-1} and θφθ1\theta\varphi\theta^{-1} are in N((L¯k)k)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}), which concludes the proof of the proposition.

Lemma 25.

For all φHam¯(Σ)\varphi\in\overline{\operatorname{Ham}}(\Sigma), (kγL¯k(φ))(\sqrt{k}\gamma_{\underline{L}^{k}}(\varphi)) is bounded.

Proof.

Writing φ=φ1φN\varphi=\varphi_{1}...\varphi_{N} where each φi\varphi_{i} is supported in a disk DiD_{i}, and choosing some ψi\psi_{i} displacing DiD_{i} away from α¯\underline{\alpha}, we get as in the proof of Proposition 23 that

kγL¯k(φ)\displaystyle\sqrt{k}\gamma_{\underline{L}^{k}}(\varphi) ki=1N(γL¯k(ψiφiψi1)+2γL¯k(ψi))\displaystyle\leqslant\sqrt{k}\sum\limits_{i=1}^{N}(\gamma_{\underline{L}^{k}}(\psi_{i}\varphi_{i}\psi_{i}^{{}_{1}})+2\gamma_{\underline{L}^{k}}(\psi_{i}))
ki=1N(k+1k(k+g)+O(1k))\displaystyle\leqslant\sqrt{k}\sum\limits_{i=1}^{N}\left(\frac{k+1}{k(k+g)}+\operatorname{O}\left(\frac{1}{\sqrt{k}}\right)\right)
k+1k(k+g)N+O(1)\displaystyle\leqslant\frac{k+1}{\sqrt{k}(k+g)}N+\operatorname{O}(1)

where we used the Calabi property (Proposition 21), with diam(L¯k)=O(1k)\operatorname{diam}(\underline{L}^{k})=\operatorname{O}\left(\frac{1}{\sqrt{k}}\right). This shows that kγL¯k(φ)\sqrt{k}\gamma_{\underline{L}^{k}}(\varphi) is bounded for every φ\varphi. ∎

Remark 26.

We see in this proof that the terms γL¯k(ψi)\gamma_{\underline{L}^{k}}(\psi_{i}) are of higher order than the other ones. If we manage to show that for smooth elements, kγL¯kk\gamma_{\underline{L}^{k}} is bounded, then it would be the case for all φHam¯(Σ)\varphi\in\overline{\operatorname{Ham}}(\Sigma), and we could define an even smaller normal subgroup by considering kγL¯kk\gamma_{\underline{L}^{k}} instead of kγL¯k\sqrt{k}\gamma_{\underline{L}^{k}}.

It remains to show that for a certain choice of (L¯k)(\underline{L}^{k}), this subgroup is proper. The proof is similar to the one in [CGHM+22] in the case of the disk. There are three steps:

  • We show that N((L¯k)k)Ker(Cal)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}})\cap\operatorname{Ker}(\operatorname{Cal}) contains all the smooth elements;

  • We construct a hameomorphism TT, and choose an equidistributed sequence (L¯k)(\underline{L}^{k}), such that TT does not belong to N((L¯k)k)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}).

  • From this TT we can construct another hameomorphism with the same property that lies in Ker(Cal)\operatorname{Ker}(\operatorname{Cal}).

Lemma 27.

N((L¯k)k)Ker(Cal)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}})\cap\operatorname{Ker}(\operatorname{Cal}) contains all the smooth elements.

Proof.

It is a corollary of the Calabi property (Proposition 21). ∎

Now we construct the hameomorphism and the sequence of links. Fix gg disjoint and homologically independant circles α1,,αg\alpha_{1},...,\alpha_{g}. Let DD be a disk of area 1/21/2 in Σ\Sigma away from α¯\underline{\alpha}, and pick a point z0z_{0} in its interior. We fix a symplectomorphism Φ:(D{z0},ω)(S1×(0,12π],rdrdθ)\Phi:(D\setminus\{z_{0}\},\omega)\xrightarrow{\sim}\left(S^{1}\times\left(0,\frac{1}{\sqrt{2\pi}}\right],rdr\wedge d\theta\right).

We define an autonomous Hamiltonian HH on Σ{z0}\Sigma\setminus\{z_{0}\} as follow:

  • HH is supported inside Φ1(S1×(0,12π])D{z0}\Phi^{-1}\left(S^{1}\times\left(0,\frac{1}{2\sqrt{\pi}}\right]\right)\subset D\setminus\{z_{0}\};

  • H(θ,r)=h(πr2)H(\theta,r)=h(\pi r^{2}) is radial;

  • h:(0,14][0,+)h:\left(0,\frac{1}{4}\right]\to[0,+\infty) is decreasing, h(r)rah(r)\leqslant r^{-a} with equality on (0,18]\left(0,\frac{1}{8}\right] for some 12+122<a<1\frac{1}{2}+\frac{1}{2\sqrt{2}}<a<1.

Then, φH1\varphi^{1}_{H} defines a Hamiltonian diffeomorphism on Σ{z0}\Sigma\setminus\{z_{0}\} which acts as a rotation around the origin inside D{z0}D\setminus\{z_{0}\}. Therefore, it extends continuously to a homeomorphism TT that fixes z0z_{0}. We claim the following:

Proposition 28.

THameo(Σ)T\in\operatorname{Hameo}(\Sigma)

Proof.

We have to find a sequence of Hamiltonians (Kn)(K_{n}), supported in a compact subset of the interior of Σ\Sigma, such that:

  • ϕKn1C0T\phi^{1}_{K_{n}}\xrightarrow{C^{0}}T;

  • (ϕKnt)(\phi^{t}_{K_{n}}) is Cauchy for the C0C^{0} distance, uniformly in t[0,1]t\in[0,1];

  • the sequence (Kn)(K_{n}) is Cauchy for the Hofer norm.

Let Dn:={z0}Φ1(S1×(0,1π2n/a))ΣD_{n}:=\{z_{0}\}\cup\Phi^{-1}\left(S^{1}\times\left(0,\frac{1}{\sqrt{\pi 2^{n/a}}}\right)\right)\subset\Sigma. It has area 12n/a\frac{1}{2^{n/a}}.

We start with a sequence of smooth Hamiltonians (Hn)(H_{n}) such that:

  • HnH_{n} coincides with HH outside of DnD_{n};

  • Hn2nH_{n}\approx 2^{n} in DnD_{n}

  • Hn+1HnHof2n||H_{n+1}-H_{n}||_{\operatorname{Hof}}\leqslant 2^{n}

To construct such a sequence, we flatten HH inside DnD_{n}.

Since HnH_{n} coincides with HH outside of DnD_{n}, we have that ϕHn1T1=Id\phi^{1}_{H_{n}}\circ T^{-1}=\operatorname{Id} outside of DnD_{n}, and therefore ϕHn1C0T\phi^{1}_{H_{n}}\xrightarrow{C^{0}}T.

We will now construct a sequence (Kn)(K_{n}) such that ϕKn1=ϕHn1\phi^{1}_{K_{n}}=\phi^{1}_{H_{n}}, (Kn)(K_{n}) is Cauchy for the Hofer norm, and (ϕKnt)(\phi^{t}_{K_{n}}) is Cauchy for the C0C^{0} distance uniformly in tt.

We will use a lemma from [CGHM+22, Lemma 4.5]:

Lemma 29.

Let Δ\Delta be a Euclidean 22-disk equipped with an area form ω\omega of total area A. Suppose DΔD\subset\Delta is diffeomorphic to D2D^{2} and that Area(D)<AN\operatorname{Area}(D)<\frac{A}{N} some integer N>0N>0. Let F be a smooth Hamiltonian supported in the interior of DD. Then, we have:

dH(ϕF1,Id)FHofN+2Ad_{H}(\phi^{1}_{F},\operatorname{Id})\leqslant\frac{||F||_{\operatorname{Hof}}}{N}+2A

where dHd_{H} denotes the Hofer distance on Hamc(Δ,ω)\operatorname{Ham}_{c}(\Delta,\omega).

Let bb be a real number such that 1<b<1a1<b<\frac{1}{a}. Let N=2bnN=2^{\left\lfloor bn\right\rfloor}, and An=(N+1)2naA_{n}=(N+1)2^{-\frac{n}{a}}. If nn0n\geqslant n_{0} where n0n_{0} is large enough, A<12A<\frac{1}{2}, and we can define Δn:={z0}Φ1(S1×(0,1πAn))\Delta_{n}:=\{z_{0}\}\cup\Phi^{-1}\left(S^{1}\times\left(0,\frac{1}{\sqrt{\pi A_{n}}}\right)\right). It is a disk of area AnA_{n}. Hn+1HnH_{n+1}-H_{n} is supported inside DnD_{n}, which has area 2na<AnN2^{-\frac{n}{a}}<\frac{A_{n}}{N}, so we can apply the lemma and get that:

dH(ϕHn+1Hn1,Id)Hn+1HnHofN+2And_{H}(\phi^{1}_{H_{n+1}-H_{n}},\operatorname{Id})\leqslant\frac{||H_{n+1}-H_{n}||_{\operatorname{Hof}}}{N}+2A_{n}

Therefore there exists GnG_{n} supported in Δn\Delta_{n} such that ϕGn1=ϕHn+1Hn1=ϕHn1ϕHn+11\phi^{1}_{G_{n}}=\phi^{1}_{H_{n+1}-H_{n}}=\phi^{-1}_{H_{n}}\circ\phi^{1}_{H_{n+1}} and GnHof2n(1b)n+(2bn+1)2an||G_{n}||_{\operatorname{Hof}}\leqslant 2^{n-\left\lceil(1-b)n\right\rceil}+(2^{\left\lfloor bn\right\rfloor}+1)2^{-\frac{a}{n}}.

By definition of bb, the series n=n0GnHof\sum\limits_{n=n_{0}}^{\infty}||G_{n}||_{\operatorname{Hof}} is summable. Since GnG_{n} is supported inside Δn\Delta_{n}, dC0(ϕGnt,Id)diamΔn=O(2(b1a)n)d_{C^{0}}(\phi^{t}_{G_{n}},\operatorname{Id})\leqslant\operatorname{diam}\Delta_{n}=\operatorname{O}(2^{(b-\frac{1}{a})n}), so n=n0dC0(ϕGnt,Id)\sum\limits_{n=n_{0}}^{\infty}d_{C^{0}}(\phi^{t}_{G_{n}},\operatorname{Id}) is also summable, uniformly in tt.

Then, we define (Kn)(K_{n}) recursively by:

  • Kn=HnK_{n}=H_{n} for nn0n\leqslant n_{0};

  • Kn+1=Kn#GnK_{n+1}=K_{n}\#G_{n} for nn0n\geqslant n_{0}

We get that ϕKn1=ϕHn1\phi^{1}_{K_{n}}=\phi^{1}_{H_{n}} for nn0n\leqslant n_{0}, and for n>n0n>n_{0}:

ϕKn1\displaystyle\phi^{1}_{K_{n}} =ϕHn01ϕGn01ϕGn11\displaystyle=\phi^{1}_{H_{n_{0}}}\phi^{1}_{G_{n_{0}}}...\phi^{1}_{G_{n-1}}
=ϕHn01ϕHn01ϕHn0+11ϕHn11ϕHn1\displaystyle=\phi^{1}_{H_{n_{0}}}\phi^{-1}_{H_{n_{0}}}\phi^{1}_{H_{n_{0}+1}}...\phi^{-1}_{H_{n-1}}\phi^{1}_{H_{n}}
=ϕHn1\displaystyle=\phi^{1}_{H_{n}}

Moreover, the summability of n=n0GnHof\sum\limits_{n=n_{0}}^{\infty}||G_{n}||_{\operatorname{Hof}} and n=n0dC0(ϕGnt,Id)\sum\limits_{n=n_{0}}^{\infty}d_{C^{0}}(\phi^{t}_{G_{n}},\operatorname{Id}) implies that (Kn)(K_{n}) is Cauchy for the Hofer norm, and (ϕKnt)(\phi^{t}_{K_{n}}) is Cauchy for the C0C^{0} distance uniformly in tt.

This concludes the proof that THameo(Σ)T\in\operatorname{Hameo}(\Sigma).

We now construct an equidistributed sequence of admissible links as follow:

Fix an integer k1k\geqslant 1. For 0ik20\leqslant i\leqslant\lfloor\frac{\sqrt{k}}{2}\rfloor, denote by AiA_{i} the annulus S1×(iπk,i+1πk)D{z0}S^{1}\times\left(\frac{i}{\sqrt{\pi k}},\frac{i+1}{\sqrt{\pi k}}\right)\subset D\setminus\{z_{0}\}.

Let L1kL^{k}_{1} be the circle S1×{1π(k+1)}S^{1}\times\left\{\sqrt{\frac{1}{\pi(k+1)}}\right\}. It bounds a disk of area 1k+1\frac{1}{k+1}. For i1i\geqslant 1, each annulus AiA_{i} has area 1k((i+1)2i2)=2i+1k\frac{1}{k}((i+1)^{2}-i^{2})=\frac{2i+1}{k}, hence we can fit inside AiA_{i} 2i+12i+1 disjoint circles Li2+1k,,L(i+1)2kL^{k}_{i^{2}+1},...,L^{k}_{(i+1)^{2}} that bound disjoint disks of area 1k+1\frac{1}{k+1} and of diameter bounded by Ck\frac{C}{\sqrt{k}} where CC is a constant that does not depend on kk.

The union of all the annuli cover a disk of area (k2+1)2k\frac{\left(\lfloor\frac{\sqrt{k}}{2}\rfloor+1\right)^{2}}{k}. The remaining area in Σ\Sigma is k(k2+1)2k\frac{k-\left(\lfloor\frac{\sqrt{k}}{2}\rfloor+1\right)^{2}}{k}, which is enough to fit k(k2+1)2k-\left(\lfloor\frac{\sqrt{k}}{2}\rfloor+1\right)^{2} disjoint circles L(k2+1)2+1k,,LkkL^{k}_{\left(\lfloor\frac{\sqrt{k}}{2}\rfloor+1\right)^{2}+1},...,L^{k}_{k} that bound disjoint disks of area 1k+1\frac{1}{k+1} and of diameter bounded by Ck\frac{C^{\prime}}{\sqrt{k}} where CC^{\prime} is a constant that does not depend on kk.

Refer to caption
Figure 3. The sequence L¯k\underline{L}^{k}

Let L¯k:=L1kLkkα1αg\underline{L}^{k}:=L^{k}_{1}\cup...\cup L^{k}_{k}\cup\alpha_{1}\cup...\alpha_{g}. Then (L¯k)(\underline{L}^{k}) is an equidistributed sequence of monotone Lagrangian links, and moreover:

Proposition 30.

For this choice of equidistributed sequence, TN((L¯k)k)T\notin N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}).

Proof.

We want to show that k(fL¯k(T)Cal(T))\sqrt{k}(f_{\underline{L}^{k}}(T)-\operatorname{Cal}(T)) is unbounded.

First, we observe that:

Cal(T)\displaystyle\operatorname{Cal}(T) =limnCal(ϕKn1)=limnCal(ϕHn1)=limnΣHnω=ΣHω\displaystyle=\lim\limits_{n\to\infty}\operatorname{Cal}(\phi^{1}_{K_{n}})=\lim\limits_{n\to\infty}\operatorname{Cal}(\phi^{1}_{H_{n}})=\lim\limits_{n\to\infty}\int_{\Sigma}H_{n}\omega=\int_{\Sigma}H\omega

and

fL¯k(T)=limnfL¯k(ϕKn1)=limnfL¯k(ϕHn1)=limncL¯k(Hn)\displaystyle f_{\underline{L}^{k}}(T)=\lim\limits_{n\to\infty}f_{\underline{L}^{k}}(\phi^{1}_{K_{n}})=\lim\limits_{n\to\infty}f_{\underline{L}^{k}}(\phi^{1}_{H_{n}})=\lim\limits_{n\to\infty}c_{\underline{L}^{k}}(H_{n})

Since HnH_{n} coincides with HH outside of DnD_{n}, for nn sufficiently large, HnH_{n} coincides with HH on L¯k\underline{L}^{k} and therefore cL¯k(Hn)=ck(H)c_{\underline{L}^{k}}(H_{n})=c_{k}(H). Thus, fL¯k(T)=cL¯k(H)f_{\underline{L}^{k}}(T)=c_{\underline{L}^{k}}(H).

We start by estimating cL¯k(H)c_{\underline{L}^{k}}(H). Since HH is supported inside S1×(0,12π]S^{1}\times\left(0,\frac{1}{2\sqrt{\pi}}\right], and by Lagrangian control:

cL¯k(H)\displaystyle c_{\underline{L}^{k}}(H) 1k+gi=1(k2+1)2maxLikH\displaystyle\leqslant\frac{1}{k+g}\sum\limits_{i=1}^{\left(\lfloor\frac{\sqrt{k}}{2}\rfloor+1\right)^{2}}\max\limits_{L^{k}_{i}}H
1k(maxL1kH+i=1k2(2i+1)maxAiH)\displaystyle\leqslant\frac{1}{k}\left(\max\limits_{L^{k}_{1}}H+\sum\limits_{i=1}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}(2i+1)\max\limits_{A_{i}}H\right)
1k(h(1k+1)+i=1k2(2i+1)h(i2k))\displaystyle\leqslant\frac{1}{k}\left(h\left(\frac{1}{k+1}\right)+\sum\limits_{i=1}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}(2i+1)h\left(\frac{i^{2}}{k}\right)\right)
1k(h(1k+1)+3h(1k)+2i=2k2h(i2k)+i=2k2(2i1)h(i2k))\displaystyle\leqslant\frac{1}{k}\left(h\left(\frac{1}{k+1}\right)+3h\left(\frac{1}{k}\right)+2\sum\limits_{i=2}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}h\left(\frac{i^{2}}{k}\right)+\sum\limits_{i=2}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}(2i-1)h\left(\frac{i^{2}}{k}\right)\right)

Using that hh is decreasing, and comparing the sums with integrals, we get:

cL¯k(H)\displaystyle c_{\underline{L}^{k}}(H) (k+1)ak+3ka1+2k1k12h(r2)𝑑r+i=1k212i+1kminAiH\displaystyle\leqslant\frac{(k+1)^{a}}{k}+3k^{a-1}+\frac{2}{\sqrt{k}}\int_{\frac{1}{\sqrt{k}}}^{\frac{1}{2}}h(r^{2})dr+\sum\limits_{i=1}^{\lfloor\frac{\sqrt{k}}{2}\rfloor-1}\frac{2i+1}{k}\min\limits_{A_{i}}H
(k+1k)aka1+3ka1+2k[112ar12a]1k12+ΣA0Hω\displaystyle\leqslant\left(\frac{k+1}{k}\right)^{a}k^{a-1}+3k^{a-1}+\frac{2}{\sqrt{k}}\left[\frac{1}{1-2a}r^{1-2a}\right]_{\frac{1}{\sqrt{k}}}^{\frac{1}{2}}+\int_{\Sigma\setminus A_{0}}H\omega
(1+ak)ka1+3ka1+22a1ka1+ΣA0Hω\displaystyle\leqslant\left(1+\frac{a}{k}\right)k^{a-1}+3k^{a-1}+\frac{2}{2a-1}k^{a-1}+\int_{\Sigma\setminus A_{0}}H\omega

Therefore, for k8k\geqslant 8,

k(fL¯k(T)Cal(T))\displaystyle\sqrt{k}(f_{\underline{L}^{k}}(T)-\operatorname{Cal}(T)) =k(cL¯k(H)kΣHω)\displaystyle=\sqrt{k}(c_{\underline{L}^{k}}(H)-\sqrt{k}\int_{\Sigma}H\omega)
(4+ak+22a1)ka12kA0Hω\displaystyle\leqslant\left(4+\frac{a}{k}+\frac{2}{2a-1}\right)k^{a-\frac{1}{2}}-\sqrt{k}\int_{A_{0}}H\omega
(4+ak+22a1)ka12k01kh(r)𝑑r\displaystyle\leqslant\left(4+\frac{a}{k}+\frac{2}{2a-1}\right)k^{a-\frac{1}{2}}-\sqrt{k}\int_{0}^{\frac{1}{k}}h(r)dr
(4+ak+22a1)ka12k11aka1\displaystyle\leqslant\left(4+\frac{a}{k}+\frac{2}{2a-1}\right)k^{a-\frac{1}{2}}-\sqrt{k}\frac{1}{1-a}k^{a-1}
(4+ak+22a111a)ka12\displaystyle\leqslant\left(4+\frac{a}{k}+\frac{2}{2a-1}-\frac{1}{1-a}\right)k^{a-\frac{1}{2}}
(8a2+8a1(2a1)(1a)+ak)ka12\displaystyle\leqslant\left(\frac{-8a^{2}+8a-1}{(2a-1)(1-a)}+\frac{a}{k}\right)k^{a-\frac{1}{2}}

Since 12+122<a<1\frac{1}{2}+\frac{1}{2\sqrt{2}}<a<1, 8a2+8a1(2a1)(1a)+ak<0\frac{-8a^{2}+8a-1}{(2a-1)(1-a)}+\frac{a}{k}<0 for kk large enough, and k(fk(T)Cal(T))\sqrt{k}(f_{k}(T)-\operatorname{Cal}(T)) goes to -\infty.

In a similar fashion, we also compute a lower bound for cL¯k(H)ΣHωc_{\underline{L}^{k}}(H)-\int_{\Sigma}H\omega (when kk is large enough) that we will need in the following section:

cL¯k(H)ΣHω\displaystyle c_{\underline{L}^{k}}(H)-\int_{\Sigma}H\omega\geqslant 1k+gi=1(k2+1)2minLikHΣHω\displaystyle\frac{1}{k+g}\sum\limits_{i=1}^{\left(\lfloor\frac{\sqrt{k}}{2}\rfloor+1\right)^{2}}\min\limits_{L^{k}_{i}}H-\int_{\Sigma}H\omega
\displaystyle\geqslant 1k+g(minL1kH+i=1k2(2i+1)minAiH)ΣHω\displaystyle\frac{1}{k+g}\left(\min\limits_{L^{k}_{1}}H+\sum\limits_{i=1}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}(2i+1)\min\limits_{A_{i}}H\right)-\int_{\Sigma}H\omega
\displaystyle\geqslant 1k+g(h(1k+1)+i=1k2(2i+1)h((i+1)2k))ΣHω\displaystyle\frac{1}{k+g}\left(h\left(\frac{1}{k+1}\right)+\sum\limits_{i=1}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}(2i+1)h\left(\frac{(i+1)^{2}}{k}\right)\right)-\int_{\Sigma}H\omega
\displaystyle\geqslant 1k+g(h(1k+1)3h(1k)2i=2k2h(i2k)+i=1k2(2i+1)h(i2k))ΣHω\displaystyle\frac{1}{k+g}\left(h\left(\frac{1}{k+1}\right)-3h\left(\frac{1}{k}\right)-2\sum\limits_{i=2}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}h\left(\frac{i^{2}}{k}\right)+\sum\limits_{i=1}^{\lfloor\frac{\sqrt{k}}{2}\rfloor}(2i+1)h\left(\frac{i^{2}}{k}\right)\right)-\int_{\Sigma}H\omega
kk+g((k+1)ak3ka12k1k12h(r2)𝑑r+i=1k212i+1kmaxAiHΣHω)\displaystyle\frac{k}{k+g}\left(\frac{(k+1)^{a}}{k}-3k^{a-1}-\frac{2}{\sqrt{k}}\int_{\frac{1}{\sqrt{k}}}^{\frac{1}{2}}h(r^{2})dr+\sum\limits_{i=1}^{\lfloor\frac{\sqrt{k}}{2}\rfloor-1}\frac{2i+1}{k}\max\limits_{A_{i}}H-\int_{\Sigma}H\omega\right)
gk+gΣHω\displaystyle-\frac{g}{k+g}\int_{\Sigma}H\omega

Once again, we used that hh is decreasing and compared the first sum with an integral. Doing the same for the second sum, we obtain:

cL¯k(H)ΣHω\displaystyle c_{\underline{L}^{k}}(H)-\int_{\Sigma}H\omega\geqslant 11+gk((k+1k)aka13ka12k[112ar12a]1k12+ΣA0HωΣHω)\displaystyle\frac{1}{1+\frac{g}{k}}\left(\left(\frac{k+1}{k}\right)^{a}k^{a-1}-3k^{a-1}-\frac{2}{\sqrt{k}}\left[\frac{1}{1-2a}r^{1-2a}\right]_{\frac{1}{\sqrt{k}}}^{\frac{1}{2}}+\int_{\Sigma\setminus A_{0}}H\omega-\int_{\Sigma}H\omega\right)
gk014h(r)𝑑r\displaystyle-\frac{g}{k}\int_{0}^{\frac{1}{4}}h(r)dr
\displaystyle\geqslant (1gk)((1+1k)aka13ka122a1ka1A0Hω)gk014rα𝑑r\displaystyle\left(1-\frac{g}{k}\right)\left(\left(1+\frac{1}{k}\right)^{a}k^{a-1}-3k^{a-1}-\frac{2}{2a-1}k^{a-1}-\int_{A_{0}}H\omega\right)-\frac{g}{k}\int_{0}^{\frac{1}{4}}r^{-\alpha}dr
\displaystyle\geqslant (1gk)((1322a1)ka101kh(r)𝑑r)gk[r1α1α]014\displaystyle\left(1-\frac{g}{k}\right)\left(\left(1-3-\frac{2}{2a-1}\right)k^{a-1}-\int_{0}^{\frac{1}{k}}h(r)dr\right)-\frac{g}{k}\left[\frac{r^{1-\alpha}}{1-\alpha}\right]^{\frac{1}{4}}_{0}
\displaystyle\geqslant (1gk)((222a1)ka111aka1)4α1g(1α)k\displaystyle\left(1-\frac{g}{k}\right)\left(\left(-2-\frac{2}{2a-1}\right)k^{a-1}-\frac{1}{1-a}k^{a-1}\right)-\frac{4^{\alpha-1}g}{(1-\alpha)k}
\displaystyle\geqslant (1gk)(222a111a)ka14α1g(1α)k\displaystyle\left(1-\frac{g}{k}\right)\left(-2-\frac{2}{2a-1}-\frac{1}{1-a}\right)k^{a-1}-\frac{4^{\alpha-1}g}{(1-\alpha)k}

Now, it remains to construct a hameomorphism in Ker(Cal)\operatorname{Ker}(\operatorname{Cal}). Choose a smooth Hamiltonian diffeomorphism θ\theta such that Cal(θ)=Cal(T)\operatorname{Cal}(\theta)=\operatorname{Cal}(T). Then, T:=Tθ1Hameo(Σ)Ker(Cal)T^{\prime}:=T\theta^{-1}\in\operatorname{Hameo}(\Sigma)\cap\operatorname{Ker}(\operatorname{Cal}). Since θ\theta is smooth, by Lemma 27, θN(Σ)\theta\in N(\Sigma), and therefore TN((L¯k)k)T^{\prime}\notin N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}).

Hence, N((L¯k)k)Ker(Cal)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}})\cap\operatorname{Ker}(\operatorname{Cal}) is a proper normal subgroup of Hameo(Σ)Ker(Cal)\operatorname{Hameo}(\Sigma)\cap\operatorname{Ker}(\operatorname{Cal}), which concludes the proof of Theorem 3.

We give more precision on the subgroups we defined:

Note that since the Hamiltonian HH we constructed in this section is radial, we have H#n=nHH^{\#n}=nH, and therefore

μL¯k(H)=limncL¯k(H#n)n=cL¯k(H)\mu_{\underline{L}^{k}}(H)=\lim\limits_{n\to\infty}\frac{c_{\underline{L}^{k}}(H^{\#n})}{n}=c_{\underline{L}^{k}}(H)

Then, by Theorem 18, cL¯k(H)c_{\underline{L}^{k}}(H) only depends on kk, α¯\underline{\alpha} and the monotonicity constant of the link. Therefore, for any other equidistributed sequence of links (L¯k)(\underline{L}^{\prime k}) having the same non-contractible components α¯\underline{\alpha}, and the same monotonicity constant, we have that TN(L¯k)T\notin N_{(\underline{L}^{\prime k})}.

Since we fixed an arbitrary α¯\underline{\alpha} at the start of the proof, and since we could modify the definition of (L¯k)(\underline{L}^{k}) to change its monotonicity constant while keeping similar inequalities for cL¯k(H)c_{\underline{L}^{k}}(H), we get the following proposition:

Proposition 31.

Let (L¯k)(\underline{L}^{\prime k}) be an equidistributed sequence of admissible links, satisfying diam(L¯k)=O(1k)\operatorname{diam}(\underline{L}^{\prime k})=\operatorname{O}\left(\frac{1}{\sqrt{k}}\right). Then, N((L¯k)k)N((\underline{L}^{\prime k})_{k\in\mathbb{N}^{*}}) is a proper normal subgroup of Hameo(Σ)Ker(Cal)\operatorname{Hameo}(\Sigma)\cap\operatorname{Ker}(\operatorname{Cal}).

Moreover, taking the intersection of those subgroups over all such sequences of links, we get an even smaller proper normal subgroup NN, which contains all smooth elements.

Remark 32.

We need the assumption on the diameter to ensure that N((L¯k)k)N((\underline{L}^{\prime k})_{k\in\mathbb{N}^{*}}) is a normal subgroup (Proposition 24).

4.3. Proof of Theorem 4

This time we consider a connected, closed, oriented surface (Σ,ω)(\Sigma,\omega).

Fix an equidistributed sequence of admissible links (L¯k)(\underline{L}^{k}) with diam(L¯k)=O(1k)\operatorname{diam}(\underline{L}^{k})=\operatorname{O}\left(\frac{1}{\sqrt{k}}\right).

Fix aa such that 12+122<a<1\frac{1}{2}+\frac{1}{2\sqrt{2}}<a<1.

Then, for k0k_{0} large enough, 11a4a22k022a1>0\frac{1}{1-a}-4-\frac{a}{2^{2k_{0}}}-\frac{2}{2a-1}>0. Fix such a k0k_{0}, then for NN large enough, 222a111a+2N(1a)(11a4a22k022a1)>0-2-\frac{2}{2a-1}-\frac{1}{1-a}+2^{N(1-a)}\left(\frac{1}{1-a}-4-\frac{a}{2^{2k_{0}}}-\frac{2}{2a-1}\right)>0.

Fix such an integer, and define gk:=cL¯22kcL¯22kNg_{k}:=c_{\underline{L}^{2^{2k}}}-c_{\underline{L}^{2^{2k-N}}}.

Let N((L¯k)k):={φHameo(Σ),(2kgk(φ)) is bounded}N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}):=\{\varphi\in\operatorname{Hameo}(\Sigma),(2^{k}g_{k}(\varphi))\text{ is bounded}\}

Proposition 33.

N((L¯k)k)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}) is a normal subgroup of Hameo(Σ)\operatorname{Hameo}(\Sigma).

Proof.

Let φ,ψN((L¯k)k)\varphi,\psi\in N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}), and θHameo(Σ)\theta\in\operatorname{Hameo}(\Sigma). Using the triangle inequality, we have:

γL¯22k(φ)gk(φψ)gk(φ)gk(ψ)γL¯22kN(φ)\gamma_{\underline{L}^{2^{2k}}}(\varphi)\leqslant g_{k}(\varphi\psi)-g_{k}(\varphi)-g_{k}(\psi)\leqslant\gamma_{\underline{L}^{2^{2k-N}}}(\varphi)
gk(φ1)=γL¯22k(φ)γL¯22kN(φ)gk(φ)g_{k}(\varphi^{-1})=\gamma_{\underline{L}^{2^{2k}}}(\varphi)-\gamma_{\underline{L}^{2^{2k-N}}}(\varphi)-g_{k}(\varphi)

and

(γL¯22k(θ)+γL¯22kN(θ))gk(θφθ1)gk(φ)γL¯22k(θ)+γL¯22kN(θ)-(\gamma_{\underline{L}^{2^{2k}}}(\theta)+\gamma_{\underline{L}^{2^{2k-N}}}(\theta))\leqslant g_{k}(\theta\varphi\theta^{-1})-g_{k}(\varphi)\leqslant\gamma_{\underline{L}^{2^{2k}}}(\theta)+\gamma_{\underline{L}^{2^{2k-N}}}(\theta)

Therefore, Lemma 25 implies that φψ\varphi\psi, φ1\varphi^{-1} and θφθ1\theta\varphi\theta^{-1} are in N(Σ)N(\Sigma). ∎

We claim that for a certain choice of link, N((L¯k)k)N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}) is a proper subgroup of Hameo(Σ)\operatorname{Hameo}(\Sigma).

In fact, once again Proposition 21 shows that it contains all the smooth elements.

We define a hameomorphism TT and a sequence of links (L¯k)(\underline{L}^{k}) as in the previous section, with the parameter 12+122<a<1\frac{1}{2}+\frac{1}{2\sqrt{2}}<a<1 we fixed earlier.

We claim that TN((L¯k)k)T\notin N((\underline{L}^{k})_{k\in\mathbb{N}^{*}}).

Let kk be a sufficiently large integer. Then, by the estimates of the previous section:

2kgk(T)\displaystyle 2^{k}g_{k}(T) =2k(cL¯22k(H)cL¯22kN(H))\displaystyle=2^{k}\left(c_{\underline{L}^{2^{2k}}}(H)-c_{\underline{L}^{2^{2k-N}}}(H)\right)
=2k((cL¯22k(H)ΣHω)+(ΣHωcL¯22kN(H)))\displaystyle=2^{k}\left(\left(c_{\underline{L}^{2^{2k}}}(H)-\int_{\Sigma}H\omega\right)+\left(\int_{\Sigma}H\omega-c_{\underline{L}^{2^{2k-N}}}(H)\right)\right)
2k((1g22k)(222a111a)(22k)a14α1g(1α)22k)\displaystyle\geqslant 2^{k}\left(\left(1-\frac{g}{2^{2k}}\right)\left(-2-\frac{2}{2a-1}-\frac{1}{1-a}\right)(2^{2k})^{a-1}-\frac{4^{\alpha-1}g}{(1-\alpha)2^{2k}}\right)
+2k(11a4a22kN22a1)(22kN)a1\displaystyle+2^{k}\left(\frac{1}{1-a}-4-\frac{a}{2^{2k-N}}-\frac{2}{2a-1}\right)(2^{2k-N})^{a-1}
2k(2a1(222a111a+2N(1a)(11a4a22kN22a1))\displaystyle\geqslant 2^{k(2a-1}\left(-2-\frac{2}{2a-1}-\frac{1}{1-a}+2^{N(1-a)}\left(\frac{1}{1-a}-4-\frac{a}{2^{2k-N}}-\frac{2}{2a-1}\right)\right)
2k(2a3)g(222a111a)2k4α1g1α\displaystyle-2^{k(2a-3)}g\left(-2-\frac{2}{2a-1}-\frac{1}{1-a}\right)-2^{-k}\frac{4^{\alpha-1}g}{1-\alpha}

By definition of NN, for kk0+N2k\geqslant k_{0}+\frac{N}{2}, 222a111a+2N(1a)(11a4a22kN22a1)>0-2-\frac{2}{2a-1}-\frac{1}{1-a}+2^{N(1-a)}\left(\frac{1}{1-a}-4-\frac{a}{2^{2k-N}}-\frac{2}{2a-1}\right)>0, and therefore 2kgk(T)2^{k}g_{k}(T) goes to infinity, which concludes the proof of Theorem 4.

Using the same argument as in the previous section, we also get:

Proposition 34.

Let (L¯k)(\underline{L}^{\prime k}) be an equidistributed sequence of admissible links satisfying diam(L¯k)=O(1k)\operatorname{diam}(\underline{L}^{\prime k})=\operatorname{O}\left(\frac{1}{\sqrt{k}}\right). Then, N((L¯k)k)N((\underline{L}^{\prime k})_{k\in\mathbb{N}^{*}}) is a proper normal subgroup of Hameo(Σ)\operatorname{Hameo}(\Sigma).

Moreover, taking the intersection of those subgroups over all such sequences of links, we get an even smaller proper normal subgroup NN, which contains all smooth elements.

5. The Künneth formula for connected sums

This section is devoted to the proof of Theorem 6.

5.1. Heeagaard Floer Homology

Let us start by discussing how we define Heegaard Floer Homology. Indeed, we decided to use the original construction, which computes Lagrangian Floer Homology in the symmetric product. Alternatively, one could work in a cylindrical setting and define Heegaard Floer Homology by counting pseudo-holomorphic curves in the 4-manifold Σ×[0,1]×\Sigma\times[0,1]\times\mathbb{R}. We believe that this cylindrical reformulation (formulated by Lipshitz in [Lip06]) could be used to prove the statements of this paper since similar results are proved in [Lip06] and [OS08] in the cylindrical setting. Indeed, a cylindrical reformulation of the Lagrangian link spectral invariants is considered in [Che21], [Che22] so it is likely that the cylindrical approach together with our arguments in the previous sections can be combined to obtain Theorem 2, 3 and 4. However, since our main results are inspired by [CGHM+21], [CGHM+22] and [OS04], which are all using the symmetric product setting, we will do the same. This setting is the following.

Consider a closed symplectic surface (Σ,ω)(\Sigma,\omega), with a compatible complex structure jj. Let L¯=L1Lk\underline{L}=L_{1}\cup...\cup L_{k} be a Lagrangian link in Σ\Sigma. Denote by SymL¯\operatorname{Sym}\underline{L} the image of L1××LkL_{1}\times...\times L_{k} in the symmetric product Symk(Σ):=Σk/𝔖k\operatorname{Sym}^{k}(\Sigma):=\Sigma^{k}/\mathfrak{S}_{k}.

Denote by π\pi the projection ΣkSymk(Σ)\Sigma^{k}\to\operatorname{Sym}^{k}(\Sigma). The symmetric product is naturally endowed with a singular symplectic form Sym(ω):=π(ωk)\operatorname{Sym}(\omega):=\pi_{*}(\omega^{\oplus k}), which is smooth away from the diagonal Δ:=π({(x1,,xk),ij,xi=xj})\Delta:=\pi\left(\{(x_{1},...,x_{k}),\exists i\neq j,x_{i}=x_{j}\}\right). It is also endowed with a complex structure Symk(j):=πj×k\operatorname{Sym}^{k}(j):=\pi_{*}j^{\times k}.

Since the circles composing L¯\underline{L} are disjoint, SymL¯\operatorname{Sym}\underline{L} does not intersect Δ\Delta. Let VV be a neighborhood of Δ\Delta that does not intersect SymL¯\operatorname{Sym}\underline{L}. Then, one can find a smooth symplectic form ωV\omega_{V} on Symk(Σ)\operatorname{Sym}^{k}(\Sigma) which agrees with Sym(ω)\operatorname{Sym}(\omega) away from VV, and compatible with Symk(j)\operatorname{Sym}^{k}(j). Then, SymL¯\operatorname{Sym}\underline{L} is a Lagrangian submanifold inside (Symk(Σ),ωV)(\operatorname{Sym}^{k}(\Sigma),\omega_{V}).

Let (L¯,K¯)(\underline{L},\underline{K}) be a pair of Lagrangian links with kk components. Let H:S1×ΣH:S^{1}\times\Sigma\to\mathbb{R} be a smooth Hamiltonian. Then one can define a Hamiltonian Symk(H)\operatorname{Sym}^{k}(H) on Symk(Σ)\operatorname{Sym}^{k}(\Sigma) by the formula Symk(H)t({x1,,xk}):=Ht(x1)++Ht(xk)\operatorname{Sym}^{k}(H)_{t}(\{x_{1},...,x_{k}\}):=H_{t}(x_{1})+...+H_{t}(x_{k}).

Let VV be a neighborhood of Δ\Delta that does not intersect Sym(φHt(L¯))\operatorname{Sym}(\varphi_{H}^{t}(\underline{L})), Sym(φHt(K¯))\operatorname{Sym}(\varphi_{H}^{t}(\underline{K})) for t[1,1]t\in[-1,1].

Definition 35.

An almost complex structures JJ over Symk(Σ)\operatorname{Sym}^{k}(\Sigma) is VV-nearly symmetric if it agrees with Symk(j)\operatorname{Sym}^{k}(j) over VV, and tames Sym(ω)\operatorname{Sym}(\omega) outside of VV.

If SymL¯\operatorname{Sym}\underline{L} and SymK¯\operatorname{Sym}\underline{K} are both monotone Lagrangian submanifolds that are Hamiltonian isotopic and Sym(φH1(L¯))SymK¯\operatorname{Sym}(\varphi_{H}^{1}(\underline{L}))\pitchfork\operatorname{Sym}\underline{K}, then given a path of VV-nearly symmetric almost complex structures (Jt)t[0,1](J_{t})_{t\in[0,1]}, one can define the Lagrangian Floer cohomology

HF(SymL¯,SymK¯,Symk(H))HF^{*}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K},\operatorname{Sym}^{k}(H))

in the standard way (cf. [CGHM+21, Section 6]). One can show that this construction will not depend on the choice of VV and JtJ_{t}.

In order to clarify some notations, we will recall briefly how HF(L,K,H)HF^{*}(L,K,H) is constructed for two Hamiltonian isotopic monotone Lagrangians LL and KK inside a closed monotone symplectic manifold (M,ω)(M,\omega), and a smooth Hamiltonian HH.

We define the space 𝒫(L,K)\mathcal{P}(L,K) of smooth paths γ:[0,1]M\gamma:[0,1]\to M with γ(0)L\gamma(0)\in L and γ(1)K\gamma(1)\in K.

Fix η\eta in 𝒫(L,K)\mathcal{P}(L,K), and let P~η(L,K)\widetilde{P}_{\eta}(L,K) be the universal cover of the connected component of η\eta (with base point η\eta).

Given a Hamiltonian HH, we can define an action functional on P~η(L,K)\widetilde{P}_{\eta}(L,K), by :

𝒜H([γ,w]):=wω+H(t,γ(t))𝑑t\mathcal{A}_{H}([\gamma,w]):=-\int w^{*}\omega+\int H\left(t,\gamma(t)\right)dt

Here, ww is a homotopy from γ\gamma to η\eta in 𝒫(L,K)\mathcal{P}(L,K).

By the definition of P~η(L,K)\widetilde{P}_{\eta}(L,K), two cappings [γ,w][\gamma,w] and [γ,w][\gamma^{\prime},w^{\prime}] are isomorphic if γ=γ\gamma=\gamma^{\prime} and ww and ww^{\prime} coincide in the set π2(η,γ)\pi_{2}(\eta,\gamma) of homotopy classes of cappings from η\eta to γ\gamma with boundary in LL and KK.

Definition 36.

Two cappings [γ,w][\gamma,w] and [γ,w][\gamma^{\prime},w^{\prime}] are defined to be equivalent if γ=γ\gamma=\gamma^{\prime} and ww and ww^{\prime} have the same image in π2(η,γ)Kerω\frac{\pi_{2}(\eta,\gamma)}{\operatorname{Ker}\omega}.

Let CF(L,K,H;η)CF^{*}_{\circ}(L,K,H;\eta) be the \mathbb{C}-vector space generated by the equivalent classes of critical points of the action functional 𝒜H\mathcal{A}_{H}. It is naturally a [T±1]\mathbb{C}[T^{\pm 1}]-module where TT acts by adjoining the smallest positive area disk class in π2(η,η)\pi_{2}(\eta,\eta) to the capping. The Lagrangian Floer complex is

CF(L,K,H)\displaystyle CF^{*}(L,K,H) :=ηπ0(𝒫(L,K))CF(L,K,H;η)\displaystyle:=\oplus_{\eta\in\pi_{0}(\mathcal{P}(L,K))}CF^{*}(L,K,H;\eta)
CF(L,K,H;η)\displaystyle CF^{*}(L,K,H;\eta) :=CF(L,K,H;η)[T±1][[T]][T1].\displaystyle:=CF^{*}_{\circ}(L,K,H;\eta)\otimes_{\mathbb{C}[T^{\pm 1}]}\mathbb{C}[[T]][T^{-1}].

One can also think of CF(L,K,H)CF^{*}(L,K,H) as a [[T]][T1]\mathbb{C}[[T]][T^{-1}]-vector space generated by the critical points of the circle-valued action functional on 𝒫(L,K)\mathcal{P}(L,K) descended from 𝒜H\mathcal{A}_{H}. These critical points are trajectories of φHt\varphi_{H}^{t} from LL to KK, and are in one-to-one correspondance with φH1(L)K\varphi^{1}_{H}(L)\cap K. This complex is graded by the Maslov index, and the Novikov parameter carries a grading given by the minimal Maslov number of LL.

To define the differential, we fix a path of ω\omega-compatible almost complex structure JtJ_{t} on MM. Then, given two paths γ\gamma and γ\gamma^{\prime}, and a homotopy class β\beta of Maslov index 1, we define the space ~Jt,β(γ,γ)\widetilde{\mathcal{M}}_{J_{t},\beta}(\gamma,\gamma^{\prime}) of smooth maps u:×[0,1]Mu:\mathbb{R}\times[0,1]\to M satisfying:

  • us+Jt(utXH(u))=0\frac{\partial u}{\partial s}+J_{t}\left(\frac{\partial u}{\partial t}-X_{H}(u)\right)=0 (where (s,t)×[0,1](s,t)\in\mathbb{R}\times[0,1]);

  • uu has finite energy;

  • u(,0)Lu(\mathbb{R},0)\subset L and u(,1)Ku(\mathbb{R},1)\subset K;

  • [u]=β[u]=\beta;

  • uu is asymptotic to γ\gamma at s=s=-\infty, and to γ\gamma^{\prime} at s=+s=+\infty.

Denote by Jt,β(γ,γ)\mathcal{M}_{J_{t},\beta}(\gamma,\gamma^{\prime}) the quotient of this moduli space by the action of \mathbb{R} by translation, and by π2(γ,γ)\pi_{2}(\gamma,\gamma^{\prime}) the set of homotopy classes of Floer trajectories between γ\gamma and γ\gamma^{\prime}. The differential is then given by

[γ,w]:=[γ,w#β]Crit(𝒜H)#Jt,β(γ,γ)[γ,w#β]\partial[\gamma^{\prime},w]:=\sum\limits_{[\gamma,w\#\beta]\in Crit(\mathcal{A}_{H})}\#\mathcal{M}_{J_{t},\beta}(\gamma,\gamma^{\prime})[\gamma,w\#\beta]

5.2. Identification of the vector spaces

Let L¯\underline{L} and K¯\underline{K} be two transverse Lagrangian links with kk components on a closed surface (Σ,ω)(\Sigma,\omega). Let (E,ωE,jE)(E,\omega_{E},j_{E}) denote the two-dimension torus with complex structure jEj_{E} and Kähler form ωE\omega_{E}. Let α\alpha be a non-contractible circle on EE and α\alpha^{\prime} be a small Hamiltonian deformation of α\alpha, such that α\alpha and α\alpha^{\prime} are transverse. Let σ1Σ(L¯K¯)\sigma_{1}\in\Sigma\setminus(\underline{L}\cup\underline{K}), and σ2\sigma_{2} be a point in EE away from the isotopy between α\alpha and α\alpha^{\prime}.

We denote by Σ(T)\Sigma^{\prime}(T) the connected sum of Σ\Sigma and EE along the points σ1\sigma_{1} and σ2\sigma_{2}, that we construct in the following way:

Pick small real numbers r1r_{1} and r2r_{2}, and fix conformal identifications Φ1:Br1(σ1){σ1}[0,)×S1\Phi_{1}:B_{r_{1}}(\sigma_{1})\setminus\{\sigma_{1}\}\xrightarrow{\simeq}[0,\infty)\times S^{1} and Φ2:Br2(σ2){σ2}[0,)×S1\Phi_{2}:B_{r_{2}}(\sigma_{2})\setminus\{\sigma_{2}\}\xrightarrow{\simeq}[0,\infty)\times S^{1} (where Br(z)B_{r}(z) denotes the closed ball of radius rr centered at zz). Let Σ(2T):=ΣΦ11((2T,)×S1)\Sigma(2T):=\Sigma\setminus\Phi_{1}^{-1}\left((2T,\infty)\times S^{1}\right) and E(2T):=EΦ21((2T,)×S1)E(2T):=E\setminus\Phi_{2}^{-1}\left((2T,\infty)\times S^{1}\right). Then, Σ(T)\Sigma^{\prime}(T) is the union of Σ(2T)\Sigma(2T) and E(2T)E(2T) modulo the identification of the cylinders [0,2T]×S1Σ(2T)[0,2T]\times S^{1}\subset\Sigma(2T) and [0,2T]×S1E(2T)[0,2T]\times S^{1}\subset E(2T) via the involution (t,θ)(2Tt,θ)(t,\theta)\sim(2T-t,\theta). We denote the resulting complex structure on Σ(T)\Sigma^{\prime}(T) by j(T)j^{\prime}(T), which agrees with jj over ΣBr1(σ1)\Sigma\setminus B_{r_{1}}(\sigma_{1}), agrees with jEj_{E} over EBr2(σ2)E\setminus B_{r_{2}}(\sigma_{2}) and agrees with the standard complex structure over the tube [0,2T]×S1[0,2T]\times S^{1}.

We assume the Hamiltonian isotopy from α\alpha to α\alpha^{\prime} is small enough such that the area of its support is less than ω(Br1(σ1))\omega(B_{r_{1}}(\sigma_{1})). In this case, we can equip Σ(T)\Sigma^{\prime}(T) with a symplectic form ω(T)\omega^{\prime}(T) which agrees with ω\omega over ΣBr1(σ1)\Sigma\setminus B_{r_{1}}(\sigma_{1}), agrees with ωE\omega_{E} over the support of the Hamiltonian isotopy from α\alpha to α\alpha^{\prime}, is compatible with j(T)j^{\prime}(T), and ω(T)(Σ(T))=ω(Σ)\omega^{\prime}(T)(\Sigma^{\prime}(T))=\omega(\Sigma).

Let W:={σ1}×Symk1(Σ)Symk(Σ)W:=\{\sigma_{1}\}\times\operatorname{Sym}^{k-1}(\Sigma)\subset\operatorname{Sym}^{k}(\Sigma). Let σ\sigma be a point that lies in the same connected component of Σ(L¯K¯)\Sigma\setminus(\underline{L}\cup\underline{K}) as σ1\sigma_{1}, but away from Br1(σ1)B_{r_{1}}(\sigma_{1}).

For any zΣ(L¯K¯)z\in\Sigma\setminus(\underline{L}\cup\underline{K}) and φH2(Symk(Σ),Sym(L¯)Sym(K¯))\varphi\in H_{2}(\operatorname{Sym}^{k}(\Sigma),\operatorname{Sym}(\underline{L})\cup\operatorname{Sym}(\underline{K})), we denote by nz(φ)n_{z}(\varphi) the intersection number of φ\varphi with {z}×Symk1(Σ)Symk(Σ)\{z\}\times\operatorname{Sym}^{k-1}(\Sigma)\subset\operatorname{Sym}^{k}(\Sigma). Similarly, for zΣ(L¯K¯αα)z^{\prime}\in\Sigma^{\prime}\setminus(\underline{L}\cup\underline{K}\cup\alpha\cup\alpha^{\prime}) and φH2(Symk+1(Σ),Sym(L¯α)Sym(K¯α))\varphi^{\prime}\in H_{2}(\operatorname{Sym}^{k+1}(\Sigma^{\prime}),\operatorname{Sym}(\underline{L}\cup\alpha)\cup\operatorname{Sym}(\underline{K}\cup\alpha^{\prime})), we denote by nz(φ)n^{\prime}_{z^{\prime}}(\varphi^{\prime}) the intersection number of φ\varphi^{\prime} with {z}×Symk(Σ)Symk+1(Σ)\{z^{\prime}\}\times\operatorname{Sym}^{k}(\Sigma^{\prime})\subset\operatorname{Sym}^{k+1}(\Sigma^{\prime}) where . For zEE(αα)z_{E}\in E\setminus(\alpha\cup\alpha^{\prime}), and φE\varphi_{E} in H2(E,αα)H_{2}(E,\alpha\cup\alpha^{\prime}), we denote by nzEE(φE)n^{E}_{z_{E}}(\varphi_{E}) the intersection number of φE\varphi_{E} with zEz_{E}.

In order to prove Theorem 6, we start by establishing an isomorphism of vector spaces between the Floer complexes. We will show that there is a one-to-one correspondence between generators.

Given an intersection point x={x1,,xk}Symk(Σ)x=\{x_{1},...,x_{k}\}\in\operatorname{Sym}^{k}(\Sigma) between SymL¯\operatorname{Sym}\underline{L} and SymK¯\operatorname{Sym}\underline{K}, and an intersection point cEc\in E between α\alpha and α\alpha^{\prime}, we get an intersection point x×{c}Symk(ΣBr1(σ1))×Sym1(EBr2(σ2))Symk+1(Σ#TE)x\times\{c\}\in\operatorname{Sym}^{k}(\Sigma\setminus B_{r_{1}}(\sigma_{1}))\times\operatorname{Sym}^{1}(E\setminus B_{r_{2}}(\sigma_{2}))\subset\operatorname{Sym}^{k+1}(\Sigma\#_{T}E) between Sym(L¯α)\operatorname{Sym}(\underline{L}\cup\alpha) and Sym(K¯α)\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}).

Since α\alpha and α\alpha^{\prime} do not intersect L¯\underline{L} and K¯\underline{K}, any intersection point between Sym(L¯α)\operatorname{Sym}(\underline{L}\cup\alpha) and Sym(K¯α)\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}) can be decomposed in a single way as x×{c}x\times\{c\} where xSym(L¯)Sym(K¯)x\in\operatorname{Sym}(\underline{L})\cap\operatorname{Sym}(\underline{K}) and cααc\in\alpha\cap\alpha^{\prime}, and therefore there is a one-to-one correspondence between (Sym(L¯)SymK¯))×(αα)\left(\operatorname{Sym}(\underline{L})\cap\operatorname{Sym}\underline{K})\right)\times(\alpha\cap\alpha^{\prime}) and Sym(L¯α)Sym(K¯α)\operatorname{Sym}(\underline{L}\cup\alpha)\cap\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}).

Now we need to consider the cappings. In fact, recall that a generator of the Floer complex is an equivalence class of an intersection point xx together with a capping ww (cf. Definition 36).

For each connected component of 𝒫(SymL¯,SymK¯)\mathcal{P}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K}) that contains an intersection point of SymL¯\operatorname{Sym}\underline{L} and SymK¯\operatorname{Sym}\underline{K}, we choose the reference path η\eta for that connected component to be a constant path at one the intersection points that are contained in that component. On EE, since we assumed that α\alpha^{\prime} is a Hamiltonian perturbation of α\alpha, we can assume that all intersection points between them are in the same connected component of 𝒫(α,α)\mathcal{P}(\alpha,\alpha^{\prime}). We choose as a reference path ηE\eta_{E} to be a constant path equal to an intersection point between α\alpha and α\alpha^{\prime}. For every η\eta chosen above, We define η:=η×ηE\eta^{\prime}:=\eta\times\eta_{E}, and choose it as a reference path in Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime}). Clearly, the constant path at any intersection point between Sym(L¯α)\operatorname{Sym}(\underline{L}\cup\alpha) and Sym(K¯α)\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}) is homotopic to some η\eta^{\prime}. Moreover, if η1\eta_{1} and η2\eta_{2} are two reference paths in 𝒫(SymL¯,SymK¯)\mathcal{P}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K}) that are not in the same connected component, then η1=η1×ηE\eta_{1}^{\prime}=\eta_{1}\times\eta_{E} and η2=η2×ηE\eta_{2}^{\prime}=\eta_{2}\times\eta_{E} do not lie in the same connected component too.

Definition 37.

Let L¯\underline{L} and K¯\underline{K} be two Lagrangian links on Σ\Sigma away from a point zz. Let xx be a path between SymL¯\operatorname{Sym}\underline{L} and SymK¯\operatorname{Sym}\underline{K}. A class φ\varphi in π2(x,x)\pi_{2}(x,x) is said to be periodic if nz(φ)=0n_{z}(\varphi)=0. The set of periodic classes will be denoted Πx(z)\Pi_{x}(z).

Then, we show the following lemma (which is a generalization of [OS04, Proposition 2.15], which corresponds to the case k=g):

Lemma 38.

Let (Σ,z)(\Sigma,z) be a pointed surface. Let L¯\underline{L} and K¯\underline{K} be two Lagrangian links on Σ\Sigma, away from zz, with kk components. Then, for any path xx from SymL¯\operatorname{Sym}\underline{L} to SymK¯\operatorname{Sym}\underline{K},

π2(x,x)π2(Symk(Σ))Πx(z)\pi_{2}(x,x)\cong\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\oplus\Pi_{x}(z)
Proof.

π2(x,x)\pi_{2}(x,x) is the fundamental group of 𝒫(SymL¯,SymK¯)\mathcal{P}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K}) based at the point xx. The evaluation at both ends of the path gives rise to a fibration

ΩSymk(Σ)𝒫(SymL¯,SymK¯)SymL¯×SymK¯\Omega\operatorname{Sym}^{k}(\Sigma)\to\mathcal{P}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K})\to\operatorname{Sym}\underline{L}\times\operatorname{Sym}\underline{K}

The corresponding long exact sequence gives

0π2(Symk(Σ))π1(𝒫(SymL¯,SymK¯))π1(SymL¯×SymK¯)𝑓π1(Symk(Σ))0\to\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\to\pi_{1}(\mathcal{P}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K}))\to\pi_{1}(\operatorname{Sym}\underline{L}\times\operatorname{Sym}\underline{K})\xrightarrow{f}\pi_{1}(\operatorname{Sym}^{k}(\Sigma))

One can rewrite it as a short exact sequence

0π2(Symk(Σ))π2(x,x)Ker(f)00\to\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\to\pi_{2}(x,x)\to\operatorname{Ker}(f)\to 0

Since kgk\geq g, we have either π2(Symk(Σ))\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\cong\mathbb{Z} or π2(Symk(Σ))=0\pi_{2}(\operatorname{Sym}^{k}(\Sigma))=0, and π2(Symk(Σ))=0\pi_{2}(\operatorname{Sym}^{k}(\Sigma))=0 happens only when k=g=1k=g=1 (cf. [BRa14, Theorem 5.4]). When π2(Symk(Σ))\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\cong\mathbb{Z}, the map π2(x,x)nz\pi_{2}(x,x)\xrightarrow{n_{z}}\mathbb{Z} gives a splitting of the short exact sequence, and π2(x,x)π2(Symk(Σ))Ker(f)\pi_{2}(x,x)\cong\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\oplus\operatorname{Ker}(f). When π2(Symk(Σ))=0\pi_{2}(\operatorname{Sym}^{k}(\Sigma))=0, we also have π2(x,x)π2(Symk(Σ))Ker(f)\pi_{2}(x,x)\cong\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\oplus\operatorname{Ker}(f). Since nzn_{z} gives a splitting of the sequence, Ker(f)\operatorname{Ker}(f) can be identified with Ker(nz)=Πx(z)\operatorname{Ker}(n_{z})=\Pi_{x}(z), which shows that π2(x,x)π2(Symk(Σ))Πx(z)\pi_{2}(x,x)\cong\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\oplus\Pi_{x}(z).

Lemma 39.

Given an intersection point x={x1,,xk}Symk(Σ)x=\{x_{1},...,x_{k}\}\in\operatorname{Sym}^{k}(\Sigma) between SymL¯\operatorname{Sym}\underline{L} and SymK¯\operatorname{Sym}\underline{K}, all cappings are of the form [x,nS#w][x,nS\#w] where SS is a generator of π2(Symk(Σ))\pi_{2}(\operatorname{Sym}^{k}(\Sigma)) whose intersection number with WW is nσ1(S)=1n_{\sigma_{1}}(S)=1, ww is a capping from xx to η\eta that does not intersect WW, and nn is an integer.

Proof.

Fix a capping wxw_{x} from η\eta to xx that does not intersect WW. Then, π2(x,η)=wx#π2(η,η)\pi_{2}(x,\eta)=w_{x}\#\pi_{2}(\eta,\eta). By the previous lemma, π2(η,η)π2(Symk(Σ))Πη(σ1)\pi_{2}(\eta,\eta)\cong\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\oplus\Pi_{\eta}(\sigma_{1}). Therefore,

π2(x,η)π2(Symk(Σ))(wx#Πη(σ1))\pi_{2}(x,\eta)\cong\pi_{2}(\operatorname{Sym}^{k}(\Sigma))\oplus(w_{x}\#\Pi_{\eta}(\sigma_{1}))

and elements of Πη(σ1)#wx\Pi_{\eta}(\sigma_{1})\#w_{x} do not intersect WW. ∎

Given an intersection point cEc\in E between α\alpha and α\alpha^{\prime}, since EE is aspherical, all cappings from cc to ηE\eta_{E} do not intersect σ2\sigma_{2}. Moreover, since α\alpha and α\alpha^{\prime} are Hamiltonian isotopic to each other, any two cappings would have the same areas and hence descend to a unique equivalence class.

Lemma 40.

Given an intersection point x×{c}x\times\{c\} between Sym(L¯α)\operatorname{Sym}(\underline{L}\cup\alpha) and Sym(K¯α)\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}), a capping ww from xx to η\eta that does not intersect WW, and wEw_{E} from cc to ηE\eta_{E} that does not intersect σ2\sigma_{2}, w×wEw\times w_{E} is a capping from x×{c}x\times\{c\} to η\eta^{\prime} such that nσ(w×wE)=0n^{\prime}_{\sigma}(w\times w_{E})=0. Moreover, all cappings from x×{c}x\times\{c\} to η\eta^{\prime} are of the form [x×{c},nS#(w×wE)][x\times\{c\},nS^{\prime}\#(w\times w_{E})] for some integer nn and cappings ww and wEw_{E} as above, and where SS^{\prime} is a generator of π2(Symk+1(Σ))\pi_{2}(\operatorname{Sym}^{k+1}(\Sigma^{\prime})) with nσ(S)=1n^{\prime}_{\sigma}(S^{\prime})=1.

Proof.

The first part of the lemma is straightforward. The proof of the second part is identical to that of the previous lemma. ∎

Proposition 41.

The linear map defined by

Φ:CF(SymL¯,SymK¯)CF(α,α)\displaystyle\Phi:CF^{*}(\operatorname{Sym}\underline{L},\operatorname{Sym}\underline{K})\otimes CF^{*}(\alpha,\alpha^{\prime}) CF(Sym(L¯α),Sym(K¯α))\displaystyle\to CF^{*}(\operatorname{Sym}(\underline{L}\cup\alpha),\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}))
[x,nS#w][c,wE]\displaystyle[x,nS\#w]\otimes[c,w_{E}] [x×{c},nS#(w×wE)]\displaystyle\mapsto[x\times\{c\},nS^{\prime}\#(w\times w_{E})]

is an isomorphism of vector spaces.

Proof.

We already know there is a one-to-one correspondence between (Sym(L¯)SymK¯))×(αα)\left(\operatorname{Sym}(\underline{L})\cap\operatorname{Sym}\underline{K})\right)\times(\alpha\cap\alpha^{\prime}) and Sym(L¯α)Sym(K¯α)\operatorname{Sym}(\underline{L}\cup\alpha)\cap\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}), and according to the previous lemma, the mapping Ψ:π2(x,η)×π2(c,ηE)π2(x×{c},η)\Psi:\pi_{2}(x,\eta)\times\pi_{2}(c,\eta_{E})\to\pi_{2}(x\times\{c\},\eta^{\prime}) is also a one-to-one correspondence. It remains to show that this mapping descends to a one-to-one correspondence between equivalence classes of cappings

π2(x,η)/Ker(Sym(ω))×π2(c,ηE)/Ker(ωE)π2(x×{c},η)/Ker(Sym(ω(T))).\pi_{2}(x,\eta)/\operatorname{Ker}(\operatorname{Sym}(\omega))\times\pi_{2}(c,\eta_{E})/\operatorname{Ker}(\omega_{E})\to\pi_{2}(x\times\{c\},\eta^{\prime})/\operatorname{Ker}(\operatorname{Sym}(\omega^{\prime}(T))).

First note that in the torus, π2(c,ηE)/Ker(ωE)\pi_{2}(c,\eta_{E})/\operatorname{Ker}(\omega_{E}) is trivial (see the paragraph before Lemma 40).

Then recall that ω(T)\omega^{\prime}(T) is chosen such that it agrees with ω\omega over ΣBr1(σ1)\Sigma\setminus B_{r_{1}}(\sigma_{1}), agrees with ωE\omega_{E} over the support of the Hamiltonian isotopy from α\alpha to α\alpha^{\prime}, is compatible with j(T)j^{\prime}(T) and ω(T)(Σ(T))=ω(Σ)\omega^{\prime}(T)(\Sigma^{\prime}(T))=\omega(\Sigma). These conditions guarantee that for all ww and wEw_{E}, we have Sym(ω(T))(Ψ(w,wE))=Sym(ω)(w)+ωE(wE)\operatorname{Sym}(\omega^{\prime}(T))(\Psi(w,w_{E}))=\operatorname{Sym}(\omega)(w)+\omega_{E}(w_{E}). It implies the result. ∎

To show that it is an isomorphism of chain complexes, we need to show that it preserves the differential, i.e. that for all such xx and cc, (x×{c})=Φ((x)c+x(c))\partial(x\times\{c\})=\Phi((\partial x)\otimes c+x\otimes(\partial c))222More precisely, we need to identify the differentials of the capped intersection points, but the identification of the cappings is straightforward so we focus on the intersection points..

In order to do this, we compare the moduli space (x,y)\mathcal{M}(x,y) of Maslov index 11 Floer trajectories from xx to some yy in Symk(Σ)\operatorname{Sym}^{k}(\Sigma) to (x×{c},y×{c})\mathcal{M}(x\times\{c\},y\times\{c\}), and (c,c)\mathcal{M}(c,c^{\prime}) to (x×{c},x×{c})\mathcal{M}(x\times\{c\},x\times\{c^{\prime}\}).

In fact, we will show that these moduli spaces can be identified when considering a complex structure on Σ=Σ#E\Sigma^{\prime}=\Sigma\#E that stretches sufficiently the connected sum tube.

This is a generalization of a statement in [OS04] which only considers the case of links with gg components (where gg is the genus of Σ\Sigma), and circles α\alpha and α\alpha^{\prime} with a single intersection point. The proof of this statement still works in our setting. We will recall the main steps of this proof and emphasize where one have to be careful when generalizing.

Before discussing the moduli spaces of Floer trajectories, which are pseudo-holomorphic disks, we have to fix paths of almost complex structures on each manifold.

We fixa path of VV-nearly symmetric almost complex structures (Jt)t[0,1](J_{t})_{t\in[0,1]} on Symk(Σ)\operatorname{Sym}^{k}(\Sigma), for some neighborhood VV of ΔSymk1(Σ)×{σ1}Symk(Σ)\Delta\cup\operatorname{Sym}^{k-1}(\Sigma)\times\{\sigma_{1}\}\subset\operatorname{Sym}^{k}(\Sigma).

Recall that j(T)j^{\prime}(T) is the complex structure on Σ=Σ#TE\Sigma^{\prime}=\Sigma\#_{T}E that coincides with jj on ΣBr1(σ1)\Sigma\setminus B_{r_{1}}(\sigma_{1}), with jEj_{E} on EBr2(σ2)E\setminus B_{r_{2}}(\sigma_{2}), and is the standard cylindrical complex structure on the connected sum tube [0,2T]×S1[0,2T]\times S^{1} between Σ\Sigma and EE.

Then, the symmetric product Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime}) endowed with the complex structure Symk+1(j(T))\operatorname{Sym}^{k+1}(j^{\prime}(T)) admits an open subset holomorphically identified with

Symk(ΣBr1(σ1))×Sym1(EBr2(σ2))\operatorname{Sym}^{k}(\Sigma-B_{r_{1}}(\sigma_{1}))\times\operatorname{Sym}^{1}(E-B_{r_{2}}(\sigma_{2}))

Fix R1>r1R_{1}>r_{1} and R2>r2R_{2}>r_{2}. We choose a path of almost complex structures Jt(T)J^{\prime}_{t}(T) on Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime}) satisfying the following conditions:

  • Jt(T)Jt×jEJ^{\prime}_{t}(T)\equiv J_{t}\times j_{E} on Symk(ΣBR1(σ1))×Sym1(EBR2(σ2))\operatorname{Sym}^{k}(\Sigma-B_{R_{1}}(\sigma_{1}))\times\operatorname{Sym}^{1}(E-B_{R_{2}}(\sigma_{2}))

  • Jt(T)=Jt,r×jEJ^{\prime}_{t}(T)=J_{t,r}\times j_{E} on Symk(ΣBr1(σ1))×Sym1(BR2(σ2)Br2(σ2))\operatorname{Sym}^{k}(\Sigma-B_{r_{1}}(\sigma_{1}))\times\operatorname{Sym}^{1}(B_{R_{2}}(\sigma_{2})-B_{r_{2}}(\sigma_{2})), where Jt,rJ_{t,r} is VV-nearly symmetric for all rr and connects JtJ_{t} to Symk(j)\operatorname{Sym}^{k}(j) as rr, the normal parameter to σ2\sigma_{2}, goes from R2R_{2} to r2r_{2}.

  • Jt(T)Symk+1(j(T))J^{\prime}_{t}(T)\equiv\operatorname{Sym}^{k+1}(j^{\prime}(T)) on the rest of Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime})

In particular, Jt(T)J^{\prime}_{t}(T) is VV^{\prime}-nearly symmetric for some neighborhood VV^{\prime} of the diagonal ΔSymk+1(Σ)\Delta^{\prime}\subset\operatorname{Sym}^{k+1}(\Sigma^{\prime}).

Let x,ySym(L¯)Sym(K¯)x,y\in\operatorname{Sym}(\underline{L})\cap\operatorname{Sym}(\underline{K}), and c,cααc,c^{\prime}\in\alpha\cap\alpha^{\prime}.

Given φπ2(x,y)\varphi\in\pi_{2}(x,y), there is a single class φcπ2(x×{c},y×{c})\varphi^{\prime}_{c}\in\pi_{2}(x\times\{c\},y\times\{c\}) such that for any zΣ(L¯K¯)z\in\Sigma\setminus(\underline{L}\cup\underline{K}), nz(φ)=nz(φc)n_{z}(\varphi)=n^{\prime}_{z}(\varphi^{\prime}_{c}). Similarly, for any φEπ2(c,c)\varphi_{E}\in\pi_{2}(c,c^{\prime}), there is a single class φE,xπ2(x×{c},x×{c})\varphi^{\prime}_{E,x}\in\pi_{2}(x\times\{c\},x\times\{c^{\prime}\}) such that for any zEE(αα)z_{E}\in E\setminus(\alpha\cup\alpha^{\prime}), nzEE(φE)=nzE(φE,x)n^{E}_{z_{E}}(\varphi_{E})=n^{\prime}_{z_{E}}(\varphi^{\prime}_{E,x}).

Then, Theorem 6 is a consequence of the following statement:

Theorem 42.

Let φπ2(x,y)\varphi\in\pi_{2}(x,y) and φEπ2(c,c)\varphi_{E}\in\pi_{2}(c,c^{\prime}) be two classes of Maslov index 11. Then, for sufficiently large TT, Jt,φ(x,y)Jt(T),φc(x×{c},y×{c})\mathcal{M}_{J_{t},\varphi}(x,y)\simeq\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{c}}(x\times\{c\},y\times\{c\}) and jE,φE(c,c)Jt(T),φE,x(x×{c},x×{c})\mathcal{M}_{j_{E},\varphi_{E}}(c,c^{\prime})\simeq\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{E,x}}(x\times\{c\},x\times\{c^{\prime}\}).

Remark 43.

The isomorphisms above are identifications between deformation theories, and therefore μ(φc)=μ(φ)=1\mu(\varphi^{\prime}_{c})=\mu(\varphi)=1, and μ(φE,x)=μ(φE)=1\mu(\varphi^{\prime}_{E,x})=\mu(\varphi_{E})=1.

The proof of this theorem consists of two steps:

  • Given a pseudo-holomorphic disk in Symk(Σ)\operatorname{Sym}^{k}(\Sigma), we construct a corresponding disk in Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime}) by gluing spheres;

  • By a Gromov compactness argument, we show that all Maslov index 11 pseudo-holomorphic disks in Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime}) can be constructed in this way.

These two steps will be addressed in the next two subsections respectively.

5.3. Gluing

Let uu be a pseudo-holomorphic disk in jE,φE(c,c)\mathcal{M}_{j_{E},\varphi_{E}}(c,c^{\prime}). Then, uu does not intersect σ2\sigma_{2}, and therefore x×ux\times u defines a trajectory in Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime}) between x×{c}x\times\{c\} and x×{c}x\times\{c^{\prime}\}, which is Jt(T)J^{\prime}_{t}(T)-holomorphic. Moreover, for any zEE(αα)z_{E}\in E\setminus(\alpha\cup\alpha^{\prime}), nzE(u)=nzEE(φE)n^{\prime}_{z_{E}}(u)=n^{E}_{z_{E}}(\varphi_{E}), so x×uJt(T),φE,x(x×{c},x×{c})x\times u\in\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{E,x}}(x\times\{c\},x\times\{c^{\prime}\}).

Let uu be a pseudo-holomorphic disk in Jt,φ(x,y)\mathcal{M}_{J_{t},\varphi}(x,y). When nσ1(φ)=0n_{\sigma_{1}}(\varphi)=0, we can construct u:=u×{c}u^{\prime}:=u\times\{c\}, and as before it lives in Jt(T),φc(x×{c},y×{c})\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{c}}(x\times\{c\},y\times\{c\}).

However when n:=nσ1(φ)0n:=n_{\sigma_{1}}(\varphi)\neq 0, we need to glue nn spheres to the disk uu to construct a disk in Symk+1(Σ)\operatorname{Sym}^{k+1}(\Sigma^{\prime}). We follow the construction of [OS04], which was done in the case k=gk=g, but still works in this more general case. We will only give an outline of the proof without going into the more technical details, which are exactly the same as in [OS04, Section 10.2 and 10.3].

Suppose uu meets W={σ1}×Symk1(Σ)W=\{\sigma_{1}\}\times\operatorname{Sym}^{k-1}(\Sigma) transversally in nn distinct points q1,,qnq_{1},...,q_{n}. We identify ×[0,1]\mathbb{R}\times[0,1] with D{±i}D\setminus\{\pm i\}, where DD denotes the unit disk in \mathbb{C}. Then, uu extends continuously to DD by setting u(i)=x,u(i)=yu(-i)=x,u(i)=y.

We fix constants 0<r1<R10<r_{1}<R_{1} such that u(D)(Br1(σ1)×Symk1(ΣBr1(σ1)))Br1(σ1)×Symk1(ΣBR1(σ1))u(D)\cap\left(B_{r_{1}}(\sigma_{1})\times\operatorname{Sym}^{k-1}(\Sigma-B_{r_{1}}(\sigma_{1}))\right)\subset B_{r_{1}}(\sigma_{1})\times\operatorname{Sym}^{k-1}(\Sigma-B_{R_{1}}(\sigma_{1})).

There exists ϵ>0\epsilon>0 such that for every 1in1\leqslant i\leqslant n, Bϵ(qi)B_{\epsilon}(q_{i}) is mapped by uu into this subset.

We fix conformal identifications Br1(σ1)σ1[0,)×S1B_{r_{1}}(\sigma_{1})-\sigma_{1}\cong[0,\infty)\times S^{1}, and Bϵ(qi)[0,)×S1B_{\epsilon}(q_{i})\cong[0,\infty)\times S^{1}.

We will use Sobolev spaces with weight function eδτ1e^{\delta\tau_{1}}, where:

  • δ\delta is a positive constant;

  • τ1:D{q1,,qn}[0,)\tau_{1}:D-\{q_{1},...,q_{n}\}\to[0,\infty) is supported inside the Bϵ(qi)B_{\epsilon}(q_{i});

  • τ1(s,φ)=s\tau_{1}(s,\varphi)=s for s1s\geqslant 1 in each Bϵ(qi)[0,)×S1B_{\epsilon}(q_{i})\cong[0,\infty)\times S^{1}.

Then, for each ii there exists wiSymk1(Σ)w_{i}\in\operatorname{Sym}^{k-1}(\Sigma), and (ti,θi)×S1(t_{i},\theta_{i})\in\mathbb{R}\times S^{1} such that the restriction of uu to Bϵ(qi){qi}[0,)×S1B_{\epsilon}(q_{i})-\{q_{i}\}\cong[0,\infty)\times S^{1} differs by a L1,δpL^{p}_{1,\delta} map from the smooth map

ati,θi,wi:[0,)×S1Symk1(Σ)×[0,)×S1Symk(Σ)a_{t_{i},\theta_{i},w_{i}}:[0,\infty)\times S^{1}\to\operatorname{Sym}^{k-1}(\Sigma)\times[0,\infty)\times S^{1}\subset\operatorname{Sym}^{k}(\Sigma)

defined by

ati,θi,wi(s,φ)=(wi,s+ti,φ+θi)a_{t_{i},\theta_{i},w_{i}}(s,\varphi)=(w_{i},s+t_{i},\varphi+\theta_{i})

where we cut-off s+tis+t_{i} if it is negative ([OS04, Section 10.2]).

Given T>0T>0, we define X1(T):=τ11([0,T])X_{1}(T):=\tau_{1}^{-1}([0,T]) and X1()=D{q1,,qn}X_{1}(\infty)=D-\{q_{1},...,q_{n}\}.

Let h:[0,1]h:\mathbb{R}\to[0,1] be a smooth, increasing function such that h(t)=0h(t)=0 for t<0t<0 and h(t)=1h(t)=1 for t>1t>1.

We can define a map u~T:X1()Symk(Σ)\widetilde{u}_{T}:X_{1}(\infty)\to\operatorname{Sym}^{k}(\Sigma) which agrees with uu away from the Bϵ(qi)B_{\epsilon}(q_{i}), and defined by

u~T(s,φ)=h(sT)ati,θi,wi(s,φ)+(1h(sT))u(s,φ)\widetilde{u}_{T}(s,\varphi)=h(s-T)a_{t_{i},\theta_{i},w_{i}}(s,\varphi)+(1-h(s-T))u(s,\varphi)

over Bϵ(qi){qi}[0,)×S1B_{\epsilon}(q_{i})\setminus\{q_{i}\}\cong[0,\infty)\times S^{1}, and extends smoothly over qiq_{i} (where the convex combination is to be interpreted using the exponential map).

We also fix a constant δ0>0\delta_{0}>0, and define τ0:×[0,1]D\tau_{0}:\mathbb{R}\times[0,1]\cong D\to\mathbb{R} supported away from the Bϵ(qi)B_{\epsilon}(q_{i}), and such that τ0(s,t)=|s|\tau_{0}(s,t)=|s| for sufficiently large ss333In [OS04], they use (s,t)[0,1]×(s,t)\in[0,1]\times\mathbb{R} and require that τ0\tau_{0} equals to |t||t| for sufficiently large tt, but we follow the more standard convention that (s,t)×[0,1](s,t)\in\mathbb{R}\times[0,1]..

Then, according to [OS04, Lemma 10.6], for the Sobolev norm with weight eδ0τ0+δτ1e^{\delta_{0}\tau_{0}+\delta\tau_{1}}, there are constants κ>0\kappa>0, S0>0S_{0}>0 and C>0C>0 such that for all S>S0S>S_{0}

¯Jtu~SLδ,δ0p(Λ0,1u~S)CeκS||\bar{\partial}_{J_{t}}\widetilde{u}_{S}||_{L^{p}_{\delta,\delta_{0}}(\Lambda^{0,1}\widetilde{u}_{S})}\leqslant Ce^{-\kappa S}

Now we consider spheres in Sym2(E)\operatorname{Sym}^{2}(E). Let vv be a holomorphic map from S2S^{2} to Symk1(Σ)×Sym2(E)\operatorname{Sym}^{k-1}(\Sigma)\times\operatorname{Sym}^{2}(E), constant on the first factor, and such that nσ2([v])=1n_{\sigma_{2}}([v])=1. Denote by (S2Symk1(Σ)×Sym2(E))\mathcal{M}(S^{2}\to\operatorname{Sym}^{k-1}(\Sigma)\times\operatorname{Sym}^{2}(E)) the moduli space of such maps, modulo holomorphic reparametrization. According to [OS04, Lemma 10.7], we have:

Lemma 44.

For such a map vv, there exists a unique pair (w,c)(w,c) in Symk1(Σ)×E\operatorname{Sym}^{k-1}(\Sigma)\times E such that (w,{c,σ2})Im(v)(w,\{c,\sigma_{2}\})\in\operatorname{Im}(v).444In [OS04], they use the notation (w,y)(w,y) instead of (w,c)(w,c). We use yy to denote an intersection point between SymL¯\operatorname{Sym}\underline{L} and SymK¯\operatorname{Sym}\underline{K} so we use cc here.

The map [v](w,c)[v]\mapsto(w,c) is then a one-to-one correspondence between (S2Symk1(Σ)×Sym2(E))\mathcal{M}(S^{2}\to\operatorname{Sym}^{k-1}(\Sigma)\times\operatorname{Sym}^{2}(E)) and Symk1(Σ)×E\operatorname{Sym}^{k-1}(\Sigma)\times E.

We fix vv as above, and normalize it so that v(0)=w×{c,σ2}v(0)=w\times\{c,\sigma_{2}\} (where we view S2S^{2} as {}\mathbb{C}\cup\{\infty\}).

We will only be interested in the case that cααc\in\alpha\cap\alpha^{\prime}. The intuitive reason is that we are going to glue u×{c}:DSymk(Σ)×Eu\times\{c\}:D\to\operatorname{Sym}^{k}(\Sigma)\times E and nn many v:S2Symk1(Σ)×Sym2(E)v:S^{2}\to\operatorname{Sym}^{k-1}(\Sigma)\times\operatorname{Sym}^{2}(E) together (one vv for each qiq_{i}), so cc has to be an intersection point between α\alpha and α\alpha^{\prime} for u×{c}u\times\{c\} to satisfy the Lagrangian boundary conditions (cf. the degeneration (3) where the Gromov limit lives). Therefore, we assume from now on that cσ2c\neq\sigma_{2}.

We identify a neighborhood of v(0)v(0) with

Symk1(Σ)×Br2(σ2)×(EBR2(σ2))Symk1(Σ)×Sym2(E)\operatorname{Sym}^{k-1}(\Sigma)\times B_{r_{2}}(\sigma_{2})\times(E-B_{R_{2}}(\sigma_{2}))\subset\operatorname{Sym}^{k-1}(\Sigma)\times\operatorname{Sym}^{2}(E)

for some 0<r2<R20<r_{2}<R_{2}.

Fix ϵ>0\epsilon>0 such that Bϵ(0)B_{\epsilon}(0) is sent by vv to this neighborhood. Fix conformal identifications [0,)×S1Bϵ(0){0}[0,\infty)\times S^{1}\cong B_{\epsilon}(0)-\{0\} and [0,)×S1Br2(σ2){σ2}[0,\infty)\times S^{1}\cong B_{r_{2}}(\sigma_{2})-\{\sigma_{2}\}.

Then, there are unique wSymk1(Σ)w\in\operatorname{Sym}^{k-1}(\Sigma), cEc\in E, (t,θ)[0,)×S1(t,\theta)\in[0,\infty)\times S^{1} such that vv restricted to [0,)×S1Bϵ(0){0}[0,\infty)\times S^{1}\cong B_{\epsilon}(0)-\{0\} differs by a L1,δpL^{p}_{1,\delta} map from the map

b(t,θ,w,y)(s,φ)=(w,s+t,θ+φ,c)b_{(t,\theta,w,y)}(s,\varphi)=(w,s+t,\theta+\varphi,c)

(where L1,δpL^{p}_{1,\delta} is defined with a weight function eδτ2e^{\delta}\tau_{2} with τ2:S2{0}+\tau_{2}:S^{2}-\{0\}\to\mathbb{R}^{+} defined in a similar fashion as τ1\tau_{1}).

For S>0S>0, let S2(S):=τ21([0,S])S^{2}(S):=\tau_{2}^{-1}([0,S]). We define a map555In [OS04], the domain of vSv_{S} is S2{}S^{2}-\{\infty\} which we believe is a typo.

vS:S2{0}Symk1(Σ)×Sym2(E{σ2})v_{S}:S^{2}-\{0\}\to\operatorname{Sym}^{k-1}(\Sigma)\times\operatorname{Sym}^{2}(E-\{\sigma_{2}\})

which agrees with vv over S2(S)S^{2}(S), and such that over [0,)×S1Bϵ(0){0}[0,\infty)\times S^{1}\cong B_{\epsilon}(0)-\{0\},

vS(s,φ)=h(sS)b(t,θ,w,c)(s,φ)+(1h(sS))v(s,φ)v_{S}(s,\varphi)=h(s-S)b_{(t,\theta,w,c)}(s,\varphi)+(1-h(s-S))v(s,\varphi)

In [OS04, Definition 10.8], the authors define a normalization condition on holomorphic spheres called being ’centered’. They show that the moduli space of centered maps cent(S2Symk1(Σ)×Sym2(E))\mathcal{M}^{\text{cent}}(S^{2}\to\operatorname{Sym}^{k-1}(\Sigma)\times\operatorname{Sym}^{2}(E)) is diffeomorphic to Symk1(Σ)××S1×E\operatorname{Sym}^{k-1}(\Sigma)\times\mathbb{R}\times S^{1}\times E through the assignment v(w,t,θ,c)v\mapsto(w,t,\theta,c).

Denote by v(w,t,θ,c)v_{(w,t,\theta,c)} the pre-image of (w,t,θ,c)(w,t,\theta,c) by this diffeomorphism.

Using the conformal identifications Br1(σ1)σ1[0,)×S1B_{r_{1}}(\sigma_{1})-\sigma_{1}\cong[0,\infty)\times S^{1} and Br2(σ2)σ2[0,)×S1B_{r_{2}}(\sigma_{2})-\sigma_{2}\cong[0,\infty)\times S^{1}, one can think of Σ(T)\Sigma^{\prime}(T) as the union of Σ(2T)\Sigma(2T) and E(2T)E(2T) modulo the identification of the cylinders [0,2T]×S1Σ(2T)[0,2T]\times S^{1}\subset\Sigma(2T) and [0,2T]×S1E(2T)[0,2T]\times S^{1}\subset E(2T) via the involution (t,θ)(2Tt,θ)(t,\theta)\sim(2T-t,\theta).

Let X2(T):=i=1nS2(T)iX_{2}(T):=\bigsqcup\limits_{i=1}^{n}S^{2}(T)_{i} and X1TX2X_{1}\cup_{T}X_{2} be the union of X1(T)X_{1}(T) and X2(T)X_{2}(T) glued at their common boundary. We have that X1TX2D2X_{1}\cup_{T}X_{2}\cong D^{2}.

There exists some constant t>0t>0 such that for any real numbers SS and TT such that 0<S<Tt0<S<T-t, given the pseudo-holomorphic disk uJs,φ(x,y)u\in\mathcal{M}_{J_{s},\varphi}(x,y) we fixed earlier, and the intersection point cααEc\in\alpha\cap\alpha^{\prime}\subset E, one can define a map

γ^c(u,S,T):DX1TX2Symk+1(Σ#TE)\hat{\gamma}_{c}(u,S,T):D\cong X_{1}\cup_{T}X_{2}\to\operatorname{Sym}^{k+1}(\Sigma\#_{T}E)

which agrees with u~S×{c}\widetilde{u}_{S}\times\{c\} over X1(T)X_{1}(T) and with v(wi,ti,θi,c),Sv_{(w_{i},-t_{i},\theta_{i},c),S} on S2(T)iX2(T)S^{2}(T)_{i}\subset X_{2}(T).

Following [OS04, Lemma 10.9], if SS is sufficiently large, then for large TT this map is smooth, and for an appropriate Sobolev norm there are some positive constants CC and aa such that

¯Jt(T)γ^c(u,S,T)CeaS||\bar{\partial}_{J^{\prime}_{t}(T)}\hat{\gamma}_{c}(u,S,T)||\leqslant Ce^{-aS}

One can show ([OS04, Proposition 10.12]) that when TT is sufficiently large, the linearization of ¯Jt(T)\bar{\partial}_{J^{\prime}_{t}(T)} for the spliced map from X1TX2X_{1}\cup_{T}X_{2} admits a right inverse whose norm is bounded independent of TT.

Then, applying the inverse theorem function ([OS04, Proposition 10.14]), there is an ϵ>0\epsilon>0 such that for sufficiently large TT, there is a unique holomorphic curve γc(u)\gamma_{c}(u) which lies in an ϵ\epsilon-neighborhood of γ^c(u,S,T)\hat{\gamma}_{c}(u,S,T) (measured in the appropriate Sobolev norm).

This γc(u)\gamma_{c}(u) lives in Jt(T),φc(x×{c},y×{c})\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{c}}(x\times\{c\},y\times\{c\}).

5.4. Gromov compactness

Now we need to show that every map in Jt(T),φc(x×{c},y×{c})\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{c}}(x\times\{c\},y\times\{c\}) and Jt(T),φE,x(x×{c},x×{c})\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{E,x}}(x\times\{c\},x\times\{c^{\prime}\}) can be attained by the construction of the previous section. Once again, the argument is similar to the one in [OS04].

Let x,ySymk+1(Σ)x^{\prime},y^{\prime}\in\operatorname{Sym}^{k+1}(\Sigma^{\prime}) be two critical points of the action functional (i.e. intersection points of the Lagrangians Sym(L¯α)Sym(K¯α)\operatorname{Sym}(\underline{L}\cup\alpha)\cap\operatorname{Sym}(\underline{K}\cup\alpha^{\prime})). Let φ\varphi^{\prime} be a Maslov index 11 class in π2(x,y)\pi_{2}(x^{\prime},y^{\prime}).

Then, according to [OS04, Proposition 10.15], any sequence uTJt(T),φ(x,y)u_{T}\in\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}}(x^{\prime},y^{\prime}) with TT going to infinity has, up to passing to a subsequence, a Gromov limit uu_{\infty} mapping to

(3) Symk+1(ΣE)=i=0k+1Symk+1i(Σ)×Symi(E).\displaystyle\operatorname{Sym}^{k+1}(\Sigma\vee E)=\bigcup\limits_{i=0}^{k+1}\operatorname{Sym}^{k+1-i}(\Sigma)\times\operatorname{Sym}^{i}(E).

We can think of the wedge sum ΣE\Sigma\vee E as the degeneration of the connected sum Σ#TE\Sigma\#_{T}E when the neck length TT goes to infinity. The limit uu_{\infty} is analyzed in Lemma 48 and 46 below.

Remark 45.

An alternative way to think about this Gromov compactness is to consider the relative Hilbert scheme Hilbk+1(π)\operatorname{Hilb}^{k+1}(\pi) of a Lefschetz fibration π:ED\pi:E\to D over a disk DD, where generic fibres are smooth and the singular fibre is ΣE\Sigma\vee E. The relative Hilbert scheme is smooth [Per07, Proposition 3.7] and one can equip Hilbk+1(π)\operatorname{Hilb}^{k+1}(\pi) with a one-parameter family of almost complex structures such that the projection to DD are pseudo-holomorphic, they are fibrewise VV-nearly symmetric, and they agree with Jt(T)J^{\prime}_{t}(T) over some fibres such that TT\to\infty corresponds to degenerating to the central fibre. The central fibre of Hilbk+1(π)\operatorname{Hilb}^{k+1}(\pi) has a canonical ‘cycle map’ to Symk+1(ΣE)\operatorname{Sym}^{k+1}(\Sigma\vee E) (cf. [Per07, Section 1.5.1]) and uu_{\infty} is the same as applying Gromov compactness inside Hilbk+1(π)\operatorname{Hilb}^{k+1}(\pi) and then applying the cycle map to Symk+1(ΣE)\operatorname{Sym}^{k+1}(\Sigma\vee E).

Lemma 46 (cf. Proposition 10.16 of [OS04]).

If (x,y)(x^{\prime},y^{\prime}) is of the form (x×{c},y×{c})(x\times\{c\},y\times\{c\}), then uu_{\infty} consists of a main component of the form u×{c}u\times\{c\}, where uu is a Maslov index 1 trajectory from xx to yy in Symk(Σ)\operatorname{Sym}^{k}(\Sigma), together with possibly some sphere components of the form {w}×v\{w\}\times v, where wSymk1(Σ)w\in\operatorname{Sym}^{k-1}(\Sigma) and vv is a holomorphic sphere in Sym2(E)\operatorname{Sym}^{2}(E) with Chern number 22 (i.e. a sphere in the ruling).

Proof.

Since the uTu_{T} are Floer trajectories between xx^{\prime} and yy^{\prime}, uu_{\infty} consists of a (possibly) broken Floer trajectory between xx^{\prime} and yy^{\prime}, as well as sphere bubbles and disk bubbles. Let uiu_{\infty}^{i} be the components of uu_{\infty} in Symk+1i(Σ)×Symi(E)\operatorname{Sym}^{k+1-i}(\Sigma)\times\operatorname{Sym}^{i}(E). By projection to factors, we can write it as ui=(uΣ,i,uE,i)u_{\infty}^{i}=(u_{\infty}^{\Sigma,i},u_{\infty}^{E,i}).

Let DΣ,i=Symki(Σ)×{σ1}Symk+1i(Σ)D_{\Sigma,i}=\operatorname{Sym}^{k-i}(\Sigma)\times\{\sigma_{1}\}\subset\operatorname{Sym}^{k+1-i}(\Sigma) and DE,i=Symi1(E)×{σ2}Symi(E)D_{E,i}=\operatorname{Sym}^{i-1}(E)\times\{\sigma_{2}\}\subset\operatorname{Sym}^{i}(E). We define the adjusted Maslov index μ~(ui)\tilde{\mu}(u_{\infty}^{i}) of uiu_{\infty}^{i} relative to DΣ,i×Symi(E)+Symk+1i(Σ)×DE,iD_{\Sigma,i}\times\operatorname{Sym}^{i}(E)+\operatorname{Sym}^{k+1-i}(\Sigma)\times D_{E,i} as the Maslov index of uu with respect to the log canonical line bundle with a simple pole along DΣ,i×Symi(E)+Symk+1i(Σ)×DE,iD_{\Sigma,i}\times\operatorname{Sym}^{i}(E)+\operatorname{Sym}^{k+1-i}(\Sigma)\times D_{E,i}. In other words, the adjusted Maslov index of uiu_{\infty}^{i} is its usual Maslov index viewed as a map to Symk+1i(Σ)×Symi(E)\operatorname{Sym}^{k+1-i}(\Sigma)\times\operatorname{Sym}^{i}(E) subtracted by 2[ui](DΣ,i×Symi(E)+Symk+1i(Σ)×DE,i)2[u_{\infty}^{i}]\cdot(D_{\Sigma,i}\times\operatorname{Sym}^{i}(E)+\operatorname{Sym}^{k+1-i}(\Sigma)\times D_{E,i}). The additivity of Maslov index under passing to a Gromov limit implies that the sum of the adjusted Maslov indices of the components of uu_{\infty} equals the Maslov index of φ\varphi^{\prime}, which is 11. This is because DΣ,i×Symi(E)+Symk+1i(Σ)×DE,iD_{\Sigma,i}\times\operatorname{Sym}^{i}(E)+\operatorname{Sym}^{k+1-i}(\Sigma)\times D_{E,i} is precisely the locus where Symk+1i(Σ)×Symi(E)\operatorname{Sym}^{k+1-i}(\Sigma)\times\operatorname{Sym}^{i}(E) intersects other irreducible components of Symk+1(ΣE)\operatorname{Sym}^{k+1}(\Sigma\vee E).

Since Sym(L¯α)\operatorname{Sym}(\underline{L}\cup\alpha) and Sym(K¯α)\operatorname{Sym}(\underline{K}\cup\alpha^{\prime}) are contained in Symk(Σ)×ESymk+1(ΣE)\operatorname{Sym}^{k}(\Sigma)\times E\subset\operatorname{Sym}^{k+1}(\Sigma\vee E), the broken Floer trajectory and disk bubbles are contained in Symk(Σ)×E\operatorname{Sym}^{k}(\Sigma)\times E. We denote the spherical components of u1u_{\infty}^{1} by u,S1u_{\infty,S}^{1} and the other components of u1u_{\infty}^{1} by u,D1u_{\infty,D}^{1}.

Now, we analyze the adjusted Maslov indices of the spherical components of uu_{\infty}. Recall from [BT01, Theorem 9.2] that the rank of π2(Symj(Σ))H2(Symj(Σ))\pi_{2}(\operatorname{Sym}^{j}(\Sigma))\to H_{2}(\operatorname{Sym}^{j}(\Sigma)) is 11 when j2j\geq 2 or Σ\Sigma has genus 0, and 0 otherwise. Moreover, the Chern number of a spherical class uu is given by (j+1g)[u][Symj1(Σ)×{σ1}](j+1-g)[u]\cdot[\operatorname{Sym}^{j-1}(\Sigma)\times\{\sigma_{1}\}] (cf. [CGHM+21, Remark 4.18]). Therefore, its adjusted Maslov index relative to Symj1(Σ)×{σ1}\operatorname{Sym}^{j-1}(\Sigma)\times\{\sigma_{1}\} is given by

(4) 2c1[u]2[u][Symj1(Σ)×{σ1}]=2(jg)[u][Symj1(Σ)×{σ1}]\displaystyle 2c_{1}\cdot[u]-2[u]\cdot[\operatorname{Sym}^{j-1}(\Sigma)\times\{\sigma_{1}\}]=2(j-g)[u]\cdot[\operatorname{Sym}^{j-1}(\Sigma)\times\{\sigma_{1}\}]

The adjusted Maslov index μ~(ui)\tilde{\mu}(u_{\infty}^{i}) is the sum of the adjusted Maslov indices μ~(uΣ,i)+μ~(uE,i)\tilde{\mu}(u_{\infty}^{\Sigma,i})+\tilde{\mu}(u_{\infty}^{E,i}) where the two terms in the sum are relative to DΣ,i×Symi(E)D_{\Sigma,i}\times\operatorname{Sym}^{i}(E) and Symk+1i(Σ)×DE,i\operatorname{Sym}^{k+1-i}(\Sigma)\times D_{E,i}, respectively. For spherical components, they can be computed by the formula (4).

For i1i\neq 1, we define

NΣ,i:=[uΣ,i]DΣ,i,NE,i:=[uE,i]DE,iN_{\Sigma,i}:=[u_{\infty}^{\Sigma,i}]\cdot D_{\Sigma,i},\quad N_{E,i}:=[u_{\infty}^{E,i}]\cdot D_{E,i}

For i=1i=1, we define

NΣ,1:=[u,SΣ,1]DΣ,1,NE,1:=[u,SE,1]DE,1N_{\Sigma,1}:=[u_{\infty,S}^{\Sigma,1}]\cdot D_{\Sigma,1},\quad N_{E,1}:=[u_{\infty,S}^{E,1}]\cdot D_{E,1}

and

PΣ,1:=[u,DΣ,1]DΣ,1,PE,1:=[u,DE,1]DE,1P_{\Sigma,1}:=[u_{\infty,D}^{\Sigma,1}]\cdot D_{\Sigma,1},\quad P_{E,1}:=[u_{\infty,D}^{E,1}]\cdot D_{E,1}

The terms PΣ,1P_{\Sigma,1} and PE,1P_{E,1} make sense because the Lagrangian boundary condition splits as a product, and they are disjoint from the divisor DΣ,1D_{\Sigma,1}, DE,1D_{E,1}. Clearly, NE,0=NE,1=0N_{E,0}=N_{E,1}=0 and NΣ,k+1=0N_{\Sigma,k+1}=0 because the spherical class is trivial. Note that DΣ,i×Symi(E)Symk+1i(Σ)×Symi(E)D_{\Sigma,i}\times\operatorname{Sym}^{i}(E)\subset\operatorname{Sym}^{k+1-i}(\Sigma)\times\operatorname{Sym}^{i}(E) and Symki(Σ)×DE,i+1Symki(Σ)×Symi+1(E)\operatorname{Sym}^{k-i}(\Sigma)\times D_{E,i+1}\subset\operatorname{Sym}^{k-i}(\Sigma)\times\operatorname{Sym}^{i+1}(E) are precisely the locus where these two components of Symk+1(ΣE)\operatorname{Sym}^{k+1}(\Sigma\vee E) meet each other. Therefore, we have

(5) NΣ,i=NE,i+1\displaystyle N_{\Sigma,i}=N_{E,i+1}

for i2i\geq 2, and

NΣ,1+PΣ,1=NE,2,NΣ,0=NE,1+PE,1=PE,1N_{\Sigma,1}+P_{\Sigma,1}=N_{E,2},\quad N_{\Sigma,0}=N_{E,1}+P_{E,1}=P_{E,1}

Recall that α\alpha^{\prime} is a Hamiltonian push-off of α\alpha such that the Hamiltonian isotopy from α\alpha to α\alpha^{\prime} does not pass through σ2\sigma_{2}. Therefore, any Floer trajectory between two intersection points of α\alpha and α\alpha^{\prime} does not pass through σ2\sigma_{2} in EE. Also, there is no non-constant disk bubble in EE so PE,1=0P_{E,1}=0. Therefore, NΣ,0=0N_{\Sigma,0}=0 and uu_{\infty} does not intersect the component Symk+1(Σ)Symk+1(ΣE)\operatorname{Sym}^{k+1}(\Sigma)\subset\operatorname{Sym}^{k+1}(\Sigma\vee E) at all.

The total sum of the adjusted Maslov indices of the components of uu_{\infty} is given by

μ~(u1)+i=2k+1μ~(ui)\displaystyle\tilde{\mu}(u_{\infty}^{1})+\sum_{i=2}^{k+1}\tilde{\mu}(u_{\infty}^{i})
=\displaystyle= μ~(u,D1)+μ~(u,SΣ,1)+μ~(u,SE,1)+i=2k+1μ~(uΣ,i)+i=2k+1μ~(uE,i)\displaystyle\tilde{\mu}(u_{\infty,D}^{1})+\tilde{\mu}(u_{\infty,S}^{\Sigma,1})+\tilde{\mu}(u_{\infty,S}^{E,1})+\sum_{i=2}^{k+1}\tilde{\mu}(u_{\infty}^{\Sigma,i})+\sum_{i=2}^{k+1}\tilde{\mu}(u_{\infty}^{E,i})
=\displaystyle= (μ(u,D1)2PΣ,1)+2(kg)NΣ,1+0+i=2k+12(k+1ig)NΣ,i+i=2k+12(i1)NE,i\displaystyle(\mu(u_{\infty,D}^{1})-2P_{\Sigma,1})+2(k-g)N_{\Sigma,1}+0+\sum_{i=2}^{k+1}2(k+1-i-g)N_{\Sigma,i}+\sum_{i=2}^{k+1}2(i-1)N_{E,i}
=\displaystyle= μ(u,D1)2PΣ,1+2(kg)NΣ,1+i=2k2(k+1ig)NΣ,i+2NE,2+i=2k2iNE,i+1\displaystyle\mu(u_{\infty,D}^{1})-2P_{\Sigma,1}+2(k-g)N_{\Sigma,1}+\sum_{i=2}^{k}2(k+1-i-g)N_{\Sigma,i}+2N_{E,2}+\sum_{i=2}^{k}2iN_{E,i+1}
=\displaystyle= μ(u,D1)2PΣ,1+2(kg)NΣ,1+2(NΣ,1+PΣ,1)+i=2k2(k+1g)NΣ,i\displaystyle\mu(u_{\infty,D}^{1})-2P_{\Sigma,1}+2(k-g)N_{\Sigma,1}+2(N_{\Sigma,1}+P_{\Sigma,1})+\sum_{i=2}^{k}2(k+1-g)N_{\Sigma,i}
=\displaystyle= μ(u,D1)+2(kg+1)NΣ,1+i=2k2(k+1g)NΣ,i\displaystyle\mu(u_{\infty,D}^{1})+2(k-g+1)N_{\Sigma,1}+\sum_{i=2}^{k}2(k+1-g)N_{\Sigma,i}

As we said earlier, this total sum has to be 11. By regularity, each Floer trajectory component of u,D1u_{\infty,D}^{1} contributes at least 11 to the Maslov index. Any non-constant disk bubble in u,D1u_{\infty,D}^{1} contributes at least 22 to the Maslov index. Since kgk\geq g, we have (kg+1)>0(k-g+1)>0. Therefore, the sum is 11 only if NΣ,i=0N_{\Sigma,i}=0 for all ii, and u,D1u_{\infty,D}^{1} consists of a single component. It implies that PΣ,1=NE,2P_{\Sigma,1}=N_{E,2}.

By genericity, we can assume that the component u,D1u_{\infty,D}^{1} intersects DΣ,1×ED_{\Sigma,1}\times E transversely. Therefore, u2u_{\infty}^{2} intersects Symk1(Σ)×DE,1\operatorname{Sym}^{k-1}(\Sigma)\times D_{E,1} transversely. Note that every component of u2u_{\infty}^{2} projects to a constant in Symk1(Σ)\operatorname{Sym}^{k-1}(\Sigma) because NΣ,2=0N_{\Sigma,2}=0. Since the domain has genus 0, the bubbling is modeled on a tree and hence no component of u2u_{\infty}^{2} can be a multiple cover of an underlying holomorphic sphere. It implies that the sphere components of uu_{\infty} are of the form {w}×v\{w\}\times v, where wSymk1(Σ)w\in\operatorname{Sym}^{k-1}(\Sigma) and vv is a holomorphic sphere in Sym2(E)\operatorname{Sym}^{2}(E) with Chern number 22. The Floer trajectory component of uu_{\infty} is u1u_{\infty}^{1}, which goes between x×{c}x\times\{c\} and y×{c})y\times\{c\}). Therefore, it is of the form u×{c}u\times\{c\}, where uu is a Maslov index 1 trajectory from xx to yy in Symk(Σ)\operatorname{Sym}^{k}(\Sigma).

Remark 47.

The bubbling analysis in the proof of Lemma 46 provides the details of the phrase ‘by a dimension count’ in the proof of [OS04, Proposition 10.16] and at the same time confirms that we can generalize it to all kgk\geq g.

Lemma 48.

If (x,y)(x^{\prime},y^{\prime}) is of the form (x×{c},x×{c})(x\times\{c\},x\times\{c^{\prime}\}), then uu_{\infty} consists of a single component and it is of a product form {x}×u\{x\}\times u where uu is a Maslov index 1 trajectory from cc to cc^{\prime} in EE.

Proof.

The proof is the same as Lemma 46. The only difference is that when (x,y)(x^{\prime},y^{\prime}) is of the form (x×{c},x×{c})(x\times\{c\},x\times\{c^{\prime}\}), the Floer trajectory u,D1u_{\infty,D}^{1} is of the form {x}×u\{x\}\times u where uu is a Maslov index 1 trajectory from cc to cc^{\prime}. Therefore, we have PΣ,1=0P_{\Sigma,1}=0 and hence there is no sphere bubbles.

Now we show that for sufficiently large TT, any map in Jt(T),φ(x×{c},y×{c})\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}}(x\times\{c\},y\times\{c\}) is attained by the construction of the previous section.

We proceed by contradiction: suppose that there is a sequence (Tm)(T_{m}) going to infinity, and a sequence of disks uTmJs(Tm),φ(x×{c},y×{c})u_{T_{m}}\in\mathcal{M}_{J^{\prime}_{s}(T_{m}),\varphi^{\prime}}(x\times\{c\},y\times\{c\}) that are not attained by the gluing construction.

By what precedes, we can extract a subsequence converging to a bubbletree uu_{\infty}, consisting of a disk u×{c}u\times\{c\} and nn spheres {wi}×vi\{w_{i}\}\times v_{i}.

Then, the authors of [OS04] show that for sufficiently large mm, uTmu_{T_{m}} is in an ϵ\epsilon-neighborhood of the nearly holomorphic map γ^c(u,S,Tm)\hat{\gamma}_{c}(u,S,T_{m}) for some 0<S<Tmt0<S<T_{m}-t (for the suitable Sobolev distance).

But we showed in the previous section that there was a single Jt(T)J^{\prime}_{t}(T) holomorphic curve in such a neighborhood, namely the curve γc(u,Tm)\gamma_{c}(u,T_{m}). Therefore, for large mm, uTm=γc(u,Tm)u_{T_{m}}=\gamma_{c}(u,T_{m}), which contradicts our assumption.

Hence by contradiction for large TT, we have Js,φ(x,y)Jt(T),φc(x×{c},y×{c})\mathcal{M}_{J_{s},\varphi}(x,y)\simeq\mathcal{M}_{J^{\prime}_{t}(T),\varphi^{\prime}_{c}}(x\times\{c\},y\times\{c\}).

This concludes the proof of Theorem 42.

References

  • [Ban97] Augustin Banyaga. The structure of classical diffeomorphism groups. 1997.
  • [BRa14] Marcel Bökstedt and Nuno M. Romão. On the curvature of vortex moduli spaces. Math. Z., 277(1-2):549–573, 2014.
  • [BT01] Aaron Bertram and Michael Thaddeus. On the quantum cohomology of a symmetric product of an algebraic curve. Duke Math. J., 108(2):329–362, 2001.
  • [Buh22] Lev Buhovsky. On two remarkable groups of area-preserving homeomorphisms. arXiv:2204.08020, 2022.
  • [CGHM+21] Daniel Cristofaro-Gardiner, Vincent Humilière, Cheuk Yu Mak, Sobhan Seyfaddini, and Ivan Smith. Quantitative heegaard floer cohomology and the calabi invariant. arXiv:2105.11026, 2021.
  • [CGHM+22] Dan Cristofaro-Gardiner, Vincent Humilière, Cheuk Yu Mak, Sobhan Seyfaddini, and Ivan Smith. Subleading asymptotics of link spectral invariants and homeomorphism groups of surfaces. arXiv:2206.10749, June 2022.
  • [CGHR15] Daniel Cristofaro-Gardiner, Michael Hutchings, and Vinicius Gripp Barros Ramos. The asymptotics of ECH capacities. Invent. Math., 199(1):187–214, 2015.
  • [CGHS20] Daniel Cristofaro-Gardiner, Vincent Humilière, and Sobhan Seyfaddini. Proof of the simplicity conjecture. arXiv:2001.01792, 2020.
  • [Che21] Guanheng Chen. Closed-open morphisms on periodic Floer homology. arXiv:2111.11891, 2021.
  • [Che22] Guanheng Chen. On PFH and HF spectral invariants. arXiv:2209.11071, 2022.
  • [EP] Michael Entov and Leonid Polterovich. Calabi quasimorphism and quantum homology. Intern. Math. Res. Notices, pages 1635–1676.
  • [Fat80] A. Fathi. Structure of the group of homeomorphisms preserving a good measure on a compact manifold. Ann. Sci. École Norm. Sup. (4), 13(1):45–93, 1980.
  • [Hum17] Vincent Humilière. Géométrie symplectique C0C^{0} et sélecteurs d’action, 2017.
  • [Hut11] Michael Hutchings. Quantitative embedded contact homology. J. Differential Geom., 88(2):231–266, 2011.
  • [Lip06] Robert Lipshitz. A cylindrical reformulation of Heegaard Floer homology. Geometry & Topology, 10(2):955–1096, aug 2006.
  • [LZ18] Rémi Leclercq and Frol Zapolsky. Spectral invariants for monotone Lagrangians. J. Topol. Anal., 10(3):627–700, 2018.
  • [MS21] Cheuk Yu Mak and Ivan Smith. Non-displaceable Lagrangian links in four-manifolds. Geom. Funct. Anal., 31(2):438–481, 2021.
  • [OM07] Yong-Geun Oh and Stefan Müller. The group of Hamiltonian homeomorphisms and C0C^{0}-symplectic topology. J. Symplectic Geom., 5(2):167–219, 2007.
  • [OS04] Peter Ozsváth and Zoltán Szabó. Holomorphic disks and topological invariants for closed three-manifolds. Ann. of Math. (2), 159(3):1027–1158, 2004.
  • [OS08] Peter Ozsváth and Zoltán Szabó. Holomorphic disks, link invariants and the multi-variable alexander polynomial. Algebraic & Geometric Topology, 8(2):615–692, may 2008.
  • [Per07] Tim Perutz. Lagrangian matching invariants for fibred four-manifolds. I. Geom. Topol., 11:759–828, 2007.
  • [PS21] Leonid Polterovich and Egor Shelukhin. Lagrangian configurations and Hamiltonian maps. arXiv:2102.06118, 2021.
  • [Sey13] Sobhan Seyfaddini. C0C^{0}-limits of Hamiltonian paths and the Oh-Schwarz spectral invariants. Int. Math. Res. Not. IMRN, (21):4920–4960, 2013.