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

Fiber Floer cohomology and conormal stops

Johan Asplund Department of Mathematics, Uppsala University, 751 06 Uppsala, Sweden johan.asplund@math.uu.se
Abstract.

Let SS be a closed orientable spin manifold. Let KSK\subset S be a submanifold and denote its complement by MKM_{K}. In this paper we prove that there exists an isomorphism between partially wrapped Floer cochains of a cotangent fiber stopped by the unit conormal ΛK\varLambda_{K} and chains of a Morse theoretic model of the based loop space of MKM_{K}, which intertwines the AA_{\infty}-structure with the Pontryagin product. As an application, we restrict to codimension 2 spheres KSnK\subset S^{n} where n=5n=5 or n7n\geq 7. Then we show that there is a family of knots KK so that the partially wrapped Floer cohomology of a cotangent fiber is related to the Alexander invariant of KK. A consequence of this relation is that the link ΛKΛx\varLambda_{K}\cup\varLambda_{x} is not Legendrian isotopic to ΛunknotΛx\varLambda_{\mathrm{unknot}}\cup\varLambda_{x} where xMKx\in M_{K}.

1. Introduction

In this paper we consider the wrapped Floer cohomology of a cotangent fiber with wrapping stopped by a conormal. We relate it to chains of based loops on the complement of a submanifold. Then we show that the Legendrian conormal knows about the smooth topology of the submanifold beyond the fundamental group.

Let SS be a closed orientable spin manifold. Let KSK\subset S be a submanifold and denote its complement by MKM_{K}. Consider the disk cotangent bundle DTSDT^{\ast}S equipped with the canonical Liouville form λ=pdq\lambda=pdq. The ideal contact boundary of the Weinstein domain DTSDT^{\ast}S is the unit cotangent bundle STSST^{\ast}S. Associated to KK are the conormal bundle

LK={(q,p)TS|qK,p,TqK=0}DTS,L_{K}=\left\{(q,p)\in T^{\ast}S\;|\;q\in K,\,\langle p,T_{q}K\rangle=0\right\}\subset DT^{\ast}S\,,

and the unit conormal ΛK=LKSTS\varLambda_{K}=L_{K}\cap ST^{\ast}S. Consider a cotangent fiber F=DTξSF=DT^{\ast}_{\xi}S at ξMK\xi\in M_{K} and let CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) be the partially wrapped Floer cochains on FF with wrapping stopped by ΛK\varLambda_{K}. Let BMKBM_{K} denote the space of piecewise geodesic loops in MKM_{K} based at ξ\xi. Consider the space Ccell(BMK)C^{\text{cell}}_{-\ast}(BM_{K}) of cellular chains of BMKBM_{K} equipped with the Pontryagin product. Then we have the following result:

Theorem 1.1 (4.12 and 5.3).

There exists a geometrically defined isomorphism of AA_{\infty}-algebras Ψ:CWΛK(F,F)Ccell(BMK)\varPsi\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow C^{\text{cell}}_{-\ast}(BM_{K}).

Moreover, Ψ\varPsi induces an isomorphism HWΛK(F,F)H(ΩξMK)HW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow H_{-\ast}(\varOmega_{\xi}M_{K}) of [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-modules.

We define CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) using a surgery approach similar to [EL17, Appendix B] and [ENS16, Section 6] (see Section 3.1 for details). The outline of the surgery approach is the following. We attach a handle modeled on DεT([0,)×ΛK)D_{\varepsilon}T^{\ast}([0,\infty)\times\varLambda_{K}) to DTSDT^{\ast}S along a neighborhood of ΛK\varLambda_{K}. We denote the resulting Liouville sector by WKW_{K} (with terminology as in [GPS20]). Then CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) is the wrapped Floer cochain complex of FF in WKW_{K}. The skeleton of WKW_{K} is LKSL_{K}\cup S with clean intersection LKS=KL_{K}\cap S=K. By performing Lagrangian surgery along the clean intersection, we obtain an exact Lagrangian submanifold MKWKM_{K}\subset W_{K} which is diffeomorphic to the complement SKS\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}K (see Section 3.1 for details).

Let ΩξMK\varOmega_{\xi}M_{K} denote the space of loops in MKM_{K} based at ξ\xi. Consider singular chains on the space of based loops C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K}). We give it the structure of an AA_{\infty}-algebra by equipping it with the Pontryagin product and all higher products equal to zero. See Section 3.2 and Section 4.2 for a more detailed discussion about the model of the based loop space we use.

In the spirit of Cieliebak–Latschev [CL09] and Abouzaid [Abo12b], we have a geometrically defined AA_{\infty}-homomorphism Ψ:CWΛK(F,F)C(ΩξMK)\varPsi\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow C_{-\ast}(\varOmega_{\xi}M_{K}). By analyzing the action filtrations, we show that Ψ\varPsi is diagonal with respect to the action filtrations. A key point in proving that Ψ\varPsi is an isomorphism is showing that the disks contributing to the diagonal are transversely cut out. The solutions of the linearized Floer equation are precisely those vector fields along the disk that restricts to broken Jacobi fields along γ\gamma on which the Hessian of the energy functional is negative definite.

In the surgery approach we attach a handle modeled on DεT([0,)×ΛK)D_{\varepsilon}T^{\ast}([0,\infty)\times\varLambda_{K}), with skeleton [0,)×ΛK[0,\infty)\times\varLambda_{K}. We consider a generic product metric on [0,)×ΛK[0,\infty)\times\varLambda_{K} such that the metric on ΛK\varLambda_{K} is scaled by a positive function with strictly negative derivative (warped product metric), see (A.1). By the genericity of the metric, there is a natural one-to-one correspondence between Reeb chords and geodesics, see 4.9 for details.

Since WKW_{K} and MKM_{K} are non-compact we use monotonicity of JJ-holomorphic curves to prove that relevant moduli spaces of JJ-holomorphic curves are compact, see Appendix A for details.

1.1. Applications

Let QQ be a smooth manifold and let KQK\subset Q be a submanifold. Consider the cotangent bundle TQT^{\ast}Q and the unit conormal bundle ΛK\varLambda_{K}. It is known in certain cases that the symplectic topology of TQT^{\ast}Q knows about the smooth topology of QQ [Abo12a, ES16, EKS16]. In some cases the contact topology of ΛK\varLambda_{K} knows about the smooth topology of KK. For instance, it is known that conormal tori ΛKST3\varLambda_{K}\subset ST^{\ast}\mathbb{R}^{3} of knots K3K\subset\mathbb{R}^{3} are complete knot invariants [She16, ENS16]. The results of Ekholm–Ng–Shende fit nicely into the broader picture of partially wrapped Floer cohomology that we consider in this paper, and is summarized in [ENS16, Section 1.3]. Specifically, in [ENS16] it is proven that there is a ring isomorphism

HWΛK0(F,F)[π1(MK)],HW^{0}_{\varLambda_{K}}(F,F)\cong\mathbb{Z}[\pi_{1}(M_{K})]\,,

which is also obtained from 1.1 by restricting to degree 0. Furthermore there is a relation between the knot contact homology of K3K\subset\mathbb{R}^{3} and the Alexander polynomial of KK [Ng08, ENS16].

Let KSnK\subset S^{n} be a codimension 2 sphere. In this paper we show that the partially wrapped Floer cohomology of the fiber is related to the Alexander invariant. The Alexander invariant is H(M~K)H_{\ast}(\widetilde{M}_{K}) regarded as a [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-module, where M~K\widetilde{M}_{K} denotes the infinite cyclic cover of MKM_{K}, see Section 5.4 for details. Denote by Λunknot\varLambda_{\mathrm{unknot}} the unit conormal of the standard embedded Sn2SnS^{n-2}\subset S^{n}. As an application of 1.1 we have the following theorem.

Theorem 1.2 (5.9).

Let n=5n=5 or n7n\geq 7. Let xMKx\in M_{K} be a point. Then there exists a codimension 2 knot KSnK\subset S^{n} with π1(MK)\pi_{1}(M_{K})\cong\mathbb{Z}, such that ΛKΛx\varLambda_{K}\cup\varLambda_{x} is not Legendrian isotopic to ΛunknotΛx\varLambda_{\mathrm{unknot}}\cup\varLambda_{x}.

1.2. Relation to other results

Let QQ be a closed smooth manifold and consider the exact symplectic manifold (TQ,dλ)(T^{\ast}Q,d\lambda) where λ\lambda is the canonical Liouville form λ=pdq\lambda=pdq. Abbondandolo–Schwarz proved that the wrapped Floer cohomology of a cotangent fiber TξQT_{\xi}^{\ast}Q is isomorphic to the homology of the based loop space of QQ [AS06]. Abouzaid extended this to an AA_{\infty}-quasi-isomorphism in [Abo12b] where the loop space is equipped with the Pontryagin product. Recently, Ganatra–Pardon–Shende proved that this result continues to hold even when QQ is not assumed to be compact as a consequence of a deeper relationship between the wrapped Fukaya category of a Liouville sector and a certain category of sheaves [GPS18a].

In this paper, we consider a similar JJ-holomorphic curve setup to the one used by Abouzaid in [Abo12b], but instead we work in the context of the partially wrapped Fukaya category of TST^{\ast}S stopped by the unit conormal ΛK\varLambda_{K}.

Remark 1.3.

Another interesting geometric point of view which motivates 1.1 is the following. Consider the wrapped Fukaya category of TMKT^{\ast}M_{K} [GPS20, BKO19]. By [GPS18a, Corollary 6.1] we have an AA_{\infty}-quasi-isomorphism

CW(F,F)C(ΩξMK),CW^{\ast}(F,F)\cong C_{-\ast}(\varOmega_{\xi}M_{K})\,,

where FTMKF\subset T^{\ast}M_{K} is the cotangent fiber at ξMK\xi\in M_{K}. We realize WKW_{K} as the result of attaching a handle to TMKT^{\ast}M_{K} as follows: Take a tubular neighborhood N(K)SN(K)\subset S of KK and consider N(K) . . =N(K)MKN^{\prime}(K)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=N(K)\cap M_{K}. Then remove LN(K)TMKL_{N^{\prime}(K)}\subset T^{\ast}M_{K} and replace it with LN(K)L_{N(K)}, identifying their common boundaries ΛN(K)=ΛN(K)\varLambda_{N(K)}=\varLambda_{N^{\prime}(K)}.

Refer to caption
Figure 1. The figure shows the construction of WKW_{K} via handle attachment on TMKT^{\ast}M_{K}.

From the point of view of handle attachment, there is a new generator of the wrapped Fukaya category, namely the cocore disk CC. Because of this, the wrapped Floer cohomology of the fiber FF will change on the level of chains. However, if we push CC very far out in the punctured handle by a Lagrangian isotopy, we look at filtered AA_{\infty}-algebras and yield a chain isomorphism

LCW(F,F)WKLCW(F,F)TMK,\mathcal{F}_{L}CW^{\ast}(F,F)_{W_{K}}\cong\mathcal{F}_{L}CW^{\ast}(F,F)_{T^{\ast}M_{K}}\,,

where L\mathcal{F}_{L} means we only consider generators of action less than LL. A standard filtration argument then shows that the wrapped Floer cohomology of FF is unaffected by this type of handle attachment and thus HW(F,F)WKHW(F,F)TMKHW^{\ast}(F,F)_{W_{K}}\cong HW^{\ast}(F,F)_{T^{\ast}M_{K}}. Hence we obtain an indirect proof of the isomorphism

HWΛK(F,F)=HW(F,F)WKHW(F,F)TMKH(ΩMK)HW^{\ast}_{\varLambda_{K}}(F,F)=HW^{\ast}(F,F)_{W_{K}}\cong HW^{\ast}(F,F)_{T^{\ast}M_{K}}\cong H_{-\ast}(\varOmega M_{K})

in 1.1.

1.3. Organization of the paper

In Section 2 we describe the version of wrapped Floer cohomology defined without Hamiltonian perturbations which we use in this paper. In Section 3 we first discuss the surgery approach to define partially wrapped Floer cohomology. Then we define the operations Ψ={Ψm}k=1\varPsi=\left\{\varPsi_{m}\right\}_{k=1}^{\infty} between CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K}) and show that Ψ\varPsi is an AA_{\infty}-homomorphism. Section 4 is devoted to proving that Ψ\varPsi is a isomorphism between CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and the Morse theoretic model of chains of based loops. Lastly, in Section 5 we equip CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K}) with [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-module structures relate HWΛK(F,F)HW^{\ast}_{\varLambda_{K}}(F,F) to the Alexander invariant H(M~K)H_{\ast}(\widetilde{M}_{K}) for certain families of codimension 2 knots KSnK\subset S^{n}. Then we show that this relation is used to show that ΛKΛx\varLambda_{K}\cup\varLambda_{x} is not Legendrian isotopic to ΛunknotΛx\varLambda_{\mathrm{unknot}}\cup\varLambda_{x}, where xSnKx\in S^{n}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}K is a point.

Acknowledgments

The author would like to thank his PhD advisor Tobias Ekholm for all his guidance and helpful discussions. He would also like to thank Georgios Dimitroglou Rizell for useful discussions regarding 1.3. Finally, the author would like to thank the anonymous referee whose many comments has improved the exposition of the paper. The author was supported by the Knut and Alice Wallenberg Foundation.

2. Wrapped Floer cohomology without Hamiltonian

In this paper, we consider a version of wrapped Floer cohomology defined without Hamiltonian perturbations. Wrapped Floer cohomology without Hamiltonian has been studied in e.g. [Ekh12, DR16, EL17] and in particular it is useful in proving various surgery formulas involving the wrapped Floer cohomology [BEE12, EL17, Ekh19]. It has also been used to study knots via knot contact homology from which there is a relationship to string topology and the cord algebra [ENS16, EENS13, CELN17].

Remark 2.1.

The relationship between wrapped Floer cohomology defined with and without Hamiltonians has also been studied. The version without Hamiltonian is known to be quasi-isomorphic to the version defined with Hamiltonians by counting strips with a Hamiltonian term that is turned on as one goes from the positive end to the negative end [EHK16, Theorem 7.2]. Such JJ-holomorphic maps with a Hamiltonian term that turns on has been more systematically studied in [EO17] and it is proven in [EL17, Lemma 68, 69] that the two versions of wrapped Floer cohomology are AA_{\infty}-quasi-isomorphic.

When working with wrapped Floer cohomology without Hamiltonian we have a priori bubbling issues. This is circumvented by considering parallel copies, which also removes the possibility of having multiply covered curves, see [EL17, Section 3.3]. Furthermore we need to count anchored curves [BEE12, Section 2.2] [EL17, Section A.1]. A specific perturbation scheme involving anchored curves is constructed in [Ekh19], and we fix such perturbation scheme so that all relevant moduli spaces are transversely cut out.

We give a brief description of the wrapped Floer cohomology without Hamiltonian by following [EL17, Appendix A-B]. We consider a Weinstein domain MM together with a smooth exact Lagrangian submaniold (M,ω . . =dλ,L)(M,\omega\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=d\lambda,L). Let Y . . =MY\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\partial M and Λ . . =LY\varLambda\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=L\cap Y be its Legendrian boundary. The boundary (Y,α . . =λ|Y)(Y,\alpha\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mathinner{\lambda\rvert}_{Y}) is a contact manifold. We consider the completion of MM and LL by attaching cylindrical ends [0,)×Y[0,\infty)\times Y to YY and [0,)×Λ[0,\infty)\times\varLambda to Λ\varLambda. Then we pick a system of parallel copies of LL as in [EL17, Section 3.3]. Consider a family (Hk,hk)k=1(H_{k},h_{k})_{k=1}^{\infty} of pairs of Morse functions, Hk:LH_{k}\colon\thinspace L\longrightarrow\mathbb{R} and hk:Λh_{k}\colon\thinspace\varLambda\longrightarrow\mathbb{R}. Let LkL_{k} be the time-1 flow of LL of the Hamiltonian vector field XHkX_{H_{k}}, and let Λk . . =LkY\varLambda_{k}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=L_{k}\cap Y. Then we call {Lk}k=0\left\{L_{k}\right\}_{k=0}^{\infty} a system of parallel copies of LL where L0 . . =LL_{0}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=L. Let L¯ . . =k=0Lk\overline{L}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\bigcup_{k=0}^{\infty}L_{k} and Λ¯ . . =k=0Λk\overline{\varLambda}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\bigcup_{k=0}^{\infty}\varLambda_{k}.

Note that in this paper, LL is a cotangent fiber. Therefore we choose the Morse functions HkH_{k} in such a way that all of them only have one minimum, since LDnL\cong D^{n}.

2.1. AA_{\infty}-structure and moduli space of disks

Let (M,λ)(M,\lambda) be a spin Weinstein domain. Let LML\subset M be an orientable exact Lagrangian with vanishing Maslov class (see [Arn67] for a definition of the Maslov class). Let L¯={Lk}k=0\overline{L}=\left\{L_{k}\right\}_{k=0}^{\infty} be the corresponding system of parallel copies of LL as in the previous section.

First we define CW(L,L)CW^{\ast}(L,L) as a \mathbb{Z}-graded module over \mathbb{Z}. Note that, for each Reeb chord cc^{\prime} starting at Λi\varLambda_{i} and ending at Λj\varLambda_{j}, there is a unique Reeb chord cc of Λ\varLambda close to cc^{\prime}. Similarly, for each transverse intersection point aa^{\prime} in LiLjL_{i}\cap L_{j}, there is a unique transverse intersection point aL0L1a\in L_{0}\cap L_{1}. We implicitly fix an identification of cc^{\prime} with cc, and aa^{\prime} with aa. We then define CW(L,L)CW^{\ast}(L,L) to be the \mathbb{Z}-graded module over \mathbb{Z}, which is generated by Reeb chords of Λ\varLambda and intersection points L0L1L_{0}\cap L_{1}. The grading is given by the Maslov index (see 2.3 below for a more precise definition).

We now describe how we equip CW(L,L)CW^{\ast}(L,L) with a AA_{\infty}-structure {μi}i=1\left\{\mu^{i}\right\}_{i=1}^{\infty} which is defined by JJ-holomorphic curve counts. Let DmD_{m}\subset\mathbb{C} denote the positively oriented unit disk, with mm points along the boundary removed. We denote the boundary punctures in DmD_{m} by ζ1,,ζm\zeta_{1},\ldots,\zeta_{m}, one of which is distinguished. These boundary punctures subdivide the boundary of DmD_{m} into mm arcs. We enumerate these arcs by κ1,,κm\kappa_{1},\ldots,\kappa_{m}, according to the boundary orientation, starting from the distinguished boundary puncture. We call κ . . ={κi}i=1m\kappa\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{\kappa_{i}\right\}_{i=1}^{m} a boundary numbering of DmD_{m}. If the sequence κ\kappa is decreasing (increasing), we say that the disk DmD_{m} has decreasing (increasing) boundary numbering κ\kappa. If κi1κi\kappa_{i-1}\leq\kappa_{i} (κi1κi\kappa_{i-1}\geq\kappa_{i}), we say that the puncture ζi\zeta_{i} is increasing (decreasing), and if κi1=κi\kappa_{i-1}=\kappa_{i} we say that ζi\zeta_{i} is a constant puncture.

We equip the boundary punctures ζjDm\zeta_{j}\in\partial D_{m} with both a positive and a negative strip-like end. Namely, we pick biholomorphisms

{ε+i:(0,)×[0,1]N(ζi)εi:(,0)×[0,1]N(ζi)i{1,,m},\begin{cases}\varepsilon_{+}^{i}\colon\thinspace(0,\infty)\times[0,1]\longrightarrow N(\zeta_{i})\\ \varepsilon_{-}^{i}\colon\thinspace(-\infty,0)\times[0,1]\longrightarrow N(\zeta_{i})\end{cases}\forall i\in\left\{1,\ldots,m\right\}\,,

where N(ζi)N(\zeta_{i}) is a neighborhood of the boundary puncture ζiDm\zeta_{i}\in\partial D_{m}.

Using notation as in [EL17], we are interested in the moduli spaces of JJ-holomorphic disks which are denoted by fi(𝒄;κ)\mathcal{M}^{\text{fi}}(\boldsymbol{c};\kappa), sy(𝒄;κ)\mathcal{M}^{\text{sy}}(\boldsymbol{c};\kappa) and pb(𝒄;κ)\mathcal{M}^{\text{pb}}(\boldsymbol{c};\kappa). These moduli spaces consist of filling disks, symplectization disks and partial holomorphic buildings respectively, and we define them below.

Filling disks:

Consider DmD_{m} equipped with a strictly decreasing boundary numbering. Note that every puncture is strictly decreasing except for the distinguished puncture, which is strictly increasing. We let 𝒄=c1cm\boldsymbol{c}=c_{1}\cdots c_{m} be a word of generators of CW(L,L)CW^{\ast}(L,L). Then we define fi(𝒄;κ)\mathcal{M}^{\text{fi}}(\boldsymbol{c};\kappa) to be the moduli space of JJ-holomorphic maps u:(Dm,Dm)(M,L¯)u\colon\thinspace(D_{m},\partial D_{m})\longrightarrow(M,\overline{L}) such that

  • •:

    near the boundary puncture ζi\zeta_{i}, uu is asymptotic to the generator cic_{i}, that is

    {lims±u(ε±i(s,t))=ci, if ci is an intersection generatorlims±u(ε±i(s,t))=(,ci), if ci is a Reeb chord generator.\begin{cases}\lim_{s\to\pm\infty}u(\varepsilon_{\pm}^{i}(s,t))=c_{i},&\text{ if }c_{i}\text{ is an intersection generator}\\ \lim_{s\to\pm\infty}u(\varepsilon_{\pm}^{i}(s,t))=(\infty,c_{i}),&\text{ if }c_{i}\text{ is a Reeb chord generator.}\end{cases}

    The sign in the above formulas is equal to - if i=ji=j, and ++ otherwise.

  • •:

    uu maps the boundary arc labeled by κj\kappa_{j} to the component LκjL_{\kappa_{j}} of L¯\overline{L}.

Refer to caption
Figure 2. A JJ-holomorphic disk in fi(𝒄;κ)\mathcal{M}^{\text{fi}}(\boldsymbol{c};\kappa). The dot on the right hand side indicates that ζj\zeta_{j} (near which, uu is asymptotic to cjc_{j}) is the distinguished puncture.
Symplectization disks:

Consider DmD_{m} equipped with a decreasing boundary numbering (not necessarily strictly decreasing). Let Dm,k . . =Dm{ζ1in,,ζkin}D_{m,k}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=D_{m}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{\zeta^{\text{in}}_{1},\ldots,\zeta^{\text{in}}_{k}\right\}, where each ζiin\zeta^{\text{in}}_{i} is a point in the interior of DmD_{m}. We equip each ζiin\zeta^{\text{in}}_{i} with a negative cylinder-like end. That is a biholomorphism

φi:(0,)×S1N(ζiin)i{1,,k}.\varphi_{-}^{i}\colon\thinspace(0,\infty)\times S^{1}\longrightarrow N(\zeta^{\text{in}}_{i})\quad\forall i\in\left\{1,\ldots,k\right\}\,.

We let 𝒄=c1σ1cmσm\boldsymbol{c}=c_{1}^{\sigma_{1}}\cdots c_{m}^{\sigma_{m}} be a word of signed Reeb chord generators of CW(L,L)CW^{\ast}(L,L), where σi{+,}\sigma_{i}\in\left\{+,-\right\} for every ii. We also let 𝜸=γ1γk\boldsymbol{\gamma}=\gamma_{1}\cdots\gamma_{k} be a word of Reeb orbits in YY, each of which is equipped with an asymptotic marker, i.e. a point piimγip_{i}\in\operatorname{im}\gamma_{i}. The distinguished boundary puncture ζj\zeta_{j} induces an asymptotic marker for each interior puncture ζiin\zeta^{\text{in}}_{i}, which is a half-ray φ1((,0)×{xi})\varphi^{-1}((-\infty,0)\times\left\{x_{i}\right\}) near ζi\zeta_{i} [EO17, Section 2.1]. By abuse of notation we say that xiS1x_{i}\in S^{1} is the asymptotic marker of ζiin\zeta^{\text{in}}_{i}. Then we define neg(𝒄,𝜸;κ)\mathcal{M}^{\text{neg}}(\boldsymbol{c},\boldsymbol{\gamma};\kappa) to be the moduli space of JJ-olomorphic maps u:(Dm,k,Dm,k)(×Y,×Λ¯)u\colon\thinspace(D_{m,k},\partial D_{m,k})\longrightarrow(\mathbb{R}\times Y,\mathbb{R}\times\overline{\varLambda}) such that

  • •:

    near the boundary puncture ζi\zeta_{i}, uu is asymptotic to the Reeb chord ciσic_{i}^{\sigma_{i}} of Λ\varLambda at ±\pm\infty, depending on the sign σi\sigma_{i}, that is

    lims±u(ε±i(s,t))=(±,ci).\lim_{s\to\pm\infty}u(\varepsilon_{\pm}^{i}(s,t))=(\pm\infty,c_{i})\,.
  • •:

    near the interior puncture ζiin\zeta^{\text{in}}_{i}, uu is asymptotic to the Reeb orbit γi\gamma_{i} in YY at -\infty respecting the asymptotic markers, that is

    {limsu(φi(s,t))=(,γi)limsu(φi(s,xi))=(,pi).\begin{cases}\lim_{s\to-\infty}u(\varphi_{-}^{i}(s,t))=(-\infty,\gamma_{i})\\ \lim_{s\to-\infty}u(\varphi_{-}^{i}(s,x_{i}))=(-\infty,p_{i})\,.\end{cases}
  • •:

    uu maps the boundary arc labeled by κj\kappa_{j} to the component ×Λκj\mathbb{R}\times\varLambda_{\kappa_{j}} of ×Λ¯\mathbb{R}\times\overline{\varLambda}, and

  • •:

    if ζi\zeta_{i} is a constant puncture, we require ζi\zeta_{i} to be a negative puncture (i.e. asymptotic to a Reeb chord of Λ\varLambda at -\infty).

Refer to caption
Figure 3. A JJ-holomorphic disk in neg(𝒄,γ1;κ)\mathcal{M}^{\text{neg}}(\boldsymbol{c},\gamma_{1};\kappa). The dot on the right hand side indicates that the puncture ζj\zeta_{j} is the distinguished puncture. The \ast on the right hand side is the asymptotic marker p1imγ1p_{1}\in\operatorname{im}\gamma_{1}.

Let γ\gamma be a Reeb orbit in YY, equipped with the asymptotic marker pimγp\in\operatorname{im}\gamma. Let SS denote S2S^{2} with one puncture ζS2\zeta\in S^{2}, with a fixed choice of asymptotic marker xx at ζ\zeta. Equip ζ\zeta with a positive cylinder-like end

φ+:(0,)×S1N(ζ).\varphi_{+}\colon\thinspace(0,\infty)\times S^{1}\longrightarrow N(\zeta)\,.

Let λ(γ)\mathcal{M}^{\lambda}(\gamma) be the λ\lambda-perturbed moduli space of JλJ_{\lambda}-holomorphic maps u:SXu\colon\thinspace S\longrightarrow X with notation as in [Ekh19, Theorem 1.1], satisfying

{limsu(φ+(s,t))=(,γ)limsu(φ+(s,x))=(,p).\begin{cases}\lim_{s\to\infty}u(\varphi_{+}(s,t))=(\infty,\gamma)\\ \lim_{s\to\infty}u(\varphi_{+}(s,x))=(\infty,p)\,.\end{cases}

Then we define

sy(𝒄;κ) . . =𝜸(neg(𝒄,𝜸;κ)×γi𝜸λ(γi)),\mathcal{M}^{\text{sy}}(\boldsymbol{c};\kappa)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\bigcup_{\boldsymbol{\gamma}}\left(\mathcal{M}^{\text{neg}}(\boldsymbol{c},\boldsymbol{\gamma};\kappa)\times\prod_{\gamma_{i}\in\boldsymbol{\gamma}}\mathcal{M}^{\lambda}(\gamma_{i})\right)\,,

See [Ekh19] and [EL17, Appendix A.1] for more deatils. Each curve in sy(𝒄;κ)\mathcal{M}^{\text{sy}}(\boldsymbol{c};\kappa) should be interpreted as curves shown in Fig. 3, but with all Reeb orbits capped off by punctured JλJ_{\lambda}-holomorphic spheres.

Refer to caption
Figure 4. A JJ-holomorphic disk in sy(𝒄;κ)\mathcal{M}^{\text{sy}}(\boldsymbol{c};\kappa). The dot on the right hand side indicates that the puncture ζj\zeta_{j} is the distinguished puncture. The \ast on the right hand side is the asymptotic marker p1imγ1p_{1}\in\operatorname{im}\gamma_{1}.
Partial holomorphic buildings:

The domain of a partial holomorphic building is a possibly broken disk with m+1m+1 boundary punctures, see Fig. 5. We denote this (possibly broken) disk by Dm+1D_{m+1} and equip it with a decreasing boundary numbering κ\kappa. In the target, the partial holomorphic building consists of a two-level JJ-holomorphic building, with exactly one symplectization disk (called the primary disk), and multiple filling disks (called secondary disks). We require that the distinguished puncture (which is the only increasing puncture), is a negative puncture of the primary disk. If the primary disk only has one negative puncture, the primary disk is the only component, and the disk is not broken. If the primary disk has more than 1 negative puncture, each additional negative puncture has a secondary disk attached to it, at the distinguished puncture of the secondary disks disk. If 𝒄=c0c1cm\boldsymbol{c}=c_{0}c_{1}\cdots c_{m} is a word of generators of CW(L,L)CW^{\ast}(L,L), where c0c_{0} is the generator to which the distinguished puncture is asymptotic to, we denote the moduli space of partial holomorphic buildings by pb(𝒄;κ)\mathcal{M}^{\text{pb}}(\boldsymbol{c};\kappa).

Refer to caption
Figure 5. A partial holomorphic building in pb(𝒄;κ)\mathcal{M}^{\text{pb}}(\boldsymbol{c};\kappa). The dots on the right hand side indicate the distinguished punctures of the corresponding disks. The signs on the left hand side indiciate the sign of the punctures of the primary disk.
Remark 2.2.

Take note that we might have additional negative punctures of the symplectization disk, at which there are constant filling disks with only 1 positive puncture attached. We have not depicted these above, but they should nonetheless be taken into account.

By [EL17, Theorem 63,65] and [Ekh19, Theorem 1.1], fi(𝒄;κ)\mathcal{M}^{\mathrm{fi}}(\boldsymbol{c};\kappa) and sy(𝒄;κ)\mathcal{M}^{\text{sy}}(\boldsymbol{c};\kappa) are transversely cut out smooth manifolds that are independent of the boundary numbering κ\kappa up to diffeomorphism. This follows from the observation that disks in fi(𝒄;κ)\mathcal{M}^{\text{fi}}(\boldsymbol{c};\kappa) or sy(𝒄;κ)\mathcal{M}^{\text{sy}}(\boldsymbol{c};\kappa) can not be multiply covered for topological reasons. Transversality is then proved using standard techniques as in [EES07]. Furthermore the moduli spaces admit compactifications that consists of JJ-holomorphic buildings of several levels.

Remark 2.3.

For Reeb chord generators the grading |a|\mathinner{\!\left\lvert a\right\rvert} is more explicitly described as follows. Suppose that a:[0,]Ya\colon\thinspace[0,\ell]\longrightarrow Y, then we first define the Conley–Zehnder index CZ(a)\operatorname{CZ}(a) by following [EES05, Section 2.2]. Namely, let aa^{-} and a+a^{+} be the start and endpoints of the Reeb chord aa, respectively. Then pick a capping path γc:[0,1]ΛY\gamma_{c}\colon\thinspace[0,1]\longrightarrow\varLambda\subset Y so that γc(0)=a+\gamma_{c}(0)=a^{+}, γc(1)=a\gamma_{c}(1)=a^{-}. Let α=λ|M\alpha=\mathinner{\lambda\rvert}_{\partial M} and ξ=kerα\xi=\ker\alpha. Then Ta+Λξa+T_{a^{+}}\varLambda\subset\xi_{a^{+}} is a Lagrangian submanifold. By parallel transport along γc\gamma_{c} and via the linearized Reeb flow we get a path of Lagrangian submanifolds in the contact planes ξTY\xi\subset TY. If we close this path up by positive close-up in the contact planes we obtain a loop of Lagrangian submanifolds in ξ\xi denoted by Γa\varGamma_{a}. We then define the Conley–Zehnder index of aa to be the Maslov index of Γa\varGamma_{a} (in the sense of [RS93]),

CZ(a) . . =μ(Γa).\operatorname{CZ}(a)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mu(\varGamma_{a})\,.

Then we define

|a|=CZ(a)+(n1).\mathinner{\!\left\lvert a\right\rvert}=-\operatorname{CZ}(a)+(n-1)\,.

For Lagrangian intersection generators xL0L1x\in L_{0}\cap L_{1} we use the choice of graded lifts of L0L_{0} and L1L_{1} to obtain a path starting at TxL1T_{x}L_{1} and ending at TxL0T_{x}L_{0}. We close this path up in TxMT_{x}M by a positive rotation. This gives a loop of Lagrangian submanifolds denoted by Γx\varGamma_{x}, which starts and ends at TxL0TxMT_{x}L_{0}\subset T_{x}M. Then define the grading of xx as the Maslov index of this loop [EL17, p. 89] [CEL10, Appendix A]

|x| . . =μ(Γx).\mathinner{\!\left\lvert x\right\rvert}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mu(\varGamma_{x})\,.

The dimension of the moduli space fi(𝒂;κ)\mathcal{M}^{\text{fi}}(\boldsymbol{a};\kappa) is dependent on whether the distinguished puncture is a Reeb chord or a Lagrangian intersection puncture. To emphasize the differences, we introduce some more notation.

  • If the distinguished puncture is a Reeb chord generator we denote it by fi,Reeb(𝒂;κ)\mathcal{M}^{\text{fi,Reeb}}(\boldsymbol{a};\kappa), and

  • if the distinguished puncture is an intersection generator we denote it by fi,Lag(𝒂;κ)\mathcal{M}^{\text{fi,Lag}}(\boldsymbol{a};\kappa).

Theorem 2.4.

Let 𝐚=ca2am\boldsymbol{a}=ca_{2}\cdots a_{m} be a word of generators of CW(L,L)CW^{\ast}(L,L). Assume that cc is the distinguished puncture and that it is a Reeb chord generator. Then the dimension of the moduli space fi,Reeb(𝐚;κ)\mathcal{M}^{\mathrm{fi,Reeb}}(\boldsymbol{a};\kappa) is

dim(fi,Reeb(𝒂;κ))=(n3)+m|c|j=2m|aj|.\dim\left(\mathcal{M}^{\mathrm{fi,Reeb}}(\boldsymbol{a};\kappa)\right)=(n-3)+m-\mathinner{\!\left\lvert c\right\rvert}-\sum_{j=2}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}\,.

Let 𝐚=xa2am\boldsymbol{a}=xa_{2}\cdots a_{m} be a word of generators of CW(L,L)CW^{\ast}(L,L). Assume that xx is the distinguished puncture and that it is a Lagrangian intersection generator. Then the dimension of the moduli space fi,Lag(𝐚;κ)\mathcal{M}^{\mathrm{fi,Lag}}(\boldsymbol{a};\kappa) is

dim(fi,Lag(𝒂;κ))=3+m|x|j=2m|aj|.\dim\left(\mathcal{M}^{\mathrm{fi,Lag}}(\boldsymbol{a};\kappa)\right)=-3+m-\mathinner{\!\left\lvert x\right\rvert}-\sum_{j=2}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}\,.

For any word of Reeb chord generators 𝐜=c1cm\boldsymbol{c}=c_{1}\cdots c_{m}, the dimension of the moduli space sy(𝐜;κ)\mathcal{M}^{\mathrm{sy}}(\boldsymbol{c};\kappa) is

dim(sy(𝒄;κ))=(n3)+m+σj=(|cj|(n1))σj=+|cj|.\dim\left(\mathcal{M}^{\mathrm{sy}}(\boldsymbol{c};\kappa)\right)=(n-3)+m+\sum_{\sigma_{j}=-}(\mathinner{\!\left\lvert c_{j}\right\rvert}-(n-1))-\sum_{\sigma_{j}=+}\mathinner{\!\left\lvert c_{j}\right\rvert}\,.

For any word of Reeb chord generators 𝐜=c0cm\boldsymbol{c}=c_{0}\cdots c_{m}, the dimension of the moduli space pb(𝐜;κ)\mathcal{M}^{\mathrm{pb}}(\boldsymbol{c};\kappa) is

dim(pb(𝒄;κ))=1+m+|c0|j=1m|cj|.\dim\left(\mathcal{M}^{\mathrm{pb}}(\boldsymbol{c};\kappa)\right)=-1+m+\mathinner{\!\left\lvert c_{0}\right\rvert}-\sum_{j=1}^{m}\mathinner{\!\left\lvert c_{j}\right\rvert}\,.
Proof.

The theorem follows from applying [CEL10, Theorem A.1], and the fact that the index of a several-level JJ-holomorphic building is the sum of the indices of the disks at each level. Let 𝒂=a1am\boldsymbol{a}=a_{1}\cdots a_{m} be a word of generators of CW(L,L)CW^{\ast}(L,L). Let either usy(𝒂;κ)u\in\mathcal{M}^{\text{sy}}(\boldsymbol{a};\kappa) or ufi(𝒂;κ)u\in\mathcal{M}^{\text{fi}}(\boldsymbol{a};\kappa). Let D^m\widehat{D}_{m} be the unit disk in \mathbb{C} together with ζ1,,ζm\zeta_{1},\ldots,\zeta_{m} regarded as marked points (and not punctures). The boundary of D^m\widehat{D}_{m} is equal to the union of closed boundary arcs CC such that the interiors of all the boundary arcs CC are pairwise disjoint, and only missing the marked points {ζ1,,ζm}\left\{\zeta_{1},\ldots,\zeta_{m}\right\}.

  1. (T1)

    For all Reeb chord generators ai𝒂a_{i}\in\boldsymbol{a}, fix a complex trivialization ZaiZ_{a_{i}} of the contact structure ξ\xi along aia_{i}, such that the linearized Reeb flow along the chord aia_{i} expressed in ZaiZ_{a_{i}} is constantly equal to the identity.

  2. (T2)

    For each boundary arc CC in D^m\widehat{D}_{m}, fix a complex trivialization ZCZ_{C} of uTMu^{\ast}TM (if ufi(𝒂;κ)u\in\mathcal{M}^{\text{fi}}(\boldsymbol{a};\kappa)) or uT(×Y)u^{\ast}T(\mathbb{R}\times Y) (if usy(𝒂;κ)u\in\mathcal{M}^{\text{sy}}(\boldsymbol{a};\kappa)) with the following properties:

    1. (a)

      If an endpoint of CC is a puncture ζi\zeta_{i} asymptotic to a Reeb chord aia_{i}, then ZC=ZaiZ_{C}=Z_{a_{i}}.

    2. (b)

      If an endpoint of CC is a puncture ζi\zeta_{i} asymptotic to an intersection generator xiLκiLκi+1x_{i}\in L_{\kappa_{i}}\cap L_{\kappa_{i+1}}, then ZC=ZCZ_{C}=Z_{C^{\prime}} where ζi\zeta_{i} is the common endpoint of the boundary arcs CC and CC^{\prime}.

Items (T1) and (T2) above give a complex trivialization ZjuZ_{\partial_{j}u} of uTMu^{\ast}TM (or uT(×Y)u^{\ast}T(\mathbb{R}\times Y)) over the jthj^{\text{th}} boundary arc CjC_{j} of D^m\widehat{D}_{m}. For each boundary arc CjC_{j}, let CjC^{\prime}_{j} be the complement of its endpoints in CjC_{j}. The tangent planes of LL along all f(Cj)f(C^{\prime}_{j}) expressed in the trivialization ZjuZ_{\partial_{j}u} gives a collection of paths of Lagrangian subspaces in n\mathbb{C}^{n}. We close up this path to a loop as follows. For each Reeb chord ai𝒂a_{i}\in\boldsymbol{a}, denote its start and endpoints by ai±a_{i}^{\pm} respectively.

  1. (C1)

    For each positive puncture ζi\zeta_{i} near which uu is asymptotic to the Reeb chord aia_{i}, the tangent planes of L=×ΛL=\mathbb{R}\times\varLambda are connected by the product of the linearized Reeb flow along aia_{i} in ξ\xi, and the identity in the \mathbb{R}-factor, followed by negative close-up in the contact plane in ξai+×\xi_{a_{i}^{+}}\times\mathbb{C} (cf. 2.3). Denote this path of Lagrangian subspaces by gai+g_{a_{i}}^{+}.

  2. (C2)

    For each negative puncture ζi\zeta_{i} near which uu is asymptotic to the Reeb chord aia_{i}, the tangent planes of L=×ΛL=\mathbb{R}\times\varLambda are connected by the product of the backwards linearized Reeb flow along aia_{i} in ξ\xi, and the identity in the \mathbb{R}-factor, followed by negative close-up in the contact plane in ξai×\xi_{a_{i}^{-}}\times\mathbb{C} (cf. 2.3). Denote this path of Lagrangian subspaces by gaig_{a_{i}}^{-}.

  3. (C3)

    For each puncture ζi\zeta_{i} near which uu is asymptotic to the intersection generator xiLκiLκi+1x_{i}\in L_{\kappa_{i}}\cap L_{\kappa_{i+1}}, connect the planes TxiLκiT_{x_{i}}L_{\kappa_{i}} and TxiLκi+1T_{x_{i}}L_{\kappa_{i+1}} by a negative rotation taking TxiLκiT_{x_{i}}L_{\kappa_{i}} to TxiLκi+1T_{x_{i}}L_{\kappa_{i+1}} in n\mathbb{C}^{n} (cf. 2.3 and [CEL10, Remark A.1]). Denote this path of Lagrangian subspaces by gxig_{x_{i}}^{\cap}.

Define μ(u,Zu)\mu(\partial u,Z_{\partial u}) to be the Maslov index of the loop of Lagrangian subspaces in n\mathbb{C}^{n} which is constructed by closing up paths of Lagrangian subspaces as described in (C1), (C2) and (C3). For the moduli spaces of filling disks and symplectization disks, we then have by [CEL10, Theorem A.1.] that

dim(fi(𝒂;κ))\displaystyle\dim\left(\mathcal{M}^{\text{fi}}(\boldsymbol{a};\kappa)\right) =(n3)+m+μ(u,Zu)\displaystyle=(n-3)+m+\mu(\partial u,Z_{\partial u})
dim(sy(𝒂;κ))\displaystyle\dim\left(\mathcal{M}^{\text{sy}}(\boldsymbol{a};\kappa)\right) =(n3)+m+μ(u,Zu).\displaystyle=(n-3)+m+\mu(\partial u,Z_{\partial u})\,.

Since LL is assumed to have vanishing Maslov class, the contribution to μ(u,Zu)\mu(\partial u,Z_{\partial u}) is equal to the sum of each contribution at every boundary puncture of DmD_{m}. Next we describe each of these contributions in terms of the grading of each generator. First let usy(𝒂;κ)u\in\mathcal{M}^{\text{sy}}(\boldsymbol{a};\kappa).

  1. (sy1)

    If ζi\zeta_{i} is a positive puncture near which uu is asymptotic to the Reeb chord aia_{i} then

    μ(gai+(Γai)1)=(n1)μ(gai+)=μ(Γai)(n1)=|ai|.\mu(g_{a_{i}}^{+}\circ(\varGamma_{a_{i}})^{-1})=-(n-1)\Leftrightarrow\mu(g_{a_{i}}^{+})=\mu(\varGamma_{a_{i}})-(n-1)=-\mathinner{\!\left\lvert a_{i}\right\rvert}\,.
  2. (sy2)

    If ζi\zeta_{i} is a negative puncture near which uu is asymptotic to the Reeb chord aia_{i} then

    μ(gaiΓai)=0μ(gai)=μ(Γai)=|ai|(n1).\mu(g_{a_{i}}^{-}\circ\varGamma_{a_{i}})=0\Leftrightarrow\mu(g_{a_{i}}^{-})=-\mu(\varGamma_{a_{i}})=\mathinner{\!\left\lvert a_{i}\right\rvert}-(n-1)\,.

Then let ufi(𝒂;κ)u\in\mathcal{M}^{\text{fi}}(\boldsymbol{a};\kappa).

  1. (fi1)

    Let ζi\zeta_{i} be a puncture near which uu is asymptotic to the Reeb chord aia_{i}.

    1. (a)

      If ζi\zeta_{i} is the distinguished puncture then

      μ(gai+Γai)=0μ(gai+)=μ(Γai)=|ai|(n1).\mu(g_{a_{i}}^{+}\circ\varGamma_{a_{i}})=0\Leftrightarrow\mu(g_{a_{i}}^{+})=-\mu(\varGamma_{a_{i}})=\mathinner{\!\left\lvert a_{i}\right\rvert}-(n-1)\,.
    2. (b)

      If ζi\zeta_{i} is not the distinguished puncture then

      μ(gai+(Γai)1)=(n1)μ(gai+)=μ(Γai)(n1)=|ai|.\mu(g_{a_{i}}^{+}\circ(\varGamma_{a_{i}})^{-1})=-(n-1)\Leftrightarrow\mu(g_{a_{i}}^{+})=\mu(\varGamma_{a_{i}})-(n-1)=-\mathinner{\!\left\lvert a_{i}\right\rvert}\,.
  2. (fi2)

    Let ζi\zeta_{i} be a puncture near which uu is asymptotic to the intersection generator xix_{i}

    1. (a)

      If ζi\zeta_{i} is the distinguished puncture then

      μ(gxi(Γxi)1)=nμ(gxi)=μ(Γxi)n=|xi|n.\mu(g_{x_{i}}^{\cap}\circ(\varGamma_{x_{i}})^{-1})=-n\Leftrightarrow\mu(g_{x_{i}}^{\cap})=\mu(\varGamma_{x_{i}})-n=\mathinner{\!\left\lvert x_{i}\right\rvert}-n\,.
    2. (b)

      If ζi\zeta_{i} is not the distinguished puncture then

      μ(gxiΓxi)=0μ(gxi)=μ(Γxi)=|xi|.\mu(g_{x_{i}}^{\cap}\circ\varGamma_{x_{i}})=0\Leftrightarrow\mu(g_{x_{i}}^{\cap})=-\mu(\varGamma_{x_{i}})=-\mathinner{\!\left\lvert x_{i}\right\rvert}\,.

From (sy1) and (sy2) we obtain

dim(sy(𝒄;κ))\displaystyle\dim\left(\mathcal{M}^{\text{sy}}(\boldsymbol{c};\kappa)\right) =(n3)+m+σj=(|cj|(n1))σj=+|cj|.\displaystyle=(n-3)+m+\sum_{\sigma_{j}=-}\left(\mathinner{\!\left\lvert c_{j}\right\rvert}-(n-1)\right)-\sum_{\sigma_{j}=+}\mathinner{\!\left\lvert c_{j}\right\rvert}\,.

From (fi1)(a), (fi1)(b) and (fi2)(b) we obtain

dim(fi,Reeb(𝒂;κ))=(n3)+m|c|j=2m|aj|.\dim\left(\mathcal{M}^{\text{fi,Reeb}}(\boldsymbol{a};\kappa)\right)=(n-3)+m-\mathinner{\!\left\lvert c\right\rvert}-\sum_{j=2}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}\,.

From (fi2)(a), (fi1)(b) and (fi2)(b) we obtain

dim(fi,Lag(𝒂;κ))=(n3)+m+(|x|n)j=2m|aj|=3+m+|x|j=2m|aj|.\dim\left(\mathcal{M}^{\text{fi,Lag}}(\boldsymbol{a};\kappa)\right)=(n-3)+m+(\mathinner{\!\left\lvert x\right\rvert}-n)-\sum_{j=2}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}=-3+m+\mathinner{\!\left\lvert x\right\rvert}-\sum_{j=2}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}\,.

For a partial holomorphic building, let a1,,apa_{1},\ldots,a_{p} be the positive punctures of the primary disk, let d0d_{0} be the distinguished negative puncture of the primary disk and let d1,,dqd_{1},\ldots,d_{q} be the remaining negative punctures. Let b1,,brb_{1},\ldots,b_{r} be all the non-distinguished punctures of all the secondary disks, see Fig. 6. Each secondary disk lies in fi,Reeb(𝒂;κ)\mathcal{M}^{\text{fi,Reeb}}(\boldsymbol{a};\kappa). We may then compute the dimension by taking sums, that is

dim(pb(𝒄;κ))\displaystyle\dim\left(\mathcal{M}^{\text{pb}}(\boldsymbol{c};\kappa)\right) =[(n3)j=1p(|aj|1)+j=0q(|dj|(n2))]\displaystyle=\left[(n-3)-\sum_{j=1}^{p}\left(\mathinner{\!\left\lvert a_{j}\right\rvert}-1\right)+\sum_{j=0}^{q}\left(\mathinner{\!\left\lvert d_{j}\right\rvert}-(n-2)\right)\right]
+[j=1q((n3)(|dj|1))+j=1r(|bj|1)].\displaystyle\qquad+\left[\sum_{j=1}^{q}\left((n-3)-\left(\mathinner{\!\left\lvert d_{j}\right\rvert}-1\right)\right)+\sum_{j=1}^{r}-\left(\mathinner{\!\left\lvert b_{j}\right\rvert}-1\right)\right]\,.
Refer to caption
Figure 6.

After canceling we get

dim(pb(𝒄;κ))\displaystyle\dim\left(\mathcal{M}^{\text{pb}}(\boldsymbol{c};\kappa)\right) =(n3)j=1p(|aj|1)+(|d0|(n2))j=1r(|bj|1)\displaystyle=(n-3)-\sum_{j=1}^{p}\left(\mathinner{\!\left\lvert a_{j}\right\rvert}-1\right)+\left(\mathinner{\!\left\lvert d_{0}\right\rvert}-(n-2)\right)-\sum_{j=1}^{r}\left(\mathinner{\!\left\lvert b_{j}\right\rvert}-1\right)
=1+(p+r)+|d0|j=1p|aj|j=1r|bj|.\displaystyle=-1+(p+r)+\mathinner{\!\left\lvert d_{0}\right\rvert}-\sum_{j=1}^{p}\mathinner{\!\left\lvert a_{j}\right\rvert}-\sum_{j=1}^{r}\mathinner{\!\left\lvert b_{j}\right\rvert}\,.

Now let c0 . . =d0c_{0}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=d_{0} and let 𝒄\boldsymbol{c} be the word of m . . =p+rm\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=p+r letters corresponding to all the generators a1,,ap,b1,,bra_{1},\ldots,a_{p},b_{1},\ldots,b_{r} in the appropriate order. Therefore

dim(pb(𝒄;κ))=m1+|c0|j=1m|cj|.\dim\left(\mathcal{M}^{\text{pb}}(\boldsymbol{c};\kappa)\right)=m-1+\mathinner{\!\left\lvert c_{0}\right\rvert}-\sum_{j=1}^{m}\mathinner{\!\left\lvert c_{j}\right\rvert}\,.

We now define operations, one for each i1i\geq 1,

μi:CW(Lκi1,Lκi)CW(Lκ1,Lκ2)CW(Lκ1,Lκi),\mu^{i}\colon\thinspace CW^{\ast}(L_{\kappa_{i-1}},L_{\kappa_{i}})\otimes\cdots\otimes CW^{\ast}(L_{\kappa_{1}},L_{\kappa_{2}})\longrightarrow CW^{\ast}(L_{\kappa_{1}},L_{\kappa_{i}})\,,

that counts various JJ-holomorphic disks discussed above. We split it as a sum μi=μLagi+μReebi\mu^{i}=\mu^{i}_{\text{Lag}}+\mu^{i}_{\text{Reeb}}, where μLagi\mu^{i}_{\text{Lag}} takes values in Lagrangian intersection generators and μReebi\mu^{i}_{\text{Reeb}} takes values in Reeb chord generators.

First we consider μLagi\mu^{i}_{\text{Lag}}. Let 𝒄=c1ci\boldsymbol{c}^{\prime}=c_{1}\cdots c_{i} be a word of generators of CW(L,L)CW^{\ast}(L,L). Then

μLagi(cic1) . . =|c0|=|𝒄|+(2i)|fi,Lag(c0𝒄;κ)|c0.\mu^{i}_{\text{Lag}}(c_{i}\otimes\cdots\otimes c_{1})\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\sum_{\mathinner{\!\left\lvert c_{0}\right\rvert}=\mathinner{\!\left\lvert\boldsymbol{c}^{\prime}\right\rvert}+(2-i)}\mathinner{\!\left\lvert\mathcal{M}^{\text{fi,Lag}}(c_{0}\boldsymbol{c}^{\prime};\kappa)\right\rvert}c_{0}\,.

The sum is taken over all Lagrangian intersection generators c0c_{0} so that dim(fi,Lag(c0𝒄;κ))=0\dim\left(\mathcal{M}^{\text{fi,Lag}}(c_{0}\boldsymbol{c}^{\prime};\kappa)\right)=0.

To define μReebi\mu^{i}_{\text{Reeb}}, consider a word of generators 𝒄=c1ci\boldsymbol{c}^{\prime}=c_{1}\cdots c_{i}. Then

μReebi(cic1) . . =|c0|=|𝒄|+(2i)|pb(c0𝒄;κ)|c0.\mu^{i}_{\text{Reeb}}(c_{i}\otimes\cdots\otimes c_{1})\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\sum_{\mathinner{\!\left\lvert c_{0}\right\rvert}=\mathinner{\!\left\lvert\boldsymbol{c}^{\prime}\right\rvert}+(2-i)}\mathinner{\!\left\lvert\mathcal{M}^{\text{pb}}(c_{0}\boldsymbol{c}^{\prime};\kappa)\right\rvert}c_{0}\,.

The sum is taken over all Reeb chords c0c_{0} so that dim(pb(c0𝒄;κ))=0\dim\left(\mathcal{M}^{\text{pb}}(c_{0}\boldsymbol{c}^{\prime};\kappa)\right)=0. The total operation μi\mu^{i} is then defined as

(2.1) μi(cic1)=(1)(μLagi(cic1)+μReebi(cic1))\mu^{i}(c_{i}\otimes\cdots\otimes c_{1})=(-1)^{\diamond}\left(\mu^{i}_{\text{Lag}}(c_{i}\otimes\cdots\otimes c_{1})+\mu^{i}_{\text{Reeb}}(c_{i}\otimes\cdots\otimes c_{1})\right)

where

=j=1ij|cj|.\diamond=\sum_{j=1}^{i}j\mathinner{\!\left\lvert c_{j}\right\rvert}\,.
Lemma 2.5.

With the sign conventions as in [Sei08], (CW(L,L),{μi}i=1)\left(CW^{\ast}(L,L),\left\{\mu^{i}\right\}_{i=1}^{\infty}\right) forms an AA_{\infty}-algebra, that is

d1+d2=d+10k<d1(1)kμd1(cd,,ck+d2+1,μd2(ck+d2,,ck+1),ck,,c1)=0,\sum_{\begin{subarray}{c}d_{1}+d_{2}=d+1\\ 0\leq k<d_{1}\end{subarray}}(-1)^{\maltese_{k}}\mu^{d_{1}}(c_{d},\ldots,c_{k+d_{2}+1},\mu^{d_{2}}(c_{k+d_{2}},\ldots,c_{k+1}),c_{k},\ldots,c_{1})=0\,,

where

k=k+j=1k|cj|.\maltese_{k}=k+\sum_{j=1}^{k}\mathinner{\!\left\lvert c_{j}\right\rvert}\,.
Proof.

See [EL17, Lemma 67]. ∎

3. Partially wrapped Floer cohomology and chains of based loops

Let SS be any closed orientable spin manifold and KSK\subset S any submanifold. The purpose of this section is to describe the surgery approach to compute the partially wrapped Floer cohomology of a cotangent fiber in the Weinstein domain (DTS,λ=pdq)(DT^{\ast}S,\lambda=pdq) stopped by the unit conormal ΛK\varLambda_{K}. We then define a chain map relating the partially wrapped Floer cohomology of a fiber to chains of based loops on a Lagrangian submanifold MKM_{K} that is diffeomorphic to the complement SKS\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}K.

In Section 3 we describe the surgery approach in more detail, and also construct the Lagrangian MKM_{K}. In Section 3.2 we describe the model we use for the chains of based loops on MKM_{K}, and equip it with the Pontryagin product. Then in Section 3.3 we describe the moduli space of half strips which we need in order to to define an AA_{\infty}-homomorphism between the partially wrapped Floer cocomplex and the chains of based loops on MKM_{K}. The construction of the AA_{\infty}-homomorphism is carried out in Section 3.4.

3.1. Partially wrapped Floer cohomology using a surgery approach

Following [EL17, Appendix B] and [ENS16, Section 6] we will now describe the surgery approach. We consider the disk cotangent bundle DTSDT^{\ast}S the conormal bundle of KK

LK={(q,p)DTS|qK,p,TqK=0}.L_{K}=\left\{(q,p)\in DT^{\ast}S\;|\;q\in K,\,\left\langle p,T_{q}K\right\rangle=0\right\}\,.

Let ΛK . . =LKSTS\varLambda_{K}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=L_{K}\cap ST^{\ast}S be the unit conormal of KK. We take a tubular neighborhood UU of ΛK\varLambda_{K} in STSST^{\ast}S and we attach a handle modeled on DεT([0,)×Λ)D_{\varepsilon}T^{\ast}([0,\infty)\times\varLambda) to UU. After handle attachment and after smoothing out corners, the Liouville vector field is equal to ppp\partial_{p} in DεT([T,)×Λ)D_{\varepsilon}T^{\ast}([T,\infty)\times\varLambda) (for T0T\geq 0 large enough) for coordinates (q,p)(q,p) in the handle. We call the resulting manifold WKW_{K}, see Fig. 7.

We then consider a cotangent fiber FDTξSF\cong DT^{\ast}_{\xi}S at ξMK\xi\in M_{K} in WKW_{K}. Denote the wrapped Floer cochains of FF in WKW_{K} as described in Section 2 by CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F).

Remark 3.1.

In the language of Sylvan [Syl19], we obtain a stop σΛK\sigma_{\varLambda_{K}} from ΛK\varLambda_{K} as follows. Pick a tubular neighborhood UΛKU\supset\varLambda_{K} in STSST^{\ast}S, and a strict contactomorphism φ:(U,λ|U)(V,dzydx)\varphi\colon\thinspace(U,\mathinner{\lambda\rvert}_{U})\longrightarrow(V,dz-ydx) where VV is a tubular neighborhood of ΛKJ1(ΛK)=TΛK×\varLambda_{K}\subset J^{1}(\varLambda_{K})=T^{\ast}\varLambda_{K}\times\mathbb{R}, viewed as the zero section. Then the Liouville hypersurface σΛK . . =φ1(TΛKV)U\sigma_{\varLambda_{K}}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\varphi^{-1}(T^{\ast}\varLambda_{K}\cap V)\subset U is a stop.

Another point of view, is to remove the tubular neighborhood UU from STSST^{\ast}S, and take the Liouville completion of (DTS)U\left(DT^{\ast}S\right)\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}U to obtain a Liouville sector as defined [GPS20]. The wrapped Fukaya category of this Liouville sector coincides with the wrapped Fukaya category associated to the pair (M,σΛK)(M,\sigma_{\varLambda_{K}}), and also with the Fukaya category associated to WKW_{K} [EL17, GPS20, GPS18b].

Refer to caption
Figure 7. The Liouville sector WKW_{K}.

To construct the complement Lagrangian MKM_{K}, we perform Lagrangian surgery of LKL_{K} and SS which intersect cleanly along KK. Above each point of KK, the intersection LKSL_{K}\cap S looks like the transverse intersection of two Lagrangian disks of dimension kk. We perform Lagrangian surgery along KK as in [MW18, Section 2.2.2] [AENV14]. We denote the result of the surgery by MKSKM_{K}\cong S\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}K (cf. [AENV14]).

Remark 3.2.

Note that the Maslov class of MKM_{K} vanishes, because it is the result of surgery of SWS\subset W and LKWL_{K}\subset W, both of which have vanishing Maslov class. In particular, consider the following model. We pick a n\mathbb{C}^{n}-neighborhood around pKp\in K such that LK=inL_{K}=i\mathbb{R}^{n} and S=nS=\mathbb{R}^{n}. Following the discussion in [ES16, Section 2.2], we have a phase function ϕ:H\phi\colon\thinspace H\longrightarrow\mathbb{R} which is unique up to an additive constant on the handle HH, so that ϕ|SH=0\mathinner{\phi\rvert}_{S\cap H}=0 and ϕ|LKH=n1\mathinner{\phi\rvert}_{L_{K}\cap H}=n-1.

Any loop that is based at any point outside of the handle pass through the entire handle an even number of times, which means that the total Maslov index of the loop is zero.

3.2. Based loops on MKM_{K}

Consider the Moore loop space of MKM_{K}, based at ξ\xi

ΩξMK={γ:[0,R]MK|γ(0)=γ(R)=ξ}.\varOmega_{\xi}M_{K}=\left\{\gamma\colon\thinspace[0,R]\longrightarrow M_{K}\;|\;\gamma(0)=\gamma(R)=\xi\right\}\,.

We use a cubical model for chains of based loops as in [Abo12b, EL17].

A singular kk-cube is a smooth map σ:[0,1]kΩξMK\sigma\colon\thinspace[0,1]^{k}\longrightarrow\varOmega_{\xi}M_{K} and it is called degenerate if σ(x1,,xk)\sigma(x_{1},\ldots,x_{k}) is constant in at least one of the coordinates. We define the space of cubical kk-chains by

Ck(ΩξMK)=[singular k-cubes][degenerate singular k-cubes].C_{k}(\varOmega_{\xi}M_{K})=\frac{\mathbb{Z}[\text{singular $k$-cubes}]}{\mathbb{Z}[\text{degenerate singular $k$-cubes}]}\,.

We equip C(ΩξMK)C_{\ast}(\varOmega_{\xi}M_{K}) with the differential

(3.1) σ . . =i=1kε=01(1)i+εσ(δi,ε(x1,,xk)),\partial\sigma\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\sum_{i=1}^{k}\sum_{\varepsilon=0}^{1}(-1)^{i+\varepsilon}\sigma(\delta_{i,\varepsilon}(x_{1},\ldots,x_{k}))\,,

where

δi,ε(x1,,xk)=(x1,,xi1,ε,xi+1,,xk),ε{0,1}\delta_{i,\varepsilon}(x_{1},\ldots,x_{k})=(x_{1},\ldots,x_{i-1},\varepsilon,x_{i+1},\ldots,x_{k}),\;\varepsilon\in\left\{0,1\right\}

is the map that replaces the ii-th coordinate with ε\varepsilon.

The Pontryagin product PP is defined as the following composition:

(3.2) Ck(ΩξMK)C(ΩξMK){C_{k}(\varOmega_{\xi}M_{K})\otimes C_{\ell}(\varOmega_{\xi}M_{K})}Ck+((ΩξMK)2){C_{k+\ell}\left(\left(\varOmega_{\xi}M_{K}\right)^{2}\right)}Ck+(ΩξMK){C_{k+\ell}(\varOmega_{\xi}M_{K})}σ2σ1{\sigma_{2}\otimes\sigma_{1}}(1)|σ1|σ2×σ1{(-1)^{\mathinner{\!\left\lvert\sigma_{1}\right\rvert}}\sigma_{2}\times\sigma_{1}}(1)|σ1|σ1σ2{(-1)^{\mathinner{\!\left\lvert\sigma_{1}\right\rvert}}\sigma_{1}\circ\sigma_{2}}

The cross product of a singular ii-cube σ1\sigma_{1} and a jj-cube σ2\sigma_{2} is the (i+j)(i+j)-cube

σ1×σ2:[0,1]i+j\displaystyle\sigma_{1}\times\sigma_{2}\colon\thinspace[0,1]^{i+j} ΩξMK×ΩξMK\displaystyle\longrightarrow\varOmega_{\xi}M_{K}\times\varOmega_{\xi}M_{K}
(x1,,xi+j)\displaystyle(x_{1},\ldots,x_{i+j}) (σ1(x1,,xi),σ2(xi+1,,xi+j)).\displaystyle\longmapsto(\sigma_{1}(x_{1},\ldots,x_{i}),\sigma_{2}(x_{i+1},\ldots,x_{i+j}))\,.

The map \circ is pointwise concatenation of loops where we first follow σ1(x1,,xi)\sigma_{1}(x_{1},\ldots,x_{i}), and then σ2(xi+1,,xi+j)\sigma_{2}(x_{i+1},\ldots,x_{i+j}). That is (σ1σ2)(x)=σ1(x1,,xi)σ2(xi+1,,xi+j)(\sigma_{1}\circ\sigma_{2})(x)=\sigma_{1}(x_{1},\ldots,x_{i})\circ\sigma_{2}(x_{i+1},\ldots,x_{i+j}), where

(σ1(x1,,xi)σ2(xi+1,,xi+j))(t) . . ={σ1(x1,,xi)(t),t[0,R1]σ2(xi+1,,xi+j)(tR1),t[R1,R1+R2].(\sigma_{1}(x_{1},\ldots,x_{i})\circ\sigma_{2}(x_{i+1},\ldots,x_{i+j}))(t)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\begin{cases}\sigma_{1}(x_{1},\ldots,x_{i})(t),&t\in[0,R_{1}]\\ \sigma_{2}(x_{i+1},\ldots,x_{i+j})(t-R_{1}),&t\in[R_{1},R_{1}+R_{2}]\end{cases}\,.

From the definitions of PP and \partial we see that for any two singular cubes σ1Ck(ΩξMK)\sigma_{1}\in C_{k}(\varOmega_{\xi}M_{K}) and σ2C(ΩξMK)\sigma_{2}\in C_{\ell}(\varOmega_{\xi}M_{K}) we have

(σ1σ2)=(1)k(σ1σ2)+σ1σ2.\partial(\sigma_{1}\circ\sigma_{2})=(-1)^{k}(\sigma_{1}\circ\partial\sigma_{2})+\partial\sigma_{1}\circ\sigma_{2}\,.

This leads via (3.2) to

P(σ2σ1)+P(σ2σ1)+(1)k+1P(σ2σ1)=0.\partial P(\sigma_{2}\otimes\sigma_{1})+P(\sigma_{2}\otimes\partial\sigma_{1})+(-1)^{k+1}P(\partial\sigma_{2}\otimes\sigma_{1})=0\,.

Hence (C(ΩξMK),,P)(C_{\ast}(\varOmega_{\xi}M_{K}),\partial,P) is an AA_{\infty}-algebra with all higher operations being zero with sign conventions as in [AS10, Sei08].

3.3. Moduli space of half strips

Consider the cotangent fiber FTξSWKF\cong T^{\ast}_{\xi}S\subset W_{K} at ξMK\xi\in M_{K} defined in Section 3.1 and consider a system of parallel copies of FF as in Section 2. In this section we construct a moduli spaces of JJ-holomorphic half strips similar to [Abo12b]. This moduli space is used to define a chain map between CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K}). By non-compactness of WKW_{K} in the horizontal direction, we use monotonicity for JJ-holomorphic half strips to establish compactness of moduli spaces, see Appendix A for details.

Let D3D_{3}\subset\mathbb{C} be the positively oriented unit disk with three boundary punctures ζ+,ζ,ζ1\zeta_{+},\zeta_{-},\zeta_{1}. Then D3D_{3} is biholomorphic to

T . . =([0,)×[0,1]){ζ+,ζ},T\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=([0,\infty)\times[0,1])\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{\zeta_{+},\zeta_{-}\right\}\subset\mathbb{C}\,,

where ζ+=(0,1)\zeta_{+}=(0,1)\in\mathbb{C} and ζ=(0,0)\zeta_{-}=(0,0)\in\mathbb{C}. The boundary segment between ζ+\zeta_{+} and ζ\zeta_{-} is called the outgoing segment.

[Uncaptioned image]

Define

{Z=(,0)×[0,1]Z+=(0,)×[0,1],\begin{cases}Z_{-}=(-\infty,0)\times[0,1]\subset\mathbb{C}\\ Z_{+}=(0,\infty)\times[0,1]\subset\mathbb{C}\,,\end{cases}

equipped with the standard complex structure jj on \mathbb{C}. We pick a positive strip-like end ε+\varepsilon^{+} near ζ+\zeta_{+}, and a negative strip-like end ε\varepsilon_{-} near ζ\zeta_{-}. That is, ε±\varepsilon_{\pm} are maps

ε+:Z+T\displaystyle\varepsilon_{+}\colon\thinspace Z_{+}\longrightarrow T
ε:ZT\displaystyle\varepsilon_{-}\colon\thinspace Z_{-}\longrightarrow T

defined in neighborhoods of ζ+\zeta_{+} and ζ\zeta_{-} respectively. Fix a family {Jt}t[0,1]𝒥(Wk,ω)\left\{J_{t}\right\}_{t\in[0,1]}\subset\mathcal{J}(W_{k},\omega) of ω\omega-compatible almost complex structures, parametrized by t[0,1]t\in[0,1]. Then consider a map

JT:T𝒥(WK,ω)J_{T}\colon\thinspace T\longrightarrow\mathcal{J}(W_{K},\omega)

which satisfies

{JT(s,t)=Jt,s>N for some N>0(ε)JT=Jt,near ζ(ε+)JT=Jt,near ζ+.\begin{cases}J_{T}(s,t)=J_{t},&s>N\text{ for some }N>0\\ (\varepsilon_{-})^{\ast}J_{T}=J_{t},&\text{near }\zeta_{-}\\ (\varepsilon_{+})^{\ast}J_{T}=J_{t},&\text{near }\zeta_{+}\,.\end{cases}

Given a generator aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) we consider maps

u:TWKu\colon\thinspace T\longrightarrow W_{K}

that satisfies the following Floer equation:

(3.3) {du+JTduj=0limsu(s,t)=a(t),t[0,1]limsu(ε+(s,t))=ξ,t[0,1]limsu(ε(s,t))=ξ,t[0,1]\begin{cases}du+J_{T}\circ du\circ j=0\\ \lim_{s\to\infty}u(s,t)=a(t),&\forall t\in[0,1]\\ \lim_{s\to\infty}u(\varepsilon_{+}(s,t))=\xi,&\forall t\in[0,1]\\ \lim_{s\to-\infty}u(\varepsilon_{-}(s,t))=\xi,&\forall t\in[0,1]\end{cases}

where the boundary conditions on uu is indicated in Fig. 8 below.

Refer to caption
Figure 8. A JTJ_{T}-holomorphic disk in (a)\mathcal{M}(a).

For a generator aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) we define (a)\mathcal{M}(a) to be the moduli space of JTJ_{T}-holomorphic maps u:TWKu\colon\thinspace T\longrightarrow W_{K} that satisfies (3.3).

Analogous to [EL17, Theorem 63] and [Abo12b, Lemma 4.2] we have

Lemma 3.3.

For generic choices of almost complex structure JTJ_{T}, the moduli space (a)\mathcal{M}(a) is a smooth orientable manifold of dimension

dim(a)=|a|.\dim\mathcal{M}(a)=-\mathinner{\!\left\lvert a\right\rvert}\,.
Proof.

See the proof of 3.4 for the proof of the statement about the dimension. Note that, because we work with a system of parallel copies of FF, JTJ_{T}-holomorphic curves can not be multiply covered, and transversality for such is achieved using standard methods as in [EES07, EL17]. ∎

Let Dm+2D_{m+2}\subset\mathbb{C} be the positively oriented unit disk with m+2m+2 boundary punctures which we denote by ζ,ζ1,,ζm,ζ+\zeta_{-},\zeta_{1},\ldots,\zeta_{m},\zeta_{+}. Let m\mathcal{R}_{m} be the Deligne–Mumford space of unit disks in the complex plane with m+1m+1 boundary punctures that are oriented counterclockwise. Let ¯m\overline{\mathcal{R}}_{m} denote the Deligne–Mumford compactification of m\mathcal{R}_{m} as in [Abo10, Section C.1] and [Sei08, Section (9f)]. Also define m\mathcal{H}_{m} to be the Deligne–Mumford space of unit disks in the complex plane with m+2m+2 boundary punctures that are oriented counterclockwise. Its Deligne–Mumford compactification is denoted by ¯m\overline{\mathcal{H}}_{m}. The boundary of ¯m\overline{\mathcal{H}}_{m} is obtained by adding broken disks and hence the codimension one boundary of ¯m\overline{\mathcal{H}}_{m} is covered by the following spaces

(3.4) ¯m1\displaystyle\overline{\mathcal{H}}_{m_{1}} ׯm2,m1+m2=m\displaystyle\times\overline{\mathcal{H}}_{m_{2}},\,m_{1}+m_{2}=m
(3.5) ¯m1\displaystyle\overline{\mathcal{H}}_{m_{1}} ׯm2,m1+m2=m+1\displaystyle\times\overline{\mathcal{R}}_{m_{2}},\,m_{1}+m_{2}=m+1

where we regard each stratum as being included in ¯m\overline{\mathcal{H}}_{m} via the natural inclusion.

Consider a word of generators akCWΛK(Fk1,Fk)a_{k}\in CW^{\ast}_{\varLambda_{K}}(F_{k-1},F_{k})

𝒂=a1am.\boldsymbol{a}=a_{1}\cdots a_{m}\,.

Then we define the moduli space (𝒂)\mathcal{M}(\boldsymbol{a}) to be maps

u:TWK,u\colon\thinspace T\longrightarrow W_{K}\,,

where T¯mT\in\overline{\mathcal{H}}_{m}, and so that uu satisfies the following Floer equation

{du+JTduj=0limsu(εk(s,t))=ak(t),t[0,1] and k{1,,m}limsu(ε+(s,t))=ξ,t[0,1]limsu(ε(s,t))=ξ,t[0,1]\begin{cases}du+J_{T}\circ du\circ j=0\\ \lim_{s\to\infty}u(\varepsilon^{k}(s,t))=a_{k}(t),&\forall t\in[0,1]\text{ and }k\in\left\{1,\ldots,m\right\}\\ \lim_{s\to\infty}u(\varepsilon_{+}(s,t))=\xi,&\forall t\in[0,1]\\ \lim_{s\to-\infty}u(\varepsilon_{-}(s,t))=\xi,&\forall t\in[0,1]\end{cases}

where ε±:Z±T\varepsilon{\pm}\colon\thinspace Z_{\pm}\longrightarrow T and εk:Z+T\varepsilon^{k}\colon\thinspace Z_{+}\longrightarrow T are strip-like ends near each puncture ζ±\zeta^{\pm} and ζk\zeta^{k} for k{1,,m}k\in\left\{1,\ldots,m\right\}. The boundary conditions of uu is indicated in Fig. 9 below

Refer to caption
Figure 9. A JTJ_{T}-holomorphic disk in (𝒂)\mathcal{M}(\boldsymbol{a}).

Again, analogous to [EL17, Theorem 63] and [Abo12b, Lemma 4.7] we have the following standard transversality result.

Lemma 3.4.

For a generic choice of almost complex structure, (𝐚)\mathcal{M}(\boldsymbol{a}) is a smooth orientable manifold of dimension

dim(𝒂)=1+mj=1m|aj|.\dim\mathcal{M}(\boldsymbol{a})=-1+m-\sum_{j=1}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}\,.
Proof.

We first observe that disks in (𝒂)\mathcal{M}(\boldsymbol{a}) have switching boundary condition which implies that they can not be multiply covered for topological reasons. Then transversality is proved using standard techniques as in [EES07, EL17].

We now prove the statement about the dimension. The proof is similar to the proof of 2.4. By [CEL10, Theorem A.1.] we have

dim(𝒂)=(n3)+m+2+μ(u,Zu),\dim\mathcal{M}(\boldsymbol{a})=(n-3)+m+2+\mu(\partial u,Z_{\partial u})\,,

where μ(u,Zu)\mu(\partial u,Z_{\partial u}) is defined as in the proof of 2.4. There is a new type of contribution coming from the Lagrangian intersection punctures ζ±\zeta_{\pm}. By definition of μ(u,Zu)\mu(\partial u,Z_{\partial u}) we see that the sum of the contributions from both ζ±\zeta_{\pm} is equal to n-n. The Maslov class of MKM_{K} vanishes (see 3.2), so the only contributions to μ(u,Zu)\mu(\partial u,Z_{\partial u}) comes from the generators 𝒂\boldsymbol{a} and the Lagrangian intersection pucntures ζ±\zeta_{\pm}. Therefore

dim(𝒂)=(n1)+mnj=1m|aj|=1+mj=1m|aj|.\dim\mathcal{M}(\boldsymbol{a})=(n-1)+m-n-\sum_{j=1}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}=-1+m-\sum_{j=1}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}\,.

Furthermore, by vanishing of the Maslov class of MKM_{K} (see 3.2) it allows us to find a coherent orientation of the moduli spaces. See Appendix B for a more general discussion about orientations. ∎

Since WKW_{K} is non-compact, we use monotonicity together with a generically chosen metric to make sure JTJ_{T}-holomorphic half strips do not escape to horizontal infinity, see Appendix A and in particular A.2. This gives that (𝒂)\mathcal{M}(\boldsymbol{a}) can be compactified by adding several-level curves and we denote the compactification by ¯(𝒂)\overline{\mathcal{M}}(\boldsymbol{a}). Similar to [Abo12b, Lemma 4.9] and by (3.4), (3.5) the codimension one boundary of ¯(𝒂)\overline{\mathcal{M}}(\boldsymbol{a}) is stratified as

(3.6) ¯(𝒂)=𝒂~𝒂t+s+r=m¯(𝒂𝒂~)×cw(𝒂~)𝒂𝒂′′=𝒂m1+m2=m¯(𝒂)ׯ(𝒂′′).\partial\overline{\mathcal{M}}(\boldsymbol{a})=\coprod_{\begin{subarray}{c}\tilde{\boldsymbol{a}}\subset\boldsymbol{a}\\ t+s+r=m\end{subarray}}\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})\amalg\coprod_{\begin{subarray}{c}\boldsymbol{a}^{\prime}\boldsymbol{a}^{\prime\prime}=\boldsymbol{a}\\ m_{1}+m_{2}=m\end{subarray}}\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})\,.

Note that we define cw(𝒂~)\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) to mean either pb(𝒂~)\mathcal{M}^{\mathrm{pb}}(\tilde{\boldsymbol{a}}) or fi,Lag(𝒂~)\mathcal{M}^{\mathrm{fi,Lag}}(\tilde{\boldsymbol{a}}) as in Section 2, depending on whether the breaking happens at a Reeb chord or a Lagrangian intersection generator.

To be more precise, (3.6) means that the codimension one boundary of ¯(𝒂)\partial\overline{\mathcal{M}}(\boldsymbol{a}) is covered by images of the natural inclusions of products ¯(𝒂𝒂~)×cw(𝒂~)\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) for subwords 𝒂~𝒂\tilde{\boldsymbol{a}}\subset\boldsymbol{a} and ¯(𝒂)ׯ(𝒂′′)\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}) for partitions

𝒂=a1am1=𝒂am1+1am1+m2=𝒂′′.\boldsymbol{a}=\underbrace{a_{1}\cdots a_{m_{1}}}_{=\boldsymbol{a}^{\prime}}\underbrace{a_{m_{1}+1}\cdots a_{m_{1}+m_{2}}}_{=\boldsymbol{a}^{\prime\prime}}\,.

Note that 𝒂𝒂~\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}} is the word of generators obtained by starting with the word 𝒂\boldsymbol{a} and replacing the subword 𝒂~\tilde{\boldsymbol{a}} with an auxiliary generator yy, see Fig. 10 and Fig. 11. If 𝒂=a1am\boldsymbol{a}=a_{1}\cdots a_{m} and 𝒂~=at+1at+s𝒂\tilde{\boldsymbol{a}}=a_{t+1}\cdots a_{t+s}\subset\boldsymbol{a} then

(3.7) 𝒂𝒂~ . . =a1atyat+s+1am=(𝒂𝒂~)1y(𝒂𝒂~)2.\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=a_{1}\cdots a_{t}ya_{t+s+1}\cdots a_{m}=(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{1}y(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2}\,.

In this case, where the auxiliary generator yy is placed at position t+1t+1 in (3.7) we say that 𝒂~𝒂\tilde{\boldsymbol{a}}\subset\boldsymbol{a} is a subword of 𝐚\boldsymbol{a} at position t+1t+1.

We now have a lemma of how the orientation of the different strata compares to the boundary orientation. See Appendix B for a general discussion about orientations of moduli spaces.

Lemma 3.5.

The product orientation on ¯(𝐚)ׯ(𝐚′′)\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}) differs from the boundary orientation on ¯(𝐚)\partial\overline{\mathcal{M}}(\boldsymbol{a}) by (1)1(-1)^{{\ddagger}_{1}} where

(3.8) 1=(m2+1)(i=1m1|ai|)+m1,{\ddagger}_{1}=(m_{2}+1)\left(\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+m_{1}\,,

while the product orientation on ¯(𝐚𝐚~)×cw(𝐚~)\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) differs from the boundary orientation on ¯(𝐚)\partial\overline{\mathcal{M}}(\boldsymbol{a}) by (1)2(-1)^{{\ddagger}_{2}} where

(3.9) 2=s(|ξ|+i=1t+s|ai|)+s(mt)+t+s,{\ddagger}_{2}=s\left(\mathinner{\!\left\lvert\xi\right\rvert}+\sum_{i=1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+s(m-t)+t+s\,,

whenever cw(𝐚~)\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) is rigid. Here 𝐚~\tilde{\boldsymbol{a}} is a subword of 𝐚\boldsymbol{a} at position t+1t+1 as in (3.7).

Proof.

See Appendix B. ∎

Lemma 3.6.

There exists a family of fundamental chains [¯(𝐚)]C(¯(𝐚))[\overline{\mathcal{M}}(\boldsymbol{a})]\in C_{\ast}(\overline{\mathcal{M}}(\boldsymbol{a})) such that

(3.10) [¯(𝒂)]=𝒂𝒂′′=𝒂(1)1[¯(𝒂)]×[¯(𝒂′′)]+𝒂~𝒂(1)2[¯(𝒂𝒂~)]×[cw(𝒂~)],\partial[\overline{\mathcal{M}}(\boldsymbol{a})]=\sum_{\boldsymbol{a}^{\prime}\boldsymbol{a}^{\prime\prime}=\boldsymbol{a}}(-1)^{{\ddagger}_{1}}[\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})]\times[\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})]+\sum_{\tilde{\boldsymbol{a}}\subset\boldsymbol{a}}(-1)^{{\ddagger}_{2}}[\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})]\times[\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})]\,,

where 1{\ddagger}_{1} and 2{\ddagger}_{2} are as in (3.8) and (3.9) respectively.

Proof.

See [Abo12b, Lemma 4.11]. ∎

3.4. The evaluation map and construction of the AA_{\infty}-homomorphism

In this section we construct the evaluation map used to define the AA_{\infty}-homomorphism between CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K}).

First pick any smooth, orientation reversing map r:Dm+2r\colon\thinspace\mathbb{R}\longrightarrow D_{m+2} which parametrizes the outgoing segment. (That is, the boundary arc of Dm+2D_{m+2} that lies between ζ+\zeta_{+} and ζ\zeta_{-}.)

Pick two strip-like ends

ε±:(0,)×[0,1]U±,\displaystyle\varepsilon_{\pm}\colon\thinspace(0,\infty)\times[0,1]\longrightarrow U_{\pm}\,,

where U±Dm+2U_{\pm}\subset D_{m+2} are neighborhoods of ζ±Dm+2\zeta_{\pm}\in D_{m+2}. We pick the strip-like ends so that ε±((0,)×{0})U±\varepsilon_{\pm}((0,\infty)\times\left\{0\right\})\subset U_{\pm} are the parts of the boundary of Dm+2D_{m+2} that points towards ζ±\zeta_{\pm} (according to the boundary orientation on Dm+2D_{m+2}), and ε±((0,)×{1})U±\varepsilon_{\pm}((0,\infty)\times\left\{1\right\})\subset U_{\pm} are the parts of the boundary of Dm+2D_{m+2} that points away from ζ±\zeta_{\pm}.

Assume that r:Dm+2r\colon\thinspace\mathbb{R}\longrightarrow D_{m+2} satisfies the following

(3.11) {limt±r(t)=ζsup|t|t~|(ε±1r)(n)(t)|<,t~>0,n1.\begin{cases}\lim_{t\to\pm\infty}r(t)=\zeta_{\mp}\\ \sup_{\mathinner{\!\left\lvert t\right\rvert}\geq\tilde{t}}\mathinner{\!\left\lvert(\varepsilon_{\pm}^{-1}\circ r)^{(n)}(t)\right\rvert}<\infty,\quad\exists\tilde{t}>0,\;\forall n\geq 1\,.\end{cases}

Then ur:MKu\circ r\colon\thinspace\mathbb{R}\longrightarrow M_{K} is a map so that limt±(ur)(t)=ξ\lim_{t\to\pm\infty}(u\circ r)(t)=\xi. We reparametrize rr by arc length with respect to a Riemannian metric on MKM_{K} (see (A.1)), and compactify the domain. In doing so, we get a smooth, orientation reversing map

r~:[0,R]Dm+2,\tilde{r}\colon\thinspace[0,R]\longrightarrow D_{m+2}\,,

that satisfies (ur~)(0)=(ur~)(R)=ξ(u\circ\tilde{r})(0)=(u\circ\tilde{r})(R)=\xi, which means ur~ΩξMKu\circ\tilde{r}\in\varOmega_{\xi}M_{K}. We then define the evaluation map as

(3.12) ev:(𝒂)\displaystyle\operatorname{ev}\colon\thinspace\mathcal{M}(\boldsymbol{a}) ΩξMK\displaystyle\longrightarrow\varOmega_{\xi}M_{K}
u\displaystyle u ur~.\displaystyle\longmapsto u\circ\tilde{r}\,.
Lemma 3.7.

Let u:Dm+2WKu\colon\thinspace D_{m+2}\longrightarrow W_{K} be a JJ-holomorphic disk and take r:Dm+2r\colon\thinspace\mathbb{R}\longrightarrow D_{m+2} so that (3.11) holds. Then sur\partial_{s}u\circ r decays exponentially in the CC^{\infty}-topology.

Proof.

Pick strip-like ends

ε±:(0,)×[0,1]U±,\displaystyle\varepsilon_{\pm}\colon\thinspace(0,\infty)\times[0,1]\longrightarrow U_{\pm}\,,

as above. By [RS01, Theorem A] we have that suε±\partial_{s}u\circ\varepsilon_{\pm} decays exponentially in the CC^{\infty}-topology. When we say that a function decays exponentially in the CC^{\infty}-toplogy we mean that there are constants δ,c0,c1,c2,>0\delta,c_{0},c_{1},c_{2},\ldots>0 so that k\forall k\in\mathbb{N} and for every t0(0,)t_{0}\in(0,\infty) we have

(3.13) suε±Ck([t0,)×[0,1])ckeδt0.\mathinner{\!\left\lVert\partial_{s}u\circ\varepsilon_{\pm}\right\rVert}_{C^{k}([t_{0},\infty)\times[0,1])}\leq c_{k}e^{-\delta t_{0}}\,.

Next consider r:Dm+2r\colon\thinspace\mathbb{R}\longrightarrow D_{m+2} which satisfies (3.11), where t~>0\tilde{t}>0 is large enough so that r(t)U±r(t)\in U_{\pm} for |t|>t~\mathinner{\!\left\lvert t\right\rvert}>\tilde{t}. This also gives

(ur)(t)={(uε)(ε1r)(t),tt~(uε+)(ε+1r)(t),tt~(u\circ r)(t)=\begin{cases}\left(u\circ\varepsilon_{-}\right)\circ(\varepsilon_{-}^{-1}\circ r)(t),&t\geq\tilde{t}\\ \left(u\circ\varepsilon_{+}\right)\circ(\varepsilon_{+}^{-1}\circ r)(t),&t\leq-\tilde{t}\end{cases}

where ε±1r:(0,)×[0,1]\varepsilon_{\pm}^{-1}\circ r\colon\thinspace\mathbb{R}\longrightarrow(0,\infty)\times[0,1] are maps so that

{(ε1r)(t)(0,)×{1}(ε+1r)(t)(0,)×{0},\begin{cases}(\varepsilon_{-}^{-1}\circ r)(t)\subset(0,\infty)\times\left\{1\right\}\\ (\varepsilon_{+}^{-1}\circ r)(t)\subset(0,\infty)\times\left\{0\right\}\,,\end{cases}

and

{limt(ε1r)(t)=(,1)limt(ε+1r)(t)=(,0).\begin{cases}\lim_{t\to\infty}(\varepsilon_{-}^{-1}\circ r)(t)=(\infty,1)\\ \lim_{t\to-\infty}(\varepsilon_{+}^{-1}\circ r)(t)=(\infty,0)\,.\end{cases}

Then we have constants δ,c0,c1,c2,>0\delta,c_{0},c_{1},c_{2},\ldots>0 so that k0\forall k\geq 0

surCk([t~,))\displaystyle\mathinner{\!\left\lVert\partial_{s}u\circ r\right\rVert}_{C^{k}([\tilde{t},\infty))} =|α|ksup|t|t~|Dα(sur)|\displaystyle=\sum_{\mathinner{\!\left\lvert\alpha\right\rvert}\leq k}\sup_{\mathinner{\!\left\lvert t\right\rvert}\geq\tilde{t}}\mathinner{\!\left\lvert D^{\alpha}(\partial_{s}u\circ r)\right\rvert}
=|α|ksup|t|t~|Dα[suε±](ε±1(r(t)))Dα[ε±1r](t)|\displaystyle=\sum_{\mathinner{\!\left\lvert\alpha\right\rvert}\leq k}\sup_{\mathinner{\!\left\lvert t\right\rvert}\geq\tilde{t}}\mathinner{\!\left\lvert D^{\alpha}\left[\partial_{s}u\circ\varepsilon_{\pm}\right](\varepsilon_{\pm}^{-1}(r(t)))\cdot D^{\alpha}\left[\varepsilon_{\pm}^{-1}\circ r\right](t)\right\rvert}
=|α|k[sup[t0,)×[0,1]|Dα(suε±)|][sup|t|t~|Dα(ε±1r)|].\displaystyle=\sum_{\mathinner{\!\left\lvert\alpha\right\rvert}\leq k}\left[\sup_{[t_{0},\infty)\times[0,1]}\mathinner{\!\left\lvert D^{\alpha}(\partial_{s}u\circ\varepsilon_{\pm})\right\rvert}\right]\left[\sup_{\mathinner{\!\left\lvert t\right\rvert}\geq\tilde{t}}\mathinner{\!\left\lvert D^{\alpha}(\varepsilon_{\pm}^{-1}\circ r)\right\rvert}\right]\,.

Here DαD^{\alpha} denotes derivative with respect to the multi-index α\alpha. Because of (3.11) we have

sup|t|t~|Dα(ε±1r)|Aα,\sup_{\mathinner{\!\left\lvert t\right\rvert}\geq\tilde{t}}\mathinner{\!\left\lvert D^{\alpha}(\varepsilon_{\pm}^{-1}\circ r)\right\rvert}\leq A_{\alpha}\,,

where AαA_{\alpha} is some constant depending on α\alpha. We conclude

surCk([t~,))Aksuε±Ck([t0,)×[0,1])Akckeδt0,\mathinner{\!\left\lVert\partial_{s}u\circ r\right\rVert}_{C^{k}([\tilde{t},\infty))}\leq A_{k}\mathinner{\!\left\lVert\partial_{s}u\circ\varepsilon_{\pm}\right\rVert}_{C^{k}([t_{0},\infty)\times[0,1])}\leq A_{k}\cdot c_{k}e^{-\delta t_{0}}\,,

by (3.13), where Ak . . =max|a|kAαA_{k}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\max_{\mathinner{\!\left\lvert a\right\rvert}\leq k}A_{\alpha}. Furthermore we note that t~>0\tilde{t}>0 is large enough so that for |t|t~\mathinner{\!\left\lvert t\right\rvert}\geq\tilde{t} we have

{(ε1r)(t)[t0,)×{1}(ε+1r)(t)[t0,)×{0}.\begin{cases}(\varepsilon_{-}^{-1}\circ r)(t)\subset[t_{0},\infty)\times\left\{1\right\}\\ (\varepsilon_{+}^{-1}\circ r)(t)\subset[t_{0},\infty)\times\left\{0\right\}\,.\end{cases}

The previous lemma enables us to extend the evaluation map to the compactification of the moduli space of half strips.

Lemma 3.8.

There is an extension of the evaluation map ev\operatorname{ev} to a continuous map on the compactification of (𝐚)\mathcal{M}(\boldsymbol{a}),

ev:¯(𝒂)ΩξMK,\operatorname{ev}\colon\thinspace\overline{\mathcal{M}}(\boldsymbol{a})\longrightarrow\varOmega_{\xi}M_{K}\,,

such that the following diagram commutes up to an overall sign of (1)1(-1)^{{\ddagger}_{1}}, where 1{\ddagger}_{1} is defined in (3.8).

¯(𝒂)ׯ(𝒂′′){\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})}¯(𝒂){\overline{\mathcal{M}}(\boldsymbol{a})}ΩξMK×ΩξMK{\varOmega_{\xi}M_{K}\times\varOmega_{\xi}M_{K}}ΩξMK,{\varOmega_{\xi}M_{K}\,,}ι\scriptstyle{\iota}ev×ev\scriptstyle{\operatorname{ev}\times\operatorname{ev}}ev\scriptstyle{\operatorname{ev}}\scriptstyle{\circ}

The map ι\iota in the top row is inclusion as in (3.6). The map in the bottom row is concatenation of loops.

Proof.

For this proof, we follow the idea outlined in [Abo12b, p. 37].

Extension of ev\operatorname{ev} to the compactification:

It is obvious how to extend it to the boundary strata ¯(𝒂𝒂~)×cw(𝒂~)\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\text{cw}}(\tilde{\boldsymbol{a}}); we define the evaluation map of such broken disk to be the same as the evaluation map when we forget about the factor cw(𝒂~)\mathcal{M}^{\text{cw}}(\tilde{\boldsymbol{a}}). However, if we have a sequence {uν}ν=0(𝒂)\left\{u^{\nu}\right\}_{\nu=0}^{\infty}\subset\mathcal{M}(\boldsymbol{a}) which Gromov converges to a broken disk in any of the boundary strata ¯(𝒂)ׯ(𝒂′′)\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}), then the Gromov limit is a stable JJ-holomorphic map (a broken disk), consisting of two JJ-holomorphic disks ui:DkiWKu_{i}\colon\thinspace D_{k_{i}}\longrightarrow W_{K} where k1+k22=m+2k_{1}+k_{2}-2=m+2, and two boundary punctures z1Dk1z_{1}\in\partial D_{k_{1}}, z2Dk2z_{2}\in\partial D_{k_{2}} so that we either have (z1,z2)=(ζ,ζ+)(z_{1},z_{2})=(\zeta_{-},\zeta_{+}) or (z1,z2)=(ζ+,ζ)(z_{1},z_{2})=(\zeta_{+},\zeta_{-}) [Fra08]. More precisely, it means that there are two families of Möbius transformations of the unit disk DD\subset\mathbb{C}

φ1ν,φ2ν:DD, where ν,\varphi_{1}^{\nu},\varphi_{2}^{\nu}\colon\thinspace D\longrightarrow D,\text{ where }\nu\in\mathbb{N}\,,

so that

(3.14) {uνφ1νu1in Cloc(Dk1{z1})uνφ2νu2in Cloc(Dk2{z2}),\begin{cases}u^{\nu}\circ\varphi_{1}^{\nu}\longrightarrow u_{1}&\text{in }C^{\infty}_{\text{loc}}(D_{k_{1}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{1}\right\})\\ u^{\nu}\circ\varphi_{2}^{\nu}\longrightarrow u_{2}&\text{in }C^{\infty}_{\text{loc}}(D_{k_{2}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{2}\right\})\,,\end{cases}

and

{(φ1ν)1φ2νz1in Cloc(Dk1{z1})(φ2ν)1φ1νz2in Cloc(Dk2{z2}).\begin{cases}(\varphi_{1}^{\nu})^{-1}\circ\varphi_{2}^{\nu}\longrightarrow z_{1}&\text{in }C^{\infty}_{\text{loc}}(D_{k_{1}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{1}\right\})\\ (\varphi_{2}^{\nu})^{-1}\circ\varphi_{1}^{\nu}\longrightarrow z_{2}&\text{in }C^{\infty}_{\text{loc}}(D_{k_{2}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{2}\right\})\,.\end{cases}

Recall that convergence in Cloc(X)C^{\infty}_{\text{loc}}(X) means CC^{\infty}-convergence on every compact subset KXK\subset X.

Define parametrizations

r1:\displaystyle r_{1}\colon\thinspace\mathbb{R} Dk1\displaystyle\longrightarrow D_{k_{1}}
r2:\displaystyle r_{2}\colon\thinspace\mathbb{R} Dk2\displaystyle\longrightarrow D_{k_{2}}

so that r1r_{1} and r2r_{2} satisfy (3.11). Then the two maps uiri:MKu_{i}\circ r_{i}\colon\thinspace\mathbb{R}\longrightarrow M_{K} are smooth maps so that suiri\partial_{s}u_{i}\circ r_{i} decay exponentially in the CC^{\infty}-topology by 3.7. Hence the composition of two smooth loops uiriu_{i}\circ r_{i} is again a smooth loop. There are two cases, depending on whether the two components of the broken disk have the puncture ζ+\zeta_{+} or ζ\zeta_{-} in common. That is, we either have (z1,z2)=(ζ,ζ+)(z_{1},z_{2})=(\zeta_{-},\zeta_{+}) or (z1,z2)=(ζ+,ζ)(z_{1},z_{2})=(\zeta_{+},\zeta_{-}). In the first case when (z1,z2)=(ζ,ζ+)(z_{1},z_{2})=(\zeta_{-},\zeta_{+}), we define a map γ:MK\gamma\colon\thinspace\mathbb{R}\longrightarrow M_{K} as

(3.15) γ(t) . . ={(u1r1)(t1t),t<0ξ,t=0(u2r2)(t1t),t>0.\gamma(t)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\begin{cases}(u_{1}\circ r_{1})\left(t-\frac{1}{t}\right),&t<0\\ \xi,&t=0\\ (u_{2}\circ r_{2})\left(t-\frac{1}{t}\right),&t>0\,.\end{cases}

In the second case when (z1,z2)=(ζ+,ζ)(z_{1},z_{2})=(\zeta_{+},\zeta_{-}) we swap places of u1r1u_{1}\circ r_{1} and u2r2u_{2}\circ r_{2} in the above definition of γ\gamma.

We then claim that this map is smooth and has exponentially decaying derivatives in the CC^{\infty}-topology as t±t\to\pm\infty. Since u1r1u_{1}\circ r_{1} and u2r2u_{2}\circ r_{2} are smooth maps with exponentially decaying derivatives in the CC^{\infty}-topology as t±t\to\pm\infty, it suffices to show that all derivatives of γ\gamma at t=0t=0 exists. This follows from the exponential decay of every derivative of u1r1u_{1}\circ r_{1} and u2r2u_{2}\circ r_{2} in the CC^{\infty}-topology. We may then reparametrize γ\gamma by arc length and compactify the domain to obtain a map γ~:[0,R]MK\tilde{\gamma}\colon\thinspace[0,R]\longrightarrow M_{K} so that γ~(0)=γ~(R)=ξ\tilde{\gamma}(0)=\tilde{\gamma}(R)=\xi, that is γ~ΩξMK\tilde{\gamma}\in\varOmega_{\xi}M_{K}, and we define ev((u1,u2)) . . =γ~\operatorname{ev}((u_{1},u_{2}))\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\tilde{\gamma}.

Commutativity of the diagram:

It follows almost immediately from the definition of the evaluation map

ev:¯(𝒂)ΩξMK\operatorname{ev}\colon\thinspace\overline{\mathcal{M}}(\boldsymbol{a})\longrightarrow\varOmega_{\xi}M_{K}

that the diagram

¯(𝒂)ׯ(𝒂){\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})}¯(𝒂){\overline{\mathcal{M}}(\boldsymbol{a})}ΩξMK×ΩξMK{\varOmega_{\xi}M_{K}\times\varOmega_{\xi}M_{K}}ΩξMK{\varOmega_{\xi}M_{K}}ι\scriptstyle{\iota}ev×ev\scriptstyle{\operatorname{ev}\times\operatorname{ev}}ev\scriptstyle{\operatorname{ev}}\scriptstyle{\circ}

commutes, since γ\gamma in (3.15) is essentially defined as the concatenation of u1r1u_{1}\circ r_{1} and u2r2u_{2}\circ r_{2}. More precisely, we consider uiri:MKu_{i}\circ r_{i}\colon\thinspace\mathbb{R}\longrightarrow M_{K} for i{1,2}i\in\left\{1,2\right\} as above. Then reparametrize r1r_{1} and r2r_{2} by arc length so that we obtain two maps

uir~i:[0,Ri]MK.u_{i}\circ\tilde{r}_{i}\colon\thinspace[0,R_{i}]\longrightarrow M_{K}\,.

These maps are so that (uir~i)(0)=(uir~i)(Ri)=ξ(u_{i}\circ\tilde{r}_{i})(0)=(u_{i}\circ\tilde{r}_{i})(R_{i})=\xi for i{1,2}i\in\left\{1,2\right\}, and the concatenation of these maps yields a map ψ:[0,R1+R2]MK\psi\colon\thinspace[0,R_{1}+R_{2}]\longrightarrow M_{K} defined by

ψ(t)={(u1r~1)(t),t[0,R1](u2r~2)(tR1),t[R1,R1+R2]\displaystyle\psi(t)=\begin{cases}(u_{1}\circ\tilde{r}_{1})(t),&t\in[0,R_{1}]\\ (u_{2}\circ\tilde{r}_{2})(t-R_{1}),&t\in[R_{1},R_{1}+R_{2}]\end{cases}

which coincides with the map γ~:[0,R]MK\tilde{\gamma}\colon\thinspace[0,R]\longrightarrow M_{K} obtained by parametrizing γ\gamma defined in (3.15) by arc length. The overall sign (1)1(-1)^{{\ddagger}_{1}} comes from 3.5, see Appendix B for a discussion about sign and orientations.

Continuity of ev\operatorname{ev}:

We claim that ev\operatorname{ev} is a continuous map, meaning that if {uν}ν=0(𝒂)\left\{u^{\nu}\right\}_{\nu=0}^{\infty}\subset\mathcal{M}(\boldsymbol{a}) is a Gromov convergent sequence of JJ-holomorphic disks, then the map γ~(t)\tilde{\gamma}(t) defined in (3.15) is realized as a limit of loops in the compact-open topology of ΩξMK\varOmega_{\xi}M_{K}.

Pick a family of smooth maps {rν:Dm+2}ν=0\left\{r^{\nu}\colon\thinspace\mathbb{R}\longrightarrow D_{m+2}\right\}_{\nu=0}^{\infty} which satisfies (3.11). Then we have that {uνrν}ν=0\left\{u^{\nu}\circ r^{\nu}\right\}_{\nu=0}^{\infty} is a family of smooth maps with exponentially decaying derivatives as t±t\to\pm\infty in the CC^{\infty}-topology by 3.7. From (3.14) we have two families of Möbius transformations {φ1ν}ν=0\left\{\varphi_{1}^{\nu}\right\}_{\nu=0}^{\infty} and {φ2ν}ν=0\left\{\varphi_{2}^{\nu}\right\}_{\nu=0}^{\infty} such that

uνφiνui,in Cloc(Dki{zi}),u^{\nu}\circ\varphi_{i}^{\nu}\to u_{i},\quad\text{in }C^{\infty}_{\text{loc}}(D_{k_{i}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{i}\right\})\,,

for i{1,2}i\in\left\{1,2\right\}. We also have that φiν\varphi_{i}^{\nu} preserves the boundary of DmD_{m} and that (φiν)1\left(\varphi_{i}^{\nu}\right)^{-1} preserves boundary marked points in the sense that limν(φiν)1(ζj)=ζj\lim_{\nu\to\infty}\left(\varphi_{i}^{\nu}\right)^{-1}(\zeta_{j})=\zeta_{j}. Then we have

(3.16) {(φ1ν)1rνr1,in Cloc(<0)(φ2ν)1rνr2,in Cloc(>0).\begin{cases}\left(\varphi_{1}^{\nu}\right)^{-1}\circ r^{\nu}\longrightarrow r_{1},\quad\text{in }C^{\infty}_{\text{loc}}(\mathbb{R}_{<0})\\ \left(\varphi_{2}^{\nu}\right)^{-1}\circ r^{\nu}\longrightarrow r_{2},\quad\text{in }C^{\infty}_{\text{loc}}(\mathbb{R}_{>0})\,.\end{cases}

Hence for any multi-index α\alpha and i{1,2}i\in\left\{1,2\right\} we have

(3.17) |Dα(uνrν)Dα(uiri)|\displaystyle\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ r^{\nu})-D^{\alpha}(u_{i}\circ r_{i})\right\rvert} =|Dα(uνφiν(φiν)1rν)Dα(uiri)|\displaystyle=\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ\varphi_{i}^{\nu}\circ\left(\varphi_{i}^{\nu}\right)^{-1}\circ r^{\nu})-D^{\alpha}(u_{i}\circ r_{i})\right\rvert}
=|Dα(uνφiν)Dα[(φiν)1rν]Dα(ui)Dα(ri)|\displaystyle=\mathinner{\!\left\lvert D^{\alpha}\left(u^{\nu}\circ\varphi_{i}^{\nu}\right)D^{\alpha}\left[\left(\varphi_{i}^{\nu}\right)^{-1}\circ r^{\nu}\right]-D^{\alpha}(u_{i})D^{\alpha}(r_{i})\right\rvert}
|Dα(uνφiν)(Dα[(φiν)1rν]Dα(ri))|\displaystyle\leq\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ\varphi_{i}^{\nu})\cdot\left(D^{\alpha}\left[\left(\varphi_{i}^{\nu}\right)^{-1}\circ r^{\nu}\right]-D^{\alpha}(r_{i})\right)\right\rvert}
+|Dα(ri)[Dα(uνφiν)Dα(ui)]|.\displaystyle\qquad+\mathinner{\!\left\lvert D^{\alpha}(r_{i})\cdot\left[D^{\alpha}\left(u^{\nu}\circ\varphi_{i}^{\nu}\right)-D^{\alpha}(u_{i})\right]\right\rvert}\,.

Let i{1,2}i\in\left\{1,2\right\} and define 1 . . =<0\mathbb{R}_{1}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mathbb{R}_{<0} and 2 . . =>0\mathbb{R}_{2}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mathbb{R}_{>0}. Inserting suprema over suitable compact sets AiA\subset\mathbb{R}_{i} and KDki{zi}K\subset D_{k_{i}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{i}\right\} gives

(3.18) supAi|Dα(uνrν)Dα(uiri)|\displaystyle\sup_{A\subset\mathbb{R}_{i}}\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ r^{\nu})-D^{\alpha}(u_{i}\circ r_{i})\right\rvert}
(3.17)supKDki{zi}|Dα(uνφiν)|supAi|Dα((φiν)1rν)Dα(ri)|\displaystyle\overset{\eqref{eq:alpha_derivatives_difference}}{\leq}\sup_{K\subset D_{k_{i}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{i}\right\}}\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ\varphi_{i}^{\nu})\right\rvert}\sup_{A\subset\mathbb{R}_{i}}\mathinner{\!\left\lvert D^{\alpha}\left(\left(\varphi_{i}^{\nu}\right)^{-1}\circ r^{\nu}\right)-D^{\alpha}(r_{i})\right\rvert}
+supAi|Dα(ri)|supKDki{zi}|Dα(uνφiν)Dα(ui)|\displaystyle\qquad+\sup_{A\subset\mathbb{R}_{i}}\mathinner{\!\left\lvert D^{\alpha}(r_{i})\right\rvert}\sup_{K\subset D_{k_{i}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{i}\right\}}\mathinner{\!\left\lvert D^{\alpha}\left(u^{\nu}\circ\varphi_{i}^{\nu}\right)-D^{\alpha}(u_{i})\right\rvert}
C1supAi|Dα((φiν)1rν)Dα(ri)|0+C2supKDki{zi}|Dα(uνφiν)Dα(ui)|0,\displaystyle\leq C_{1}\underbrace{\sup_{A\subset\mathbb{R}_{i}}\mathinner{\!\left\lvert D^{\alpha}\left(\left(\varphi_{i}^{\nu}\right)^{-1}\circ r^{\nu}\right)-D^{\alpha}(r_{i})\right\rvert}}_{\longrightarrow 0}+C_{2}\underbrace{\sup_{K\subset D_{k_{i}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{i}\right\}}\mathinner{\!\left\lvert D^{\alpha}\left(u^{\nu}\circ\varphi_{i}^{\nu}\right)-D^{\alpha}(u_{i})\right\rvert}}_{\longrightarrow 0}\,,

Here we have used that uνφiνuiu^{\nu}\circ\varphi_{i}^{\nu}\longrightarrow u_{i} in Cloc(Dki{zi})C^{\infty}_{\mathrm{loc}}(D_{k_{i}}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\left\{z_{i}\right\}) and hence that uνφiνu^{\nu}\circ\varphi_{i}^{\nu} is also bounded in this topology. Furthermore we have used that (φiν)1rνri\left(\varphi_{i}^{\nu}\right)^{-1}\circ r^{\nu}\longrightarrow r_{i} in Cloc(i)C^{\infty}_{\text{loc}}(\mathbb{R}_{i}) by (3.16).

Then by recalling the definition of γ(t)\gamma(t) in (3.15), we have

supA|Dα(uνrν)Dα(γ)|\displaystyle\sup_{A\subset\mathbb{R}}\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ r^{\nu})-D^{\alpha}(\gamma)\right\rvert}
supA<0|Dα(uνrν)Dα(γ)|+supA>0|Dα(uνrν)Dα(γ)|\displaystyle\leq\sup_{A\subset\mathbb{R}_{<0}}\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ r^{\nu})-D^{\alpha}(\gamma)\right\rvert}+\sup_{A\subset\mathbb{R}_{>0}}\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ r^{\nu})-D^{\alpha}(\gamma)\right\rvert}
=supA<0|Dα(uνrν)Dα(u1r1)|+supA>0|Dα(uνrν)Dα(u2r2)|\displaystyle=\sup_{A\subset\mathbb{R}_{<0}}\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ r^{\nu})-D^{\alpha}(u_{1}\circ r_{1})\right\rvert}+\sup_{A\subset\mathbb{R}_{>0}}\mathinner{\!\left\lvert D^{\alpha}(u^{\nu}\circ r^{\nu})-D^{\alpha}(u_{2}\circ r_{2})\right\rvert}

By (3.18), we get uνrνγu^{\nu}\circ r^{\nu}\longrightarrow\gamma in Cloc()C^{\infty}_{\mathrm{loc}}(\mathbb{R}), and thus by passing to arc length parametrizations we get ev(uν)ev(u)\operatorname{ev}(u^{\nu})\longrightarrow\operatorname{ev}(u) in the compact-open topology on ΩξMK\varOmega_{\xi}M_{K}.

The evaluation map

ev:¯(𝒂)ΩξMK\operatorname{ev}\colon\thinspace\overline{\mathcal{M}}(\boldsymbol{a})\longrightarrow\varOmega_{\xi}M_{K}

induces a map on chains ev:C(¯(𝒂))C(ΩξMK)\operatorname{ev}_{\ast}\colon\thinspace C_{-\ast}(\overline{\mathcal{M}}(\boldsymbol{a}))\longrightarrow C_{-\ast}(\varOmega_{\xi}M_{K}). We then pick a fundamental chain [¯(𝒂)][\overline{\mathcal{M}}(\boldsymbol{a})] by 3.6 so that (3.10) holds, and define a family of maps {Ψm}m=1\left\{\varPsi_{m}\right\}_{m=1}^{\infty}

(3.19) Ψm:CWΛK(Fm1,Fm)CWΛK(F0,F1)\displaystyle\varPsi_{m}\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F_{m-1},F_{m})\otimes\cdots\otimes CW^{\ast}_{\varLambda_{K}}(F_{0},F_{1}) C(ΩξMK)\displaystyle\longrightarrow C_{-\ast}(\varOmega_{\xi}M_{K})
ama1\displaystyle a_{m}\otimes\cdots\otimes a_{1} (1)§ev[¯(𝒂)],\displaystyle\longmapsto(-1)^{\S}\operatorname{ev}_{\ast}[\overline{\mathcal{M}}(\boldsymbol{a})]\,,

where

§=j=1mj|aj|+(m+1)|ξ|+(|ξ|+m)dim¯(𝒂)=j=1mj|aj|+(|ξ|+m)j=1m|aj|(mod2).\S=\sum_{j=1}^{m}j\mathinner{\!\left\lvert a_{j}\right\rvert}+(m+1)\mathinner{\!\left\lvert\xi\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m)\dim\overline{\mathcal{M}}(\boldsymbol{a})=\sum_{j=1}^{m}j\mathinner{\!\left\lvert a_{j}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m)\sum_{j=1}^{m}\mathinner{\!\left\lvert a_{j}\right\rvert}\pmod{2}\,.

Note that |ξ|\mathinner{\!\left\lvert\xi\right\rvert} means the grading of ξ\xi regarded as an intersection generator of CWΛK(F0,F1)CW^{\ast}_{\varLambda_{K}}(F_{0},F_{1}) as in Section 2.1.

Lemma 3.9.

The following diagram commutes,

Ck(¯(𝒂)){C_{-k}(\overline{\mathcal{M}}(\boldsymbol{a}))}Ck+1(¯(𝒂)){C_{-k+1}(\overline{\mathcal{M}}(\boldsymbol{a}))}Ck(ΩξMK){C_{-k}(\varOmega_{\xi}M_{K})}Ck+1(ΩξMK){C_{-k+1}(\varOmega_{\xi}M_{K})}ev\scriptstyle{\operatorname{ev}_{\ast}}\scriptstyle{\partial}ev\scriptstyle{\operatorname{ev}_{\ast}}\scriptstyle{\partial}

and the following diagram commutes up to an overall sign of (1)1+dim¯(𝐚)(-1)^{{\ddagger}_{1}+\dim\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})}, where 1{\ddagger}_{1} is defined in (3.8).

Ck(¯(𝒂))C(¯(𝒂′′)){C_{-k}(\overline{\mathcal{M}}(\boldsymbol{a}^{\prime}))\otimes C_{-\ell}(\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}))}C(k+)(¯(𝒂)){C_{-(k+\ell)}(\overline{\mathcal{M}}(\boldsymbol{a}))}Ck(ΩξMK)C(ΩξMK){C_{-k}(\varOmega_{\xi}M_{K})\otimes C_{-\ell}(\varOmega_{\xi}M_{K})}C(k+)(ΩξMK){C_{-(k+\ell)}(\varOmega_{\xi}M_{K})}ι×\scriptstyle{\iota_{\ast}\circ\times}evev\scriptstyle{\operatorname{ev}_{\ast}\otimes\operatorname{ev}_{\ast}}ev\scriptstyle{\operatorname{ev}_{\ast}}P\scriptstyle{P}

In the latter diagram we have the subdivision 𝐚=𝐚𝐚′′\boldsymbol{a}=\boldsymbol{a}^{\prime}\boldsymbol{a}^{\prime\prime}, and the map ι\iota_{\ast} is the composition of the map induced by the inclusion

ι:¯(𝒂)ׯ(𝒂′′)¯(𝒂).\iota\colon\thinspace\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})\longrightarrow\overline{\mathcal{M}}(\boldsymbol{a})\,.
Proof.

That the first diagram commutes follows more or less by definition. Namely, let ACk(¯(𝒂))A\in C_{-k}(\overline{\mathcal{M}}(\boldsymbol{a})). Then ev(A)=evA\operatorname{ev}_{\ast}(A)=\operatorname{ev}\circ A, and by using the definition of \partial in (3.1) and the definition of ev\operatorname{ev} in (3.12) we get

(ev(A))\displaystyle\partial(\operatorname{ev}_{\ast}(A)) =i=1kε=01(1)i+ε(evA)(δi,ε(x))=i=1kε=01(1)i+ε(A(δi,ε(x))r~)\displaystyle=\sum_{i=1}^{k}\sum_{\varepsilon=0}^{1}(-1)^{i+\varepsilon}(\operatorname{ev}\circ A)(\delta_{i,\varepsilon}(x))=\sum_{i=1}^{k}\sum_{\varepsilon=0}^{1}(-1)^{i+\varepsilon}(A(\delta_{i,\varepsilon}(x))\circ\tilde{r})
=(i=1kε=01(1)i+εA(δi,ε(x)))r~\displaystyle=\left(\sum_{i=1}^{k}\sum_{\varepsilon=0}^{1}(-1)^{i+\varepsilon}A(\delta_{i,\varepsilon}(x))\right)\circ\tilde{r}
=ev(A).\displaystyle=\operatorname{ev}_{\ast}(\partial A)\,.

The second diagram is split up into the following digram

Ck(¯(𝒂))C(¯(𝒂′′)){C_{-k}(\overline{\mathcal{M}}(\boldsymbol{a}^{\prime}))\otimes C_{-\ell}(\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}))}C(k+)(¯(𝒂)ׯ(𝒂′′)){C_{-(k+\ell)}(\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}))}C(k+)(¯(𝒂)){C_{-(k+\ell)}(\overline{\mathcal{M}}(\boldsymbol{a}))}Ck(ΩξMK)C(ΩξMK){C_{-k}(\varOmega_{\xi}M_{K})\otimes C_{-\ell}(\varOmega_{\xi}M_{K})}C(k+)((ΩξMK)2){C_{-(k+\ell)}((\varOmega_{\xi}M_{K})^{2})}C(k+)(ΩξMK).{C_{-(k+\ell)}(\varOmega_{\xi}M_{K})\,.}evev\scriptstyle{\operatorname{ev}_{\ast}\otimes\operatorname{ev}_{\ast}}×\scriptstyle{\times}ι\scriptstyle{\iota_{\ast}}ev×ev\scriptstyle{\operatorname{ev}_{\ast}\times\operatorname{ev}_{\ast}}ev\scriptstyle{\operatorname{ev}_{\ast}}×\scriptstyle{\times}P\scriptstyle{P}\scriptstyle{\circ}

The right square commutes, since the corresponding diagram before application of CC_{-\ast} commutes, by 3.8, and the maps ι\iota_{\ast}, ev\operatorname{ev}_{\ast} and \circ on chains are defined pointwise. The left square also commutes, because evev\operatorname{ev}_{\ast}\otimes\operatorname{ev}_{\ast} and ev×ev\operatorname{ev}_{\ast}\times\operatorname{ev}_{\ast} act componentwise. Hence the outer square also commutes.

The overall sign (1)1+dim(¯(𝒂))(-1)^{{\ddagger}_{1}+\dim(\overline{\mathcal{M}}(\boldsymbol{a}^{\prime}))} comes from the definition of PP in (3.2), and from the inclusion

ι:¯(𝒂)ׯ(𝒂′′)¯(𝒂),\iota\colon\thinspace\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})\longrightarrow\overline{\mathcal{M}}(\boldsymbol{a})\,,

of ¯(𝒂)ׯ(𝒂′′)\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}) as a boundary stratum of ¯(𝒂)\overline{\mathcal{M}}(\boldsymbol{a}) as in (3.10). ∎

Lemma 3.10.

The maps {Ψm}m=1\left\{\varPsi_{m}\right\}_{m=1}^{\infty} form an AA_{\infty}-homomorphism. That is,

Ψm+m1+m2=mP(Ψm2Ψm1)=r+s+t=m(1)tΨr+1+t(idrμsidt),\partial\varPsi_{m}+\sum_{m_{1}+m_{2}=m}P(\varPsi_{m_{2}}\otimes\varPsi_{m_{1}})=\sum_{r+s+t=m}(-1)^{\maltese_{t}}\varPsi_{r+1+t}(\operatorname{id}^{\otimes r}\otimes\mu^{s}\otimes\operatorname{id}^{\otimes t})\,,

where

t=t+j=1t|xj|.\maltese_{t}=t+\sum_{j=1}^{t}\mathinner{\!\left\lvert x_{j}\right\rvert}\,.
Proof.

From 3.4 it is clear that Ψm\varPsi_{m} has degree 1m1-m.

We first ignore signs and prove the statement modulo 2. We look at the codimension one boundary of ¯(𝒂)\overline{\mathcal{M}}(\boldsymbol{a}) of dimension dd. It consists of two types of broken JJ-holomorphic curves as in (3.10), and we analyze each boundary term separately.

  1. (1)

    The first boundary term is

    𝒂~𝒂¯(𝒂𝒂~)×cw(𝒂~),\coprod_{\tilde{\boldsymbol{a}}\subset\boldsymbol{a}}\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})\,,

    where 𝒂~𝒂\tilde{\boldsymbol{a}}\subset\boldsymbol{a} is a subword at position t+1t+1 of 𝒂\boldsymbol{a}.

  2. (2)

    The second boundary term is

    𝒂𝒂′′=𝒂¯(𝒂)ׯ(𝒂′′),\coprod_{\boldsymbol{a}^{\prime}\boldsymbol{a}^{\prime\prime}=\boldsymbol{a}}\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})\,,

    and it consists of broken half strips that is broken at the Lagrangian intersection point ξ\xi.

In view of (3.10), we consider the fundamental chain of ¯(𝒂)\partial\overline{\mathcal{M}}(\boldsymbol{a}). Consider the natural inclusions of the boundary strata

ι:¯(𝒂)ׯ(𝒂′′)\displaystyle\iota\colon\thinspace\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}) ¯(𝒂)\displaystyle\longrightarrow\overline{\mathcal{M}}(\boldsymbol{a})
ι:¯(𝒂𝒂~)×cw(𝒂~)\displaystyle\iota\colon\thinspace\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) ¯(𝒂).\displaystyle\longrightarrow\overline{\mathcal{M}}(\boldsymbol{a})\,.

We consider Ψm(ama1)\partial\varPsi_{m}(a_{m}\otimes\cdots\otimes a_{1}) and use 3.9. Then

(3.20) Ψm(ama1)\displaystyle\partial\varPsi_{m}(a_{m}\otimes\cdots\otimes a_{1}) =ev[¯(𝒂)]=ev[¯(𝒂)]\displaystyle=\partial\operatorname{ev}_{\ast}[\overline{\mathcal{M}}(\boldsymbol{a})]=\operatorname{ev}_{\ast}\partial[\overline{\mathcal{M}}(\boldsymbol{a})]
=𝒂𝒂′′=𝒂ev(ι([¯(𝒂)]×[¯(𝒂′′)]))\displaystyle=\sum_{\boldsymbol{a}^{\prime}\boldsymbol{a}^{\prime\prime}=\boldsymbol{a}}\operatorname{ev}_{\ast}\left(\iota_{\ast}\left([\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})]\times[\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})]\right)\right)
+𝒂~𝒂ev(ι([¯(𝒂𝒂~)]×[cw(𝒂~)]))\displaystyle\quad+\sum_{\tilde{\boldsymbol{a}}\subset\boldsymbol{a}}\operatorname{ev}_{\ast}\left(\iota_{\ast}\left([\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})]\times[\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})]\right)\right)

We start by considering boundary terms of type (1). The evaluation applied to these terms is

evι([¯(𝒂𝒂~)]×[cw(𝒂~)])=ev[¯((𝒂𝒂~)1μs(𝒂~)(𝒂𝒂~)2)],\operatorname{ev}_{\ast}\iota_{\ast}\left([\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})]\times[\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})]\right)=\operatorname{ev}_{\ast}[\overline{\mathcal{M}}((\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{1}\mu^{s}(\tilde{\boldsymbol{a}})(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2})]\,,

because of the definition of ev\operatorname{ev} on these boundary strata. Note that if cw(𝒂~)\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) is not rigid, then the image evι([¯(𝒂𝒂~)]×[cw(𝒂~)])\operatorname{ev}_{\ast}\iota_{\ast}\left([\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})]\times[\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})]\right) would be degenerate in C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K}), and hence does not contribute. In figures we illustrate this equality as follows

evι[[Uncaptioned image]]=ev[[Uncaptioned image]].\operatorname{ev}_{\ast}\iota_{\ast}\left[\raisebox{-0.4pt}{\includegraphics{equality_small_1.pdf}}\right]=\operatorname{ev}_{\ast}\left[\raisebox{-0.4pt}{\includegraphics{equality_small_2.pdf}}\right]\,.

The word (𝒂𝒂~)1μs(𝒂~)(𝒂𝒂~)2(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{1}\mu^{s}(\tilde{\boldsymbol{a}})(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2} is the word obtained from 𝒂\boldsymbol{a}, by replacing the word 𝒂~\tilde{\boldsymbol{a}} with μs(𝒂~)\mu^{s}(\tilde{\boldsymbol{a}}). Therefore

(3.21) ev(ι([¯(𝒂𝒂~)]×[cw(𝒂~)]))=Ψr+1+t(amat+s+1μs(𝒂~)ata1),\operatorname{ev}_{\ast}\left(\iota_{\ast}\left([\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})]\times[\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})]\right)\right)=\varPsi_{r+1+t}\left(a_{m}\otimes\cdots\otimes a_{t+s+1}\otimes\mu^{s}(\tilde{\boldsymbol{a}})\otimes a_{t}\otimes\cdots\otimes a_{1}\right)\,,

where

{t=length of the word (𝒂𝒂~)1s=length of the word 𝒂~r=length of the word (𝒂𝒂~)2.\displaystyle\begin{cases}t=\text{length of the word }(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{1}\\ s=\text{length of the word }\boldsymbol{\tilde{a}}\\ r=\text{length of the word }(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2}\,.\end{cases}

This means that the broken disks of type (1) correspond to terms of the form Ψr+1+t(idrμsidt)\varPsi_{r+1+t}(\operatorname{id}^{\otimes r}\otimes\mu^{s}\otimes\operatorname{id}^{\otimes t}) where r+s+t=mr+s+t=m.

Refer to caption
Figure 10. All broken disks of type (1) in the case of m=3m=3.

Similarly, for the first terms in (3.20) which correspond to broken disks of type (2) we apply 3.9 to get

(3.22) ev(ι([¯(𝒂)]×[¯(𝒂′′)]))=P(ev[¯(𝒂′′)]ev[¯(𝒂)])=P(Ψm2(𝒂′′)Ψm1(𝒂)),\operatorname{ev}_{\ast}\left(\iota_{\ast}\left([\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})]\times[\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})]\right)\right)=P\left(\operatorname{ev}_{\ast}[\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})]\otimes\operatorname{ev}_{\ast}[\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})]\right)=P(\varPsi_{m_{2}}(\boldsymbol{a}^{\prime\prime})\otimes\varPsi_{m_{1}}(\boldsymbol{a}^{\prime}))\,,

so that the broken disks of type (2) correspond to terms of the form P(Ψm2Ψm1)P(\varPsi_{m_{2}}\otimes\varPsi_{m_{1}}) where m1+m2=mm_{1}+m_{2}=m.

Refer to caption
Figure 11. All broken disks of type (2) in the case of m=3m=3.

Therefore via (3.21) and (3.22), equation (3.20) becomes

Ψm+m1+m2=mP(Ψm2Ψm1)=r+s+t=mΨr+1+t(idrμsidt),\partial\varPsi_{m}+\sum_{m_{1}+m_{2}=m}P(\varPsi_{m_{2}}\otimes\varPsi_{m_{1}})=\sum_{r+s+t=m}\varPsi_{r+1+t}(\operatorname{id}^{\otimes r}\otimes\mu^{s}\otimes\operatorname{id}^{\otimes t})\,,

and these are precisely the AA_{\infty}-relations modulo 2. For confirmation of signs we refer the reader to Appendix B. ∎

4. The chain map is an isomorphism

This section is dedicated to the proof of 1.1.

Theorem 4.1 (1.1).

There exists a geometrically defined isomorphism of AA_{\infty}-algebras Ψ:CWΛK(F,F)Ccell(BMK)\varPsi\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow C^{\text{cell}}_{-\ast}(BM_{K}).

The first step is to replace the full Moore loop space with a Morse theoretic model of it. It is the space of piecewise geodesic loops and we denote it by BMKBM_{K} (see Section 4.2). In the Morse theoretic model of the loop space, we have that the geodesics on MKM_{K} are precisely critical points of the energy functional, with finite dimensional unstable manifolds, and infinite dimensional stable manifolds. There is a one-to-one correspondence between Reeb chords and oriented geodesics. Assuming that the metric is generic gives moreover that Reeb chords of degree λ-\lambda are in one-to-one correspondence with geodesics of index λ\lambda (see 4.9). We will show that the evaluation map defined in Section 3.4 is transverse to the infinite dimensional stable manifolds, and that the kernel of the linearized operator DuD_{u} has the same dimension as the unstable manifold.

In Section 4.1 we will define the action filtration on CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F), followed by Section 4.2 where we first replace the full Moore loop space with the Morse theoretic model consisting of piecewise geodesic loops, and then we filter the space of loops by length. In Section 4.3 we prove that Ψ1\varPsi_{1} respects the action filtrations and in fact that Ψ1\varPsi_{1} is diagonal with respect to the action filtrations. In Section 4.5 we prove that CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) is isomorphic to the Morse theoretic model of the loop space in each filtration level, which allows us to pass to colimits.

Consider MKWKM_{K}\subset W_{K} and fix a generic Riemannian metric gg on MKM_{K} such that in the handle DεT([0,)×ΛK)D_{\varepsilon}T^{\ast}([0,\infty)\times\varLambda_{K}) of WKW_{K}, the metric has the form

dt2+f(t)g,dt^{2}+f(t)g\,,

where tt is the coordinate in the [0,)[0,\infty)-factor, and f:[0,)[0,)f\colon\thinspace[0,\infty)\longrightarrow[0,\infty) satisfies f(0)=1f^{\prime}(0)=-1, f(t)<0f^{\prime}(t)<0 and f′′(t)0f^{\prime\prime}(t)\geq 0, see (A.1) for details.

4.1. Length filtration on CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F)

For a Reeb chord generator cCWΛK(F,F)c\in CW^{\ast}_{\varLambda_{K}}(F,F) define its action by

𝔞(c) . . =0cλ.\mathfrak{a}(c)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\int_{0}^{\ell}c^{\ast}\lambda\,.

In our case with FTξSWKF\cong T^{\ast}_{\xi}S\subset W_{K} for ξMK\xi\in M_{K} we only have a single Lagrangian intersection generator ξ\xi, whose action we define explicitly as 𝔞(ξ) . . =0\mathfrak{a}(\xi)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=0. We then filter CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) by this action, and use the notation

pCWΛK(F,F) . . ={cCWΛK(F,F)|𝔞(c)<p}.\mathcal{F}_{p}CW^{\ast}_{\varLambda_{K}}(F,F)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{c\in CW^{\ast}_{\varLambda_{K}}(F,F)\;|\;\mathfrak{a}(c)<p\right\}\,.

Now, by applying Stokes’ theorem to any JJ-holomorphic disk which contributes to μ1(c)\mu^{1}(c) we get the following lemma. (Compare with e.g. [Ekh06, Lemma B.3].)

Lemma 4.2.

The differential μ1:CWΛK(F,F)CWΛK(F,F)\mu^{1}\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow CW^{\ast}_{\varLambda_{K}}(F,F) does not increase the action of generators. That is,

𝔞(c)𝔞(μ1(c)),\mathfrak{a}(c)\geq\mathfrak{a}(\mu^{1}(c))\,,

for any cCWΛK(F,F)c\in CW^{\ast}_{\varLambda_{K}}(F,F).

4.2. Length filtration on C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K})

In this section we review basic material on the Morse theory of loop spaces from [Mil63].

One goal in this section is to replace the full Moore loop space ΩξMK\varOmega_{\xi}M_{K} with a homotopy equivalent Morse theoretic model by approximating Moore loops by piecewise geodesic loops. The second goal is to in detail define the filtration on the model of chains of based loops we use.

By abuse of notation, we denote by ΩξMK\varOmega_{\xi}M_{K} the space of continuous based loops γ:[0,1]MK\gamma\colon\thinspace[0,1]\longrightarrow M_{K} with fixed domain [0,1][0,1]. It is homotopy equivalent with the space of Moore loops as defined in Section 3.2. With respect to the generic Riemannian metric hh on MKM_{K} as described in (A.1), equip ΩξMK\varOmega_{\xi}M_{K} with the supremum metric

d(γ,β) . . =supt[0,1]h(γ(t),β(t)),γ,βΩξMK.d^{\ast}(\gamma,\beta)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\sup_{t\in[0,1]}h(\gamma(t),\beta(t)),\quad\gamma,\beta\in\varOmega_{\xi}M_{K}\,.

The metric topology on ΩξMK\varOmega_{\xi}M_{K} induced by dd^{\ast} then agrees with the compact-open topology. Define ΩpwMK\varOmega^{\text{pw}}M_{K} as the space of piecewise smooth loops, and equip it with the metric

d(γ,β) . . =d(γ,β)+(01|γ˙|2|β˙|2dt)12,γ,βΩpwMK.d(\gamma,\beta)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=d^{\ast}(\gamma,\beta)+\left(\int_{0}^{1}\mathinner{\!\left\lvert\dot{\gamma}\right\rvert}^{2}-\mathinner{\!\left\lvert\dot{\beta}\right\rvert}^{2}dt\right)^{\frac{1}{2}},\quad\gamma,\beta\in\varOmega^{\text{pw}}M_{K}\,.

By [Mil63, Theorem 17.1], we have that the inclusion i:ΩpwMKΩξMKi\colon\thinspace\varOmega^{\text{pw}}M_{K}\longrightarrow\varOmega_{\xi}M_{K} is a homotopy equivalence. We define the energy of γΩpwMK\gamma\in\varOmega^{\text{pw}}M_{K} by

(4.1) E(γ) . . =01|γ˙|2dt.E(\gamma)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\int_{0}^{1}\mathinner{\!\left\lvert\dot{\gamma}\right\rvert}^{2}dt\,.

Similarly we define the length of γΩpwMK\gamma\in\varOmega^{\text{pw}}M_{K} as

L(γ) . . =01|γ˙|dt.L(\gamma)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\int_{0}^{1}\mathinner{\!\left\lvert\dot{\gamma}\right\rvert}dt\,.

Define

Ωpw,cMK . . ={γΩpwMK|E(γ)<c2}.\varOmega^{\text{pw},c}M_{K}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{\gamma\in\varOmega^{\text{pw}}M_{K}\;|\;E(\gamma)<c^{2}\right\}\,.

Fix a subdivision of [0,1][0,1],

0=t0<t1<t2<<tm=1.0=t_{0}<t_{1}<t_{2}<\cdots<t_{m}=1\,.

Then define BMKBM_{K} to be the set of loops in ΩpwMK\varOmega^{\text{pw}}M_{K} that are geodesic in the time interval [ti,ti+1][t_{i},t_{i+1}] for each i{0,,m1}i\in\left\{0,\ldots,m-1\right\}. Let

BcMK . . ={γBMK|E(γ)<c2}.B^{c}M_{K}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{\gamma\in BM_{K}\;|\;E(\gamma)<c^{2}\right\}\,.

Applying [Mil63, Lemma 16.1] then gives that for a sufficiently fine subdivision, BcMKB^{c}M_{K} is a smooth finite dimensional manifold which is a natural submanifold of (MK)m1(M_{K})^{m-1}. Moreover by [Mil63, Theorem 16.2], BcMKB^{c}M_{K} is a deformation retract of Ωpw,cMK\varOmega^{\text{pw},c}M_{K}, and critical points of E|Ωpw,cMK\mathinner{E\rvert}_{\varOmega^{\text{pw},c}M_{K}} are the same as the critical points of E|BcMK\mathinner{E\rvert}_{B^{c}M_{K}}, and E|BcMK\mathinner{E\rvert}_{B^{c}M_{K}} is furthermore a Morse function.

We consider another increasing filtration on BMKBM_{K} by filtering by length. Namely, define the length filtration of BMKBM_{K} by

cBMK={γBMK|L(γ)<c},\mathcal{F}_{c}BM_{K}=\left\{\gamma\in BM_{K}\;|\;L(\gamma)<c\right\}\,,

and correspondingly

cΩpwMK={γΩpwMK|L(γ)<c}.\mathcal{F}_{c}\varOmega^{\text{pw}}M_{K}=\left\{\gamma\in\varOmega^{\text{pw}}M_{K}\;|\;L(\gamma)<c\right\}\,.

By the same proof as [Mil63, Theorem 16.2], we construct an explicit deformation retract of cΩpwMK\mathcal{F}_{c}\varOmega^{\text{pw}}M_{K} onto cBMK\mathcal{F}_{c}BM_{K} (see 4.3 below).

If σCk(BMK)\sigma\in C_{-k}(BM_{K}) is a cubical kk-chain of piecewise geodesic loops, we define the action of σ\sigma as

𝔞(σ) . . =maxx[0,1]kL(σ(x)).\mathfrak{a}(\sigma)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\max_{x\in[0,1]^{k}}L(\sigma(x))\,.

We then define

cC(BMK) . . ={σC(BMK)|𝔞(σ)<c},\mathcal{F}_{c}C_{-{\ast}}(BM_{K})\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{\sigma\in C_{-\ast}(BM_{K})\;|\;\mathfrak{a}(\sigma)<c\right\}\,,

which gives us an increasing filtration on C(BMK)C_{-{\ast}}(BM_{K}). Futhermore, we see by definition that 𝔞(σ)𝔞(σ)\mathfrak{a}(\partial\sigma)\leq\mathfrak{a}(\sigma).

Lemma 4.3.

There is a deformation retract

r:cΩpwMKcBMK,r\colon\thinspace\mathcal{F}_{c}\varOmega^{\text{pw}}M_{K}\longrightarrow\mathcal{F}_{c}BM_{K}\,,

which therefore induces a quasi-isomorphism

r:cC(ΩpwMK)cC(BMK).r_{\ast}\colon\thinspace\mathcal{F}_{c}C_{-\ast}(\varOmega^{\text{pw}}M_{K})\longrightarrow\mathcal{F}_{c}C_{-\ast}(BM_{K})\,.
Proof.

From the proof of [Mil63, Theorem 16.2], we first define a retraction

r:cΩpwMKcBMK,r\colon\thinspace\mathcal{F}_{c}\varOmega^{\text{pw}}M_{K}\longrightarrow\mathcal{F}_{c}BM_{K}\,,

as follows. Consider the closed ball with center ξMK\xi\in M_{K} and radius cc

B(ξ,c)={xMK|h(x,ξ)c}.B(\xi,c)=\left\{x\in M_{K}\;|\;h(x,\xi)\leq c\right\}\,.

For any γcΩpwMK\gamma\in\mathcal{F}_{c}\varOmega^{\text{pw}}M_{K}, fix a fine enough subdivision of [0,1][0,1]

0=t0<t1<<tk1<1=tk,0=t_{0}<t_{1}<\cdots<t_{k-1}<1=t_{k}\,,

so that h(γ(ti1),γ(ti))<εh(\gamma(t_{i-1}),\gamma(t_{i}))<\varepsilon for some ε>0\varepsilon>0 small enough so that there is a unique geodesic between γ(ti1)\gamma(t_{i-1}) and γ(ti)\gamma(t_{i}). Because γ\gamma is contained in the ball B(ξ,c)B(\xi,c), we have by [Mil63, Corollary 10.8] that there is a unique minimal geodesic between γ(ti1)\gamma(t_{i-1}) and γ(ti)\gamma(t_{i}) of length less than ε\varepsilon. Define r(γ)r(\gamma) so that for each i{1,,k1}i\in\left\{1,\ldots,k-1\right\} we have

r(γ)|[ti1,ti]= unique minimal geodesic of length less than ε from γ(ti1) to γ(ti).\mathinner{r(\gamma)\rvert}_{[t_{i-1},t_{i}]}=\text{ unique minimal geodesic of length less than }\varepsilon\text{ from }\gamma(t_{i-1})\text{ to }\gamma(t_{i})\,.

Since geodesics are locally length minimizing, it is clear that L(γ)L(r(γ))L(\gamma)\geq L(r(\gamma)) and therefore that rr takes values in cBMK\mathcal{F}_{c}BM_{K}. For each s[0,1]s\in[0,1] we define

rs:cΩpwMKcBMK,r_{s}\colon\thinspace\mathcal{F}_{c}\varOmega^{\text{pw}}M_{K}\longrightarrow\mathcal{F}_{c}BM_{K}\,,

in such a way that for s[ti1,ti]s\in[t_{i-1},t_{i}] and any i{1,,k1}i\in\left\{1,\ldots,k-1\right\} the map rsr_{s} is so that

{rs(γ)|[0,ti1]=r(γ)|[0,ti1]rs(γ)|[ti1,s]=unique minimal geodesic from γ(ti1) to γ(s)rs(γ)|[s,1]=γ|[s,1].\begin{cases}\mathinner{r_{s}(\gamma)\rvert}_{[0,t_{i-1}]}=\mathinner{r(\gamma)\rvert}_{[0,t_{i-1}]}\\ \mathinner{r_{s}(\gamma)\rvert}_{[t_{i-1},s]}=\text{unique minimal geodesic from }\gamma(t_{i-1})\text{ to }\gamma(s)\\ \mathinner{r_{s}(\gamma)\rvert}_{[s,1]}=\mathinner{\gamma\rvert}_{[s,1]}\,.\end{cases}

Then {r0(γ)=γr1(γ)=r(γ)\begin{cases}r_{0}(\gamma)=\gamma\\ r_{1}(\gamma)=r(\gamma)\end{cases} and it is continuous in both ss and γ\gamma. Hence it shows that cBMK\mathcal{F}_{c}BM_{K} is a deformation retract of cΩpwMK\mathcal{F}_{c}\varOmega^{\text{pw}}M_{K}.

It is now straightforward to see that this map is defined on singular chains. Namely, for any fixed c>0c>0, we pick a fine enough subdivision of [0,1][0,1]

0=t0<t1<<tN1<1=tN,0=t_{0}<t_{1}<\cdots<t_{N-1}<1=t_{N}\,,

so that for every i{1,,N1}i\in\left\{1,\ldots,N-1\right\} we have

maxx[0,1]kL(σ(x)|[ti1,ti])<ε.\max_{x\in[0,1]^{k}}L\left(\mathinner{\sigma(x)\rvert}_{[t_{i-1},t_{i}]}\right)<\varepsilon\,.

Hence for any x[0,1]kx\in[0,1]^{k}, there is a unique geodesic from σ(x)(ti1)\sigma(x)(t_{i-1}) to σ(x)(ti)\sigma(x)(t_{i}). Then rr induces a map

(4.2) r:cC(ΩpwMK)\displaystyle r_{\ast}\colon\thinspace\mathcal{F}_{c}C_{-\ast}(\varOmega^{\text{pw}}M_{K}) cC(BMK)\displaystyle\longrightarrow\mathcal{F}_{c}C_{-\ast}(BM_{K})
σ\displaystyle\sigma rσ.\displaystyle\longmapsto r\circ\sigma\,.

By [Mil63, Theorem 16.3], BMKBM_{K} is a CW-complex with one cell of dimension λ\lambda for each closed geodesic on MKM_{K} of index λ\lambda. We consider the cellular chain complex Ccell(BMK)C^{\mathrm{cell}}_{-\ast}(BM_{K}). We think of the generators of Cλcell(BMK)C_{-\lambda}^{\mathrm{cell}}(BM_{K}) as the unstable manifolds of geodesics of index λ\lambda with respect to the energy functional EE on BMKBM_{K}. We define the action of a λ\lambda-cell eλe_{\lambda} as

𝔞(eλ)=maxx[0,1]λL(eλ(x)).\mathfrak{a}(e_{\lambda})=\max_{x\in[0,1]^{\lambda}}L(e_{\lambda}(x))\,.

It is well known that singular chains and cellular chains on a CW-complex are homotopy equivalent. Denote the induced isomorphism on homology by

(4.3) s:H(BMK)Hcell(BMK).s\colon\thinspace H_{-\ast}(BM_{K})\xrightarrow{\cong}H_{-\ast}^{\text{cell}}(BM_{K})\,.

In particular by 4.3 the map rr_{\ast} in (4.2) induces an isomorphism

(4.4) r:H(ΩpwMK)H(BMK).r_{\ast}\colon\thinspace H_{-\ast}(\varOmega^{\text{pw}}M_{K})\xrightarrow{\cong}H_{-\ast}(BM_{K})\,.

4.3. The chain map Ψ1\varPsi_{1} respects the filtration

The goal for this section is to prove that the chain map

Ψ1:CWΛK(F,F)C(ΩξMK),\varPsi_{1}\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow C_{-\ast}(\varOmega_{\xi}M_{K})\,,

respects the filtrations c\mathcal{F}_{c} defined on CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and C(ΩξMK)C_{-\ast}(\varOmega_{\xi}M_{K}) in Sections 4.1 and 4.2 respectively. The plan is to follow and adapt the proof of [CELN17, Proposition 8.9] to the current situation. The outline of the proof is to consider any JJ-holomorphic disk u¯(a)u\in\overline{\mathcal{M}}(a) contributing to Ψ1(a)\varPsi_{1}(a) and integrate the 2-form dλτd\lambda_{\tau} (defined in (4.7) below) over the disk. Using Stokes’ theorem we show that 0u1(WK)u𝑑λτ=𝔞(a)L(γ)0\leq\int_{u^{-1}(W_{K})}u^{\ast}d\lambda_{\tau}=\mathfrak{a}(a)-L(\gamma).

Consider a generator aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) and pick some loop γ=Ψ1(a)(x):[0,1]MK\gamma=\varPsi_{1}(a)(x)\colon\thinspace[0,1]\longrightarrow M_{K}. Then pick a tubular neighborhood N(MK)N(M_{K}) of MKM_{K} in WKW_{K} and a symplectomorphism

(4.5) φ:N(MK)DδTMK,\varphi\colon\thinspace N(M_{K})\longrightarrow D_{\delta}T^{\ast}M_{K}\,,

by the Lagrangian neighborhood theorem for some positive constant δ\delta. By a similar argument to that of the proof of [Wei71, Theorem 7.1], we may assume that φ\varphi sends the fiber FN(MK)F\cap N(M_{K}) to a fiber of DδTMKD_{\delta}T^{\ast}M_{K}.

Recall that we use the metric on MKM_{K} defined in (A.1). Pick coordinates (q,p)(q,p) in TMKT^{\ast}M_{K}, and define the canonical 1-form β=pdq\beta=pdq. Then let β1\beta_{1} be a 11-form on TMKT^{\ast}M_{K} that is given by

β1=δpdq|p|.\beta_{1}=\frac{\delta pdq}{\mathinner{\!\left\lvert p\right\rvert}}\,.

When we restrict to SδTMKS_{\delta}T^{\ast}M_{K}, the Reeb vector field R=pqR=p\partial_{q} and the contact structure ξ=kerβ1\xi=\ker\beta_{1} have the following expressions in these coordinates

R=i=1npiqi,ξ=kerβ1ker(pdp)=(span{R,pp})dβ1.R=\sum_{i=1}^{n}p_{i}\partial_{q_{i}},\quad\xi=\ker\beta_{1}\cap\ker(pdp)=\left(\mathrm{span}\left\{R,p\partial_{p}\right\}\right)^{\perp_{d\beta_{1}}}\,.

Then we have the splitting T(q,p)TMK=span{R,pp}ξT_{(q,p)}T^{\ast}M_{K}=\mathrm{span}\left\{R,p\partial_{p}\right\}\oplus\xi. We have picked an almost complex structure JJ on WKW_{K} which is compatible with dλd\lambda. The almost complex structure JJ induces an almost complex structure JJ^{\prime} on TMKT^{\ast}M_{K} defined as

J . . =(dφ)J(dφ)1,J^{\prime}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=(d\varphi)\circ J\circ(d\varphi)^{-1}\,,

which satisfies the following:

  1. (1)

    JJ^{\prime} is compatible with dpdqdp\wedge dq, and

  2. (2)

    JJ^{\prime} preserves the splitting T(q,p)TMK=span{R,pp}ξT_{(q,p)}T^{\ast}M_{K}=\mathrm{span}\left\{R,p\partial_{p}\right\}\oplus\xi.

These two conditions ensure that the map

φu:(u1(N(MK)),j)(DδTMK,J)\varphi\circ u\colon\thinspace(u^{-1}(N(M_{K})),j)\longrightarrow(D_{\delta}T^{\ast}M_{K},J^{\prime})

is JJ^{\prime}-holomorphic.

By the proof of [CELN17, Lemma 8.8] we have that dβ1(v,Jv)0d\beta_{1}(v,J^{\prime}v)\geq 0. However, if we integrate dβ1d\beta_{1} over the domain of u¯(a)u\in\overline{\mathcal{M}}(a) we can not use Stokes’ theorem directly since β1\beta_{1} is singular along the zero section, so we have to make some further modifications to get rid of this singularity.

Let

(4.6) τ:[0,)[0,1],\tau\colon\thinspace[0,\infty)\longrightarrow[0,1]\,,

be a smooth function so that

  • τ(s)=0\tau(s)=0 near s=0s=0, and

  • τ(s)0\tau^{\prime}(s)\geq 0 for every ss,

  • τ(s)=1\tau(s)=1 for sεs\geq\varepsilon for some small ε<δ\varepsilon<\delta.

Then define

βτ . . =δτ(|p|)|p|pdq.\beta_{\tau}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\frac{\delta\tau(\mathinner{\!\left\lvert p\right\rvert})}{\mathinner{\!\left\lvert p\right\rvert}}pdq\,.
Lemma 4.4 ([CELN17, Lemma 8.8]).

For any vT(q,p)TMKv\in T_{(q,p)}T^{\ast}M_{K} outside of the zero section we have

dβτ(v,Jv)0.d\beta_{\tau}(v,J^{\prime}v)\geq 0\,.

For τ(|p|)>0\tau(\mathinner{\!\left\lvert p\right\rvert})>0 and τ(|p|)>0\tau^{\prime}(\mathinner{\!\left\lvert p\right\rvert})>0 equality holds if and only if v=0v=0, whereas at points where τ(|p|)>0\tau(\mathinner{\!\left\lvert p\right\rvert})>0 and τ(|p|)=0\tau^{\prime}(\mathinner{\!\left\lvert p\right\rvert})=0 equality holds if and only if vv is a linear combination of the Liouville vector field ppp\partial_{p} and the Reeb vector field R=pqR=p\partial_{q}.

Let aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) be a generator and consider u:D3WKu\colon\thinspace D_{3}\longrightarrow W_{K} in ¯(a)\overline{\mathcal{M}}(a). Denote by γ . . =ev(u)\gamma\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\operatorname{ev}(u). Using the symplectomorphism φ\varphi in (4.5) we define an exact 2-form on WKW_{K}. Define

(4.7) dλτ . . =φdβτ,d\lambda_{\tau}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\varphi^{\ast}d\beta_{\tau}\,,

on N(MK)WKN(M_{K})\subset W_{K}. We may extend dλτd\lambda_{\tau} to the whole of WKW_{K} by defining it to be

dλτ={φdβτ,in N(MK)ω,otherwise.d\lambda_{\tau}=\begin{cases}\varphi^{\ast}d\beta_{\tau},&\text{in }N(M_{K})\\ \omega,&\text{otherwise.}\end{cases}
Lemma 4.5.

The 2-form dλτd\lambda_{\tau} on WKW_{K} defined above satisfies

dλτ(v,Jv)0.d\lambda_{\tau}(v,Jv)\geq 0\,.
Proof.

In N(MK)N(M_{K}) we have dλτ=φdβτd\lambda_{\tau}=\varphi^{\ast}d\beta_{\tau}, in which case the conclusion follows from 4.4. Otherwise we have dλτ=ωd\lambda_{\tau}=\omega which is non-negative on complex lines, because JJ is ω\omega-compatible. ∎

Lemma 4.6.

Consider the exact Lagrangian fiber FN(MK)F\cap N(M_{K}). Its image F . . =φ(FN(MK))DδTMKF^{\prime}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\varphi(F\cap N(M_{K}))\subset D_{\delta}T^{\ast}M_{K} under φ\varphi is exact with respect to β1\beta_{1}.

Proof.

This follows immediately from the assumption that φ\varphi maps FN(MK)F\cap N(M_{K}) to a fiber of DδTMKD_{\delta}T^{\ast}M_{K}, say F=DδTxMKF^{\prime}=D_{\delta}T^{\ast}_{x}M_{K} for xMKx\in M_{K}. ∎

Proposition 4.7.

Let aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) be any generator and uu be any JJ-holomorphic half strip with positive puncture at aa. Letting γ . . =ev(u)\gamma\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\operatorname{ev}(u) we have

𝔞(a)L(γ),\mathfrak{a}(a)\geq L(\gamma)\,,

with equality if and only if uu is a branched covering of a half strip over a Reeb chord.

Proof.

Since dλτ(u,Ju)0d\lambda_{\tau}(u,Ju)\geq 0 by 4.5 we integrate it over the disk u:D3WKu\colon\thinspace D_{3}\longrightarrow W_{K} and use Stokes’ theorem:

0u1(WK)u𝑑λτ\displaystyle 0\leq\int_{u^{-1}(W_{K})}u^{\ast}d\lambda_{\tau} =u1(WKN(MK))u𝑑λτ+u1(N(MK))u𝑑λτ\displaystyle=\int_{u^{-1}(W_{K}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(M_{K}))}u^{\ast}d\lambda_{\tau}+\int_{u^{-1}(N(M_{K}))}u^{\ast}d\lambda_{\tau}
(4.8) =u1(WKN(MK))uω= . . I1+(φu)1(DδTMK)(φu)𝑑βτ= . . I2.\displaystyle=\underbrace{\int_{u^{-1}(W_{K}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(M_{K}))}u^{\ast}\omega}_{=\mathrel{\vbox{\hbox{$\raisebox{0.3014pt}{\scriptsize.}$}\hbox{\scriptsize.}}}I_{1}}+\underbrace{\int_{(\varphi\circ u)^{-1}(D_{\delta}T^{\ast}M_{K})}(\varphi\circ u)^{\ast}d\beta_{\tau}}_{=\mathrel{\vbox{\hbox{$\raisebox{0.3014pt}{\scriptsize.}$}\hbox{\scriptsize.}}}I_{2}}\,.

For the remainder of this proof we follow the proof of [CELN17, Proposition 8.9]. We start by computing I2I_{2}. To do this we consider β1=δpdq|p|\beta_{1}=\frac{\delta pdq}{\mathinner{\!\left\lvert p\right\rvert}}. Then pick a biholomorphism

ψ:[0,δ0]×[0,1]UD3,\psi\colon\thinspace[0,\delta_{0}]\times[0,1]\longrightarrow U\subset D_{3}\,,

where UD3U\subset D_{3} is a neighborhood of the boundary arc between the boundary punctures ζ±\zeta_{\pm} both of which are mapped to ξMKWK\xi\in M_{K}\subset W_{K}, so that ψ(0,t)\psi(0,t) is a parametrization of the boundary arc between ζ\zeta_{-} and ζ+\zeta_{+}. We choose δ0\delta_{0} small enough so that (φuψ)(δ0,t)(\varphi\circ u\circ\psi)(\delta_{0},t) does not hit MKTMKM_{K}\subset T^{\ast}M_{K}. Let

q(t) . . =φuψ(0,t).q(t)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\varphi\circ u\circ\psi(0,t)\,.

Since we have a non-flat metric hh on MKM_{K} (see (A.1)) we consider the splitting T(TMK)VHT(T^{\ast}M_{K})\cong V\oplus H and geodesic normal coordinates (q,p)(q,p) on TMKT^{\ast}M_{K}. The almost complex structure JJ then takes the vertical subspace to the horizontal and vice versa. Consider the Levi-Civita connection on T(TMK)T(T^{\ast}M_{K}), and denote its associated Christoffel symbols by Γijk\varGamma^{k}_{ij}. Recall that in geodesic normal coordinates, the metric tensor at (q,p)(q,p) has components hij(q,p)=δijh_{ij}(q,p)=\delta_{ij}, where δij\delta_{ij} is the Kronecker delta. In particular the Christoffel symbols vanish at (q,p)(q,p). For any xx in a neighborhood of (q,p)(q,p) it follows that Γijk(x)=O(|x|)\varGamma^{k}_{ij}(x)=O(\mathinner{\!\left\lvert x\right\rvert}).

The almost complex structure in a neighborhood of (q,p)(q,p) is

{J(pi)=qiΓijkpjpkJ(qi)=pi+ΓijkpjqkΓijmΓmnkpjpnpk.\begin{cases}J(\partial_{p_{i}})=\partial_{q_{i}}-\varGamma^{k}_{ij}p^{j}\partial_{p_{k}}\\ J(\partial_{q_{i}})=-\partial_{p_{i}}+\varGamma^{k}_{ij}p^{j}\partial_{q_{k}}-\varGamma^{m}_{ij}\varGamma^{k}_{mn}p^{j}p^{n}\partial_{p_{k}}\,.\end{cases}

Since uu is JJ-holomorphic, we write

u~(s,t) . . =(φuψ)(s,t)=(Q(s,t),P(s,t)),\tilde{u}(s,t)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=(\varphi\circ u\circ\psi)(s,t)=(Q(s,t),P(s,t))\,,

where

{sQkΓijktQiPj+tPk=0sPktQk+ΓijktPiPjΓijmΓmnktQitQiPjPn=0.\begin{cases}\partial_{s}Q^{k}-\varGamma^{k}_{ij}\partial_{t}Q^{i}P^{j}+\partial_{t}P^{k}=0\\ \partial_{s}P^{k}-\partial_{t}Q^{k}+\varGamma^{k}_{ij}\partial_{t}P^{i}P^{j}-\varGamma^{m}_{ij}\varGamma^{k}_{mn}\partial_{t}Q^{i}\partial_{t}Q^{i}P^{j}P^{n}=0\,.\end{cases}

Recall that in our geodesic normal coordinates we have Γijk(x)=O(|x|)\varGamma^{k}_{ij}(x)=O(\mathinner{\!\left\lvert x\right\rvert}) where xx is in a neighborhood of (q,p)(q,p) , and hence with x=u~(s,t)x=\tilde{u}(s,t) we have

(4.9) {sQ+tP+O(|x|)=0sPtQ+O(|x|)=0.\begin{cases}\partial_{s}Q+\partial_{t}P+O(\mathinner{\!\left\lvert x\right\rvert})=0\\ \partial_{s}P-\partial_{t}Q+O(\mathinner{\!\left\lvert x\right\rvert})=0\,.\end{cases}

If we write Q(s,t)=q(t)+v(s,t)Q(s,t)=q(t)+v(s,t) we get from the the second equation in (4.9) that

P(s,t)=s(q˙(t)+O(|x|))+w(s,t),P(s,t)=s\left(\dot{q}(t)+O(\mathinner{\!\left\lvert x\right\rvert})\right)+w(s,t)\,,

where w(s,t) . . =0stv(σ,t)dσw(s,t)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\int_{0}^{s}\partial_{t}v(\sigma,t)d\sigma. We now have v(0,t)=0=w(0,t)v(0,t)=0=w(0,t) and hence vt(0,t)=0=wt(0,t)\dfrac{\partial{}v}{\partial{t}}(0,t)=0=\dfrac{\partial{}w}{\partial{t}}(0,t). Setting s=0s=0 in (4.9) gives vs(0,t)=O(|u~(0,t)|)=ws(0,t)\dfrac{\partial{}v}{\partial{s}}(0,t)=O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})=\dfrac{\partial{}w}{\partial{s}}(0,t). Next, from Taylor’s formula we get vt(δ0,t)=O(δ0)\dfrac{\partial{}v}{\partial{t}}(\delta_{0},t)=O(\delta_{0}) and w(δ0,t)=δ0O(|u~(0,t)|)+O(δ02)w(\delta_{0},t)=\delta_{0}O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\delta_{0}^{2}). Then we get

(4.10) u~β1|s=δ0\displaystyle\mathinner{\tilde{u}^{\ast}\beta_{1}\rvert}_{s=\delta_{0}} =(δ0q˙(t)+w(δ0,t))(q˙(t)+vt(δ0,t))|δ0q˙(t)+w(δ0,t)|dt\displaystyle=\frac{(\delta_{0}\dot{q}(t)+w(\delta_{0},t))(\dot{q}(t)+\dfrac{\partial{}v}{\partial{t}}(\delta_{0},t))}{\mathinner{\!\left\lvert\delta_{0}\dot{q}(t)+w(\delta_{0},t)\right\rvert}}dt
=δ0q˙(t)+δ0O(|u~(0,t)|)+O(δ02),q˙(t)+O(δ0)|δ0q˙(t)+δ0O(|u~(0,t)|)+O(δ02)|dt\displaystyle=\frac{\left\langle\delta_{0}\dot{q}(t)+\delta_{0}O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\delta_{0}^{2}),\dot{q}(t)+O(\delta_{0})\right\rangle}{\mathinner{\!\left\lvert\delta_{0}\dot{q}(t)+\delta_{0}O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\delta_{0}^{2})\right\rvert}}dt
=q˙(t)+O(|u~(0,t)|)+O(δ0),q˙(t)+O(δ0)|q˙(t)+O(|u~(0,t)|)+O(δ0)|dt=(|q˙(t)|+O(δ0))dt\displaystyle=\frac{\left\langle\dot{q}(t)+O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\delta_{0}),\dot{q}(t)+O(\delta_{0})\right\rangle}{\mathinner{\!\left\lvert\dot{q}(t)+O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\delta_{0})\right\rvert}}dt=\left(\mathinner{\!\left\lvert\dot{q}(t)\right\rvert}+O(\delta_{0})\right)dt

Next, pick ε>0\varepsilon>0 so that it is smaller than the minimal norm of the pp-components of (φuψ)(δ0,t)(\varphi\circ u\circ\psi)(\delta_{0},t) and pick a function τ:[0,)[0,1]\tau\colon\thinspace[0,\infty)\longrightarrow[0,1] as in (4.6). Namely, τ\tau satisfies

{τ(s)0,s[0,)τ(s)=0,near s=0τ(s)=1,sε.\begin{cases}\tau^{\prime}(s)\geq 0,&\forall s\in[0,\infty)\\ \tau(s)=0,&\text{near }s=0\\ \tau(s)=1,&s\geq\varepsilon\,.\end{cases}

Consider βτ=δτ(|p|)pdq|p|\beta_{\tau}=\frac{\delta\tau(\mathinner{\!\left\lvert p\right\rvert})pdq}{\mathinner{\!\left\lvert p\right\rvert}}. By 4.4 we have (φu)dβτ0(\varphi\circ u)^{\ast}d\beta_{\tau}\geq 0, and also that βτ\beta_{\tau} agrees with β1\beta_{1} in the set {|p|ε}TMK\left\{\mathinner{\!\left\lvert p\right\rvert}\geq\varepsilon\right\}\subset T^{\ast}M_{K}. Then we get

limδ00{δ0}×[0,1](φuψ)βτ=limδ00{δ0}×[0,1]|q˙(t)|+O(δ0)dt=limδ00L(γ)+O(δ0)=L(γ).\lim_{\delta_{0}\to 0}\int_{\left\{\delta_{0}\right\}\times[0,1]}(\varphi\circ u\circ\psi)^{\ast}\beta_{\tau}=\lim_{\delta_{0}\to 0}\int_{\left\{\delta_{0}\right\}\times[0,1]}\mathinner{\!\left\lvert\dot{q}(t)\right\rvert}+O(\delta_{0})dt=\lim_{\delta_{0}\to 0}L(\gamma)+O(\delta_{0})=L(\gamma)\,.
Refer to caption
Figure 12. Domain of the JJ-holomorphic disk uu with neighborhoods around the outgoing segment between ζ\zeta_{-} and ζ+\zeta_{+} and the positive puncture ζ1\zeta_{1} marked in white.

By 4.6 we have that F=φ(FN(MK))DδTMKF^{\prime}=\varphi(F\cap N(M_{K}))\subset D_{\delta}T^{\ast}M_{K} is exact with respect to βτ\beta_{\tau}. Therefore we get

I2\displaystyle I_{2} =(φu)1(DδTMK)(φu)𝑑βτ\displaystyle=\int_{(\varphi\circ u)^{-1}(D_{\delta}T^{\ast}M_{K})}(\varphi\circ u)^{\ast}d\beta_{\tau}
=(φu)1(SδTMK)(φu)βτlimδ00{δ0}×[0,1](φuψ)βτ\displaystyle=\int_{(\varphi\circ u)^{-1}(S_{\delta}T^{\ast}M_{K})}(\varphi\circ u)^{\ast}\beta_{\tau}-\lim_{\delta_{0}\to 0}\int_{\left\{\delta_{0}\right\}\times[0,1]}(\varphi\circ u\circ\psi)^{\ast}\beta_{\tau}
(4.11) =(φu)1(SδTMK)(φu)βτL(γ).\displaystyle=\int_{(\varphi\circ u)^{-1}(S_{\delta}T^{\ast}M_{K})}(\varphi\circ u)^{\ast}\beta_{\tau}-L(\gamma)\,.

Finally, the integral I1I_{1} in (4.3) is computed by using Stokes’ theorem and that dλτ=ωd\lambda_{\tau}=\omega outside of N(MK)N(M_{K}) by definition.

(4.12) I1\displaystyle I_{1} =u1(WKN(MK))uω=𝔞(a)u1(N(MK))uλ\displaystyle=\int_{u^{-1}(W_{K}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(M_{K}))}u^{\ast}\omega=\mathfrak{a}(a)-\int_{u^{-1}(\partial N(M_{K}))}u^{\ast}\lambda

By combining (4.12) with (4.3) we get

(4.13) 0u1(WK)u𝑑λτ=u1(N(MK))uλ+(φu)1(SδTMK)(φu)βτ+𝔞(a)L(γ).0\leq\int_{u^{-1}(W_{K})}u^{\ast}d\lambda_{\tau}=-\int_{u^{-1}(\partial N(M_{K}))}u^{\ast}\lambda+\int_{(\varphi\circ u)^{-1}(S_{\delta}T^{\ast}M_{K})}(\varphi\circ u)^{\ast}\beta_{\tau}+\mathfrak{a}(a)-L(\gamma)\,.

Note that along SδTMKS_{\delta}T^{\ast}M_{K} we have βτ=β=pdq\beta_{\tau}=\beta=pdq. Furthermore φ\varphi is an exact symplectomorphism so we have φβλ=dθ\varphi^{\ast}\beta-\lambda=d\theta. Hence

u1(N(MK))u(φβλ)=u1(N(MK))u𝑑θ=0,\int_{u^{-1}(\partial N(M_{K}))}u^{\ast}(\varphi^{\ast}\beta-\lambda)=\int_{u^{-1}(\partial N(M_{K}))}u^{\ast}d\theta=0\,,

and therefore (4.13) turns into

0u1(WK)u𝑑λτ=𝔞(a)L(γ)𝔞(a)L(γ).0\leq\int_{u^{-1}(W_{K})}u^{\ast}d\lambda_{\tau}=\mathfrak{a}(a)-L(\gamma)\Leftrightarrow\mathfrak{a}(a)\geq L(\gamma)\,.

Corollary 4.8.

Let aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) be any generator and let u¯(a)u\in\overline{\mathcal{M}}(a). Then

𝔞(a)𝔞(Ψ1a)𝔞((rΨ1)(a)).\mathfrak{a}(a)\geq\mathfrak{a}(\varPsi_{1}a)\geq\mathfrak{a}((r_{\ast}\circ\varPsi_{1})(a))\,.
Proof.

Fix a generator aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) and consider the moduli space ¯(a)\overline{\mathcal{M}}(a). The action of Ψ1(a)C(ΩξMK)\varPsi_{1}(a)\in C_{-\ast}(\varOmega_{\xi}M_{K}) is

𝔞(Ψ1(a))=maxx[0,1]L(Ψ1(a)(x)).\mathfrak{a}(\varPsi_{1}(a))=\max_{x\in[0,1]^{\ast}}L(\varPsi_{1}(a)(x))\,.

Note that the maximum is well-defined by the compactness of [0,1][0,1]^{\ast}. Let xmax[0,1]x_{\mathrm{max}}\in[0,1]^{\ast} be such that L(Ψ1(a)(xmax))=L(Ψ1(a))L(\varPsi_{1}(a)(x_{\mathrm{max}}))=L(\varPsi_{1}(a)), and let γmax . . =Ψ1(a)(xmax)\gamma_{\mathrm{max}}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\varPsi_{1}(a)(x_{\mathrm{max}}). Since 4.7 holds for any γΩξMK\gamma\in\varOmega_{\xi}M_{K}, in particular it holds for γmax\gamma_{\mathrm{max}}. Therefore

𝔞(a)L(γmax)=maxx[0,1]L(Ψ1(a)(x))=𝔞(Ψ1a).\mathfrak{a}(a)\geq L(\gamma_{\mathrm{max}})=\max_{x\in[0,1]^{\ast}}L(\varPsi_{1}(a)(x))=\mathfrak{a}(\varPsi_{1}a)\,.

Moreover, the inequality 𝔞(Ψ1a)𝔞((rΨ1)(a))\mathfrak{a}(\varPsi_{1}a)\geq\mathfrak{a}((r_{\ast}\circ\varPsi_{1})(a)) holds because rr_{\ast} does not increase filtration (see the proof of 4.3). ∎

4.4. The chain map Ψ1\varPsi_{1} is diagonal with respect to the action filtrations

In this section we prove that Ψ1\varPsi_{1} is diagonal with respect to the action filtrations, that is 𝔞(a)=𝔞(Ψ1a)\mathfrak{a}(a)=\mathfrak{a}(\varPsi_{1}a).

We first give a brief outline of the proof. We consider the trivial JJ-holomorphic half strip u0u_{0} whose image is the cone over the Reeb chord aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) and whose tangent space at (q,p)(q,p) is spanned by pqp\partial_{q} and ppp\partial_{p} in geodesic normal coordinates. We show that u0u_{0} is transversely cut out, and therefore by 4.4 together with the proof of 4.7 we get

0=u01(WK)u0𝑑λτ=𝔞(a)L(γ0).0=\int_{u_{0}^{-1}(W_{K})}u_{0}^{\ast}d\lambda_{\tau}=\mathfrak{a}(a)-L(\gamma_{0})\,.

To prove that u0u_{0} is transversely cut out, we choose a generic Riemannian metric gg on MKM_{K} (see (A.1)). Then we have a one-to-one correspondence between Reeb chords of degree λ-\lambda and geodesics of index λ\lambda (see 4.9 below). We consider vector fields vkerDu0v\in\ker D_{u_{0}} in the kernel of the linearized Cauchy–Riemann operator at u0u_{0}. Then we show that vv restricts to broken Jacobi fields along γ\gamma for which the Hessian of the energy functional (4.1) is negative definite.

The following lemma is essentially found and proven in [RS95, Prop 6.38] and [Dui76]. Recall that the degree of Reeb chords is defined via the Conley–Zehnder index, see 2.3 for details.

Lemma 4.9.

Let a<ba<b be two real numbers. There is a one-to-one correspondence between Reeb chords aa of degree λ-\lambda with action 𝔞(a)=A\mathfrak{a}(a)=A and geodesics γ\gamma in BMKBM_{K} of index λ\lambda with length L(γ)=AL(\gamma)=A.

Proof.

It is a consequence of the first part of 4.11 and in particular (4.18) (which do not depend on the current lemma) that Reeb chords with action AA are in one-to-one correspondence with geodesics in BMKBM_{K} with length AA. What is left to show is that this one-to-one correspondence also preserves degree/index.

Let γBMK\gamma\in BM_{K} be a (non-broken) geodesic. By Morse theory on the loop space, the index of γ\gamma is defined as the Morse index of the energy functional (4.1). Morse’s index theorem [Mil63, Theorem 15.1] says that the index of γ\gamma is equal to the number of points γ(t)\gamma(t) for t(0,1)t\in(0,1), which is conjugate to γ(0)\gamma(0) along γ\gamma, counted with multiplicity. Recall that γ(t)\gamma(t) is conjugate to γ(0)\gamma(0) along γ\gamma by definition, if there is a Jacobi field KK along γ\gamma so that K(t)=K(0)=0K(t)=K(0)=0. A Jacobi field is a vector field along γ\gamma satisfying the Jacobi equation

ddtddtK+R(γ˙,K)γ˙=0,\nabla_{\frac{d}{dt}}\nabla_{\frac{d}{dt}}K+R(\dot{\gamma},K)\dot{\gamma}=0\,,

where \nabla is the Levi-Civita connection on the bundle γTMK\gamma^{\ast}TM_{K}, and RR is the corresponding curvature tensor. The geodesic flow on MKM_{K} lifts to the Reeb flow on STMKST^{\ast}M_{K}. Therefore Jacobi fields — which are seen as linearizations of the geodesic flow — lift to the linearized Reeb flow.

Assume that t1(0,1)t_{1}\in(0,1) so that γ(t1)\gamma(t_{1}) is a point that is conjugate to γ(0)\gamma(0). Then let {ei(t)}i=2n\left\{e_{i}(t)\right\}_{i=2}^{n} be a parallel orthonormal frame of γ˙(t)Tγ(t)MK\dot{\gamma}(t)^{\perp}\subset T_{\gamma(t)}M_{K}, and let K(t)=i=2nKi(t)ei(t)K(t)=\sum_{i=2}^{n}K_{i}(t)e_{i}(t) be a Jacobi field so that K(t1)=K(0)=0K(t_{1})=K(0)=0. Defining e1(t) . . =γ˙(t)e_{1}(t)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\dot{\gamma}(t) we thus have that {ei(t)}i=1n\left\{e_{i}(t)\right\}_{i=1}^{n} is a parallel orthonormal frame of Tγ(t)MKT_{\gamma(t)}M_{K} along γ\gamma. Using the notation K˙ . . =ddtK\dot{K}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\nabla_{\frac{d}{dt}}K, we define L . . =K˙L\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\dot{K}. Then we have the system

{K˙=LL˙=R(γ˙,K)γ˙.\begin{cases}\dot{K}=L\\ \dot{L}=-R(\dot{\gamma},K)\dot{\gamma}\,.\end{cases}

By using K=i=1nKieiK=\sum_{i=1}^{n}K_{i}e_{i} and L=i=1nLieiL=\sum_{i=1}^{n}L_{i}e_{i} with K1=L1=0K_{1}=L_{1}=0, we get the system of differential equations

{K˙i(t)=Li(t)L˙i(t)=j=1nRji(t)Kj(t) for all i{1,,n}\begin{cases}\dot{K}_{i}(t)=L_{i}(t)\\ \dot{L}_{i}(t)=-\sum_{j=1}^{n}R^{i}_{j}(t)K_{j}(t)\end{cases}\text{ for all }i\in\left\{1,\ldots,n\right\}

where R(t)={Rji(t)}i,j=1n={R(γ˙,ej)γ˙,ei}i,j=1nR(t)=\left\{R^{i}_{j}(t)\right\}_{i,j=1}^{n}=\left\{\langle R(\dot{\gamma},e_{j})\dot{\gamma},e_{i}\rangle\right\}_{i,j=1}^{n} is a symmetric matrix. This is equivalent to say that

ddt(KL)=(0InR0)(KL),\frac{d}{dt}\begin{pmatrix}K\\ L\end{pmatrix}=\begin{pmatrix}0&I_{n}\\ -R&0\end{pmatrix}\begin{pmatrix}K\\ L\end{pmatrix}\,,

An explicit fundamental solution to this system is

Φ(t)=exp(t(0InR0))=(C(t2R)tS(t2R)(tR)S(t2R)C(t2R)),\varPhi(t)=\exp\left(t\begin{pmatrix}0&I_{n}\\ -R&0\end{pmatrix}\right)=\begin{pmatrix}C(-t^{2}R)&tS(-t^{2}R)\\ (-tR)S(-t^{2}R)&C(-t^{2}R)\end{pmatrix}\,,

where

C(A)=k=0Ak(2k)!,S(A)=k=0Ak(2k+1)!,C(A)=\sum_{k=0}^{\infty}\frac{A^{k}}{(2k)!},\qquad S(A)=\sum_{k=0}^{\infty}\frac{A^{k}}{(2k+1)!}\,,

for AEnd(n)A\in\operatorname{End}(\mathbb{R}^{n}). The fundamental solution Φ\varPhi satisfies Φ(0)=I\varPhi(0)=I and in particular

(K(t)L(t))=Φ(t)(K(0)L(0)).\begin{pmatrix}K(t)\\ L(t)\end{pmatrix}=\varPhi(t)\begin{pmatrix}K(0)\\ L(0)\end{pmatrix}\,.

We use that the Jacobi field KK vanishes at t=0t=0. Then we plug in t=t1(0,1)t=t_{1}\in(0,1) for which we also have K(t1)=0K(t_{1})=0. Thus

(0L(t1))=Φ(t)(0L(0))=(t1S(t12R)L(0)C(t12R)L(0)).\begin{pmatrix}0\\ L(t_{1})\end{pmatrix}=\varPhi(t)\begin{pmatrix}0\\ L(0)\end{pmatrix}=\begin{pmatrix}t_{1}S(-t_{1}^{2}R)L(0)\\ C(-t_{1}^{2}R)L(0)\end{pmatrix}\,.

From this, we have that γ(t1)\gamma(t_{1}) is conjugate to γ(0)\gamma(0) if and only if S(t2R)S(-t^{2}R) is singular at t=t1t=t_{1}.

We consider (K,L)T(Tγ(t)MK)(K,L)\in T(T_{\gamma(t)}M_{K}), and using the metric isomorphism and scaling KK properly, we consider (K,L)T(STγ(t)MK)(K,L)\in T(ST^{\ast}_{\gamma(t)}M_{K}) for γ(t)MK\gamma(t)\in M_{K}. Since we assumed that KK (and hence LL) was orthogonal to γ\gamma, we regard the lift as (K,L)ξT(STMK)(K,L)\in\xi\subset T(ST^{\ast}M_{K}) along the lifted geodesic. Since ξzn1n1in1\xi_{z}\cong\mathbb{C}^{n-1}\cong\mathbb{R}^{n-1}\oplus i\mathbb{R}^{n-1} is symplectic with the standard symplectic form in these coordinates, we have that in1ξi\mathbb{R}^{n-1}\subset\xi is Lagrangian. We then consider the path of Lagrangians

(t)=Φ(t)(in1)=Φ(t)(0ζ)=(tS(t2R)ζC(t2R)ζ),ζin1.\ell(t)=\varPhi(t)(i\mathbb{R}^{n-1})=\varPhi(t)\begin{pmatrix}0\\ \zeta\end{pmatrix}=\begin{pmatrix}tS(-t^{2}R)\zeta\\ C(-t^{2}R)\zeta\end{pmatrix},\quad\forall\zeta\in i\mathbb{R}^{n-1}\,.

Whenever γ(t)\gamma(t) is conjugate to γ(0)\gamma(0), the matrix S(t2R)S(-t^{2}R) is singular. Hence it has non-trivial kernel which contributes to the Maslov index exactly the dimension of the kernel. The dimension of the kernel also correspond to the multiplicity of γ(t)\gamma(t) as a conjugate point to γ(0)\gamma(0). By closing up the loop positively, we find an extra contribution of n1n-1. Hence

μ()=(n1)+t:S(t2R) singulardimkerS(t2R)=(n1)+ind(γ),\mu(\ell)=(n-1)+\sum_{t\;\mathrel{\mathop{\ordinarycolon}}\;S(-t^{2}R)\text{ singular}}\dim\ker S(-t^{2}R)=(n-1)+\operatorname{ind}(\gamma)\,,

from which we conclude ind(γ)=|a|\operatorname{ind}(\gamma)=-\mathinner{\!\left\lvert a\right\rvert}. ∎

Proposition 4.10.

Let aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) be any generator. Let u0u_{0} be a JJ-holomorphic half strip as in Section 3.3, and let vkerDu0v\in\ker D_{u_{0}}, where Du0D_{u_{0}} is the linearized Cauchy–Riemann operator at u0u_{0}. Then consider the linearized solution

uε . . =expu0(εv),u_{\varepsilon}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\exp_{u_{0}}(\varepsilon v)\,,

for any ε>0\varepsilon>0. Then we have

𝔞(a)>L(γε),\mathfrak{a}(a)>L(\gamma_{\varepsilon})\,,

where γε . . =ev(uε)\gamma_{\varepsilon}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\operatorname{ev}(u_{\varepsilon}).

Proof.

We modify the proof of 4.7. Since uεu_{\varepsilon} is not JJ-holomorphic, we first prove that the estimate 0<uε1(WK)uε𝑑λ10<\int_{u_{\varepsilon}^{-1}(W_{K})}u_{\varepsilon}^{\ast}d\lambda_{1} holds.

In a neighborhood of (q,p)imuεWK(q,p)\in\operatorname{im}u_{\varepsilon}\subset W_{K}, we have a splitting

T(q,p)WK=span{pp,pq}ξ2×2n2.T_{(q,p)}W_{K}=\mathrm{span}\left\{p\partial_{p},p\partial_{q}\right\}\oplus\xi\cong\mathbb{R}^{2}\times\mathbb{R}^{2n-2}\,.

We pick a small ball B(ρ)B(\rho) of radius ρ>0\rho>0 around (0,0)2×2n2(0,0)\in\mathbb{R}^{2}\times\mathbb{R}^{2n-2}. Let J0J_{0} be the product almost complex structure on 2×2n2\mathbb{R}^{2}\times\mathbb{R}^{2n-2}, which extends JJ over B(ρ)B(\rho), and let ||\mathinner{\!\left\lvert-\right\rvert} be the norm determined by dλ1d\lambda_{1} and J0J_{0}. Then we pick some coordinate z=(s,t)Tz=(s,t)\in T in the domain of uεu_{\varepsilon}, and define

q(z) . . =(J(uε(z))+J0(uε(z)))1(J(uε(z))J0(uε(z))).q(z)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=(J(u_{\varepsilon}(z))+J_{0}(u_{\varepsilon}(z)))^{-1}(J(u_{\varepsilon}(z))-J_{0}(u_{\varepsilon}(z)))\,.

In the operator norm we have q(z)=O(ρ)\mathinner{\!\left\lVert q(z)\right\rVert}=O(\rho) as ρ0\rho\to 0. Then we have

(4.14) 2(J0+J)1J¯Juε=¯J0uε+q(z)J0uε= . . A(z).2(J_{0}+J)^{-1}J\overline{\partial}_{J}u_{\varepsilon}=\overline{\partial}_{J_{0}}u_{\varepsilon}+q(z)\partial_{J_{0}}u_{\varepsilon}=\mathrel{\vbox{\hbox{$\raisebox{0.43057pt}{\scriptsize.}$}\hbox{\scriptsize.}}}A(z)\,.

Let ρ\rho be small enough so that |J0uε|23|q(z)J0uε|2>0\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}-3\mathinner{\!\left\lvert q(z)\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}>0. Then we have

|J0uε|2+|¯J0uε|2\displaystyle\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}+\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2} <|J0uε|2+|¯J0uε|2+(|J0uε|23|q(z)J0uε|2)+|A(z)|2\displaystyle<\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}+\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2}+\left(\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}-3\mathinner{\!\left\lvert q(z)\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}\right)+\mathinner{\!\left\lvert A(z)\right\rvert}^{2}
=2|J0uε|2+|¯J0uε|23(|A(z)|2+|q(z)J0uε|2)+4|A(z)|2\displaystyle=2\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}+\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2}-3\left(\mathinner{\!\left\lvert A(z)\right\rvert}^{2}+\mathinner{\!\left\lvert q(z)\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}\right)+4\mathinner{\!\left\lvert A(z)\right\rvert}^{2}
2(|J0uε|2|¯J0uε|2)+4|A(z)|2.\displaystyle\leq 2\left(\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}-\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2}\right)+4\mathinner{\!\left\lvert A(z)\right\rvert}^{2}\,.

Next we note that

uεdλ1=(|duε|22|¯J0uε|2)dvolT=(|J0uε|2|¯J0uε|2)dvolT,u_{\varepsilon}^{\ast}d\lambda_{1}=\left(\mathinner{\!\left\lvert du_{\varepsilon}\right\rvert}^{2}-2\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2}\right)\mathrm{dvol}_{T}=\left(\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}-\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2}\right)\mathrm{dvol}_{T}\,,

and hence

|duε|2dvolT\displaystyle\mathinner{\!\left\lvert du_{\varepsilon}\right\rvert}^{2}\mathrm{dvol}_{T} =(|J0uε|2+|¯J0uε|2)dvolT<2(|J0uε|2|¯J0uε|2)dvolT+4|A(z)|2dvolT\displaystyle=\left(\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}+\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2}\right)\mathrm{dvol}_{T}<2\left(\mathinner{\!\left\lvert\partial_{J_{0}}u_{\varepsilon}\right\rvert}^{2}-\mathinner{\!\left\lvert\overline{\partial}_{J_{0}}u_{\varepsilon}\right\rvert}^{2}\right)\mathrm{dvol}_{T}+4\mathinner{\!\left\lvert A(z)\right\rvert}^{2}\mathrm{dvol}_{T}
=2uεdλ1+4|A(z)|2dvolT.\displaystyle=2u_{\varepsilon}^{\ast}d\lambda_{1}+4\mathinner{\!\left\lvert A(z)\right\rvert}^{2}\mathrm{dvol}_{T}\,.

In view of the definition of A(z)A(z) in (4.14), we have that |A(z)|2=O(ε4)\mathinner{\!\left\lvert A(z)\right\rvert}^{2}=O(\varepsilon^{4}).

Let πξ:TWKTWK\pi_{\xi}\colon\thinspace TW_{K}\longrightarrow TW_{K} be the projection onto the contact plane ξ\xi. By 4.4 the only contribution to uε1(B(ρ))uε𝑑λ1\int_{u_{\varepsilon}^{-1}(B(\rho))}u_{\varepsilon}^{\ast}d\lambda_{1} comes from the restriction to ξ\xi. Summing over all balls B(ρ)B(\rho) covering the image of u0u_{0} gives

(4.15) 2uε1(WK)uε𝑑λ1=2uε1(WK)πξ(uεdλ1)πξ(duε)24πξ(A(z))2.2\int_{u_{\varepsilon}^{-1}(W_{K})}u_{\varepsilon}^{\ast}d\lambda_{1}=2\int_{u_{\varepsilon}^{-1}(W_{K})}\pi_{\xi}(u_{\varepsilon}^{\ast}d\lambda_{1})\geq\mathinner{\!\left\lVert\pi_{\xi}(du_{\varepsilon})\right\rVert}^{2}-4\mathinner{\!\left\lVert\pi_{\xi}(A(z))\right\rVert}^{2}\,.

The Taylor expansion of uεu_{\varepsilon} around ε=0\varepsilon=0 is

uε=u0+εv+O(ε2),u_{\varepsilon}=u_{0}+\varepsilon v+O(\varepsilon^{2})\,,

where vkerDu0v\in\ker D_{u_{0}}. Because vv is a non-zero solution of the linearized equation Du0v=0D_{u_{0}}v=0, we rescale vv in such a way that

(4.16) vWκ2,22=vLκ22+dvWκ1,22=1\mathinner{\!\left\lVert v\right\rVert}^{2}_{W^{2,2}_{\kappa}}=\mathinner{\!\left\lVert v\right\rVert}^{2}_{L^{2}_{\kappa}}+\mathinner{\!\left\lVert dv\right\rVert}^{2}_{W^{1,2}_{\kappa}}=1

where Wκk,pW^{k,p}_{\kappa} is the weighted Sobolev space Wk,p([0,)×[0,1])W^{k,p}([0,\infty)\times[0,1]) with weight eκse^{\kappa s} for some small κ>0\kappa>0 and where ss is the coordinate in the [0,)[0,\infty)-factor. Let ZT=[0,T]×[0,1][0,)×[0,1]Z_{T}=[0,T]\times[0,1]\subset[0,\infty)\times[0,1] for some T>0T>0. We use the Poincaré inequality vLκ2(ZT)2C1dvLκ2(ZT)2\mathinner{\!\left\lVert v\right\rVert}^{2}_{L^{2}_{\kappa}(Z_{T})}\leq C_{1}\mathinner{\!\left\lVert dv\right\rVert}^{2}_{L^{2}_{\kappa}(Z_{T})}, where C1>0C_{1}>0 (given that κ>0\kappa>0 is small enough), together with (4.16). This gives that dvWκ1,2(ZT)2C0\mathinner{\!\left\lVert dv\right\rVert}^{2}_{W^{1,2}_{\kappa}(Z_{T})}\geq C_{0} for some C0>0C_{0}>0 and some T>0T>0. Hence dvWκ1,22C\mathinner{\!\left\lVert dv\right\rVert}^{2}_{W^{1,2}_{\kappa}}\geq C for some C>0C>0.

The same argument applied to πξ(v)\pi_{\xi}(v) and πξ(dv)\pi_{\xi}(dv) gives πξ(dv)Wκ1,22C\mathinner{\!\left\lVert\pi_{\xi}(dv)\right\rVert}_{W^{1,2}_{\kappa}}^{2}\geq C^{\prime} for some C>0C^{\prime}>0. Hence

πξ(duε)24πξ(A(z))2=επξ(dv)2+O(ε3)Cε2+O(ε3)>0,\mathinner{\!\left\lVert\pi_{\xi}(du_{\varepsilon})\right\rVert}^{2}-4\mathinner{\!\left\lVert\pi_{\xi}(A(z))\right\rVert}^{2}=\mathinner{\!\left\lVert\varepsilon\pi_{\xi}(dv)\right\rVert}^{2}+O(\varepsilon^{3})\geq C^{\prime}\varepsilon^{2}+O(\varepsilon^{3})>0\,,

for small enough ε>0\varepsilon>0. By (4.15) we therefore have uε1(WK)uε𝑑λ1>0\int_{u_{\varepsilon}^{-1}(W_{K})}u_{\varepsilon}^{\ast}d\lambda_{1}>0.

Next, we show uε1(WK)uε𝑑λ1=𝔞(a)L(γε)\int_{u_{\varepsilon}^{-1}(W_{K})}u_{\varepsilon}^{\ast}d\lambda_{1}=\mathfrak{a}(a)-L(\gamma_{\varepsilon}). The proof is similar to the computation in the proof of 4.7. The only difference is the computation of I2I_{2} (with notation as in 4.7). Since ¯Juε=O(ε2)\overline{\partial}_{J}u_{\varepsilon}=O(\varepsilon^{2}), equation (4.9) becomes

(4.17) {sQ+tP+O(|x|)=O(ε2)sPtP+O(|x|)=O(ε2),\begin{cases}\partial_{s}Q+\partial_{t}P+O(\mathinner{\!\left\lvert x\right\rvert})=O(\varepsilon^{2})\\ \partial_{s}P-\partial_{t}P+O(\mathinner{\!\left\lvert x\right\rvert})=O(\varepsilon^{2})\,,\end{cases}

where x=u~(s,t)x=\tilde{u}(s,t). Then from the second equation we get

P(s,t)=s(q˙(t)+O(|x|)+O(ε2))+w(s,t).P(s,t)=s\left(\dot{q}(t)+O(\mathinner{\!\left\lvert x\right\rvert})+O(\varepsilon^{2})\right)+w(s,t)\,.

Setting s=0s=0 in (4.17) gives

vs(0,t)=O(|u~(0,t)|))+O(ε2)=ws(0,t),\frac{\partial v}{\partial s}(0,t)=O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert}))+O(\varepsilon^{2})=\frac{\partial w}{\partial s}(0,t)\,,

and hence

{vt(δ0,t)=O(δ0)w(δ0,t)=δ0(O(|u~(0,t)|)+O(ε2))+O(δ02).\begin{cases}\frac{\partial v}{\partial t}(\delta_{0},t)=O(\delta_{0})\\ w(\delta_{0},t)=\delta_{0}\left(O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\varepsilon^{2})\right)+O(\delta_{0}^{2})\,.\end{cases}

By repeating the same calculation as in (4.10) we end up at

u~β1|s=δ0=q˙(t)+O(|u~(0,t)|)+O(ε2)+O(δ0),q˙(t)+O(δ0)|q˙(t)+O(|u~(0,t)|)+O(ε2)+O(δ0)|dt=(|q˙(t)|+O(δ0))dt.\mathinner{\tilde{u}^{\ast}\beta_{1}\rvert}_{s=\delta_{0}}=\frac{\left\langle\dot{q}(t)+O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\varepsilon^{2})+O(\delta_{0}),\dot{q}(t)+O(\delta_{0})\right\rangle}{\mathinner{\!\left\lvert\dot{q}(t)+O(\mathinner{\!\left\lvert\tilde{u}(0,t)\right\rvert})+O(\varepsilon^{2})+O(\delta_{0})\right\rvert}}dt=\left(\mathinner{\!\left\lvert\dot{q}(t)\right\rvert}+O(\delta_{0})\right)dt\,.

The rest of the proof of 4.7 (which does not require holomorphicity) gives us the result. ∎

Proposition 4.11.

Let aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) be any generator and consider Ψ1(a)C(ΩξMK)\varPsi_{1}(a)\in C_{-\ast}(\varOmega_{\xi}M_{K}). Then

𝔞(a)=𝔞(Ψ1a).\mathfrak{a}(a)=\mathfrak{a}(\varPsi_{1}a)\,.

The same is also true for the chain map

rΨ1:CWΛK(F,F)C(BMK).r_{\ast}\circ\varPsi_{1}\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow C_{-\ast}(BM_{K})\,.
Proof.

By 4.8 we have that

0𝔞(a)𝔞(Ψ1a),0\leq\mathfrak{a}(a)-\mathfrak{a}(\varPsi_{1}a)\,,

and to prove equality, it is enough to show that for any aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F), there exists a transversely cut out JJ-holomorphic disk u¯(a)u\in\overline{\mathcal{M}}(a) with

u1(WK)u𝑑λτ=0.\int_{u^{-1}(W_{K})}u^{\ast}d\lambda_{\tau}=0\,.

We let

u0:TWK,u_{0}\colon\thinspace T\longrightarrow W_{K}\,,

be the JJ-holomorphic half strip that is the cone over the Reeb chord aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F). In geodesic normal coordinates at (q,p)=u0(s,t)(q,p)=u_{0}(s,t), we have that the tangent space of imu0\operatorname{im}u_{0} at (q,p)(q,p) is Tu0(s,t)imu0=span{pp,pq}T_{u_{0}(s,t)}\operatorname{im}u_{0}=\mathrm{span}\left\{p\partial_{p},p\partial_{q}\right\} which means that u0dλτ=0u_{0}^{\ast}d\lambda_{\tau}=0 by 4.4 and hence by 4.7 we have

(4.18) 𝔞(a)=L(γ).\mathfrak{a}(a)=L(\gamma)\,.

What is left to show is that u0u_{0} is transversely cut out. Consider the following space of vector fields along u0u_{0}

V={ηkerDu0|I(πη,πη)<0}.V=\left\{\eta\in\ker D_{u_{0}}\;|\;I(\pi_{\ast}\eta,\pi_{\ast}\eta)<0\right\}\,.

where π\pi is the projection π:WKMK\pi\mathrel{\mathop{\ordinarycolon}}W_{K}\longrightarrow M_{K} along the Liouville flow, and II is the index form (see (4.21) below for a definition). By 3.4 we have

indDu0=indγ.\operatorname{ind}D_{u_{0}}=\operatorname{ind}\gamma\,.

That is, indDu0\operatorname{ind}D_{u_{0}} is equal to the dimension of the maximal subspace of the space of sections of γTMK\gamma^{\ast}TM_{K} on which II is negative definite. The projection

π|kerDu0:kerDu0γTMK\mathinner{\pi_{\ast}\rvert}_{\ker D_{u_{0}}}\colon\thinspace\ker D_{u_{0}}\longrightarrow\gamma^{\ast}TM_{K}

is injective by unique continuation (cf. [Wen16, Corollary 2.27]), which implies that we have

(4.19) dimVindDu0.\dim V\leq\operatorname{ind}D_{u_{0}}\,.

For vkerDu0v\in\ker D_{u_{0}} we have that uε=expu0(εv)u_{\varepsilon}=\exp_{u_{0}}(\varepsilon v) is a disk that is near to u0u_{0} for small ε>0\varepsilon>0. In particular, it is a solution of the Floer equation (3.3) up to the first order. By 4.10 and (4.18) we get

0>L(γε)𝔞(a)=L(γε)L(γ),0>L(\gamma_{\varepsilon})-\mathfrak{a}(a)=L(\gamma_{\varepsilon})-L(\gamma)\,,

which in turn implies

(4.20) 0>E(γε)E(γ),0>E(\gamma_{\varepsilon})-E(\gamma)\,,

where E(γ)=01|γ˙(t)|2𝑑tE(\gamma)=\int_{0}^{1}\mathinner{\!\left\lvert\dot{\gamma}(t)\right\rvert}^{2}dt is the energy of the curve γ\gamma.

Now, by defining E(s) . . =E(ev(expu0(sv)))E(s)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=E(\operatorname{ev}(\exp_{u_{0}}(sv))) we compute

(4.21) d2ds2E(s)|s=0\displaystyle\frac{d^{2}}{ds^{2}}\mathinner{E(s)\biggr{\rvert}}_{s=0} =d2ds2E(ev(expu0(sv)))devexpu0(sv)(dexpu0(sv))|s=0\displaystyle=\frac{d^{2}}{ds^{2}}\mathinner{E(\operatorname{ev}(\exp_{u_{0}}(sv)))d\operatorname{ev}_{\exp_{u_{0}}(sv)}(d\exp_{u_{0}}(sv))\biggr{\rvert}}_{s=0}
=d2ds2E(ev(u0))devu0(v)=I(πv,πv),\displaystyle=\frac{d^{2}}{ds^{2}}E(\operatorname{ev}(u_{0}))d\operatorname{ev}_{u_{0}}(v)=I(\pi_{\ast}v,\pi_{\ast}v)\,,

where II is the index form, see e.g. [Jos08, Section 4.1]. The Taylor expansion of E(ε)E(\varepsilon) around ε=0\varepsilon=0 is

E(ε)E(0)=ε2I(πv,πv)+O(ε3)<(4.20)0.E(\varepsilon)-E(0)=\varepsilon^{2}I(\pi_{\ast}v,\pi_{\ast}v)+O(\varepsilon^{3})\overset{\eqref{eq:difference_of_energies_negative}}{<}0\,.

Hence for small enough ε>0\varepsilon>0, we obtain I(πv,πv)<0I(\pi_{\ast}v,\pi_{\ast}v)<0 and consequently vVv\in V. Therefore we have

dimkerDu0dimV(4.19)indDu0,\dim\ker D_{u_{0}}\leq\dim V\overset{\eqref{eq:dim_jacobi_equals_index_of_lin_op}}{\leq}\operatorname{ind}D_{u_{0}}\,,

which concludes dimcokerDu0=0\dim\operatorname{coker}D_{u_{0}}=0 and therefore u0u_{0} is transversely cut out.

The same proof shows that rΨ1r_{\ast}\circ\varPsi_{1} is diagonal with respect to the action and length filtrations of Reeb chords and chains of broken geodesic loops, respectively, because of [Mil63, Lemma 15.4]. Namely, the index of the Hessian EE_{\ast\ast} is equal to the index of EE_{\ast\ast} restricted to the tangent space TγBMKT_{\gamma}BM_{K}. ∎

4.5. Isomorphism between CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and Ccell(BMK)C^{\mathrm{cell}}_{-\ast}(BM_{K})

The goal of this section is to show that there is a chain isomorphism between CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and Ccell(BMK)C^{\mathrm{cell}}_{-\ast}(BM_{K}). The outline of the proof is the following. Given a generator aCWΛK(F,F)a\in CW^{\ast}_{\varLambda_{K}}(F,F) we consider the trivial JJ-holomorphic half strip u0(a)u_{0}\in\mathcal{M}(a) as in Section 4.4. By the genericity of the metric as in (A.1) we show that the evaluation map defined in Section 3.4 is transverse to the infinite dimensional stable manifold of the geodesic γ\gamma in BMKBM_{K}. This gives a chain isomorphism between CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and Ccell(BMK)C^{\mathrm{cell}}_{-\ast}(BM_{K}) by identifying a neighborhood of u0(a)u_{0}\in\mathcal{M}(a) with the unstable manifold of the geodesic γBMK\gamma\in BM_{K} which correspnds to the generator aa.

We use the notation

[x1,x2)C . . =x1C/x2C,\mathcal{F}_{[x_{1},x_{2})}C\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mathcal{F}_{x_{1}}C/\mathcal{F}_{x_{2}}C\,,

and order the generators of CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) by their action

0\displaystyle 0 𝔞(a1)<𝔞(a2)<\displaystyle\leq\mathfrak{a}(a_{1})<\mathfrak{a}(a_{2})<\cdots

Pick a strictly increasing sequence of numbers {ai}i=1\left\{a_{i}\right\}_{i=1}^{\infty} so that

0𝔞(a1)<A1<𝔞(a2)<A2<,\displaystyle 0\leq\mathfrak{a}(a_{1})<A_{1}<\mathfrak{a}(a_{2})<A_{2}<\cdots\,,

and define

AiCWΛK(F,F)\displaystyle\mathcal{F}_{A_{i}}CW^{\ast}_{\varLambda_{K}}(F,F) . . ={cCWΛK(F,F)|𝔞(c)<Ai}\displaystyle\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{c\in CW^{\ast}_{\varLambda_{K}}(F,F)\;|\;\mathfrak{a}(c)<A_{i}\right\}
AiC(BMK)\displaystyle\mathcal{F}_{A_{i}}C_{-\ast}(BM_{K}) . . ={σC(BMK)|𝔞(σ)<Ai}\displaystyle\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{\sigma\in C_{-\ast}(BM_{K})\;|\;\mathfrak{a}(\sigma)<A_{i}\right\}
AiCcell(BMK)\displaystyle\mathcal{F}_{A_{i}}C_{-\ast}^{\text{cell}}(BM_{K}) . . ={σCcell(BMK)|𝔞(σ)<Ai}.\displaystyle\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{\sigma\in C_{-\ast}^{\text{cell}}(BM_{K})\;|\;\mathfrak{a}(\sigma)<A_{i}\right\}\,.

We extend the filtration to all of \mathbb{Z} by letting Ai=0A_{i}=0 for every i0i\leq 0.

Note that the ordering of the generators of CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) gives an ordering of the generators of C(BMK)C_{-\ast}(BM_{K}) and Ccell(BMK)C_{-\ast}^{\text{cell}}(BM_{K}) by 4.11.

Recall the definition of the retraction

r:AiΩpwMKAiBMK,r\colon\thinspace\mathcal{F}_{A_{i}}\varOmega^{\text{pw}}M_{K}\longrightarrow\mathcal{F}_{A_{i}}BM_{K}\,,

defined in the proof of 4.3: Let γAiΩpwMK\gamma\in\mathcal{F}_{A_{i}}\varOmega^{\text{pw}}M_{K} be any loop with L(γ)=01|γ˙|𝑑t<aiL(\gamma)=\int_{0}^{1}\mathinner{\!\left\lvert\dot{\gamma}\right\rvert}dt<a_{i}. Pick a subdivision of the domain of [0,1][0,1]

0=t0<t1<<tN1<1=tN,0=t_{0}<t_{1}<\cdots<t_{N-1}<1=t_{N}\,,

which is fine enough so that ρ(γ(ti1),γ(ti))<ε\rho(\gamma(t_{i-1}),\gamma(t_{i}))<\varepsilon for some ε>0\varepsilon>0 small enough. Then r(γ)r(\gamma) is defined so that

r(γ)|[ti1,ti]= unique minimal geodesic of length<ε from γ(ti1) to γ(ti).\mathinner{r(\gamma)\rvert}_{[t_{i-1},t_{i}]}=\text{ unique minimal geodesic of length}<\varepsilon\text{ from }\gamma(t_{i-1})\text{ to }\gamma(t_{i})\,.

Then we define

r:AiCcell(ΩpwMK)\displaystyle r_{\ast}\colon\thinspace\mathcal{F}_{A_{i}}C_{-\ast}^{\mathrm{cell}}(\varOmega^{\text{pw}}M_{K}) AiCcell(BMK)\displaystyle\longrightarrow\mathcal{F}_{A_{i}}C_{-\ast}^{\mathrm{cell}}(BM_{K})
σ\displaystyle\sigma rσ.\displaystyle\longmapsto r\circ\sigma\,.
Theorem 4.12.

The map

rΨ1:CWΛK(F,F)\displaystyle r_{\ast}\circ\varPsi_{1}\colon\thinspace CW^{\ast}_{\varLambda_{K}}(F,F) Ccell(BMK)\displaystyle\longrightarrow C_{-\ast}^{\mathrm{cell}}(BM_{K})
a\displaystyle a rev[¯(a)],\displaystyle\longmapsto r_{\ast}\circ\operatorname{ev}_{\ast}[\overline{\mathcal{M}}(a)]\,,

is an isomorphism

Proof.

We first show that for any ii\in\mathbb{Z} the map

rΨ1:[Ai1,Ai)CWΛK(F,F)\displaystyle r_{\ast}\circ\varPsi_{1}\colon\thinspace\mathcal{F}_{[A_{i-1},A_{i})}CW^{\ast}_{\varLambda_{K}}(F,F) [Ai1,Ai)Ccell(BMK)\displaystyle\longrightarrow\mathcal{F}_{[A_{i-1},A_{i})}C_{-\ast}^{\mathrm{cell}}(BM_{K})
a\displaystyle a rev[¯(a)],\displaystyle\longmapsto r_{\ast}\circ\operatorname{ev}_{\ast}[\overline{\mathcal{M}}(a)]\,,

is an isomorphism.

By the definition of the numbers {Ai}i=0\left\{A_{i}\right\}_{i=0}^{\infty} there is only one generator a[Ai1,Ai)CWΛK(F,F)a\in\mathcal{F}_{[A_{i-1},A_{i})}CW^{\ast}_{\varLambda_{K}}(F,F). Denote its degree by λ\lambda. By 4.9, there is exactly one generator σ[Ai1,Ai)Cλcell(BMK)\sigma\in\mathcal{F}_{[A_{i-1},A_{i})}C^{\mathrm{cell}}_{-\lambda}(BM_{K}) that corresponds to aa. We think of σ\sigma as the unstable manifold of the geodesic γ\gamma corresponding to aa. Since both [Ai1,Ai)CWΛK(F,F)\mathcal{F}_{[A_{i-1},A_{i})}CW^{\ast}_{\varLambda_{K}}(F,F) and [Ai1,Ai)Ccell(BMK)\mathcal{F}_{[A_{i-1},A_{i})}C^{\mathrm{cell}}_{-\ast}(BM_{K}) only contains one generator each, we only need to show that

(4.22) σ=(rΨ1)(a).\sigma=(r_{\ast}\circ\varPsi_{1})(a)\,.

By 4.11 we already know that the trivial JJ-holomorphic half strip u0¯(a)u_{0}\in\overline{\mathcal{M}}(a) over aa is so that ev(u0)=γ\operatorname{ev}(u_{0})=\gamma. To prove that equation (4.22) holds it is enough to consider the map

rev:¯(a)BMK,r\circ\operatorname{ev}\colon\thinspace\overline{\mathcal{M}}(a)\longrightarrow BM_{K}\,,

and show that it is locally surjective at γimσBMK\gamma\in\operatorname{im}\sigma\subset BM_{K}. We do this by showing that it is a submersion. That is, we consider

d(rev)u0:Tu0¯(a)TγBMK,d(r\circ\operatorname{ev})_{u_{0}}\colon\thinspace T_{u_{0}}\overline{\mathcal{M}}(a)\longrightarrow T_{\gamma}BM_{K}\,,

and we show that it is surjective onto the image of σ\sigma. As noted above, σ\sigma should be thought of as the unstable manifold of γ\gamma inside BMKBM_{K} with respect to the energy functional EE. The following composition

Tu0¯(a){T_{u_{0}}\overline{\mathcal{M}}(a)}TγΩpwMK{T_{\gamma}\varOmega^{\mathrm{pw}}M_{K}}TγBMK{T_{\gamma}BM_{K}}devu0\scriptstyle{d\operatorname{ev}_{u_{0}}}drγ\scriptstyle{dr_{\gamma}}

is described as follows. Pick a subdivision of the domain of γ\gamma

0t0<t1<<tN1,0\leq t_{0}<t_{1}<\cdots<t_{N}\leq 1\,,

for some N+N\in\mathbb{Z}_{+}. The tangent space of ΩpwMK\varOmega^{\mathrm{pw}}M_{K} at γ\gamma has the following splitting

TγΩpwMK=TγBMKT,T_{\gamma}\varOmega^{\mathrm{pw}}M_{K}=T_{\gamma}BM_{K}\oplus T^{\prime}\,,

by [Mil63, Lemma 15.3, 15.4]. Here TγBMKT_{\gamma}BM_{K} is the space of broken Jacobi fields vanishing at the endpoints, and TT^{\prime} is the space of all vector fields WW along γ\gamma so that W(tk)=0W(t_{k})=0 for every k{1,,N}k\in\left\{1,\ldots,N\right\} (cf. [Mil63, Section 15]). Furthermore we write

TγBMK=TγσT+,T_{\gamma}BM_{K}=T_{\gamma}\sigma\oplus T^{+}\,,

where TγσT_{\gamma}\sigma is the (maximal) subspace of TγBMKT_{\gamma}BM_{K} on which the Hessian EE_{\ast\ast} is negative definite, and T+TγBMKT^{+}\subset T_{\gamma}BM_{K} is the subspace on which EE_{\ast\ast} is positive semidefinite. We will show that for any non-zero vTu0¯(a)v\in T_{u_{0}}\overline{\mathcal{M}}(a), its image devu0(v)d\operatorname{ev}_{u_{0}}(v) does not lie in TT^{\prime}, and that the image d(rev)u0(v)d(r\circ\operatorname{ev})_{u_{0}}(v) does not lie in T+T^{+}.

devu0d\operatorname{ev}_{u_{0}} is transverse to TT^{\prime}:

Consider any non-zero vTu0¯(a)=kerDu0v\in T_{u_{0}}\overline{\mathcal{M}}(a)=\ker D_{u_{0}}, where

Du0:Wκ2,2(D3,u0TWK)Wκ1,2(D3,Λ0,1Ju0TWK)D_{u_{0}}\colon\thinspace W^{2,2}_{\kappa}(D_{3},u_{0}^{\ast}TW_{K})\longrightarrow W^{1,2}_{\kappa}(D_{3},\varLambda^{0,1}\otimes_{J}u_{0}^{\ast}TW_{K})

is the linearization of ¯J\overline{\partial}_{J} at u0u_{0}. Here Wκ2,2W^{2,2}_{\kappa} is the Sobolev space W2,2W^{2,2} with weight eκse^{\kappa s} for some small κ>0\kappa>0 at the positive punctures in the domain D3D_{3}. The differential of the evaluation map ev\operatorname{ev} is a trace operator on Wκ2,2(D3,u0TWK)W^{2,2}_{\kappa}(D_{3},{u_{0}}^{\ast}TW_{K}), so devu0(v)d\operatorname{ev}_{u_{0}}(v) is a vector field in WKW_{K} along γMKWK\gamma\subset M_{K}\subset W_{K}. Assume that vv is such that v . . =devu0(v)Tv^{\prime}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=d\operatorname{ev}_{u_{0}}(v)\in T^{\prime}, that is v(γ(tk))=0v^{\prime}(\gamma(t_{k}))=0 for every k{1,,N}k\in\left\{1,\ldots,N\right\}. Since devu0d\operatorname{ev}_{u_{0}} is a restriction to γTMK\gamma^{\ast}TM_{K}, we assume that vv is so that v(γ(tk))=0v(\gamma(t_{k}))=0 for every k{1,,N}k\in\left\{1,\ldots,N\right\}. We consider the subspace

A . . ={vWκ2,2(D3,u0TWK)|v(γ(tk))=0,k{1,,N}}.A\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\left\{v\in W^{2,2}_{\kappa}(D_{3},{u_{0}}^{\ast}TW_{K})\;|\;v(\gamma(t_{k}))=0,\,k\in\left\{1,\ldots,N\right\}\right\}\,.

It is closed and has codimension NN. The restricted linearized operator Du0|A\mathinner{D_{u_{0}}\rvert}_{A} is therefore a Fredholm operator with index

indDu0|A=indDu0N.\operatorname{ind}\mathinner{D_{u_{0}}\rvert}_{A}=\operatorname{ind}D_{u_{0}}-N\,.

If we pick NN large enough by making the subdivision of the domain of loops fine enough, the index indDu0|A\operatorname{ind}\mathinner{D_{u_{0}}\rvert}_{A} is negative. Hence kerDu0A\ker D_{u_{0}}\cap A is empty for generic choices of almost complex structures on WKW_{K}. This means that imdevu0T={0}\operatorname{im}d\operatorname{ev}_{u_{0}}\cap T^{\prime}=\left\{0\right\}.

d(rev)u0d(r\circ\operatorname{ev})_{u_{0}} is transverse to T+T^{+}:

Next, we show that for any non-zero vkerDu0v\in\ker D_{u_{0}}, its projection

v′′ . . =d(rev)u0(v)TγBMKv^{\prime\prime}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=d(r\circ\operatorname{ev})_{u_{0}}(v)\in T_{\gamma}BM_{K}

does not lie in T+T^{+}. We consider the path sexpu0(sv)s\longmapsto\exp_{u_{0}}(sv) for s(0,ε)s\in(0,\varepsilon) with ε\varepsilon small enough. Then by the proof of 4.11 we have for every s(0,ε)s\in(0,\varepsilon) that

E((rev)(u0))>E((rev)(expu0(sv))).E((r\circ\operatorname{ev})(u_{0}))>E((r\circ\operatorname{ev})(\exp_{u_{0}}(sv)))\,.

Repeating the argument in the proof of 4.11 gives I(v′′,v′′)<0I(v^{\prime\prime},v^{\prime\prime})<0, which shows that v′′=d(rev)u0(ξ)v^{\prime\prime}=d(r\circ\operatorname{ev})_{u_{0}}(\xi) does not lie in T+T^{+}.

Therefore rΨ1:[Ai1,Ai)CWΛK(F,F)[Ai1,Ai)Ccell(BMK)r_{\ast}\circ\varPsi_{1}\colon\thinspace\mathcal{F}_{[A_{i-1},A_{i})}CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow\mathcal{F}_{[A_{i-1},A_{i})}C_{-\ast}^{\mathrm{cell}}(BM_{K}) is an isomorphism.

The filtrations on CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) and Ccell(BMK)C_{-\ast}^{\text{cell}}(BM_{K}) are both bounded from below which gives an isomorphism AiCWΛK(F,F)AiCcell(BMK)\mathcal{F}_{A_{i}}CW^{\ast}_{\varLambda_{K}}(F,F)\cong\mathcal{F}_{A_{i}}C_{-\ast}^{\text{cell}}(BM_{K}) for every ii\in\mathbb{Z}. Thus every square in the following diagram commutes.

{\cdots}AiCWΛK(F,F){\mathcal{F}_{A_{i}}CW^{\ast}_{\varLambda_{K}}(F,F)}ai+1CWΛK(F,F){\mathcal{F}_{a_{i+1}}CW^{\ast}_{\varLambda_{K}}(F,F)}{\cdots}{\cdots}AiCcell(BMK){\mathcal{F}_{A_{i}}C_{-\ast}^{\text{cell}}(BM_{K})}ai+1Ccell(BMK){\mathcal{F}_{a_{i+1}}C_{-\ast}^{\text{cell}}(BM_{K})}{\cdots}{\subset}rΨ1\scriptstyle{r_{\ast}\circ\varPsi_{1}}\scriptstyle{\cong}{\subset}rΨ1\scriptstyle{r_{\ast}\circ\varPsi_{1}}\scriptstyle{\cong}{\subset}{\subset}{\subset}{\subset}

We then pass to colimits to obtain the isomorphism

CWΛK(F,F)Ccell(BMK).CW^{\ast}_{\varLambda_{K}}(F,F)\cong C_{-\ast}^{\text{cell}}(BM_{K})\,.

Proof of 4.1..

4.1 is now an immediate corollary of 4.12, because there is a chain homotopy equivalence Ccell(BMK)C(ΩξMK)C_{-\ast}^{\text{cell}}(BM_{K})\simeq C_{-\ast}(\varOmega_{\xi}M_{K}). So in particular we have H(ΩξMK)Hcell(BMK)H_{-\ast}(\varOmega_{\xi}M_{K})\cong H^{\mathrm{cell}}_{-\ast}(BM_{K}) via srs\circ r_{\ast} defined in (4.3) and (4.4). Hence

Ψ1:HWΛK(F,F)H(ΩξMK),\varPsi_{1}\colon\thinspace HW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow H_{-\ast}(\varOmega_{\xi}M_{K})\,,

is an isomorphism. ∎

5. Applications

The first goal of this section is to equip HWΛK(F,F)HW^{\ast}_{\varLambda_{K}}(F,F) and H(ΩξMK)H_{-\ast}(\varOmega_{\xi}M_{K}) with the structure of [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-modules. The second goal is then to consider the case when S=SnS=S^{n} and exhibit examples of codimension 2 knots KSnK\subset S^{n} where the Alexander invariant is related to CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) as [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-modules. From this we draw the conclusion that the unit conormal of KK knows about the smooth topology of KK beyond the fundamental group.

After we have discussed the [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-module structures in Section 5.1, we will provide background material surrounding the Alexander invariant in Sections 5.2 and 5.3. Then, in Section 5.4 we use the Leray–Serre spectral sequence associated with the path-loop fibration to relate the Alexander invariant to CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) as a [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-module.

5.1. [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-module structures on HWΛK(F,F)HW^{\ast}_{\varLambda_{K}}(F,F) and H(ΩMK)H_{-\ast}(\varOmega M_{K})

Consider any homotopy class [γ]π1(MK)[\gamma]\in\pi_{1}(M_{K}) represented by the unique minimizing geodesic γ\gamma in the given homotopy class. Via the cell structure of BMKBM_{K}, we associate to γ\gamma a generator σγH(ΩMK)\sigma_{\gamma}\in H_{-\ast}(\varOmega M_{K}). Then consider the map

π1(MK)×H(ΩMK)\displaystyle\pi_{1}(M_{K})\times H_{-\ast}(\varOmega M_{K}) H(ΩMK)\displaystyle\longrightarrow H_{-\ast}(\varOmega M_{K})
(γ,σ)\displaystyle(\gamma,\sigma) γσ . . =(1)σγP(σσγ),\displaystyle\longmapsto\gamma\sigma\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=(-1)^{\sigma_{\gamma}}P(\sigma\otimes\sigma_{\gamma})\,,

where PP denotes the Pontryagin product as in (3.2).

Lemma 5.1.

The map (γ,σ)γσ(\gamma,\sigma)\longmapsto\gamma\sigma defines a group action of π1(MK)\pi_{1}(M_{K}) on H(ΩMK)H_{-\ast}(\varOmega M_{K}).

Proof.

Let [γ1],[γ2]π1(MK)[\gamma_{1}],[\gamma_{2}]\in\pi_{1}(M_{K}). As above, we assign to γ1\gamma_{1} and γ2\gamma_{2} the cohomology classes σγ1,σγ2H(ΩMK)\sigma_{\gamma_{1}},\sigma_{\gamma_{2}}\in H_{-\ast}(\varOmega M_{K}). Assign to the composition γ1γ2\gamma_{1}\gamma_{2} the cohomology class

σγ1γ2 . . =σγ1σγ2=(1)|σγ1|P(σγ2σγ1)H(ΩMK).\sigma_{\gamma_{1}\gamma_{2}}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\sigma_{\gamma_{1}}\circ\sigma_{\gamma_{2}}=(-1)^{\mathinner{\!\left\lvert\sigma_{\gamma_{1}}\right\rvert}}P(\sigma_{\gamma_{2}}\otimes\sigma_{\gamma_{1}})\in H_{-\ast}(\varOmega M_{K})\,.

Since PP is associative up to a sign in cohomology we have

γ1(γ2σ)\displaystyle\gamma_{1}(\gamma_{2}\sigma) =P(P(σσγ2)σγ1)=(1)|σγ1|P(σP(σγ2σγ1))\displaystyle=P(P(\sigma\otimes\sigma_{\gamma_{2}})\otimes\sigma_{\gamma_{1}})=(-1)^{\mathinner{\!\left\lvert\sigma_{\gamma_{1}}\right\rvert}}P(\sigma\otimes P(\sigma_{\gamma_{2}}\otimes\sigma_{\gamma_{1}}))
=(1)|σγ1|+|σγ2|P(σσγ1γ2)=(γ1γ2)σ.\displaystyle=(-1)^{\mathinner{\!\left\lvert\sigma_{\gamma_{1}}\right\rvert}+\mathinner{\!\left\lvert\sigma_{\gamma_{2}}\right\rvert}}P(\sigma\otimes\sigma_{\gamma_{1}\gamma_{2}})=(\gamma_{1}\gamma_{2})\sigma\,.

By linearity we extend the action to a [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-module structure on H(ΩMK)H_{-\ast}(\varOmega M_{K}).

Consider a generator aγHWΛK(F,F)a_{\gamma}\in HW^{\ast}_{\varLambda_{K}}(F,F), and denote by γ\gamma the geodesic that aγa_{\gamma} corresponds to. Via γ\gamma, we let σγH(ΩMK)\sigma_{\gamma}\in H_{-\ast}(\varOmega M_{K}) be the cohomology class corresponding to aγHWΛK(F,F)a_{\gamma}\in HW^{\ast}_{\varLambda_{K}}(F,F). Then define

π1(MK)×HWΛK(F,F)\displaystyle\pi_{1}(M_{K})\times HW^{\ast}_{\varLambda_{K}}(F,F) HWΛK(F,F)\displaystyle\longrightarrow HW^{\ast}_{\varLambda_{K}}(F,F)
(γ,a)\displaystyle(\gamma,a) γa . . =(1)|aγ|μ2(aaγ).\displaystyle\longmapsto\gamma a\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=(-1)^{\mathinner{\!\left\lvert a_{\gamma}\right\rvert}}\mu^{2}(a\otimes a_{\gamma})\,.
Lemma 5.2.

The map (γ,a)γa(\gamma,a)\longmapsto\gamma a defines a group action of π1(MK)\pi_{1}(M_{K}) on HWΛK(F,F)HW^{\ast}_{\varLambda_{K}}(F,F).

Proof.

Let σγ1,σγ2\sigma_{\gamma_{1}},\sigma_{\gamma_{2}} and σγ1γ2\sigma_{\gamma_{1}\gamma_{2}} be as in the proof of 5.1 above. Let aHWΛK(F,F)a\in HW^{\ast}_{\varLambda_{K}}(F,F) be any generator. Because μ2\mu^{2} is associative up to a sign in cohomology we have

(5.1) γ1(γ2a)=(1)|aγ1|+|aγ2|μ2(μ2(aaγ2)aγ1)=(1)|aγ2|μ2(aμ2(aγ2aγ1)).\gamma_{1}(\gamma_{2}a)=(-1)^{\mathinner{\!\left\lvert a_{\gamma_{1}}\right\rvert}+\mathinner{\!\left\lvert a_{\gamma_{2}}\right\rvert}}\mu^{2}(\mu^{2}(a\otimes a_{\gamma_{2}})\otimes a_{\gamma_{1}})=(-1)^{\mathinner{\!\left\lvert a_{\gamma_{2}}\right\rvert}}\mu^{2}(a\otimes\mu^{2}(a_{\gamma_{2}}\otimes a_{\gamma_{1}}))\,.

Because {Ψm}k=1\left\{\varPsi_{m}\right\}_{k=1}^{\infty} is an AA_{\infty}-homomorphism, we glue the two disks contributing to Ψ1(aγ1)=σγ1\varPsi_{1}(a_{\gamma_{1}})=\sigma_{\gamma_{1}} and Ψ1(aγ2)=σγ2\varPsi_{1}(a_{\gamma_{2}})=\sigma_{\gamma_{2}} to obtain

(5.2) P(Ψ1(aγ2)Ψ1(aγ1))=Ψ1(μ2(aγ2aγ1)).P(\varPsi_{1}(a_{\gamma_{2}})\otimes\varPsi_{1}(a_{\gamma_{1}}))=\varPsi_{1}(\mu^{2}(a_{\gamma_{2}}\otimes a_{\gamma_{1}}))\,.

Hence there exists a JJ-holomorphic disk in the symplectization of WK\partial W_{K} with two positive punctures aγ1a_{\gamma_{1}} and aγ2a_{\gamma_{2}}. Define aγ1γ2 . . =(1)|σγ1|μ2(aγ2,aγ1)a_{\gamma_{1}\gamma_{2}}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=(-1)^{\mathinner{\!\left\lvert\sigma_{\gamma_{1}}\right\rvert}}\mu^{2}(a_{\gamma_{2}},a_{\gamma_{1}}). Then (5.2) says that

Ψ1(aγ1γ2)=(1)|σγ1|P(σγ2σγ1)=σγ1γ2.\varPsi_{1}(a_{\gamma_{1}\gamma_{2}})=(-1)^{\mathinner{\!\left\lvert\sigma_{\gamma_{1}}\right\rvert}}P(\sigma_{\gamma_{2}}\otimes\sigma_{\gamma_{1}})=\sigma_{\gamma_{1}\gamma_{2}}\,.

Thus aγ1γ2HWΛK(F,F)a_{\gamma_{1}\gamma_{2}}\in HW^{\ast}_{\varLambda_{K}}(F,F) is the generator corresponding to the concatenation [γ1γ2]π1(MK)[\gamma_{1}\gamma_{2}]\in\pi_{1}(M_{K}). Combining this with (5.1) gives

γ1(γ2a)\displaystyle\gamma_{1}(\gamma_{2}a) =(1)|aγ1|+|aγ2|μ2(μ2(aaγ2)aγ1)\displaystyle=(-1)^{\mathinner{\!\left\lvert a_{\gamma_{1}}\right\rvert}+\mathinner{\!\left\lvert a_{\gamma_{2}}\right\rvert}}\mu^{2}(\mu^{2}(a\otimes a_{\gamma_{2}})\otimes a_{\gamma_{1}})
=(1)|aγ2|μ2(aμ2(aγ2aγ1))=(1)|aγ1γ2|μ2(aaγ1γ2)=(γ1γ2)a\displaystyle=(-1)^{\mathinner{\!\left\lvert a_{\gamma_{2}}\right\rvert}}\mu^{2}(a\otimes\mu^{2}(a_{\gamma_{2}}\otimes a_{\gamma_{1}}))=(-1)^{\mathinner{\!\left\lvert a_{\gamma_{1}\gamma_{2}}\right\rvert}}\mu^{2}(a\otimes a_{\gamma_{1}\gamma_{2}})=(\gamma_{1}\gamma_{2})a

Lastly, we need to prove that if γconstπ1(MK)\gamma_{\text{const}}\in\pi_{1}(M_{K}) is the constant loop, then μ2(aaγconst)=a\mu^{2}(a\otimes a_{\gamma_{\text{const}}})=a. This follows from the definition of Ψ1\varPsi_{1}. The generator of HWΛK(F,F)HW^{\ast}_{\varLambda_{K}}(F,F) which corresponds to the (0-chain of) the constant loop is the unique Lagrangian intersection generator in the compact part of WKW_{K} which corresponds to the unique intersection point of FiF_{i} with FjF_{j}, call it xHWΛK(F,F)x\in HW^{\ast}_{\varLambda_{K}}(F,F). Therefore

γconsta=μ2(ax)=a.\gamma_{\text{const}}a=\mu^{2}(a\otimes x)=a\,.

By linearity we extend the action to a [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-module structure on HWΛK(F,F)HW^{\ast}_{\varLambda_{K}}(F,F).

Theorem 5.3 (1.1).

The isomorphism

Ψ1:HWΛK(F,F)H(ΩMK),\varPsi_{1}\colon\thinspace HW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow H_{-\ast}(\varOmega M_{K})\,,

is an isomorphism of [π1(MK)]\mathbb{Z}[\pi_{1}(M_{K})]-modules.

Proof.

Let aHWΛK(F,F)a\in HW^{\ast}_{\varLambda_{K}}(F,F) be a generator and let [γ]π1(MK)[\gamma]\in\pi_{1}(M_{K}) be a homotopy class represented by a unique minimizing geodesic γ\gamma. Then consider a generator aγHWΛK(F,F)a_{\gamma}\in HW^{\ast}_{\varLambda_{K}}(F,F) so that Ψ1(aγ)=σγ\varPsi_{1}(a_{\gamma})=\sigma_{\gamma}, where σγH(ΩMK)\sigma_{\gamma}\in H_{-\ast}(\varOmega M_{K}) is the cohomology class corresponding to γ\gamma. Then we have

Ψ1(γa)\displaystyle\varPsi_{1}(\gamma a) =(1)|aγ|Ψ1(μ2(aaγ))=(5.2)(1)|aγ|P(Ψ1(a)Ψ1(aγ))\displaystyle=(-1)^{\mathinner{\!\left\lvert a_{\gamma}\right\rvert}}\varPsi_{1}(\mu^{2}(a\otimes a_{\gamma}))\overset{\eqref{eq:pontryagin_product_giving_group_action}}{=}(-1)^{\mathinner{\!\left\lvert a_{\gamma}\right\rvert}}P(\varPsi_{1}(a)\otimes\varPsi_{1}(a_{\gamma}))
=(1)|σγ|P(Ψ1(a)σγ)=γΨ1(a).\displaystyle=(-1)^{\mathinner{\!\left\lvert\sigma_{\gamma}\right\rvert}}P(\varPsi_{1}(a)\otimes\sigma_{\gamma})=\gamma\varPsi_{1}(a)\,.

Remark 5.4.

Note that [π1(MK)]H0(ΩMK)\mathbb{Z}[\pi_{1}(M_{K})]\cong H_{0}(\varOmega M_{K}), and consider C0cell(BMK)C_{0}^{\text{cell}}(BM_{K}) as an AA_{\infty}-algebra with operations {mi}i=1\left\{m_{i}\right\}_{i=1}^{\infty} where m1=0m_{1}=0, m2=Pm_{2}=P is the Pontryagin product, and mi=0m_{i}=0 for i3i\geq 3. We observe that CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F) can be equipped with the structure of a left A3A_{3}-module over C0cell(BMK)C^{\text{cell}}_{0}(BM_{K}). More precisely we define this left A3A_{3}-module structure as a sequence of maps

νCWr:(C0cell(BMK))(r1)CWΛK(F,F)CWΛK(F,F),\nu^{r}_{CW}\colon\thinspace(C^{\text{cell}}_{0}(BM_{K}))^{\otimes(r-1)}\otimes CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow CW^{\ast}_{\varLambda_{K}}(F,F)\,,

defined by

{νCW1(c) . . =μ1(c)νCW2(xc) . . =μ2(axc)νCW3(x2x1c) . . =μ3(ax2ax1c)νCWk(xk1x1c) . . =0,k4,\begin{cases}\nu^{1}_{CW}(c)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mu^{1}(c)\\ \nu^{2}_{CW}(x\otimes c)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mu^{2}(a_{x}\otimes c)\\ \nu^{3}_{CW}(x_{2}\otimes x_{1}\otimes c)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mu^{3}(a_{x_{2}}\otimes a_{x_{1}}\otimes c)\\ \nu^{k}_{CW}(x_{k-1}\otimes\cdots\otimes x_{1}\otimes c)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=0,&k\geq 4\,,\end{cases}

where axia_{x_{i}} is the unique Reeb chord corresponding to xix_{i} via Ψ1\varPsi_{1}. Then by computation we have that {νCWr}r=1\left\{\nu^{r}_{CW}\right\}_{r=1}^{\infty} satisfies the following equation for n{1,2,3}n\in\left\{1,2,3\right\}.

(5.3) i=0n1mni(xn1xi+1νCWi+1(xix1c))\displaystyle\sum_{i=0}^{n-1}m_{n-i}(x_{n-1}\otimes\cdots\otimes x_{i+1}\otimes\nu^{i+1}_{CW}(x_{i}\otimes\cdots\otimes x_{1}\otimes c))
++k<r1k0(1)|c|+kνCWr+1(xr1xk++1μ(xk+xk+1)xkx1c)=0\displaystyle+\sum_{\begin{subarray}{c}\ell+k<r\\ \ell\geq 1\\ k\geq 0\end{subarray}}(-1)^{\mathinner{\!\left\lvert c\right\rvert}+\maltese_{k}}\nu^{r-\ell+1}_{CW}(x_{r-1}\otimes\cdots\otimes x_{k+\ell+1}\otimes\mu^{\ell}(x_{k+\ell}\otimes\cdots\otimes x_{k+1})\otimes x_{k}\otimes\cdots\otimes x_{1}\otimes c)=0

Note that this means that there is a group action up to homotopy of C0cell(BMK)C_{0}^{\text{cell}}(BM_{K}) on CWΛK(F,F)CW^{\ast}_{\varLambda_{K}}(F,F), but there are no higher coherent homotopies. However, this is enough to directly obtain 5.1 and 5.2.

Since Ccell(BMK)C^{\text{cell}}_{-\ast}(BM_{K}) is an AA_{\infty}-algebra, it can be regarded as a left AA_{\infty}-module over itself, and therefore also as a left AA_{\infty}-module over C0cell(BMK)C^{\text{cell}}_{0}(BM_{K}) via the sequence of maps

νcellr:(C0cell(BMK))(r1)Ccell(BMK)Ccell(BMK),\nu^{r}_{\text{cell}}\colon\thinspace(C^{\text{cell}}_{0}(BM_{K}))^{\otimes(r-1)}\otimes C^{\text{cell}}_{-\ast}(BM_{K})\longrightarrow C^{\text{cell}}_{-\ast}(BM_{K})\,,

defined by

{νcell1(y) . . =m1(y)νcell2(xy) . . =P(xy)νcellk(xk1x1y) . . =0,k3.\begin{cases}\nu^{1}_{\text{cell}}(y)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=m_{1}(y)\\ \nu^{2}_{\text{cell}}(x\otimes y)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=P(x\otimes y)\\ \nu^{k}_{\text{cell}}(x_{k-1}\otimes\cdots\otimes x_{1}\otimes y)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=0,&k\geq 3\,.\end{cases}

By a computation we see that {νcellr}r=1\left\{\nu^{r}_{\text{cell}}\right\}_{r=1}^{\infty} satisfies (5.3) for every n+n\in\mathbb{Z}_{+}. For n4n\geq 4 the equation is trivial.

Furthermore we have that the AA_{\infty}-homomorphism {Ψk}k=1\left\{\varPsi_{k}\right\}_{k=1}^{\infty} induces an isomorphism of A3A_{3}-modules over C0cell(BMK)C^{\text{cell}}_{0}(BM_{K}) as follows. The isomorphism of A3A_{3}-modules over C0cell(BMK)C^{\text{cell}}_{0}(BM_{K}) is a sequence of maps

ψr:(C0cell(BMK))(r1)CWΛK(F,F)Ccell(BMK)\psi_{r}\colon\thinspace(C^{\text{cell}}_{0}(BM_{K}))^{\otimes(r-1)}\otimes CW^{\ast}_{\varLambda_{K}}(F,F)\longrightarrow C^{\text{cell}}_{-\ast}(BM_{K})

defined by

ψr(xr1x1c) . . =Ψr(axr1ax1c),\psi_{r}(x_{r-1}\otimes\cdots\otimes x_{1}\otimes c)\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\varPsi_{r}(a_{x_{r-1}}\otimes\cdots\otimes a_{x_{1}}\otimes c)\,,

where axia_{x_{i}} is the unique generator corresponding to xix_{i} via Ψ1\varPsi_{1}. Then by computation we have that {ψr}r=1\left\{\psi_{r}\right\}_{r=1}^{\infty} satisfies the following equation for n{1,2,3}n\in\left\{1,2,3\right\}.

i=0r1νcellri(xr1xi+1ψi+1(xix1c))\displaystyle\sum_{i=0}^{r-1}\nu^{r-i}_{\text{cell}}(x_{r-1}\otimes\cdots\otimes x_{i+1}\otimes\psi_{i+1}(x_{i}\otimes\cdots\otimes x_{1}\otimes c))
=i=0r1ψri(xr1xi+1νCWi+1(xix1c))\displaystyle=\sum_{i=0}^{r-1}\psi_{r-i}(x_{r-1}\otimes\cdots\otimes x_{i+1}\otimes\nu^{i+1}_{CW}(x_{i}\otimes\cdots\otimes x_{1}\otimes c))
+s+t+k=rt,k1s0(1)|c|+kψr+1(xr1xk++1m(xk+xk+1)xkx1c)=0,\displaystyle+\sum_{\begin{subarray}{c}s+t+k=r\\ t,k\geq 1\\ s\geq 0\end{subarray}}(-1)^{\mathinner{\!\left\lvert c\right\rvert}+\maltese_{k}}\psi_{r-\ell+1}(x_{r-1}\otimes\cdots\otimes x_{k+\ell+1}\otimes m_{\ell}(x_{k+\ell}\otimes\cdots\otimes x_{k+1})\otimes x_{k}\otimes\cdots\otimes x_{1}\otimes c)=0\,,

The fact that this is an A3A_{3}-module isomorphism directly implies 5.3.

5.2. Plumbings and infinite cyclic covers

In this section we review standard background material from [Rol76].

Let p,q2p,q\geq 2 and n=p+q+1n=p+q+1. We consider the plumbing of SpS^{p} with SqS^{q}. That is, consider Sp×DqS^{p}\times D^{q} and Sq×DpS^{q}\times D^{p}. By identifying DpD^{p} with the upper hemisphere of SpS^{p}, we have

Dp×DqSp×Dq\displaystyle D^{p}\times D^{q}\subset S^{p}\times D^{q}
Dq×DpSq×Dp.\displaystyle D^{q}\times D^{p}\subset S^{q}\times D^{p}\,.

We then take the disjoint union of Sp×DqS^{p}\times D^{q} with Sq×DpS^{q}\times D^{p} and identify their common submanifolds Dp×DqDq×DpD^{p}\times D^{q}\cong D^{q}\times D^{p} via f:(x,y)(y,x)f\colon\thinspace(x,y)\longmapsto(y,x). We call the resulting space the plumbing of SpS^{p} and SqS^{q}, denoted by Sp#plumbSqS^{p}\#_{\text{plumb}}S^{q}. In short we write

Σ=Sp#plumbSq . . =(Sp×Dq)f(Sq×Dp).\varSigma=S^{p}\#_{\text{plumb}}S^{q}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=(S^{p}\times D^{q})\sqcup_{f}(S^{q}\times D^{p})\,.

We note that SpSqS^{p}\vee S^{q} is the deformation retract of Σ\varSigma. Let K . . =ΣK\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\partial\varSigma and note that it is a (p+q1)(p+q-1)-dimensional sphere. Embed Σ\varSigma into SnS^{n} and consider the complement of its boundary MK . . =SnKM_{K}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=S^{n}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}K; denote its infinite cyclic cover by M~K\widetilde{M}_{K}.

Following [Rol76, Section 5.C] we find the simplicial structure of M~K\widetilde{M}_{K} by cutting along Σ\varSigma. More precisely, let Σ±Σ×(1,1)\varSigma^{\pm}\cong\overset{\circ}{\varSigma}\times(-1,1) be an open bicollar of the interior of Σ\varSigma, and let

Σ+\displaystyle\varSigma^{+} . . =Σ×(0,1)Sn\displaystyle\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\overset{\circ}{\varSigma}\times(0,1)\subset S^{n}
Σ\displaystyle\varSigma^{-} . . =Σ×(1,0)Sn\displaystyle\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\overset{\circ}{\varSigma}\times(-1,0)\subset S^{n}
MΣ\displaystyle M_{\varSigma} . . =SnΣ.\displaystyle\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=S^{n}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\varSigma\,.

Consider infinitely many copies of each of Σ+\varSigma^{+}, Σ\varSigma^{-} and MΣM_{\varSigma}. Denote the copies by MΣ;iM_{\varSigma;i}, Σi+\varSigma^{+}_{i} and Σi\varSigma^{-}_{i} for ii\in\mathbb{Z}. Then consider the disjoint union of all the MΣ;iM_{\varSigma;i} and glue them together by identifying Σi+MΣ;i\varSigma_{i}^{+}\subset M_{\varSigma;i} with Σi+1MΣ;i+1\varSigma_{i+1}^{-}\subset M_{\varSigma;i+1} via the map

Σi+=Σ×(0,1)\displaystyle\varSigma_{i}^{+}=\overset{\circ}{\varSigma}\times(0,1) Σ×(1,0)=Σi+1\displaystyle\longrightarrow\overset{\circ}{\varSigma}\times(-1,0)=\varSigma_{i+1}^{-}
(σ,t)\displaystyle(\sigma,t) (σ,t1).\displaystyle\longmapsto(\sigma,t-1)\,.

Define

(5.4) M~K . . =i=MΣ;i/(Σi+Σi+1).\widetilde{M}_{K}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\coprod_{i=-\infty}^{\infty}M_{\varSigma;i}/(\varSigma^{+}_{i}\sim\varSigma^{-}_{i+1})\,.
[Uncaptioned image]

5.3. The Alexander invariant

In this section we review standard material on the Alexander invariant from [Rol76].

Associated to the open cover 𝒰=(MΣ;i)i=\mathcal{U}=\left(M_{\varSigma;i}\right)_{i=-\infty}^{\infty} of M~K\widetilde{M}_{K} is the sequence of inclusions

i=MΣ;iMΣ;i+1{\coprod_{i=-\infty}^{\infty}M_{\varSigma;i}\cap M_{\varSigma;i+1}}i=MΣ;i{\coprod_{i=-\infty}^{\infty}M_{\varSigma;i}}M~K{\widetilde{M}_{K}}ιi+1\scriptstyle{\iota_{i+1}}ιi\scriptstyle{\iota_{i}}κ\scriptstyle{\kappa}

from which we get a short exact sequence in singular chains (cf. [BT13, Section 8])

(5.5) 0C(i=MΣ;iMΣ;i+1)C(i=MΣ;i)C(M~K)0αβ.\leavevmode\hbox to339.6pt{\vbox to19.75pt{\pgfpicture\makeatletter\hbox{\hskip 169.7984pt\lower-8.84026pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-169.7984pt}{-2.31946pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\quad\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-2.5pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${0}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\quad\hfil&\hfil\hskip 77.84195pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-55.53636pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${C_{\ast}\left(\coprod_{i=-\infty}^{\infty}M_{\varSigma;i}\cap M_{\varSigma;i+1}\right)}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 59.8419pt\hfil&\hfil\hskip 57.2884pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-34.98282pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${C_{\ast}(\coprod_{i=-\infty}^{\infty}M_{\varSigma;i})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 39.28836pt\hfil&\hfil\hskip 39.057pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-16.75142pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${C_{\ast}(\widetilde{M}_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 21.05696pt\hfil&\hfil\hskip 24.80559pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-2.5pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${0}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\quad\hfil\cr}}}\pgfsys@invoke{ }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope}}} { {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-155.98732pt}{0.18054pt}\pgfsys@lineto{-138.78723pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-138.58725pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-18.30347pt}{0.18054pt}\pgfsys@lineto{-1.10338pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-0.9034pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.19312pt}{3.53331pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\alpha_{\ast}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{78.2733pt}{0.18054pt}\pgfsys@lineto{95.47339pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{95.67337pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{82.71582pt}{3.89441pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\beta_{\ast}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{138.38727pt}{0.18054pt}\pgfsys@lineto{155.58736pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{155.78734pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope \pgfsys@invoke{ }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}\,.

Let x=(xi)iC(i=MΣ;iMΣ;i+1)x=(x_{i})_{i\in\mathbb{Z}}\in C_{\ast}\left(\coprod_{i=-\infty}^{\infty}M_{\varSigma;i}\cap M_{\varSigma;i+1}\right). Then

αx=((ιi)(xi)(ιi)(xi1))i,\alpha_{\ast}x=\left((\iota_{i})_{\ast}(x_{i})-(\iota_{i})_{\ast}(x_{i-1})\right)_{i\in\mathbb{Z}}\,,

and for any y=(yi)iC(i=MΣ;i)y=(y_{i})_{i\in\mathbb{Z}}\in C_{\ast}(\coprod_{i=-\infty}^{\infty}M_{\varSigma;i}) we have

βy=i=κ(yi).\beta_{\ast}y=\sum_{i=-\infty}^{\infty}\kappa_{\ast}(y_{i})\,.

Since

MΣ;iMΣ;i+1=Σi+SpSq,M_{\varSigma;i}\cap M_{\varSigma;i+1}=\varSigma_{i}^{+}\simeq S^{p}\vee S^{q}\,,

the short exact sequence (5.5) induces a long exact sequence in homology

i=Hj(SpSq)i=Hj(MΣ;i)Hj(M~K)i=Hj1(SpSq)i=Hj1(MΣ;i)αα.\leavevmode\hbox to328.46pt{\vbox to67.64pt{\pgfpicture\makeatletter\hbox{\hskip 164.23027pt\lower-33.81956pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-164.23027pt}{-20.15974pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\quad\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-3.75pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\cdots}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\quad\hfil&\hfil\hskip 64.15886pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-41.85327pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\bigoplus_{i=-\infty}^{\infty}H_{j}(S^{p}\vee S^{q})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 46.15881pt\hfil&\hfil\hskip 59.53337pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-37.22778pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\bigoplus_{i=-\infty}^{\infty}H_{j}(M_{\varSigma;i})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 41.53333pt\hfil&\hfil\hskip 39.49641pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-17.19083pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{j}(\widetilde{M}_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 21.49637pt\hfil&\hfil\hskip 22.30559pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${{}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\enspace\hfil\cr\vskip 18.00005pt\cr\hfil\enspace\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${{}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\enspace\hfil&\hfil\hskip 67.49915pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-45.19356pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\bigoplus_{i=-\infty}^{\infty}H_{j-1}(S^{p}\vee S^{q})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\hskip 49.4991pt\hfil&\hfil\hskip 62.87367pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-40.56808pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\bigoplus_{i=-\infty}^{\infty}H_{j-1}(M_{\varSigma;i})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\hskip 44.87363pt\hfil&\hfil\hskip 26.05559pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-3.75pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\cdots}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\quad\hfil\cr}}}\pgfsys@invoke{ }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope}}} { {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-147.91919pt}{25.15985pt}\pgfsys@lineto{-127.37881pt}{25.15985pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-127.17883pt}{25.15985pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-34.26123pt}{25.15985pt}\pgfsys@lineto{-10.38055pt}{25.15985pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-10.18057pt}{25.15985pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-26.81058pt}{28.51262pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\alpha_{\ast}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-80.62004pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{73.48605pt}{25.15985pt}\pgfsys@lineto{94.02644pt}{25.15985pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{94.22643pt}{25.15985pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope{}{}{ {}{}{}}{}{{}{}{}}{{{}}{{}}} {}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{ {}{}{}}{}{}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{}{}{{{}{}}}{ {}{}{}}{{{}}{{}}}{}{}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{{}{}{{}}}{{}{}{{}}}{}{}{{}{}{{}}}{{}{}{{}}}{}{}{{}{}{{}}}{{}{}{{}}}{}{}{{}{}{{}}}{{}{}{{}}}{}{}{{}{}{{}}}{{}{}{{}}}{}{}{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{137.81914pt}{25.15985pt}\pgfsys@lineto{142.43022pt}{25.15985pt}\pgfsys@curveto{144.63939pt}{25.15985pt}{146.43022pt}{23.36902pt}{146.43022pt}{21.15985pt}\pgfsys@lineto{146.43022pt}{4.0pt}\pgfsys@curveto{146.43022pt}{1.79083pt}{144.63939pt}{0.0pt}{142.43022pt}{0.0pt}\pgfsys@lineto{-76.62004pt}{0.0pt}\pgfsys@curveto{-78.82921pt}{0.0pt}{-82.41087pt}{0.0pt}{-84.62004pt}{0.0pt}\pgfsys@lineto{-134.93022pt}{0.0pt}\pgfsys@curveto{-137.13939pt}{0.0pt}{-138.93022pt}{-1.79083pt}{-138.93022pt}{-4.0pt}\pgfsys@lineto{-138.93022pt}{-13.65974pt}\pgfsys@curveto{-138.93022pt}{-15.86891pt}{-137.13939pt}{-17.65974pt}{-134.93022pt}{-17.65974pt}\pgfsys@lineto{-130.7191pt}{-17.65974pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-130.51912pt}{-17.65974pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}}{}{}{}{}{} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-80.62004pt}{-2.35277pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-30.92094pt}{-17.65974pt}\pgfsys@lineto{-13.72086pt}{-17.65974pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-13.52087pt}{-17.65974pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-26.8106pt}{-14.30698pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\alpha_{\ast}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{76.82635pt}{-17.65974pt}\pgfsys@lineto{107.46727pt}{-17.65974pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{107.66725pt}{-17.65974pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}\,.

We have

H~j(SpSq)H~j(Sp)H~j(Sq),\widetilde{H}_{j}(S^{p}\vee S^{q})\cong\widetilde{H}_{j}(S^{p})\oplus\widetilde{H}_{j}(S^{q})\,,

where H~\widetilde{H} denotes reduced homology. Since MΣ=SnΣM_{\varSigma}=S^{n}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\varSigma, Alexander duality gives that

H~j(MΣ;i)H~nj1(Σ)H~nj1(SpSq)H~nj1(Sp)H~nj1(Sq).\widetilde{H}_{j}(M_{\varSigma;i})\cong\widetilde{H}^{n-j-1}(\varSigma)\cong\widetilde{H}^{n-j-1}(S^{p}\vee S^{q})\cong\widetilde{H}^{n-j-1}(S^{p})\oplus\widetilde{H}^{n-j-1}(S^{q})\,.

Since n=p+q+1n=p+q+1, we get

Hj(SpSq)Hj(MΣ;i){,j=0,p,q0,otherwise,H_{j}(S^{p}\vee S^{q})\cong H_{j}(M_{\varSigma;i})\cong\begin{cases}\mathbb{Z},&j=0,p,q\\ 0,&\text{otherwise,}\end{cases}

which means that Hj(M~K)0H_{j}(\widetilde{M}_{K})\cong 0 unless j{0,p,q}j\in\left\{0,p,q\right\}.

Since the group of deck transformations of M~K\widetilde{M}_{K} is infinite cyclic, we choose a generator τAut(M~K,π)\tau\in\mathrm{Aut}(\widetilde{M}_{K},\pi) which induces an automorphism

τ:H(M~K)H(M~K).\tau_{\ast}\colon\thinspace H_{\ast}(\widetilde{M}_{K})\longrightarrow H_{\ast}(\widetilde{M}_{K})\,.

This gives a [t±1]\mathbb{Z}[t^{\pm 1}]-module structure on H(M~K)H_{\ast}(\widetilde{M}_{K}) as follows. Let p(t)=i=srciti[t±1]p(t)=\sum_{i=-s}^{r}c_{i}t^{i}\in\mathbb{Z}[t^{\pm 1}], then for any αH(M~K)\alpha\in H_{\ast}(\widetilde{M}_{K}) let

p(t)α=i=srciτi(α),p(t)\alpha=\sum_{i=-s}^{r}c_{i}\tau_{\ast}^{i}(\alpha)\,,

where τi\tau_{\ast}^{i} is the ii-fold composition power of τ\tau_{\ast}. The Alexander invariant is then defined as H(M~K)H_{\ast}(\widetilde{M}_{K}) considered as a [t±1]\mathbb{Z}[t^{\pm 1}]-module.

Lemma 5.5 ([Rol76, Theorem 7.G.1]).

There exist non-trivial knots KSn+2K\subset S^{n+2} with infinite cyclic knot group, π1(MK)\pi_{1}(M_{K})\cong\mathbb{Z}.

Proof.

Let p,q2p,q\geq 2 and let n=p+q1n=p+q-1. We then consider any KK obtained as Σ\partial\varSigma, where

Σ=Sp#plumbSq.\varSigma=S^{p}\#_{\text{plumb}}S^{q}\,.

Now we have that M~Σ\widetilde{M}_{\varSigma} is simply connected: Every loop in Sn+2S^{n+2} shrinks missing Σ\varSigma since Σ\varSigma is homotopy equivalent to SpSqS^{p}\vee S^{q}. This is because codim(Sp)3\operatorname{codim}(S^{p})\geq 3 and codim(Sq)3\operatorname{codim}(S^{q})\geq 3 in Sn+2S^{n+2}. From the construction of M~K\widetilde{M}_{K} in (5.4) we thus have π1(M~K)1\pi_{1}(\widetilde{M}_{K})\cong 1. Hence, because the group of deck transformations of M~KMK\widetilde{M}_{K}\longrightarrow M_{K} is \mathbb{Z}, we have π1(MK)\pi_{1}(M_{K})\cong\mathbb{Z}.

Refer to caption
Figure 13. The core of the self plumbing of two knotted S2S^{2} embedded in S5S^{5}.

To see that such non-trivial KK exists, we may consider K=(S2#plumbS2)S5K=\partial(S^{2}\#_{\text{plumb}}S^{2})\subset S^{5}, where the core of the plumbing is shown in Fig. 13. The Alexander invariant of KK is non-trivial by a computation (cf. [Rol76, Exercise 7.F.5]). ∎

5.4. Using the Leray–Serre spectral sequence

Consider the knot K=Sp#plumbSqSnK=S^{p}\#_{\mathrm{plumb}}S^{q}\subset S^{n} with p+q=n1p+q=n-1 as in Section 5.2. We use the notation π . . =[π1(MK)]\mathbb{Z}\pi\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mathbb{Z}[\pi_{1}(M_{K})].

Associated to the path-loop fibration

ΩMK{\varOmega M_{K}}PMK{PM_{K}}MK{M_{K}}

is the Leray–Serre spectral sequence. It is first quadrant spectral sequence {Ei,jr,di,jr}i,j\left\{E^{r}_{i,j},d^{r}_{i,j}\right\}_{i,j\in\mathbb{N}} of π\mathbb{Z}\pi-modules which converges:

Ei,j2Hi(MK;Hj(ΩMK))Hi+j(PMK)={π/(t1),i+j=00,otherwise,E^{2}_{i,j}\cong H_{i}(M_{K};H_{j}(\varOmega M_{K}))\Longrightarrow H_{i+j}(PM_{K})=\begin{cases}\mathbb{Z}\pi/(t-1),&i+j=0\\ 0,&\text{otherwise,}\end{cases}

Note that π1(MK)\pi_{1}(M_{K}) is Abelian and hence we can consider Hi(MK;Hj(ΩMK))H_{i}(M_{K};H_{j}(\varOmega M_{K})) as a π\mathbb{Z}\pi-module.

Since C(M~K)C_{\ast}(\widetilde{M}_{K}) is only supported in degrees 0, pp and qq we have the following facts

  • Following [Hat02, Section 3.H] and [Shu10] we have the following identification

    Hi(MK;Hj(ΩMK))Hi(C(M~K)πHj(ΩMK)).H_{i}(M_{K};H_{j}(\varOmega M_{K}))\cong H_{i}\left(C_{\ast}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}H_{j}(\varOmega M_{K})\right)\,.

    Assume that |pq|1\mathinner{\!\left\lvert p-q\right\rvert}\neq 1. Then we trivially have

    Hi(MK;Hj(ΩMK))={Hj(ΩMK),i=0Hp(M~K)πHj(ΩMK),i=pHq(M~K)πHj(ΩMK),i=q0,otherwise,H_{i}(M_{K};H_{j}(\varOmega M_{K}))=\begin{cases}H_{j}(\varOmega M_{K}),&i=0\\ H_{p}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}H_{j}(\varOmega M_{K}),&i=p\\ H_{q}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}H_{j}(\varOmega M_{K}),&i=q\\ 0,&\text{otherwise,}\end{cases}

    because C(M~K)C_{\ast}(\widetilde{M}_{K}) is only supported in {0,p,q}\ast\in\left\{0,p,q\right\}.

  • Ei,j2E^{2}_{i,j} is only supported on the vertical lines i{0,p,q}i\in\left\{0,p,q\right\}.

  • The bottom row is Ei,02=Hi(M~K)E^{2}_{i,0}=H_{i}(\widetilde{M}_{K}), since H0(ΩMK)πH_{0}(\varOmega M_{K})\cong\mathbb{Z}\pi.

Example 5.6.

Consider the case when KK is obtained as the boundary of Sp#plumbSpS2p+1S^{p}\#_{\text{plumb}}S^{p}\subset S^{2p+1} where the core of the plumbing is depicted in Fig. 13. In this case, Ei,j2E^{2}_{i,j} is only supported at the vertical lines i{0,p}i\in\left\{0,p\right\}. For this spectral sequence, the pp-th page is the first page after page 1 that has non-zero differentials. Namely, the pp-th page of this spectral sequence is

[Uncaptioned image]

Where Γ=π/(t1)\varGamma=\mathbb{Z}\pi/(t-1) and every tensor product is taken over π\mathbb{Z}\pi. The differentials at every page succeeding the pp-th page is zero, so in particular we get

Hp(M~K)π/(t1)πHp1(ΩMK)π/(t1)πHWΛK1p(F,F).H_{p}(\widetilde{M}_{K})\cong\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}H_{p-1}(\varOmega M_{K})\cong\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}HW^{1-p}_{\varLambda_{K}}(F,F)\,.
Example 5.7.

Suppose KK is obtained as the boundary of Sp#plumbS2pS3p+1S^{p}\#_{\text{plumb}}S^{2p}\subset S^{3p+1} for p2p\geq 2 where the core of the plumbing is depicted in Fig. 13. In this case the second page of the spectral sequence is only supported at the lines i{0,p,2p}i\in\left\{0,p,2p\right\}. The pp-th page of the spectral sequence is the first page after page 1 that has non-zero differentials and it looks as follows:

[Uncaptioned image]

Immediately from this page, we get an isomorphism of π\mathbb{Z}\pi-modules

(5.6) Hp(M~K)π/(t1)πHp1(ΩMK)π/(t1)πHWΛK1p(F,F).H_{p}(\widetilde{M}_{K})\cong\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}H_{p-1}(\varOmega M_{K})\cong\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}HW^{1-p}_{\varLambda_{K}}(F,F)\,.

Furthermore, the next page with non-zero differentials is page 2p2p, which looks as follows

[Uncaptioned image]

This is the last page with non-zero differentials, so cokerαp1{0}\operatorname{coker}\alpha_{p-1}\cong\left\{0\right\} and hence from page pp, we obtain an exact sequence

(5.7) H2p(M~K)Hp(M~K)πHp1(ΩMK)ΓπH2p2(ΩMK)0β0αp1,\leavevmode\hbox to325.15pt{\vbox to19.75pt{\pgfpicture\makeatletter\hbox{\hskip 162.57669pt\lower-8.84026pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-162.57669pt}{-2.31946pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 23.69383pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-19.38829pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{2p}(\widetilde{M}_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 23.69383pt\hfil&\hfil\hskip 76.67725pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-54.37166pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{p}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}H_{p-1}(\varOmega M_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 58.6772pt\hfil&\hfil\hskip 64.4001pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-42.09451pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\varGamma\otimes_{\mathbb{Z}\pi}H_{2p-2}(\varOmega M_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 46.40005pt\hfil&\hfil\hskip 24.80559pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-2.5pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${0}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\quad\hfil\cr}}}\pgfsys@invoke{ }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-114.98903pt}{0.18054pt}\pgfsys@lineto{-97.78894pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-97.58896pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-110.40762pt}{3.89441pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\beta_{0}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{20.36542pt}{0.18054pt}\pgfsys@lineto{37.5655pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{37.76549pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{21.5848pt}{4.50551pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\alpha_{p-1}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{131.16557pt}{0.18054pt}\pgfsys@lineto{148.36566pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{148.56564pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope \pgfsys@invoke{ }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}\,,

and for each i+i\in\mathbb{Z}_{+} we have an exact sequence

H2p(M~K)Hi(ΩMK)Hp(M~K)πHp1+i(ΩMK)ΓπH2p2+i(ΩMK)βiαp1+i.\leavevmode\hbox to369.66pt{\vbox to19.75pt{\pgfpicture\makeatletter\hbox{\hskip 184.82825pt\lower-8.84026pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-184.82825pt}{-2.31946pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 52.78275pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-48.4772pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{2p}(\widetilde{M}_{K})\otimes H_{i}(\varOmega M_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 52.78275pt\hfil&\hfil\hskip 81.16135pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-58.85576pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{p}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}H_{p-1+i}(\varOmega M_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\hskip 63.1613pt\hfil&\hfil\hskip 68.8842pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-46.57861pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\varGamma\otimes_{\mathbb{Z}\pi}H_{2p-2+i}(\varOmega M_{K})}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\hskip 50.88416pt\hfil\cr}}}\pgfsys@invoke{ }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-79.06276pt}{0.18054pt}\pgfsys@lineto{-61.86267pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-61.66269pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-74.11386pt}{3.89441pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\beta_{i}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{65.25989pt}{0.18054pt}\pgfsys@lineto{82.45998pt}{0.18054pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{82.65996pt}{0.18054pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{62.57585pt}{4.50551pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\alpha_{p-1+i}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope \pgfsys@invoke{ }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}\,.

Furthermore, from page 2p2p we get isomorphisms

kerβjcokerαp+j,\ker\beta_{j}\cong\operatorname{coker}\alpha_{p+j}\,,

for every jj\in\mathbb{N}. In particular, in view of (5.7), we have kerβ0cokerαp\ker\beta_{0}\cong\operatorname{coker}\alpha_{p} where

αp:Hp(M~K)πHp(ΩMK)ΓπH2p1(ΩMK).\alpha_{p}\colon\thinspace H_{p}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}H_{p}(\varOmega M_{K})\longrightarrow\varGamma\otimes_{\mathbb{Z}\pi}H_{2p-1}(\varOmega M_{K})\,.

So if H(ΩMK)H_{\ast}(\varOmega M_{K}) and αp\alpha_{p} is known, we compute Hp(M~K)H_{p}(\widetilde{M}_{K}) by (5.6) but also a quotient of H2p(M~K)H_{2p}(\widetilde{M}_{K}) by exactness of (5.7)

H2p(M~K)/cokerαpkerαp1.H_{2p}(\widetilde{M}_{K})/\operatorname{coker}\alpha_{p}\cong\ker\alpha_{p-1}\,.

Let us summarize what we have.

{Hp(M~K)π/(t1)πHWΛK1p(F,F)H2p(M~K)/cokerαpkerαp1,\displaystyle\begin{cases}H_{p}(\widetilde{M}_{K})\cong\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}HW^{1-p}_{\varLambda_{K}}(F,F)\\ H_{2p}(\widetilde{M}_{K})/\operatorname{coker}\alpha_{p}\cong\ker\alpha_{p-1}\,,\end{cases}

where

αp:Hp(M~K)πHWΛKp(F,F)\displaystyle\alpha_{p}\colon\thinspace H_{p}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}HW^{-p}_{\varLambda_{K}}(F,F) π/(t1)πHWΛK12p(F,F)\displaystyle\longrightarrow\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}HW_{\varLambda_{K}}^{1-2p}(F,F)
αp1:Hp(M~K)πHWΛK1p(F,F)\displaystyle\alpha_{p-1}\colon\thinspace H_{p}(\widetilde{M}_{K})\otimes_{\mathbb{Z}\pi}HW_{\varLambda_{K}}^{1-p}(F,F) π/(t1)πHWΛK22p(F,F).\displaystyle\longrightarrow\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}HW_{\varLambda_{K}}^{2-2p}(F,F)\,.
Example 5.8.

For a slightly more general case, where p2p\geq 2 and q>p+1q>p+1 we consider again KK to be the boundary of Sp#plumbSqSp+q+1S^{p}\#_{\text{plumb}}S^{q}\subset S^{p+q+1} where the core of the plumbing is depicted in Fig. 13. The Leray–Serre spectral sequence is supported at the lines i{0,p,q}i\in\left\{0,p,q\right\}. Exactly like in 5.7, we compute

Hp(M~K)π/(t1)πHWΛK1p(F,F).H_{p}(\widetilde{M}_{K})\cong\mathbb{Z}\pi/(t-1)\otimes_{\mathbb{Z}\pi}HW^{1-p}_{\varLambda_{K}}(F,F)\,.

Let Λunknot\varLambda_{\mathrm{unknot}} denote the unit conormal of the standard embedded Sn2SnS^{n-2}\subset S^{n}. A consequence of these computations is the following theorem.

Theorem 5.9 (1.2).

Let n=5n=5 or n7n\geq 7. Let xMKx\in M_{K} be a point. Then there exists a codimension 2 knot KSnK\subset S^{n} with π1(MK)\pi_{1}(M_{K})\cong\mathbb{Z}, such that ΛKΛx\varLambda_{K}\cup\varLambda_{x} is not Legendrian isotopic to ΛunknotΛx\varLambda_{\mathrm{unknot}}\cup\varLambda_{x}.

Proof.

For the case n=5n=5, consider the knot K=(S2#plumbS2)S5K=\partial(S^{2}\#_{\mathrm{plumb}}S^{2})\subset S^{5} where the core of the plumbing is depicted in Fig. 13. In the case n7n\geq 7 we let p2p\geq 2 and q>p+1q>p+1 and consider K=(Sp#plumbSq)Sp+q+1K=\partial(S^{p}\#_{\mathrm{plumb}}S^{q})\subset S^{p+q+1}, where the core of the plumbing is again depicted in Fig. 13.

We note that for dimensional reasons we have π1(MK)\pi_{1}(M_{K})\cong\mathbb{Z}, but the Alexander invariant shows that KK is non-trivial [Rol76, Section 7.G] (see also 5.5).

The computations in 5.6 and 5.8 show that in particular

Hp(M~K)[π1(MK)]/(t1)[π1(MK)]HWΛK1p(F,F).H_{p}(\widetilde{M}_{K})\cong\mathbb{Z}[\pi_{1}(M_{K})]/(t-1)\otimes_{\mathbb{Z}[\pi_{1}(M_{K})]}HW^{1-p}_{\varLambda_{K}}(F,F)\,.

Since we use classical methods to show that Hp(M~K)H_{p}(\widetilde{M}_{K}) is non-trivial (see 5.5, it follows that HWΛK1p(F,F)HW^{1-p}_{\varLambda_{K}}(F,F) is non-trivial. Consider the unknot Sn2SnS^{n-2}\subset S^{n}, then the complement MunknotM_{\mathrm{unknot}} is homotopy equivalent to a circle, which means that H(ΩMunknot)HWΛunknot(F,F)H_{-\ast}(\varOmega M_{\mathrm{unknot}})\cong HW^{\ast}_{\varLambda_{\mathrm{unknot}}}(F,F) is only supported in degree 0. Therefore we have HWΛK1p(F,F)≇HWΛunknot1p(F,F)HW^{1-p}_{\varLambda_{K}}(F,F)\not\cong HW^{1-p}_{\varLambda_{\mathrm{unknot}}}(F,F) and so ΛKΛx\varLambda_{K}\cup\varLambda_{x} is not Legendrian isotopic to ΛunknotΛx\varLambda_{\mathrm{unknot}}\cup\varLambda_{x}. ∎

Appendix A Monotonicity of JJ-holomorphic half strips

To establish compactness of the moduli spaces (𝒂)\mathcal{M}(\boldsymbol{a}) in Section 3.3 we need to make sure that JJ-holomorphic half strips in (𝒂)\mathcal{M}(\boldsymbol{a}) does not escape to horizontal infinity. Pick a tubular neighborhood of KMKK\subset M_{K} and call it N(K)N(K). Then we decompose MKM_{K} as

MK(SN(K))N(K)([0,)×N(K)),M_{K}\cong(S\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(K))\cup_{\partial N(K)}\left([0,\infty)\times\partial N(K)\right)\,,

where we identify (SN(K))N(K)\partial(S\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(K))\cong\partial N(K) with {0}×N(K)N(K)\left\{0\right\}\times\partial N(K)\cong\partial N(K). Pick a generic Riemannian metric gg on SN(K)S\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(K) such that geodesics are non-degenerate critical points of the length and energy functionals. Define a function

f:[0,)[0,)f\colon\thinspace[0,\infty)\longrightarrow[0,\infty)

so that

(A.1) {f(0)=1f(t)>c0>0,t[0,)f(0)=1f(t)<0,t[0,)f′′(t)0,t[0,)\begin{cases}f(0)=1\\ f(t)>c_{0}>0,&\forall t\in[0,\infty)\\ f^{\prime}(0)=-1\\ f^{\prime}(t)<0,&\forall t\in[0,\infty)\\ f^{\prime\prime}(t)\geq 0,&\forall t\in[0,\infty)\end{cases} [Uncaptioned image]

Define a metric hh on MKM_{K} as

h={g,in SN(K)dt2+f(t)g|N(K),in [0,)×N(K),h=\begin{cases}g,&\text{in }S\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(K)\\ dt^{2}+f(t)\mathinner{g\rvert}_{\partial N(K)},&\text{in }[0,\infty)\times\partial N(K)\ ,\end{cases}

where tt is the coordinate in the [0,)[0,\infty)-factor. Similar to the situation in [EL17, Appendix C], if x,yN(K)x,y\in\partial N(K) are two points and c:[0,]MKc\colon\thinspace[0,\ell]\longrightarrow M_{K} a geodesic with c(s1)=xN(K)c(s_{1})=x\in\partial N(K) and c(s2)=yN(K)c(s_{2})=y\in\partial N(K), then there is a unique geodesic (t(s),c(s))[0,)×N(K)(t(s),c(s))\in[0,\infty)\times\partial N(K) so that

  • (0,c(s1))=(0,x)(0,c(s_{1}))=(0,x) and (t(),c(s2))=(0,y)(t(\ell),c(s_{2}))=(0,y), and

  • t:[0,][0,)t\colon\thinspace[0,\ell]\longrightarrow[0,\infty) is a Morse function with a unique maximum at some interior point s0(0,){s_{0}\in(0,\ell)}.

If we define

Ni . . =[0,i]×N(K),N_{i}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=[0,i]\times\partial N(K)\,,

then

N0N1N2[0,)×N(K),N_{0}\subset N_{1}\subset N_{2}\subset\cdots\subset[0,\infty)\times\partial N(K)\,,

is an exhaustion of [0,)×N(K)[0,\infty)\times\partial N(K) by compacts. Then given any geodesic c:[0,]MKc\colon\thinspace[0,\ell]\longrightarrow M_{K}, there exists some m0m\geq 0 so that c(t)Nmc(t)\in N_{m} for every t[0,]t\in[0,\ell]. In particular, if we restrict to the present situation in this paper, where every geodesic is a loop based at ξSN(K)\xi\in S\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(K). To this end fix some constant L0>0L_{0}>0 and assume γL0BMK\gamma\in\mathcal{F}_{L_{0}}BM_{K}, that is γ\gamma is a piecewise geodesic loop based at ξSN(K)\xi\in S\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N(K) with length bounded above by L0L_{0} (for details see Section 4.2). Then there is some m=m(L0,h)>0m=m(L_{0},h)>0 depending only on L0L_{0} and the metric hh so that γ(t)Nm\gamma(t)\in N_{m} for every tt. We prove that there exists some m0>0m_{0}>0 (depending on mm and the metric hh) so that the JJ-holomorphic strips lie inside of Nm0N_{m_{0}} by using the monotonicity lemma [Sik94, Proposition 4.7.2] (see also [CEL10, Lemma 3.4]).

Our metric hh defined in (A.1) extends to a metric on WKW_{K} such that it has bounded geometry in the terminology of [Sik94, Section 4]. Furthermore, since MKWKM_{K}\subset W_{K} is Lagrangian, the tuple (WK,J,MK,h)(W_{K},J,M_{K},h) is tame in the sense of [Sik94, Definition 4.1.1]. Let rW,CW>0r_{W},C_{W}>0 be constants so that for any x,yMKx,y\in M_{K}

dMK(x,y)rWdWK(x,y)CWdMK(x,y),d_{M_{K}}(x,y)\leq r_{W}\ \Rightarrow\ d_{W_{K}}(x,y)\leq C_{W}d_{M_{K}}(x,y)\,,

where dMKd_{M_{K}} and dWKd_{W_{K}} are the metrics induced by hh on MKM_{K} and WKW_{K} respectively. If we denote the lower bound on the injectivity radius by ρ\rho, we may assume rWρr_{W}\leq\rho.

Lemma A.1 ([Sik94, Proposition 4.7.2 (ii)]).

Let (V,J,W,μ)(V,J,W,\mu) be tame. Then there exist a positive constant C4(W)>0C_{4}(W)>0 with the following property. Let u:TVu\colon\thinspace T\longrightarrow V be a JJ-holomorphic curve so that u(T)B(x,r)Wu(\partial T)\subset\partial B(x,r)\cup W where xu(T)x\in u(T) and r<rWr<r_{W}. Then

area(u(T)B(x,r))C4(W)r2.\operatorname{area}(u(T)\cap B(x,r))\geq C_{4}(W)r^{2}\,.

We use this lemma with V=WKV=W_{K}, W=FMKW=F\cup M_{K} and μ=h\mu=h.

Theorem A.2.

Let A>0A>0 be arbitrary and consider a generator aACWΛK(F,F)a\in\mathcal{F}_{A}CW^{\ast}_{\varLambda_{K}}(F,F). Then there exists m>0m>0 so that imuNm\operatorname{im}u\subset N_{m} for any u(a)u\in\mathcal{M}(a).

Proof.

Consider a generator aACWΛK(F,F)a\in\mathcal{F}_{A}CW^{\ast}_{\varLambda_{K}}(F,F) and pick some u¯(a)u\in\overline{\mathcal{M}}(a). Then by 4.11 we have

L(ev(u))=𝔞(a)<A,L(\operatorname{ev}(u))=\mathfrak{a}(a)<A\,,

Because the JJ-holomorphic disk u(a)u\in\mathcal{M}(a) has boundary on the Reeb chord aa, the exact Lagrangian FDTξSF\cong DT^{\ast}_{\xi}S for ξMK\xi\in M_{K} and the geodesic γ . . =ev(u)\gamma\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\operatorname{ev}(u). Therefore there is some m>0m^{\prime}>0 (depending only on AA) so that imuNm\partial\operatorname{im}u\subset N_{m^{\prime}} for any u(a)u\in\mathcal{M}(a). Then pick some m>m>0m>m^{\prime}>0 (which a priori can be equal to \infty) and assume that imuNm\operatorname{im}u\subset N_{m}. We consider U . . =imu(NmNm)U\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\operatorname{im}u\cap(N_{m}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}N_{m^{\prime}}) and then we prove that mm is finite. Namely, fix some r<rWr<r_{W} and let v1,,vμUv_{1},\ldots,v_{\mu}\in U be the maximal number of points so that dWK(vi,vj)>2rd_{W_{K}}(v_{i},v_{j})>2r. Then we apply A.1 to each Ui . . =UB(vi,r)U_{i}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=U\cap B(v_{i},r) so that area(Ui)C4r2\operatorname{area}(U_{i})\geq C_{4}r^{2} for each i{1,,μ}i\in\left\{1,\ldots,\mu\right\}. Therefore

area(U)μarea(U1)μC4r2μarea(U)C4r2.\operatorname{area}(U)\geq\mu\operatorname{area}(U_{1})\geq\mu C_{4}r^{2}\Leftrightarrow\mu\leq\frac{\operatorname{area}(U)}{C_{4}r^{2}}\,.

Since 𝔞(a)\mathfrak{a}(a) is bounded by AA, so is the area of UU. Hence

μ<AC4r2<.\mu<\frac{A}{C_{4}r^{2}}<\infty\,.

This shows that there is some finite m>0m>0 such that imuNm\operatorname{im}u\subset N_{m} for every u(a)u\in\mathcal{M}(a). ∎

Appendix B Signs, gradings and orientations of moduli spaces

In this section, we use the same conventions and setup as in [Sei08, Section (11)] and [FOOO10, Section 8]. Pick some T¯mT\in\overline{\mathcal{H}}_{m} and consider the collection of Lagrangian branes F0#,,Fm#F_{0}^{\#},\ldots,F_{m}^{\#} of a cotangent fiber F0TξSWKF_{0}\cong T^{\ast}_{\xi}S\subset W_{K} at ξMK\xi\in M_{K} and a system of parallel copies F¯\overline{F} as in Section 3.3 and Section 2. Pick a word of generators 𝒂=a1am\boldsymbol{a}=a_{1}\cdot\cdots\cdot a_{m} where akCW(Fk1,Fk)a_{k}\in CW^{\ast}(F_{k-1},F_{k}), and pick abstract perturbation data so that (𝒂)\mathcal{M}(\boldsymbol{a}) is regular. Then for some u(𝒂)u\in\mathcal{M}(\boldsymbol{a}), denote the linearization of the operator ¯JT\overline{\partial}_{J_{T}} at the JJ-holomorphic disk uu by DuD_{u}. Then we have the following:

Lemma B.1 ([Abo12b, Lemma 6.1]).

With the choice as above there is a canonical up to homotopy isomorphism

detDuoξoa1oamoξ\det D_{u}\cong o_{\xi}\otimes o_{a_{1}}^{\vee}\otimes\cdots\otimes o_{a_{m}}^{\vee}\otimes o_{\xi}^{\vee}

and in particular

top(T¯(𝒂))top(T¯m)oξoa1oamoξ.\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{M}}(\boldsymbol{a}))\cong\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m})\otimes o_{\xi}\otimes o_{a_{1}}^{\vee}\otimes\cdots\otimes o_{a_{m}}^{\vee}\otimes o_{\xi}^{\vee}\,.

Since the orientation lines oxo_{x} are naturally graded by the indices of the linearized operators DxD_{x}, we have a nautral isomorphism coming from reordering tensor products of orientation lines which produces a Koszul sign

ox1ox2(1)|x1||x2|ox2ox1.o_{x_{1}}\otimes o_{x_{2}}\cong(-1)^{\mathinner{\!\left\lvert x_{1}\right\rvert}\mathinner{\!\left\lvert x_{2}\right\rvert}}o_{x_{2}}\otimes o_{x_{1}}\,.

Furthermore there are natural non-degnerate pairings

oxox.o_{x}\otimes o_{x}^{\vee}\cong\mathbb{R}\,.

From now on we use the following abbreviation: For the word 𝒂=a1am\boldsymbol{a}=a_{1}\cdots a_{m}, we let

o𝒂 . . =oa1oam.o_{\boldsymbol{a}}\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=o_{a_{1}}\otimes\cdots\otimes o_{a_{m}}\,.

As in (3.4) and (3.5) denote by m\mathcal{H}_{m} the moduli space of abstract JJ-holomorphic disks with m+2m+2 boundary punctures, and its Deligne–Mumford compactification by ¯m\overline{\mathcal{H}}_{m}. Then the codimension one boundary ¯m\partial\overline{\mathcal{H}}_{m} is covered by the natural inclusions of the following strata

(B.1) ¯m1\displaystyle\overline{\mathcal{H}}_{m_{1}} ׯm2,m1+m2=m\displaystyle\times\overline{\mathcal{H}}_{m_{2}},\,m_{1}+m_{2}=m
(B.2) ¯m1\displaystyle\overline{\mathcal{H}}_{m_{1}} ׯm2,m1+m2=m+1.\displaystyle\times\overline{\mathcal{R}}_{m_{2}},\,m_{1}+m_{2}=m+1\,.

Here m\mathcal{R}_{m} is the Deligne–Mumford space of unit disks in the complex plane with m+1m+1 boundary punctures that are oriented counterclockwise. We would like to compare the product orientation of each of the strata with the boundary orientation on ¯m\partial\overline{\mathcal{H}}_{m}. The orientation of the boundary is determined as follows. Any orientation on a manifold XX induces an orientation on its boundary via the outward normal first-rule. More precisely via the canonical isomorphism

topTXνXtopTX,\bigwedge\nolimits^{\mathrm{top}}TX\cong\nu_{\partial X}\otimes\bigwedge\nolimits^{\mathrm{top}}T\partial X\,,

where νX\nu_{\partial X} is the normal bundle of X\partial X which is canonically trivialized by the outwards normal vector along the boundary. Following the conventions in [Sei08, Abo10, Abo12b] there is a choice of coherent orientations on ¯m\overline{\mathcal{H}}_{m} such that the boundary strata (B.1) and (B.2) differs from the boundary orientation on ¯m\partial\overline{\mathcal{H}}_{m} by a sign (1)1(-1)^{{\dagger}_{1}} and (1)2(-1)^{{\dagger}_{2}} respectively where we have

(B.3) 1\displaystyle{\dagger}_{1} =m1\displaystyle=m_{1}
(B.4) 2\displaystyle{\dagger}_{2} =m2(mk)+k+m2,\displaystyle=m_{2}(m-k)+k+m_{2}\,,

and ¯m2\overline{\mathcal{R}}_{m_{2}} is attached to the (k+1)(k+1)-th outgoing leaf of ¯m1\overline{\mathcal{H}}_{m_{1}} (cf. [Sei08, (12.22)]). The first sign 1{\dagger}_{1} is obtained from [Sei08, (12.22)] by using m=m2+1m=m_{2}+1, d=m1+m2+1d=m_{1}+m_{2}+1 and n=dn=d since ¯m2\overline{\mathcal{H}}_{m_{2}} is attached to ¯m1\overline{\mathcal{H}}_{m_{1}} at the last outgoing leaf.

The second sign 2{\dagger}_{2} is obtained from [Sei08, (12.22)] by using m=m2m=m_{2}, d=m1+m2d=m_{1}+m_{2} and n=kn=k.

Proof of 3.5.

We consider the moduli space ¯(𝒂)\overline{\mathcal{M}}(\boldsymbol{a}) and the stratification of its codimension one boundary as in (3.6). We first consider the strata of the form ¯(𝒂)ׯ(𝒂′′)\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}) where 𝒂𝒂′′=𝒂\boldsymbol{a}^{\prime}\boldsymbol{a}^{\prime\prime}=\boldsymbol{a}. Then, using B.1 we have

top(T¯(𝒂))top(T¯(𝒂′′))\displaystyle\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{M}}(\boldsymbol{a}^{\prime}))\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime})) =top(T¯m1)oξo𝒂oξ\displaystyle=\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m_{1}})\otimes o_{\xi}\otimes o_{\boldsymbol{a}^{\prime}}^{\vee}\otimes o_{\xi}^{\vee}
top(T¯m2)oξo𝒂′′oξ.\displaystyle\quad\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m_{2}})\otimes o_{\xi}\otimes o_{\boldsymbol{a}^{\prime\prime}}^{\vee}\otimes o_{\xi}^{\vee}\,.

Reordering the factors so that top(T¯m2)\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m_{2}}) becomes adjacent to top(T¯m1)\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m_{1}}) introduces the Koszul sign (1)1(-1)^{\rotatebox[origin={c}]{180.0}{${\dagger}$}_{1}} where

1=(m2+1)(i=1m1|ai|),\rotatebox[origin={c}]{180.0}{${\dagger}$}_{1}=(m_{2}+1)\left(\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)\,,

since dim¯m2=m2+1\dim\overline{\mathcal{H}}_{m_{2}}=m_{2}+1. Canceling the adjacent factors oξo_{\xi}^{\vee} and oξo_{\xi} then gives

top(T¯m1)top(T¯m2)oξo𝒂o𝒂′′oξ.\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m_{1}})\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m_{2}})\otimes o_{\xi}\otimes o_{\boldsymbol{a}^{\prime}}^{\vee}\otimes o_{\boldsymbol{a}^{\prime\prime}}^{\vee}\otimes o_{\xi}^{\vee}\,.

Then by (B.3) we get a sign (1)1(-1)^{{\dagger}_{1}} when comparing the product orientation of ¯m1ׯm2\overline{\mathcal{H}}_{m_{1}}\times\overline{\mathcal{H}}_{m_{2}} with the boundary orientation of ¯m\partial\overline{\mathcal{H}}_{m}. After these reorderings we arrive at

top(T¯m)oξo𝒂o𝒂′′oξ=top(T¯m)oξo𝒂oξ,\bigwedge\nolimits^{\mathrm{top}}(T\partial\overline{\mathcal{H}}_{m})\otimes o_{\xi}\otimes o_{\boldsymbol{a}^{\prime}}^{\vee}\otimes o_{\boldsymbol{a}^{\prime\prime}}^{\vee}\otimes o_{\xi}^{\vee}=\bigwedge\nolimits^{\mathrm{top}}(T\partial\overline{\mathcal{H}}_{m})\otimes o_{\xi}\otimes o_{\boldsymbol{a}}^{\vee}\otimes o_{\xi}^{\vee}\,,

which is canonically isomorphic to top(T¯(𝒂))\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{M}}(\boldsymbol{a})). The total sign difference between the product orientation on ¯(𝒂)ׯ(𝒂′′)\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})\times\overline{\mathcal{M}}(\boldsymbol{a}^{\prime\prime}) and the boundary orientation on ¯(𝒂)\partial\overline{\mathcal{M}}(\boldsymbol{a}) is therefore

1=1+1=(m2+1)(i=1m1|ai|)+m1.{\ddagger}_{1}=\rotatebox[origin={c}]{180.0}{${\dagger}$}_{1}+{\dagger}_{1}=(m_{2}+1)\left(\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+m_{1}\,.

Similarly, we compare the product orientation of ¯(𝒂𝒂~)×cw(𝒂~)\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) with the boundary orientation on ¯(𝒂)\partial\overline{\mathcal{M}}(\boldsymbol{a}). Recall from (3.7) that if 𝒂~𝒂\tilde{\boldsymbol{a}}\subset\boldsymbol{a} is a subword at position t+1t+1, then 𝒂𝒂~\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}} denotes the word 𝒂\boldsymbol{a} with the subword 𝒂~\tilde{\boldsymbol{a}} replaced by an auxiliary generator yy. Again by B.1 we therefore have

top(T¯(𝒂𝒂~))top(Tcw(𝒂~))\displaystyle\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}}))\otimes\bigwedge\nolimits^{\mathrm{top}}(T\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}})) =top(T¯t+1+r)oξo𝒂𝒂~oξ\displaystyle=\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{t+1+r})\otimes o_{\xi}\otimes o_{\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}}}^{\vee}\otimes o_{\xi}^{\vee}
top(T¯s)oyo𝒂~\displaystyle\quad\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{R}}_{s})\otimes o_{y}\otimes o_{\tilde{\boldsymbol{a}}}^{\vee}

Assuming that cw(𝒂~)\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) is rigid means especially that |y|=2s+|y|i=1s|at+i|\mathinner{\!\left\lvert y\right\rvert}=2-s+\mathinner{\!\left\lvert y\right\rvert}-\sum_{i=1}^{s}\mathinner{\!\left\lvert a_{t+i}\right\rvert} and so we move o(𝒂𝒂~)2oξo_{(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2}}^{\vee}\otimes o_{\xi}^{\vee} past top(T¯s)oyo𝒂~\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{R}}_{s})\otimes o_{y}\otimes o_{\tilde{\boldsymbol{a}}}^{\vee} without introducing any sign and arrive at

top(T¯t+1+r)oξo(𝒂𝒂~)1oytop(T¯s)oyo𝒂~o(𝒂𝒂~)2oξ.\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{t+1+r})\otimes o_{\xi}\otimes o_{(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{1}}^{\vee}\otimes o_{y}^{\vee}\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{R}}_{s})\otimes o_{y}\otimes o_{\tilde{\boldsymbol{a}}}^{\vee}\otimes o_{(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2}}^{\vee}\otimes o_{\xi}^{\vee}\,.

Because dim¯s=s\dim\overline{\mathcal{R}}_{s}=s, moving top(T¯s)\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{R}}_{s}) to the front and adjacent to top(T¯t+1+r)\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{t+1+r}) gives

top(T¯t+1+r)top(T¯s)oξo(𝒂𝒂~)1oyoyo𝒂~o(𝒂𝒂~)2oξ,\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{t+1+r})\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{R}}_{s})\otimes o_{\xi}\otimes o_{(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{1}}^{\vee}\otimes o_{y}^{\vee}\otimes o_{y}\otimes o_{\tilde{\boldsymbol{a}}}^{\vee}\otimes o_{(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{1.80835pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2}}^{\vee}\otimes o_{\xi}^{\vee}\,,

with a sign difference of (1)2(-1)^{\rotatebox[origin={c}]{180.0}{${\dagger}$}_{2}} where

2=s(|ξ|+|y|+i=1t|ai|)=s(|ξ|+i=1t+s|ai|).\rotatebox[origin={c}]{180.0}{${\dagger}$}_{2}=s\left(\mathinner{\!\left\lvert\xi\right\rvert}+\mathinner{\!\left\lvert y\right\rvert}+\sum_{i=1}^{t}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)=s\left(\mathinner{\!\left\lvert\xi\right\rvert}+\sum_{i=1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)\,.

Recall from the assumptions in 3.5 that 𝒂~𝒂\tilde{\boldsymbol{a}}\subset\boldsymbol{a} is a subword at position t+1t+1.

Then using oyoyo_{y}^{\vee}\otimes o_{y}\cong\mathbb{R} and 𝒂=(𝒂𝒂~)1𝒂~(𝒂𝒂~)2\boldsymbol{a}=(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{1}\tilde{\boldsymbol{a}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})_{2} this collapses to

top(T¯t+1+r)top(T¯s)oξo𝒂oξ,\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{t+1+r})\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{R}}_{s})\otimes o_{\xi}\otimes o_{\boldsymbol{a}}^{\vee}\otimes o_{\xi}^{\vee}\,,

and using (B.4), top(T¯t+1+r)top(T¯s)top(T¯m)\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{t+1+r})\otimes\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{R}}_{s})\cong\bigwedge\nolimits^{\mathrm{top}}(T\overline{\mathcal{H}}_{m}) with a sign difference of (1)2(-1)^{{\dagger}_{2}}. The total sign difference between the product orientation of ¯(𝒂𝒂~)×cw(𝒂~)\overline{\mathcal{M}}(\boldsymbol{a}\mathbin{\vbox{\hbox{$\rotatebox{20.0}{\raisebox{2.58334pt}{$\scriptscriptstyle\mathrlap{\setminus}{\hskip 0.2pt}\setminus$}}$}}}\tilde{\boldsymbol{a}})\times\mathcal{M}^{\mathrm{cw}}(\tilde{\boldsymbol{a}}) and the boundary orientation on ¯(𝒂)\partial\overline{\mathcal{M}}(\boldsymbol{a}) is therefore

2=2+2=s(|ξ|+i=1t+s|ai|)+s(mt)+t+s.{\ddagger}_{2}=\rotatebox[origin={c}]{180.0}{${\dagger}$}_{2}+{\dagger}_{2}=s\left(\mathinner{\!\left\lvert\xi\right\rvert}+\sum_{i=1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+s(m-t)+t+s\,.

Proof of 3.10 (continued).

To confirm that the signs match up in the AA_{\infty}-relation

Ψm+m1+m2=mP(Ψm2Ψm1)=r+s+t=m(1)tΨr+1+t(idrμsidt),\partial\varPsi_{m}+\sum_{m_{1}+m_{2}=m}P(\varPsi_{m_{2}}\otimes\varPsi_{m_{1}})=\sum_{r+s+t=m}(-1)^{\maltese_{t}}\varPsi_{r+1+t}(\operatorname{id}^{\otimes r}\otimes\mu^{s}\otimes\operatorname{id}^{\otimes t})\,,

we look at the terms one by one and compute the sign that is in front of each term. In the first term Ψm\partial\varPsi_{m} it is only the sign from Ψm\varPsi_{m} that is taken into account, namely (1)§(-1)^{\S} where

§=i=1mi|ai|+(|ξ|+m)i=1m|ai|.\S=\sum_{i=1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m)\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}\,.

The second term has a sign coming from:

  1. (1)

    The definition of the Pontryagin product PP in (3.2) contributes with a sign (1)(-1)^{\circ} where

    =|Ψm1(am1a1)|=dim¯(𝒂)=1+m1i=1m1|ai|,\circ=\mathinner{\!\left\lvert\varPsi_{m_{1}}(a_{m_{1}}\otimes\cdots\otimes a_{1})\right\rvert}=\dim\overline{\mathcal{M}}(\boldsymbol{a}^{\prime})=-1+m_{1}-\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}\,,
  2. (2)

    the difference between the product orientation on ¯m1ׯm2\overline{\mathcal{H}}_{m_{1}}\times\overline{\mathcal{H}}_{m_{2}} and the boundary orientation on ¯m\partial\overline{\mathcal{H}}_{m} is (1)1(-1)^{{\ddagger}_{1}}, where m1+m2=mm_{1}+m_{2}=m and

    1=(m2+1)(i=1m1|ai|)+m1,{\ddagger}_{1}=(m_{2}+1)\left(\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+m_{1}\,,

    as in 3.5,

  3. (3)

    the definition of Ψm1(am1a1)\varPsi_{m_{1}}(a_{m_{1}}\otimes\cdots\otimes a_{1}) in (3.19) contributes with a sign (1)§1(-1)^{\S_{1}} where

    §1=i=1m1i|ai|+(|ξ|+m1)i=1m1|ai|,\S_{1}=\sum_{i=1}^{m_{1}}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m_{1})\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}\,,

    and

  4. (4)

    the definition of Ψm2(amam1+1)\varPsi_{m_{2}}(a_{m}\otimes\cdots\otimes a_{m_{1}+1}) in (3.19) contributes with a sign (1)§2(-1)^{\S_{2}} where

    §2=i=m1+1m(im1)|ai|+(|ξ|+m2)i=m1+1m|ai|.\S_{2}=\sum_{i=m_{1}+1}^{m}(i-m_{1})\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m_{2})\sum_{i=m_{1}+1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}\,.

Now it is straightforward to check that +1+§1+§2=1+§(mod2)\circ+{\ddagger}_{1}+\S_{1}+\S_{2}=1+\S\pmod{2}.

+1+§1+§2\displaystyle\circ+{\ddagger}_{1}+\S_{1}+\S_{2} =1+m1i=1m1|ai|+(m2+1)(i=1m1|ai|)+m1+i=1m1i|ai|\displaystyle=-1+m_{1}-\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}+(m_{2}+1)\left(\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+m_{1}+\sum_{i=1}^{m_{1}}i\mathinner{\!\left\lvert a_{i}\right\rvert}
+(|ξ|+m1)i=1m1|ai|+i=m1+1m(im1)|ai|+(|ξ|+m2)i=m1+1m|ai|\displaystyle\quad+(\mathinner{\!\left\lvert\xi\right\rvert}+m_{1})\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}+\sum_{i=m_{1}+1}^{m}(i-m_{1})\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m_{2})\sum_{i=m_{1}+1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}
=1+m2i=1m1|ai|+i=1mi|ai|+(|ξ|+m1)i=1m1|ai|+m1i=m1+1m|ai|\displaystyle=1+m_{2}\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}+\sum_{i=1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m_{1})\sum_{i=1}^{m_{1}}\mathinner{\!\left\lvert a_{i}\right\rvert}+m_{1}\sum_{i=m_{1}+1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}
+(|ξ|+m2)i=m1+1m|ai|\displaystyle\quad+(\mathinner{\!\left\lvert\xi\right\rvert}+m_{2})\sum_{i={m_{1}+1}}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}
=1+i=1mi|ai|+(|ξ|+m)i=1m|ai|=1+§(mod2).\displaystyle=1+\sum_{i=1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m)\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}=1+\S\pmod{2}\,.

Next we consider the term in the right hand side. Let y . . =μs(at+sat+1)y\mathrel{\vbox{\hbox{$\raisebox{-0.86108pt}{\scriptsize.}$}\hbox{\scriptsize.}}}=\mu^{s}(a_{t+s}\otimes\cdots\otimes a_{t+1}). This sum has a sign coming from:

  1. (1)

    The difference between the product orientation on ¯r+1+tׯs\overline{\mathcal{H}}_{r+1+t}\times\overline{\mathcal{R}}_{s} and the boundary orientation on ¯m\partial\overline{\mathcal{H}}_{m} is (1)2(-1)^{{\ddagger}_{2}} where r+s+t=mr+s+t=m and

    2=s(|ξ|+i=1t+s|ai|)+s(mt)+t+s,{\ddagger}_{2}=s\left(\mathinner{\!\left\lvert\xi\right\rvert}+\sum_{i=1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+s(m-t)+t+s\,,
  2. (2)

    the definition of μs(at+sat+1)\mu^{s}(a_{t+s}\otimes\cdots\otimes a_{t+1}) in (2.1) contributes with a sign (1)(-1)^{\diamond} where

    =i=t+1t+s(it)|ai|,\diamond=\sum_{i=t+1}^{t+s}(i-t)\mathinner{\!\left\lvert a_{i}\right\rvert}\,,
  3. (3)

    the definition of Ψr+1+t(amat+s+1yata1)\varPsi_{r+1+t}(a_{m}\otimes\cdots\otimes a_{t+s+1}\otimes y\otimes a_{t}\otimes\cdots\otimes a_{1}) in (3.19) contributes with a sign (1)§~(-1)^{\tilde{\S}} where

    §~\displaystyle\tilde{\S} =i=1ti|ai|+(t+1)|y|+i=t+s+1m(is+1)|ai|+(|ξ|+r+t+1)(i=1t|ai|+|y|+i=t+s+1m|ai|).\displaystyle=\sum_{i=1}^{t}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(t+1)\mathinner{\!\left\lvert y\right\rvert}+\sum_{i=t+s+1}^{m}(i-s+1)\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+r+t+1)\left(\sum_{i=1}^{t}\mathinner{\!\left\lvert a_{i}\right\rvert}+\mathinner{\!\left\lvert y\right\rvert}+\sum_{i=t+s+1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)\,.

    Note that since we assume that cw(at+1at+s)\mathcal{M}^{\mathrm{cw}}(a_{t+1}\cdots a_{t+s}) is rigid, we have

    |y|=2s+i=t+1t+s|ai|=s+i=t+1t+s|ai|(mod2),\mathinner{\!\left\lvert y\right\rvert}=2-s+\sum_{i=t+1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}=s+\sum_{i=t+1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}\pmod{2}\,,

    hence we get

    §~\displaystyle\tilde{\S} =i=1ti|ai|+(t+1)(s+i=t+1t+s|ai|)+i=t+s+1m(is+1)|ai|\displaystyle=\sum_{i=1}^{t}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(t+1)\left(s+\sum_{i=t+1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+\sum_{i=t+s+1}^{m}(i-s+1)\mathinner{\!\left\lvert a_{i}\right\rvert}
    +(|ξ|+r+t+1)(s+i=1m|ai|)\displaystyle\quad+(\mathinner{\!\left\lvert\xi\right\rvert}+r+t+1)\left(s+\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)
    =i=1ti|ai|+(t+1)i=t+1t+s|ai|+(s+1)i=t+s+1m|ai|\displaystyle=\sum_{i=1}^{t}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(t+1)\sum_{i=t+1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}+(s+1)\sum_{i=t+s+1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}
    +i=t+s+1mi|ai|+(|ξ|+r)s+(|ξ|+r+t+1)i=1m|ai|(mod2).\displaystyle\quad+\sum_{i=t+s+1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+r)s+(\mathinner{\!\left\lvert\xi\right\rvert}+r+t+1)\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}\pmod{2}\,.

    It is then again a straightforward calculation to show that 2++§~+t=§(mod2){\ddagger}_{2}+\diamond+\tilde{\S}+\maltese_{t}=\S\pmod{2}.

    2++§~+t\displaystyle\textdaggerdbl_{2}+\diamond+\tilde{\S}+\maltese_{t} =s(|ξ|+i=1t+s|ai|)+s(mt)+t+s+i=t+1t+s(it)|ai|\displaystyle=s\left(\mathinner{\!\left\lvert\xi\right\rvert}+\sum_{i=1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}\right)+s(m-t)+t+s+\sum_{i=t+1}^{t+s}(i-t)\mathinner{\!\left\lvert a_{i}\right\rvert}
    +i=1ti|ai|+(t+1)i=t+1t+s|ai|+(s+1)i=t+s+1m|ai|\displaystyle\quad+\sum_{i=1}^{t}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(t+1)\sum_{i=t+1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}+(s+1)\sum_{i=t+s+1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}
    +i=t+s+1mi|ai|+(|ξ|+r)s+(|ξ|+r+t+1)i=1m|ai|+i=1t|ai|+t\displaystyle\quad+\sum_{i=t+s+1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+r)s+(\mathinner{\!\left\lvert\xi\right\rvert}+r+t+1)\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}+\sum_{i=1}^{t}\mathinner{\!\left\lvert a_{i}\right\rvert}+t
    =si=1t+s|ai|+s(mt)+t+s+i=t+1t+si|ai|+i=1ti|ai|\displaystyle=s\sum_{i=1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}+s(m-t)+t+s+\sum_{i=t+1}^{t+s}i\mathinner{\!\left\lvert a_{i}\right\rvert}+\sum_{i=1}^{t}i\mathinner{\!\left\lvert a_{i}\right\rvert}
    +i=t+1t+s|ai|+(s+1)i=t+s+1m|ai|+i=t+s+1mi|ai|\displaystyle\quad+\sum_{i=t+1}^{t+s}\mathinner{\!\left\lvert a_{i}\right\rvert}+(s+1)\sum_{i=t+s+1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}+\sum_{i=t+s+1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}
    +rs+(|ξ|+r+t+1)i=1m|ai|+i=1t|ai|+t\displaystyle\quad+rs+(\mathinner{\!\left\lvert\xi\right\rvert}+r+t+1)\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}+\sum_{i=1}^{t}\mathinner{\!\left\lvert a_{i}\right\rvert}+t
    =mi=1m|ai|+rs+s2+s+i=1mi|ai|+rs+|ξ|i=1m|a|i\displaystyle=m\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}+rs+s^{2}+s+\sum_{i=1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}+rs+\mathinner{\!\left\lvert\xi\right\rvert}\sum_{i=1}^{m}\mathinner{\!\left\lvert a\right\rvert}_{i}
    =i=1mi|ai|+(|ξ|+m)i=1m|ai|=§(mod2).\displaystyle=\sum_{i=1}^{m}i\mathinner{\!\left\lvert a_{i}\right\rvert}+(\mathinner{\!\left\lvert\xi\right\rvert}+m)\sum_{i=1}^{m}\mathinner{\!\left\lvert a_{i}\right\rvert}=\S\pmod{2}\,.

References

  • [Abo10] M. Abouzaid, A geometric criterion for generating the Fukaya category, Publications mathématiques de l’IHÉS, 112 (2010), pp. 191–240.
  • [Abo12a] ———, Framed bordism and Lagrangian embeddings of exotic spheres, Annals of Mathematics, (2012), pp. 71–185.
  • [Abo12b] ———, On the wrapped Fukaya category and based loops, Journal of Symplectic Geometry, 10 (2012), pp. 27–79.
  • [AENV14] M. Aganagic, T. Ekholm, L. Ng, and C. Vafa, Topological strings, D-model, and knot contact homology, Advances in Theoretical and Mathematical Physics, 18 (2014), pp. 827–956.
  • [Arn67] V. I. Arnol’d, Characteristic class entering in quantization conditions, Functional Analysis and its applications, 1 (1967), pp. 1–13.
  • [AS06] A. Abbondandolo and M. Schwarz, On the Floer homology of cotangent bundles, Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 59 (2006), pp. 254–316.
  • [AS10] M. Abouzaid and P. Seidel, An open string analogue of Viterbo functoriality, Geometry & Topology, 14 (2010), pp. 627–718.
  • [BEE12] F. Bourgeois, T. Ekholm, and Y. Eliashberg, Effect of Legendrian surgery, Geometry & Topology, 16 (2012), pp. 301–389.
  • [BKO19] Y. Bae, S. Kim, and Y.-G. Oh, A wrapped Fukaya category of knot complement, arXiv:1901.02239, (2019).
  • [BT13] R. Bott and L. W. Tu, Differential forms in algebraic topology, volume 82, Springer Science & Business Media (2013).
  • [CEL10] K. Cieliebak, T. Ekholm, and J. Latschev, Compactness for holomorphic curves with switching Lagrangian boundary conditions, Journal of Symplectic Geometry, 8 (2010), pp. 267–298.
  • [CELN17] K. Cieliebak, T. Ekholm, J. Latschev, and L. Ng, Knot contact homology, string topology, and the cord algebra, J. Éc. polytech. Math, 4 (2017), pp. 661–780.
  • [CL09] K. Cieliebak and J. Latschev, The role of string topology in symplectic field theory, New perspectives and challenges in symplectic field theory, CRM Proc. Lecture Notes, 49 (2009), pp. 113–146.
  • [DR16] G. Dimitroglou Rizell, Lifting pseudo-holomorphic polygons to the symplectisation of P×P\times\mathbb{R} and applications, Quantum Topology, 7 (2016), pp. 29–105.
  • [Dui76] J. Duistermaat, On the Morse index in variational calculus, Advances in Mathematics, 21 (1976), pp. 173–195.
  • [EENS13] T. Ekholm, J. B. Etnyre, L. Ng, and M. G. Sullivan, Knot contact homology, Geometry & Topology, 17 (2013), pp. 975–1112.
  • [EES05] T. Ekholm, J. Etnyre, and M. Sullivan, The contact homology of Legendrian submanifolds in 2n+1\mathbb{R}^{2n+1}, Journal of Differential Geometry, 71 (2005), pp. 177–305.
  • [EES07] ———, Legendrian contact homology in P×P\times\mathbb{R}, Transactions of the American Mathematical Society, 359 (2007), pp. 3301–3335.
  • [EHK16] T. Ekholm, K. Honda, and T. Kálmán, Legendrian knots and exact Lagrangian cobordisms, Journal of the European Mathematical Society, 18 (2016), pp. 2627–2689.
  • [Ekh06] T. Ekholm, Rational symplectic field theory over 2\mathbb{Z}_{2} for exact Lagrangian cobordisms, math/0612029, (2006).
  • [Ekh12] ———, Rational SFT, linearized Legendrian contact homology, and Lagrangian Floer cohomology, in Perspectives in analysis, geometry, and topology, Springer (2012), pp. 109–145.
  • [Ekh19] ———, Holomorphic curves for Legendrian surgery, arXiv:1906.07228, (2019).
  • [EKS16] T. Ekholm, T. Kragh, and I. Smith, Lagrangian exotic spheres, Journal of Topology and Analysis, 8 (2016), pp. 375–397.
  • [EL17] T. Ekholm and Y. Lekili, Duality between Lagrangian and Legendrian invariants, arXiv:1701.01284, (2017).
  • [ENS16] T. Ekholm, L. Ng, and V. Shende, A complete knot invariant from contact homology, arXiv:1606.07050, (2016).
  • [EO17] T. Ekholm and A. Oancea, Symplectic and contact differential graded algebras, Geometry & Topology, 21 (2017), pp. 2161–2230.
  • [ES16] T. Ekholm and I. Smith, Exact Lagrangian immersions with a single double point, Journal of the American Mathematical Society, 29 (2016), pp. 1–59.
  • [FOOO10] K. Fukaya, Y.-G. Oh, H. Ohta, and K. Ono, Lagrangian intersection Floer theory: anomaly and obstruction, Part II, volume 2, American Mathematical Soc. (2010).
  • [Fra08] U. Frauenfelder, Gromov convergence of pseudoholomorphic disks, Journal of Fixed Point Theory and Applications, 3 (2008), pp. 215–271.
  • [GPS18a] S. Ganatra, J. Pardon, and V. Shende, Microlocal Morse theory of wrapped Fukaya categories, arXiv:1809.08807, (2018).
  • [GPS18b] ———, Sectorial descent for wrapped Fukaya categories, arXiv:1809.03427, (2018).
  • [GPS20] ———, Covariantly functorial wrapped Floer theory on Liouville sectors, Publications mathématiques de l’IHÉS, 131 (2020), pp. 73–200.
  • [Hat02] A. Hatcher, Algebraic Topology, Cambridge University Press (2002).
  • [Jos08] J. Jost, Riemannian geometry and geometric analysis, Springer Science & Business Media (2008).
  • [Mil63] J. Milnor, Morse Theory, Annals of Mathematic Studies, 51 (1963).
  • [MW18] C. Y. Mak and W. Wu, Dehn twist exact sequences through Lagrangian cobordism, Compositio Mathematica, 154 (2018), pp. 2485–2533.
  • [Ng08] L. Ng, Framed knot contact homology, Duke Mathematical Journal, 141 (2008), pp. 365–406.
  • [Rol76] D. Rolfsen, Knots and links, volume 346, American Mathematical Soc. (1976).
  • [RS93] J. Robbin and D. Salamon, The Maslov index for paths, Topology, 32 (1993), pp. 827–844.
  • [RS95] ———, The spectral flow and the Maslov index, Bulletin of the London Mathematical Society, 27 (1995), pp. 1–33.
  • [RS01] J. W. Robbin and D. A. Salamon, Asymptotic behaviour of holomorphic strips, in Annales de l’Institut Henri Poincare (C) Non Linear Analysis, volume 18(5), Elsevier Masson (2001), pp. 573–612.
  • [Sei08] P. Seidel, Fukaya categories and Picard-Lefschetz theory, European Mathematical Society (2008).
  • [She16] V. Shende, The conormal torus is a complete knot invariant, arXiv:1604.03520, (2016).
  • [Shu10] M. G. Shulman, Equivariant spectral sequences for local coefficients, arXiv:1005.0379, (2010).
  • [Sik94] J.-C. Sikorav, Some properties of holomorphic curves in almost complex manifolds, in Holomorphic curves in symplectic geometry, Springer (1994), pp. 165–189.
  • [Syl19] Z. Sylvan, On partially wrapped Fukaya categories, Journal of Topology, 12 (2019), pp. 372–441.
  • [Wei71] A. Weinstein, Symplectic manifolds and their Lagrangian submanifolds, Advances in mathematics, 6 (1971), pp. 329–346.
  • [Wen16] C. Wendl, Lectures on Symplectic Field Theory, arXiv:1612.01009, (2016).