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

\DeclareDelimFormat

[bib,biblist]nametitledelim,

A link invariant from higher-dimensional Heegaard Floer homology

Tianyu Yuan Beijing International Center for Mathematical Research, Peking University, Beijing 100871, China [email protected]
Abstract.

We define a higher-dimensional analogue of symplectic Khovanov homology. Consider the standard Lefschetz fibration p:WDp\colon W\to D\subset\mathbb{C} of a 2n2n-dimensional Milnor fiber of the A2κ1A_{2\kappa-1} singularity. We represent a link by a κ\kappa-strand braid, which is expressed as an element hh of the symplectic mapping class group Symp(W,W)\mathrm{Symp}(W,\partial W). We then apply the higher-dimensional Heegaard Floer homology machinery to the pair (𝒂,h(𝒂))(\boldsymbol{a},h(\boldsymbol{a})), where 𝒂\boldsymbol{a} is a collection of κ\kappa unstable manifolds of WW which are Lagrangian spheres. We prove its invariance under arc slides and Markov stabilizations, which shows that it is a link invariant. This work constitutes part of the author’s PhD thesis.

Key words and phrases:
Higher-dimensional Heegaard Floer homology, Khovanov homology
2010 Mathematics Subject Classification:
Primary 53D40; Secondary 57M27.

1. Introduction

Many powerful Floer-theoretic invariants of knots and links have emerged over the past two decades. These include Heegaard Floer homology [OS04] and knot Floer homology in dimension 1; symplectic Khovanov homology [SS06] and knot contact homology [EENS13] in dimension 2. Here when we say “dimension nn”, we are taking the ambient symplectic manifold to be 2n2n-dimensional and the Lagrangian submanifolds (if we are talking about Lagrangian intersection Floer thoeries) to be nn-dimensional. In [Man07] Manolescu also used quiver varieties to define a higher-dimensional analogue of 𝔰𝔩(n)\mathfrak{sl}(n)-homologies.

Along similar lines, the aim of this paper is to construct a link invariant using higher-dimensional Heegaard Floer homology, which is defined in [CHT20] as a higher-dimensional analogue of Heegaard Floer homology in a cylindrical setting.

More specifically, in dimension 1, Lipschitz [Lip06] proved the equivalence between Ozsváth and Szabó’s Heegaard Floer homology [OS04] and its cylindrical analogue. In dimension 2, Mak and Smith [MS19] established the equivalence of symplectic Khovanov homology and its cylindrical interpretation. Colin, Honda, and Tian [CHT20] then defined a higher-dimensional analogue of cylindrical Heegaard Floer homology, which helps place the cylindrical symplectic Khovanov homology in a more general framework. In this paper, the ambient manifold is a 2n2n-dimensional Milnor fiber of the A2κ1A_{2\kappa-1}-singularity p:WDp:W\to D\subset\mathbb{C}, extending the case of n=1n=1 considered in [CHT20]. Given a link, we consider its κ\kappa-strand braid representation σ\sigma, which corresponds to an element hh of the symplectic mapping class group Symp(W,W)\mathrm{Symp}(W,\partial W). There is a natural collection of κ\kappa Lagrangian spheres 𝒂\boldsymbol{a} by the matching cycle construction between pairs of critical points of pp. We then apply the higher-dimensional Heegaard Floer homology machinery to the pair (𝒂,h(𝒂))(\boldsymbol{a},h(\boldsymbol{a})) to define the link invariant and denote the homology group by Kh(σ^)Kh^{\sharp}(\widehat{\sigma}).

Though cylindrical versions of Heegaard Floer theories are more convenient for visualizing pseudoholomorphic curves, the original theories defined in the symmetric products Symκ(M)\mathrm{Sym}^{\kappa}(M) have their advantages: In dimension 1, Perutz [Per08] proved that Lagrangians in Symκ(Σ)\mathrm{Sym}^{\kappa}(\Sigma) related by a handle slide are in fact Hamiltonian isotopic for some specific symplectic form, which directly implies the handle slide invariance property without curve counting techniques in [OS04]. In dimension 2, Seidel and Smith [SS06] considered nilpotent slices instead of Symκ(M)\mathrm{Sym}^{\kappa}(M), which was shown to be a subset of Hilbκ(M)\mathrm{Hilb}^{\kappa}(M) by Manolescu [Man06]. Inside the nilpotent slice, matching cycles as Lagrangians related by arc slides are also Hamiltonian isotopic, which is not obvious in the cylindrical formulation. Mak and Smith [MS19] then showed that the cylindrical version is equivalent to the original symplectic Khovanov homology in nilpotent slices.

However, in higher dimensions, the cylindrical symplectic Khovanov homology with vanishing cycles κ\kappa-tuples of SnS^{n} does not have its “original” version; at this moment we do not know how to put the theory inside a nilpotent slice setting. Another problem is that Hilbert schemes of points in higher dimensions are not smooth, so we need new ways to resolve the singularities along the diagonal of Symκ(M)\mathrm{Sym}^{\kappa}(M). One possible solution is to restrict to some smooth stratum of Hilbκ(M)\mathrm{Hilb}^{\kappa}(M). For example, the subset of subschemes where each support point is of length at most 3 (at most triple point) is smooth. However we will not continue the discussion in this paper further. It would be interesting to study this problem in future.  

Therefore, we will adopt the curve counting approach as in [OS04] and [CHT20] to prove the invariance under arc slides and Markov stabilizations.

In Section 2, we begin with a brief review of Section 9 of [CHT20], providing necessary notations, definitions and prerequisite theorems. Then we state the main result. We use a subsection to explain the Morse flow tree theory and its relation to pseudoholomorphic curves originated from [FO97], which is crucial in our proof.

In Section 3, we show the arc slide invariance by counting pseudoholomorphic curves with certain boundary Lagrangians. The idea is to stretch the curve into several parts so that each one is easy to count by elementary model calculation.

In Section 4, we perform a model calculation of pseudoholomorphic quadrilaterals, which will be used in Section 5 and Section 6 for several times.

In Section 5, we translate the Markov stabilization into the gluing of pseudoholomorphic curves, which we count by Morse flow tree arguments instead.

In Section 6, we compute Kh(σ^)Kh^{\sharp}(\widehat{\sigma}) for unknots, Hopf links and trefoils.  

Acknowledgements. I would like to thank Ko Honda for countless discussions and introducing this project to me. I also thank Yin Tian for numerous ideas and suggestions, and thank Eilon Reisin-Tzur for his patient revision. I am partially supported by China Postdoctoral Science Foundation 2023T160002.

2. Definitions and main results

2.1. Higher-dimensional analogue of symplectic Khovanov homology

This subsection works as a review of [CHT20], so most proofs and details are omitted.

Let D~={2Rez,Imz2}z\widetilde{D}=\{-2\leq\mathrm{Re}\,z,\mathrm{Im}\,z\leq 2\}\subset\mathbb{C}_{z}. We consider the standard 2n2n-dimensional Lefschetz fibration

p~:W~D~z\widetilde{p}\colon\widetilde{W}\to\widetilde{D}\subset\mathbb{C}_{z}

for a Milnor fiber of the A2k1A_{2k-1} singularity, where the regular fiber is TSn1T^{*}S^{n-1}. There are 2κ2\kappa critical values 𝒛~={z1,,z2κ}\boldsymbol{\widetilde{z}}=\{z_{1},\dots,z_{2\kappa}\}, where Rezi=Rezi+κ\mathrm{Re}\,z_{i}=\mathrm{Re}\,z_{i+\kappa}, Imzi=1\mathrm{Im}\,z_{i}=-1 and Imzi+κ=1\mathrm{Im}\,z_{i+\kappa}=1, i=1,,κi=1,\dots,\kappa. Let

p:Wp~1(D)Dp\colon W\coloneqq\widetilde{p}^{-1}(D)\to{D}

be the restriction of p~\widetilde{p} to D=D~{Imz0}D=\widetilde{D}\cap\{\mathrm{Im}\,z\leq 0\}. For i=1,,κi=1,\dots,\kappa, connect ziz_{i} and zi+κz_{i+\kappa} by straight arcs γ~i\widetilde{\gamma}_{i}. Then the matching cycles a~={a~1,,a~κ}\widetilde{a}=\{\widetilde{a}_{1},\dots,\widetilde{a}_{\kappa}\} over {γ~1,,γ~κ}\{\widetilde{\gamma}_{1},\dots,\widetilde{\gamma}_{\kappa}\} are Lagrangian spheres. Let 𝒛\boldsymbol{z}, aia_{i}, and γi\gamma_{i} be “half” of 𝒛~\boldsymbol{\widetilde{z}}, a~i\widetilde{a}_{i}, and γ~i\widetilde{\gamma}_{i}, i.e., their restrictions to WW.

Given a κ\kappa-strand braid σDiff+(D,D,𝒛)\sigma\in\mathrm{Diff}^{+}(D,\partial D,\boldsymbol{z}), let hσSymp(W,W)h_{\sigma}\in\mathrm{Symp}(W,\partial W) be the monodromy on WW which descends to σ\sigma and let h~σ\widetilde{h}_{\sigma} be the extension of hσh_{\sigma} to W~\widetilde{W} by identity.

In this paper we always do cohomology. The variant CKh(σ^)CKh^{\sharp}(\widehat{\sigma}) of the symplectic Khovanov cochain complex is defined as the higher-dimensional Heegaard Floer cochain complex, in the sense of [Lip06] and [CHT20], denoted by CF^(W~,h~σ(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}). Specifically, a κ\kappa-tuple of intersection points of h~σ(𝒂~)\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}) and 𝒂~\widetilde{\boldsymbol{a}} is a κ\kappa-tuple 𝒚={y1,,yκ}\boldsymbol{y}=\{y_{1},\dots,y_{\kappa}\} where yia~ih~σ(a~β(i))y_{i}\in\widetilde{a}_{i}\cap\widetilde{h}_{\sigma}(\widetilde{a}_{\beta(i)}) and β\beta is some permutation of {1,,κ}\{1,\dots,\kappa\}. Then CF^(W~,h~σ(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) is the free 𝔽[𝒜],1]\mathbb{F}[\mathcal{A}]\llbracket\hbar,\hbar^{-1}]-module generated by all such κ\kappa-tuples 𝒚\boldsymbol{y}, where the coefficient ring is discussed below.

To define the differential, let FF be a surface with boundary of Euler characteristic χ\chi and F˙\dot{F} be the surface with boundary punctures. We start with the split almost complex structure J×[0,1]×JW~J_{\mathbb{R}\times[0,1]}\times J_{\widetilde{W}} on ×[0,1]×W~\mathbb{R}\times[0,1]\times\widetilde{W}, and apply a perturbation to achieve transversality, denoted by JJ^{\lozenge}.

For 𝒚,𝒚CF^(W~,h~σ(𝒂~),𝒂~)\boldsymbol{y},\boldsymbol{y}^{\prime}\in\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}), let Jind=1,A,χ(𝒚,𝒚)\mathcal{M}^{\mathrm{ind}=1,A,\chi}_{J^{\lozenge}}(\boldsymbol{y},\boldsymbol{y}^{\prime}) be the moduli space of u:F˙×[0,1]×W~u\colon\dot{F}\to\mathbb{R}\times[0,1]\times\widetilde{W} satisfying

  1. (1)

    duJ=Jdudu\circ J^{\lozenge}=J^{\lozenge}\circ du;

  2. (2)

    u(F˙)×(({1}×𝒂~)({0}×h~σ(𝒂~)))u(\partial\dot{F})\subset\mathbb{R}\times((\{1\}\times\widetilde{\boldsymbol{a}})\cup(\{0\}\times\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}})));

  3. (3)

    As πu\pi_{\mathbb{R}}\circ u tends to ++\infty (resp. -\infty), πW~\pi_{\widetilde{W}} tends to 𝒚\boldsymbol{y} (resp. 𝒚\boldsymbol{y}^{\prime}), where π,πW~\pi_{\mathbb{R}},\pi_{\widetilde{W}} are the projections of uu to \mathbb{R} and W~\widetilde{W}.

The differential is then defined as

(2.1) d𝒚=𝒚,χκ,A𝒜#Jind=1,A,χ(𝒚,𝒚)/κχeA𝒚.d\boldsymbol{y}=\sum_{\boldsymbol{y}^{\prime},\chi\leq\kappa,A\in\mathcal{A}}\#\mathcal{M}^{\mathrm{ind}=1,A,\chi}_{J^{\lozenge}}(\boldsymbol{y},\boldsymbol{y}^{\prime})/\mathbb{R}\cdot\hbar^{\kappa-\chi}\cdot e^{A}\cdot\boldsymbol{y}^{\prime}.

We write Kh(σ^)Kh^{\sharp}(\widehat{\sigma}) for the cohomology group HF^(W~,h~σ(𝒂~),𝒂~)\widehat{HF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}).

The coefficient ring 𝔽[𝒜],1]\mathbb{F}[\mathcal{A}]\llbracket\hbar,\hbar^{-1}] (power series in \hbar and polynomial in 1\hbar^{-1}) keeps track of the relative homology class and Euler characteristic of the domain, where

𝒜=H2([0,1]×W~,({1}×𝒂~)({0}×h~σ(𝒂~));).\mathcal{A}=H_{2}([0,1]\times\widetilde{W},(\{1\}\times\widetilde{\boldsymbol{a}})\cup(\{0\}\times\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}));\mathbb{Z}).

The following lemmas justify the use of coefficient 𝔽[𝒜],1]\mathbb{F}[\mathcal{A}]\llbracket\hbar,\hbar^{-1}] for n=2n=2 and 𝔽,1]\mathbb{F}\llbracket\hbar,\hbar^{-1}] for n>3n>3:

Lemma 2.1.

For fixed 𝐲{\boldsymbol{y}},𝐲{\boldsymbol{y}^{\prime}} and χ\chi, #Jind=1,χ(y,y)/\#\mathcal{M}^{\operatorname{ind}=1,\chi}_{J^{\lozenge}}(\textbf{y},\textbf{y}^{\prime})/\mathbb{R} is finite.

Lemma 2.2.

Suppose W~\widetilde{W} is of dimension 2n2n. For n=2n=2, 𝒜r1\mathcal{A}\simeq\mathbb{Z}^{r-1}, where rr is the number of connected components of σ^\widehat{\sigma}; for n>2n>2, 𝒜{0}\mathcal{A}\simeq\{0\}.

Proof.

Denote X=[0,1]×W~X=[0,1]\times\widetilde{W}, Y=({1}×𝒂~)({0}×h~σ(𝒂~)))Y=(\{1\}\times\widetilde{\boldsymbol{a}})\cup(\{0\}\times\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}))), and i:YXi\colon Y\to X for the inclusion.

If n=2n=2, H3(X,Y)H_{3}(X,Y) and H1(Y)H_{1}(Y) are trivial. By the long exact sequence for relative singular homology, H2(X,Y)H2(X)/iH2(Y)H_{2}(X,Y)\simeq H_{2}(X)/i_{*}H_{2}(Y). Since there are 2κ2\kappa critical points of p~:W~D~\widetilde{p}\colon\widetilde{W}\to\widetilde{D}, H2(X)H_{2}(X) is generated by 2κ12\kappa-1 matching cycles over arcs connecting pairs of critical values in D~\widetilde{D}. H2(Y)H_{2}(Y) is simply generated by collections 𝒂~\widetilde{\boldsymbol{a}} and h~σ(𝒂~)\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}). Therefore, H2(X,Y)H_{2}(X,Y) is generated by 2κ12\kappa-1 arcs quotient by collapsing arcs corresponding to 𝒂~\widetilde{\boldsymbol{a}} and h~σ(𝒂~)\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}). The number of remaining nontrivial arcs is r1r-1, where rr is the number of connected components of the braid σ^\widehat{\sigma}.

If n>2n>2, H2(X)H_{2}(X) and H1(Y)H_{1}(Y) are trivial, so H2(X,Y){0}H_{2}(X,Y)\simeq\{0\} by the relative homology exact sequence. ∎

The following lemma computes the Fredholm index, which is a generalization from [CGH12] and the proof is omitted:

Lemma 2.3.

The Fredholm index of uu is

ind(u)=(n2)(χκ)+μ(u),\mathrm{ind}(u)=(n-2)(\chi-\kappa)+\mu(u),

where μ(u)\mu(u) is the Maslov index.

The Floer homology group Kh(σ^)Kh^{\sharp}(\widehat{\sigma}) is well-defined and now we state our main result, which is a higher-dimensional version of Theorem 1.2.1 of [CHT20]:

Theorem 2.4.

For n=2n=2 or n>3n>3, Kh(σ^)Kh^{\sharp}(\widehat{\sigma}) is a link invariant, i.e., it is independent of the choice of arcs {γ~1,,γ~κ}\{\widetilde{\gamma}_{1},\dots,\widetilde{\gamma}_{\kappa}\} and Lagrangian thimbles {a~1,,a~κ}\{\widetilde{a}_{1},\dots,\widetilde{a}_{\kappa}\}, and is invariant under Markov stabilizations.

Remark 2.5.

Kh(σ^)Kh^{\sharp}(\widehat{\sigma}) is a relative graded module over 𝔽[𝒜],1]\mathbb{F}[\mathcal{A}]\llbracket\hbar,\hbar^{-1}]. From (2.1) and Lemma 2.3, \hbar is of degree 2n2-n.

Remark 2.6.

Note that in Theorem 2.4, the case of n=3n=3 is excluded. The reason is explained by the remark after Lemma 3.2 in Section 3.

The computation of Kh(σ^)Kh^{\sharp}(\widehat{\sigma}) for simple links in Section 6 gives results highly similar to Khovanov homology [Kho00, Wil08]. We will not discuss their relation further in this paper. Instead, we leave it as a conjecture:

Conjecture 2.7.
Kh,k(σ^)ij=kmodn2Khi,j(σ).Kh^{\sharp,k}(\widehat{\sigma})\simeq\bigoplus_{i-j=k\,\operatorname{mod}\,n-2}Kh^{i,j}(\sigma).

Note that Conjecture 2.7 implies that we have found nothing new other than the symplectic Khovanov homology [SS06]. While it may serve as a new approach to compute the symplectic Khovanov homology.

2.2. Moduli space of Morse gradient trees

If two closed Lagrangian submanifolds L0,L1L_{0},L_{1} of a symplectic manifold (M,ω)(M,\omega) are C1C^{1}-close to each other, and moreover they are exact Lagrangian isotopic, then by the Lagrangian neighbourhood theorem, one can view L1L_{1} as the graph ΓdfTL0\Gamma_{df}\subset T^{*}L_{0}, where f:L0f\colon L_{0}\to\mathbb{R}. In this case, the count of pseudoholomorphic disks bounding L0,L1L_{0},L_{1} can be reduced to the count of Morse gradient trajectories on L0L_{0}, which is much more convenient to compute. More generally, if we have exact Lagrangian isotopic L0,,Lk1L_{0},\dots,L_{k-1}, where Li=ΓdfiTL0L_{i}=\Gamma_{df_{i}}\subset T^{*}L_{0} for fi:L0,i=0,,k1f_{i}\colon L_{0}\to\mathbb{R},\,i=0,\dots,k-1, then we count gradient trees, to be explained below. This alternative method of counting appears several times in Section 3 and 5, so we make the statement precise here. The main reference is [FO97], which generalizes Floer’s correspondence between gradient trajectories and pseudoholomorphic strips [Flo88]. There is also a simpler proof of [FO97] by Iacovino [Iac08], which considers the perturbation of Floer equations.

Refer to caption
Figure 1. An embedding of TT into D2D^{2}.
Definition 2.8.

A ribbon tree is a directed tree TT which satisfies:

  1. (1)

    TT has no degree 2 vertex.

  2. (2)

    For each internal vertex vTv\in T, there is a cyclic order of the edges attached to vv, i.e., we have a labeling bijection v:Ev{0,1,,kv1}\ell_{v}\colon E_{v}\to\{0,1,\dots,k_{v}-1\}, where EvE_{v} is the set of edges attached to vv and kvk_{v} is the degree of vv such that there is a unique outgoing edge and it is labeled 0. Therefore, there are two labels for each internal edge coming from the two labeling functions corresponding to the endpoints of the edge and one label for each external edge.

  3. (3)

    There is a cyclic order of external vertices of TT such that the root has label 0, i.e., we have a bijection :Vext{0,,k1}\ell\colon V_{ext}\to\{0,\dots,k-1\} where VextV_{ext} is the set of external vertices and the one adjacent to the unique outgoing edge is labeled by 0.

  4. (4)

    There exists an embedding ϕ:TD2\phi\colon T\to D^{2} so that at each internal vertex vv, ϕ(v1(0)),,ϕ(v1(kv1))\phi(\ell_{v}^{-1}(0)),\dots,\phi(\ell_{v}^{-1}(k_{v}-1)) are in counterclockwise order around ϕ(v)\phi(v). Also, ϕ(Vext)D2\phi(V_{ext})\subset\partial D^{2} and ϕ(1(0)),,ϕ(1(k1))\phi(\ell^{-1}(0)),\dots,\phi(\ell^{-1}(k-1)) are in counterclockwise order on D2\partial D^{2}. Note that the way we embed TT is not important.

In the definition univalent vertices are called external and the others are internal; edges adjacent to external vertices are called external and the others are internal.

The edges of TT will represent gradient flows. Specifically, let (M,g)(M,g) be a Riemannian manifold and f0,,fk1f_{0},\dots,f_{k-1} be CC^{\infty}-functions on MM so that fifi+1f_{i}-f_{i+1} is Morse for each ii (where fk=f0f_{k}=f_{0}). We define the moduli space of gradient trees g(M;𝒇,𝒑)\mathcal{M}_{g}(M;\boldsymbol{f},\boldsymbol{p}) to consist of pairs (T,I)(T,I) where TT is a ribbon tree and I:T\VextMI\colon T\backslash V_{ext}\to M is a continuous function satisfying:

  1. (1)

    limxviI(x)=pi\lim_{x\to v_{i}}I(x)=p_{i}, where pip_{i} is a critical point of fifi+1f_{i}-f_{i+1}.

  2. (2)

    External edges are identified with (,0](-\infty,0] except the outgoing one which is identified with [0,)[0,\infty); each internal edge ee is identified with [0,t(e)][0,t(e)] where t:Eint[0,)t:E_{int}\to[0,\infty) is a length function defined on internal edges. Now II takes the above intervals to MM. For each edge ee of TT,

    I˙|e=g(fl(e)fr(e)),\dot{I}|_{e}=-\nabla_{g}(f_{l(e)}-f_{r(e)}),

    where l(e)l(e) and r(e)r(e) are defined with respect to the direction of ee, i.e. if one looks in the positive direction of ee on D2D^{2}, then l(e)l(e) (resp. r(e)r(e)) is the component of D2\TD^{2}\backslash T on the left (resp. right) side of ee.

Figure 1 shows an element TT, together with an embedding of TT in D2D^{2}. Note that we assign each function fif_{i} to a corresponding region of D2\TD^{2}\backslash T.

Lemma 2.9 ([FO97]).

Fixing a generic choice of 𝐟\boldsymbol{f}, g(M,𝐟,𝐩)\mathcal{M}_{g}(M,\boldsymbol{f},\boldsymbol{p}) is a smooth manifold of dimension

(2.2) vind(v)(k1)n+(k3),\sum_{v}\mathrm{ind}(v)-(k-1)n+(k-3),

where kk is the number of boundary vertices and nn is the dimension of the manifold.

Next we define the moduli space of pseudoholomorphic disks bounding exact Lagrangian submanifolds in TMT^{*}M. Let (M,g)(M,g) be a Riemannian manifold. Let 𝒇=(f0,,fk1)\boldsymbol{f}=(f_{0},\dots,f_{k-1}) be a generic collection of functions on MM, and let 𝚪=(Γdf0,,Γdfk1)\boldsymbol{\Gamma}=(\Gamma_{df_{0}},\dots,\Gamma_{df_{k-1}}) be the graphs of their differentials. We associate each critical point pip_{i} of fifi+1f_{i}-f_{i+1} with xi=(pi,ϵdfi(pi))ϵΓdfi+1ϵΓdfix_{i}=(p_{i},\epsilon df_{i}(p_{i}))\in\epsilon\Gamma_{df_{i+1}}\cap\epsilon\Gamma_{df_{i}} where ϵ>0\epsilon>0 is small.

We then fix a canonical almost complex structure JgJ_{g} on TMT^{*}M associated to the metric gg on MM such that

  1. (1)

    JgJ_{g} is compatible with the canonical symplectic form ω\omega on TMT^{*}M.

  2. (2)

    JgJ_{g} maps vertical tangent vectors to horizontal tangent vectors of TMT^{*}M with respect to gg.

  3. (3)

    On the zero section of TMT^{*}M, for vTqMT(q,0)(TM)v\in T_{q}M\subset T_{(q,0)}(T^{*}M), let Jg(v)=g(v,)TqMT(q,0)(TM)J_{g}(v)=g(v,\cdot)\in T_{q}^{*}M\subset T_{(q,0)}(T^{*}M).

Definition 2.10.

Let Jg(TM;ϵ𝚪,𝐱)\mathcal{M}_{J_{g}}(T^{*}M;\epsilon\boldsymbol{\Gamma},\boldsymbol{x}) consist of pairs (u,D𝐳2)(u,D^{2}_{\boldsymbol{z}}) where D𝐳2D^{2}_{\boldsymbol{z}} is the domain D2D^{2} with marked points 𝐳=(z0,,zk1)\boldsymbol{z}=(z_{0},\dots,z_{k-1}) arranged in counterclockwise order on D2\partial D^{2} and u:D2\{z0,,zk1}TMu\colon D^{2}\backslash\{z_{0},\dots,z_{k-1}\}\to T^{*}M satisfies

  1. (1)

    u(zi)=piu(z_{i})=p_{i},

  2. (2)

    u(iD2)ϵΓdfiu(\partial_{i}D^{2})\subset\epsilon\Gamma_{df_{i}},

  3. (3)

    Jgdu=duJgJ_{g}\circ du=du\circ J_{g},

where iD2\partial_{i}D^{2} is the shortest counterclockwise arc between ziz_{i} and zi+1z_{i+1} on D2\partial D^{2}. Here we are identifying pairs (u,D𝐳2)(u,D^{2}_{\boldsymbol{z}}) and (v,D𝐳2)(v,D^{2}_{\boldsymbol{z}^{\prime}}) which are related by an isomorphism of the domain.

The following theorem relates Morse gradient trees on MM to pseudoholomorphic disks on TMT^{*}M:

Theorem 2.11 ([FO97]).

For ϵ>0\epsilon>0 sufficiently small, there is an oriented diffeomorphism g(M;𝐟,𝐩)Jg(TM;ϵ𝚪,𝐱)\mathcal{M}_{g}(M;\boldsymbol{f},\boldsymbol{p})\cong\mathcal{M}_{J_{g}}(T^{*}M;\epsilon\boldsymbol{\Gamma},\boldsymbol{x}).

Roughly speaking, for each gradient tree (T,I)(T,I), we can construct a pseudoholomorphic curve near I(T)I(T) inside TMT^{*}M by some gluing techniques. We should note that the theorem depends on a specific choice of almost complex structure JgJ_{g}.

Refer to caption
Figure 2. The gradient tree TT viewed inside D2D^{2} (left), the perturbation of Lagrangians in the special case of TS1T^{*}S^{1} (middle) and the image of TT on SnS^{n} (right). We abuse notation and label both the domain and image of a vertex by viv_{i} since there is no ambiguity.

The following example counts pseudoholomorphic triangles with 3 Lagrangian boundaries, which we will meet again in Section 3.

Example 2.12.

We give an example of a gradient tree with 3 vertices here. Consider functions f0,f1,f2f_{0},f_{1},f_{2} on SnS^{n} so that the domain D2D^{2} looks like the left side of Figure 2. We can perturb f0,f1,f2f_{0},f_{1},f_{2} generically so that v1v_{1} is a source (top generator) of (f2f1)\nabla(f_{2}-f_{1}), v2v_{2} is a sink (bottom generator) of (f0f2)\nabla(f_{0}-f_{2}) and v0v_{0} is a sink (bottom generator) of (f0f1)\nabla(f_{0}-f_{1}). The middle of Figure 2 shows a possible perturbation when n=1n=1 for illustration. Sources and sinks are denoted by blue arrows. The right side of Figure 2 shows what happens on SnS^{n}: v1v_{1} is the unique top generator of f2f1f_{2}-f_{1} on SnS^{n}, so the flows from v1v_{1} form a Sn1S^{n-1}-family and pass through all the points of SnS^{n} except two critical points. v2v_{2} is the bottom generator of f0f2f_{0}-f_{2}, so the flow from v2v_{2} of (f0f2)\nabla(f_{0}-f_{2}) is of length 0. Therefore the flow from v1v_{1} should pass v2v_{2} and there is a unique such flow line. Similarly, there is a unique flow line from v2v_{2} to v0v_{0} of (f0f1)\nabla(f_{0}-f_{1}). To conclude, there is a unique gradient tree. Thus there is a unique pseudoholomorphic disk bounded by the Lagrangians ϵΓdf0,ϵΓdf1,ϵΓdf2\epsilon\Gamma_{df_{0}},\epsilon\Gamma_{df_{1}},\epsilon\Gamma_{df_{2}} for small ϵ\epsilon by Theorem 2.11.

3. Invariance under arc slides

Given {γ~1,,γ~κ}\{\widetilde{\gamma}_{1},\dots,\widetilde{\gamma}_{\kappa}\} from Section 2.1, consider the arc slide of γ~1\widetilde{\gamma}_{1} over γ~2\widetilde{\gamma}_{2}: Let {γ~1,,γ~κ}\{\widetilde{\gamma}^{\prime}_{1},\dots,\widetilde{\gamma}^{\prime}_{\kappa}\} be the new set of arcs as in Figure 3. Let 𝒂~={a~1,,a~κ}\widetilde{\boldsymbol{a}}^{\prime}=\{\widetilde{a}^{\prime}_{1},\dots,\widetilde{a}^{\prime}_{\kappa}\} be the new tuple of Lagrangians over {γ~1,,γ~κ}\{\widetilde{\gamma}^{\prime}_{1},\dots,\widetilde{\gamma}^{\prime}_{\kappa}\}. For i=1,,κi=1,\dots,\kappa, let Θi\Theta_{i} (resp. Ξi\Xi_{i}) be the intersection point of a~i\widetilde{a}_{i} and a~i\widetilde{a}^{\prime}_{i} that lies over ziz_{i} (resp. zi+κz_{i+\kappa}). Denote 𝚯={Θ1,,Θκ}\boldsymbol{\Theta}=\{\Theta_{1},\dots,\Theta_{\kappa}\} and 𝚵={Ξ1,,Ξκ}\boldsymbol{\Xi}=\{\Xi_{1},\dots,\Xi_{\kappa}\}.

Refer to caption
Figure 3. Arc sliding of γ~1\widetilde{\gamma}_{1} over γ~2\widetilde{\gamma}_{2}.

The purpose of this section is to prove:

Theorem 3.1.

CF^(W~,h~σ(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) and CF^(W~,h~σ(𝐚~),𝐚~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}^{\prime}),\widetilde{\boldsymbol{a}}^{\prime}) are quasi-isomorphic.

Proof.

It suffices to show the quasi-isomorphism between CF^(W~,h~σ(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) and CF^(W~,h~σ(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}^{\prime}). One can show the quasi-isomorphism between CF^(W~,h~σ(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}^{\prime}) and CF^(W~,h~σ(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}^{\prime}),\widetilde{\boldsymbol{a}}^{\prime}) similarly.

We define the cochain map by the μ2\mu_{2} composition map of the AA_{\infty}-relation:

Φ:CF^(W~,h~σ(𝒂~),𝒂~)CF^(W~,h~σ(𝒂~),𝒂~),\Phi\colon\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}})\to\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}^{\prime}),
𝐲μ2(𝚵𝐲),\mathrm{\boldsymbol{y}}\mapsto\mu_{2}(\boldsymbol{\Xi}\otimes\mathrm{\boldsymbol{y}}),

where μ2\mu_{2} is the product map

μ2:CF^(W~,𝒂~,𝒂~)CF^(W~,h~σ(𝒂~),𝒂~)CF^(W~,h~σ(𝒂~),𝒂~).\mu_{2}\colon\widehat{CF}(\widetilde{W},\widetilde{\boldsymbol{a}},\widetilde{\boldsymbol{a}}^{\prime})\otimes\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}})\to\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}^{\prime}).

Similarly we define the cochain map going back by

Ψ:CF^(W~,h~σ(𝒂~),𝒂~)CF^(W~,h~σ(𝒂~),𝒂~),\Psi\colon\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}^{\prime})\to\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}),
𝐲μ2(𝚯𝐲).\mathrm{\boldsymbol{y}}\mapsto\mu^{\prime}_{2}(\boldsymbol{\Theta}\otimes\mathrm{\boldsymbol{y}}).

The following lemma justifies that Φ\Phi, Ψ\Psi are well-defined:

Lemma 3.2.

For n=2n=2 or n>3n>3, 𝚵\boldsymbol{\Xi} is a cocycle in CF^(W~,𝐚~,𝐚~)\widehat{CF}(\widetilde{W},\widetilde{\boldsymbol{a}},\widetilde{\boldsymbol{a}}^{\prime}) and 𝚯\boldsymbol{\Theta} is a cocycle in CF^(W~,𝐚~,𝐚~)\widehat{CF}(\widetilde{W},\widetilde{\boldsymbol{a}}^{\prime},\widetilde{\boldsymbol{a}}).

Proof of Lemma 3.2: For simplicity assume κ=2\kappa=2. By Lemma 2.3 and a Maslov index calculation,

(3.1) ind\displaystyle\mathrm{ind} (u;𝚯,𝚵)=(n2)(χ2)+4n4,\displaystyle(u;\boldsymbol{\Theta},\boldsymbol{\Xi})=(n-2)(\chi-2)+4n-4,
(3.2) ind\displaystyle\mathrm{ind} (u;𝚯,{Ξ2,Θ1})=(n2)(χ2)+n,\displaystyle(u;\boldsymbol{\Theta},\{\Xi_{2},\Theta_{1}\})=(n-2)(\chi-2)+n,
(3.3) ind\displaystyle\mathrm{ind} (u;𝚯,{Ξ1,Θ2})=(n2)(χ2)+3n4,\displaystyle(u;\boldsymbol{\Theta},\{\Xi_{1},\Theta_{2}\})=(n-2)(\chi-2)+3n-4,

where for example, ind(u;𝚯,𝚵)\mathrm{ind}(u;\boldsymbol{\Theta},\boldsymbol{\Xi}) denotes the Fredholm index of curves u:F˙×[0,1]×W~u\colon\dot{F}\to\mathbb{R}\times[0,1]\times\widetilde{W} such that

  1. (1)

    duJ=Jdudu\circ J^{\lozenge}=J^{\lozenge}\circ du,

  2. (2)

    u(F˙)×(({1}×𝒂~)({0}×𝒂~))u(\partial\dot{F})\subset\mathbb{R}\times((\{1\}\times\widetilde{\boldsymbol{a}})\cup(\{0\}\times\widetilde{\boldsymbol{a}}^{\prime})),

  3. (3)

    As πu\pi_{\mathbb{R}}\circ u tends to ++\infty (resp. -\infty), πW~\pi_{\widetilde{W}} tends to 𝚯\boldsymbol{\Theta} (resp. 𝚵\boldsymbol{\Xi}), where π\pi_{\mathbb{R}} and πW~\pi_{\widetilde{W}} are the projections of uu to \mathbb{R} and W~\widetilde{W}.

We explain (3.1)-(3.3) now. First, the Maslov index of a closed path over γ~2\widetilde{\gamma}_{2}^{\prime} and γ~2\widetilde{\gamma}_{2} is nn by definition, which implies (3.2). To get (3.3), observe that the projection to D~\widetilde{D} of the curve from Θ1\Theta_{1} to Ξ1\Xi_{1} is over the region surrounded by γ~1\widetilde{\gamma}_{1}^{\prime} and γ~1\widetilde{\gamma}_{1}, which contains two critical values of the Lefschetz fibration. The region surrounded by γ~1\widetilde{\gamma}_{1}^{\prime} and γ~1\widetilde{\gamma}_{1} can be viewed as the outcome after applying Lagrangian surgery twice on a trivial strip to incorporate the two critical values. By Theorem 55.5 of [FOOO09], each Lagrangian surgery increases the Fredholm index by n2n-2. Therefore, ind(u;𝚯,{Ξ1,Θ2})=(n2)(χ2)+3n4\mathrm{ind}(u;\boldsymbol{\Theta},\{\Xi_{1},\Theta_{2}\})=(n-2)(\chi-2)+3n-4 by comparing with (3.2). Finally, (3.1) follows by adding the Maslov index terms of (3.2) and (3.3).

For n=2n=2, the Fredholm index does not depend on χ\chi: ind(u;𝚯,𝚵)=4\mathrm{ind}(u;\boldsymbol{\Theta},\boldsymbol{\Xi})=4 and ind(u;𝚯,{Ξ2,Θ1})=ind(u;𝚯,{Ξ1,Θ2})=2\mathrm{ind}(u;\boldsymbol{\Theta},\{\Xi_{2},\Theta_{1}\})=\mathrm{ind}(u;\boldsymbol{\Theta},\{\Xi_{1},\Theta_{2}\})=2.

For n>3n>3, observe that in all 3 cases the domain FF has 2 punctures and thus χ\chi is even. In particular, ind(u;𝚯,𝚵)\mathrm{ind}(u;\boldsymbol{\Theta},\boldsymbol{\Xi}) is even and ind(u;𝚯,𝚵)1\mathrm{ind}(u;\boldsymbol{\Theta},\boldsymbol{\Xi})\neq 1; If ind(u;𝚯,{Ξ1,Θ2})=1\mathrm{ind}(u;\boldsymbol{\Theta},\{\Xi_{1},\Theta_{2}\})=1, then χ=2n1n2\chi=2-\frac{n-1}{n-2}, which is not a integer for n>3n>3; If ind(u;𝚯,{Ξ2,Θ1})=1\mathrm{ind}(u;\boldsymbol{\Theta},\{\Xi_{2},\Theta_{1}\})=1, then χ=11n2\chi=-1-\frac{1}{n-2}, which is also not a integer for n>3n>3.

Therefore, for n=2n=2 or n>3n>3, there is no index 1 curve from 𝚯\boldsymbol{\Theta}, i.e., 𝚯\boldsymbol{\Theta} is a cocycle in CF^(W~,𝒂~,𝒂~)\widehat{CF}(\widetilde{W},\widetilde{\boldsymbol{a}}^{\prime},\widetilde{\boldsymbol{a}}). The case of 𝚵\boldsymbol{\Xi} is similar. ∎

Remark 3.3.

Lemma 3.2 does not deal with the case of n=3n=3, where the index could be 1 for some χ\chi. Therefore, to show 𝚯\boldsymbol{\Theta} is a cocycle, we still need to understand the index 1 pseudoholomorphic curves from Θ2\Theta_{2} to Ξ2\Xi_{2}, whose projection to D~\widetilde{D} is the thin strip surrounded by γ~2\widetilde{\gamma}^{\prime}_{2} and γ~2\widetilde{\gamma}_{2}, and also the curves from Θ1\Theta_{1} to Ξ1\Xi_{1}, whose projection to D~\widetilde{D} is the fat strip surrounded by γ~1\widetilde{\gamma}^{\prime}_{1} and γ~1\widetilde{\gamma}_{1}.

It remains to show that ΨΦ\Psi\circ\Phi induces identity on cohomology (with some nonzero coefficient). The composition of Ψ\Psi and Φ\Phi is given by the left-hand side of Figure 4, which is viewed as a degeneration of a family of curves. The right-hand side of Figure 4 shows another degeneration. Observe that the right-hand part of the right degeneration is of index 0. In fact the right degeneration corresponds to the identity map times the count of its left-hand part, which is chain homotopic to ΨΦ\Psi\circ\Phi.

Refer to caption
Figure 4. Two possible degenerations.

Now we count the left-hand part of the right degeneration of Figure 4. For convenience, assume κ=2\kappa=2, i.e., focus on the arc sliding of γ~1\widetilde{\gamma}_{1} over γ~2\widetilde{\gamma}_{2}. The moduli space J(𝚵,𝚯)\mathcal{M}_{J}(\boldsymbol{\Xi},\boldsymbol{\Theta}) contains curves uu with boundary condition which maps F˙\partial\dot{F} to (×{1}×𝒂~)(×{0}×𝒂~)(\mathbb{R}\times\{1\}\times\boldsymbol{\widetilde{a}^{\prime}})\cup(\mathbb{R}\times\{0\}\times\boldsymbol{\widetilde{a}}). There is another restriction on the right degeneration: uu passes through (0,0,w1)(0,0,w_{1}) and (0,0,w2)(0,0,w_{2}), where (0,0)×[0,1](0,0)\in\mathbb{R}\times[0,1] and wi𝒂~𝒊w_{i}\in\boldsymbol{\widetilde{a}_{i}}. Passing through a generic 𝒘={w1,w2}\boldsymbol{w}=\{w_{1},w_{2}\} is a codimension 2n2n condition. By Lemma 2.3, ind(𝚵,𝚯)=2n\mathrm{ind}(\boldsymbol{\Xi},\boldsymbol{\Theta})=2n if and only if n=2n=2 or χ=0\chi=0 for n>3n>3.

By Theorem 3.5 below, the count of Jχ=0(𝚵,𝚯)\mathcal{M}_{J}^{\chi=0}(\boldsymbol{\Xi},\boldsymbol{\Theta}) passing a generic 𝒘\boldsymbol{w} is 1 (mod 2). Thus ΨΦ\Psi\circ\Phi is cochain homotopic to identity with some nonzero coefficient. The case of ΦΨ\Phi\circ\Psi is similar. ∎

3.1. Half of curve counting

There is a half version of the counting problem, where the base is shown as Figure 5. The base DD can be extended to \mathbb{C} by adding cylindrical ends, denoted by D¯\bar{D}. Let a¯i\bar{a}_{i} and a¯j\bar{a}_{j}^{\prime} be the cylindrical completion of aia_{i} and aja_{j}^{\prime}. The asymptotic Reeb chords from a¯i\bar{a}_{i} to a¯j\bar{a}_{j}^{\prime} form a Sn1S^{n-1}-family. We then perturb the contact form so that the Sn1S^{n-1}-family becomes Morse-Bott and let cˇij,c^ij\check{c}_{ij},\hat{c}_{ij} be the longer and shorter asymptotic Reeb chords. Let 𝒄={cˇ12,cˇ21}\boldsymbol{c}=\{\check{c}_{12},\check{c}_{21}\}. The problem is to count Jχ=1,𝒘(𝒄,𝚯)\mathcal{M}_{J^{\lozenge}}^{\chi=1,\boldsymbol{w}}(\boldsymbol{c},\boldsymbol{\Theta}), which is over the cylindrical extension of the region surrounded by γ1,γ2,γ1,γ2\gamma_{1},\gamma_{2},\gamma_{1}^{\prime},\gamma_{2}^{\prime}.

Refer to caption
Figure 5. Half of the base. We count pseudoholomorphic disks surrounded by γ1,γ2,γ1,γ2\gamma_{1},\gamma_{2},\gamma_{1}^{\prime},\gamma_{2}^{\prime}.
Theorem 3.4.

#Jχ=1,𝒘(𝒄,𝚯)=1\#\mathcal{M}_{J^{\lozenge}}^{\chi=1,\boldsymbol{w}}(\boldsymbol{c},\boldsymbol{\Theta})=1 mod 22 for generic 𝐰\boldsymbol{w}.

Proof.

For uJχ=1,𝒘(𝒄,𝚯)u\in\mathcal{M}_{J^{\lozenge}}^{\chi=1,\boldsymbol{w}}(\boldsymbol{c},\boldsymbol{\Theta}), let v=πW¯u:F˙W¯v=\pi_{\bar{W}}\circ u\colon\dot{F}\to\bar{W}, where W¯\bar{W} is the cylindrical extension of WW and FF is the unit disk in \mathbb{C}.

Recall that p:W¯D¯p\colon\bar{W}\to\bar{D} is the projection. The map pv:F˙D¯p\circ v\colon\dot{F}\to\bar{D} has degree 2 (resp. 1) over region A (resp. B) of Figure 5, which are connected components of Dγ1γ1γ2γ2D-{\gamma}_{1}\cup{\gamma}^{\prime}_{1}\cup{\gamma}_{2}\cup{\gamma}^{\prime}_{2} with extension. There are two possible branching behaviors: Type intint has a branch point bb that maps to the interior of region A; Type \partial is more obscure which has two switch points b1,b2b_{1},b_{2} on the boundary of region A instead of a branch point. Denote Type 1\partial_{1} for the case b1,b2b_{1},b_{2} on γ2\gamma_{2}^{\prime} and Type 2\partial_{2} for b1,b2b_{1},b_{2} on γ2\gamma_{2}. Choose b1b_{1} as the point closer to the puncture c12c_{12} (resp. c21c_{21}) on F˙\partial\dot{F} that maps to γ2\gamma_{2}^{\prime} (resp. γ2\gamma_{2}).

There is another codimension-1 constraint. Denote the preimage of * under uu as q()q(*), and we will often leave this notation out if there is no ambiguity. A possible arrangement of points on F˙\partial\dot{F} is shown as Figure 6. Since the projection of uu to ×[0,1]\mathbb{R}\times[0,1] is a double branched cover and under this projection, {q(w1),q(w2)}\{q(w_{1}),q(w_{2})\}, {q(Θ1),q(Θ2)}\{q(\Theta_{1}),q(\Theta_{2})\}, {q(cˇ12),q(cˇ21)}\{q(\check{c}_{12}),q(\check{c}_{21})\} are mapped to (0,0),,+(0,0),-\infty,+\infty, respectively, the deck transformation requires the involution:

(3.4) q(Θ1)q(Θ2),q(cˇ21)q(cˇ12),q(w2)q(w1).q(\Theta_{1})\mapsto q(\Theta_{2}),\,q(\check{c}_{21})\mapsto q(\check{c}_{12}),\,q(w_{2})\mapsto q(w_{1}).
Refer to caption
Figure 6.

The counting strategy uses the fact that the mod 2 count #Jχ=1,𝒘(𝒄,𝚯)\#\mathcal{M}_{J^{\lozenge}}^{\chi=1,\boldsymbol{w}}(\boldsymbol{c},\boldsymbol{\Theta}) does not depend on 𝒘\boldsymbol{w}, which allows us to stretch the curve by letting 𝒘\boldsymbol{w} tend to some limit:

  1. (1)

    Let |p(w2)|0|p(w_{2})|\gg 0;

  2. (2)

    Let the width between γ2,γ2,m0\gamma_{2},\gamma_{2}^{\prime},m\to 0;

  3. (3)

    Choose w1w_{1} such that p(w1)z1p(w_{1})\to z_{1}.

StepStep 1. Denote ι=Imp\iota=\mathrm{Im}\circ p. We first observe that as p(w1)z1p(w_{1})\to z_{1}, either ιv(b)ι(w2)\iota\circ v(b)\gg\iota(w_{2}) or ιv(b2)ι(w2)\iota\circ v(b_{2})\gg\iota(w_{2}). If this is not true, then by Gromov compactness there exists a limiting curve which contradicts the involution condition 3.4. We refer the reader to [CHT20] for details.


StepStep 2. Write Jχ=1,𝒘,(𝒄,𝚯)\mathcal{M}_{J^{\lozenge}}^{\chi=1,\boldsymbol{w},\sharp}(\boldsymbol{c},\boldsymbol{\Theta}) for the curves satisfying ιv(b)ι(w2)\iota\circ v(b)\gg\iota(w_{2}) or ιv(b1)ι(w2)C\iota\circ v(b_{1})\geq\iota(w_{2})-C. The case of ιv(b)ι(w2)\iota\circ v(b)\gg\iota(w_{2}) is shown as the left of Figure 7, where the limiting 2-level curve contains v(1)v(2)v(3)v^{(1)}\cup v^{(2)}\cup v^{(3)}. The case of ιv(b1)ι(w2)C\iota\circ v(b_{1})\geq\iota(w_{2})-C is similar. We push off the bottom singular point a little so that a1a_{1} and a1a_{1}^{\prime} have clean Sn1S^{n-1}-intersection, which does not change the moduli space of v(1)v^{(1)} by [FOOO09].

Refer to caption
Figure 7.

The region of pv(1)p\circ v^{(1)} contains a singular point of a Lefschetz fibration, which can be viewed as the outcome after a Lagrangian surgery from the trivial region without singular points. By Theorem 55.5 of [FOOO09], the moduli space of v(1)v^{(1)} is diffeomorphic to Sn2S^{n-2}. Therefore the evaluation map of v(1)v^{(1)} at cˇ11\check{c}_{11} sweeps Sn2S^{n-2} inside the Sn1S^{n-1}-family of Reeb chords, of which the homology class vanishes. The result is #Jχ=1,𝒘,(𝒄,𝚯)=0\#\mathcal{M}_{J^{\lozenge}}^{\chi=1,\boldsymbol{w},\sharp}(\boldsymbol{c},\boldsymbol{\Theta})=0 mod 2.


StepStep 3. The remaining case is the curve satisfying ιv(b2)ι(w2)\iota\circ v(b_{2})\gg\iota(w_{2}) and ιv(b1)ι(w2)C\iota\circ v(b_{1})\leq\iota(w_{2})-C. The stretched limiting curves are shown as the middle and the right of Figure 7. We expect that such curves exist and contribute 1 (mod 2) to Jχ=1,𝒘(𝒄,𝚯)\mathcal{M}_{J^{\lozenge}}^{\chi=1,\boldsymbol{w}}(\boldsymbol{c},\boldsymbol{\Theta}).

First we treat Type 1\partial_{1}. v(2)v^{(2)} is uniquely determined since there is a unique gradient trajectory from cˇ22\check{c}_{22} to Θ2\Theta_{2} passing through w2w_{2}. Denote the bottom-left Reeb chord of v(3)v^{(3)} by d^21\hat{d}_{21}. We show that the pseudoholomorphic triangle v(3)v^{(3)} exists uniquely: The Lefschetz fibration around v(3)v^{(3)} is the trivial one p:×TSn1p\colon\mathbb{C}\times T^{*}S^{n-1}\to\mathbb{C}. We can perturb the Lagrangians a1,a1,a2,a2a_{1},a_{1}^{\prime},a_{2},a_{2}^{\prime} in the fiber direction. The case of n=2n=2 is shown as Figure 8, where there exists a unique pseudoholomorphic triangle.

Refer to caption
Figure 8. The fiber TS1T^{*}S^{1} with perturbed Lagrangians. The sides are identified.

More generally, the case of n2n\geq 2 is done by the gradient tree argument. Put the perturbed Lagrangians as Figure 8 on TS1TSn1T^{*}S^{1}\subset T^{*}S^{n-1} and extend to TSn1T^{*}S^{n-1}, where we view S1Sn1S^{1}\subset S^{n-1} as a equator. After a further small perturbation, we consider the moduli space of Morse gradient trees with 3 vertices cˇ22,cˇ21,d^21\check{c}_{22},\check{c}_{21},\hat{d}_{21}. Viewing all vertices as sources of gradient flow, we check that ind(cˇ22)=0\mathrm{ind}(\check{c}_{22})=0, ind(cˇ21)=ind(d^21)=n1\mathrm{ind}(\check{c}_{21})=\mathrm{ind}(\hat{d}_{21})=n-1.

By Lemma 2.9, the dimension of this moduli space is 0 and it contains a unique gradient tree which corresponds to a single pseudoholomorphic triangle by Theorem 2.11. Note that this is almost the same as Example 2.12.

It remains to consider v(1)v^{(1)}, which is a pseudoholomorphic disk with a slit and a singular point inside. The singular point can be viewed as coming from a Lagrangian surgery [FOOO09], which increases the Fredholm index by n2n-2. Now the moduli space of v(1)v^{(1)} passing through a generic w1w_{1} is of dimension n1n-1. Consider the evaluation map ev21ev_{21} of v(1)v^{(1)} on its top-left end, where the image lies in a Sn1S^{n-1}-family of Reeb chords. The gluing condition on both ends says ev21ev_{21} should take the value of d^21\hat{d}_{21}. All we need is that ev21ev_{21} intersects d^21\hat{d}_{21} at a unique point. We will show in Type 1\partial_{1}, ev21ev_{21} sweeps half of Sn1S^{n-1} and the other half is dealt with by Type 2\partial_{2}. This is done by a model calculation of explicit pseudoholomorphic curves in Step 3’ below.


StepStep 3’. A model calculation. All notations are limited to this step.

We replace the base of v(1)v^{(1)} in Type 1\partial_{1} by a standard one, i.e., the unit disk in z\mathbb{C}_{z} with a slit {1Rez0}{Imz=0}\{-1\leq\mathrm{Re}\,z\leq 0\}\cap\{\mathrm{Im}\,z=0\}. The Lefschetz fibration over the unit disk is

(3.5) p:(z1,z2,,zn)z12+z22++zn2,p\colon(z_{1},z_{2},\dots,z_{n})\mapsto z_{1}^{2}+z_{2}^{2}+\dots+z_{n}^{2},

with a critical value at 0z0\in\mathbb{C}_{z}. It is however more convenient to think of the case n=2n=2 first, and the Lefschetz fibration is

(3.6) p:(z1,z2)z1z2,p^{\prime}\colon(z_{1}^{\prime},z_{2}^{\prime})\mapsto z_{1}^{\prime}z_{2}^{\prime},

where z1=z1+iz2,z2=z1iz2z_{1}^{\prime}=z_{1}+iz_{2},\,z_{2}^{\prime}=z_{1}-iz_{2}. Let TT be the Clifford torus {|z1|=1}×{|z2|=1}\{|z_{1}^{\prime}|=1\}\times\{|z_{2}^{\prime}|=1\} over |z|=1|z|=1 and let LL be the Lagrangian thimble over {1Rez0}{Imz=0}\{-1\leq\mathrm{Re}\,z\leq 0\}\cap\{\mathrm{Im}\,z=0\}. TLT\cap L is a clean S1S^{1}-intersection over 1z-1\in\mathbb{C}_{z}.

We consider curves with boundary on TLT\cup L: let 2\mathcal{M}_{2} (2 stands for n=2n=2) be the moduli space of holomorphic disks

u=(u1,u2):×[0,1]z1,z22u=(u^{\prime}_{1},u^{\prime}_{2})\colon\mathbb{R}\times[0,1]\to\mathbb{C}^{2}_{z_{1}^{\prime},z_{2}^{\prime}}

with standard complex structure JstdJ_{std} satisfying

  1. (1)

    u(×{0})Tu(\mathbb{R}\times\{0\})\subset T and u(×{1})Lu(\mathbb{R}\times\{1\})\subset L;

  2. (2)

    pup^{\prime}\circ u has degree 1 over {|z|<1}{1Rez0}{Imz=0}\{|z|<1\}-\{-1\leq\mathrm{Re}\,z\leq 0\}\cap\{\mathrm{Im}\,z=0\} and degree 0 otherwise;

  3. (3)

    u(0,0)=w1=(w11,w12)=(1,1)u(0,0)=w_{1}=(w_{11}^{\prime},w_{12}^{\prime})=(1,1).

Refer to caption
Figure 9. The pictorial description of 2\mathcal{M}_{2}. The left column describes the base z\mathbb{C}_{z}. The middle and right column are for two sets of coordinates. Dashed curves denote the unit circle. The red dots indicate ei(π+ϵ)e^{i(\pi+\epsilon)} on the left and for their preimages on the middle and right. The arrows indicate images of the arrow from (0,1)(0,1) to (0,0)(0,0) on the domain ×[0,1]\mathbb{R}\times[0,1]. Note that the drawings are not necessarily accurate.

Condition (3) is essentially the same as that v(1)v^{(1)} passes through w1w_{1} in Step 3. It is not hard to see that 2\mathcal{M}_{2} is homeomorphic to a line segment where 2\partial\mathcal{M}_{2} consists of two curves z(z,1)z\mapsto(z,1) and z(1,z)z\mapsto(1,z). Figure 9 gives a schematic description of 2\mathcal{M}_{2}, from the top row of z(z,1)z\mapsto(z,1) to the bottom row of z(1,z)z\mapsto(1,z), where the right-hand column changes the coordinates to (z1,z2)(z_{1},z_{2}).

We then consider the evaluation map. The top-left end of v(1)v^{(1)} in Type 1\partial_{1} is translated to ei(π+ϵ)ze^{i(\pi+\epsilon)}\in\mathbb{C}_{z} for small ϵ>0\epsilon>0. Define ev1:2S|z1|=11ev^{\prime}_{1}\colon\mathcal{M}_{2}\to S^{1}_{|z^{\prime}_{1}|=1} as the z1z^{\prime}_{1} projection of the intersection between uu and p1(ei(π+ϵ))p^{\prime-1}(e^{i(\pi+\epsilon)}), which is shown by red dots in the z1z^{\prime}_{1} column of Figure 9. Clearly ev1ev^{\prime}_{1} is a homeomorphism between 2\mathcal{M}_{2} and {eiθ1|π+ϵ<θ1<2π}\{e^{i\theta_{1}}|\pi+\epsilon<\theta_{1}<2\pi\}.

For general n2n\geq 2, define n\mathcal{M}_{n} similar to 2\mathcal{M}_{2}, where T,LT,L in condition (1) are still the Lagrangian vanishing cycles over the unit circle and {1Rez0}{Imz=0}\{-1\leq\mathrm{Re}\,z\leq 0\}\cap\{\mathrm{Im}\,z=0\}. Condition (3) is modified to

  1. (3’)

    u(0,0)=w1=(w11,w12,,w1n)=(1,0,,0)u(0,0)=w_{1}=(w_{11},w_{12},\dots,w_{1n})=(1,0,\dots,0).

The moduli space 2\mathcal{M}_{2} in coordinate (z1,z2)(z_{1},z_{2}) is viewed as the slice of n\mathcal{M}_{n} that restricts to 0 on z3,,znz_{3},\dots,z_{n}. Observe that for n\mathcal{M}_{n}, the coordinates z2,z3,,znz_{2},z_{3},\dots,z_{n} are symmetric. Thus we can recover n\mathcal{M}_{n} from 2\mathcal{M}_{2} by a symmetric rotation of z2z_{2}-coordinate. Each u=(z1,z2)2u=(z_{1},z_{2})\in\mathcal{M}_{2} corresponds to a Sn2S^{n-2}-family of curves in n\mathcal{M}_{n}:

uλ1,,λn1=(z1,λ1z2,,λn1z2),u_{\lambda_{1},\dots,\lambda_{n-1}}=(z_{1},\lambda_{1}z_{2},\dots,\lambda_{n-1}z_{2}),

where λ12++λn12=1\lambda_{1}^{2}+\dots+\lambda_{n-1}^{2}=1 and λi\lambda_{i}\in\mathbb{R} for i=1,,n1i=1,\dots,n-1. Figure 10 shows u1/n1,,1/n1u_{1/\sqrt{n-1},\dots,1/\sqrt{n-1}} recovered from the first row of Figure 9.

Refer to caption
Figure 10. u1/n1,,1/n1u_{1/\sqrt{n-1},\dots,1/\sqrt{n-1}} of the first row in Figure 9.

The new evaluation map evev is defined as the nn-tuple of coordinates which projects to ei(π+ϵ)ze^{i(\pi+\epsilon)}\in\mathbb{C}_{z}. Therefore, n\mathcal{M}_{n} is homeomorphic to Dn1D^{n-1}. One can check that ev:nSn1ev\colon\mathcal{M}_{n}\to S^{n-1} is a homeomorphism to its image, which is half of the vanishing cycle Sn1S^{n-1} over ei(π+ϵ)e^{i(\pi+\epsilon)}.

In case JstdJ_{std} is not regular, we apply small perturbation JJ^{\lozenge} of JstdJ_{std}. One can show that for any 𝒛evJstd(n)\boldsymbol{z}\in ev_{J_{std}}(\mathcal{M}_{n}), #evJ1(𝒛)=1\#ev^{-1}_{J^{\lozenge}}(\boldsymbol{z})=1 mod 22. Therefore the argument of Step 3 still works. This finishes Step 3’.


Finally we glue v(1),v(2)v^{(1)},v^{(2)} and v(3)v^{(3)}. Still assume we are in Type 1\partial_{1}. The involution condition (3.4) will fix the neck length: As we take ιv(b2)\iota\circ v(b_{2})\to\infty, q(Θ2)q(\Theta_{2}) approaches q(cˇ21)q(\check{c}_{21}) but |q(w2)q(Θ2)||q(w2)q(cˇ21)||q(w_{2})-q(\Theta_{2})|\ll|q(w_{2})-q(\check{c}_{21})|. Assume w1w_{1} is close to Θ1\Theta_{1}, there is a unique value of ιv(b2)\iota\circ v(b_{2}) for which there exists an involution of FF. This completes the proof of Theorem 3.4. ∎

3.2. Curve counting

The goal of this section is to count the full version of pseudoholomorphic annuli from Ξ1,Ξ2\Xi_{1},\Xi_{2} to Θ1,Θ2\Theta_{1},\Theta_{2} over the region in Figure 3 and prove the following theorem:

Theorem 3.5.

#Jχ=0,𝒘(𝚵,𝚯)=1\#\mathcal{M}^{\chi=0,\boldsymbol{w}}_{J^{\lozenge}}(\boldsymbol{\Xi},\boldsymbol{\Theta})=1 mod 22 for generic 𝐰\boldsymbol{w}.

Proof. The strategy is the same as the proof of Theorem 3.4: We stretch the curve into several levels by choosing some extreme 𝒘\boldsymbol{w}, and then use the restriction of domain involution and gluing conditions to find a unique (mod 22) curve.

We closely follow the proof of Theorem 9.3.7 of [CHT20] and some details are omitted. As before, we write u:F˙×[0,1]×W~u\colon\dot{F}\to\mathbb{R}\times[0,1]\times\widetilde{W} for an element in Jχ=0,𝒘(𝚵,𝚯)\mathcal{M}^{\chi=0,\boldsymbol{w}}_{J^{\lozenge}}(\boldsymbol{\Xi},\boldsymbol{\Theta}) and let vv be its projection to W~\widetilde{W}.

The main idea is to stretch the base D~\widetilde{D} in Imz\mathrm{Im}\,z direction as Figure 11: Let Imzi=2K\mathrm{Im}\,z_{i}=-2K and Imzi+κ=2K\mathrm{Im}\,z_{i+\kappa}=2K, i=1,,κi=1,\dots,\kappa and K+K\to+\infty. The region \mathcal{R} bounded by γ1,γ1\gamma_{1},\gamma_{1}^{\prime} is split into 3 parts: 1={ImzK}\mathcal{R}_{1}=\mathcal{R}\cap\{\mathrm{Im}\,z\leq-K\}, 2={KImzK}\mathcal{R}_{2}=\mathcal{R}\cap\{-K\leq\mathrm{Im}\,z\leq K\} and 3={ImzK}\mathcal{R}_{3}=\mathcal{R}\cap\{\mathrm{Im}\,z\geq K\}. The mod 22 count is independent of 𝒘\boldsymbol{w}. We choose w~1\widetilde{w}_{1} close to z1z_{1}, w~2\widetilde{w}_{2} close to Imz=0\mathrm{Im}\,z=0 and the thin strip \mathcal{R}^{\prime} between γ2,γ2\gamma_{2},\gamma_{2}^{\prime} with width m0m\to 0.

Refer to caption
Figure 11. The stretched base as K0K\gg 0. Imz\mathrm{Im}\,z is the horizontal direction.

The curve vv has degree 2 over \mathcal{R}^{\prime} and degree 1 over \mathcal{R}-\mathcal{R}^{\prime}. The types of branching behaviors are denoted by 22, 1L1L, 1R1R, 0LL0LL, 0LR0LR, 0RR0RR. Take 0LR0LR for example: 0 means the number of interior branch points is 0; LL means one pair of switch points is over γ~2\widetilde{\gamma}_{2}^{\prime}; RR means one pair of switch points is over γ~2\widetilde{\gamma}_{2}. Denote (if exist) interior branch points by b,bint(F˙)b,b^{\prime}\in\mathrm{int}(\dot{F}) and switch points by b1,b2,b3,b4F˙b_{1},b_{2},b_{3},b_{4}\in\partial\dot{F}. We assume ι(b)>ι(b)\iota(b^{\prime})>\iota(b), ι(b2)>ι(b1)\iota(b_{2})>\iota(b_{1}) and ι(b4)>ι(b3)\iota(b_{4})>\iota(b_{3}).


StepStep 11. Suppose K0K\gg 0. Take a sequence of u(i)Jχ=0,𝒘(𝚵,𝚯)u^{(i)}\in\mathcal{M}^{\chi=0,\boldsymbol{w}}_{J^{\lozenge}}(\boldsymbol{\Xi},\boldsymbol{\Theta}) so that v(i):F(i)W~v^{(i)}\colon F^{(i)}\to\widetilde{W} with m(i)0m^{(i)}\to 0. In the limit ii\to\infty, the thin strip tends to a slit, the limiting curve splits into vδ+δv_{\infty}\cup\delta_{+}\cup\delta_{-}, where δ+\delta_{+} is a gradient trajectory from Ξ2\Xi_{2} and δ\delta_{-} is a gradient trajectory to Θ2\Theta_{2}, and vv_{\infty} is a pseudoholomorphic annulus. Figure 12 describes the limiting procedure in the case of Type 0LR0LR. In fact we will show that Type 0LR0LR is the only nontrivial case.

Refer to caption
Figure 12. The limiting proceduce of Type 0LR0LR.

StepStep 22. We claim that the limiting slit is long enough, i.e., for K0K\gg 0 and m0m\to 0,

(3.7) max{ι(b),ι(b2),ι(b4)}K,\mathrm{max}\{\iota(b^{\prime}),\iota(b_{2}),\iota(b_{4})\}\geq K,
(3.8) min{ι(b),ι(b1),ι(b3)}K.\mathrm{min}\{\iota(b),\iota(b_{1}),\iota(b_{3})\}\leq-K.

which are the two endpoints of the slit.

If (3.8) is not true, i.e., min{ι(b),ι(b1),ι(b3)}>K\mathrm{min}\{\iota(b),\iota(b_{1}),\iota(b_{3})\}>-K, then the part of vv_{\infty} in region 1\mathcal{R}_{1} has no slit, which can be viewed as the outcome of a Lagrangian surgery on a trivial pseudoholomorphic disk. Similar to Step 2 in the proof of Theorem 3.4, the moduli space of pseudoholomorphic disks passing through a generic w1w_{1} is diffeomorphic to Sn2S^{n-2}, of which the evaluation map at the cylindrical end has a Sn2S^{n-2}-intersection with the Sn1S^{n-1}-family of Reeb chords. The evaluation map vanishes at homology level, which contributes 0 (mod 2) to the curve count. The argument for (3.7) is similar.  

StepStep 33. We claim that for K0K\gg 0 and mm small, if (3.7) and (3.8) hold, the mod 22 contribution of Type 2,1L,0LL,0RR2,1L,0LL,0RR is 0. The reason is that if one considers the involution condition

(3.9) q(Θ1)q(Θ2),q(Ξ1)q(Ξ2),q(w1)q(w2)q(\Theta_{1})\mapsto q(\Theta_{2}),\,q(\Xi_{1})\mapsto q(\Xi_{2}),\,q(w_{1})\mapsto q(w_{2})

for a pseudoholomorphic annulus, there is a constraint on the position of Θ2,Ξ2,w2\Theta_{2},\Xi_{2},w_{2} on the slit in vv_{\infty}. One can refer to [CHT20] for detailed discussion that all but Type 0LR0LR contradict with (3.9).


StepStep 44. It remains to consider Type 0LR0LR. Assuming (3.7) and (3.8) are satisfied, we claim that the contribution of Type 0LR0LR is 1 (mod 2).

Although there are other possible arrangements of b1,b2,b3,b4b_{1},b_{2},b_{3},b_{4} on the slit, we just consider the case in Figure 12 for illustration. As shown in Figure 13, vv_{\infty} is the gluing of two regions: vl,v_{l,\infty} with ImzK\mathrm{Im}\,z\ll-K and vr,v_{r,\infty} with ImzK\mathrm{Im}\,z\gg-K, which can be viewed as two pseudoholomorphic disks similar to Step 3’ of the previous section. The conditions that w1w_{1} is close to Θ1\Theta_{1} and that w2w_{2} sits on the slit will be translated to an explicit model calculation in Step 4’ below. For suitable choices of w1w_{1} and w2w_{2}, we will show that the space of two disks with c21=c21,c12=c12c_{21}=c^{*}_{21},\,c_{12}=c^{*}_{12} is homeomorphic to a line segment \mathcal{I}. Let MM_{\mathcal{I}} be the set of vv_{\infty} glued from the \mathcal{I}-family of (vl,,vr,)(v_{l,\infty},v_{r,\infty}).

Refer to caption
Figure 13. vv_{\infty} with a long slit. 1\mathcal{R}_{1} on the left and 23\mathcal{R}_{2}\cup\mathcal{R}_{3} on the right.

The involution condition (3.9) determines a unique curve inside MM_{\mathcal{I}}. In conclusion, #Jχ=0,𝒘(𝚵,𝚯)=1\#\mathcal{M}^{\chi=0,\boldsymbol{w}}_{J^{\lozenge}}(\boldsymbol{\Xi},\boldsymbol{\Theta})=1 mod 22. We have proved Theorem 3.5 modulo the model calculation below:


StepStep 4’. A model calculation. We use similar notations as in Step 3’ of the previous section. Define l\mathcal{M}_{l} for the space of vlv_{l} over the left side of Figure 13 and r\mathcal{M}_{r} for those of vrv_{r} over the right side. Both are viewed as maps from ×[0,1]\mathbb{R}\times[0,1] to n\mathbb{C}^{n} over the unit circle with one slit {1Rez0}{Imz=0}z\subset\{-1\leq\mathrm{Re}\,z\leq 0\}\cap\{\mathrm{Im}\,z=0\}\subset\mathbb{C}_{z}. Consider the evaluation maps

evl,evr:l,rSn1×Sn1,ev_{l},ev_{r}\colon\mathcal{M}_{l},\mathcal{M}_{r}\to S^{n-1}\times S^{n-1},

which are defined below, corresponding to c12,c21,c12,c21c_{12},c_{21},c^{*}_{12},c^{*}_{21} in Figure 13. The gluing condition is evl(vl)=evr(vr)ev_{l}(v_{l})=ev_{r}(v_{r}).

As usual, we first consider the case of n=2n=2 with Lefschetz fibration (3.6).

The moduli space l\mathcal{M}_{l} is defined as holomorphic disks passing through w1=(1,1)w_{1}=(1,1) over 1z1\in\mathbb{C}_{z}. Note that l\mathcal{M}_{l} is the same as \mathcal{M} in Step 3’ of the previous section. Let w±=ei(π±ϵ)zw_{\pm}=e^{i(\pi\pm\epsilon)}\in\mathbb{C}_{z}. Define the evaluation map evlev_{l} as

evl:lS1×S1,\displaystyle ev_{l}\colon\mathcal{M}_{l}\to S^{1}\times S^{1},
vl(evl+(vl),evl(vl)),\displaystyle v_{l}\mapsto(ev_{l+}(v_{l}),ev_{l-}(v_{l})),

where evl±(vl)ev_{l\pm}(v_{l}) is the z1z_{1}^{\prime}-coordinate of the point that projects to w±w_{\pm}. Figure 14 gives a schematic description of l\mathcal{M}_{l}, homeomorphic to a line segment, and its evaluation maps evl±ev_{l\pm} denoted by red and violet dots. We also show the maps in coordinates (z1,z2)(z_{1},z_{2}) with z1=zi+iz2,z2=ziiz2z_{1}^{\prime}=z_{i}+iz_{2},\,z_{2}^{\prime}=z_{i}-iz_{2}.

Refer to caption
Figure 14. The pictorial description of l\mathcal{M}_{l}. The left column describes the base z\mathbb{C}_{z}. The middle and right columns are for two sets of coordinates. The red and violet dots indicate ei(π±ϵ)e^{i(\pi\pm\epsilon)} on the left and their preimages on the middle and right.

Since w1=(1,0)w_{1}=(1,0) in coordinate (z1,z2)(z_{1},z_{2}), define w1=(1,0,0,,0)w_{1}=(1,0,0,\dots,0) for n2n\geq 2. Then l,2\mathcal{M}_{l,2}, the moduli space for n=2n=2, is viewed as the slice of l,n\mathcal{M}_{l,n} that restricts to 0 on z3,,znz_{3},\dots,z_{n}. Now l,n\mathcal{M}_{l,n} is a symmetric rotation of l,2\mathcal{M}_{l,2} which contains

v=(z1,λ1z2,,λn1z2),v^{\prime}=(z_{1},\lambda_{1}z_{2},\dots,\lambda_{n-1}z_{2}),

where λ12++λn12=1\lambda_{1}^{2}+\dots+\lambda_{n-1}^{2}=1, λ1,,λn1\lambda_{1},\dots,\lambda_{n-1}\in\mathbb{R}. Thus l,2\mathcal{M}_{l,2} is homeomorphic to Dn1D^{n-1}.

Next we consider r\mathcal{M}_{r} of v′′:×[0,1]nv^{\prime\prime}\colon\mathbb{R}\times[0,1]\to\mathbb{C}^{n} and first let n=2n=2. We put the constraint that vrv_{r} passes through (r,r)L(-r,r)\in L for some r[0,1]r\in[0,1], which corresponds to w2w_{2} sitting on the slit in Figure 13. Let evr±(vr)ev_{r\pm}(v_{r}) be the z1z_{1}^{\prime}-coordinate of the point that projects to w±w_{\pm} and define the map

evr:rS1×S1,\displaystyle ev_{r}\colon\mathcal{M}_{r}\to S^{1}\times S^{1},
vr(evr(vr),evr+(vr)),\displaystyle v_{r}\mapsto(ev_{r-}(v_{r}),ev_{r+}(v_{r})),

where ++ and - are switched because we want to identify w±w_{\pm} of vlv_{l} with ww_{\mp} of vrv_{r}.

Refer to caption
Figure 15. The pictorial description of part of r\mathcal{M}_{r}. The notations are as before, while the red and violet dots indicate ei(πϵ)e^{i(\pi\mp\epsilon)} on the left and their preimages on the middle and right.

Observe that r,2\mathcal{M}_{r,2} is of dimension 2. Figure 15 describes some of the curves inside r,2\mathcal{M}_{r,2}. (r,r)(-r,r) in (z1,z2)(z_{1}^{\prime},z_{2}^{\prime}) equals (0,ir)(0,ir) in (z1,z2)(z_{1},z_{2}). Thus for n2n\geq 2, if r,2\mathcal{M}_{r,2} is viewed as the slice of r,n\mathcal{M}_{r,n} with z3==zn=0z_{3}=\dots=z_{n}=0, then r,n\mathcal{M}_{r,n} contains curves of

v′′=(λ1z1,z2,λ2z1,,λn1z1),v^{\prime\prime}=(\lambda_{1}z_{1},z_{2},\lambda_{2}z_{1},\dots,\lambda_{n-1}z_{1}),

where λ12++λn12=1\lambda_{1}^{2}+\dots+\lambda_{n-1}^{2}=1, λ1,,λn1\lambda_{1},\dots,\lambda_{n-1}\in\mathbb{R}.

For n=2n=2, we have the following observation:

Claim 3.6.

For n=2n=2 and ϵ0\epsilon\to 0, the evaluation map of l\mathcal{M}_{l} and r\mathcal{M}_{r} with images in S1×S1S^{1}\times S^{1} is described in Figure 16. evl(l)ev_{l}(\mathcal{M}_{l}) is the blue line segment and evr(r)ev_{r}(\mathcal{M}_{r}) is the 2-dimensional pink region. Their intersection is a line segment \mathcal{I}.

Refer to caption
Figure 16. The description of evlev_{l} (blue) and evrev_{r} (pink) for n=2n=2. The sides are identified. aa-gg correspond to curves in Figure 15.

For general n2n\geq 2, we have the same result:

Lemma 3.7.

For n2n\geq 2 and ϵ0\epsilon\to 0, the intersection between evl(l)ev_{l}(\mathcal{M}_{l}) and evr(r)ev_{r}(\mathcal{M}_{r}) is still \mathcal{I}.

Proof of Lemma 3.7. Suppose vl,𝝀=(zl1,λ1zl2,,λn1zl2)v_{l,\boldsymbol{\lambda}}=(z_{l1},\lambda_{1}z_{l2},\dots,\lambda_{n-1}z_{l2}) and vr,𝝀=(λ1zr1,zr2,λ2zr1,,λn1zr1)v_{r,\boldsymbol{\lambda^{\prime}}}=(\lambda^{\prime}_{1}z_{r1},\linebreak z_{r2},\lambda^{\prime}_{2}z_{r1},\dots,\lambda^{\prime}_{n-1}z_{r1}) satisfy evl(vl,𝝀)=evr(vr,𝝀)ev_{l}(v_{l,\boldsymbol{\lambda}})=ev_{r}(v_{r,\boldsymbol{\lambda^{\prime}}}) where the 3rd to nn-th coordinates are not all zero. Denote evl±(vl,𝝀)=(zl1±,λ1zl2±,,λn1zl2±)ev_{l\pm}(v_{l,\boldsymbol{\lambda}})=(z_{l1\pm},\lambda_{1}z_{l2\pm},\dots,\lambda_{n-1}z_{l2\pm}) and evr(vr,𝝀)=(λ1zr1,zr2,λ2zr1,,λn1zr1)ev_{r\mp}(v_{r,\boldsymbol{\lambda^{\prime}}})=(\lambda^{\prime}_{1}z_{r1\mp},z_{r2\mp},\lambda^{\prime}_{2}z_{r1\mp},\dots,\lambda^{\prime}_{n-1}z_{r1\mp}).

Observe that Imλ1zl2+=Imλ1zl2\mathrm{Im}\,\lambda_{1}z_{l2+}=\mathrm{Im}\,\lambda_{1}z_{l2-} and then zr2=zr2+z_{r2-}=z_{r2+}. From cc and dd in Figure 15 we see that evr((zr1,zr2))ev_{r}((z_{r1},z_{r2})) (in coordinate (z1,z2)(z_{1}^{\prime},z_{2}^{\prime})) must lie in {θ1+θ2=2π}{πθ13π/2}\{\theta_{1}+\theta_{2}=2\pi\}\cap\{\pi\leq\theta_{1}\leq 3\pi/2\} of Figure 16. For such curves vv (as (a,b,ea,b,e) in Figure 15), Imzr1=Imzr1+\mathrm{Im}\,z_{r1-}=-\mathrm{Im}\,z_{r1+} and Rezr1=Rezr1+\mathrm{Re}\,z_{r1-}=\mathrm{Re}\,z_{r1+}. Since Rezl2+=Rezl2\mathrm{Re}\,z_{l2+}=-\mathrm{Re}\,z_{l2-} and Imzl2+=Imzl2\mathrm{Im}\,z_{l2+}=\mathrm{Im}\,z_{l2-}, the consequence is that λ2zl2±==λn1zl2±=0\lambda_{2}z_{l2\pm}=\dots=\lambda_{n-1}z_{l2\pm}=0, which is a contradiction. ∎


Now we go back to the curve counting problem. We want to pick a single curve from the \mathcal{I}-family of vv_{\infty} and then glue it to get a unique v(i)v^{(i)} for each ii.

First we show the position of b2b_{2}^{\infty} determines vv_{\infty} uniquely in \mathcal{I}: Consider the slit in vv_{\infty} of Figure 12. In the limit q(w1)=q(Θ1)q(w_{1})=q(\Theta_{1}), so q(w2)=q(Θ2)=q(b2)q(w_{2})=q(\Theta_{2})=q(b_{2}^{\infty}). If we fix b2b_{2}^{\infty}, then v′′v_{\infty}^{\prime\prime} passes through (r,r)(-r,r) for some r[0,1]r\in[0,1], corresponding to fixing a hypersurface in Sn1×Sn1S^{n-1}\times S^{n-1}, whose intersection with S1×S1S^{1}\times S^{1} is the dotted arc in Figure 16. The dotted arc intersects the blue line at a single point, which determines vv_{\infty}. Moreover, the length of the slit in vv_{\infty}^{\prime} and v′′v_{\infty}^{\prime\prime} are determined and thus b1b_{1}^{\infty} and b4b_{4}^{\infty} are fixed. Finally b3b_{3}^{\infty} is fixed by the involution of FF^{\infty}.

Consider then v(i)v^{(i)} for large ii. From the previous paragraph ι(b2(i))\iota(b_{2}^{(i)}) will fix a unique vv_{\infty} in \mathcal{I}. By Implicit Function Theorem, it will fix a unique v(i)v^{(i)} as well, which is close to vv_{\infty}. Then observe that the distance between q(Θ2)q(\Theta_{2}) and q(w2)q(w_{2}) is a monotone function of ι(b2(i))\iota(b_{2}^{(i)}): As ι(b2(i))\iota(b_{2}^{(i)}) increases, the slit gets longer, b1(i)b_{1}^{(i)} moves left and b4(i)b_{4}^{(i)} moves right. Therefore q(w2)q(w_{2}) leaves q(Θ2)q(\Theta_{2}) and approaches q(Ξ2)q(\Xi_{2}) on F(i)F^{(i)}. The involution of F(i)F^{(i)} determines a unique ι(b2(i))\iota(b_{2}^{(i)}) and thus a unique v(i)v^{(i)}.

This finishes the proof of Theorem 3.5. ∎

4. A model calculation of quadrilaterals

We make a model calculation which will be used in Section 5 and 6. Consider the trivial fibration p^:×TSn1\hat{p}\colon\mathbb{C}\times T^{*}S^{n-1}\to\mathbb{C} and Lagrangian submanifolds ai={y=i}×Sn1a_{i}=\{y=i\}\times S^{n-1}, i=1,2i=1,2, bj={x=j}×Sn1b_{j}=\{x=j\}\times S^{n-1}, j=1,2j=1,2, where Sn1S^{n-1} is the zero section of TSn1T^{*}S^{n-1}. We further modify ai,bja_{i},b_{j} to ai,bja^{\prime}_{i},b^{\prime}_{j} by a Hamiltonian perturbation in the fiber direction so that they intersect transversely. Specifically, we choose the restriction of Euclidean metric on Sn1S^{n-1} and identify TSn1T^{*}S^{n-1} with TSn1TS^{n-1}. Choose Morse functions f1,f2,f3,f4f_{1},f_{2},f_{3},f_{4} on Sn1S^{n-1} each with 2 critical points and all of the critical points are disjoint (as the right of Figure 17). We can then rescale these Morse functions so that the difference of each pair is still Morse with 2 critical points:

Refer to caption
Figure 17. The base \mathbb{C} on the left and the fiber TSn1T^{*}S^{n-1} on the right.
Lemma 4.1.

For small enough ϵ>0\epsilon>0, the difference of each pair of functions in {ϵ3f1,ϵ2f2,ϵf3,f4}\{\epsilon^{3}f_{1},\epsilon^{2}f_{2},\epsilon f_{3},f_{4}\} is Morse with 2 critical points.

Proof. For small enough ϵ>0\epsilon>0, f4ϵf3f_{4}-\epsilon f_{3} is a small perturbation of f4f_{4}. Since Morse condition is CC^{\infty}-stable, f4ϵf3f_{4}-\epsilon f_{3} is still Morse with 2 critical points. By a simple induction the proof is finished. ∎

Denote fa1=ϵ3f1,fa2=ϵ2f2,fb1=ϵf3,fb2=f4f_{a_{1}^{\prime}}=\epsilon^{3}f_{1},\,f_{a_{2}^{\prime}}=\epsilon^{2}f_{2},\,f_{b_{1}^{\prime}}=\epsilon f_{3},\,f_{b_{2}^{\prime}}=f_{4}, the gradients of which correspond to the fiber projection of a1,a2,b1,b2a_{1}^{\prime},a_{2}^{\prime},b_{1}^{\prime},b_{2}^{\prime}. Let xˇij,x^ij\check{x}_{ij},\hat{x}_{ij} over xijx_{ij} be the top and bottom critical points of fbjfaif_{b_{j}^{\prime}}-f_{a_{i}^{\prime}}, i,j{1,2}i,j\in\{1,2\}.

Now we compute the differentials of CF^(𝒃,𝒂)\widehat{CF}(\boldsymbol{b^{\prime}},\boldsymbol{a^{\prime}}), which is generated by 8 elements {x12,x21}\{x_{12}^{\dagger},x_{21}^{\dagger}\} and {x11,x22}\{x_{11}^{\dagger},x_{22}^{\dagger}\}, where {\dagger} denotes a check or hat.

Lemma 4.2.

The differential of CF^(𝐛,𝐚)\widehat{CF}(\boldsymbol{b^{\prime}},\boldsymbol{a^{\prime}}) is given by \hbar times the arrows in Figure 18. Moreover, a relative grading by Maslov index is denoted in Figure 18.

Refer to caption
Figure 18. The differentials of CF^(𝒃,𝒂)\widehat{CF}(\boldsymbol{b},\boldsymbol{a}). The generators in the top row have 2 checks, those in the middle row have 1 check and those in the bottom row have no check.
Proof.

Let u:F˙×[0,1]×(×TSn1)u\colon\dot{F}\to\mathbb{R}\times\left[0,1\right]\times(\mathbb{C}\times T^{*}S^{n-1}) be a pseudoholomorphic disk with positive ends {x12,x21}\{{x}^{{\dagger}}_{12},{x}^{{\dagger}}_{21}\} and negative ends {x11,x22}\{{x}^{{\dagger}}_{11},{x}^{{\dagger}}_{22}\}. Suppose the complex structure is split, then its projection to \mathbb{C} is a degree 1 map over [1,2]×[1,2][1,2]\times[1,2], which fixes the cross ratio of the 4 punctures on F˙\partial\dot{F}.

Then we consider the projection of uu to the fiber direction TSn1T^{*}S^{n-1}, denoted by w:F˙TSn1w\colon\dot{F}\to T^{*}S^{n-1}. By the construction above a1,a2,b1,b2a^{\prime}_{1},a^{\prime}_{2},b^{\prime}_{1},b^{\prime}_{2} are graphical near Sn1TSn1S^{n-1}\subset T^{*}S^{n-1}, with respect to Morse functions fai,fbjf_{a^{\prime}_{i}},f_{b^{\prime}_{j}}, i,j=1,2i,j={1,2}. The domain of the Morse moduli space is shown in Figure 19, where inner edges are ignored and arrows denote the direction of (frightfleft)-\nabla(f_{\mathrm{right}}-f_{\mathrm{left}}).

Refer to caption
Figure 19.

Viewing all boundary vertices as sources of gradient flow, observe that xˇij\check{x}_{ij} is of Morse index n1n-1 if iji\neq j and 0 if i=ji=j; x^ij\hat{x}_{ij} is of Morse index 0 if iji\neq j and n1n-1 if i=ji=j. By Lemma 2.9, we can further perturb fai,fbjf_{a^{\prime}_{i}},f_{b^{\prime}_{j}}, i,j=1,2i,j={1,2} such that all Morse gradient trees ww we consider are transversely cut out, and

(4.1) ind(w)=(#checksin{x12,x21}+#hatsin{x11,x22}3)(n1)+1.\mathrm{ind}(w)=\left(\#\mathrm{checks\,\,in\,\,}\{x^{{\dagger}}_{12},x^{{\dagger}}_{21}\}+\#\mathrm{hats\,\,in\,\,}\{x^{{\dagger}}_{11},x^{{\dagger}}_{22}\}-3\right)(n-1)+1.

Therefore, the Morse moduli space with respect to the arrows in Figure 18 is of ind(w)=1\mathrm{ind}(w)=1 and the case of gradient tree on Sn1S^{n-1} from {xˇ12,xˇ21}\{\check{x}_{12},\check{x}_{21}\} to {x^11,xˇ22}\{\hat{x}_{11},\check{x}_{22}\} is shown in Figure 20.

Refer to caption
Figure 20. Two possible gradient trees on Sn1S^{n-1}. The moduli space is parametrized by the length of the inner (blue) edge.

By Theorem 2.11, the moduli space of pseudoholomorphic curves is diffeomorphic to the Morse moduli space, so we can think of gradient trees instead of pseudoholomorphic disks. Taking the base direction into consideration, one checks that

(4.2) ind(u)=ind(w).\mathrm{ind}(u)=\mathrm{ind}(w).

For example, we still consider the case of Figure 20: The two gradient trees are parametrized by the length of their inner edges. As the inner length tends to zero, the left and right gradient trees tend to the same one. As the inner edge of the left one tends to the bottom generator of fb2fb1f_{b^{\prime}_{2}}-f_{b^{\prime}_{1}}, its length tends to infinity and q(xˇ21)q(\check{x}_{21}) approaches q(xˇ22)q(\check{x}_{22}). Similarly, as the inner edge of the right one tends to the top generator of fa2fa1f_{a^{\prime}_{2}}-f_{a^{\prime}_{1}}, its length tends to infinity and q(xˇ12)q(\check{x}_{12}) approaches q(x^22)q(\hat{x}_{22}). Since the cross ratio on the domain is fixed by the base direction, the result is that the algebraic count of ww is one. This verifies ind(u)=1\mathrm{ind}(u)=1 and the arrows in Figure 18.

If we set {xˇ12,xˇ21}\{\check{x}_{12},\check{x}_{21}\} to be of grading 0, we can verify the relative grading of generators in Figure 18 by Lemma 2.3, (4.1), (4.2) and the convention that ||=2n|\hbar|=2-n, where the difference of grading is given by the Maslov index. ∎

5. Invariance under Markov stabilization

A Markov stabilization is shown in Figure 21: σ\sigma is a κ\kappa-strand braid which intersects DD along 𝒛={z1,,zk}\boldsymbol{z}=\{z_{1},\dots,z_{k}\}. On the base DD, σ\sigma is viewed as an element of Diff+(D,D,z)\mathrm{Diff^{+}}(D,\partial D,z), which restricts to identity near γ0\gamma_{0}. Without loss of generality, we construct a positive Markov stabilization between γ0\gamma_{0} and γ1\gamma_{1}: Let cc be an arc from z0z_{0} to z1z_{1} which is disjoint from other γj\gamma_{j}, perform a positive half twist along cc, then we get a (κ+1)(\kappa+1)-strand braid given by σσc\sigma\circ\sigma_{c}.

Now we consider the fiber and Lagrangians. Let p:WDp^{\prime}\colon W^{\prime}\to D be the standard Lefschetz fibration with regular fiber TSn1T^{*}S^{n-1} and critical values 𝒛={z0,,zκ}\boldsymbol{z^{\prime}}=\{z_{0},\dots,z_{\kappa}\} and p:WDN(γ0)p\colon W\to D-N(\gamma_{0}) be its restriction to DN(γ0)D-N(\gamma_{0}). Let aja_{j} denote the Lagrangian thimble over γj\gamma_{j}. Let hσh_{\sigma} be an element of Symp(W,W)\mathrm{Symp}(W,\partial W) which descends to σ\sigma and hσSymp(W,W)h^{\prime}_{\sigma}\in\mathrm{Symp}(W^{\prime},\partial W^{\prime}) be its extension to WW^{\prime} by identity. Finally, let τcSymp(W,W)\tau_{c}\in\mathrm{Symp}(W^{\prime},\partial W^{\prime}) be the Dehn twist along the Lagrangian sphere over cc.

Refer to caption
Figure 21. Markov stabilization along cc.
Refer to caption
Figure 22. The red half-arcs are σσc(γ0)\sigma\circ\sigma_{c}(\gamma_{0}) and σσc(γ1)\sigma\circ\sigma_{c}(\gamma_{1}). The shaded region denotes the curve we are gluing.

The proof of invariance under Markov stabilization is the same as Theorem 9.4.2 of [CHT20], and we briefly restate its proof here:

Theorem 5.1.

CF^(W,hσ(𝒂),𝒂)\widehat{CF}(W,h_{\sigma}(\boldsymbol{a}),\boldsymbol{a}) and CF^(W,hστc(𝐚),𝐚)\widehat{CF}(W^{\prime},h^{\prime}_{\sigma}\circ\tau_{c}(\boldsymbol{a}^{\prime}),\boldsymbol{a}^{\prime}) are isomorphic cochain complexes for specific choices of almost complex structure and hσ(𝐚)h_{\sigma}(\boldsymbol{a}) and hστc(𝐚)h^{\prime}_{\sigma}\circ\tau_{c}(\boldsymbol{a}^{\prime}) after a Hamiltonian isotopy.

Proof.

We directly construct a homomorphism Φs:CF^(W,hσ(𝒂),𝒂)CF^(W,hστc(𝒂),𝒂)\Phi_{s}:\widehat{CF}(W,h_{\sigma}(\boldsymbol{a}),\boldsymbol{a})\to\widehat{CF}(W^{\prime},h^{\prime}_{\sigma}\circ\tau_{c}(\boldsymbol{a}^{\prime}),\boldsymbol{a}^{\prime}) and show it is a cochain isomorphism. The notations are as in Figure 22.

Consider the κ\kappa-tuples in CF^(W,hσ(𝒂),𝒂)\widehat{CF}(W,h_{\sigma}(\boldsymbol{a}),\boldsymbol{a}): It may or may not contain {x1}\{x_{1}\}. Similarly, the κ\kappa-tuples in CF^(W,hστc(𝒂),𝒂)\widehat{CF}(W^{\prime},h^{\prime}_{\sigma}\circ\tau_{c}(\boldsymbol{a}^{\prime}),\boldsymbol{a}^{\prime}) may contain either {x0,x1}\{x_{0},x_{1}\} or {Θ01}\{\Theta_{01}\}. In fact there is a linear isomorphism:

Φs:CF^(W,hσ(𝒂),𝒂)CF^(W,hστc(𝒂),𝒂),\displaystyle\Phi_{s}\colon\widehat{CF}(W,h_{\sigma}(\boldsymbol{a}),\boldsymbol{a})\to\widehat{CF}(W^{\prime},h^{\prime}_{\sigma}\circ\tau_{c}(\boldsymbol{a}^{\prime}),\boldsymbol{a}^{\prime}),
{x1}𝒚{x0,x1}𝒚,𝒚{Θ01}𝒚,\displaystyle\{x_{1}\}\cup\boldsymbol{y}^{\prime}\mapsto\{x_{0},x_{1}\}\cup\boldsymbol{y}^{\prime},\,\,\boldsymbol{y}\mapsto\{\Theta_{01}\}\cup\boldsymbol{y},

For convenience of gluing below, we put hσ(a1)h_{\sigma}(a_{1}) and hστc(a0)h^{\prime}_{\sigma}\circ\tau_{c}(a_{0}) in a position so that they go over the same arc near γ1\gamma_{1}. Let γ1={Rez1}×[1,0]D\gamma_{1}=\{\mathrm{Re}\,z_{1}\}\times[-1,0]\subset D and ϵ>0\epsilon>0 be small. The projections σ(γ1)\sigma(\gamma_{1}) and σσc(γ0)\sigma\circ\sigma_{c}(\gamma_{0}) are written as ζ1ζ2ζ3\zeta_{1}\cup\zeta_{2}\cup\zeta_{3} and ζ1ζ2ζ3\zeta^{\prime}_{1}\cup\zeta_{2}\cup\zeta_{3}. The main requirement is that ζ2\zeta_{2} be a common neck, for example, set ζ2={Rez1ϵ}×[2/3,1/3]\zeta_{2}=\{\mathrm{Re}\,z_{1}-\epsilon\}\times[-2/3,-1/3]. Refer to Figure 22 for details.

We compare the differential of CF^(W,hστc(𝒂),𝒂)\widehat{CF}(W^{\prime},h^{\prime}_{\sigma}\circ\tau_{c}(\boldsymbol{a}^{\prime}),\boldsymbol{a}^{\prime}) and CF^(W,hσ(𝒂),𝒂)\widehat{CF}(W,h_{\sigma}(\boldsymbol{a}),\boldsymbol{a}). If uu^{\prime} goes from {x0,x1}y1\{x_{0},x_{1}\}\cup{y_{1}^{\prime}} to {x0,x1}y2\{x_{0},x_{1}\}\cup{y_{2}^{\prime}}, then uu^{\prime} is in bijection with uu that goes from {x1}y1\{x_{1}\}\cup{y_{1}^{\prime}} to {x1}y2\{x_{1}\}\cup{y_{2}^{\prime}} since the strip from x0x_{0} to x0x_{0} is trivial. Similarly, uu^{\prime} that goes from {Θ01}y1\{\Theta_{01}\}\cup{y_{1}} to {Θ01}y2\{\Theta_{01}\}\cup{y_{2}} is in bijection with uu that goes from y1{y_{1}} to y2{y_{2}}. There are no curves from {x0,x1}y1\{x_{0},x_{1}\}\cup{y_{1}^{\prime}} to {Θ01}y2\{\Theta_{01}\}\cup{y_{2}} and no curves from {x1}y1\{x_{1}\}\cup{y_{1}^{\prime}} to y2{y_{2}}. The nontrivial case is that, if uu^{\prime} goes from {Θ01}y1\{\Theta_{01}\}\cup{y_{1}} to {x0,x1}y2\{x_{0},x_{1}\}\cup{y_{2}^{\prime}}, then it is in bijection with uu that goes from y1{y_{1}} to {x0}y2\{x_{0}\}\cup{y_{2}^{\prime}}, where uu^{\prime} comes from uu by replacing the end containing x1x_{1} by the shaded region in Figure 22. The bijection comes from Lemma 4.2 which says that the curve over the shaded region has algebraic count 1. The gluing details are omitted. ∎

6. Examples

In this section we consider some simple links and compute their cohomology groups Kh(σ^)Kh^{\sharp}(\widehat{\sigma}) in the sense of Theorem 2.4, with coefficient ring 𝔽[𝒜],1]\mathbb{F}[\mathcal{A}]\llbracket\hbar,\hbar^{-1}] when n=2n=2 and 𝔽,1]\mathbb{F}\llbracket\hbar,\hbar^{-1}] when n>3n>3, where \hbar is of degree 2n2-n. We assume that 𝔽\mathbb{F} is of characteristic 2 for simplicity.

6.1. Unknots

Figure 23 shows the 2-strand braid representation of an unknot. In the Morse-Bott family of Reeb chords, x1,x2x_{1},x_{2} are viewed as longer Reeb chords (top generators) and z1,z2z_{1},z_{2} are viewed as shorter Reeb chords (bottom generators). The only possible differential in CF^(W~,h~σunknot(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma_{\mathrm{unknot}}}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) counts quadrilaterals uu with {x1,x2}\{x_{1},x_{2}\} at positive ends and {z1,z2}\{z_{1},z_{2}\} at negative ends. The projection of uu to DD has degree 1 over the region bounded by the loop x1z2x2z1x1x_{1}\to z_{2}\to x_{2}\to z_{1}\to x_{1} and degree 0 over its complement.

Refer to caption
Figure 23. The braid representation of an unknot on DD.
Proposition 6.1.

Kh(σ^unknot)Kh^{\sharp}(\widehat{\sigma}_{\mathrm{unknot}}) is freely generated by {x1,x2}\{x_{1},x_{2}\} and {z1,z2}\{z_{1},z_{2}\}, where the difference of grading between these two generators is 22 (mod n2n-2):

generators grading
{x1,x2}\{x_{1},x_{2}\} 0
{z1,z2}\{z_{1},z_{2}\} 22
Proof.

Similar to the proof of Lemma 4.2, we check that the pseudoholomorphic disk uu with {x1,x2}\{x_{1},x_{2}\} at positive ends and {z1,z2}\{z_{1},z_{2}\} at negative ends is of Fredholm index nn and Maslov index 2n22n-2. Therefore, {x1,x2}\{x_{1},x_{2}\} and {z1,z2}\{z_{1},z_{2}\} are both cocycles and hence generators of Kh(σ^unknot)Kh^{\sharp}(\widehat{\sigma}_{\mathrm{unknot}}). ∎

6.2. Hopf links

We then consider the 2-strand braid representation of a left-handed Hopf link as the left side of Figure 24. ‘AA’, ‘BB’ and ‘CC’ denote the corresponding regions on the base DD. For example, we use ‘A+BA+B’ to represent the union of region `A`A\textrm{'} and `B`B\textrm{'}.

Refer to caption
Figure 24. The braid representation of the left-handed Hopf link (left) and the right-handed Hopf link (right).
Lemma 6.2.

CF^(W~,h~σleftHopf(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma_{\mathrm{left\,Hopf}}}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) contains the following (mod 2) differential relations:

(6.1) d\displaystyle d {x1,z2}={y^1,yˇ2}+{yˇ1,y^2},\displaystyle\{x_{1},z_{2}\}=\hbar\{\hat{y}_{1},\check{y}_{2}\}+\hbar\{\check{y}_{1},\hat{y}_{2}\},
(6.2) d\displaystyle d {x2,z1}={y^1,yˇ2}+{yˇ1,y^2},\displaystyle\{x_{2},z_{1}\}=\hbar\{\hat{y}_{1},\check{y}_{2}\}+\hbar\{\check{y}_{1},\hat{y}_{2}\},
(6.3) d\displaystyle d {yˇ1,y^2}={z1,z2},\displaystyle\{\check{y}_{1},\hat{y}_{2}\}=\hbar\{z_{1},z_{2}\},
(6.4) d\displaystyle d {y^1,yˇ2}={z1,z2}.\displaystyle\{\hat{y}_{1},\check{y}_{2}\}=\hbar\{z_{1},z_{2}\}.
Proof.

Note that x1,x2,z1,z2x_{1},x_{2},z_{1},z_{2} should be viewed as top generators at positive ends or as bottom generators at negative ends. Specifically they give no constraints on the Morse-Bott family.

All nontrivial index 1 curves with count 1 (mod 2) are listed as follows, where we label the regions with positive weights after projection to the base DD:

regions ends of generators
AA {x1,z2}{yˇ1,y^2}/{y^1,yˇ2}\{x_{1},z_{2}\}\to\{\check{y}_{1},\hat{y}_{2}\}/\{\hat{y}_{1},\check{y}_{2}\}
BB {x2,z1}{yˇ1,y^2}/{y^1,yˇ2}\{x_{2},z_{1}\}\to\{\check{y}_{1},\hat{y}_{2}\}/\{\hat{y}_{1},\check{y}_{2}\}
CC {yˇ1,y^1}/{y^1,yˇ1}{z1,z2}\{\check{y}_{1},\hat{y}_{1}\}/\{\hat{y}_{1},\check{y}_{1}\}\to\{z_{1},z_{2}\}

To see this, after projection to the base, the domain with positive weights is bounded by one of the following loops:

  1. (1)

    y1x1y2z2y1y_{1}\to x_{1}\to y_{2}\to z_{2}\to y_{1}. This corresponds to region ‘AA’, where uu is a quadrilateral with {x1,z2}\{x_{1},z_{2}\} at positive ends and {yˇ1,y^2}\{\check{y}_{1},\hat{y}_{2}\} or {y^1,yˇ2}\{\hat{y}_{1},\check{y}_{2}\} at negative ends. By Lemma 4.2, ind(u)=1\mathrm{ind}(u)=1 and we get (6.1).

  2. (2)

    y1z1y2x2y1y_{1}\to z_{1}\to y_{2}\to x_{2}\to y_{1}. This corresponds to region ‘BB’, where uu is a quadrilateral with {x2,z1}\{x_{2},z_{1}\} at positive ends and {yˇ1,y^2}\{\check{y}_{1},\hat{y}_{2}\} or {y^1,yˇ2}\{\hat{y}_{1},\check{y}_{2}\} at negative ends. This is similar to (1). Therefore, ind(u)=1\mathrm{ind}(u)=1 and we get (6.2).

  3. (3)

    y1z2y2z1y1y_{1}\to z_{2}\to y_{2}\to z_{1}\to y_{1}. This corresponds to region ‘CC’, where uu is a quadrilateral with {z1,z2}\{z_{1},z_{2}\} at negative ends. Since z1,z2z_{1},z_{2} are bottom generators, we need one check and one hat at positive ends due to Lemma 4.2 so that there is a nontrivial count of uu. Therefore we get (6.3) and (6.4).

  4. (4)

    y1x1y2x2y1y_{1}\to x_{1}\to y_{2}\to x_{2}\to y_{1}. This corresponds to region ‘A+B+CA+B+C’, where uu is a quadrilateral with {x1,x2}\{x_{1},x_{2}\} at positive ends and {y1,y2}\{{y}_{1},{y}_{2}\} at negative ends. x1,x2x_{1},x_{2} are viewed as top generators and there are two critical points inside the domain. We check that ind(u)=n\mathrm{ind}(u)=n for {yˇ1,yˇ2}\{\check{y}_{1},\check{y}_{2}\} at negative ends; ind(u)=2n1\mathrm{ind}(u)=2n-1 for {y^1,yˇ2}\{\hat{y}_{1},\check{y}_{2}\} and {yˇ1,y^2}\{\check{y}_{1},\hat{y}_{2}\} at negative ends; ind(u)=3n2\mathrm{ind}(u)=3n-2 for {y^1,y^2}\{\hat{y}_{1},\hat{y}_{2}\} at negative ends. Therefore there is no such uu with index 1.

Corollary 6.3.

If we set |{x1,x2}|=0|\{x_{1},x_{2}\}|=0, then Kh(σ^leftHopf)Kh^{\sharp}(\widehat{\sigma}_{\mathrm{left\,Hopf}}) is freely generated by the following generators with the corresponding relative grading (mod n2n-2):

generators grading
{x1,x2}\{x_{1},x_{2}\} 0
{x2,z1}+{x1,z2}\{x_{2},z_{1}\}+\{x_{1},z_{2}\} 22
{yˇ1,yˇ2}\{\check{y}_{1},\check{y}_{2}\} 22
{y^1,y^2}\{\hat{y}_{1},\hat{y}_{2}\} 44

The computation for the right-handed Hopf link is similar. As shown in Figure 24, the braid representation of the right-handed Hopf link is the mirror of the left-handed one. Note that there is a bijection of curves between these two Hopf links: Each curve in the left-handed moduli space corresponds to a curve in the right-handed one, where the positive and negative ends are exchanged, as well as the checks and hats. As a consequence, the grading by Maslov indices is also reversed. Specifically, the right-handed Hopf link satisfies:

Lemma 6.4.

CF^(W~,h~σrightHopf(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma_{\mathrm{right\,Hopf}}}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) contains the following (mod 2) differential relations:

d\displaystyle d {yˇ1,y^2}={x1,z2}+{x2,z1},\displaystyle\{\check{y}_{1},\hat{y}_{2}\}=\hbar\{x_{1},z_{2}\}+\hbar\{x_{2},z_{1}\},
d\displaystyle d {y^1,yˇ2}={x1,z2}+{x2,z1},\displaystyle\{\hat{y}_{1},\check{y}_{2}\}=\hbar\{x_{1},z_{2}\}+\hbar\{x_{2},z_{1}\},
d\displaystyle d {z1,z2}={yˇ1,y^2}+{y^1,yˇ2}.\displaystyle\{z_{1},z_{2}\}=\hbar\{\check{y}_{1},\hat{y}_{2}\}+\hbar\{\hat{y}_{1},\check{y}_{2}\}.
Corollary 6.5.

If we set |{x1,x2}|=0|\{x_{1},x_{2}\}|=0, then Kh(σ^rightHopf)Kh^{\sharp}(\widehat{\sigma}_{\mathrm{right\,Hopf}}) is freely generated by the following generators with the corresponding relative grading (mod n2n-2):

generators grading
{x1,x2}\{x_{1},x_{2}\} 0
{x1,z2}\{x_{1},z_{2}\} 2-2
{y^1,y^2}\{\hat{y}_{1},\hat{y}_{2}\} 2-2
{yˇ1,yˇ2}\{\check{y}_{1},\check{y}_{2}\} 4-4

6.3. Trefoils

The left side of Figure 25 shows the 2-strand braid representation of a left-handed trefoil, where `A`A\textrm{'} to `E`E\textrm{'} denote the corresponding regions on the base.

Refer to caption
Figure 25. The braid representation of the left-handed trefoil (left) and the right-handed trefoil (right).

The curve counting for trefoils is more interesting than unknots and Hopf links. We first do a model calculation. Recall the notations in Step 3’ of the proof of Theorem 3.4. We consider curves with boundary on TLT\cup L: Let 2\mathcal{M}^{\prime}_{2} (2 stands for n=2n=2) be the moduli space of holomorphic disks

u:×[0,1]z1,z22u\colon\mathbb{R}\times[0,1]\to\mathbb{C}^{2}_{z_{1}^{\prime},z_{2}^{\prime}}

with standard complex structure JstdJ_{std} satisfying

  1. (1)

    u(×{0})Tu(\mathbb{R}\times\{0\})\subset T and u(×{1})Lu(\mathbb{R}\times\{1\})\subset L;

  2. (2)

    pup^{\prime}\circ u has degree 1 over {|z|<1}{1Rez0}{Imz=0}\{|z|<1\}-\{-1\leq\mathrm{Re}\,z\leq 0\}\cap\{\mathrm{Im}\,z=0\} and degree 0 otherwise;

  3. (3)

    u(,0)=w1=(w11,w12)=(1,1)u(-\infty,0)=w_{1}=(w_{11}^{\prime},w_{12}^{\prime})=(-1,1).

It is easy to see that 2\mathcal{M}^{\prime}_{2} is homeomorphic to a line segment where 2\partial\mathcal{M}^{\prime}_{2} consists of two curves z(z,1)z\mapsto(z,1) and z(1,z)z\mapsto(-1,-z). Figure 26 gives a schematic description of 2\mathcal{M}^{\prime}_{2} in the coordinates (z1,z2)(z_{1}^{\prime},z_{2}^{\prime}). For a given θ(π,π)\theta\in(-\pi,\pi), define the evaluation map

evθ:2S|z1|=11ev^{\prime}_{\theta}\colon\mathcal{M}^{\prime}_{2}\to S^{1}_{|z_{1}^{\prime}|=1}

as the z1z^{\prime}_{1} projection of the intersection between uu and p1(eiθ)p^{\prime-1}(e^{i\theta}), which is shown by violet dots in the z1z^{\prime}_{1} column of Figure 26.

Refer to caption
Figure 26. The left column denotes the base, where the red dot indicates 1-1, and the violet dot indicates eiθe^{i\theta}.

Note that (1,1)(-1,1) in (z1,z2)(z_{1}^{\prime},z_{2}^{\prime})-coordinates equals (0,i)(0,i) in (z1,z2)(z_{1},z_{2})-coordinates. For general n2n\geq 2, define n\mathcal{M}^{\prime}_{n} similar to 2\mathcal{M}^{\prime}_{2}, where T,LT,L in condition (1) are still the Lagrangian vanishing cycles over the unit circle and {1Rez0}{Imz=0}\{-1\leq\mathrm{Re}\,z\leq 0\}\cap\{\mathrm{Im}\,z=0\}. Condition (3) is modified to

  1. (3’)

    u(,0)=w1=(w11,w12,,w1n)=(0,i,0,,0)u(-\infty,0)=w_{1}=(w_{11},w_{12},\dots,w_{1n})=(0,i,0,\dots,0).

The moduli space 2\mathcal{M}^{\prime}_{2} in coordinate (z1,z2)(z_{1},z_{2}) is viewed as the slice of n\mathcal{M}^{\prime}_{n} that restricts to 0 on z3,,znz_{3},\dots,z_{n}. Observe that for n\mathcal{M}^{\prime}_{n}, the coordinates z1,z3,,znz_{1},z_{3},\dots,z_{n} are symmetric. Thus we can recover n\mathcal{M}^{\prime}_{n} from 2\mathcal{M}^{\prime}_{2} by a symmetric rotation of z1z_{1}-coordinate. Each u=(z1,z2)2u=(z_{1},z_{2})\in\mathcal{M}^{\prime}_{2} corresponds to a Sn2S^{n-2}-family of curves in n\mathcal{M}^{\prime}_{n}:

uλ1,,λn1=(λ1z1,z2,λ2z1,,λn1z1),u_{\lambda_{1},\dots,\lambda_{n-1}}=(\lambda_{1}z_{1},z_{2},\lambda_{2}z_{1},\dots,\lambda_{n-1}z_{1}),

where λ12++λn12=1\lambda_{1}^{2}+\dots+\lambda_{n-1}^{2}=1 and λi\lambda_{i}\in\mathbb{R} for i=1,,n1i=1,\dots,n-1.

The new evaluation map evθev_{\theta} is defined as the nn-tuple of coordinates which projects to eiθze^{i\theta}\in\mathbb{C}_{z}. Therefore, n\mathcal{M}^{\prime}_{n} is homeomorphic to Dn1D^{n-1}. One can check that evθ:nSn1ev_{\theta}\colon\mathcal{M}^{\prime}_{n}\to S^{n-1} is a homeomorphism to its image. As θ\theta increases from π-\pi to π\pi, the ratio of area of evθ(n)Sn1ev_{\theta}(\mathcal{M}^{\prime}_{n})\subset S^{n-1} increases from 0 to 1.

Lemma 6.6.

CF^(W~,h~σlefttrefoil(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma_{\mathrm{left\,trefoil}}}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) contains the following (mod 2) differential relations:

d\displaystyle d {x1,wˇ2}={yˇ1,y^2}+{y^1,yˇ2},\displaystyle\{x_{1},\check{w}_{2}\}=\hbar\{\check{y}_{1},\hat{y}_{2}\}+\hbar\{\hat{y}_{1},\check{y}_{2}\},
d\displaystyle d {x1,w^2}={y^1,y^2}+{yˇ2,z1}+{yˇ1,z2},\displaystyle\{x_{1},\hat{w}_{2}\}=\hbar\{\hat{y}_{1},\hat{y}_{2}\}+\hbar\{\check{y}_{2},z_{1}\}+\hbar\{\check{y}_{1},z_{2}\},
d\displaystyle d {x2,wˇ1}={yˇ1,y^2}+{y^1,yˇ2},\displaystyle\{x_{2},\check{w}_{1}\}=\hbar\{\check{y}_{1},\hat{y}_{2}\}+\hbar\{\hat{y}_{1},\check{y}_{2}\},
d\displaystyle d {x2,w^1}={y^1,y^2}+{yˇ1,z2}+{yˇ2,z1},\displaystyle\{x_{2},\hat{w}_{1}\}=\hbar\{\hat{y}_{1},\hat{y}_{2}\}+\hbar\{\check{y}_{1},z_{2}\}+\hbar\{\check{y}_{2},z_{1}\},
d\displaystyle d {yˇ1,z2}={wˇ1,w^2}+{w^1,wˇ2},\displaystyle\{\check{y}_{1},z_{2}\}=\hbar\{\check{w}_{1},\hat{w}_{2}\}+\hbar\{\hat{w}_{1},\check{w}_{2}\},
d\displaystyle d {y^1,z2}={w^1,w^2},\displaystyle\{\hat{y}_{1},z_{2}\}=\hbar\{\hat{w}_{1},\hat{w}_{2}\},
d\displaystyle d {yˇ2,z1}={wˇ1,w^2}+{w^1,wˇ2},\displaystyle\{\check{y}_{2},z_{1}\}=\hbar\{\check{w}_{1},\hat{w}_{2}\}+\hbar\{\hat{w}_{1},\check{w}_{2}\},
d\displaystyle d {y^2,z1}={w^1,w^2},\displaystyle\{\hat{y}_{2},z_{1}\}=\hbar\{\hat{w}_{1},\hat{w}_{2}\},
d\displaystyle d {wˇ1,w^2}={z1,z2},\displaystyle\{\check{w}_{1},\hat{w}_{2}\}=\hbar\{z_{1},z_{2}\},
d\displaystyle d {w^1,wˇ2}={z1,z2},\displaystyle\{\hat{w}_{1},\check{w}_{2}\}=\hbar\{z_{1},z_{2}\},
d\displaystyle d {yˇ1,y^2}={wˇ1,wˇ2},\displaystyle\{\check{y}_{1},\hat{y}_{2}\}=\hbar\{\check{w}_{1},\check{w}_{2}\},
d\displaystyle d {y^1,yˇ2}={wˇ1,wˇ2}.\displaystyle\{\hat{y}_{1},\check{y}_{2}\}=\hbar\{\check{w}_{1},\check{w}_{2}\}.
Proof.

All nontrivial index 1 curves with count 1 (mod 2) are listed as follows:

regions ends of generators
AA {x1,wˇ2}{yˇ1,y^2}/{y^1,yˇ2}\{x_{1},\check{w}_{2}\}\to\{\check{y}_{1},\hat{y}_{2}\}/\{\hat{y}_{1},\check{y}_{2}\}, {x1,w^2}{y^1,y^2}\{x_{1},\hat{w}_{2}\}\to\{\hat{y}_{1},\hat{y}_{2}\}
BB {x2,wˇ1}{yˇ1,y^2}/{y^1,yˇ2}\{x_{2},\check{w}_{1}\}\to\{\check{y}_{1},\hat{y}_{2}\}/\{\hat{y}_{1},\check{y}_{2}\}, {x2,w^1}{y^1,y^2}\{x_{2},\hat{w}_{1}\}\to\{\hat{y}_{1},\hat{y}_{2}\}
CC {yˇ1,z2}{wˇ1,w^2}/{w^1,wˇ2}\{\check{y}_{1},z_{2}\}\to\{\check{w}_{1},\hat{w}_{2}\}/\{\hat{w}_{1},\check{w}_{2}\}, {y^1,z2}{w^1,w^2}\{\hat{y}_{1},z_{2}\}\to\{\hat{w}_{1},\hat{w}_{2}\}
DD {yˇ2,z1}{wˇ1,w^2}/{w^1,wˇ2}\{\check{y}_{2},z_{1}\}\to\{\check{w}_{1},\hat{w}_{2}\}/\{\hat{w}_{1},\check{w}_{2}\}, {y^2,z1}{w^1,w^2}\{\hat{y}_{2},z_{1}\}\to\{\hat{w}_{1},\hat{w}_{2}\}
EE {wˇ1,w^2}{z1,z2}\{\check{w}_{1},\hat{w}_{2}\}\to\{z_{1},z_{2}\}, {w^1,wˇ2}{z1,z2}\{\hat{w}_{1},\check{w}_{2}\}\to\{z_{1},z_{2}\}
A+C+EA+C+E {x1,w^2}{yˇ2,z1}\{x_{1},\hat{w}_{2}\}\to\{\check{y}_{2},z_{1}\}
B+C+EB+C+E {x2,w^1}{yˇ2,z1}\{x_{2},\hat{w}_{1}\}\to\{\check{y}_{2},z_{1}\}
A+D+EA+D+E {x1,w^2}{yˇ1,z2}\{x_{1},\hat{w}_{2}\}\to\{\check{y}_{1},z_{2}\}
B+D+EB+D+E {x2,w^1}{yˇ1,z2}\{x_{2},\hat{w}_{1}\}\to\{\check{y}_{1},z_{2}\}
C+D+EC+D+E {yˇ1,y^2}/{y^1,yˇ2}{wˇ1,wˇ2}\{\check{y}_{1},\hat{y}_{2}\}/\{\hat{y}_{1},\check{y}_{2}\}\to\{\check{w}_{1},\check{w}_{2}\}

First consider the region ‘B+D+EB+D+E’. The generators w^1,yˇ1\hat{w}_{1},\check{y}_{1} are viewed as point constraints and x2,z2x_{2},z_{2} impose no constraints on the Morse-Bott family. There are two possible slits: going down from w1{w}_{1} or going to the right from w1{w}_{1}. If the slit goes down, we can apply the model of Figure 26, where w^1\hat{w}_{1} corresponds to the red dot (0,i,0,,0)(0,i,0,\dots,0) and yˇ1\check{y}_{1} corresponds to the violet dot over eiθe^{i\theta} for some θ(π,π)\theta\in(-\pi,\pi). If w1,y1{w}_{1},{y}_{1} are close on the base, then θ\theta is close to π\pi. Therefore, evθ(n)Sn1ev_{\theta}(\mathcal{M}^{\prime}_{n})\subset S^{n-1} takes almost all of Sn1S^{n-1} and hence #evθ1(yˇ1)=1\#ev^{-1}_{\theta}(\check{y}_{1})=1 mod 22. Note that the disparity between θ\theta and π\pi prevents w^1,yˇ1\hat{w}_{1},\check{y}_{1} from being close on the fiber direction (with respect to the symplectic parallel transport away from z1,z2z_{1},z_{2}), which contributes to the complement Sn1\evθ(n)S^{n-1}\backslash ev_{\theta}(\mathcal{M}^{\prime}_{n}). However, even if w1,y1{w}_{1},{y}_{1} are not close on the base, the count remains the same: As the slit goes to the right from w1w_{1}, w^1,yˇ1\hat{w}_{1},\check{y}_{1} are getting closer on the fiber direction, and this part takes care of the complement Sn1\evθ(n)S^{n-1}\backslash ev_{\theta}(\mathcal{M}^{\prime}_{n}). The case of ‘A+C+EA+C+E’, ‘B+C+EB+C+E’ and ‘A+D+EA+D+E’ are similar.

Refer to caption
Figure 27. Horizontal stretch of region C+D+EC+D+E.

Next we consider the region ‘C+D+EC+D+E’. We stretch the region ‘C+D+EC+D+E’ as depicted in Figure 27. The dashed line is viewed as a gluing condition of a point constraint of both sides. There are two different cases:

  1. (1)

    {y^1,yˇ2}{wˇ1,wˇ2}\{\hat{y}_{1},\check{y}_{2}\}\to\{\check{w}_{1},\check{w}_{2}\}. On the left side of Figure 27, y^1\hat{y}_{1} and wˇ1\check{w}_{1} are point constraints, which fixes a unique holomorphic disk by similar reasons as case ‘B+D+EB+D+E’. Note that the slit going to the left from w1w_{1} cannot be too long (across the dashed line): If the slit is long, then it cuts the left part into a trivial quadrilateral and a disk region with a branch point z1z_{1} inside. However, the (mod 2) count of a holomorphic disk with a branch point inside is 0.

    The left disk then fixes the point constraint of the dashed line. The right part of Figure 27 now has two point constraints: wˇ2\check{w}_{2} and the dashed line. We then apply the model of Figure 26 again, where wˇ2\check{w}_{2} corresponds to the red dot and the dashed line corresponds to the violet dot. Similar to the discussion of ‘B+D+EB+D+E’, there exists a unique holomorphic disk. Note that the slit going to the left from w2w_{2} cannot be too long for the same reason as above. To conclude, the (mod 2) count is 1.

    The case of {yˇ1,y^2}{wˇ1,wˇ2}\{\check{y}_{1},\hat{y}_{2}\}\to\{\check{w}_{1},\check{w}_{2}\} is similar and the (mod 2) count is 1.

  2. (2)

    {y^1,y^2}{wˇ1,w^2}\{\hat{y}_{1},\hat{y}_{2}\}\to\{\check{w}_{1},\hat{w}_{2}\}. Similar to (1), y^1\hat{y}_{1} and wˇ1\check{w}_{1} fix a unique holomorphic disk on the left side of Figure 27 and gives the dashed line a point constraint. The dashed line and y^2\hat{y}_{2} are two point constraints on the right side. If the slit from w2w_{2} goes down, we apply the model of Figure 26 where y^2\hat{y}_{2} corresponds to the red dot and the dashed line corresponds to the violet dot. Note that evθ(n)ev_{\theta}(\mathcal{M}^{\prime}_{n}) sweeps approximately 1/41/4 of Sn1S^{n-1}. One can check that the slit going to the left from w2w_{2} also sweeps 1/41/4 of Sn1S^{n-1} (by symplectic parallel transport), and it cancels out with the contribution from evθ(n)ev_{\theta}(\mathcal{M}^{\prime}_{n}). As a result, the (mod 2) count is 0.

    The case of {y^1,y^2}{w^1,wˇ2}\{\hat{y}_{1},\hat{y}_{2}\}\to\{\hat{w}_{1},\check{w}_{2}\} is similar and the (mod 2) count is 0.

The remaining cases are similar to the discussion of Hopf links and we omit the details.

Corollary 6.7.

If we set |{x1,x2}|=0|\{x_{1},x_{2}\}|=0, then Kh(σ^lefttrefoil)Kh^{\sharp}(\widehat{\sigma}_{\mathrm{left\,trefoil}}) is freely generated by the following generators with the corresponding relative grading (mod n2n-2):

generators grading
{x1,x2}\{x_{1},x_{2}\} 0
{x1,wˇ2}+{x2,wˇ1}\{x_{1},\check{w}_{2}\}+\{x_{2},\check{w}_{1}\} 22
{yˇ1,yˇ2}\{\check{y}_{1},\check{y}_{2}\} 22
{x1,w^2}+{x2,w^1}\{x_{1},\hat{w}_{2}\}+\{x_{2},\hat{w}_{1}\} 33
{yˇ1,z2}+{yˇ2,z1}\{\check{y}_{1},z_{2}\}+\{\check{y}_{2},z_{1}\} 44
{y^1,z2}+{y^2,z1}\{\hat{y}_{1},z_{2}\}+\{\hat{y}_{2},z_{1}\} 55

The computation for the right-handed trefoil is similar and the proof is omitted:

Lemma 6.8.

CF^(W~,h~σrighttrefoil(𝒂~),𝒂~)\widehat{CF}(\widetilde{W},\widetilde{h}_{\sigma_{\mathrm{right\,trefoil}}}(\widetilde{\boldsymbol{a}}),\widetilde{\boldsymbol{a}}) contains the following (mod 2) differential relations:

d\displaystyle d {yˇ1,y^2}={x1,w^2}+{x2,w^1},\displaystyle\{\check{y}_{1},\hat{y}_{2}\}=\hbar\{x_{1},\hat{w}_{2}\}+\hbar\{x_{2},\hat{w}_{1}\},
d\displaystyle d {y^1,yˇ2}={x1,w^2}+{x2,w^1},\displaystyle\{\hat{y}_{1},\check{y}_{2}\}=\hbar\{x_{1},\hat{w}_{2}\}+\hbar\{x_{2},\hat{w}_{1}\},
d\displaystyle d {yˇ1,yˇ2}={x1,wˇ2}+{x2,wˇ1},\displaystyle\{\check{y}_{1},\check{y}_{2}\}=\hbar\{x_{1},\check{w}_{2}\}+\hbar\{x_{2},\check{w}_{1}\},
d\displaystyle d {wˇ1,w^2}={y^1,z2}+{y^2,z1},\displaystyle\{\check{w}_{1},\hat{w}_{2}\}=\hbar\{\hat{y}_{1},z_{2}\}+\hbar\{\hat{y}_{2},z_{1}\},
d\displaystyle d {w^1,wˇ2}={y^1,z2}+{y^2,z1},\displaystyle\{\hat{w}_{1},\check{w}_{2}\}=\hbar\{\hat{y}_{1},z_{2}\}+\hbar\{\hat{y}_{2},z_{1}\},
d\displaystyle d {wˇ1,wˇ2}={yˇ1,z2}+{yˇ2,z1},\displaystyle\{\check{w}_{1},\check{w}_{2}\}=\hbar\{\check{y}_{1},z_{2}\}+\hbar\{\check{y}_{2},z_{1}\},
d\displaystyle d {z1,z2}={wˇ1,w^2}+{w^1,wˇ2},\displaystyle\{z_{1},z_{2}\}=\hbar\{\check{w}_{1},\hat{w}_{2}\}+\hbar\{\hat{w}_{1},\check{w}_{2}\},
d\displaystyle d {y^1,z2}={x1,wˇ2}+{x2,wˇ1},\displaystyle\{\hat{y}_{1},z_{2}\}=\hbar\{x_{1},\check{w}_{2}\}+\hbar\{x_{2},\check{w}_{1}\},
d\displaystyle d {y^2,z1}={x1,wˇ2}+{x2,wˇ1},\displaystyle\{\hat{y}_{2},z_{1}\}=\hbar\{x_{1},\check{w}_{2}\}+\hbar\{x_{2},\check{w}_{1}\},
d\displaystyle d {w^1,w^2}={yˇ1,y^2}+{y^1,yˇ2}.\displaystyle\{\hat{w}_{1},\hat{w}_{2}\}=\hbar\{\check{y}_{1},\hat{y}_{2}\}+\hbar\{\hat{y}_{1},\check{y}_{2}\}.
Corollary 6.9.

If we set |{x1,x2}|=0|\{x_{1},x_{2}\}|=0, then Kh(σ^righttrefoil)Kh^{\sharp}(\widehat{\sigma}_{\mathrm{right\,trefoil}}) is freely generated by the following generators with the corresponding relative grading (mod n2n-2):

generators grading
{x1,x2}\{x_{1},x_{2}\} 0
{x1,w^2}\{x_{1},\hat{w}_{2}\} 2-2
{y^1,y^2}\{\hat{y}_{1},\hat{y}_{2}\} 2-2
{x1,wˇ2}\{x_{1},\check{w}_{2}\} 3-3
{y^1,z2}+{y^2,z1}\{\hat{y}_{1},z_{2}\}+\{\hat{y}_{2},z_{1}\} 4-4
{yˇ1,z2}\{\check{y}_{1},z_{2}\} 5-5

References

  • [CGH12] Vincent Colin, Paolo Ghiggini and Ko Honda “The equivalence of Heegaard Floer homology and embedded contact homology via open book decompositions I”, 2012 arXiv:1208.1074 [math.GT]
  • [CHT20] Vincent Colin, Ko Honda and Yin Tian “Applications of higher-dimensional Heegaard Floer homology to contact topology”, 2020 arXiv:2006.05701 [math.SG]
  • [EENS13] Tobias Ekholm, John B Etnyre, Lenhard Ng and Michael G Sullivan “Knot contact homology” In Geometry & Topology 17.2 Mathematical Sciences Publishers, 2013, pp. 975–1112
  • [Flo88] Andreas Floer “Morse theory for Lagrangian intersections” In J. Differential Geom. 28.3, 1988, pp. 513–547 URL: http://projecteuclid.org/euclid.jdg/1214442477
  • [FO97] Kenji Fukaya and Yong-Geun Oh “Zero-loop open strings in the cotangent bundle and Morse homotopy” In Asian J. Math. 1.1, 1997, pp. 96–180 DOI: 10.4310/AJM.1997.v1.n1.a5
  • [FOOO09] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta and Kaoru Ono “Lagrangian intersection Floer theory: anomaly and obstruction. Part II” 46, AMS/IP Studies in Advanced Mathematics American Mathematical Society, Providence, RI, 2009, pp. i–xii and 397–805 DOI: 10.1090/crmp/049/07
  • [Iac08] Vito Iacovino “A simple proof of a theorem of Fukaya and Oh”, 2008 arXiv:0812.0129 [math.SG]
  • [Kho00] Mikhail Khovanov “A categorification of the Jones polynomial” In Duke Math. J. 101.3, 2000, pp. 359–426 DOI: 10.1215/S0012-7094-00-10131-7
  • [Lip06] Robert Lipshitz “A cylindrical reformulation of Heegaard Floer homology” In Geom. Topol. 10, 2006, pp. 955–1096 DOI: 10.2140/gt.2006.10.955
  • [Man06] Ciprian Manolescu “Nilpotent slices, Hilbert schemes, and the Jones polynomial” In Duke Math. J. 132.2, 2006, pp. 311–369 DOI: 10.1215/S0012-7094-06-13224-6
  • [Man07] Ciprian Manolescu “Link homology theories from symplectic geometry” In Adv. Math. 211.1, 2007, pp. 363–416 DOI: 10.1016/j.aim.2006.09.007
  • [MS19] Cheuk Yu Mak and Ivan Smith “Fukaya-Seidel categories of Hilbert schemes and parabolic category 𝒪\mathcal{O}”, 2019 arXiv:1907.07624 [math.SG]
  • [OS04] Peter Ozsváth and Zoltán Szabó “Holomorphic disks and topological invariants for closed three-manifolds” In Ann. of Math. (2) 159.3, 2004, pp. 1027–1158 DOI: 10.4007/annals.2004.159.1027
  • [Per08] Timothy Perutz “Hamiltonian handleslides for Heegaard Floer homology”, 2008 arXiv:0801.0564 [math.SG]
  • [SS06] Paul Seidel and Ivan Smith “A link invariant from the symplectic geometry of nilpotent slices” In Duke Math. J. 134.3, 2006, pp. 453–514 DOI: 10.1215/S0012-7094-06-13432-4
  • [Wil08] Brandon Williams “Computations in Khovanov Homology” In University of Illinois at Chicago, Chicago, Illinois, 2008